title
stringlengths
31
247
content
stringlengths
15
7.12M
2401.11065v1.Extremely_strong_spin_orbit_coupling_effect_in_light_element_altermagnetic_materials.pdf
Extremely strong spin-orbit coupling effect in light element altermagnetic materials Shuai Qu1,3, Ze-Feng Gao1,2,3, Hao Sun2, Kai Liu1,3, Peng-Jie Guo1,3,∗and Zhong-Yi Lu1,3† 1. Department of Physics and Beijing Key Laboratory of Opto-electronic Functional Materials &Micro-nano Devices. Renmin University of China, Beijing 100872, China 2. Gaoling School of Artificial Intelligence, Renmin University of China, Beijing, China and 3. Key Laboratory of Quantum State Construction and Manipulation (Ministry of Education), Renmin University of China, Beijing 100872, China (Dated: January 23, 2024) Spin-orbit coupling is a key to realize many novel physical effects in condensed matter physics, but the mechanism to achieve strong spin-orbit coupling effect in light element antiferromagnetic compounds has not been explored. In this work, based on symmetry analysis and the first-principles electronic structure calculations, we demonstrate that strong spin-orbit coupling effect can be real- ized in light element altermagnetic materials, and propose a mechanism for realizing the correspond- ing effective spin-orbit coupling. This mechanism reveals the cooperative effect of crystal symmetry, electron occupation, electronegativity, electron correlation, and intrinsic spin-orbit coupling. Our work not only promotes the understanding of light element compounds with strong spin-orbit cou- pling effect, but also provides an alternative for realizing light element compounds with an effective strong spin-orbit coupling. Introduction. Spin-orbit coupling (SOC) is ubiquitous in realistic materials and crucial for many novel physical phenomena emerging in condensed matter physics, in- cluding topological physics [1–3], anomalous Hall effect [4], spin Hall effect [5, 6], magnetocrystalline anisotropy [7] and so on. For instance, quantum anomalous Hall (QAH) insulators are characterized by non-zero Chern numbers [8]. The Chern number is derived from the inte- gration of Berry curvature over the occupied state of the Brillouin zone (BZ). For collinear ferromagnetic and an- tiferromagnetic systems, the integral of Berry curvature over the occupied state of the Brillouin zone must be zero without SOC due to the spin symmetry {C⊥ 2T||T}. Here theC⊥ 2andTrepresent the 180 degrees rotation perpen- dicular to the spin direction and time-reversal operation, respectively. Therefore, QAH effect can only be real- ized in collinear magnetic systems when SOC is included [9, 10]. On the other hand, strong SOC may open up a large nontrivial bandgap, which is very important to re- alize QAH effect at high temperatures. In general, strong SOC exists in heavy element compounds. Unfortunately, the chemical bonds of heavy element compounds are weaker than those of light element compounds, which leads to more defects in heavy element compounds. Thus, the stability to realize exotic functionalities in heavy el- ement compounds is relatively weak. An interesting question is whether the strong SOC ef- fect can be achieved in light element compounds. Very recently, Li et al. demonstrated that the SOC can be en- hanced in light element ferromagnetic materials, which derives from the cooperative effects of crystal symme- try, electron occupancy, electron correlation, and intrin- sic SOC [11]. This provides a new direction for the design of light element materials with strong effective SOC. ∗guopengjie@ruc.edu.cn †zlu@ruc.edu.cnVery recently, based on spin group theory, altermag- netism is proposed as a new magnetic phase distinct from ferromagnetism and conventional collinear antiferromag- netism [12, 13]. Moreover, altermagnetic materials have a wide range of electronic properties, which cover met- als, semi-metals, semiconductors, and insulators [13, 14]. Different from ferromagnetic materials with s-wave spin polarization, altermagnetic materials have k-dependent spin polarization, which results in many exotic physical effects [12, 13, 15–21]. With spin-orbit coupling, similar to the case of ferromagnetic materials, the time-reversal symmetry-breaking macroscopic phenomena can be also realized in altermagnetic materials [10, 22–24]. Never- theless, altermagnetism is proposed based on spin group theory and the predicted altermagnetic materials basi- cally are light element compounds [13, 14]. Therefore, it is very important to propose a mechanism to enhance SOC in light element compounds with altermagnetism and predict the corresponding compounds with strong SOC effect. In this work, based on symmetry analysis and the first- principles electronic structure calculations, we predict that the light element compound NiF 3is an i-wave al- termagnetic material with extremely strong SOC effect. Then, we propose a mechanism to enhance SOC effect in light element compounds with altermagnetism, which reveals the cooperative effects of crystal symmetry, elec- tron occupation, electronegativity, electron correlation, and intrinsic SOC. We also explain the weak SOC effect in altermagnetic materials VF 3, CrF 3, FeF 3, CoF 3. Results and discussion. The NiF 3takes rhombohedral structure with nonsymmorphic R −3c (167) space group symmetry, as shown in Fig. 1 (a)and(b). The corre- sponding elementary symmetry operations are C 3z, C1 2t and I, which yield the point group D 3d. The t repre- sents (1/2, 1/2, 1/2) fractional translation. To confirm the magnetic ground state of NiF 3, we consider six dif- ferent collinear magnetic structures, including one ferro-arXiv:2401.11065v1 [cond-mat.mtrl-sci] 19 Jan 20242 FM AFM2 AFM4 AFM3 AFM5 AFM1 a b c Ni F a b c 140 °(a) (c)(e)(g) (b) (d) (f) (h) a b c FIG. 1. The crystal structure and six collinear magnetic structures of NiF 3.(a)and(b)are side and top views of the crystal structure, respectively. The cyan arrow represents the direction of easy magnetization axis. (c)-(h)are six different collinear magnetic structures including one ferromagnetic and five different collinear antiferromagnetic structures. The bond angle of Ni−F−Ni for the nearest neighbour Ni ions is 140 degrees. The primitive cell of NiF 3is shown in (d). The red and blue arrows represent spin-up and spin-down magnetic moments, respectively. magnetic and five collinear antiferromagnetic structures which are shown in Fig. 1 (c)-(h) . Then we calculate relative energies of six magnetic states with the varia- tion of correlation interaction U. With the increase of correlation interaction U, the NiF 3changes from the fer- romagnetic state to the collinear antiferromagnetic state AFM1 (Fig. 2 (a)). The AFM1 is intralayer ferromag- netism and interlayer antiferromagnetism (Fig. 1 (d)). In previous works, the correlation interaction U was selected as 6.7eV for Ni 3dorbitals [25, 26]. Thus, the magnetic ground state of NiF 3is the AFM1 state, which is consis- tent with previous works[14]. On the other hand, since the bond angle of Ni −F−Ni for the nearest neighbour Ni ions is 140 degrees, the spins of the nearest neigh- bour and next nearest neighbour Ni ions are in antipar- allel and parallel arrangement according to Goodenough- Kanamori rules [27], respectively. This will result in NiF 3 being the collinear antiferromagnetic state AFM1. Thus, the results of theoretical analysis are in agreement with those of theoretical calculation. Indeed, the structure of AFM1 is very simple and the corresponding magnetic primitive cell only contains twomagnetic atoms with opposite spin arrangement which is shown in Fig. 2 (b). From Fig. 2 (b), the two Ni atoms with opposite spin arrangement are surrounded by F-atom octahedrons with different orientations, respec- tively. Thus, the two opposite spin Ni sublattices cannot be connected by a fractional translation. Due to two Ni ions located at space-inversion invariant points, the two opposite spin Ni sublattices cannot be either connected by space-inversion symmetry. However, the two opposite spin Ni sublattices can be connected by C1 2t symmetry. Thus, the NiF 3is an altermagnetic material. The BZ of altermagnetic NiF 3is shown in Fig. 2 (c)and both the high-symmetry and non-high-symmetry lines and points are marked. In order to display the altermagnetic prop- erties more intuitively, we calculate polarization charge density of altermagnetic NiF 3, which is shown in Fig. 2(d). From Fig. 2 (d), the polarization charge densi- ties of two Ni ions with opposite spin arrangement are anisotropic and their orientations are different, result- ing from F-atom octahedrons with different orientations. The anisotropic polarization charge densities can result in k-dependent spin polarization in reciprocal space. More-3 FM AFM1 AFM2 AFM3 AFM4 AFM5 ΓFT L SHSଶmSଶHଶ(a) (c) (d)(b) abc FIG. 2. The magnetic ground state of NiF 3and the cor- responding properties. (a)Relative energies of six differ- ent magnetic states with the variation of correlation inter- action U. (b)and(c)are the magnetic primitive cell of NiF 3 and the corresponding Brillouin zone, respectively. The red and blue arrows represent spin-up and spin-down magnetic moments, respectively. The high-symmetry and non-high- symmetry lines and points are marked in the BZ. (d)The anisotropic polarization charge densities. The red and blue represent spin-up and spin-down polarization charge density, respectively. over, according to different spin group symmetries, the k-dependent spin polarization can form d-wave, g-wave, ori-wave magnetism [12]. Without SOC, the nontrivial elementary spin sym- metry operations in altermagnetic NiF 3have{E||C3z}, {C⊥ 2||M1t},{E||I}, and {C⊥ 2T||T}. The spin symme- tries{C⊥ 2||M1t},{T||TM 1t}, and {E||C3z}make alter- magnetic NiF 3being an i-wave magnetic material, as shown in Fig. 3 (a). Moreover, the spins of bands are opposite along non-high-symmetry S 2−Γ and Γ −m(S 2) directions, reflecting features of i-wave magnetism (Fig. 3(b)). In order to well understand the electronic properties, we also calculate the electronic band structures of alter- magnetic NiF 3along the high-symmetry directions. Ig- noring SOC, the NiF 3is an altermagnetic metal. There are four bands crossing the Fermi level due to spin de- generacy on the high-symmetry directions (Fig. 3 (c)). Especially, these four bands are degenerate on the Γ −T axis. In fact, any kpoint on the Γ −T axis has nontriv- ial elementary spin symmetry operations {E||C3z}and {C⊥ 2||M1t}. And the spin symmetry {E||C3z}has one one-dimensional irreducible real representation and two one-dimensional irreducible complex representations. Al- though the time-reversal symmetry is broken, altermag- (d)Cଷ CଶୄMଵt TT Mଵt(c) (b)(a) Ni-3d F-2p ΓTHଶ|HLΓS |SଶFΓ-0.50-0.250.000.250.50Energy (eV) Ni-3d F-2p ΓTHଶ|HLΓS|SଶFΓ-2-1012Energy (eV) Γ Sଶ mSଶ-0.50-0.250.000.250.50Energy (eV)FIG. 3. Schematic diagram of the i-wave magnetism and electronic band structures of altermagnetic NiF 3.(a) Schematic diagram of the i-wave magnetism. The red and blue parts represent spin up and down, respectively. (b)The electronic band structure without SOC along the non-high- symmetry directions. The red and blue lines represent spin- up and spin-down bands, respectively. (c)and(d)are the electronic band structures without and with SOC along the high-symmetry directions. netic materials have equivalent time-reversal spin sym- metry {C⊥ 2T||T}. The spin symmetry {C⊥ 2T||T}will result in two one-dimensional irreducible complex repre- sentations to form a Kramers degeneracy. Meanwhile, the spin symmetry {C⊥ 2||M1t}protects the spin degen- eracy. Therefore, there is one four-dimensional and one two-dimensional irreducible representations on the Γ −T axis. The quadruple degenerate band crossing the Fermi level is thus protected by the spin group symmetry. Fur- thermore, the orbital weight analysis shows that these four bands are contributed by both the 3dorbitals of Ni and the porbitals of F (Fig. 3 (c)). As is known to all, the F atom has the strongest electronegativity among all chemical elements, but the 2porbitals of F do not fully acquire the 3d-orbital electrons of Ni, which is very in- teresting. In our calculations, the number of valence electrons of NiF 3is 74, which makes the quadruple band only half- filled. This is the reason why the porbitals of F do not fully acquire the d-orbital electrons of Ni. When SOC is included, the spin group symmetry breaks down to mag- netic group symmetry. The reduction of symmetry will result in the quadruple band to split into multiple bands. Since the F atom has the strongest electronegativity, the 2porbitals of F will completely acquire the 3d-orbital electrons of Ni. This will result in altermagnetic NiF 3to transform from metal phase to insulator phase. In order to prove our theoretical analysis, we calculate the elec- tronic band structure of altermagnetic NiF 3with SOC. The calculation of the easy magnetization axis and sym- metry analysis based on magnetic point group are shown in Supplementary Material[28]. Just like our theoretical analysis, the 2porbitals of F indeed fully acquire the4 3d-orbital electrons of Ni and altermagnetic NiF 3trans- forms into an insulator with a bandgap of 2 .31eV (Fig. 3(d)). In general, the SOC strength of Ni is in the order of 10meV, so the SOC strength of altermagnetic NiF 3is two orders of magnitude higher than that of Ni. Thus, the SOC effect of altermagnetic NiF 3is extremely strong. On the other hand, we also examine the effect of cor- relation interaction in altermagnetic NiF 3. We calcu- late the electronic band structures of altermagnetic NiF 3 along the high-symmetry directions without SOC under correlation interaction U = 3 ,5,7eV, which are shown in Fig. 4 (a),(b), and (c), respectively. From Fig. 4 (a), (b)and(c), the correlation interaction has a slight ef- fect on the band structure around the Fermi level with- out SOC, due to the constraints of spin symmetry and electron occupancy being 74. When including SOC, al- termagnetic NiF 3transforms from a metal phase to an insulator phase under different correlation interaction U. Moreover, the bandgap of altermagnetic NiF 3in- creases linearly with the correlation interaction U. Thus, the correlation interaction can substantially enhance the bandgap opened by the SOC of altermagnetic materials. (a) (c) (b)(d)ΓTL ΓFΓ-0.50-0.250.000.250.50Energy (eV) HଶHSSଶ ΓTL ΓFΓ-0.50-0.250.000.250.50Energy (eV) HଶHSSଶ ΓTLΓ FΓ-0.50-0.250.000.250.50Energy (eV) HଶHSSଶ3 4 5 6 7 U (eV)1.01.52.02.5Bandgap (eV) FIG. 4. The electronic properties of altermagnetic NiF 3 under different correlation interaction U. (a),(b)and(c) are the electronic band structures along the high-symmetry directions without SOC under correlation interaction U = 3,5,7eV, respectively. (d)The bandgap as a function of cor- relation interaction U under SOC. Now we well understand the reason for the extremely strong SOC effect in altermagnetic NiF 3. A natural ques- tion is whether such a strong SOC effect can be realized in other altermagnetic materials. According to the above analysis, we propose four conditions for realizing such an effective strong SOC in light element altermagnetic ma- terials: First, the spin group of altermagnetic material has high-dimensional (greater than four dimensions) irre- ducible representation (crystal symmetry groups are pre- sented in the Supplementary Material[28]); Second, the band with high-dimensional representation crossing the Fermi level is half-filled by valence electrons; Third, non- metallic elements have strong electronegativity; Fourth,the altermagnetic material has strong electron correla- tion. To verify these four conditions, we also calcu- late the electronic band structures of four i-wave alter- magnetic materials (VF 3, CrF 3, FeF 3and CoF 3), which have the same crystal structure and spin group sym- metry as NiF 3[14]. The calculations show that none of the four altermagnetic materials meets the second con- dition, and the SOC effect is very weak (Detailed calcu- lations and analysis are presented in the Supplementary Material[28]). On the other hand, since high-dimensional irreducible representations can be protected by spin space group in two-dimensional altermagnetic systems, the pro- posed mechanism is also applicable to two-dimensional light element altermagnetic materials, which may be ad- vantage for realizing quantum anomalous Hall effect at high temperatures [10]. The mechanism for enhancing the SOC effect that we propose in altermagnetic materials is different from that in ferromagnetic materials [11]. First, the high- dimensional representation of the symmetry group is 2 or 3 dimensions in ferromagnetism, while in altermagnetism the high-dimension representation is 4 or 6 dimensions, so their symmetry requirements are entirely different. Sec- ond, the band with high-dimensional representation in ferromagnetism comes from dorbitals, while the band with high-dimensional representation in altermagnetism can come from the combination of porbitals and dor- bitals. Third, the enhancement of SOC effect derives from correlation interaction for ferromagnetic materials, but from both correlation interaction and electronegativ- ity of nonmetallic element for altermagnetic materials. Due to one more degree of freedom to enhance the SOC effect, a stronger SOC effect can be achieved in the al- termagnetic materials. Moreover, if electronegativity of nonmetallic element is weak, different topological phases may be realized in altermagnetic materials when includ- ing SOC. On the other hand, the mechanism for enhanc- ing SOC effect in altermagnetic materials can be also generalized to conventional antiferromagnetic materials. Due to the equivalent time-reversal symmetry, more spin groups with conventional antiferromagnetism have high- dimensional irreducible representations. Moreover, con- ventional antiferromagnetic materials are more abundant than altermagnetic materials, thus conventional antifer- romagnetic materials of light elements with strong SOC effect remain to be discovered. Summary. Based on spin symmetry analysis and the first-principles electronic structure calculations, we demonstrate that extremely strong SOC effect can be re- alized in altermagnetic material NiF 3. Then, we propose a mechanism to enhance SOC effect in altermagnetic ma- terials. This mechanism reveals the cooperative effect of crystal symmetry, electron occupation, electronega- tivity, electron correlation, and intrinsic spin-orbit cou- pling. The mechanism can explain not only the extremely strong SOC effect in altermagnetic NiF 3, but also the weak SOC in altermagnetic VF 3, CrF 3, FeF 3, CoF 3. Moreover, the mechanism for enhancing SOC effect can5 be also generalized to two-dimensional altermagnetic ma- terials.ACKNOWLEDGMENTS This work was financially supported by the Na- tional Key R&D Program of China (Grant No. 2019YFA0308603), the National Natural Science Foun- dation of China (Grant No.11934020, No.12204533, No.62206299 and No.12174443) and the Beijing Natural Science Foundation (Grant No.Z200005). Computational resources have been provided by the Physical Laboratory of High Performance Computing at Renmin University of China. [1] M. Z. Hasan and C. L. Kane, Colloquium: Topological insulators, Rev. Mod. Phys. 82, 3045 (2010). [2] X.-L. Qi and S.-C. Zhang, Topological insulators and su- perconductors, Rev. Mod. Phys. 83, 1057 (2011). [3] A. Bansil, H. Lin, and T. Das, Colloquium: Topological band theory, Rev. Mod. Phys. 88, 021004 (2016). [4] N. Nagaosa, J. Sinova, S. Onoda, A. H. MacDonald, and N. P. Ong, Anomalous hall effect, Rev. Mod. Phys. 82, 1539 (2010). [5] J. Sinova, S. O. Valenzuela, J. Wunderlich, C. H. Back, and T. Jungwirth, Spin hall effects, Rev. Mod. Phys. 87, 1213 (2015). [6] J. E. Hirsch, Spin hall effect, Phys. Rev. Lett. 83, 1834 (1999). [7] M. Cinal, D. M. Edwards, and J. Mathon, Magnetocrys- talline anisotropy in ferromagnetic films, Phys. Rev. B 50, 3754 (1994). [8] F. D. M. Haldane, Model for a quantum hall effect with- out landau levels: Condensed-matter realization of the ”parity anomaly”, Phys. Rev. Lett. 61, 2015 (1988). [9] C.-Z. Chang, C.-X. Liu, and A. H. MacDonald, Collo- quium: Quantum anomalous hall effect, Rev. Mod. Phys. 95, 011002 (2023). [10] P.-J. Guo, Z.-X. Liu, and Z.-Y. Lu, Quantum anomalous hall effect in collinear antiferromagnetism, npj Comput. Mater. 9, 70 (2023). [11] J. Li, Q. Yao, L. Wu, Z. Hu, B. Gao, X. Wan, and Q. Liu, Designing light-element materials with large effec- tive spin-orbit coupling, Nat. Commun. 13, 919 (2022). [12] L. ˇSmejkal, J. Sinova, and T. Jungwirth, Beyond conven- tional ferromagnetism and antiferromagnetism: A phase with nonrelativistic spin and crystal rotation symmetry, Phys. Rev. X 12, 031042 (2022). [13] L. ˇSmejkal, J. Sinova, and T. Jungwirth, Emerging re- search landscape of altermagnetism, Phys. Rev. X 12, 040501 (2022). [14] Z.-F. Gao, S. Qu, B. Zeng, Y. Liu, J.-R. Wen, H. Sun, P.-J. Guo, and Z.-Y. Lu, Ai-accelerated discovery of al- termagnetic materials (2023), arXiv:2311.04418 [cond- mat.mtrl-sci]. [15] H.-Y. Ma, M. Hu, N. Li, J. Liu, W. Yao, J.-F. Jia, and J. Liu, Multifunctional antiferromagnetic materials with giant piezomagnetism and noncollinear spin current, Nat. Commun. 12, 2846 (2021). [16] L. ˇSmejkal, A. B. Hellenes, R. Gonz´ alez-Hern´ andez, J. Sinova, and T. Jungwirth, Giant and tunneling mag-netoresistance in unconventional collinear antiferromag- nets with nonrelativistic spin-momentum coupling, Phys. Rev. X 12, 011028 (2022). [17] R. Gonz´ alez-Hern´ andez, L. ˇSmejkal, K. V´ yborn´ y, Y. Ya- hagi, J. Sinova, T. c. v. Jungwirth, and J. ˇZelezn´ y, Efficient electrical spin splitter based on nonrelativis- tic collinear antiferromagnetism, Phys. Rev. Lett. 126, 127701 (2021). [18] H. Bai, L. Han, X. Y. Feng, Y. J. Zhou, R. X. Su, Q. Wang, L. Y. Liao, W. X. Zhu, X. Z. Chen, F. Pan, X. L. Fan, and C. Song, Observation of spin splitting torque in a collinear antiferromagnet RuO 2, Phys. Rev. Lett. 128, 197202 (2022). [19] S. Karube, T. Tanaka, D. Sugawara, N. Kadoguchi, M. Kohda, and J. Nitta, Observation of spin-splitter torque in collinear antiferromagnetic RuO 2, Phys. Rev. Lett. 129, 137201 (2022). [20] D. Zhu, Z.-Y. Zhuang, Z. Wu, and Z. Yan, Topological superconductivity in two-dimensional altermagnetic met- als, Phys. Rev. B 108, 184505 (2023). [21] P.-J. Guo, Y. Gu, Z.-F. Gao, and Z.-Y. Lu, Altermagnetic ferroelectric LiFe 2F6and spin-triplet excitonic insulator phase (2023), arXiv:2312.13911 [cond-mat.mtrl-sci]. [22] L. ˇSmejkal, R. Gonzalez-Hernandez, T. Jungwirth, and J. Sinova, Crystal time-reversal symmetry breaking and spontaneous hall effect in collinear antiferromagnets, Sci. Adv. 6, eaaz8809 (2020). [23] X. Zhou, W. Feng, X. Yang, G.-Y. Guo, and Y. Yao, Crystal chirality magneto-optical effects in collinear an- tiferromagnets, Phys. Rev. B 104, 024401 (2021). [24] X.-Y. Hou, H.-C. Yang, Z.-X. Liu, P.-J. Guo, and Z.-Y. Lu, Large intrinsic anomalous hall effect in both Nb 2FeB 2 and Ta 2FeB 2with collinear antiferromagnetism, Phys. Rev. B 107, L161109 (2023). [25] B. ˇZemva, K. Lutar, L. Chacon, M. Fele-Beuermann, J. A. Allman, C. J. Shen, and N. Bartlett, Thermody- namically unstable fluorides of nickel: NiF 4and NiF 3 syntheses and some properties, J. Am. Chem. Soc. 117, 10025 (1995). [26] S. Mattsson and B. Paulus, Density functional theory calculations of structural, electronic, and magnetic prop- erties of the 3d metal trifluorides MF 3(M = Ti-Ni) in the solid state, J. Comput. Chem. 40, 1190 (2019). [27] J. B. Goodenough, Goodenough-Kanamori rule, Scholar- pedia 3, 7382 (2008). [28] Supplemental material, .
1101.1268v3.Quantum_phase_transitions_in_a_strongly_entangled_spin_orbital_chain__A_field_theoretical_approach.pdf
arXiv:1101.1268v3 [cond-mat.str-el] 31 May 2011Quantum phase transitions in a strongly entangled spin-orb ital chain: A field-theoretical approach Alexander Nersesyan The Abdus Salam International Centre for Theoretical Physi cs, 34100, Trieste, Italy Andronikashvili Institute of Physics, Tamarashvili 6, 017 7, Tbilisi, Georgia Center of Condensed Mater Physics, ITP, Ilia State Universi ty, 0162, Tbilisi, Georgia Gia-Wei Chern and Natalia B. Perkins Department of Physics, University of Wisconsin, Madison, W isconsin 53706, USA Motivated by recent experiments on quasi-1D vanadium oxide s, we study quantum phase transi- tions in a one-dimensional spin-orbital model describing a Haldane chain and a classical Ising chain locally coupled by the relativistic spin-orbit interactio n. By employing a field-theoretical approach, we analyze the topology of the ground-state phase diagram an d identify the nature of the phase transitions. In the strong coupling limit, a long-range N´ e el order of entangled spin and orbital angular momentum appears in the ground state. We find that, de pending on the relative scales of the spin and orbital gaps, the linear chain follows two dis tinct routes to reach the N´ eel state. First, when the orbital exchange is the dominating energy sc ale, a two-stage ordering takes place in which the magnetic transition is followed by melting of th e orbital Ising order; both transitions belong to the two-dimensional Ising universality class. In the opposite limit, the low-energy orbital modes undergo a continuous reordering transition which rep resents a line of Gaussian critical points. On this line the orbital degrees of freedom form a Tomonaga-L uttinger liquid. We argue that the emergence of the Gaussian criticality results from merging of the two Ising transitions in the strong hybridization region where the characteristic spin and orb ital energy scales become comparable. Finally, we show that, due to the spin-orbit coupling, an ext ernal magnetic field acting on the spins can induce an orbital Ising transition. I. INTRODUCTION Over the past decades, one-dimensional spin-orbital models have been a subject of intensive theoretical stud- ies. The interest is to a large extent motivated by exper- imental discovery of unusual magnetic properties in vari- ous quasi-one-dimensional Mott insulators.1,2The inter- dependence of spin and orbital degrees of freedom is usu- ally described by the so-called Kugel-Khomskii Hamil- tonian in which the effective spin exchange constant de- pends on the orbital configuration and vice versa.3,4An- other mechanism of coupling spin and orbital degrees of freedom is the on-site relativistic spin-orbit (SO) interac- tionλL·S, whereListhe orbitalangularmomentumand λis the coupling constant. In compounds with quenched orbital degrees of freedom, the presence of the SO term usuallyleadstothesingle-ionspinanisotropy DS2 zwhere D∼λ2/∆ and ∆ denotes the energy scale of the crystal field which lifts the degenerate orbital states. For systems with residual orbital degeneracy, on the other hand, the effect of the SO term is much less explored compared with the Kugel-Khomskii-type cou- pling. Due to the directional dependence of the orbital wave functions, the SU(2) symmetry of the Heisenberg spin exchange is expected to be broken in the presence of theSOinteraction. Theresultantspinanisotropyislikely to induce a long-range magnetic order in the spin sector. Amoreintriguingquestioniswhathappenstotheorbital sector. To answer this question, one needs to consider the details of the interplay between the orbital exchange and the SO coupling. Here we consider the simplestcase of a two-fold orbital degeneracy per site. Specifi- cally, the two degenerate states could be the dyzanddzx orbitals in a tetragonal crystal field observed in several transition-metal compounds. We introduce pseudospin- 1/2 operators τa(a=x,y,z) to describe the doublet orbital degrees of freedom assuming that τz=±1 corre- spond to the states |yz/an}bracketri}htandi|zx/an}bracketri}ht, respectively. Alterna- tively, one can also realize the double orbital degeneracy in the Mott-insulating phase of a 1D fermionic optical lattice where the eigenvectors of τzrefers to pxandpy orbitals in an anisotropic potential.5,6Restricted to this doublet space, the orbital angular momentum operator L= (0,0,τx). This can be easily seen by noting that the eigenstatesof τxcarryanangularmomentum /an}bracketle{tLz/an}bracketri}ht=±1. The exchange interaction between localized orbital de- grees of freedom is characterized by its highly direc- tional dependence: the interaction energy only depends on whether the relevant orbital is occupied for bonds of a given orientation. This is particularly true for inter- actions dominated by direct exchange mechanism. De- noting the relevant orbital projectors on a given bond as P= (1+τβ)/2, where τβ/2isanappropriatepseudospin- 1/2 operator ( τβbeing a Pauli matrix), the orbital inter- action is thus described by an Ising-type term τβ iτβ j. The well studied orbital compass model and Kitaev model both belong to this category.7,8The quantum nature of these models comes from the fact that different operators τβ, which do not commute with each other, are used for bonds of different types. To avoid unnecessary compli- cations coming from the details of orbital interactions, we assume that there is only one type of bond in our 1D2 system and the orbital interaction is thus governed by a classical Ising Hamiltonian. We incorporate these features into the following toy model of spin-orbital chain ( Js,Jτ>0): H=HS+Hτ+HSτ (1) =JS/summationdisplay nSn·Sn+1+Jτ/summationdisplay nτz nτz n+1+λ/summationdisplay nτx nSz n. Motivated by the recent experimental characterizations of quasi-1D vanadium oxides,14–19here we focus on the case of quantum spin with length S= 1. The above model thusdescribesaHaldanechainlocallycoupled toa classical Ising chain by the SO interaction HSτ. The role of theλ-term is two-fold: firstly it introduces anisotropy tothespin-1subsystem, andsecondlyit endowsquantum dynamics to the otherwise classical Ising chain. Before turning to a detailed study of the phase dia- gram of model (1), we first discuss its connections to real compounds. As mentioned above, the interest in the toy model is partly motivated by the recent experi- mental progresson vanadium oxides which include spinel ZnV2O414–17and quasi-1D CaV 2O4.18,19In both types of vanadates, the two delectrons of V3+ions have a spinS= 1 in accordance with Hund’s rule. In the low- temperature phase of both vanadates, the vanadium site embeddedinaflattenedVO 6octahedronhasatetragonal symmetry. This tetragonal crystal field splits the degen- eratet2gtriplet into a singlet and a doublet. As one of the twodelectrons occupies the lower-energy dxystate, a double orbital degeneracy arises as the second electron could occupy either dzxordyzorbitals. The fact that thedxyorbital is occupied everywherealso contributes to theformationofweaklycoupledquasi-1Dspin-1chainsin these compounds.20On the other hand, the details of the orbital exchange depends on the geometry of the lattice and in the case of vanadium spinel the orbital interaction is of three-dimensional nature. The Ising orbital Hamil- tonian in Eq. (1) thus should be regarded as an effective interaction in the mean-field sense. Nonetheless, the toy model provides a first step towardsunderstanding the es- sentialphysics introducedby the SO coupling. Moreover, many conclusions of this paper can be applied to the case of quasi-1Dcompound CaV 2O4where the vanadium ions form a zigzag chain. It is instructive to first establish regions of stable mas- sive phases. In the decoupling limit, λ→0, our model describes two gapped systems: a quantum spin-1 Heisen- bergchainand aclassicalorbitalIsingchain. Theground state of the spin sector is a disordered quantum spin liquid with a finite spectral gap21∆S, whereas the or- bital ground state is characterized by a classical N´ eel order along the chain: /an}bracketle{tτz n/an}bracketri}ht= (−1)nηz. Quantum ef- fects in the orbital sector induced by the SO coupling play a minor role. Obviously, just because of being gapped, both the spin-liquid phase and the orbital or- dered state are stable as long as λremains small. Con- sider now the opposite limit, λ≫JS,Jτ. In the ze- roth order approximation, the model is dominated by thesingle-ionterm HSτwhose doubly degenerateeigenstates |±/an}bracketri}ht=|Sz=±1/an}bracketri}ht⊗|τx=±1/an}bracketri}htrepresent locally entangled spin and orbital degrees of freedom. Switching on small JSandJτleads to a staggered ordering of the |+/an}bracketri}htand |−/an}bracketri}htstates alongthe chain. Physically, the large- λground state can be viewed as a simultaneous N´ eel ordering of spin and orbital angular momentum characterized by or- der parameters ζandηxsuch that /an}bracketle{tSz n/an}bracketri}ht= (−1)nζand /an}bracketle{tLz n/an}bracketri}ht=/an}bracketle{tτx n/an}bracketri}ht= (−1)nηx. The Ising order parameter ηz vanishes identically in this phase. These observations naturally lead to the following questions. How is the magnetically ordered N´ eel state at large λconnected to the disordered Haldane phase asλ→0 ? What is the scenario for the orbital re- orientation transition ηz→ηx, which is of essentially quantum nature ? In this paper we employ the field- theoretical approach to address these questions. We first note that the one-dimensional model (1) is not exactly integrable. As a consequence, the regime of strong hy- bridization of the spin and orbital excitations, which is the case when Jτ,JSandλare all of the same order, stays beyond the reach of approximate analytical meth- ods. We thus will be mainly dealing with limiting cases Jτ≫JSandJτ≪JS, in which one can integrate out the “fast” variables to obtain an effective action for the “slow” modes. Following this approach, we establish the topology and main features of the ground-state phase diagram in the accessible parts of the parameter space of the model. We were able to unambiguously identify the universality classes of quantum criticalities separat- ing different massive phases. Using plausible arguments we comment on some features of the model in the regime of strong spin-orbital hybridization. We demonstrate that the aforementioned reorientation transition ηz→ηxcan be realized in one of two possi- ble ways. In the limit of large Jτ, we find a sequence of two quantum Ising transitions and an intermediate massive phase, sandwiched between these critical lines, in which both ηzandηxare nonzero. This is consis- tent with the recent findings22based on DMRG calcu- lations and some analytical estimations. In the oppo- site limit, when the Haldane gap ∆ Sis the largest en- ergy scale, integrating out the spin excitations yields an effective lowest-energy action for the orbital degrees of freedom, which shows that the ηz→ηxcrossover takes place as a single Gaussian quantum criticality. At this critical point, the orbital degrees of freedom display an extremelyquantumbehaviour: they aregaplessand form a Tomonaga-Luttinger liquid. This is the main result of this paper. We bring about arguments suggesting that the emergence of the Gaussian critical line is the result of merging of the two Ising criticalities in the region of strong spin-orbital hybridization. Any field-theoretical treatment of the model (1) must be based on a properly chosen contiuum description of the spin-1 antiferromagnetic Heisenberg chain. Its prop- erties have been thoroughly studied, both analytically and numerically (see for a recent review Ref. 23). In3 what follows, the spin sector of the model (1) will be treated within the O(3)-symmetric Majorana field the- ory, proposed by Tsvelik:24 HM=/summationdisplay a=1,2,3/bracketleftbiggiv 2(ξa L∂xξa L−ξa R∂xξa R)−imξa Rξa L/bracketrightbigg +Hint. (2) Hereξa R,L(x) is a degenerate triplet of real (Majorana) Fermi fields with a mass m, the indices RandLlabel the chirality of the particles, and Hint=1 2g/summationdisplay a(ξa Rξa L)2 is a weak four-fermion interaction which can be treated perturbatively. The continuum theory (2) adequately de- scribesthelow-energypropertiesofthegeneralizedspin-1 bilinear-biquadratic chain HS→¯HS=JS/summationdisplay n/bracketleftBig Sn·Sn+1−β(Sn·Sn+1)2/bracketrightBig .(3) in the vicinityofthe criticalpoint β= 1.25Thisquantum criticality belongs to the universality class of the SU(2) 2 Wess-Zumino-Novikov-Witten(WZNW)modelwithcen- tral charge c= 3/2. At small deviations from criticality the Majoranamass m∼JS|β−1|determines the magnitude of the triplet gap, ∆ S=|m| ≪JS. The theory of a massive triplet of Majorana fermions is equivalent to a system of three degenerate noncritical 2D Ising models, with m∼(T−Tc)/Tc. This is one of the most appealing fea- tures of the theory because the most strongly fluctuating physical fields of the S= 1 chain, namely the staggered magnetization and dimerization operators, have a simple local representation in terms of the Ising order and dis- order parameters.24,26,27It is this fact that greatly sim- plifies the analysis of the spin-orbital model (1). While the correspondence between the models (2) and (3) is well justified at |β−1| ≪1, it is believed that the Ma- jorana model (2) captures generic properties of the Hal- danespin-liquidphaseofthe spin-1chain, eventhoughat large deviations from criticality ( |β−1| ∼1,∆S∼Js) all parameters of the model should be treated as phe- nomenological. The remainderofthe paperis organizedasfollows. We start our discussion with Sec. II which contains a brief summary of known facts about the Majorana model24 that will be used in the rest of the paper. In Sec. III we consider the limit Jτ/∆S≫1 and by integrating out the ‘fast’ orbital modes, show that on increasing the SO coupling λthe system undergoes a sequence of two consecutive quantum Ising transitions in the spin and or- bital sectors, respectively. In section IV we analyze the opposite limiting case, Jτ/∆S≪1, and, by integrating over the ‘fast’ spin modes, show that there exists a sin- gle Gaussian transition in the orbital sector accompaniedNeel ζ=0/ Orbital Ising orderIIII IIPath 1Orbital Ising order xτxS Path 2 ζ=0 Haldane spin liquid xzη =0, η =0//η =0, η =0z x xz//η =0, η =0 FIG. 1: Schematic phase diagram of the model on the ( xS, xτ)-plane, where xS= ∆S/λandxτ=Jτ/λ. by a Neel ordering of the spins. We then conjecture on the topology of the ground-state phase diagram of the model. In Sec. V we show that spin-orbital hybridiza- tion effects near the orbital Gaussian transition lead to the appearance of a non-zero spectral weight well below the Haldane gap which can be detected by inelastic neu- tron scattering experiments and NMR measurements. In Sec. VI we comment on the role of an external magnetic field. We show that, through the SO interaction, a suf- ficiently strong magnetic field affects the orbital degrees of freedom and can lead to a quantum Ising transition in the orbital sector. Sec. VII contains a summary of the obtained results and conclusions. The paper has two appendices containing certain technical details. II. SOME FACTS ABOUT MAJORANA THEORY OF SPIN-1 CHAIN In this Section, we provide some details about the O(3)-symmetric Majorana field theory,24Eq. (2), which represents the continuum limit of the biquadratic spin-1 model (3) at |β−1| ≪1. In the continuum description, the local spin density of the spin model (3) has contributions from the low-energy modes centered in momentum space at q= 0 and q=π: S(x) =IR(x)+IL(x)+(−1)x/a0N(x) (4) The smooth part of the local magnetization, I=IR+ IL, is a sum of the level-2 chiral vector currents. The SU(2)2Kac-Moody algebra of these currents is faithfully reproduced in terms of a triplet of massless Majorana fields28ξ= (ξ1,ξ2,ξ3): Iν=−i 2(ξν×ξν),(ν=R,L) (5)4 This fact is not surprising because, as already men- tioned, the central charge of the SU(2) 2WZNW theory isc= 3/2, whereas that of the theory of a massless Ma- jorana fermion (equivalently, critical 2D Ising model) is c= 1/2. At smalldeviationsfromcriticality( |β−1| ≪1) thefermionsacquireamass. Stronglyfluctuatingfieldsof the spin-1 chain, the staggered magnetization N(x) and dimerization operator ǫ(x) = (−1)nSn·Sn+1, are nonlo- cal in terms of the Majorana fields but admit a simple representation in terms of the order, σ, and disorder, µ, operators of the related noncritical Ising models: N∼(1/α)(σ1µ2µ3, µ1σ2µ3, µ1µ2σ3), ǫ∼(1/α)σ1σ2σ3, (6) whereα∼a0is a short-distance cutoff of the contin- uum theory. These expressions together with their duals (i.e. their counterparts obtained by the duality trans- formation in all Ising copies, σa↔µa) determine the vector and scalar parts of the WZNW 2 ×2 matrix field ˆgwhich is a primary scalar field with scaling dimension 3/8. It has been demonstrated in Ref. 28 that using the representation (6) and the short-distance operator product expansions for the Ising fields, one correctly re- produces all fusion rules of the SU(2) 2WZNW model. An equivalent way to make sure that this is indeed the case is to consider the four-Majorana representation of the weakly coupled spin-1/2 Heisenberg ladder26,27and take the limit of a infinite singlet Majorana mass to map the low-energy sector of the model on the O(3) theory (2). In the spin-liquid phase of the spin chain (3), which is the case β <1, the Majorana mass mis positive, implying that the degenerate triplet of2D Isingmodels is inadisorderedphase: /an}bracketle{tσa/an}bracketri}ht= 0,/an}bracketle{tµa/an}bracketri}ht /ne}ationslash= 0 (a= 1,2,3). In particular, this implies that the O(3) symmetry remains unbroken, /an}bracketle{tN/an}bracketri}ht= 0, and the ground state of the system is not spontaneously dimerized, /an}bracketle{tǫ/an}bracketri}ht= 0. The representation (6) proves to be very useful for cal- culatingthe dynamicalspincorrelationfunctionsbecause the asymptotics of the Ising correlators /an}bracketle{tσ(x,τ)σ(0,0)/an}bracketri}ht and/an}bracketle{tµ(x,τ)µ(0,0)/an}bracketri}htarewellknownbothatcriticalityand in a noncritical regime. In the disordered phase ( m >0), the leading asymptotics of the Ising correlators are: /an}bracketle{tµ(r)µ(0)/an}bracketri}ht ∼(a/ξS)1/4/bracketleftBig 1+O(e−2r/ξS)/bracketrightBig , /an}bracketle{tσ(r)σ(0)/an}bracketri}ht ∼(a/ξS)1/4/radicalbig ξS/r e−r/ξS(7) whereξS=v/mis the correlation length, and r=√ x2+v2τ2. (By duality, in the ordered phase ( m <0) the asymptotics of the correlators in (7) must be inter- changed.) Correspondingly, the dynamical correlation function /an}bracketle{tN(r)N(0)/an}bracketri}ht ∼(a/ξS)3/4/radicalbig ξS/r e−r/ξS.(8) Its Fourier transform at q∼πand small ωdescribes a coherent excitation – a triplet magnon with the mass gapm: ℑm χ(q,ω)∼m |ω|δ/parenleftBig ω−/radicalbig (q−π)2v2+m2/parenrightBig .(9) Since the single-ion anisotropy Hanis=D/summationtext n(Sz n)2 lowers the original O(3) symmetry down to O(2) ×Z2, one expects24that in the continuum theory it will induce anisotropy in the Majorana masses m1=m2/ne}ationslash=m3, as well as in the coupling constants parametrizing the four-fermion interaction: Hint→1 2/summationdisplay a/negationslash=bgab(ξa Rξa L)/parenleftbig ξb Rξb L/parenrightbig , g13=g23/ne}ationslash=g12. This can be checked by using the correspondence (4) and short-distance operator product expansions (OPE) for the physical fields. There will also appear anisotropy in the velocities, v1=v2/ne}ationslash=v3, but we will systematically neglect this effect. Thus, we have Hanis=/integraltext dxHanis, with Hanis=Dα/integraldisplay dx/bracketleftbig I3(x)I3(x+α)+N3(x)N3(x+α)/bracketrightbig ,(10) whereα∼ais a short-distance cutoff of the continuum theory. Using (5) and keeping only the Lorentz invariant terms (i.e. neglecting renormalization of the velocities) we can replace ( I3)2by 2I3 RI3 L. To treat the second term in the r.h.s. of (10), we need OPEs for the products of Ising operators:29 σ(z,¯z)σ(w,¯w) =1√ 2/parenleftbiggα |z−w|/parenrightbigg1/4/bracketleftbig 1−π|z−w|ε(w,¯w)/bracketrightbig ,(11) µ(z,¯z)µ(w,¯w) =1√ 2/parenleftbiggα |z−w|/parenrightbigg1/4/bracketleftbig 1+π|z−w|ε(w,¯w)/bracketrightbig .(12) Hereε=iξRξLis the energy density (mass bilinear) of the Ising model, z=vτ+ ixandw=vτ′+ ix′are two-dimensional complex coordinates, ¯ zand ¯ware their conjugates. From the above OPEs it follows that N3(x)N3(x+α) = i(π/α)/parenleftbig ξ1 Rξ1 L+ξ2 Rξ2 L−ξ3 Rξ3 L/parenrightbig −(π2C)[(ξ1 Rξ1 L)(ξ2 Rξ2 L)−(ξ1 Rξ1 L)(ξ3 Rξ3 L)−(ξ2 Rξ2 L)(ξ3 Rξ3 L)], whereC∼1 is a nonuniversal constant. As a result, Hanis=−i/summationdisplay a=1,2,3δmaξa Rξa L+1 2/summationdisplay a/negationslash=bδgij(ξa Rξa L)/parenleftbig ξb Rξb L/parenrightbig ,(13) where δm1=δm2=−δm3=−(πC)D (14)5 are corrections to the single-fermion masses, and δg12= (2−π2C)Dα, δg 13=δg23=π2CDαare cou- pling constants of the induced interaction between the fermions. Smallness of the Majorana masses ( |m|α/v≪ 1) implies that the additional mass renormalizations caused by the interaction in (13) are relatively small, m(Dα/v)ln(v/|m|α)≪D, so that the main effect of the single-ion anisotropy is the additive renormalization of the fermionic masses, ma=m+δma, withδmagiven by Eq.(14). The cases D >0 andD <0 correspond to an easy- plane and easy-axis anisotropy, respectively. The spin anisotropy (18) induced by the spin-orbit coupling is of the easy-axis type. At D <0 the singlet Majorana fermion, ξ3, is the lightest, m3< m1=m2. Increas- ing anisotropy drives the system towards an Ising crit- icality at D=−D∗, where m3= 0. At D <−D∗ the system occurs in a new phase where the Ising dou- blet remains disordered while the singlet Ising system becomes ordered. It then immediately follows from the representation (6) that the new phase is characterized by a N´ eel long-range order with /an}bracketle{tN3/an}bracketri}ht /ne}ationslash= 0. Transverse spin fluctuations, as well as fluctuations of dimerization, are incoherent in this phase. III. TWO ISING TRANSITIONS IN THE ∆S≪JτLIMIT Nowwe turn to ourmodel (1). Let us considerthe case when, in the absence of spin-orbit coupling, the orbital gapisthelargest: Jτ≫Js. Theorbitalpseudospinsthen represent the ‘fast’ subsystem and can be integrated out. Assuming that λ≪Jτ, we treat the spin-orbit coupling perturbatively. In this case, the zero order Hamiltonian H0=HS+Hτdescribes decoupled spin and orbital sys- tems, while the spin-orbit interaction HSτdenotes per- turbation. Defining the interaction representation for all operators according to A(τ) =eτH0Ae−τH0(hereτde- notes imaginary time), the interaction term in the Eu- clidian action is given by SSτ=λ/summationdisplay n/integraldisplay dτ τx n(τ)Sz n(τ). (15) The first nonvanishing correction to the effective action in the spin sector is of the second order in λ: ∆Ss=−λ2 2/summationdisplay nm/integraldisplay dτ1dτ2/angbracketleftbig τx n(τ1)τx m(τ2)/angbracketrightbig τSz n(τ1)Sz m(τ2). (16) Averaging in the right-hand side of (16) goes over config- urations of the classical Ising chain Hτ. The correlation function /an}bracketle{tτx n(τ1)τx m(τ2)/an}bracketri}htτis calculated in Appendix A. It is spatially ultralocal (because there are no propagating excitations in the classical Ising model) and rapidly de- caying at the characteristic time ∼1/Jτ, which is muchshorter than the spin correlation time ∼1/∆0: /an}bracketle{tτx n(τ1)τx m(τ2)/an}bracketri}htτ=δnmexp(−4Jτ|τ1−τ2|).(17) Passing to new variables, τ= (τ1+τ2)/2 andρ=τ1−τ2, and integrating over ρyields a correction to the effec- tive spin action which has the form of a single-ion spin anisotropy. Thus in the second order in λ, the spin Hamiltonian acquires an additional term Hani=−λ2 4Jτ/summationdisplay n(Sz n)2. (18) The anisotropy splits the Majorana triplet into a doublet (ξ1,ξ2) and singlet ( ξ3), with masses m1=m2=m+πCλ2 4Jτ, m3=m−πCλ2 4Jτ,(19) whereC∼1 is a nonuniversal positive constant. The anisotropy is of the easy-axis type, so that the singlet mode has a smaller mass gap. As long as all the masses maremain positive, the sys- tem maintains the properties of an anisotropic Haldane’s spin-liquid. The dynamical spin susceptibilities calcu- lated at small ωandq∼π(see Sec. II), ℑm χxx(q,ω) =ℑm χyy(q,ω) (20) ∼m1 |ω|δ/parenleftbigg ω−/radicalBig (q−π)2v2+m2 1/parenrightbigg , ℑm χzz(q,ω)∼m3 |ω|δ/parenleftbigg ω−/radicalBig (q−π)2v2+m2 3/parenrightbigg , indicate the existence of the Sz=±1 andSz= 0 optical magnons with mass gaps m1andm3, respectively. In- creasing the spin-orbital coupling leads eventually to an Isingcriticalityat λ=λc1= 2/radicalbig Jτm/πC, wherem3= 0. Atm3<0thesystemoccursinalong-rangeorderedN´ eel phase with staggered magnetization /an}bracketle{tSz n/an}bracketri}ht= (−1)nζ(λ), in whichthe Z2-symmetryofmodel (18) is spontaneously broken. Using the Ising-model representation (6) of the staggered magnetization of the spin-1 chain, we find that at 0< λ−λc1≪λc1the order parameter ζ(λ) follows a power-law increase: ζ(λ)∼/parenleftbiggλ−λc1 λc1/parenrightbigg1/8 . (21) The transverse spin fluctuations become incoherent in this phase. The situation here is entirely similar to that in the spontaneously dimerized massive phase of a two- chain spin-1/2 ladder27,30, where the dimerization kinks make spin fluctuations incoherent. In the present case, the spontaneouslybroken Z2symmetry ofthe Neel phase leadstotheexistenceofpairsofmassivetopologicalkinks contributing to a broad continuum with a threshold at ω=m1+|m3|(the details of calculation can be found in Ref.27): ℑm χxx(q,ω) (22) ∼1/radicalbig m1|m3|θ(ω2−(q−π)2v2−(m1+|m3|)2)/radicalbig ω2−(q−π)2v2−(m1+|m3|)2.6 In the N´ eel phase, the orbital sector acquires quantum dynamicsbecauseantiferromagneticorderingofthe spins generates an effective transverse magnetic field which transforms the classical Ising model Hτto a quantum Ising chain. At λ > λ c1the spin-orbit term takes the form HSτ=−h/summationdisplay n(−1)nτx n+H′ Sτ, (23) whereh=λζ(λ) andH′ Sτ=−λ/summationtext n(Sz n−/an}bracketle{tSz n/an}bracketri}ht)τx nac- counts for fluctuations. Since both the orbital and spin sectors are gapped, the main effect of this term is a renormalization of the mass gaps and group velocities. The transverse field hgives rise to quantum fluctuations which decrease the classical value of ηzand, at the same time, lead to a staggered ordering of the orbital pseu- dospins in the transverse direction. Since the orbital sec- tor has a finite susceptibility with respect to a transverse staggered field, in the right vicinity of the critical point ηxfollows the same power-law increase as ζbut with a smaller amplitude: ηx∼/parenleftbiggh Jτ/parenrightbigg ∼/radicalbigg ∆S Jτ/parenleftbiggλ−λc1 λc1/parenrightbigg1/8 .(24) This result is in a good agreement with previously obtained numerical results for order parameters (See Fig. 4(a) in Ref. 22). Performing an inhomogeneous π-rotation of the pseu- dospins around the y-axis,τx,z n→(−1)nτx,z n,τy n→τy n, we find that at λ > λc1the effective model in the orbital sectorreducesto a ferromagneticIsing chainin auniform transverse (pseudo)magnetic field: Hτ;eff=−Jτ/summationdisplay nτz nτz n+1−h/summationdisplay nτx n.(25) Notice that the restriction λ≪Jτ, which was imposed in the derivation of the effective Hamiltonian in the spin sector, now can be released because the spin sector is assumed to be in the N´ eel phase. Ath=Jτ, i.e. at λ=λc2whereλc2satisfies the equation λc2ζ(λc2) =Jτ, (26) the model (25) undergoes a 2D Ising transition27,31to a massive disordered phase with /an}bracketle{tτz n/an}bracketri}ht= 0. This quantum critical point can be reached when λis further increased in the region λ > λc1. It is clear from (26) that λc2is of theorderoforgreaterthan Jτ. Itisreasonabletoassume that for such values of λthe N´ eel magnetization is close to its nominal value, ζ∼1, implying that λc2∼Jτ. We see that the two Ising transitions are well separated: λc2/λc1∼(Jτ/∆S)1/2≫1. (27) Thus, in the limit Jτ≫∆S, the ground-state phase diagram of the model (1) consists of three gapped phasesFIG. 2: Schematic diagram of order parameters as functions of the SO coupling constant λ. (a) Two Ising transitions in theJτ≫∆Slimit. (b) A single Gaussian transition in the ∆S≫Jτlimit. These two scenarios correspond to path-1 and path-2 in the phase diagram (Fig. 1), respectively. separated by two Ising criticalities, one in the spin sector (λ=λc1) and the other in the orbital sector ( λ=λc2). At 0< λ < λ c1the spin sector represents an anisotropic spin-liquid while in the orbital sector there is a N´ eel-like ordering of the pseudospins: ( −1)n/an}bracketle{tτz n/an}bracketri}ht ≡ηz(λ)/ne}ationslash= 0. Atλc1< λ < λ c2the orbital degrees of freedom reveal their quantum nature: the onset of the spin N´ eel order (ζ/ne}ationslash= 0) is accompanied by the emergence of the trans- versecomponent of the staggered pseudospin density: (−1)n/an}bracketle{tτx n/an}bracketri}ht ≡ηx(λ)/ne}ationslash= 0. Upon increasing λ, the stag- gered orbital order parameter ηundergoes a continuous rotationfrom the z-directionto x-direction. At λ=λc2a quantum Ising transition takes place in the orbital sector whereηzvanishes. At λ > λ c2both sectors are long- range ordered, with order parameters ζ,ηx/ne}ationslash= 0. The de- pendenceoforderparameterson λisschematicallyshown in Fig. 2(a); this picture is in full qualitative agreement with the results of the recent numerical studies.22 The crossover between the small and large λlimits studied in this sectioncorrespondsto path 1 onthe phase diagram shown in Fig. 1. The path is located in the re- gionJτ≫∆S. Starting from the massive phase I and moving along this path we first observe the spin-Ising transition (I →II) to the N´ eel phase. Long-range or- dering of the spins induces quantum reconstruction of the initialy classical orbital sector (i.e. generation of a nonzeroηx). Theorbital-Ising transition (II →III) takes place inside the spin N´ eel phase. Of course, feedback ef- fects (that is, orbitaffecting spin) become inreasinglyim- portant upon deviating from the critical curve ∆ SJτ∼1 into phases II and III, especially in the vicinity of the orbital transition where the spin-orbit coupling is very strong,λ∼Jτ. In this region the behavior of the spin degrees of freedom is not expected to follow that of an isolated anisotropic spin-1 chain in the N´ eel phase since the effect of an “explicit” staggered magnetic field ∼ληx becomes important. We will see a pattern of such be- havior in the opposite limit of “heavy” spins, which is discussed in the next section.7 IV. GAUSSIAN CRITICALITY AT J τ≪∆S In this section we turn to the opposite limiting case: ∆S≫Jτ. Now the spin degrees of freedom constitute the “fast” subsystem and can be integrated out to gen- erate an effective action in the orbital sector. We will showthat, in this regime, the intermediate massivephase wherethe orbitalorderparameter ηundergoesacontinu- ousrotationfrom η= (0,0,ηz) toη= (ηx,0,0)nolonger exists. Going along path 2, Fig. 1, which is located in the region ∆ S≫Jτ, we find that the two massive phases, I and III, are separated by a single Gaussian critical line characterized by central charge c= 1. On this line the vectorηvanishes, the orbital degrees of freedom become gapless and represent a spinless Tomonaga-Luttinger liq- uid characterized by power-law orbital correlations. Atλ= 0 the spin-1 subsystem represents a disordered, isotropic spin liquid. Therefore the first nonzero correc- tiontothelow-energyeffectiveactionintheorbitalsector appears in the second order in λ: ∆S(2) τ=−1 6/an}bracketle{tS2 Sτ/an}bracketri}htS (28) =−1 2λ2/summationdisplay nm/integraldisplay dτ1/integraldisplay dτ2/an}bracketle{tSn(τ1)Sm(τ2)/an}bracketri}htSτx n(τ1)τx m(τ2), where/an}bracketle{t···/an}bracketri}htSmeans averaging over the massive spin de- grees of freedom. According to the decomposition of the spin density, Eq. (4), the correlation function in (29) has the structure: /an}bracketle{tSl(τ)S0(0)/an}bracketri}ht= (−1)lf1(r/ξS)+f2(r/ξS).(29) Hereξs=vs/∆Sis the spin correlation length and r= (vsτ,x) is the Euclidian two-dimensional radius- vector.f1andf2aresmooth functions with the following asymptotic behaviour27 f1(x) =C1x−1/2e−x, f2(x) =C2x−1e−2x(x≫1),(30) whereC1andC2are nonuniversal constants. DMRG calculations show32thatC2≪C1; for this reason the contribution of the smooth part of the spin correlation function can be neglected in (29). Integrating over the relative time τ−=τ1−τ2we find that the spin-orbit coupling generates a pseudospin xx- exchange with the following structure: H′ τ=/summationdisplay n/summationdisplay l≥1(−1)l+1J′ τ(l)τx nτx n+l (31) Here the exchange couplings exponentially decay with the separation l,J′ τ(l)∼(λ2/∆S)exp(−la0/ξS), so the summation in (31) actually extends up to l∼ξS/a0. In the Heisenberg model ξSis of the order of a few lattice spacings, so for a qualitative understanding it would be sufficient to consider the l= 1 term as the leading one and treat the l= 2 term as a correction. Making a π/2rotation in the pseudospin space, τz n→τy n,τy n→ −τz n, we passto the conventionalnotationsand write down the effective Hamiltonian for the orbital degrees of freedom as a perturbed XY spin-1/2 chain: Heff τ=/summationdisplay n/parenleftbig Jxτx nτx n+1+Jyτy nτy n+1/parenrightbig +H′ τ.(32) where H′ τ=−J′ x/summationdisplay nτx nτx n+2+···. (33) HereJy=Jτ,Jx=J′ τ(1)>0 andJ′ x=J′ τ(2)>0. By order of magnitude J′ x< Jx∼λ2/∆S. In the absence of the perturbation H′ τ, the model (32) represents a spin-1/2 XY chain which for any nonzero anisotropy in the basal plane ( Jx/ne}ationslash=Jy) has a N´ eel long- range order in the ground state and a massive excitation spectrum. This follows from the Jordan-Wigner trans- formation τz n= 2a† nan−1, τ+ n=τx n+iτy n= 2a† neiπ/summationtext j<na† jaj(34) which maps the XY chain onto a model of complex spin- less fermions with a Cooper pairing:33 Heff τ= (Jx+Jy)/summationdisplay n/parenleftbig a† nan+1+h.c./parenrightbig + (Jx−Jy)/summationdisplay n/parenleftBig a† na† n+1+h.c./parenrightBig .(35) By increasing λ(equivalently, decreasing Jτ) the model (35) can be driven to a XX quantum critical point, Jx= Jy(1), i.e. λ=λc∼√Jτ∆S, where the the system acquires a continuous U(1) symmetry. At this point the Jordan-Wignerfermionsbecomemasslessand thesystem undergoes a continuous quantum transition. The transition is associated with reorientation of the pseudospins. Away from the Gaussian criticality the effective orbital Hamiltoian is invariant under Z2×Z2 transformations: τx n→ −τx n,τz n→ −τz n. In massive phasesthis symmetry is spontaneouslybroken. Making a back rotation from τytoτzwe conclude that at Jy> Jx (λ < λ c)ηz/ne}ationslash= 0,ηx= 0, while at Jy< Jx(λ > λ c) ηz= 0,ηx/ne}ationslash= 0. Both ηzandηxvanish at the critical point, so contrary to the case Jτ≫∆S, here there is no region of their coexistence. The passage to the continuum limit for the model (32) based on Abelian bosonization is discussed in Appendix B. There we show that the perturbation H′ τadds a marginal four-fermion interaction g=J′ x(2)/πv≪1 to the free-fermion model (B3). In the spin-chain lan- guage, this is equivalent to adding a weak ferromag- neticzz-coupling. In the limit of weak XY anisotropy, |λ−λc|/λc≪1, the low-energy properties of the orbital sector are described by a quantum sine-Gordon model (all notations are explained in Appendix B) H=u 2/bracketleftbigg KΠ2+1 K(∂xΦ)2/bracketrightbigg +2γ παcos√ 4πΘ,(36)8 where γ∼Jτ/parenleftbiggλ−λc λc/parenrightbigg , K= 1+2g+O(g2).(37) The U(1) criticality is reached at λ=λcwhere, due to a finite value of g, the orbital degrees of freedom represent a Tomonaga-Luttingerliquid. Close to the criticality, the spectral gap in the orbital sector scales as the renormal- ized mass of the sine-Gordon model (36): Morb∼/vextendsingle/vextendsingle/vextendsingleλ−λc λc/vextendsingle/vextendsingle/vextendsingleK 2K−1. (38) Strongly fluctuating physical fields acquire coupling dependentscalingdimensions. Inparticular,accordingto the bosonization rules,27the staggered pseudospin den- sities are expressed in terms of the vertex operators, (−1)nτx n≡nx(x)∼sin√πΘ(x), (−1)nτz n≡nz(x)∼cos√πΘ(x),(39) both with scaling dimension d= 1/4K. This anomalous dimension determines the power-lawbehaviourof the av- erage staggered densities close to the criticality: ηz(λ)∼(λc−λ)1/4K, λ < λ c ηx(λ)∼(λ−λc)1/4K, λ > λ c. (40) A finite staggered pseudospin magnetization ηxat λ > λ cgenerates an effective external staggered mag- netic field in the spin sector: HS→¯H=HS+H′ S, H′ S=−hS/summationdisplay n(−1)nSz n,(41) wherehS=−ληx. The spectrum of the Hamiltonian ¯H is always massive. This can be easily understood within the Majorana model (2). According to (6), in the con- tinuum limit, the sign-alternating component of the spin magnetization, N3∼(−1)nSz n, canbeexpressedinterms of the order and disorder fields of the degenerate triplet of 2D disordered Ising models: N3∼µ1µ2σ3. In the leading order, the magnetic interaction H′ Sgives rise to an effective magnetic field h3=hS/an}bracketle{tµ1µ2/an}bracketri}htapplied to the third Ising system: h3σ3. The latter always stays off- critical. Since in the Haldane phase the spin correlations are short-ranged, close to the transition point the induced staggered magnetization ζcan be estimated using linear response theory. Therefore, at 0 < λ−λc≪λc,ζfollows the same power-law increase as that of ηxbut with a smaller amplitude: ζ∼hS ∆S∼/parenleftbiggJτ ∆S/parenrightbigg1/2/parenleftbiggλ−λc λc/parenrightbigg1/4K (42) So, in the part of the phase C, Fig. 1, where ∆ S≫Jτ, theηx-orbital order, being the result of a spontaneousbreakdown of a Z2symmetry τx n→ −τx n, acts as an ef- fective staggered magnetic field applied to the spins and inducestheir N´ eel alignment. This fact is reflected in a coupling dependent, nonuniversal exponent 1 /4Kchar- acterizing the increase of the staggered magnetization at λ > λ c. The order parameters as functions of λin the ∆S≫Jτlimit is schematically shown in Fig. 2(b). As already mentioned, the absence of a small parame- ter in the regime of strong hybridization, Jτ∼JS∼λ, makestheanalysisofthephasediagraminthisregionnot easily accessible by analytical tools. Nevertheless some plausible arguments can be put forward to comment on the topology of the phase diagram. It is tempting to treat the curve Jτ∆S/λ2∼1 as a single critical line go- ing throughout the whole phase plane ( Jτ/λ,∆S/λ). If so, we then can expect that there exists a special sin- gular point located in the region Jτ∆S/λ2∼1. This expectation is based on the fact that at Jτ≫∆slimit the transition is of the Ising type and the spontaneous spin magnetization below the critical curve follows the lawζ∼(λ−λc1)1/8with auniversal critical exponent, whereas at Jτ≪∆sthe spin magnetization has a differ- ent,nonuniversal exponent, ζ∼(λ−λc)1/4K. Continu- ity considerations make it very appealing to suggest that at the special point the Tomonaga-Luttinger liquid pa- rameter takes the value K= 2, and the two power laws match. Since the central charges of two Ising and one Gaussian criticalities satisfy the relation 1 /2+1/2 = 1, the singular point must be a point where the two Ising critical curves merge into a single Gaussian one. V. DYNAMICAL SIN SUSCEPTIBILITY AND NMR RELAXATION RATE IN THE VICINITY OF GAUSSIAN CRITICALITY It may seem at the first sight that, in the regime ∆S≫J, the spin degrees of freedom which have been integrated out remain massive across the orbital Gaus- sian transition, and the spectral weight of the staggered spin fluctuations is only nonzero in the high-energy re- gionω∼∆S. However, this conclusion is only correct for the zeroth-order definition of the spin field N0(x), given by Eq. (6), with respect to the spin-orbit inter- action. In fact, the staggered magnetization hybridizes with low-energyorbitalmodes viaSO couplingalreadyin the first order in λand thus acquires a low-energy pro- jection which contributes to a nonzero spectral weight displayed by the dynamical spin susceptibility at ener- gies well below the Haldane gap. Tofind the low-energyprojectionofthe field Nz(r), we must fuse the local operator Nz 0(r) with the perturbative part of the total action. Keeping in mind that close to and at the Gaussian criticality most strongly fluctuating fields are the staggered components of the orbital po- larization, we approximate the SO part of the Euclidian9 action by the expression SSτ≃λa0 vS/integraldisplay d2rNz(r)nx(r), (43) wherer= (vSτ,x) is the two-dimensional radius vector (hereτis the imaginary time). We thus construct Nz P(r) =/an}bracketle{te−SSτNz(r)/an}bracketri}ht =Nz 0(r)−λa0 vS/integraldisplay d2r1/an}bracketle{tNz 0(r)Nz 0(r1)/an}bracketri}htSnx(r1) +O(λ2), (44) where averaging is done over the unperturbed, high- energy spin modes. For simplicity, here we neglect the anisotropy of the spin-liquid phase of the S=1 chain and use formula (8). The spin correlation function is short- ranged. Treating the spin correlation length ξS∼vS/∆S as a new lattice constant (new ultraviolet cutoff) and be- ing interested in the infrared asymptotics |r| ≫ξS, we can replace in (44) nx(r1) bynx(r). The integral /integraldisplay d2ρ/an}bracketle{tNz 0(ρ)Nz 0(0)/an}bracketri}htS (45) ∼1 a2 0(a/ξS)3/4/integraldisplay∞ 0dρ ρ/radicalbig ξs/ρ e−ρ/ξS∼(ξS/a0)5/4. So the first-order low-energy projection of the staggered magnetization is proportional to Nz P(r)∼λ ∆S/parenleftbiggξS a0/parenrightbigg1/4 nx(r). (46) Thisresultclarifiestheessenceofthehybridizationeffect: close to the Gaussian criticality the spin fluctuations ac- quire a finite spectral weight in the low-energy region, ω≪∆S,q∼π, which is contributed by orbital fluctu- ations and can be probed in magnetic inelastic neutron scattering experiments and NMR measurements. Away from but close to the Gaussian criticality the behavior of the dynamical spin susceptibility ℑmχ(q,ω) is determined by the excitation spectrum of the sine- Gordon model for the dual field, Eq.(36). Since K >1, it consists of kinks, antikinks carrying the mass Morb, and their bound states (breathers) with masses (see e.g. Ref. 27) Mj= 2Morbsin(πj/2ν), j= 1,2,...ν−1, ν= 2K−1 (47) SinceK−1 = 2gis small, there will be only the first breather in the spectrum, with mass M1= 2Morb(1− 2π2g2). The sine-Gordon model is integrable, and the asymptotics of its correlation functions in the massive regime have been calculated using the form-factor ap- proach (see for a recent review 35). Here we utilize some of the known results. At λ < λ cthe operator nx∼sin√πΘ has a nonzero matrix element betweenthe vacuum and the first breather state. This form- factor contributes to a coherent peak in the dynamical spin susceptibility at frequencies much smaller than than the Haldane gap: ℑmχ(q,ω,T= 0) = A(λ/∆S)2δ[ω2−(q−π)2v2−M2 1] +ℑmχcont(q,ω,T= 0). (48) HereAis a constant and the second term is the contribu- tion of a multi-kink continuum of states with a threshold atω= 2Morb. Atλ > λ cthe spectral properties of the operator cos√πΘ coincide with those of the opera- tor sin√πΘ atλ < λ c. For symmetry reasons35, this operator does not couple to the first breather, so that atλ > λcℑmχ(q,ω) will only display the kink-antikink scattering continuum. Weseethat, due tospin-orbithybridizationeffects, the spin sector of our model loses the properties of a spin liq- uid alreadyin anoncriticalorbitalregime. Thistendency gets strongly enhanced at the orbital Gaussian criticality (Morb→0) where all multi-particle processesmerge, and the spin correlationfunction exhibits an algebraicallyde- caying asymptotics /an}bracketle{tNz(r)Nz(0)/an}bracketri}ht ≃ /an}bracketle{tNz P(r)Nz P(0)/an}bracketri}ht ∼/parenleftbiggλ ∆S/parenrightbigg2/parenleftBiga r/parenrightBig1 2K, (49) implying that the spin sector of the model becomes remi- niscent of Tomonaga-Luttinger liquid. In this limit (here for simplicity we consider the T= 0 case) the dynamical spin susceptibility is given by34 ℑmχ(q,ω,T= 0)∼(λ/∆S)2/bracketleftbig ω2−v2(q−π)2/bracketrightbig1 4K−1. (50) The NMR relaxation rate probes the spectrum of local spin fluctuations 1 T1=A2Tlim ω→01 ω/summationdisplay qℑmχzz(q,ω,T) whereAis an effective hyperfine constant. In spin- liquid regime of an isolated spin-1 chain, the existence of a Haldane gap makes 1 /T1exponentially suppressed36: 1/T1∼exp(−2∆S/T). The admixture of low-energy or- bital states in the spin-fluctuation spectrum drastically changedthisresult. Asimplepowercountingargument37 leads to a power-law temperature dependence of the NMR relaxation rate: 1 T1∼A2/parenleftbiggλ ∆s/parenrightbigg2 T1 2K−1(51) This result is valid not only exactly at the Gaussian crit- icality but also in its vicinity provided that the tempera- ture is larger than the orbital mass gap. By construction (see the preceding section) K≥1. This means that the exponent 1 /2K−1 isnegative and the NMR relaxation rateincreases on lowering the temperature. It is worth10 noticingthatsuchregimesarenotunusualforTomonaga- Luttingerphasesoffrustratedspin-1/2ladders.38Forour model, such behavior of 1 /T1would be a strong indica- tion of an extremely quantum nature of the collective orbital excitations.39 VI. BEHAVIOR IN A MAGNETIC FIELD: QUANTUM ISING TRANSITION IN ORBITAL SECTOR We have seen in Sec.III that, due to spin-orbit cou- pling, the N´ eel ordering of the spins is accompanied by the emergence of quantum effects in the orbital sector: the classical orbital Ising chain transforms to a quan- tum one. In this section we briefly comment on a similar situation that can arise upon application of a uniform external magnetic field h. Since the spin-1 chain is massive, it will acquire a finite ground-statemagnetization /an}bracketle{tSz/an}bracketri}htonly when the magnetic field,h, is higher than the critical value hc1∼∆S, corre- sponding to the commensurate-incommensurate (C-IC) transition. According to the definition (5), a uniform magnetic field along the z-axis,Hmag=−hIz, mixes up a pair of Majorana fields, ξ1andξ2, and splits the spectrum of Sz=±1 excitations (the Sz= 0 modes are unaffected by the field). At h=hc1the gap in the spec- trum of the Sz= 1 excitations closes, and at h > h c1 these modes condense giving rise to a finite magnetiza- tion. Once /an}bracketle{tSz/an}bracketri}ht /ne}ationslash= 0, the effective Hamiltonian of the τ-chain becomes ¯Hτ=Jτ/summationdisplay nτz nτz n+1−∆τ/summationdisplay nτx n,∆τ=λ/an}bracketle{tSz/an}bracketri}ht.(52) Here we ignore the fluctuation term that couples τx nto ∆Sz n=Sz n−/an}bracketle{tSz n/an}bracketri}ht. One should keep in mind that there exists the sec- ond C-IC transition at a higher field hc2associated with full polarization of the spin-1 chain. To simplify further analysis, let us assume that the range of magnetic fields hc1< h < h c2, where an isolated spin-1 chain has an in- commensurate,gaplessgroundstate,issufficientlybroad. This can be easily achieved in the biquadratic model (3) withβ∼1, in which case the Haldane gap – and hence hc1– is small, and the effects associated with the second C-IC transition can be neglected. Now, by increasing the magnetic field hin the region h > h c1, the effective orbital chain (52) can be driven to an Ising criticality. The induced transverse “magnetic field” ∆ τis proportional to a nonzero magnetization of the spin-1 chain. If λ/Jτis large enough, then upon in- creasing the field the effective quantum Ising chain (52) can reach the point ∆ τ(h∗) =Jτwhere the Ising transi- tion occurs. This will happen at some field h=h∗> hc1. In the region |h−h∗|/h∗≪1 the quantum Ising τ-chain will be slightly off-critical. Due to the SO coupling, thesemassive orbital excitations will interact with the gap- lessSz=±1 spin modes. However, this interaction can only give rise to the orbital mass renormalization (i.e. a small shift of the Ising critical point) and a group veloc- ity renormalization of the spin-doublet modes. For this reason we do not expect the aforementioned spin-orbital fluctuation term to cause any qualitative changes. The above discussion reveals an interesting fact: a suf- ficiently strong magnetic field acting on the spin degrees of freedom can affect the orbital structure of the chain anddriveit to aquantum Isingtransition. Thedifference with the situation discussedin Sec.III is that the external magnetic field induces a uniform spin polarization which, in turn, gives rise to a uniform transverse orbital order- ing/an}bracketle{tτx n/an}bracketri}ht /ne}ationslash= 0. Thus, the classical long-range orbital order /an}bracketle{tτz n/an}bracketri}ht= (−1)nηz, present at h < h∗, disappears in the regionh > h∗, where the orbital degrees of freedom are characterized by a transverse ferromagnetic polarization, /an}bracketle{tτx/an}bracketri}ht /ne}ationslash= 0. VII. CONCLUSION AND DISCUSSION In this paper, we have proposed and analyzed a 1D spin-orbital model in which a spin-1 Haldane chain is lo- cally coupled to an orbital Ising chain by an on-site term λτxSzoriginating from relativistic spin-orbit (SO) in- teraction. The SO term not only introduces anisotropy to the spin sector, but also gives quantum dynamics to the orbital degrees of freedom. We approach this prob- lem from well defined limits where either the spin or the orbital sector is strongly gapped and becomes a ‘fast’ subsystem which can be integrated out. By analyzing the resultant effective action of the remaining ‘slow’ de- grees of freedom, we have identified the stable massive and critical phases of the model which are summarized in a schematic phase diagram shown in Fig. 1. Inthe limit dominatedbyalargeorbitalgap, i.e. Jτ≫ ∆S, integrating out the orbital variables gives rise to an easy-axis spin anisotropy D(Sz)2whereD∼ −λ2/Jτ. Asλincreases, the disordered Haldane spin liquid un- dergoes an Ising transition into a magnetically ordered N´ eel state. The presence of antiferromagnetic spin order ζin the N´ eel phase in turn generates an effective trans- verse field h∼λζacting on the orbital Ising variables. The orbital sector which is described by the Hamiltonian of a quantum Ising chain reaches criticality when h=Jτ. In between the two Ising critical points lies an interme- diate phase (phase II in Fig. 1) where both Ising order parameters ηxandηzare nonzero. Such a two-stage or- deringscenarioillustratedbypath1inthephasediagram (Fig.1)hasbeenconfirmednumericallybyrecentDMRG calculations.22Interestingly, the orbital Ising transition can also be induced by applying a magnetic field to the spin sector. As the field strength is greater than the Hal- dane gap, a field-induced magnon condensation results in a finite magnetization density /an}bracketle{tSz/an}bracketri}htin the linear chain. Thanks to the SO coupling, the orbital sector again ac-11 quires a transverse field h∼λ/an}bracketle{tSz/an}bracketri}htand becomes critical whenh=Jτ. A distinct scenario of the orbital reorientation tran- sitionηz→ηxoccurs in the opposite limit ∆ S≫Jτ. This time we integrate out the fast spin subsystem and obtain a perturbed spin-1/2 XY Hamiltonian for the or- bital sector. The effective exchange constants are given byJx∼λ2/∆SandJy=Jτ. Asλis varied, the orbital sector reaches a Gaussian critical point when Jx=Jy, at which the system acquires an emergent U(1) sym- metry. The orbital order parameter goes directly from η= (0,0,ηz)to(ηx,0,0)inthissingle-transitionscenario (illustrated by path 2 in Fig. 1). Both order parameters ηxandηzvanish at the critical point. We have shown that spin-orbital hybridization effects near the Gaussian transition lead to the appearance of a non-zero spectral weight of the staggered spin density well below the Hal- dane gap – the effect which can be detected by inelastic neutronscatteringexperimentsandNMRmeasurements. The stability analysis of the orbital Gaussian criti- cality in the original lattice model (1), done in Ap- pendix B, has shown that this critical regime is pro- tected by the τz→ −τzsymmetry of the underlying microscopic model. This symmetry will be broken in the presence of an orbital field δ/summationtext nτz nwhich removes degeneracy between the local orbitals dzxanddyzand adds a ”magnetic” field along the y-axis in the effective XY model (32). Such perturbation will drive the orbital sector away from the Gaussian criticality. The same argument applies to a perturbation with the structure β/summationtext nSz nτz nwhich also breaks the aforementioned sym- metry. Integrating over the spins will generate an extra term∼λβ/summationtext n(τx nτy n+1+τy nτz n+1) which, in the contin- uum limit, translates to λβsin√ 4πΘ. As explained in Appendix B, such perturbation will keep the orbital sec- tor gapped with coexisting ηxandηzorderings. Since the analysispresentedin thispaper isdonein the limiting cases, precise predictions on the detailed shape of the phase diagram or on the behavior of correlation functions in the regime of strong hybridization of spin and orbital degrees of freedom, where all interactions in- cluded inthe modelareofthe sameorder, arebeyondour reach and require further numerical calculations. On the other hand, the continuity and scaling analysis allow us to believe that the global topology of the phase diagram and character of critical lines are given correctly. Finally the spin-orbital model Eq. (1) can be generalized to the zigzag geometrical where two parallel spin-1 chains are coupled to a zigzag Ising orbital chain via on-site SO in- teraction. The zigzag case is closely related to the quasi- 1D compound CaV 2O4. While the two-Ising-transitions scenario is expected to hold in the Jτ≫∆Sregime, the counterpart of Gaussian criticality in the zigzag chain remains to be explored and will be left for future study.Acknowledgements The authors are grateful to Andrey Chubukov, Fabian Essler, Vladimir Gritsev, Philippe Lecheminant and Alexei Tsvelik for stimulating discussions. A.N. grate- fully acknowledges hospitality of the Abdus Salam Inter- national Centre for Theoretical Physics, Trieste, where part of this work has been done. He is also supported by the grants GNSF-ST09/4-447 and IZ73Z0-128058/1. G.W.C. acknowledges the support of ICAM and NSF grant DMR-0844115. N.P. acknowledges the support from NSF grant DMR-1005932 and ASG ”Unconven- tional magnetism”. G.W.C. and N.P. also thank the hos- pitality of the visitors program at MPIPKS, where the part of the work on this manuscript has been done. Appendix A: Ising correlation function In this Appendix we estimate the correlation func- tion Γxx nm(τ) =/an}bracketle{tτx n(τ)τx m(0)/an}bracketri}ht, where the averaging is per- formed over the ground state of the Ising Hamiltonian Hτ=Jτ/summationtext nτz nτz n+1, andτx n(τ) =eτHττx ne−τHτ. It proves useful to make a duality transformation: τz nτz n+1=µx n, τx n=µz nµz n+1. The new set of Pauli matrices µa nrepresents disorder op- erators. The Hamiltonian and correlation function be- come: H→Jτ/summationdisplay nµx n, (A1) Γzz nm(τ)→ /an}bracketle{tµz n(τ)µz n+1(τ)µz m(0)µz m+1(0)/an}bracketri}ht.(A2) The most important fact about the dual representation is the additive, single-spin structure of the Hamiltonian: the latter describes noninteracting spins in an external “magnetic field” Jτ. Notice that by symmetry /an}bracketle{tµz n/an}bracketri}ht= 0. Therefore the correlationfunction in (A2) has an ultralo- cal structure: Γxx nm(τ) =δnmY2(τ), Y(τ) =/an}bracketle{tµz n(τ)µz n(0)/an}bracketri}ht.(A3) The time-dependence of the disorder operator can be ex- plicitly computed, µz n(τ) =eτJτµx nµz ne−τJτµx n=µz ncosh(2Jττ)−iµy nsinh(2Jττ). Therefore (below we assume that τ > τ′) Y(τ−τ′) = cosh2 Jτ(τ−τ′)+/an}bracketle{tµx/an}bracketri}htsinh2Jτ(τ−τ′) = exp[−2|Jτ|(τ−τ′)]. (A4) Hereweusedthefactthat, inthegroundstatetheHamil- tonianHτ,/an}bracketle{tµx/an}bracketri}ht=−sgnJτ. Thus, as expected for the 1D Ising model, the correlation function Γxx nm(τ) is local in real space and decays exponentially with τ: Γxx nm(τ) =δnmexp(−4J⊥|τ|). (A5)12 Appendix B: Perturbed XY chain, Eq. (32) In this Appendix we analyze the perturbation (33) to the XY spin chain (32) and show that at the XX point it represents a marginal perturbation which transforms the free-fermion regime to a Gaussian criticality describ- ing a Luttinger-liquid behavior of the orbital degrees of freedom. Using the Jordan-Wigner transformation (34) we rewrite (33) as H′ τ=H′ 1+H′ 2, where H′ 1=J′ x(2) 2/summationdisplay n(a† nan+2+h.c.)(a† n+1an+1−1 2),(B1) H′ 2=J′ x(2) 2/summationdisplay n(a† na† n+2+h.c.)(a† n+1an+1−1 2).(B2) Assuming that |Jx−Jy|,J′ x≪Jx+Jy, we pass to a con- tinuum description of the XY chain in terms of chiral, right (R) and left (L), fermionic fields based on the de- composition (to simplify notations we set here a0= 1): an→(−i)nR(x)+inL(x).Then the Hamiltonian density of the XY model takes the form: HXY(x) =−iv/parenleftbig R†∂xR−L†∂xL/parenrightbig −2iγ/parenleftbig R†L†−h.c./parenrightbig , (B3) whereγ=Jx−Jy. Standard rules of Abelian bosonization27transform(B3)toaquantumsine-Gordon model: HXY(x) =v 2/bracketleftBig Π2+(∂xΦ)2/bracketrightBig +2γ παcos√ 4πΘ,(B4) wherev= 2(Jx+Jy)a0is the Fermi velocity, Π( x) = ∂xΘ(x) is the momentum conjugate to the scalar field Φ(x) = Φ R(x) + ΦL(x), and Θ( x) =−ΦR(x) + ΦL(x) is the field dual to Φ( x). Here Φ R,L(x) are chiral components of the scalar field. Using the fact that the fermions are spinless, one can impose the condi- tion [Φ R(x),ΦL(x′)] =i/4 and thus make sure that the bosonization rules correctly reproduce the anticommuta- tion relations {R(x),L(x′)}={R(x),L†(x′)}= 0. An explicit introduction of the so-called Klein factors be- comes necessary when bosonizing fermions with an inter- nal degree of freedom, such as spin 1/2, chain index etc, which is not the case here. Let is find the structure of the perturbation (33) in the continuum limit. First of all we notice that a† n+1an+1−1/2≡:a† n+1an+1: →(:R†R: + :L†L:)+(−1)n+1(R†L+L†R) =1√π∂xΦ+(−1)n παsin√ 4πΦ. (B5) Similarly a† nan+2+h.c. → −2/bracketleftbig (:R†R: + :L†L:)+(−1)n(R†L+L†R)/bracketrightbig =−2/bracketleftbigg1√π∂xΦ−(−1)n παsin√ 4πΦ/bracketrightbigg . (B6)Dropping Umklapp processes R†(x)R†(x+α)L(x+ α)L(x) +h.c.∼cos√ 16πΦ as strongly irrelevant (with scaling dimension 4) at the XXcriticality and ignoring interaction of the fermions in the vicinity of the same Fermi point, we find that (a† nan+2+h.c.)(a† n+1an+1−1/2)/vextendsingle/vextendsingle/vextendsingle smooth → −8 :R†R::L†L:= 2/bracketleftbig Π2−(∂xΦ)2/bracketrightbig .(B7) We see that the perturbation H′ 1generates a marginal four-fermion interaction to the free-fermion model (B3), thus transforming the model (32) to an XYZ model with a weak ferromagnetic( zz)-coupling. This interaction can be incorporated into the Gaussian part of the bosonic theory (B4) by changing the compactification radius of the field Φ: H=HXY+H′ 1 =u 2/bracketleftbigg KΠ2+1 K(∂xΦ)2/bracketrightbigg −2γ παcos√ 4πΘ.(B8) Hereuis the renormalized velocity and Kis the inter- action constant which at J′ x≪(Jx+Jy) is given by K= 1+2g+O(g2),whereg=J′ x(2)a0/πv≪1. Now we turn to H′ 2. We have: a† na† n+2+h.c. (B9) → −/bracketleftbig R†(x)L†(x+α)+L†(x)R†(x+α)+h.c./bracketrightbig +(−1)n/bracketleftbig R†(x)R†(x+α)+L†(x)L†(x+α)+h.c./bracketrightbig . Bosonizing the smooth term in the r.h.s. of (B10) one obtains∂xΦcos√ 4πΘ. Bosonizing the staggered term yields sin√ 4πΦcos√ 4πΘ. Using the OPE sin√ 4πΦ(x)sin√ 4πΦ(x+α) = const−πα2(∂xΦ)2−1 2cos√ 16πΦ, we find that, in the continuum limit, the Hamiltonian densityH′ 2is contributed by the operatorscos√ 4πΘ and (∂xΦ)2cos√ 4πΘ (as before, we drop corrections related to Umklapp processes). The former leads to a small ad- ditive renormalization of the fermionic mass γand thus produces a shift of the critical point. The latter repre- sents an irrelevant perturbation (with scaling dimension 3) at the XX criticality. In a noncritical regime it renor- malizes the mass and four-fermion coupling constant g. Considering the structure of the remaining terms in the expansion (31) one arrives at similar conclusions. Here a remark is in order. The only dangerous perturba- tion which would dramatically affect the above picture is sin√ 4πΘ. The presence of two nonlinear terms in the Hamiltonian, γcos√ 4πΘ+δsin√ 4πΘ, would make the fermionic mass equal to/radicalbig (λ−λc)2+δ2. The Gaus- sian criticality in this case would never be reached, the13 model would always remain massive, and nonzero stag- gered pseudospin densities, ηzandηx, would coexist in the whole parameter range of the model. Fortunately, the appearance of the operator sin√ 4πΘ is forbidden by symmetry. The initial Hamiltonian (1) is invariant under global pseudospin inversion In the z- component only: τz n→ −τz n. After rotation τz→τythis translates to τy n→ −τy n. Using the bosonized ex- pressions (39) for the staggered pseudospin densities we find that the corresponding transformation of the dual field is Θ →√π−Θ and so the bosonized Hamiltonian density must be invariant under this transfomation. This explainswhythe operatorsin√ 4πΘcannotappearin the effective continuum theory. 1E. Axtell, T. Ozawa, S. Kauzlarich, and R. R. P. Singh, J. Solid State Chem. 134, 423 (1997). 2M. Isobe and Y. Ueda, J. Phys. Soc. Jpn. 65, 1178 (1996) 3K. I. Kugel andD. I. Khomsky, JETP Lett. 15, 446 (1971). 4Y. Tokura and N. Nagaosa, Science, 288, 462 (2000). 5E. Zhao and W. V. Liu, Phys. Rev. Lett. 100, 160403 (2008). 6C. Wu, Phys. Rev. Lett. 100, 200406 (2008). 7Z. Nussinov and E.Fradkin, Phys. Rev. B 71, 195120 (2005). 8A. Kitaev, Ann. Phys. (N.Y.) 321, 2 (2006). 9P. Azaria, A.O. Gogolin., P. Lecheminant, and A.A. Ners- esyan, Phys. Rev. Lett. 83, 624 (1999). 10C. Itoi, S. Qin, and I. Affleck, Phys. Rev. B 61, 6747 (2000). 11Y. Yamashita, N. Shibata, and K. Ueda, Phys. Rev. B 58, 9114 (1998). 12B. Frischmuth, F. Mila, and M. Troyer, Phys. Rev. Lett. 82, 835 (1999). 13C. Itoi, S. Qin, andI. Affleck, Phys. Rev.B 61, 6747 (2000) 14H. Mamiya, M. Onoda, T. Furubayashi, J. Tang, and I. Nakatani, J. Appl. Phys. 81, 5289 (1997). 15M. Reehuis, A. Krimmel, N. B´ uttgen, A. Loidl, A. Prokofiev, Eur. Phys. J. B 35, 311 (2003). 16M. Onoda and J. Hasegawa, J. Phys.: Condens. Matter 15, 95 (2003). 17S.-H. Lee, D. Louca, H. Ueda, S. Park, T. J. Sato, M. Isobe, Y. Ueda, S. Rosenkranz, P. Zschack, J. Iniguez, Y. Qiu, R. Osborn, Phys. Rev. Lett. 93, 15640 (2004). 18A. Niazi et al, Phys. Rev. B 79, 104432 (2009). 19O. Pieper et al, Phys. Rev. B 79,180409(R) (2009). 20G.-W. Chern and N. Perkins, Phys. Rev. B 80, 220405(R) (2009). 21F. D. M. Haldane, Phys. Lett. 93A, 464 (1983). 22G.-W. Chern, N. Perkins and G. I. Japaridze, Phys. Rev. B82, 172408 (2010). 23H.-J. Mikeska and A. K. Kolezhuk, Lect. Notes Phys. 645, 1 (2004).24A. M. Tsvelik, Phys. Rev. B 42, 10 499 (1990). 25L. Takhtajian, Phys. Lett. 87 A, 479 (1982); H. Babujian, ibid.90 A, 464 (1982). 26D. Shelton, A. A. Nersesyan and A. M. Tsvelik, Phys. Rev. B53, 8521 (1996). 27A. O. Gogolin, A. A. Nersesyan and A. M. Tsvelik, Bosonization andStronglyCorrelated Systems , Cambridge University Press, 1998. 28A. B. Zamolodchikov and V. A. Fateev, Sov. J. Nucl. Phys. 43, 657 (1986). 29P. Di Francesco, P. Mathieu and D. Senechal, Conformal Field Theory , (Springer-Verlag, 1997). 30A. A. Nersesyan and A. M. Tsvelik, Phys. Rev. Lett. 78, 3939 (1997). 31J. Kogut, Rev. Mod. Phys. 51, 659 (1979). 32E. S. Sorensen and I. Affleck, Phys. Rev. B 49, 15771 (1994). 33E. Lieb, T. D. Schultz, and D. C. Mattis, Ann. Phys. (NY) 16, 407 (1961). 34H. J. Schulz and C. Bourbonnais, Phys. Rev. B 27, 5856 (1983); H.J. Schulz, Phys. Rev. B 34, 6372 (1986). 35F.H.L. Essler and R.M. Konik, Applications of Massive Integrable Quantum Field Theories to Problems in Con- densed Matter Physics , in the Ian Kogan Memorial Vol- ume ”From Fields to Strings: Circumnavigating Theoret- ical Physics”, eds. M. Shifman, A. Vainshtein and J.F. Wheater, World Scientific, 2005, p. 684-831. 36J. Sagi and I. Affleck, Phys. Rev. B 539188 (1996). 37T. Giamarchi, Quantum Physics in One Domension , Ox- ford University Press, 2003. 38M. Sato, T. Momoi, and A. Furusaki, Phys. Rev. B 79, 060406(R) (2009). 39In the most general setting, there will be a contribution to the NMR relaxation rate coming directly from the orbital sector. However, since the coupling to orbitals is non-loca l, this contribution is thus less singular and is usually negli - gible compared with contribution from spins.
1106.5667v2.Phase_separation_in_a_polarized_Fermi_gas_with_spin_orbit_coupling.pdf
Phase separation in a polarized Fermi gas with spin-orbit coupling W. Yi and G.-C. Guo Key Laboratory of Quantum Information, University of Science and Technology of China, CAS, Hefei, Anhui, 230026, People's Republic of China (Dated: November 20, 2018) We study the phase separation of a spin polarized Fermi gas with spin-orbit coupling near a wide Feshbach resonance. As a result of the competition between spin-orbit coupling and population imbalance, the phase diagram for a uniform gas develops a rich structure of phase separated states involving topologically non-trivial gapless super uid states. We then demonstrate the phase sepa- ration induced by an external trapping potential and discuss the optimal parameter region for the experimental observation of the gapless super uid phases. PACS numbers: 03.75.Ss, 03.75.Lm, 05.30.Fk Spin-orbit coupling (SOC), common in condensed mat- ter systems for electrons, has been considered a key in- gredient for many interesting phenomena such as topo- logical insulators [1], quantum spin Hall e ects [2], etc. The recent realization of synthetic gauge eld and hence spin-orbit couplings in ultracold atomic systems opens up exciting new routes in the study of these phenomena [3, 4], allowing us to take advantage of the features of the ultracold atoms, e.g. clean environment and highly con- trollable parameters. In particular, with the Feshbach resonance technique, the e ective interaction strength between atoms can be tuned [5, 6]. This technique has been applied to study various interesting topics, e.g. the BCS-BEC crossover [7], polarize Fermi gases [8], itinerant ferromagnetism [9], etc. The introduction of spin-orbit coupling may shed new light on these strongly correlated systems. Spin-orbit coupled Fermi gas near a Feshbach reso- nance has recently attracted much theoretical attention [10{16]. The SOC has been shown to enhance pairing on the BCS side of the Feshbach resonance [12, 14, 15]. Furthermore, for a polarized Fermi gas, the SOC intro- duces competition against population imbalance, which can lead to topologically non-trivial phases [13, 16]. Re- cently, the phase diagrams for a polarized Fermi gas with spin-orbit coupling near a Feshbach resonance have been reported for a uniform gas [16]. The phase boundaries have been calculated by solving the gap equation and the number equations self-consistently. However, simi- lar to the case of a polarized Fermi gas near Feshbach resonance [17], due to the competition between di er- ent phases, the solutions of the gap equation may corre- spond to metastable or unstable states. By considering the compressibility criterion [16], the unstable solutions are correctly discarded, while the metastable solutions may survive, rendering the resulting phase boundaries, in particular those representing rst order phase transi- tions, unreliable. In this paper, we examine in detail the zero temper- ature phase diagrams for a polarized Fermi gas with Rashba spin-orbit coupling near a wide Feshbach reso-nance for both the uniform and the trapped cases. To avoid getting metastable or unstable solutions, instead of solving the gap equation, we minimize the thermody- namic potential directly as in Ref. [17]. For the uniform gas, we nd larger stability regions for the phase sepa- rated state at unitarity as compared to the results in Ref. [16]. More interestingly, we nd that SOC may induce more complicated phase separated states involving gap- less super uid phases that are topologically non-trivial, in addition to the typical phase separated state composed of normal (N) and gapped super uid (SF) phases. We calculate the stability region for the various phase sepa- rated states as well as for the gapless super uid states, SF state and normal state. We show that there are two dis- tinct gapless phases that di er by the number of crossings their excitation spectra have with the zero energy in mo- mentum space, consistent with previous results [13, 16]. These novel gapless phases are stabilized by intermedi- ate SOC strengths; whereas for large enough SOC, the system always becomes a gapped super uid of `rashbons' [12]. We show how these phases can be characterized by their di erent excitation spectra and momentum space density distributions. We then discuss the phase separa- tion in an external trapping potential, where the various phases naturally phase separate in real space. By exam- ining their respective stability regions, we demonstrate the optimal parameter region to observe the gapless su- per uid states in the presence of a trapping potential. For all of our calculations in the paper, we adopt the BCS-type mean eld treatment. Although the mean eld theory does not give quantitatively accurate results near a wide Feshbach resonance, it is a natural rst step for us to qualitatively estimate what phases may be stable, as well as to understand their respective properties. We also note that we have neglected the Fulde-Ferrell-Larkin- Ovchinnikov (FFLO) phase in our calculations. This is motivated by the fact that the FFLO phase is stable only in a narrow parameter region in the absence of SOC due to competition against other phases [8]. As SOC intro- duces new gapless phases into this competition, we do not expect a signi cant increase in its stability region.arXiv:1106.5667v2 [cond-mat.quant-gas] 30 Sep 20112 We rst consider a uniform three dimensional polarized Fermi gas with Rashba spin-orbit coupling in the plane perpendicular to the quantization axis z. The model Hamiltonian takes the form [13, 14, 16] HX N=X k;ka† k;ak; +h 2X k a† k;#ak;#a† k;"ak;" +U VX k;k0a† k;"a† k;#ak0;#ak0;" +X k k? ei'ka† k;"ak;#+h:c: ; (1) wherek=k, with the kinetic energy k=~2k2 2m; =f";#gare the atomic spins; Ndenotes the to- tal number of particles with spin ;ak;(a† k;) annihi- lates (creates) a fermion with momentum kand spin; =h=2 is the chemical potential of the correspond- ing spin species, and Vis the quantization volume. The Rashba spin-orbit coupling strength can be tuned via parameters of the gauge- eld generating lasers [4], while k?=q k2x+k2yand'k= arg (kx+iky). In writing Hamiltonian (1), we assume s-wave contact interaction between the two fermion species, with the bare interac- tion rateUrenormalized following the standard relation 1 U=1 Up1 VP k1 2k[7]. The physical interaction rate is given asUp=4~2as m, whereasis the s-wave scattering length between the two fermionic spin species. To diagonalize the Hamiltonian, we make the trans- formation: ak;"=1p 2ei'k(ak;++ak;),ak;#= 1p 2(ak;+ak;), whereak;are the annihilation op- erators for the dressed spin states with di erent helic- ities () [12{16]. Taking the pairing mean eld  = U VP khak;#ak;"ias in the standard BCS-type theory, we may diagonalize the mean eld Hamiltonian in the basis of the dressed spins:n ak;+;a† k;+;ak;;a† k;oT . The thermodynamic potential is then evaluated from =1 ln tr e (HP N) , with = 1=kBT. In this paper, we will focus on the zero temperature case, for which the thermodynamic potential has the form =1 2X k;=(Ek;)Vjj2 U; (2) with the quasi-particle excitation spectrum Ek;=r 2 k+ 2k2 ?+jj2+h2 42q (h2 4+ 2k2 ?)2 k+h2 4jj2. Before proceeding, let us examine the quasi-particle excitations rst and study the conditions for possible gapless phases. We see that at the points in the momentum space where Ek;crosses zero, the quasi- particle excitation becomes gapless while the pairing gap  remains nite. The SOC, together with the population imbalance re-arranges the topology of the 0 0.5 1−0.2−0.15−0.1Ω/h ∆/h0 0.5 1−0.4−0.35−0.3 ∆/hΩ/h 0 0.5 1−0.23−0.22−0.21−0.2 ∆/hΩ/h 0 0.5 1−0.32−0.3−0.28−0.26 ∆/hΩ/h(a) (c)(b) (d)N GP2SF SFFIG. 1. Illustration of typical shapes of the thermodynamic potential =has a function of order parameter  =hfor var- ious phases at unitarity: (a) =h = 0:52, kh=h= 0:1; (b) =h= 0:7, kh=h= 0:1; (c)=h= 0:52, kh=h= 0:3, (d) =h= 0:52, kF=h= 0:6. The chemical potential his taken to be the energy unit, while the unit of momentum khis de- ned through~2k2 h 2m=h. Fermi surfaces of the spin species [13, 16]. The points of gapless excitations lie on the kzaxis withk?= 0, and are symmetric with respect to the kz= 0 plane. More speci cally, for 0, the excitation spectrum has two gapless points 2m ~2 +q h2 4jj21 2 , so long asjhj 2>p 2+jj2. For > 0, the excitation spectrum has four gapless points( 2m ~2 +q h2 4jj21 2 ;2m ~2 q h2 4jj21 2) , withjj<jhj 2<p 2+jj2; two gapless points 2m ~2 +q h2 4jj21 2 , withjhj 2>p 2+jj2. We identify the super uid states with two excitation points (GP1) and those with four excitation points (GP2) as di erent topological phases [13, 16]. We illustrate in Fig. 1 typical shapes of the thermo- dynamic potential as a function of  with di erent pa- rameters. Notably, due to the competition between dif- ferent phases, a double-well structure appears (see Fig. 1(a-c)). Hence the solutions to the gap equation may correspond to the metastable states (local minimum) or the unstable states (local maximum). To make sure that the ground state is achieved, we directly minimize the thermodynamic potential [17]. Another complication comes from the existence of the phase separated state, which must be considered explic- itly for a uniform gas. As in the case of polarized Fermi gases without SOC [8], we introduce the mixing coe- cientx(0x1), and the thermodynamic potential becomes =x ( 1) + (1x) ( 2); (3)3 00.2 0.4 0.6 0.8 11.200.20.40.60.81 α kF/EFPGP1 N+SF GP2+SF SFGP2N GP2+GP2 FIG. 2. Zero temperature phase diagram for a uniform Fermi gas with population imbalance at ( kFas)1= 0. Within the bold phase boundaries are the various phase separated states (see text). These phase separated states can be connected with the non-phase separated states by rst order phase tran- sitions (solid bold curve). The thin curves represent various second order phase transitions (see text). Here kF= (32n)1 3, EF=~2k2 F 2m, andnis the total density of the system. where i(i= 1;2) is the pairing gap for the ith com- ponent state. Note that due to SOC, we now have the possibility of a phase separated state of two distinct su- per uid states (see Fig. 1(c)). The number equations of the phase separated state become N=x@ @ = 1+ (1x)@ @ = 2: (4) Minimizing the thermodynamic potential Eq. (3) with respect to  iandxwhile implementing the number con- straints Eq. (4), we map out the phase diagram for a uni- form polarized Fermi gas with SOC at ( kFas)1= 0. Fig. 2 illustrates the resulting phase boundaries in the plane of (P; kF=EF), where the polarization P=N"N# N"+N#. When the SOC is o ( = 0), the system remains in a phase separated state of normal and gapped super uid (PS1) up toP0:93 before it becomes a normal state via a rst order phase transition. This is consistent with pre- vious mean eld calculations for a polarized Fermi gas [8, 19], while di erent from the result in Ref. [16]. As the SOC strength increases, a rich structure of dif- ferent phases shows up, e.g. gapped super uid phase (SF), gapless super uid phases with di erent Fermi sur- face topology (GP1 and GP2), and notably, various phase separated states. These phase separated states are con- ned by a phase boundary of rst order phase transition (bold curve in Fig. 2). In addition to the typical PS1 phase, we now have a phase separated state with GP2 and SF phases (PS2), and a phase separated state of two distinct GP2 phases (PS3). As increases, the system can undergo second order phase transitions from PS1 to PS2 and then to PS3 for intermediate Pand . As- sumingj1j<j2j, the phase boundaries between them −1 0 101 kz/khEk,−/h 0 1 200.51 kz/khDensity 0 200.51 kz/khDensity 0 200.51 kz/khDensity0121 k⊥/kh0120.51 k⊥/kh 0121 k⊥/khSF GP1 GP2(a) (b) (d) (c)FIG. 3. Typical excitation spectrum and momentum space density distribution for di erent phases. (a) Lower branch of the excitation spectra for GP1 (solid), GP2 (dashed) and SF (dash-dotted) phases; (b-d) Density distribution in momen- tum space for spin-up (solid) and spin-down (dashed) species alongk?= 0 andkz= 0 (inset), for (b) kh=h= 0:35, =h = 0:5 (SF); (c) kh=h= 0:35,=h = 0:45 (GP1); (d) kh=h= 0:45,=h= 0:43 (GP2), respectively. can be determined by imposing  1= 0 (PS1 and PS2) andh 2=j2j(PS2 and PS3), respectively. These phase separated states nally become unstable and give way to single component super uid phases as becomes large. The phase boundaries between these single component states are determined by settingh 2=jj(SF and GP2), h 2=p 2+jj2(GP2 and GP1), and  = 0 (N and GP1), respectively. When is large enough, the stabil- ity region of the GP2 phase decreases and nally vanishes at a tri-critical point ( = 0), beyond which only GP1, SF and normal phase may exist. Note that beyond the tri-critical point, the chemical potential becomes neg- ative, and the phase boundary between GP1 and SF will bend upwards so that in the large limit the SF phase becomes dominant in the phase diagram. To characterize the properties of the di erent phases, we calculate the excitation spectrum and number distribution in momentum space for SF, GP1 and GP2 states (see Fig. 3). Several interest- ing observations are in order. Firstly, the gapless phases leave their signatures in the momentum space density distribution. For k?= 0 andjkzj 2" min 0;r q h2 4jj2! ;r +q h2 4jj2# , the minority spin component vanishes, and pairing does not occur in this region. This is reminiscent of the momentum space phase separation of a breached pairing phase in the polarized Fermi gas [18], though now the unpaired region lies only on the kzaxis. Away from kzaxis, the occupation of the minority spin recovers from zero gradually, leaving a signature which may be detected in the time of ight imaging experiment [16]. Secondly, for nite , both the gapless and the gapped4 0 0.5 1 1.5−0.500.5 αkh/hµ/h0.42 0.450.430.46 αkh/hµ/h SF GP1 NGP2 V FIG. 4. Phase diagram in the ( =h; k h=h) plane at (khas)1= 0. While the second order phase transitions are in dashed thin curves, the rst order phase transitions are shown in solid bold curves, which end at the point where the double-well structure in the thermodynamic potential disap- pears (inset). The boundary for vacuum (V) is determined by setting the chemical potential of the majority spin species to vanish in the normal phase. super uid phases can support population imbalance, which can be seen from the density distribution along k?(see Fig. 3 insets). Indeed as we will see later, for large enough , we may expect no phase separation even in the presence of a harmonic trapping potential. The atoms in the trap will all be in the super uid phase induced by SOC. To understand the spatial distribution of the various phases in a trapping potential, we calculate the phase diagram as a function of ( =h; kh=h) at unitarity (Fig. 4), wherekhis de ned in the caption of Fig. 1. Under the Local Density Approximation (LDA) while assuming both spin species experience the same harmonic poten- tial, the local chemical potential (r) can be related to that at the center of the trap as(r) =V(r), whereV(r) gives the trapping potential. Thus a down- ward vertical line in Fig. 4 represents a trajectory from a trap center to its edge, with the chemical potential at the trap center xed by that at the starting point of the line. In Fig. 4, consistent with Fig. 2, the GP2 phase only exists in a small parameter region in the trap, while there appears to be considerable stability regions for the GP1 phase. When is small, the Fermi gas in the trap will phase separate into two regions, SF at the core, nor- mal phase (N) towards the edge. At intermediate , the gapless phases GP2 and GP1 may appear either near the center of the trap or as a ring between the SF core and the normal edge, depending on the chemical potentials. Note that the boundary of the rst order phase transi- tion between PS3 and GP2 (dotted thin curve in Fig. 2) corresponds to a small scale structure here (Fig. 4 inset), where a rst order phase transition (bold black curve) ex- 0 1 2012 αkh/hµ/h 00.20.40.6−0.5−0.4−0.3−0.2 αkh/hµ/hGP2 GP1SF GP1SF(b) (a) N VN VFIG. 5. Phase diagram in the ( =h; k h=h) plane at (a) (khas)1=1 and (b) ( khas)1= 0:5. First order phase transitions are shown in solid bold curves, while second order phase transitions are in dashed thin curves. ists between two distinct gapless super uids, both in the GP2 phase. However, this region is found to be small at unitarity and only increases slightly towards the BCS side of the resonance. It is therefore dicult to observe this phase transition in the trapped Fermi gas in the parame- ter region that we considered. For large SOC beyond the tip of the GP1-SF phase boundary, there is only the SF phase in the phase diagram, and hence we will have only the SOC induced SF phase in the trap for large enough SOC. We have also calculated the phase diagram in (=h; kh=h) plane away from the resonance point. On the BCS side (Fig. 5(a)), the stability region for the GP2 phase increases considerably. It is therefore desirable to prepare the system on the BCS side of the resonance for the observation of GP2 phases. On the BEC side (Fig. 5(b)), the GP2 phase vanishes from the phase diagram altogether for <0, consistent with our previous discus- sion. In summary, we have calculated in detail the phase di- agrams near a wide Feshbach resonance for a polarized Fermi gas with Rashba spin-orbit coupling. We nd that the competition among pairing, polarization and SOC gives rise to a rich structure of phases and phase separa- tions involving topologically non-trivial phases. From the phase diagrams for both uniform and trapped systems, we nd that the interesting gapless super uid phases are most likely to be observed in an experiment with moder- ate polarization and SOC strength. We would like to thank L.-M. Duan for helpful discus- sions. This work was supported by NFRP 2011CB921200 and 2011CBA00200, NNSF 60921091, and The Fun- damental Research Funds for the Central Universities WK2470000001. [1] M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010).5 [2] D. Xiao, M.-C. Chang, Q. Niu, Rev. Mod. Phys. 82, 1959 (2010). [3] Y.-J. Lin, R. L. Compton, A. R. Perry, W. D. Phillips, J. V. Porto, and I. B. Spielman, Phys. Rev. Lett. 102, 130401 (2009). [4] Y.-J. Lin, K. Jim enez-Garc a and I. B. Spielman, Nature (London) 471, 83 (2011). [5] R. A. Duine and H. T. C. Stoof, Phys. Rep. 396, 115 (2004). [6] S. Inouye, M. R. Andrews, J. Stenger, H.-J. Miesner, D. M. Stamper-Kurn and W. Ketterle, Nature (London) 392, 15 (1998). [7] Q.-J. Chen, J. Stajic, S. Tan and K. Levin, Phys. Rep. 412, 1 (2005). [8] D. E Sheehy and L. Radzihovsky, Ann. Phys. 322, 1790 (2007). [9] G.-B. Jo, Y.-R. Lee, J.-H. Choi, C. A. Christensen, T. H.Kim, J. H. Thywiseen, D. E. Pritchard and W. Ketterle, Science 325, 1521 (2009). [10] S. Tewari et al. , New J. Phys. 1, 065004 (2011). [11] J. P. Vyasanakere and V. B. Shenoy, Phys. Rev. B 83, 094515 (2011). [12] J. P. Vyasanankere, S. Zhang, and V. B. Shenoy, Phys. Rev. B 84, 014512 (2011). [13] M. Gong, S. Tewari and C. Zhang, arXiv:1105.1796. [14] Z.-Q. Yu and H. Zhai, arXiv:1105.2250. [15] H. Hu, L. Jiang, X.-J. Liu, H. Pu, arXiv:1105.2488. [16] M. Iskin and A. L. Subasi, Phys. Rev. Lett. 107, 050402 (2011). [17] W. Yi and L.-M. Duan, Phys. Rev. A 73, 031604(R) (2006). [18] W. V. Liu and F. Wilczek, Phys. Rev. Lett. 90, 047002 (2003). [19] M. M. Parish, F. M. Marchetti, A. Lamacraft and B. D. Simons, Nat. Phys. 3, 124 (2007).
2012.13914v1.Microwave_spectroscopy_of_spin_orbit_coupled_states__energy_detuning_versus_interdot_coupling_modulation.pdf
arXiv:2012.13914v1 [cond-mat.mes-hall] 27 Dec 2020Microwave spectroscopy of spin-orbit coupled states: ener gy detuning versus interdot coupling modulation G. Giavaras1and Yasuhiro Tokura1,2 1Faculty of Pure and Applied Sciences, University of Tsukuba , Tsukuba 305-8571, Japan 2Tsukuba Research Center for Energy Materials Science (TREM S), Tsukuba 305-8571, Japan We study the AC field induced current peaks of a spin blockaded double quantum dot with spin-orbit interaction. The AC field modulates either the in terdot tunnel coupling or the energy detuning, and we choose the AC field frequency range to induce two singlet-triplet transitions giving rise to two current peaks. We show that for a large detu ning the two current peaks can be significantly stronger when the AC field modulates the tunnel coupling, thus making the detection of the spin-orbit gap more efficient. We also demonstrate the i mportance of the time dependence of the spin-orbit interaction. I. INTRODUCTION The singlet-triplet states of two electron spins in tunnel-coupled quantum dots can be used to define spin- qubits in semiconductor devices.1,2In the presence of a strongspin-orbitinteraction(SOI)anappliedACelectric field can give rise to singlet-triplet transitions and spin resonance can be achieved.3,4In a double quantum dot which is tuned to the spin blockade regime,5the transi- tions can be probed by the AC induced current peaks. It has been experimentally demonstrated that the two- spin energy spectra can be extracted by examining the magneticfielddependent positionofthecurrentpeaks.3,4 The exchange energy, the strength of the SOI, as well as theg-factors of the quantum dots can then be esti- mated. Microwavespectroscopy has also been performed for the investigation of charge qubits,6,7as well as other hybrid spin systems.8–10Charge localization in quantum dot systems can be controlled with AC fields,11–13while various important parameters of the spin and/or charge dynamics can be extracted from AC induced interference patterns.14–17 In this work, we study the current through a double dot(DD) fortwodifferentcasesoftheACelectricfield; in the first case the AC field modulates the interdot tunnel coupling of the DD, and in the second case the AC field modulates the energy detuning of the DD. We consider a specific energy configuration and AC frequency range which involve two SOI coupled singlet-triplet states, and a third state with (mostly) triplet character. The two SOI-coupled singlet-triplet states form an anticrossing point,3,4,18,19and in this work we focus on this point. Specifically, we perform electronic transport calcula- tions and demonstrate that for a large energy detuning the tunnel coupling modulation results in stronger AC- induced current peaks than the corresponding peaks in- duced by the detuning modulation. The stronger peaks canofferasignificantadvantagewhenthespectroscopyof the coupled spin system is performed by monitoring the magnetic field dependence of the position of the peaks. When the peaks are suppressed no reliable information can be extracted. The tunnel coupling modulation offers a similar ad-vantage when the transitions involve only the two states forming the anticrossing point.20This finding together with the results of the present work demonstrate that modulating the tunnel coupling of a DD with an AC electric field is a robust method to perform spectroscopy of spin-orbit coupled states. Furthermore, in the present work, we explore the time dependent role of the SOI, and specify the regime in which this time dependence should be taken into accountbecause it can drasticallyaffect the AC-induced current peaks. In some experimental works the interdot tunnel cou- pling has been accurately controlled and transport measurements have been performed.21,22For instance, Bertrand et al[Ref. 22] have demonstrated that the tun- nel coupling can be tuned by orders of magnitude on the nanosecond time scale. Therefore, our theoretical find- ings could be tested with existing semiconductor technol- ogy. II. DOUBLE QUANTUM DOT MODEL We focus on the spin blockade regime5for two serially tunnel-coupled quantum dots. In this regime the quan- tum dot 1 (dot 2) is coupled to the left (right) metallic lead and with the appropriate bias voltage current can flow through the DD when the blockade is partially or completely lifted. Each quantum dot has a single orbital level and dot 2 is lower in energy by an amount equal to the charging energy which is assumed to be much larger thanthetunnelcoupling. Consequently,fortheappropri- ate bias voltage a single electron can be localized in dot 2 during the electronic transport process.5If we use the notation (n,m) to indicate nelectrons on the dot 1 and melectronsonthedot2thenelectronictransportprocess throughthe DDtakesplaceviathe chargecycle: (0 ,1)→ (1,1)→(0,2)→(0,1). For the DD system there are in total six two-electron states but in the spin blockade regime the double occupation on dot 1 can be ignored because it lies much higher in energy and does not affect the dynamics. Therefore, the relevanttwo-electronstates are the triplet states |T+/angbracketright=c† 1↑c† 2↑|0/angbracketright,|T−/angbracketright=c† 1↓c† 2↓|0/angbracketright, |T0/angbracketright= (c† 1↑c† 2↓+c† 1↓c† 2↑)|0/angbracketright/√ 2 and the two singlet states2 |S02/angbracketright=c† 2↑c† 2↓|0/angbracketright,|S11/angbracketright= (c† 1↑c† 2↓−c† 1↓c† 2↑)|0/angbracketright/√ 2. The fermionic operator c† iσcreates an electron on dot i= 1, 2 with spin σ=↑,↓, and|0/angbracketrightdenotes the vacuum state. In this singlet-triplet basis the DD Hamiltonian20is HDD= ∆[|T−/angbracketright/angbracketleftT−|−|T+/angbracketright/angbracketleftT+|]−δ|S02/angbracketright/angbracketleftS02| −√ 2Tc|S11/angbracketright/angbracketleftS02|+∆−|S11/angbracketright/angbracketleftT0|+H.c. −Tso[|T+/angbracketright/angbracketleftS02|+|T−/angbracketright/angbracketleftS02|]+H.c.(1) Here,δis the energy detuning, Tcis the tunnel coupling between the two dots, and Tsois the SOI-induced tunnel coupling causing a spin-flip.23,24The magnetic field is denoted by Bwhich gives rise to the Zeeman splitting ∆i=giµBB(i= 1, 2) in each quantum dot. Then ∆ = (∆ 2+∆1)/2, and the Zeeman asymmetry is ∆−= (∆2−∆1)/2. To a good approximation, in the transport process of a spin-blockaded DD only the c† 2↑|0/angbracketright,c† 2↓|0/angbracketright single electron states are important, and HDDcan be also derived using a standard two-site Hubbard model.25 In the present work, we consider two cases for the AC field. Specifically, in the first case the AC field modu- lates the energy detuning of the DD, thus we consider the following time dependence: δ(t) =ε+Adsin(2πft), (2) whereAdis the AC amplitude and fis the AC frequency. Theconstantvalueofthedetuningisdenotedby ε. Inthe second case, the AC field modulates the interdot tunnel coupling, thus we assume the time dependent terms Tc(t) =tc+Absin(2πft), Tso(t) =tso+xsoAbsin(2πft).(3) The AC amplitude is Aband in general Ab/negationslash=Ad. For most calculations we assume that xso=tso/tc, so at any time theratio Tso(t)/Tc(t) isatimeindependent constant equaltoxso. We alsoaddressthe casewhere xso/negationslash=tso/tc, but for simplicity we assume no phase difference between the tunnel couplings Tc(t) andTso(t). For the numerical calculations the DD parameters are taken to be tc= 0.2 meV,tso= 0.02 meV,g1= 2 andg2= 2.4. The basic conclusions of this work are general enough and not specific to these numbers. III. RESULTS In this section we present the basic results of our work. Wedeterminethe ACinducedcurrentforeachcaseofthe two AC fields Eq. (2) and Eq. (3). The DD eigenenergies EisatisfyHDD|ψi/angbracketright=Ei|ψi/angbracketrightwithAd=Ab= 0, and are shown in Fig. 1(a) at B= 1 T. When tso= 0 andg1=g2 singlet and triplet states are uncoupled. The energy lev- elsE2,E3andE4correspond to the pure triplet states |T+/angbracketright,|T0/angbracketrightand|T−/angbracketrightrespectively. These levels are detun- ing independent and are Zeeman-split due to the applied magnetic field. The energy levels E1,E5correspond to-0.6-0.4-0.2 0 0.2 0 0.5 1 1.5E (meV) ε (meV)E1E2E3E4E5(a) 0 0.01 0.02 0 0.5 1 1.5 2 2.5 3 0 2 4 6∆so (meV) ∆so (GHz) ε (meV)* *(b) FIG. 1: (a) Two-electron eigenenergies as a function of the energydetuningfor themagnetic field B= 1T. The levels E4, E5anticross at ε= 0.5 meV due to the spin-orbit interaction. The two vertical arrows indicate possible transitions whic h can be induced by the AC fields defined in the main text Eq. (2) and Eq. (3). (b) Spin-orbit gap (∆ so=E5−E4at the anticrossing point) as a function of the detuning. In this ca se the magnetic field is detuning dependent. The AC induced current is computed for the marked points. pure singlet states which are |S11/angbracketright,|S02/angbracketrighthybridized due to the tunnel coupling tcand are independent of the field as can be seen from HDD. Importantly, the singlet levels E1,E5define a two-level system and for a fixed tcthe hybridization is controlled by the energy detuning. The two levelsE1,E5anticross at ε= 0 where the hybridiza- tion is maximum. This is the only anticrossing point in the energy spectrum for tso= 0. However, according to HDDwhentso/negationslash= 0 the polarized triplets |T±/angbracketrightcouple to the singlet state |S02/angbracketright. Therefore, as seen in Fig. 1, at ε≈0.5 meV the levels E4andE5form an anticrossing point due to the SOI. Another SOI-induced anticross- ing point is formed at ε <0 between the energy levels E1andE2, but here we consider ε >0 and as in the experiments3,26–28we taketso<tc. BecauseoftheSOI( tso/negationslash= 0)andthedifferenceinthe g- factors (g1/negationslash=g2) the DD eigenstates |ψi/angbracketrightare hybridized singlet-triplet states and can be written in the general form |ψi/angbracketright=αi|S11/angbracketright+βi|T+/angbracketright+γi|S02/angbracketright+ζi|T−/angbracketright+ηi|T0/angbracketright.(4)3 The coefficients denoted by Greek letters determine the character of the states, and are sensitive to the detuning. One method to probe the SOI anticrossing point is to focus on the AC frequency range 0 <hf/lessorsimilarE5−E4and determine the position of the AC induced current peak. This method has been theoretically studied in Ref. 20. Another method to probe the anticrossing point is to fo- cus on the AC frequency range E4−E2/lessorsimilarhf/lessorsimilarE5−E2, and determine the positions of the two AC induced cur- rentpeaks. The presentworkisconcernedwith the latter method and the main subject of the present work is to compare the current peaks induced separately by the two AC fields; the tunnel barrier modulation and energy de- tuningmodulation. InRef.3bothmethodshavebeenex- perimentally investigated under the assumption that the AC field modulates the energy detuning of the DD. The case where the AC field modulates simultaneouslythe in- terdot tunnel coupling and the energy detuning might be experimentally relevant,29but this case is not pursued in the present work. In Fig. 1 the SOI anticrossing is formed at ε≈0.5 meV forB= 1 T. A lower magnetic field shifts the SOI anticrossing point at larger detuning, and the degree of hybridization due to the SOI decreases. The reason is that asεincreases the |S02/angbracketrightcharacter in the original sin- glet state ( tso= 0) is gradually replaced by the |S11/angbracketright character. As a result, the SOI gap ∆ so=E5−E4, defined at the anticrossing point, decreases with εas shown in Fig. 1(b). For the parameters considered in this work, the SOI gap can be analytically determined from the expression30∆so= 2tso/radicalbig (1−cosθ)/2, with θ= arctan(2√ 2tc/ε). In ourpreviouswork20weexaminedthe transitionsbe- tween the two singlet-triplet states |ψ4/angbracketrightand|ψ5/angbracketright, whose energy levels form the SOI anticrossing point (Fig. 1). These transitions give rise to one current peak which is suppressedneartheanticrossingpoint, inagreementwith an experimental study.3In the present work, we focus on the transitions between the two pairs of states |ψ5/angbracketrightand |ψ2/angbracketrightas well as |ψ4/angbracketrightand|ψ2/angbracketright. Here, |ψ4/angbracketrightand|ψ5/angbracketrightare strongly hybridized singlet-triplet states, whereas |ψ2/angbracketright has mostly triplet character provided the detuning is large. We compute the AC-induced current flowing through the double dot within the Floquet-Markov density ma- trix equation of motion.31,32In this approach we treat the time dependence of the AC field exactly, taking ad- vantage of the fact that the DD Hamiltonian is time pe- riodic and thus it can be expanded in a Fourier series. The model uses for the basis states of the DD density matrix the periodic Floquet modes,33and consequently it is applicable for any amplitude of the AC field. In most calculations we take the parameter xso= 0.1 unless otherwise specified. To study the AC current spectra we choose two values fortheenergydetuning ε(2, 0.5meV),anddeterminethe magnetic field at which |ψ4/angbracketrightand|ψ5/angbracketrightanticross. At this specific field we plot in Fig. 2 the AC-induced current as 0.05 0.1 0.15 0.2 0.25 17 17.5 18 18.5 19 19.5 20Ι (pA) f (GHz)Ab = 10 µeV Ad = 10 µeV (a) 0.4 0.8 1.2 1.6 58 59 60 61 62 63 64 65Ι (pA) f (GHz)Ab = 10 µeV Ad = 10 µeV (b) FIG. 2: Current as a function of AC frequency, when the AC field modulates the tunnel barrier with the AC amplitude Ab= 10µeV, and the energy detuning with Ad= 10µeV. The constant value of the detuning is ε= 2, 0.5 meV and the corresponding magnetic field is B= 0.3, 1 T for (a) and (b) respectively. These fields define the singlet-triplet antic ross- ing point for each value of the detuning. a function of the AC frequency. As the energy detuning decreases the magnetic field defining the corresponding anticrossing point increases. This in turn means that the AC frequencyhasto increaseto satisfythe corresponding resonance condition hf=E5−E2(orhf=E4−E2). This increase in the frequency explains the different fre- quency range in Fig. 2. Furthermore, the off-resonant current is larger for ε= 0.5 meV due to the stronger SOI hybridization.25 In the two cases shown in Fig. 2 two peaks are formed; one peak is due to the transition between the eigenstates |ψ2/angbracketrightand|ψ4/angbracketright, and the second peak is due to the transi- tion between |ψ2/angbracketrightand|ψ5/angbracketright. Therefore, the distance be- tween the centres of the two peaks is equal to the singlet- triplet energy splitting E5−E4. For the specific choice of magnetic field this energy splitting is equal to the SOI gap of the anticrossing point. For example, for ε= 0.5 meV the gap is ∆ so≈3.5 GHz, and for ε= 2 meV the gap is ∆ so≈1.1 GHz. These numbers are in agreement with those derived from the exact energies of the time independent part of the Hamiltonian HDD. According to Fig. 2, for a given energy detuning and driving field the two peaks are almost identical. This is due to the fact, that at the anticrossing point the states |ψ4/angbracketright,|ψ5/angbracketrighthave identical characters when the driving field is off, and the relevant transition rates are almost equal. In contrast,4 0 0.5 1 1.5 2 B (T) 0.5 1 1.5 2ε (meV) 0 0.3 0.6 qb (µeV)(b) 0 0.5 1 1.5 2 0.5 1 1.5 2ε (meV) 0 0.25 0.5 qd (µeV)(a) FIG. 3: (a) Absolute value of the coupling parameter qdas a function of the energy detuning and magnetic field for the AC amplitude Ad= 10µeV. The dotted curve defines the anticrossing point for each εandB. (b) The same as (a) but forqbwithAb= 10µeV. 0 10 20 30 40 0 0.5 1 1.5 2 2.5 3 3.5qb/qd ε (meV)0.00.020.040.060.080.1xso = 0.12 FIG. 4: The ratio qb/qddefined in Eq. (9) as a function of detuning for different values of xsoandAd=Ab. away from the anticrossing point the two peaks can be very different.33 The results in Fig. 2 demonstrate that the two driving fields Eq. (2) and Eq. (3) induce different peak magni- tudes. Specifically, the peaks due to the tunnel barrier modulation are stronger than those due to the detuningmodulation. As an example, for ε= 0.5 meV [Fig. 2(b)] the tunnel barrier modulation induces a relative peak height of about 1 pA, whereas the relative peak height is only 0.1 pA for the energy detuning modulation. Some insight into this interesting behavior can be obtained by inspecting the time-scale (“Rabi” frequency) of the co- herent transitions between the eigenstates |ψ2/angbracketrightand|ψ4/angbracketright. When the AC field modulates the tunnel coupling, the transitionscanbestudiedwithintheexactFloqueteigen- value problem, but for simplicity we here employ an ap- proximate approach.20This two-level approach gives the transition frequency qb/h, with qb=hfhb 24 (hb 22−hb 44)J1/parenleftbiggAb(hb 22−hb 44) hf/parenrightbigg ,(5) whereJ1(r) is a Bessel function of the first kind, and the argument is r=Ab(hb 22−hb 44)/hf, withhf=E4−E2 and hb ij=−γj(√ 2αi+xsoβi+xsoζi) −γi(√ 2αj+xsoβj+xsoζj), i,j= 2,4(6) When the AC field modulates the energy detuning the time-scale of the coherent transitions between the eigen- states|ψ2/angbracketrightand|ψ4/angbracketrightis approximately qd/h. The cou- pling parameter qdis found by qbwith the replacements Ab→Adandhb ij→hd ij, where hd ij=−γiγj, i,j= 2,4 (7) In general, qbcan be very different from qd, even when Ab=Ad. Therefore, the two driving fields are expected to induce current peaks with different width and height. To quantify the two parameters qb,qdwe plot in Fig. 3 qb,qd, as a function of the energy detuning and the mag- netic field. Here, Ad=Ab= 10µeV, andhf=E4−E2 [in Eq. (5)] is magnetic field as well as detuning depen- dent, and is determined by the energies of HDD. If we denote byBanthe field at which the anticrossing point is formed, then as seen in Fig. 3 both qbandqdare largefor B > B an, but vanishingly small for B≪Ban. The rea- son is that the state |ψ4/angbracketrightis singlet-like for B >B an, but triplet-like for B <B an, whereas |ψ2/angbracketrighthas mostly triplet characterindependent of B, providedεisawayfromzero. Transitions between triplet-like states are in general slow leading to vanishingly small qb,qdforB≪Ban. In contrast, if we choose hf=E5−E2, then both qband qdare large for B < B an. For large enough detuning where the two spins are in the Heisenberg regime, the exchange energy is approximately 2 t2 c/εthereforeBan satisfies (g1+g2)µBBan/2≈2t2 c/ε. Most importantly Fig. 3 demonstrates that qb> qd whenε/greaterorsimilar0.2 meV. To understand this result we focus on the anticrossing point where r<1, then from Eq. (5) qb≈hb 24Ab/2 becauseJ1(r)≈r/2, and similarly qd≈ hd 24Ad/2. Moreover, away from zero detuning the state |ψ2/angbracketrighthas mostly triplet character, therefore hb 24≈ −γ4(√ 2α2+xsoβ2)−γ2(√ 2α4+xsoζ4),(8)5 and the ratio qb/qdis qb qd=Ab Ad/parenleftbigg√ 2α2 γ2+xsoβ2 γ2+√ 2α4 γ4+xsoζ4 γ4/parenrightbigg .(9) Asεincreasesβ2→1,γ2≪1 and, considering absolute values, the second term in Eq. (9) dominates β2 γ2≫α2 γ2,α4 γ4,ζ4 γ4. (10) Consequently, qbcan be much greater than qd, especially at largeε, and for a fixed tunnel coupling tcthe exact value of the ratio qb/qddepends sensitively on xso. This demonstrates the importance of the time dependence of the spin-orbit coupling. The conclusions derived from the parameters qb,qdassume that there is no ‘multi- level’ interference and only the levels Ei,Ejsatisfying hf=|Ei−Ej|are responsible for the current peaks. The approximate results are more accurate when the ar- gumentrof the Bessel function is kept small. To examine the xso-dependence, we consider Ab=Ad and plot in Fig. 4 the ratio qb/qdversus the detuning at the anticrossing point, and for different values of xso. By increasing εand for large values of xsothe coupling parameters qb,qdcan differ by over an order of mag- nitude;qb/qd>10. This leads to (very) different cur- rent peaks with the tunnel barrier modulation inducing stronger peaks. The special value xso= 0 corresponds to atimeindependent SOItunnelcoupling[seeEq.(3)], and thespecialvalue xso= 0.1correspondstoatimeindepen- dent ratioTso/Tc= 0.1. Although, the ratio qb/qdcan be computed at any ε, the regime of small ε(<0.2 meV) is not particularly interesting in this work. The reason is that with decreasing εthe character of the state |ψ2/angbracketright changes from triplet-like to singlet-triplet, which even- tually becomes approximately equally populated to |ψ4/angbracketright and|ψ5/angbracketright. Therefore, the current peaks induced by both driving fields are suppressed even when qborqdis large. In Fig. 4 the maximum value of the detuning is chosen to giveε/tc≈17.5 which can be easily achieved in dou- ble quantum dots. Some experiments1,3,4have reported values greater than ε/tc≈100, thusqbcan be even two orders of magnitude greater than qd. According to the above analysis if qb/qd≈1 then the current peaks induced by the two driving fields should approximately display the same characteristics. As an example, consider the two sets of current peaks shown in Fig. 2 both for xso= 0.1 andε= 2 meV, ε= 0.5 meV respectively. Focusing on xso= 0.1 in Fig. 4, we see that at ε= 2 meVqb/qd≈19 and atε= 0.5 meV qb/qd≈4.9. These numbers suggestthat if at ε= 2 meV we choose for the AC amplitudes the ratio Ab/Ad≈1/19 then the detuning and the barrier modulation should in- duce approximately the same peak characteristics. Like- wise atε= 0.5 meV the ratio should be Ab/Ad≈1/4.9. These arguments are quantified in Fig. 5 where we plot the current peaks for the two driving fields for different AC amplitudes satisfying the condition qb/qd≈1. The 0.05 0.1 0.15 0.2 0.25 17 17.5 18 18.5 19 19.5 20Ι (pA) f (GHz)Ad = 190 µeV Ab = 10 µeV (a) 0.4 0.8 1.2 1.6 58 59 60 61 62 63 64 65Ι (pA) f (GHz)Ad = 49 µeV Ab = 10 µeV (b) FIG. 5: As in Fig. 2, but (a) Ab= 10µeV andAd= 19Ab, (b)Ab= 10µeV andAd= 4.9Ab. The value of Adis chosen so that to approximately induce the same current peaks as those induced by Ab. 0.05 0.1 0.15 0.2 0.25 17.5 18 18.5 19 19.5Ι (pA) f (GHz) FIG. 6: Current as a function of AC frequency, when the AC field modulates the tunnel barrier, with the AC amplitude Ab= 10µeV. The detuning is ε= 2 meV and from the upper to the lower curve the parameter xso= 0.1, 0.04, 0.02, 0. results confirm that the induced current peaks display approximately the same characteristics. Inducing strong current peaks can be advantageous in order to perform spectroscopy of the singlet-triplet levels and extract the SOI anticrossing gap. However, an im- portant aspect is that the SOI gap cannot be extracted from the positions of the current peaks at arbitrary large AC amplitudes. In particular, by increasing the AC am- plitude the two peaks start to overlap and eventually6 the resonant pattern of the current changes drastically.33 Therefore, the distance between the two peaks cannot accurately predict the SOI gap. This effect has been theoretically studied for the case of a time dependent energy detuning,33and it can be readily shown that sim- ilar trends occur for a time dependent tunnel coupling. The driving regime where the two current peaks strongly overlap is not considered in the present work, since it is not appropriate for the spectroscopy of the SOI gap. Finally, in Fig. 6 we plot the current peaks when the AC field modulatesthe tunnel barrierwith the amplitude Ab= 10µeVandtheconstantdetuning ε= 2meV.With decreasing xsothe two peaks gradually weaken and for xso= 0 the peaks arevanishingly small; for this value the peaks are of the same order as the peaks induced by the detuning modulation with the same amplitude Ad= 10 µeV (for clarity these peaks are not shown). The small difference between the left and the right peaks, for exam- ple whenxso= 0.02,can be understood byinspecting the different values of qb[Eq. (5)] which involvedifferent ma- trixelementsandfrequencies. Theoveralltrendsindicate the important role of the time dependent spin-orbit term andareconsistentwiththeresultsshowninFig.4. As xso decreases the coupling parameter qbdecreases too, thus the time scale of the singlet-triplet transitions becomes longer leading to smaller peaks. Moreover, by decreasing xso,qbbecomes approximately equal to qd, therefore the tunnel barrier modulation and the detuning modulationresult in approximately the same current peaks. IV. SUMMARY In summary, we considered a double quantum dot in the spin blockade regime and studied the AC induced current peaks for a specific energy configuration which involves two hybridized singlet-triplet states as well as a third state with mostly triplet character. The two AC induced transitions which rely on the spin-orbit in- teraction, result in two current peaks. We found that for a large energy detuning the two peaks are stronger when the time periodic field modulates the interdot tun- nel coupling (barrier)instead ofthe energy detuning. We demonstrated that a time dependence in the spin-orbit coupling can significantly modify the peak characteris- tics, and should be taken into account even when the actual spin-orbit coupling is small. Our work suggests an efficient way of probing the spin-orbit energy gap in two-spin states based on transport measurements. ACKNOWLEDGEMENT Part of this work was supported by CREST JST (JP- MJCR15N2), and by JSPS KAKENHI (18K03479). 1J. R. Petta, A. C. Johnson, J. M. Taylor, E. A. Laird, A. Yacoby, M. D. Lukin, C. M. Marcus, M. P. Hanson, and A. C. Gossard, Science 309, 2180 (2005). 2J. J. L. Morton, and B. W. Lovett, Annu. Rev. Condens. Matter Phys. 2, 189 (2011). 3K. Ono, G. Giavaras, T. Tanamoto, T. Ohguro, X. Hu, and F. Nori, Phys. Rev. Lett. 119, 156802 (2017). 4S. Nadj-Perge, V. S. Pribiag, J. W. G. van den Berg, K. Zuo, S. R. Plissard, E. P. A. M. Bakkers, S. M. Frolov, and L. P. Kouwenhoven, Phys. Rev. Lett. 108, 166801 (2012). 5K. Ono, D.G. Austing, Y. Tokura, andS. Tarucha, Science 297, 1313 (2002). 6T. Hayashi, T. Fujisawa, H. D. Cheong, Y. H. Jeong, and Y. Hirayama, Phys. Rev. Lett. 91, 226804 (2003). 7Z.V.Penfold-Fitch, F.Sfigakis, andM. R.Buitelaar, Phys. Rev. Applied 7, 054017 (2017). 8B. C. Wang, G. Cao, H. O. Li, M. Xiao, G. C. Guo, X. Hu, H. W. Jiang, and G. P. Guo, Phys. Rev. Applied 8, 064035 (2017). 9X. C. Yang, G. X. Chan, and X. Wang, Phys. Rev. A 98, 032334 (2018). 10W. Song, T. Du, H. Liu, M. B. Plenio, and J. Cai, Phys. Rev. Applied 12, 054025 (2019). 11E. Paspalakis and A. Terzis, J. Appl. Phys. 95, 1603 (2004). 12A. Terzis and E. Paspalakis, J. Appl. Phys. 97, 023523 (2005). 13C. E. Creffield and G. Platero, Phys. Rev. B 65, 113304 (2002).14S. N. Shevchenko, S. Ashhab, and F. Nori, Phys. Rep. 492, 1 (2010). 15S. N. Shevchenko, S. Ashhab, and F. Nori, Phys. Rev. B 85, 094502 (2012). 16F. Gallego-Marcos, R. Sanchez, and G. Platero J. Appl. Phys.117, 112808 (2015). 17A. Chatterjee, S. N. Shevchenko, S. Barraud, R. M. Otxoa, F. Nori, J. J. L. Morton, and M. F. Gonzalez-Zalba, Phys. Rev. B97, 045405 (2018). 18S. Takahashi, R. S. Deacon, K. Yoshida, A. Oiwa, K. Shi- bata, K. Hirakawa, Y. Tokura, and S. Tarucha, Phys. Rev. Lett.104, 246801 (2010). 19Y. Kanai, R. S. Deacon, S. Takahashi, A. Oiwa, K. Yoshida, K. Shibata, K. Hirakawa, Y. Tokura, and S. Tarucha, Nature Nanotechnology 6, 511 (2011). 20G. Giavaras and Y. Tokura, Phys. Rev. B 99, 075412 (2019). 21T. H. Oosterkamp, T. Fujisawa, W. G. van der Wiel, K. Ishibashi, R. V. Hijman, S. Tarucha, and L. P. Kouwen- hoven, Nature (London) 395, 873 (1998). 22B. Bertrand, H. Flentje, S. Takada, M. Yamamoto, S. Tarucha, A. Ludwig, A. D. Wieck, C. Buerle, and T. Me- unier, Phys. Rev. Lett. 115, 096801 (2015). 23F. Mireles and G. Kirczenow, Phys. Rev. B 64, 024426 (2001). 24H. Pan and Y. Zhao, J. Appl. Phys. 111, 083703 (2012). 25G. Giavaras, N. Lambert, and F. Nori, Phys. Rev. B 87, 115416 (2013). 26S. J. Chorley, G. Giavaras, J. Wabnig, G. A. C. Jones, C.7 G. Smith, G. A. D. Briggs, and M. R. Buitelaar, Phys. Rev. Lett. 106, 206801 (2011). 27G. Xuet al., Appl. Phys. Express 13, 065002 (2020). 28M. Marx , J. Yoneda, T. Otsuka, K. Takeda, Y. Yamaoka, T. Nakajima, S. Li, A. Noiri, T. Kodera, and S. Tarucha, Jpn. J. Appl. Phys. 58, SBBI07 (2019). 29T. Nakajima, M. R. Delbecq, T. Otsuka, S. Amaha, J. Yoneda, A. Noiri, K. Takeda, G. Allison, A. Ludwig, A. D. Wieck, X. Hu, F. Nori, and S. Tarucha, Nature Communi-cations9, 2133 (2018). 30D. Stepanenko, M. Rudner, B. I. Halperin, and D. Loss, Phys. Rev. B 85, 075416 (2012). 31M. Grifoni and P. H¨ anggi, Phys. Rep. 304, 229 (1998). 32S. Kohler, J. Lehmann, and P. H¨ anggi, Phys. Rep. 406, 379 (2005). 33G. Giavaras and Y. Tokura, Phys. Rev. B 100, 195421 (2019).
2003.13181v1.Intrinsic_orbital_moment_and_prediction_of_a_large_orbital_Hall_effect_in_the_2D_transition_metal_dichalcogenides.pdf
Intrinsic orbital moment and prediction of a large orbital Hall e ect in the 2D transition metal dichalcogenides Sayantika Bhowaland S. Satpathy Department of Physics & Astronomy, University of Missouri, Columbia, MO 65211, USA Carrying information using generation and detection of the orbital current, instead of the spin current, is an emerging eld of research, where the orbital Hall e ect (OHE) is an important ingredi- ent. Here, we propose a new mechanism of the OHE that occurs in non-centrosymmetric materials. We show that the broken inversion symmetry in the 2D transition metal dichalcogenides (TMDCs) causes a robust orbital moment, which ow in di erent directions due to the opposite Berry cur- vatures under an applied electric eld, leading to a large OHE. This is in complete contrast to the inversion-symmetric systems, where the orbital moment is induced only by the external electric eld. We show that the valley-orbital locking as well as the OHE both appear even in the absence of the spin-orbit coupling. The non-zero spin-orbit coupling leads to the well-known valley-spin locking and the spin Hall e ect, which we nd to be weak, making the TMDCs particularly suitable for direct observation of the OHE, with potential application in orbitronics . Orbital Hall e ect (OHE) is the phenomenon of trans- verse ow of orbital angular momentum in response to an applied electric eld, similar to the ow of spin angu- lar momentum in the spin Hall e ect (SHE). The OHE is more fundamental in the sense that it occurs with or without the presence of the spin-orbit coupling (SOC), while in presence of the SOC, OHE leads to the addi- tional ow of the spin angular momentum resulting in the SHE. In fact, the idea of OHE has already been in- voked to explain the origin of a large anomalous and spin Hall e ect in several materials [1{3]. Because of this and the fact that OHE is expected to have a larger magnitude than its spin counterpart, there is a noticeable interest in developing the OHE [4{7], with an eye towards future \orbitronics" device applications. In this work, we propose a new mechanism of the OHE that occurs in non-centrosymmetric materials and explic- itly illustrate the ideas for monolayer transition metal dichalcogenides (TMDCs) which constitute the classic example of 2D materials with broken inversion symme- try. In complete constrast to the centrosymmetric mate- rials [4, 6], where orbital moments are quenched due to symmetry and a non-zero moment develops only due to the symmetry-breaking applied electric eld, here an in- trinsic orbital moment is already present in the Brillouin zone (BZ) even without the applied electric eld. Unlike the centrosymmetric systems, the physics here is dom- inated by the non-zero Berry curvatures, which deter- mines the magnitude of the OHE. Our work emphasizes the intrinsic nature of orbital transport in contrast to the valley Hall e ect [8{12], for example, which can only be achieved by extrinsic means (doping, light illumination, etc.). We develop the key physics of the underlying mecha- nism of the OHE using a tight-binding (TB) model as well as from density-functional calculations. The e ect is demonstrated for the selected members of the family bhowals@missouri.edu MXxy X(a) (b) x2-y2+i xy x2-y2-i xy 3z2-1 3z2-1 K (t = +1) K' (t = -1) MX2(d) K K' Eky kxkz=0Mz < 0 Mz >0E v =(e/ħ) E X Ωv v Mz < 0 Mz >0a 10 5 0 -5 -10(c) FIG. 1. Illustration of OHE in monolayer MX 2. (a) Crystal structure of MX 2, showing the triangular network of transi- tion metal M atoms as viewed from top. The two out-of-plane chalcogen atoms X occur above and below the plane. (b) The band structure near K(4=3a;0) andK0(4=3a;0), show- ing the valley dependent spin and orbital characters. (c) The orbital moment Mz(~k) in the BZ and the anomalous veloci- tiesv, indicated by the blue and the red arrows. (d) Orbital moments ow in the transverse direction leading to the OHE. of monolayer TMDCs, viz., 2H-Mo X2(X= S, Se, Te), where we nd a large OHE and at the same time a negli- gible intrinsic spin Hall e ect, making these materials an excellent platform for the direct observation of the OHE. The basic physics is illustrated in Fig. 1, where we have shown the computed intrinsic orbital moments in the BZ as well as the electron \anomalous" velocities at the K, K0valleys. Symmetry demands that in the presence of inversion (I), orbital moments satisfy the condition ~M(~k) =~M(~k), while if time-reversal ( T) symmetry is present, we have ~M(~k) =~M(~k). Thus for a non-zero ~M(~k), at least one of the two symmetries must be broken.arXiv:2003.13181v1 [cond-mat.mtrl-sci] 30 Mar 20202 In the present case, broken Ileads to a nonzero ~M(~k), while its sign changes between the KandK0points due to the presence of T. The Berry curvatures ~ (~k) follow the same symmetry properties leading to the non-zero anomalous velocity ~ v= (e=~)~E~ ~k[13] which has op- posite directions at the two valleys, and thus leads to the OHE. These arguments are only suggestive, and one must evaluate the magnitude of the e ect from the cal- culation of the orbital Berry curvatures [13] as discussed below. Tight Binding results near the valley points { The val- ley points ( K/K0) have the major contributions to the OHE in the TMDCs and this can be studied analytically using a TB model. Due to the broken I[see Fig. 1 (a)], the chalcogen atoms must be kept along with the transi- tion metal atom (M) in the TB basis set; However, their e ect may be incorporated via the L owdin downfolding [14] producing an e ective TB Hamiltonian for the M- dorbitals with modi ed Slater-Koster matrix elements [15]. The e ective Hamiltonian, valid near the KandK0 valley points reads H(~ q) = (~d~ ) Is+ 2(z+ 1) sz; (1) where only terms linear in ~ q=~k~Khave been kept, ignoring thereby the higher-order trigonal warping [16], which are unimportant for the present study. Here ~ sand ~ are respectively the Pauli matrices for the electron spin and the orbital pseudo-spins, jui= (p 2)1(jx2y2i+ ijxyi) andjdi=j3z2r2i.Isis the 22 identity operator in the electron spin space, is the SOC con- stant, and the valley index =1 for theKandK0 valleys, respectively. The TB hopping integrals appear in the parameter ~d, withdx=tqxa;dy=tqya;and dz==2, whereais the lattice constant,  is the energy gap at the K(K0) point, and tis an e ective inter-band hopping, determined by certain ddhopping matrix elements. We note that Eq. (1) is consistent with the Hamiltonian derived earlier [10] using the kpthe- ory. The TB derivation has the bene t that it directly expresses the parameters of the Hamiltonian in terms of the speci c hopping integrals. The magnitude of the orbital moment ~M(~k) can be computed for a speci c band of the Hamiltonian (1) using the modern theory of orbital moment [13, 17], viz., ~M(~k) =21Im[h~rku~kj(H"~k)j~rku~ki] + Im[h~rku~kj(F"~k)j~rku~ki]; (2) where"~kandu~kare the band energy and the Bloch wave function, and the two terms in (2) are, respectively, the angular momentum ( ~ r~ v) contribution due to the self- rotation and due to the motion of the center-of-mass of the Bloch electron wave packet. Diagonalizing the 4 4 Hamiltonian (1), we nd the energy eigenvalues: " = 21[(()2+ 4t2a2q2)1=2], where=1 are the two spin-split states within the conduction or valence band manifold, denoted by the subscript . The wave K'(a) (b)3z2-r2 x2-y2 - i xyxz + i yz xz - i yz x2-y2 + i xyM zDFTTB ModelE F x2-y2 - i xy3z2-r2 DEnergy (eV)M z (eV Å2 )FIG. 2. (a) Density-functional band structure together with the orbital characters near the valley points and (b) the com- puted sum of the orbital moments ( Mz) over all occupied bands along selected symmetry lines. functions in the basis set ( ju"i;jd"i;ju#i;andjd#i) are ju=1 (q)i=Nh 1 (Dp (D)2+d2)=d0 0iT ; ju=1  (q)i=Nh 0 0 1 (Dp (D)2+d2)=diT (3) whereD= ()=2,d=ta(qxiqy),d2= t2a2(q2 x+q2 y), andNis the appropriate normalization factor. With these wave functions, the orbital moments can be evaluated exactly within the TB model from Eq. 2. For the two valence bands ( =1), the result is Mz(~ q) =m0D(D) 2[(D)2+t2q2a2]3=2(4a) m0[1 +(32)=](16m0q2=); (4b) wherem0= 1t2a2, only the out-of-plane ^ zcomponent of the orbital moment is non-zero, and the second line is the expansion for small qand, both. Note the important result (4) that a large orbital mo- mentMzexists at the valley points ( ~ q= 0) and its sign al- ternates between the two valleys ( =1) (valley-orbital locking ). Furthermore, it exists even in absence of the SOC (= 0). For typical parameters, t= 1:22 eV,  = 1.66 eV, and = 0:08 eV, relevant for the monolayer MoS 2,m09:1 eV. A22:4B(~=e). As seen from Eq. 4 (b), there is only a weak dependence on . In fact it is interesting to note that the valley- dependent spin splitting [Fig. 1 (b)] directly follows from the valley orbital moments due to the h~L~Siterm, which favors anti-alignment of spin with the orbital moment [3]. Thus for the valence bands, the spin- #band is lower in energy atK, while the spin-"band is lower at K0, with a spin splitting of about 2 . Therefore, the well-known spin polarization of the bands at the valley points can be thought of to be driven by the robust orbital moments via the perturbative SOC.3 The orbital moment is the largest at the valley points K;K0, as seen from Eq. (4), falling o quadratically with momentum q. This is also validated by the DFT results shown in Fig. 2. The orbital moment at the center of the BZ () vanishes exactly due to symmetry reasons, and therefore is expected to be small in the neighborhood of as seen from Fig. 2 (b) as well. It is easy to argue that under an applied electric eld, the electrons in the two valleys move in opposite direc- tions, so that a net orbital Hall current is produced. To see this, we rst realize that only the Berry curvature term in the semi-classical expression [13] for the electron velocity _~ rc=~1[~rk"k+e~E~ (~k)]~kcis non-zero for the two valleys. Furthermore, only the ^ zcomponent of the Berry curvature survives, which we evaluate near the K;K0valleys within the TB model using the Kubo for- mula below. The result is z n(~ q) =2~2X n06=nIm hun~ qjvxjun0~ qihun0~ qjvyjun~ qi ("n0~ q"n~ q)2 =2Mz(~ q)  +(2)2m0 2( + 26m0q2):(5) Clearly, zhas opposite signs for the two valleys, so that ~ v/~E~ is in opposite directions for the Kand the K0valley electrons. Thus the positive orbital moment of theKvalley moves in one direction, while the negative orbital moment of K0moves in the opposite direction, leading to a net orbital Hall current. The magnitude of the orbital Hall conductivity (OHC) may be calculated using the Kubo formula by the mo- mentum sum of the orbital Berry curvatures [4, 6], viz.,  ;orb =e NkVcoccX n~k ;orb n; (~k); (6) where ; ; are the cartesian components, jorb; =  ;orb E is the orbital current density along the direc- tion with the orbital moment along , generated by the electric eld along the direction. In the 2D systems, Vc is the surface unit cell area, so that the conductivity has the dimensions of ( ~=e) Ohm1. The orbital Berry curvature ;orb n; in Eq. 6 can be evaluated as ;orb n; (~k) = 2 ~X n06=nIm[hun~kjJ ;orb jun0~kihun0~kjv jun~ki] ("n0~k"n~k)2; (7) where the orbital current operator is J ;orb =1 2fv ;L g, withv =1 ~@H @k is the velocity operator and L is the orbital angular momentum operator. It turns out that due to the simplicity of the TB Hamil- tonian (1), valid near the valley points, the orbital and the standard Berry curvatures are the same, apart from a valley-dependent sign, viz., z;orb n;yx(~ q) = z n(~ q): (8)To see this, we take the momentum derivative of (1) to get ~vx(~ q) =2 640ta 0 0 ta 0 0 0 0 0 0 ta 0 0ta 03 75=tax Is;(9) and, similarly, ~vy(~ q) =tay Isandvz(~ q) = 0. Furthermore, in the subspace of the TB Hamiltonian, Lx=Ly= 0, andLz=~(z+ 1) Is. By matrix mul- tiplication, we immediately nd that Jz;orb =~v and Jx;orb =Jy;orb = 0, which leads to the result (8). The expression for the orbital Berry curvature then follows from Eqs. (5) and (8), viz., z;orb ;yx(~ q) =2Mz(~ q)  +(2); (10) whereMz(~ q) is the orbital moment in Eq. (4). At a generalkpoint, the full expression (7) must be evaluated to obtain the OHC. This is a key result of the paper, which shows that the orbital Berry curvatures near the KandK0points are directly proportional to the respective orbital moments, and, more importantly, they have the same sign at the two valleys as both andMzchange signs simultane- ously. Thus, the contributions from these two valleys add up, leading to a non-zero OHC. Another important point is that z;orb ;yx exists even without the SOC, and it has only a weak dependence on as seen from Eq. (10). Neglecting the dependence, we see that at both valley points, the contribution to the OHC is given by z;orb ;yx = 2t2a2=2. In fact, the momentum sum in OHC can be performed analytically in this limit by integrating up to the radius qc(q2 c= BZ) to yield the result z;orb yx =2e (2)2X =1Zqc 0d2q z;orb ;yx(~ q) =e h 1q 2+ (32t2=p 3)i +O(2=2); (11) which is consistent with the anticipated result that the larger the parameter t2=2, the larger is the OHC, pri- marily because the orbital moment Mzincreases. We pause here to compare the OHE with the related phenomenon of the valley Hall e ect, which has been pro- posed in the gapped graphene as well as in the TMDCs [11, 12]. In the valley Hall e ect, electrons in the two valleys ow in opposite directions, leading to a charge current and additionally to an orbital current (the valley orbital Hall e ect [12]), if there is a valley population im- balance (e.g., created by shining light). This is in com- plete contrast to the OHE, which is an intrinsic e ect without any need for population imbalance between the valleys. More interestingly, unlike the valley Hall e ect, the OHE described here does not have any net charge4 MK K'(a) K K' GM 1.00.50-0.5 1.5G -505101520(b) 00 FIG. 3. (a) Orbital and (b) spin Berry curvatures (in units of A2), summed over the occupied states, on the kz= 0 plane for 2H-MoS 2. The contours correspond to the tick values on the color bar and the zero contours have been indicated explicitly. current but there exists only a pure orbital current. Fur- thermore, in the valley Hall e ect, the non-zero valley orbital magnetization [11] explicitly breaks the Tsym- metry, which is preserved in the present case. In this sense the OHE studied here is completely di erent from the valley Hall e ect proposed earlier. Density functional results { We now turn to the DFT results for the monolayer TMDCs. Orbital moments were computed using pseudopotential methods [18] and the Wannier functions as implemented in the Wannier90 code [19, 20] [see Supplementary Materials [21] for details]. The complementary mun-tin orbitals based method (NMTO) [22] was used to compute the orbital moment as well as the orbital and the spin Hall conductivities. In the latter method, e ective TB hopping matrix ele- ments between the M- dorbitals are obtained for several neighbors, which yields the full TB Hamiltonian valid everywhere in the BZ, using which all quantities of inter- est are computed. The BZ sums for the OHC and spin Hall conductivity (SHC) were computed with 400 400 kpoints in the 2D zone. The computed orbital moments using the Wannier90 or the NMTO method agree quite well. The DFT band structure and the corresponding or- bital moments are shown in Fig. 1 (c) and Fig. 2 for TABLE I. DFT results for the OHC of the monolayer TMDCs, including the partial contributions ( z;orb yx =K++rest), K,, andrestbeing the contributions, respectively, from the valley, -point, and the remaining regions of the BZ. OHC are in units of 103(~=e) 1, while the SHC are in units of (~=e) 1. Materials Krestz;orb yxz;spin yx MoS 2 -9.1 1.7 -3.2 -10.6 1.0 MoSe 2 -8.0 1.7 -3 -9.3 1.8 MoTe 2 -9.1 1.1 -2.5 -10.5 3.0 WTe 2 -8.6 1.0 -2.6 -10.2 9.4MoS 2. As shown in Fig. 2 (b), the orbital moments com- puted from the Hamiltonian (1) near the valley points agree quite well with the DFT results. Note that the to- tal orbital moment (summed over the BZ) vanishes due to the presence of T, though it is non-zero at individual kpoints. From the TB model (1), we had studied the or- bital moment and the OHE near the valley points. From the DFT calculations, we can compute the same over the entire BZ, the result of which is shown in Fig. 3 (a). As seen from the gure, the dominant contribution comes from thekspace near the valley points K,K0. Since the intrinsic orbital moment near the point is absent, the orbital Berry curvature in this region takes a non- zero value only due to the orbital moments induced by the applied electric eld in the Hall measurement, similar to the centrosymmetric case [4]. This results in a small contribution to the net OHC, as seen from Table I, which lists the partial contributions to the OHC coming from di erent parts of the BZ. Note that there is only one independent component of OHC, viz., z;orb yx = -z;orb xy. Spin Hall E ect { For a material to be a good can- didate for the detection of the OHE, the SHC must be small, as both carry angular momentum. To this end, we compute the SHC, rst from the model Hamiltonian and then from the full DFT calculations. Analogous to the calculation of the OHC, the SHC can be obtained by the sum of the spin Berry curvatures, z;spin ;yx (~k), evaluated by replacing the orbital current operator with the spin current operator J ;spin =1 4fv ;s gin Eq. 7. For the two spin-split valence bands near the valley points in the TB model, we nd z;spin ;yx (~ q) =Mz(~ q)  +(2)= 2 z;orb yx(~ q):(12) Note that z;spin ;yx (~ q) has opposite signs for the two spin- split bands and in the limit of = 0, they exactly cancel everywhere producing a net zero SHC. For a non-zero , these two contributions add up to produce a small net SHC. Calculating the contributions from the valley points with a similar procedure as Eq. (11), we obtain the resultz;spin yxe()1;in the limit . This is clearly much smaller than the OHC (11), by a factor of =. From the DFT results (see Table I), we do indeed nd that the SHC is about three orders of magnitude smaller than the OHC. Even in doped samples though the SHC is expected to be higher than the undoped sam- ple, the typical values [8] are nevertheless still an order of magnitude smaller than the computed OHC. These ar- guments suggest the TMDCs to be excellent candidates for the observation of OHE, since the intrinsic SHC is negligible in comparison. In conclusion, we examined the intrinsic OHE in non- centrosymmetric materials and illustrated the ideas for the monolayer TMDCs. The broken Iin TMDCs pro- duces a robust momentum-space intrinsic orbital moment ~M(~k), present even in the absence of . Due to the op- posite Berry curvatures at the valley points KandK0, these orbital moments ow in opposite directions, leading5 to a large OHC (104~=e 1). The vanishingly small intrinsic SHC in these materials make them particularly suitable for the direct observation of the OHE, which can be measured by detecting the orbital torque generated by the orbital Hall current [5]. Magneto-optical Kerr e ect may also be used to detect the orbital moments accumu- lated at the edges of the sample due to the OHE [23]. Furthermore, the valley-orbital locking can be probed in photon polarized angle-resolved photoemission measure-ments [24]. In addition, it may be possible to tune the OHC by applying a transverse electric eld [25, 26]. Ex- perimental con rmation of the OHE in the TMDC's may open up new avenues for the realization of orbitronics de- vices. We thank the U.S. Department of Energy, Oce of Basic Energy Sciences, Division of Materials Sciences and Engineering for nancial support under Grant No. DEFG02-00ER45818. [1] T. Tanaka, H. Kontani, M. Naito, T. Naito, D. S. Hi- rashima, K. Yamada and J. Inoue, Intrinsic spin Hall e ect and orbital Hall e ect in 4d and 5d transition met- als, Phys. Rev. B 77, 165117 (2008). [2] H. Kontani, T. Tanaka, D. S. Hirashima, K. Yamada and J. Inoue, Giant Intrinsic Spin and Orbital Hall E ects in Sr 2MO4(M= Ru, Rh, Mo), Phys. Rev. Lett. 100, 096601 (2008). [3] H. Kontani, T. Tanaka, D. S. Hirashima, K. Yamada and J. Inoue, Giant Orbital Hall E ect in Transition Metals: Origin of Large Spin and Anomalous Hall E ects, Phys. Rev. Lett. 102, 016601 (2009). [4] D. Go, D. Jo, C. Kim, and H.-W. Lee, Intrinsic Spin and Orbital Hall E ects from Orbital Texture, Phys. Rev. Lett. 121, 086602 (2018). [5] D. Go and H.-W. Lee, Orbital Torque: Torque Gen- eration by Orbital Current Injection, arXiv:1903.01085 (2019). [6] D. Jo, D. Go and H.-W. Lee, Gigantic intrinsic orbital Hall e ects in weakly spin-orbit coupled metals, Phys. Rev. B 98, 214405 (2018). [7] V. T. Phong, Z. Addison, S. Ahn, H. Min, R. Agarwal, and E. J. Mele, Optically Controlled Orbitronics on a Triangular Lattice, Phys. Rev. Lett. 123, 236403 (2019). [8] W. Feng, Y. Yao, W. Zhu, J. Zhou, W. Yao, and D. Xiao, Intrinsic spin Hall e ect in monolayers of group-VI dichalcogenides: A rst-principles study, Phys. Rev. B 86, 165108 (2012). [9] B. T. Zhou, K. Taguchi, Y. Kawaguchi, Y. Tanaka, and K.T. Law, Spin-orbit coupling induced valley Hall ef- fects in transition-metal dichalcogenides, Communica- tions Physics 2, 26 (2019). [10] D. Xiao, G.-B. Liu, W. Feng, X. Xu, and W. Yao, Cou- pled Spin and Valley Physics in Monolayers of MoS 2and Other Group-VI Dichalcogenides, Phys. Rev. Lett. 108, 196802 (2012). [11] D. Xiao, W. Yao, and Q. Niu, Valley-Contrasting Physics in Graphene: Magnetic Moment and Topological Trans- port, Phys. Rev. Lett. 99, 236809 (2007). [12] Z. Song, R. Quhe, S. Liu, Y. Li, J. Feng, Y. Yang, J. Lu, and J. Yang, Tunable Valley Polarization and Val- ley Orbital Magnetic Moment Hall E ect in Honeycomb Systems with Broken Inversion Symmetry, Sc. Rep. 5, 13906 (2015). [13] D. Xiao, J. Shi and Q. Niu, Berry Phase Correction to Electron Density of States in Solids, Phys. Rev. Lett. 95, 137204 (2005); D. Xiao, M.-C. Chang, and Q. Niu, Berry phase e ects on electronic properties, Rev. Mod. Phys. 82, 1959 (2010).[14] P.-O. L owdin, A Note on the QuantumMechanical Per- turbation Theory, J. Chem. Phys. 19, 1396 (1951). [15] J. C. Slater and G. F. Koster, Simpli ed LCAO Method for the Periodic Potential Problem, Phys. Rev. 94, 1498 (1954). [16] A. Korm anyos, V. Z olyomi, N. D. Drummond, P. Rakyta, G. Burkard, and V. I. Fal'ko, Monolayer MoS 2: Trigo- nal warping, the valley, and spin-orbit coupling e ects, Phys. Rev. B 88, 045416 (2013). [17] D. Ceresoli, T. Thonhauser, D. Vanderbilt, and R. Resta, Orbital magnetization in crystalline solids: Multi-band insulators, Chern insulators, and metals, Phys. Rev. B 74, 024408 (2006). [18] Giannozzi, P. et al. QUANTUM ESPRESSO: a modular and open-source software project for quantum simula- tions of materials, J. Phys. Condens. Matter 21, 395502 (2009). [19] Marzari, N. & Vanderbilt, D. Maximally localized gen- eralized Wannier functions for composite energy bands, Phys. Rev. B 56, 12847 (1997); I. Souza, N. Marzari and D. Vanderbilt, Maximally Localized Wannier Functions for Entangled Energy Bands, Phys. Rev. B 65, 035109 (2001). [20] M. G. Lopez, D. Vanderbilt, T. Thonhauser, and I. Souza Wannier-based calculation of the orbital magnetization in crystals, Phys. Rev. B 85, 014435 (2012); A. A. Mosto , J. R. Yates, G. Pizzi, Y. S. Lee, I. Souza, D. Vanderbilt and N. Marzari, An updated version of Wannier90: A Tool for Obtaining Maximally Localised Wannier Func- tions, Comput. Phys. Commun. 185, 2309 (2014). [21] Supplementary Materials describing the density func- tional methods and additional DFT results. [22] O. K. Andersen and T. Saha-Dasgupta, Mun-tin or- bitals of arbitrary order, Phys. Rev. B 62, R16219 (2000). [23] Y. K. Kato, R. C. Myers, A. C. Gossard, and D. D. Awschalom, Intrinsic Spin and Orbital Hall E ects from Orbital Texture, Science 306, 1910 (2004). [24] S. R. Park, J. Han, C. Kim, Y. Y. Koh, C. Kim, H. Lee, H. J. Choi, J. H. Han, K. D. Lee, N. J. Hur, M. Arita, K. Shimada, H. Namatame, and M. Taniguchi, Chiral Orbital-Angular Momentum in the Surface States of Bi 2Se3, Phys. Rev. Lett. 108, 046805 (2012). [25] S. Bhowal and S. Satpathy, Electric eld tuning of the anomalous Hall e ect at oxide interfaces, npj Computa- tional Materials 5(1), 61 (2019). [26] D. Go, J.-P. Hanke, P. M. Buhl, F. Freimuth, G. Bihlmayer, H.-W. Lee, Y. Mokrousov and Stefan Bl ugel, Toward surface orbitronics: giant orbital magnetism from the orbital Rashba e ect at the surface of spmetals, Sci.6 Rep.7, 46742 (2017).
2105.04172v1.Self_Bound_Quantum_Droplet_with_Internal_Stripe_Structure_in_1D_Spin_Orbit_Coupled_Bose_Gas.pdf
arXiv:2105.04172v1 [cond-mat.quant-gas] 10 May 2021Self-Bound Quantum Droplet with Internal Stripe Structure in 1D Spin-Orbit-Coupled Bose Gas Yuncheng Xiong and Lan Yin∗ Peking University (Dated: May 11, 2021) We study the quantum-droplet state in a 3-dimensional (3D) B ose gas in the presence of 1D spin-orbit-coupling and Raman coupling, especially the st ripe phase with density modulation, by numerically computing the ground state energy including th e mean-field energy and Lee-Huang- Yang correction. In this droplet state, the stripe can exist in a wider range of Raman coupling, compared with the BEC-gas state. More intriguingly, both sp in-orbit-coupling and Raman coupling strengths can be used to tune the droplet density. PACS numbers: 03.75.Hh; 03.75.Mn; 05.30.Jp; 31.15.Md; Introduction. Ultracold atoms have been excellent platforms for investigating many-body quantum phe- nomena since the experimental realization of Bose- Einstein condensation(BEC) [ 1–3]. In most cases, Gross- Pitaevskii (GP) equations, derived by minimizing mean- field energy functional with respect to condensation wavefunction, provide a good description for the BEC state of trapped Bose gases [ 4]. The next-order cor- rection to the gound state energy, i.e. the Lee-Huang- Yang(LHY) energy, is usually negligible in the dilute limit. However in a binary boson mixture, it was found that when the attractive inter-species coupling constant g↑↓is a little larger in magnitude than the geometric av- erageofthe repulsiveintra-speciescoupling constants g↑↑ andg↓↓, the repulsive LHY energy overtakes the attrac- tive MF ground-state energy, and the system becomes a self-bound quantum droplet [ 5]. The quantum droplet wasfirstobservedindipolarBosegases[ 6–10]andlaterin binary boson mixtures [ 11–13]. The self-binding mech- anism of a single-component dipolar Bose gas is simi- lar to that of a boson mixture except that the residual MF attraction arises from the counterbalance between attractive dipole-dipole interaction and repulsive contact interaction. Theoretically, the quantum droplet has been investigated with various methods, including variational HNC-EL method [ 14], ab initial diffusion Monte-Carlo [15], and extended GPEs with the LHY correction in- cluded [16]. On the other hand, Raman-induced spin-orbit- coupling(SOC)hasbeenrealizedexperimentallyinrecent years both in bosonic [ 17,18] and fermionic [ 19,20] sys- tems. Alternative scheme of SOC which is immune from heating problem has been theoretically [ 21–23] and ex- perimentally [ 24] investigated. In a two-component Bose gas with a one-dimensional (1D) SOC, a stripe struc- ture appears when the inter-species coupling constant is smaller than geometric average of intra-species cou- pling constants below a critical Raman coupling (RC) [17,25–30]. In previous studies, the stripe state has been investigated in the BEC-gas region with repulsive MF ground-state energy. Recently, theoretically stud-ies [31,32] reveal that the stripe state can also exist in the quantum droplet regime. In Ref. [ 31], the quantum droplet was found in a two-dimensional Bose gas with very weak SOC, where the LHY energy density was ap- proximated by that of a uniform system without SOC. In Ref. [32] the LHY energy of a three-dimensional system with SOC was calculated numerically, and its fitted form wasusedintheextended(GP)equation. Thephasetran- sition between a stripe gas and a stripe liquid was found by tuning coupling constants and RC. In their calcula- tion, the ultraviolet divergence in the expression of LHY energy was removed by dimensional regularization. In this work, we apply the standard regularization scheme to treat the ultraviolet divergence in the LHY energy of a three-dimensional system with SOC. We found that the droplet density can be easily tuned by RC and SOC, even to the zero limit. Compared to the case with a re- pulsive inter-species interaction, in the quantum droplet the stripe phase can exist in a bigger regime of RC and SOC. Boson mixture with SOC. Westudyatwo-component Bose gas system with total particle number Nand vol- umeV. In momentum space, its Hamiltonian is given by H=/summationdisplay k/summationdisplay ρρ′ˆφ† ρk/parenleftbigg(k−krexσz)2 2+Ω 2σx/parenrightbigg ρρ′ˆφρ′k +1 2V/summationdisplay k1,k2,q/summationdisplay ρρ′gρρ′ˆφ† ρk1+qˆφ† ρ′k2−qˆφρ′k2ˆφρk1,(1) whereˆφρkandˆφ† ρkare the annihilation and creation op- eratorsofthe ρ-component boson with the momentum k, {ρ,ρ′}={↑,↓},krand Ω are the strengths of SOC and RCrespectively, existhe unit vectorin x-directionwhich is the SOC direction. For convenience, we set ¯ hand the boson mass to be one. In this paper, we focus on the Ω<4Erregime where the lower excitation spectrum of the single-particle Hamiltonian has two degenerate min- ima [17,28]. For simplicity, the interactions are chosen to be symmetric g↑↑=g↓↓=g. The droplet regime is set by the condition g↑↓<∼−g[5].2 To implement Bogoliubov approximation to obtain ex- citation spectra, and thereby LHY correction, we need to know ground state(GS) wavefunction. To this end, we determine GS by variationally minimizing MF energy. We choose GS ansatz to be superposition of plane waves [27,28], φ(r) =/parenleftbiggφ↑ φ↓/parenrightbigg =/radicalbigg N0 V/summationdisplay m/parenleftbiggφ↑m −φ↓m/parenrightbigg eimk1·r,(2) whereN0is the particle number of the condensate, k1≡ (γkr,0,0),γis a variational parameter to be determined. In the lowest order, only m=±1 components corre- sponding to the two minima of the lower single-particle spectrum are relevent [ 27]. However, the periodic stripes induced by the condensation of ±k1will lead to the cou- plings between the momenta differing from each other by reciprocal lattice vectors. Therefore, it is necessary to include all the components with momenta K±k1in the higher order approximations [ 28], where K= 2sk1withs= 0,±1,±2,..., arereciprocallattice vectors. In short, the summation is over all the odd integer m= 2s±1 in the region −C1≤m≤C1where the cutoff C1is a posi- tive odd number. The normalization relation is given by/summationtext ρ,m|φρm|2= 1. In the BEC state, following the Bogoliubov prescrip- tion, we replace ˆφ(†) ρmk1by√N0φ(∗) ρm+ˆφ(†) ρmk1and keep terms up to the quadratic order. The first-order terms vanish due to the minimization of MF energy. The number of atoms in the condensation is given by N0= N−/summationtext′′ ρ,kˆφ† ρkˆφρkwhich can be used to rewrite N0in terms of the total atom number N. The MF energy per particle εMFand the Bogoliubov Hamiltonian HBare given by εMF=/summationdisplay m/summationdisplay ρρ′/parenleftbiggk2 r 2(mγ−σz)2−Ω 2σx/parenrightbigg ρρ′φ∗ ρmφρ′m +/summationdisplay m+l=i+j/summationdisplay ρρ′gρρ′n 2φ∗ ρmφ∗ ρ′lφρ′iφρj,(3) HB=EMF+/summationdisplay ρρ′/summationdisplay k/parenleftbigg(k−krexσz)2 2+Ω 2σx−µˆI/parenrightbigg ρρ′ˆφ† ρkˆφρ′k +/summationdisplay ρ,q/summationdisplay m+l=α+βgρρl 2/bracketleftbigg 2φ∗ ρmφ∗ ρlˆφραk1−qˆφρβk1+q+2φ∗ ρmφρ−l(ˆφ† ρ−αk1+qˆφρβk1+q+ˆφ† ρ−αk1−qˆφρβk1−q)+H.c./bracketrightbigg +/summationdisplay ρ/negationslash=ρ′,q/summationdisplay m+l=α+βgρρ′l 2/bracketleftbigg (−φ∗ ρmφ∗ ρ′l)(ˆφρ′αk1−qˆφρβk1+q+ˆφρ′αk1+qˆφρβk1−q)+(−φ∗ ρmφρ′−l)× (ˆφ† ρ′−αk1+qˆφρβk1+q+ˆφ† ρ′−αk1−qˆφρβk1−q)+φ∗ ρmφρ−l(ˆφ† ρ′−αk1+qˆφρ′βk1+q+ˆφ† ρ′−αk1−qˆφρ′βk1−q)+H.c./bracketrightbigg ,(4) wherem,l,i,j,α,βare all odd integers with −C1≤m,n,i,j ≤C1and−C2≤α,β≤C2,C2 is another cutoff necessary for numerical diagonaliza- tion of Bogoliubov Hamiltonian, qxis in the first Bril- lioun zone, 0 < qx< k1,nis total particle density, andµ=/summationtext m/summationtext ρ,ρ′/parenleftbigg k2 r 2(mγ−σz)2−Ω 2σx/parenrightbigg ρρ′φ∗ ρmφρ′m+ /summationtext m+l=i+j/summationtext ρ,ρ′gρρ′nφ∗ ρmφ∗ ρ′lφρ′iφρjis, in nature, the MF chemical potential, satisfying µ=∂EMF/∂N. In Hamiltonian Eq.( 4) there are not only terms equiv- alent to periodic potentials, but also off-diagonal terms such as ˆφραk1+qˆφραk1−q. The quasiparticle spectra are characterized by band index and quasimomentum. We define a column operator ˆAq≡/parenleftbig ···,ˆφ† ↑αk1−q,ˆφ† ↓αk1−q,ˆφ↑αk1+q,ˆφ↓αk1+q,···/parenrightbigT withα=±1,±3,...,±C2, and rewrite Eq.( 4) in a com-pact form HB=EMF+E1+/summationdisplay qˆA† qHqˆAq, where E1=−/summationdisplay m,q,±/bracketleftbigg(mk1−q±kr)2 2−µ+gn+g↑↓n 2/bracketrightbigg .(5) Thematrix Hqcanbe obtainedfromEq.( 4)andsubse- quentlydiagonalizedtoobtainquasiparticlespectra. The diagonalized Bogoliubov Hamiltonian is given by HB=EMF+E1+/summationdisplay α,q(E↑− α(q)+E↓− α(q)) +/summationdisplay α,q,±/summationdisplay ρEρ± α(q)ˆ˜φ† ραk1±qˆ˜φραk1±q, (6) where/parenleftbig ···,ˆ˜φ† ↑αk1−q,ˆ˜φ† ↓αk1−q,ˆ˜φ↑αk1+q,ˆ˜φ↓αk1+q,···/parenrightbigT= MqˆAq,Mqthe Bogoliubov transformation matrix sat- isfyingMqΣM† q= Σ, and Σ is a diagonal matrix with3 everyfourdiagonalmatrixelementsgivenby −1,−1,1,1. The quasi-particle energy Eρ± α(q) can be solved from the generalized secular equation |Hq−λΣ|= 0. Before we write down the expression of LHY en- ergy, we need to rewrite gρρ′in terms of scattering lengthaρρ′through regularization relation gρρ′=Uρρ′+ (U2 ρρ′/V)/summationtext k1/k2[33] whereUρρ′= 4πaρρ′. The LHY energy is therefore given by ELHY=E1+/summationdisplay α,q/bracketleftbigg (E↑− α(q)+E↓− α(q))+/parenleftbigg/summationdisplay m+l=i+j /summationdisplay ρ,ρ′(Uρρ′n)2 2φ∗ ρmφ∗ ρ′lφρ′iφρj/parenrightbigg/parenleftbigg/summationdisplay ±1 (αk1±q)2/parenrightbigg/bracketrightbigg ,(7) wherethe summation convergesquicklyfor largemomen- tum due to regularization. In contrast, the divergence of LHY energy was removed by dimensional regularization in Ref. [32]. Self-bound quantum droplet with stripe With the ex- plicit expressions of MF energy Eq.( 3) and LHY cor- rection Eq.( 7), we are ready to investigate the inter- play between SOC, RC and interactions in the forma- tion of droplet. We take Er≡k2 r/2 andkras en- ergy and momentum units respectively. The dimension- less version of MF and LHY energies are given in ap- pendix. In the following numerical calculations, we use the parameters a↑↑=a↓↓≡a= 89.08a0,a↑↓=−1.1a wherea0is the Bohr radius. Correspondingly, we de- fineU≡4πa=U↑↑=U↓↓. And before the effect of SOC is considered, the recoil momentum is fixed at kr= 2π×106m−1(or equivalently akr≈0.0296). When implementing the numerical calculations, we introduce two cutoffs: C1is for the ground-state ansatz, Eq.( 2), andC2is for the diagonalization of the otherwise infinite-dimensional Bogoliubov Hamiltonian, Eq.(4). We have numerically verified the convergence of wavefunctionsandLHYenergies,andfindthatthechoice ofC1= 9 and C2= 39 can produce sufficiently accurate results. The condensate fraction at m= 9 is about 10−14 of the total density. In our calculations, the two charac- teristic momenta,√gnand√ Ω, are at most of the same order of the recoil momentum kr. The momentum cut- offC2kris much larger than any of them. Therefore, throughout our calculations, we set C1= 9 and C2= 39. We first minimize the mean-field energy at fixed to- tal density nto determine variational parameters φρm andγ. It shows that, at low densities, the mean-field energy per particle εMFshows linear dependence on den- sity, and can be fitted by εMF=c0+c1(Un/E r). The density-independent background energy appears due to Raman energy in Eq.( 3), and thus c0depends strongly on the strength of RC, while the proportionality coef- ficientc1, as shown in Fig. 1, weakly relies on the RC strength in the considered regime. Both c0andc1are irrelevant to the strength of SOC, kr, since the MF en- ergy has been rescaled in the unit of Er. Moreover, c1isnegative throughout the considered regime, which indi- cates the tendency to collapse in the mean field and thus higher-order correction is necessary to stabilize such a system. In the low density region, the mean-field density dis- tribution exhibits stripes as in the low RC limit in re- pulsive BEC-gas [ 27]. As has been studied [ 17,27,29], in experiments on87Rb atoms, stripe phase exists only for very small Ω ( <∼0.2Er), and stripes cannot be de- tected directly in the absoption imaging. In contrast, the stripe phase of the quantum droplet can survive in a much larger range of RC, of the order of several Er. This result can be obtained in the variational theory [ 27] of a Bose gas with SOC, where the stripe phase and the plane-wave phase can all be described. In the low den- sity limit with strong SOC, there is a transition between these two phases at a critical RC, ΩI-II= 4Er/radicalBig 2γ 1+2γ whereγ= (g−g↑↓)/(g+g↑↓) (see Eq.(12) in [ 27]). Con- sequently, we can reach a conclusion that in a quantum droplet with Ω <4Erand strong SOC, stripe phase is favored. Compared to the BEC case with repulsive iner- species interaction where ΩI-II<∼0.2Er, one can draw the conclusion that it is the strong inter-species attrac- tion that significantly enlarges the region of stripe phase. We compute the LHY energy by solving excitation spectra and numerically performing the integration in Eq.(7). As in the case without SOC [ 5], the lowest exci- tation spectrum at qx≈0 and 2krhas small imaginary part. When performing the numerical integration, we keep all the real and imaginary contributions in the exci- tation energies. Even though the resulting LHY energy is complex, the imaginary part is at least three orders smaller in magnitude than the real part in the parameter region that we are considering, i.e. with low density and strong SOC. Consequently, the imaginary parts can be safely omitted in the ensuing calculations as in the case without SOC [ 5]. Notice that LHY energy is a function of three di- mensionless parameters, Un/E r, Ω/Erandakr(see Eq.(A.9)) whereas the MF energy depends only on the first two parameters as can be seen in Eq.( A.8). In the dilute limit, in terms of Un/E r, the LHY energy per par- ticle can be fitted by the formula εLHY=c2(Un/E r) + c3(Un/E r)3/2, wherec2andc3arepositivefittingparam- eters. Compared to the case without SOC, in addition to the usual term proportional to n3/2, a linear repulsive term appears in the LHY energy which will be discussed later. The coefficient c2, as shown in Fig. 1, has a roughly quadratic dependence on RC strength while c3remains constant in the range of 0 <Ω<3Er, in agreement with the preceding work [ 32]. Also, such a behavior occurs in the Rabi-coupled case [ 34] where the coefficient of usual n3/2term in LHY energy is free of Ω. For Ω >3Er, only low density region can be sampled, and while the4 0 1 2 3 40.000.010.020.030.04 Ω/Er|c1| c2 c3 FIG. 1. Three fitting coefficients c1,c2andc3in MF en- ergy and LHY energy vs the strength of Raman coupling Ω atakr≈0.0296.c1is negative and has been shown by its magnitude for convenient comparison with c2. fittedc1andc2remains reliable, the value of c3, which determines the behavior of the LHY energy in the higher density region, can not be trusted due to the deficiency of sampling. 0.001 0.010 0.100 1-0.008-0.006-0.004-0.0020.0000.002 Un/Er /Er =1/2=3/2=5/2=7/2 FIG. 2. Sampled points (dots) and fitting functions (solid lines) of total energy per particle ε=εMF+εLHYfor several Raman coupling strengths at akr≈0.0296. The constant energy background from the MF energy has been subtracted for comparison. Including both MF and LHY energies, the total en- ergy per particle in the droplet regime is shown in Fig. 2. Near the collapse point of the MF energy, contrary to the monotonously decreasing tendency of the MF energy with density, the total energy has a minimum, because the repulsive LHY energy overcomes the attractive MF energy at larger densities. As discussed above, the MF wavefunction shows density modulation. Therefore, in the dropletregime, the self-bound stripe phase exists and can survive for even larger Raman coupling compared to the BEC gas regime [ 17,27,29]. Although the general analytic total energy per particle0.050.100.150.200.250.301017101810191020 akrnd/m-3 Ω=1/2 Ω=3/2 Ω=5/2 FIG. 3. Droplet density ndversus SOC strength krfor several RC strengths. With RC fixed, the stronger the SOC is, the smaller the equilibrium density of droplet becomes. When SOC is strong enough, for example at akr≈0.32 for Ω = 1/2Er, the droplet state disappears and the system expands without an external trap. is inaccessible in the presence of SOC and RC, it can be approximatedbyaddingtogetherthe fitted MF andLHY energies, ε=c0+(c2−|c1|)(Un/E r)+c3(Un/E r)3/2with c0,1,2,3all fitted numerically. The equilibrium density of self-bound droplet can obtained by solving zero-pressure condition, i.e. P=∂ε/∂V= 0, yielding Und/Er= 4(|c1|−c2)2/9c2 3for|c1|−c2≥0. We now discuss the role played by RC strength on droplet formation. The dependence of droplet on RC is similar to the case in uniform Rabi-coupled binary mixture which has been reported in 3D [ 34] and lower dimensions [ 35]. The similarity stems from the gapped single-particlespectrum. Without RC,thesingle-particle spectra of both components are gapless. The finite RC induces coupling between the two components leading to the new quasi-particle spectra, with the lower one gap- lessandthe higheronegapped. Due tothe gappedmode, LHY correction per particle acquires a positive term lin- ear in density nin addition to the n3/2term, as men- tioned above. As Ω increases, the rapid increase of c2results in the decreasing of |c1| −c2in magnitude, which is mani- fest in Fig. 1, withc3remaining almost constant, mak- ing it easier to counterbalance the attractive MF energy. Thus the equilibrium density of droplet is smaller for larger Ω as shown in Fig. 2. A critical point is reached at|c1|=c2, as shown in Fig. 1where Ω c≈3.5Erat akr≈0.0296. Above this point, the total energy in- creases monotonously with density since both c2− |c1| andc3are positive, and thereby no self-bound droplet can exist. A similar droplet-gas transition has been re- ported theoretically in Rabi-coupled binary mixture in both 3D [ 34] and lower dimensions [ 35]. ItiseasiertoconsiderthedependenceonSOCstrength askr(Er≡k2 r/2) serves as the momentum (energy) unit. From the dimensionless expression of MF and LHY energies Eq.( A.8) and Eq.( A.9), it’s easy to see that the MF energy only depends on Un/E rand Ω/Er, and so5 does the excitation energy Eρ± α(q). With fixed reduced interaction Un/E rand Raman coupling Ω /Er, the LHY energy is proportional to kr, and as a consequence, the fitting parameters in LHY energy, c2andc3are both linearly proportional to kr. Since c1is independent of kr, increasing SOC strength krhas the same effect as increasing RC strength Ω. Although the energy unit Er is also increased in the same process, the overall effect of increasing SOC strength is, as shown in Fig. 3, decreasing equilibrium density of droplet. And finally above some specific value, no droplet can exist any more. Conclusion and Discussion In current experiment on 39K, the droplet state has been realized in the mixture of hyperfine states |1,−1/an}bracketri}htand|1,0/an}bracketri}htby tunning scattering lengths [5,11–13], but the artificial SOC has not been re- alizedinthissystem. Incontrast,in87Rbsystems[ 17,18] the SOC has been realized and the stripe state has been observed, but tunning the interactions in this system has not been achieved. Our results could be tested experi- mentallyiftheinteractionscanbetunedandtheartificialSOC can be generated in the same system. In conclusion, we have studied the quantum droplet state of a uniform binary Bose gas in the presence of 1D spin-orbit-coupling and Raman coupling, and find that groundstate can display density modulation of the stripe phase in the low Ω regime [ 27,28]. The density modula- tion can survive for much larger Ω than in the BEC gas state with inter-species interaction. Compared to the case without SOC, the droplet density can be tuned by changing the strength of SOC and RC. With the increase of SOC and RC, the droplet density can be reduced by several orders of magnitude, and eventually to the zero limit at a critical kror Ω. We plan to study the finite- size effect of the quantum droplet with SOC in the future work. Appendix Dimensionless MF and LHY energies per particle are given by EMF/(NEr) =/summationdisplay m/summationdisplay ρρ′/parenleftbigg (mγ−σz)2−Ω 2σx/parenrightbigg ρρ′φ∗ ρmφρ′m+/summationdisplay m+l=i+j/summationdisplay ρρ′Uρρ′n 2φ∗ ρmφ∗ ρ′lφρ′iφρj,(A.8) ELHY/(NEr) = (akr)1 π2(Un)/summationdisplay α/integraldisplayγ 0/integraldisplayγC2 −γC2/integraldisplayγC2 −γC2dqxdqydqz/bracketleftbigg (E↑− α(q)+E↓− α(q))−/summationdisplay ±/parenleftbigg (αγ−qx±1)2+q2 y +q2 z−µ+Un+U↑↓n 2/parenrightbigg +/parenleftbigg/summationdisplay m+l=i+j/summationdisplay ρρ′Uρρ′n 2φ∗ ρmφ∗ ρ′lφρ′iφρj/parenrightbigg/parenleftbigg1/2 (αγ±qx)2+q2y+q2z/parenrightbigg/bracketrightbigg , (A.9) where Ω, Uρρ′nand excitation spectra Eρ− αare in the unit ofEr. ∗yinlan@pku.edu.cn [1] M. H. Anderson, J. R. Ensher, M. R. Matthews, C. E. Wieman, and E. A. Cornell, Science269, 198 (1995) . [2] C. C. Bradley, C. A. Sackett, J. J. Tollett, and R. G. Hulet,Phys. Rev. Lett. 75, 1687 (1995) . [3] K. B. Davis, M. O. Mewes, M. R. Andrews, N. J. van Druten, D. S. Durfee, D. M. Kurn, and W. Ketterle, Phys. Rev. Lett. 75, 3969 (1995) . [4] F. Dalfovo, S. Giorgini, L. P. Pitaevskii, and S.Stringa ri, Rev. Mod. Phys. 71, 463 (1999) . [5] D. S. Petrov, Phys. Rev. Lett. 115, 155302 (2015) . [6] H. Kadau, M. Schmitt, M. Wenzel, C. Wink, T. Maier, I. Ferrier-Barbut, and T. Pfau, Nature530, 194 (2016) . [7] I. Ferrier-Barbut, H. Kadau, M. Schmitt, M. Wenzel, and T. Pfau, Phys. Rev. Lett. 116, 215301 (2016) . [8] L. Chomaz, S. Baier, D. Petter, M. J. Mark, F. W¨ achtler,L. Santos, and F. Ferlaino, Phys. Rev. X 6, 041039 (2016). [9] M. Schmitt, M. Wenzel, F. Bottcher, I. Ferrier-Barbut, and T. Pfau, Nature539, 259 (2016) . [10] A. Trautmann, P. Ilzh¨ ofer, G. Durastante, C. Politi, M. Sohmen, M. J. Mark, and F. Ferlaino, Phys. Rev. Lett.121, 213601 (2018) . [11] C. R. Cabrera, L. Tanzi, J. Sanz, B. Naylor, P. Thomas, P. Cheiney, and L. Tarruell, Science359, 301 (2018) . [12] P. Cheiney, C. R. Cabrera, J. Sanz, B. Naylor, L. Tanzi, and L. Tarruell, Phys. Rev. Lett. 120, 135301 (2018) . [13] G. Semeghini, G. Ferioli, L. Masi, C. Mazzinghi, L. Wol- swijk, F. Minardi, M. Modugno, G. Modugno, M. In- guscio, and M. Fattori, Phys. Rev. Lett. 120, 235301 (2018). [14] C. Staudinger, F. Mazzanti, andR.E. Zillich, Phys. Rev. A98, 023633 (2018) . [15] V. Cikojevi´ c, K. Dˇ zelalija, P. Stipanovi´ c, L. Vranj eˇ s Marki´ c, and J. Boronat, Phys.Rev. B 97, 140502 (2018) . [16] F. W¨ achtler and L. Santos, Phys. Rev. A 93, 061603 (2016). [17] Y. J. Lin, K.Jim´ enez-Garc´ ıa, andI.B. Spielman, Nature6 471, 83 (2011) . [18] J.-Y. Zhang, S.-C. Ji, Z. Chen, L. Zhang, Z.-D. Du, B.Yan, G.-S.Pan, B.Zhao, Y.-J.Deng, H.Zhai, S.Chen, and J.-W. Pan, Phys. Rev. Lett. 109, 115301 (2012) . [19] L. W. Cheuk, A. T. Sommer, Z. Hadzibabic, T. Yefsah, W. S. Bakr, and M. W. Zwierlein, Phys. Rev. Lett. 109, 095302 (2012) . [20] P. Wang, Z.-Q. Yu, Z. Fu, J. Miao, L. Huang, S. Chai, H. Zhai, and J. Zhang, Phys. Rev. Lett. 109, 095301 (2012). [21] Y. Zhang, G. Chen, and C. Zhang, Sci. Rep. 3, 1937 (2013). [22] Z.-F. Xu, L. You, and M. Ueda, Phys. Rev. A 87, 063634 (2013). [23] B. M. Anderson, I. B. Spielman, and G. Juzeli¯ unas, Phys. Rev. Lett. 111, 125301 (2013) . [24] X. Luo, L. Wu, J. Chen, Q. Guan, K. Gao, Z. F. Xu, L. You, and R. Wang, Sci. Rep. 6, 18983 (2016) . [25] C. Wang, C. Gao, C.-M. Jian, and H. Zhai, Phys. Rev. Lett.105, 160403 (2010) . [26] T.-L. Ho and S. Zhang, Phys. Rev. Lett. 107, 150403 (2011).[27] Y. Li, L. P. Pitaevskii, and S. Stringari, Phys. Rev. Lett. 108, 225301 (2012) . [28] Y. Li, G. I. Martone, L. P. Pitaevskii, and S. Stringari, Phys. Rev. Lett. 110, 235302 (2013) . [29] S.-C. Ji, J.-Y. Zhang, L. Zhang, Z.-D. Du, W. Zheng, Y.-J. Deng, H. Zhai, S. Chen, and J.-W. Pan, Nature Physics10, 314 (2014) . [30] J.-R. Li, J. Lee, W. Huang, S. Burchesky, B. Shteynas, F. a. Top, A. O. Jamison, and W. Ketterle, Nature543, 91 (2017) . [31] R. Sachdeva, M. N. Tengstrand, and S. M. Reimann, Phys. Rev. A 102, 043304 (2020) . [32] J. S´ anchez-Baena, J. Boronat, and F. Mazzanti, Phys. Rev. A102, 053308 (2020) . [33] L. Pitaevskii and S. Stringari, Bose-Einstein condensa- tion and superfluidity , Vol. 164 (Oxford University Press, 2016). [34] A. Cappellaro, T. Macri, G. F. Bertacco, and L. Salas- nich,Sci. Rep. 7, 13358 (2017) . [35] E. Chiquillo, Phys. Rev. A 99, 051601 (2019) .
2203.12263v1.Effects_of_a_single_impurity_in_a_Luttinger_liquid_with_spin_orbit_coupling.pdf
E ects of a single impurity in a Luttinger liquid with spin-orbit coupling M. S. Bahovadinov1, 3and S. I. Matveenko2, 3 1Physics Department, National Research University Higher School of Economics, Moscow, 101000, Russia 2L. D. Landau Institute for Theoretical Physics, Chernogolovka, Moscow region 142432, Russia 3Russian Quantum Center, Skolkovo, Moscow 143025, Russia (Dated: March 24, 2022) In quasi-1D conducting nanowires spin-orbit coupling destructs spin-charge separation, intrinsic to Tomonaga-Luttinger liquid (TLL). We study renormalization of a single scattering impurity in a such liquid. Performing bosonization of low-energy excitations and exploiting perturbative renormalization analysis we extend the phase portrait in KKspace, obtained previously for TLL with decoupled spin-charge channels. I. INTRODUCTION Low-energy excitations of the one-dimensional con- ductors have collective phonon-like character, formalized within the Tomonaga-Luttinger liquid (TLL) theory [1{ 4]. The TLL formalism correctly predicts algebraic decay of single-particle correlations with exponents depending on the strength of the (short-range) electron-electron (e- e) interactions. Due to the collective nature of these ex- citations the interparticle interactions drastically change the low-energy physics leading to fractionalization of con- stituent carriers [5{7], vanishing density of states and power law singularity of the zero-temperature momen- tum distribution at the Fermi level [8{10]. Another sur- prising prediction of the TLL theory is related to the e ects of a single scattering impurity on transport prop- erties at low temperatures, presented in prominent se- ries of works in Refs. [11{13]. Perturbative renormaliza- tion group (RG) studies have shown that if the electron- electron interactions of the single-channel TLL is attrac- tive (with the TLL parameter K > 1), impurity is irrele- vant in the RG sense due to the pinned superconducting uctuations. On the other hand, if the e-e interactions have a repulsive character ( K < 1), the backscattering e ects are relevant and grow upon the RG integration of short-distant degrees of freedom. At T= 0 the sys- tem is e ectively decoupled into two disjoint TLLs for arbitrarily weak impurity potentials. The latter implies a metal-insulator transition with vanishing two-terminal dc-conductance G= 0 atT= 0, whereas at nite temperatures it exhibits power-law dependence on T, i.e GT2=K2. These results are in sharp constrast with the result in the non-interacting limit ( K= 1), where the impurity is marginal in the RG sense (in all orders of RG) and the transmission coecient Tdepends on the impurity strength with Landauer conductance G=Te2 h. Experimental con rmation of these e ects was performed recently, using remarkable quantum simulator based on a hybrid circuit [14]. In TLLs composed of spin-1/2 electrons [15{18] spin and charge degrees of freedom are decoupled [19{23], in contrast to higher dimensional counterparts. Each chan- nel has individual bosonic excitations with the proper velocity and carry the corresponding quantum number 0.5 2 KS 0.52KC 0.5 2 KS 0.52KCweak barrier strong barrier (a) (b) IIII IIIV IIIII IIVFIG. 1. Phase portrait at T= 0 for TLL with separated spin and charge channels. Results are obtained from perturbative RG studies of (a) a weak scattering barrier or (b) a weak link. In region I the backscattering of carriers from a weak barrier (tunneling through a link) is relevant (irrelavant) and ows to the strong-coupling (to the weak-coupling) xed point in both channels. The opposite occurs in the regions IV. In the remained regions the mixed phase is realized, where one of the channels is insulating and the other is conducting. separately. However, this decoupling of modes is not characteristic for all TLLs. Particularly, the earlier stud- ies have shown that spin-charge coupling (SCC) occurs in the TLL subjected to a strong Zeeman eld [24] or with strong spin-orbit interactions [25{27]. The latter is typically expected in nanowires, where electrons in trans- verse directions are con ned, whereas in the other direc- tion they move freely. Spin-orbit coupling (SOC) in these systems plays an important role on the realization of spin- tronic devices [28, 29]. Thus, an interplay of (large) SOC and (large) e-e interactions in the TLL regime is an in- teresting question and has gained both theoretical and experimental attention in recent years [30{32]. Impurity e ects in conventional TLLs with decoupled channels were studied in the original works [12, 13], and can be summarized by phase portrait at T= 0, presented in Fig. 1. RG studies imply that electrons scattering on a weak potential barrier [Fig. 1(a)] are fully re ected when Kc+Ks<2 (region I). In region II (III) charge (spin) quanta are fully re ected, while the spin (charge) ones are transmitted. The backscattering terms scale to thearXiv:2203.12263v1 [cond-mat.str-el] 23 Mar 20222 ξ(k) ξ(k) kυ1 kF,1υ2 -kF,2-kF,1 kF,2υ1 υ2υ υ kF (a) Polarized TLL -kF (b) TLL with SOC kυ υ FIG. 2. Linearized electronic excitation spectrum corre- sponding to the (a) TLL in a strong Zeeman eld and (b) TLL with SOC. In both models spin-charge separation is vi- olated by the mixing terms given in Eq. (11) and Eq. (30). weak-coupling xed point in region IV, so electrons fully transmit through the barrier. The same physical picture is obtained from the perturbative analysis of the tun- neling events between disconnected wires [Fig. 1(b)]. In region I (1 Kc+1 Ks<2;Kc<2;Ks<2) all hopping events are irrelevant, so the system renormalizes into two dis- connected wires. In region II (region III) only tunneling of spin (charge) is relevant. Numerical con rmation of these results was presented using path-integral Monte- Carlo methods [33]. The modi cation of the phase portrait in the presence of SCC has not yet been considered, although there are several studies which partially addressed this question in di erent aspects [34, 35]. Particularly, impurity e ects on transport properties in the case of the SCC caused by SOC were recently studied [31, 32]. As our main moti- vation of this work we consider modi cation of the phase portrait in the presence of SOC. We also show that the spin- ltering e ect conjectured for this model in Ref. [35], is not exhibited. The paper is organized as follows: In Section II we present the model of our study and set the necessary for- malism and conventions. We next present a generalized approach to tackle the impurity problem for a nite SOC in Section III. In Section IV we present our main results and discussions. Concluding remarks are given in Section V. II. MODEL AND METHODS In this section we set up our used conventions and ter- minology and present the model of our study. We ap- proach the problem using the standard Abelian bosoniza- tion technique with the consequent perturbative RG analysis of impurity terms. Bosonization of low-lying modes relies on the assump- tion of linear electronic spectrum with the corresponding right (R) and left (L) movers for carriers with both spin projections (see Fig. 2). Fermionic eld operators foreach branch and spin component can be expressed via the bosonic displacement and phase elds: ;=^Fp 2aeikF;xeip((x)(x))(1) with2f" +1;# 1gand2fR+1;L1g. The dual elds satisfy the commutation relations: [(x);(x0)0] =i;0(xx0); (2) where (x) =r(x) is a conjugate to (x) momen- tum. The UV cuto aof the theory is on the order of inter-atomic distances. As in the standard litera- ture, the Klein factors ^Fguarantee anti-commutation of fermionic elds with di erent spin orientations. The second exponent can also be expressed in terms of constituent bosonic ladder operators in momentum space (lis the length of the system) ip((x)(x)) =X q>0Aq bq;eiqxby q;eiqx (3) withAq=q 2 ljqjeajqj=2and [bq;by q0] =q;q0,q=2 l. We hereafter use the spin and the charge basis as the canonical basis, i.e we work with the elds c;s="#p 2 and c;s="#p 2, which also satisfy Eq. (2). Hereafter, the subbscripts "s" and "c" stand for spin and charge degrees of freedom. The uctuations of the spin and the charge density are given by, s(x) =r 2 X @x (4) and c(x) =r 2 X @x: (5) To get similar expressions for the current, one uses the transformation @x!@x. In the following, we imply normal ordering with respect to Dirac sea, whenever it is needed and avoid : () : symbol. Model The important e ect of SOC in quantum nanowires is band distortion, which is usually considered within the two-band model [36]. This distortion causes the veloc- ity di erence  = v1v2, pronounced in Fig. 2(b) in the approximation of linearized spectrum. Initially pro- posed in Refs.[25, 26], the model can be realized by tun- ing chemical potential, lling only the lowest subband. It should be noted that the spin orientation of carriers moving in the same direction can be tuned also from the parallel [37] to anti-parallel [25] by band- lling [32]. In3 this work we consider anti-parallel spin orientation, as shown in Fig. 2(b). The non-interacting Hamiltonian of the model with lin- earized excitation spectrum is H0=iv1Z dx y R;#(x)@x R;#(x) y L;"(x)@x L;"(x) iv2Z dx y R;"(x)@x R;"(x) y L;#(x)@x L;#(x) ; (6) with distinct Fermi velocities v1andv2. As it is clear from the electronic spectrum, the model has broken chiral symmetry, but the time-reversal symmetry is preserved. Corresponding excitations can be rewritten equiva- lently in the bosonic language bq(neglecting zero-energy modes) H0=v1 X q>0jqjby q;#bq;#+X q<0jqjby q;"bq;"! +v2 X q>0jqjby q;"bq;"+X q<0jqjby q;#bq;#! :(7) Collecting the common terms, we obtain H0=vF 2X q6=0;jqjby q;bq;+ 2X q6=0;qby q;bq; (8) with  =v1v2andvF=v1+v2. The second term vanishes for vanishing SOC and has the following form in the spin-charge basis: Hmix= 2X q6=0q by q;sbq;c+by q;cbq;s : (9) This term with nite  6= 0 violates the spin-charge sep- aration and demands more general approach. We consider the interacting theory within the general- ized g-ology approach and do not impose any constraint on TLL parameters K,2(s;c). Within this general- ization, the coordinate space representation of the inter- acting Hamiltonian in the spin-charge basis has quadratic Gaussian form: HSOC =X =s;cv 2Z1 K(@x)2+K(@x)2 dx+Hmix: (10) The SCC term is expressed as follows: Hmix= 2Z [@xs@xc+@xc@xs]dx: (11) For vanishing SOC, the chiral symmetry is restored, with fullSU(2) symmetry and Ks= 1. In principle, one has to include also the backscattering term to the Hamiltonian, HBS=gs 2(a)2Z dxcos(p 8s): (12)In the parameter space of its relevancy, this term opens a gap in the spin channel via the Berezinskii-Kosterlitz- Thouless mechanism. There are several works which have addressed the relevancy of this term for repulsive e-e in- teractions [25, 38{40]. However, in a recent experiment on InAs nanowires with a strong SOC, no sign of spin gap is observed [31]. Based on the phenomenology, we neglect all such terms within the whole parameter space. We also emphasize that impurity e ects in the system with gapped spin channel were also previously studied [40, 41]. For perturbative analysis of impurity e ects, the imaginary-time Euclidean actions for the displacement elds can be obtained by integrating out the quadratic phase elds. In the ~ x= (x;) space it takes the following form: S =1 2vZ dxd (@)2+ ~v2 (@x)2 (13) withv=vsKs=vcKc, guaranteed by Galilean invari- ance of the model with  = 0. The renormalized veloci- ties are: ~v2 =v2 d; (14) d= 12 4v2 : (15) The contribution from the mixing term is, S mix=i 2vZ dxd (@xs@c+@xc@s): (16) Euclidean actions for the phase eld uctuations can be obtained in a similar fashion, S =v 2Z dxd1 v2(@)2+d(@x)2 (17) along with the mixing -term S mix=iv 2Z dxd1 v2c(@xs@c) +1 v2s(@xc@s) : (18) For  = 0, one correctly obtains the standard expres- sions for actions S andS . Impurity bosonization We consider a single impurity embedded at the origin x= 0. This impurity causes backscattering of carriers Hbs=VbsX  y R; L;+h:c: ; (19) whereVbsis the Fourier component of the backscattering potentialV(kF;1+kF;2). The forward-scattering term can be always gauged out [3].4 The bosonized expression of the backscattering term leads to the boundary sine-Gordon model in the spin- charge basis: Hbs=2Vbs acos( Is(x= 0)) cos( Ic(x= 0)) (20) with 2 I= 2. This term also couples spin and charge degrees of freedom at the impurity point. We note that in the bosonization process of this term we do not take into account Klein factors ^F, since their e ect is irrelevant within our model with all terms included terms in this work. The corresponding backscattering action takes the following form: Sb=2Vbs aZ 0dcos( Is(0;)) cos( Ic(0;)) (21) =1 Tis the inverse temperature and should not be confused with I. We hereafter assume Vbs=1, where  in the e ective bandwidth for both channels. For a complete analysis, pertubative study of the backscatteting action Eq. (21) should be accompanied with the RG study in the strong barrier limit. For this purpose, we consider two disjointed wires and treat in- terwire tunneling term perturbatively. The most relevant tunneling term has the following contribution to the total action, St=t aZ 0dcos( Is(0;)) cos( Ic(0;));(22) wheretis the bare tunneling amplitude through the weak link. III. EFFECTIVE BATHS ACTIONS At this point it is worth noting that the previous bosonization studies of TLL with SCC [24, 32, 35] were performed using basis rotation approach, where one transforms the initial basis to the new spin-charge basis with decoupled channels and new renormalized velocities andKparameters. This approach usually results on redundant expressions for model parameters and com- plicates further analysis. Instead, we follow a generic approach [3] without performing a basis rotation. As we show below, this procedure is not required. We next trace out all space degrees of freedom except for the im- purity point. The low-lying excitations of the elds away from the impurity act as a dissipative bath reducing the problem to an e ective 0D eld theory of a single Brow- nian quantum particle in a 2D harmonic impurity poten- tial [42]. We follow the standard procedure [3, 12] to get e ective bath actions for the displacement uctuation. The same procedure applies for the phase elds.The integration of the bulk degrees of freedom is done using the standard trick of introducing Lagrange multi- pliers with the real auxilliary elds  Z=Z DsDcDcDsDsDceST(23) with ST=S[c;s] +iZ 0dX [()(0;)](): (24) One needs consequently integrate out the andto get the nal baths actions: S =X !nj!nj ^F det(F)! j(!n)j2(25) with the bosonic Matsubara frequencies !n=2n ;(n2 Z). The mixing term is: S mix=X !nj!nj ^Fmix det(F)! s(!n)c(!n);(26) where theFmatrix and ^Fare given in Eqs. (A.1)-(A.3). Importantly, the mixing action S mixvanishes (see Ap- pendix) and one is left with the decoupled set: S =X !nj!nj ~K j(!n)j2(27) with ~K given by Eq. (A.5) and 2(s;c). The resulted Caldeira-Legett type actions are common and describe dynamics of a single quantum Brownian particle in the regime of Ohmic dissipation [42, 43]. These actions rep- resent the weak-coupling xed point of pure Luttinger liquid. Similarly, the e ective baths actions for  elds can be obtained as S =X !nj!nj~K j(!n)j2(28) with the new TLL parameters ~K given by Eq. (A.6). The mixing action also vanishes in this case. Vanishing mixing terms and decoupled baths map the problem onto the one with conventional TLL baths, but with the new parameters ~Kandv;+given by Eqs. (A.5)-(A.7). Hence, the usual basis-rotation proce- dure is excessive here. Obviously, for the vanishing spin- charge mixing term  = 0, the standard TLL parameters are correctly recovered, ~K!K. The characteristic velocitiesv+!max(vs;vc) andv!min(vs;vc). This implies that in the presence of SOC, one has new modes, with carriers composed of spin and charge quanta. These new modes have excitation velocities v;+and posses new TLL parameters ~K. Prior to the discussion of our main5 results, we emphasize two important features arisen al- ready at this stage of analysis. Mode freezing . As the strength of SOC is increased, v+ monotonically increases, whereas vsimilarly decreases and vanishes at the critical c . As it was mentioned in previous works [25, 27], at the critical SOC strength "freezing" of the corresponding mode (phase separation) with diverging spin (or charge) susceptibility is exhibited. For repulsive e-e interactions, it was shown [25] that it is the spin susceptibility which diverges at the critical SOC as, =0 1 cs2!1 (29) with spin-susceptibility 0at zero SOC and c s= 2=Ks. Similar divergence can be observed in the charge channel, if one considers full KcKsspace. These divergences in- dicate a phase transition, which occurs in the spin/charge channel (see Ref. [27] and references therein). In this work, we consider SOC strengths causing the velocity dif- ference =v < 0:8, which restricts the parameter space toK<5=2. Absence of spin- ltering e ect . In the case of the spin-charge mixing caused by the strong Zeeman eld [Fig. 2(a)] the mixing bath actions Smixdoes not vanish, due to the di erent type of SCC mechanism. The mix- ing term Eq. (11) in the Hamiltonian takes the following form: Hmix=B 2Z [@xs@xc+@xs@xc]dx (30) with B=v"v#. As it was demonstrated earlier in Refs. [24, 35], a single impurity as in Eq. (19) embedded to such TLL causes polarization of the spin current with a ratio of tunneling amplitudes, t" t#=T  ; (31) and with a nite exponent (B) for nite  B. Observation of this e ect in our model was also pro- posed earlier in Ref. [35]. However, an important conse- quence of vanishing Smixaction terms of Eq. (26) (and similar term for  eld) is the absence of such spin- ltering e ect, albeit the spin-charge separations is vi- olated. IV. RESULTS AND DISCUSSIONS Weak potential barrier Once the e ective bath actions are obtained as in Eq. (27) and Eq. (28) , the standard pRG analysis in the limit of weak and strong impurity potential can be performed.We rst consider the scattering of electrons on a weak barrier. In this limit, the partition function is, Z=Z DsDceST; (32) with the total action, ST=S c+S s+Vm;n aZ 0dcos(m Ic) cos(n Is): (33) To analyze the relevance of the last term within the Kc Ksparameter space, we generalize the impurity term to take di erent values of mandnwith the amplitudes Vm;n, since such terms are necessarily generated during the RG process. Treating the backscattering terms perturbatively, one obtains the standard set of rst-order RG equations [12, 13], dV1;1(l) dl= 1~K s+~K c 2! V1;1(l); (34) dV0;2(l) dl= 12~K s V0;2(l); (35) dV2;0(l) dl= 12~K c V2;0(l) (36) withdl=d . The rst equation for V1;1Vbscorresponds to the kF;1+kF;2backscattering of a single electron, whereas the second and the third equations correspond to the backscattering of spin or charge (electron pair) degree of freedom. The last processes have the fermionic expres- sions y R;" L;" y L;# R;#+h:c:and y R;" L;" y R;# L;#+ h:c:, respectively. The sketches of these scattering pro- cesses in the momentum space are presented in Fig. 3(b)- (c). All higher-order decendent terms are neglected, since the regions of their relevancy are covered with the ones of the last two equations, even for the nite . Thus, the main low-temperature processes are dictated by these three equations Eqs.(34)-(36). For vanishing SOC  = 0 ( ~K!K), the marginal xed line de ned in Eq. (34) is determined by the con- ditionKs+Kc= 2. The region I in Fig. 3(a) corre- sponds to a parameter space where the backscattering of a single electron in a weak potential becomes a relevant process and the backscattering amplitude V1;1 ows to the strong-coupling regime. This leads to the blocked transport in both charge and spin channels and the elds s;care pinned on the minima of cosine functions. Qual- itatively, in this region either single electron ( "or#) is fractionalized by e-e interactions, which is responsible for full backscattering of single carrier and hence, charge and spin carriers, too. Thus, one has vanishing conductance G= 0 atT= 0.6 (b) (c)charge backscattering spin backscattering(a) IV I IIIII KF,1 KF,2 KF,2K K KF,1ξ(k) ξ(k) 0 0.5 1 1.5 2 2.5 Ks00.511.522.5Kc∆=0 ∆=0.4 ∆=0.8 FIG. 3. (a) The modi cation of the phase portrait for the TLL with nite SOC. For  >0 the phase boundary of region I is modi ed. In (b) and (c) the backscattering processes of charge and spin are sketched, which correspond to Eqs. (35)- (36), respectively. As dictated by Eqs. (35)-(36), the backscattering pro- cesses of the charge/spin carriers also become relevant whenKc=s<0:5. For the charge backscattering, this im- plies the charge quanta to carry smaller than 0 :5 unit, to make it a relevant process. In the region of relevance II (III for spin channel), the charge (spin) channel is insu- lating since the the  c(s) is pinned. The spin (charge) fully transmit the barrier in this region, implying the re- alization of mixed phases in these regions. In region IV both channels have dominating superconducting uctu- ations [27]. We emphasize again that in this region with attractive e-e interactions, other type of instabilities in the bulk may arise, which usually open a gap in one of the channels. Here, we neglect all such processes and present the simplest picture. To investigate the e ect of SOC on backscattering of carriers, we represent the new TLL parameters in a sym- metric way: ~K =Kv ~v~vc+ ~vs v++v ; (37) where ~vandvare given in Eq. (14) and Eq. (A.7), respectively. In the presence of SOC the rst e ect is the renormalization of spin and charge velocities, which is manifested in the rst factor of Eq. (37). The second factor is due to the emerged modes with velocities v+;. As it was mentioned in the previous section, we limit considiration of the parameter space K<5 2and =v< 0:8. For non-interacting electrons, the e ect of SOC is lim- ited to the breaking of chiral symmetry and the renor- malization of excitation velocities. Indeed, one has ~K = K = 1, since renormalizing factors in Eq. (37) cancel each other. As shown in Fig. 3a this is also valid for weakly-interacting electrons, K1. For the given nite SOC strengths, where 0 <=v < 0:8, the second factor ~vs+~vc v+v+ 1 is xed for both K, whereas the rst one de nes scattering of carriersand determines the boundary of region I. The strongest e ect of SOC is exhibited when in one of the channels K1. In this limit one can lock the corresponding  eld to the minimum of cosine function and reexpress the scaling dimension of the impurity term for m= 1 andn= 1 as follows: 21;1=K+K(1 +2 8v2 ): (38) The resulting marginality line is presented in Fig. 3(a) for =v= 0:4 and =v= 0:8. At the critical K=5 2, the excitations in the spin/charge channel become frozen (v= 0) and the bulk of the channel becomes insulating. The e ect of nite SOC on the boundaries of regions II/III with the region IV is negligibally small. The largest correction to the scaling dimension 0;1(1;0) is of the order of 102for the largest  =v= 0:8. Thereby it can be safely neglected. Finally, one can straightforwardly generalize the ex- pressions for corrections to (bulk) conductances obtained in Refs. [12][13] to the case with nite SOC by K!~K: G=e2 hX m;ncm;njVm;nj2T(m2~Kc+n2~Ks)=22(39) with dimensionless coecients cm;n. Strong barrier For analysis of tunneling term in Eq. (22) one considers the following total action ST, ST=S c+S s+tm;n aZ 0dcos(m Ic) cos(n Is): (40) Similar to the previous case, the impurity term is gener- alized for di erent mandn. The RG transformation of the impurity term leads to the following set of equations, dt1;1(l) dl= 11 21 ~Kc+1 ~Ks t1;1(l); (41) dt2;0(l) dl= 12 ~Kc t2;0(l); (42) dt0;2(l) dl= 12 ~Ks t0;2(l) (43) with the TLL prameters rewritten as: 1 ~K=1 Kv ^v^vc+ ^vs v+v+ ; (44)7 KSKC 0 0.5 1 1.5 2 2.5IIVIII II∆=0 ∆=0.4 ∆=0.8 00.511.522.5 FIG. 4. (a) The modi cation of the phase portrait obtained from the renormalization analysis of tunneling events. The e ect of SOC is manifested on phase boundaries between the regions I-II and I-III. In the vicinity of mode freezing, hopping events of a single ( ")=(#) carrier become irrelevant. All shaded regions correspond to disconnected wires with no transport of carriers. and ^v=vq 12 4v2: (45) An amplidute t1;1corresponds to the interwire hopping event of a single ( ")=(#) electron, while t2;0andt0;2are tunneling amplitude of charge and spin, respectively. Based on the set of equations, one obtains phase por- trait atT= 0. It consists of four regions, exhibited also in the limit of weak scattering potential and it is shown in Fig. 4(a). For vanishing SOC ~K!K, the bound- aries of regions are de ned by1 Kc+1 Ks= 2;Kc= 2, andKs= 2. In region I (IV), hopping event of a single (")=(#) electron is irrelevant (relevant) and one eventu- ally renormalizes onto the xed point of disjoined (con- nected) wires in the RG process. In region II, the tunnel- ing amplitude t0;2for spin is relevant and grows upon RG transformation, however charge carriers can not tunnel. An opposite situation with a conducting charge channel and an insulating spin channel occurs in the region III. We note that the results obtained in the opposite limits of weak and strong barrier are consistent. For relatively weak SOC ( = 0 :4) the only pro- nounced e ect is modi cation of boundaries between re- gions I-II and I-III, as shown in Fig. 4(a). These e ects are dictated by Eqs. (42)-(43). The regions II and III are extended towards the region I with the new bound-ariesK1:8<2. The area of the extended region is larger for the larger value of  = 0 :8 with boundaries de ned via K1:6. Remarkably, these e ects are also exhibited in the weak barrier analysis, in Fig. 3(a). At the largest  = 0 :8, the tunneling of single carri- ers with amplitude t1;1becomes irrelevant near the phase separation point K= 5=2. This is consistent with the picture of mode freezing, discussed in the previous sec- tions. On the other hand, for weakly-interacting elec- trons withK1 the e ect of SOC is negligibely small, even for large values of , also compatible with the re- sults of the previous section. Conductance for insulating regions at nite tempera- tureTcan be also generalized as follows: G=e2 hX m;ndm;nt2 m;nT2(m2=~Kc+n2=~Ks2)(46) with dimensionless coecients dm;n. Experimentally relevant values of  =v0:2 [25, 26, 31] are smaller than the maximum value considered in this work. For this range of  values, one can neglect mode-freezing e ects safely. The results of the RG anal- ysis on the last sections arm that for weak and mod- erate e-e interactions e ects of SOC are negligible small. This results are in accrodance with recent theoretical and expertimental studies [31, 32]. V. CONCLUSIONS We studied carrier scattering e ects upon a single im- purity embedded to a Luttinger liquid with spin-orbit coupling using Abelian bosonization and pertubative renormalization techniques. Spin-orbit interaction de- grades spin-charge separation and renormalizes the TLL parameters Ks;cand the excitation velocities vs;c. We demonstrated that the scaling dimension of impurity op- erator is identical for both ( ")=(#) carriers. This im- plies the absence of conjectured spin- ltering e ect in Ref.[35]. The strongest e ects of spin-orbit coupling are pronounced for strong e-e interactions, whereas these ef- fects are negligibly small for moderate e-e interactions. Our main results are summarized by the phase portrait modi cations presented in Figs (3)-(4). ACKNOWLEDGMENTS We thank F. Yilmaz and V. I. Yudson for useful com- ments and discussions.8 Appendix: Expressions for new TLL parameters and excitations velocities TheFmatrix has the following form, F= 2^Fc^Fmix ^Fmix 2^Fs : (A.1) The matrix elements ^F=j!jFare expressed via the following integrals, F=Z+1 1dk 2 [G ]1 det(Q)! (A.2) and for the mixing element, Fmix=Z+1 1dk 2 [G mix]1 det(Q)! (A.3) where [G ]1are the inverse propogators of Eqs. (13)- (16) in~ q= (k;!) space, and the Qmatrix is de ned as, Q= 2[G c]1[G mix]1 [G mix]12[G s]1 : (A.4)For the mixing parameter Fmixthe numerator of the kernel is odd function of k, i.e [G mix]1=k!, whereas the denominator is an even function of k. This leads to the vanishing action S mix. Similarly, S mixvanishes and one is left with fully decoupled actions [Eqs. (27)-(28)]. Evaluation of integrals leads to the following results for new TLL parameters for displacement uctuation elds, ^K =v (v++v)~v2  v+v+ 1 ; (A.5) and for the phase elds, 1 ^K=v2  v(v+v+)v2 ~v v2~v+ 1 : (A.6) The sound velocities for the newly emerged modes are given by, v2 =1 2h ~v2 c+ ~v2 s+ 2p (~v2c~v2s+ 2)2+ 4~v2s2i : (A.7) For vanishing SOC  = 0, the sound velocities in the spin/charge channel are recovered. [1] S.-i. Tomonaga, Progress of Theoretical Physics 5, 544 (1950). [2] J. Luttinger, Journal of mathematical physics 4, 1154 (1963). [3] A. O. Gogolin, A. A. Nersesyan, and A. M. Tsvelik, Bosonization and strongly correlated systems (Cambridge university press, 2004). [4] F. D. M. Haldane, 14, 2585 (1981). [5] M. P. Fisher and L. I. Glazman, in Mesoscopic Electron Transport (Springer, 1997) pp. 331{373. [6] H. Steinberg, G. Barak, A. Yacoby, L. N. Pfei er, K. W. West, B. I. Halperin, and K. Le Hur, Nature Physics 4, 116 (2008). [7] H. Kamata, N. Kumada, M. Hashisaka, K. Muraki, and T. Fujisawa, Nature Nanotechnology 9, 177 (2014). [8] J. Voit, Reports on Progress in Physics 58, 977 (1995). [9] T. Giamarchi, Quantum physics in one dimension , Vol. 121 (Clarendon press, 2003). [10] E. Jeckelmann, Journal of Physics: Condensed Matter 25, 014002 (2012). [11] C. Kane and M. P. Fisher, Phys. Rev. Lett. 68, 1220 (1992). [12] A. Furusaki and N. Nagaosa, Phys. Rev. B 47, 4631 (1993). [13] C. Kane and M. P. Fisher, Phys. Rev. B 46, 15233 (1992). [14] A. Anthore, Z. Iftikhar, E. Boulat, F. D. Parmentier, A. Cavanna, A. Ouerghi, U. Gennser, and F. Pierre, Phys. Rev. X 8, 031075 (2018). [15] A. M. Chang, Rev. Mod. Phys. 75, 1449 (2003). [16] X. G. Wen, Phys. Rev. B 41, 12838 (1990). [17] M. Bockrath, D. H. Cobden, J. Lu, A. G. Rinzler, R. E. Smalley, L. Balents, and P. L. McEuen, Nature 397, 598(1999). [18] C. Blumenstein, J. Sch afer, S. Mietke, S. Meyer, A. Dollinger, M. Lochner, X. Y. Cui, L. Patthey, R. Matzdorf, and R. Claessen, Nature Physics 7, 776 (2011). [19] Y. Jompol, C. J. B. Ford, J. P. Griths, I. Farrer, G. A. C. Jones, D. Anderson, D. A. Ritchie, T. W. Silk, and A. J. Scho eld, Science 325, 597 (2009). [20] E. Bocquillon, V. Freulon, J.-. M. Berroir, P. Degiovanni, B. Pla cais, A. Cavanna, Y. Jin, and G. F eve, Nature Communications 4, 1839 (2013). [21] M. Hashisaka, N. Hiyama, T. Akiho, K. Muraki, and T. Fujisawa, Nature Physics 13, 559 (2017). [22] T. Vekua, S. I. Matveenko, and G. V. Shlyapnikov, JETP Letters 90, 289 (2009). [23] S. Brazovskii, S. Matveenko, and P. Nozieres, Journal de Physique I 4, 571 (1994). [24] T. Kimura, K. Kuroki, and H. Aoki, Phys. Rev. B 53, 9572 (1996). [25] A. V. Moroz, K. V. Samokhin, and C. H. W. Barnes, Phys. Rev. B 62, 16900 (2000). [26] A. V. Moroz, K. V. Samokhin, and C. H. W. Barnes, Phys. Rev. Lett. 84, 4164 (2000). [27] A. Iucci, Phys. Rev. B 68, 075107 (2003). [28] S. Datta and B. Das, Applied Physics Letters 56, 665 (1990), https://doi.org/10.1063/1.102730. [29] S. Nadj-Perge, S. M. Frolov, E. P. A. M. Bakkers, and L. P. Kouwenhoven, Nature 468, 1084 (2010). [30] C. Fasth, A. Fuhrer, L. Samuelson, V. N. Golovach, and D. Loss, Phys. Rev. Lett. 98, 266801 (2007). [31] Y. Sato, S. Matsuo, C.-H. Hsu, P. Stano, K. Ueda, Y. Takeshige, H. Kamata, J. S. Lee, B. Shojaei, K. Wick-9 ramasinghe, J. Shabani, C. Palmstrm, Y. Tokura, D. Loss, and S. Tarucha, Phys. Rev. B 99, 155304 (2019). [32] C.-H. Hsu, P. Stano, Y. Sato, S. Matsuo, S. Tarucha, and D. Loss, Phys. Rev. B 100, 195423 (2019). [33] Y. Hamamoto, K.-I. Imura, and T. Kato, Phys. Rev. B 77, 165402 (2008). [34] T. Hikihara, A. Furusaki, and K. A. Matveev, Phys. Rev. B 72, 035301 (2005). [35] K. Kamide, Y. Tsukada, and S. Kurihara, Phys. Rev. B 73, 235326 (2006). [36] M. Governale and U. Z ulicke, Solid State Communica- tions 131, 581 (2004), new advances on collective phe- nomena in one-dimensional systems. [37] M. Governale and U. Z ulicke, Phys. Rev. B 66, 073311(2002). [38] V. Gritsev, G. Japaridze, M. Pletyukhov, and D. Baeriswyl, Phys. Rev. Lett. 94, 137207 (2005). [39] A. Schulz, A. De Martino, P. Ingenhoven, and R. Egger, Phys. Rev. B 79, 205432 (2009). [40] N. Kainaris and S. T. Carr, Phys. Rev. B 92, 035139 (2015). [41] N. Kainaris, S. T. Carr, and A. D. Mirlin, Phys. Rev. B 97, 115107 (2018). [42] M. P. A. Fisher and W. Zwerger, Phys. Rev. B 32, 6190 (1985). [43] A. Caldeira and A. Leggett, Annals of Physics 149, 374 (1983).
0809.5119v1.Multi_terminal_Spin_Transport__Non_applicability_of_linear_response_and_Equilibrium_spin_currents.pdf
arXiv:0809.5119v1 [cond-mat.mes-hall] 30 Sep 2008Multi-terminal Spin Transport: Non applicability of linea r response and Equilibrium spin currents T. P. Pareek Harish Chandra Research Institute Chhatnag Road, Jhusi, Allahabad - 211019, India We present generalized scattering theory for multi-termin al spin transport in systems with broken SU(2) symmetry either due to spin-orbit interaction,magne tic impurities or magnetic leads. We derive equation for spin current consistent with charge con servation. It is shown that resulting spin current equations can not be expressed as difference of poten tial pointing to non applicability of linear response for spin currents and as a consequence equilibrium spin currents(ESC) in the l eads are non zero. We illustrate the theory by calculating ESC in t wo terminal normal system in presence of Rashba spin orbit coupling and show that it leads to spin re ctification consistent with the non linear nature of spin transport. PACS numbers: 75.60.Jk, 72.25-b,72.25.Dc, 72.25.Mk Spin transport has emerged as an important subfield of research in bulk condensed matter system as well in mesoscopic and nano system[1]. In macroscopic systems, the very definition of spin currents is still debated due to non conservation of spin in presence of SO interaction[2] . On the other hand in mesoscopic hybrid system since current is defined in the leads where SO coupling is absent,therefore, it has been assumed that Landauer- B¨ uttiker formula for charge current[3] (Eq. (9) in this manuscript), which determines current in leads in terms ofapplied voltagedifference multiplied by totaltransmis- sionprobability, can be straightawaygeneralizedforspin currents by replacing total transmission probability with some particular combination of spin resolved transmis- sion probabilities[4, 5, 6]. This simple generalization has been widely used in the literature to study spin depen- dent phenomena in nanosystems[4, 5, 6]. The Landauer- B¨ uttiker formula in its widely used form has inbuilt cur- rent conservation , i.e., total current is divergence less (divj=0) which follows from basic Maxwell equations of electrodynamics. Physically this implies that total cur- rent has neither sources nor sinks this would be true for spin currents as well if spin is conserved. However, in presence of spin-orbit interaction, magnetic impurities or non-collinear magnetization in leads, spin is no longer a conserved quantity, hence the spin currents can not be divergence less. Therfore a straight forward generaliza- tion of Landauer-B¨ uttiker charge current formula to spin currents can not be correct for spin non-conserving sys- tems. In view of the above discussion in this work we de- velopaconsistentscatteringtheoreticformulationofcou- pled spin and charge transport in multiterminal systems with broken SU(2) symmetry in spin space following B¨ uttiker’s work on charge transport [3]. The SU(2) sym- metry in spin space can be broken due to either SO inter- action, magnetic impurities or non-collinear magnetiza- tion in leads[7, 8]. Our analysis provides a correct gener- alization of Landauer-B¨ uttiker theory for spin transport. In particular we derive a spin currents equation (Eq. (8)inthismanuscript)consistentwithchargecurrentconser- vation. However, the resulting spin current equation can not be cast in terms of spin resolved transmission and re- flectionprobabilitiesmultiplied byvoltagedifferenceasis the casefor chargecurrent(Eq. (9) in this manuscript)[3]. Therefore, equilibrium spin currents are generically non zero.Moreover, it implies that linear response theory with respect to electric field is not applicable to the spin currents (equilibrium as well non-equilibrium).Thus spin currents are intrinsically non-linear in electrical circu its. However, this is not surprising since linear response is valid for thermodynamically conjugate variable. In an electrical circuit thermodynamically conjugate variable to electric field is charge current not the spin currents. We illustrate the theory by calculating ESC analytically for two-dimensional electron system with Rashba SO in- teraction in contact with two unpolarized metallic con- tacts. Our analytical formula for ESC clearly demon- strates that it is transfer of angular momentum per unit time from SO coupled sample to leads where SO inter- action is zero. Therefore,it is truly a transport current in contrast to equilibrium spin currents in macroscopic Rashba medium(see ref.[2, 9]). To formulate scattering theory for spin transport we consider a mesoscopicconductor with brokenSU(2) sym- metry in spin space connected to a number of ideal mag- netic and non-magnetic leads (without SO interaction) which in turn are connected to electron reservoirs. To include the effect of broken SU(2) symmetry, it is neces- saryto write spin scattering state in eachlead along local spin quantizationaxis. For magnetic leadslocal magneti- zation direction provides a natural spin quantization axis which we denote in a particular lead αbyˆmα(ϑα,ϕα) whereϑαandϕαis polar and azimuthal angle respec- tively. For nonmagnetic leads since there is no preferred spin quantization axis, hence we choose an arbitrary spin quantization axis ˆu(θ,φ) which is same for all nonmag- netic leads. Thus the most general spin scattering state in leadαwhich can be either magnetic or nonmagnetic2 is given by, ˆΨσ α(r,t) =/integraldisplay dENσ α(E)/summationdisplay n=1Φαn(r⊥)χα(σ)/radicalbig 2π/planckover2pi1vσαn(E) (aσ αn(E)eikσ αn(E)x+bσ αn(E)e−ikσ αn(E)x) (1) where Φ αn(r⊥) is transverse wavefunction of channel n andχα(σ) is corresponding spin wave function along chosen spin quantization axis, ˆuorˆmαsuch that S· ˆuχ(σ)=(σ/planckover2pi1/2)χ(σ) orS·ˆmαϕ(σ)=(σ/planckover2pi1/2)ϕ(σ) withσ= σ(σ=±1,representing local up or down spin compo- nents) for nonmagnetic and magnetic leads respectively. HereS= (/planckover2pi1/2)σis a vector of Pauli spin matrices and Nσ αisnumberofchannelswith spin σinleadα. The rela- tion between spin dependent wavevector kσ αn(E) and en- ergyEis specified by, E=/bracketleftbig /planckover2pi12k2 αnσ/2m+εαn+σ∆α/bracketrightbig , whereεαnis energy due to transverse motion, ∆ αis stoner exchange splitting in the magnetic lead α. The stoner exchange splitting is zero for nonmagnetic leads. The operators aσ αnandbσ αnare annihilation operator for incoming and outgoing spin channels in lead αand are related via the scattering matrix, bσ αm=/summationdisplay βnσ′Sσσ′ αm;βnaσ′ βn (2) The scattering matrix elements Sσσ′ αm;βnprovides scatter- ing amplitude between spin channel nσ′in leadβto spin channelmσin leadα. These scattering matrix elements will be function of energy E as well angles, ϑandϕ. The angular dependence of scattering matrix elements on polar and azimuthal angle arise due to broken SU(2) symmetry. Note that for noncollinear magnetization in leads and in absence of SO interaction and magnetic im- purities, the angular dependence is purely of geometric origin and is related to the angular variation of various magnetoresistance phenomena[10]. The current in spin channel σalong longitudinal direc- tionˆx(through a cross section of lead α) and the local spin quantization axis ˆuis defined as, ˆIˆxσ αˆu(t) =/planckover2pi1 2mi/integraldisplay/bracketleftBig ˆΨ†σ α(S·ˆu)∇xˆΨσ α−∇xˆΨ†σ α(S·ˆu)ˆΨσ α/bracketrightBig dr⊥. (3) Substituting for ˆΨσ αfrom Eq.(1) into Eq.(3, we get an ex- pression for spin current in terms of creation and annihi- lation operators. On the resulting expression we perform quantum statistical averages and after a lengthy algebra we obtain followingexpression for averagecurrentin spin channelσ(for brevity of notation we suppress the super- scriptˆxwritten in Eq. (3), /an}b∇acketle{tIσ αˆu/an}b∇acket∇i}ht=g h/integraldisplay∞ 0dE/summationdisplay βfβ(E)/bracketleftBigg Nσ αδαβ−/summationdisplay σ′mnS†σ′σ βm;αnSσσ′ αn;βm/bracketrightBigg (4) Wherefβ= 1/exp[(E−µβ)/kT]+1 is Fermi distribu- tion function with chemical potential µβand the pre- factorgequalsσ/planckover2pi1/2. The summation over σ′in Eq. (4)can take on values ±σcorresponding to two spin pro- jections along local spin quantization axis. The second term of Eq.(4) can be written explicitly in terms of spin resolved reflection and transmission probabilities as, /summationdisplay βσ′;mnS†σ′σ βm;αnSσσ′ αn;βm=/summationdisplay σ′;mnS†σ′σ αm;αnSσσ′ αn;αm +/summationdisplay β/negationslash=ασ′;mnS†σ′σ βm;αnSσσ′ αn;βm ≡/summationdisplay σ′Rσσ′ αα+/summationdisplay β/negationslash=ασ′Tσσ′ αβ(5) WhereRσσ′ ααandTσσ′ αβare spin resolved reflection and transmission probability in the same probe and between different probes respectively. In Eq.5 on right hand side spin resolvedreflectionandtransmissionprobabilitiesare summed over all possible input modes for a fixed output spin mode σin leadα. Because partial scattering matrix in spin subspaceis not unitary due to non conservationof spin hence this summation need not to be equal to num- ber of spin σchannels in lead α,i.e.Nσ α, rather it can have any value lying between zero and Nσ α. To determine Nσ αin terms of spin resolved reflection and transmission probabilities, consider a situation where current is in- jected from reservoir only in spin channels σin leadα. In this casechargeconservationrequiresthat this current should leavethe spin channel σthrough all other possible channels in the same lead as well in differing leads, which implies, Nσ α=/summationdisplay σ′Rσ′σ αα+/summationdisplay β/negationslash=ασ′Tσ′σ βα. (6) As we can see that Eq.(6) differs from Eq.(5) in a sub- tle way and are not equal because in general spin re- solved transmission or reflection probabilities can not be related among themselves by interchanging spin indices, i.e.,Tσ′σ αβ/ne}ationslash=Tσσ′ βαandRσ−σ αα/ne}ationslash=R−σσ αα( we will discuss constraints due to time reversal symmetry below). If we demand that sum in Eq.(5) also equals to Nσ αthen it would imply spin conservation which is incorrect in pres- ence of spin flip scattering or broken SU(2) symmetry. The inadvertent use of this charge conservation sum rule for spin degrees of freedom in Ref. [5, 6, 11] has led to incorrect spin current equation. Though the partial scat- tering matrix in spin subspace is not unitary,however, the full scattering matrix is unitary,i.e., SS†=S†S=I, therefore, if we sum over σalso in Eq.(5) or Eq.(6) then it should give total number of channels in leads α,i.e. N=Nσ α+N−σ α, and as a result we get the following sum rule for total transmission probability, Tβα=/summationdisplay σ′σTσ′σ βα=/summationdisplay σσ′Tσσ′ αβ=Tαβ (7) whereTαβis total tranmission probability. The net spin current flowing in lead αis defined as IS ˆuα=/an}b∇acketle{tIσ ˆuα/an}b∇acket∇i}ht+/an}b∇acketle{tI−σ ˆuα/an}b∇acket∇i}htwhile the net charge current flowing3 is given by sum of absolute values, i.e., Iq ˆuα=| /an}b∇acketle{tIσ ˆuα/an}b∇acket∇i}ht |+| /an}b∇acketle{tI−σ ˆuα/an}b∇acket∇i}ht |with pre-factor greplace by the electronic charge ein Eq. (4). Using Eqs.( 5),(6) in Eq.(4) we obtain net spin and charge current as, Is αˆu= (/planckover2pi1 2h)/integraldisplay∞ 0dE2fα(E)(R−σσ αα−Rσ−σ αα)+ /summationdisplay β/negationslash=ασ′/bracketleftBig fα(E)(Tσ′σ βα−Tσ′−σ βα)−fβ(E)(Tσσ′ αβ−T−σσ′ αβ)/bracketrightBig (8) Iq αˆu= (e h)/integraldisplay∞ 0dE/summationdisplay β/negationslash=ασ′[fα(E)−fβ(E)]Tαβ(9) Equation (8) is the central result of this work. We stress that Eqs. (8) and (9) are valid under most general condi- tions as we have not made any assumptions about sym- metries of the scattering region . It is instructive to note that in general Tσ′σ αβ/ne}ationslash=Tσ′−σ αβandRσ−σ αα/ne}ationslash=R−σσ ααthere- fore, spin current equation can not be simplified further and written in terms of difference of Fermi function mul- tiplied by transmission or reflection probabilities as is the case for charge current in Eq. (9) which is stan- dard Landauer-B¨ uttikerresult[3]. Hence the spin current given by Eq. (8) will be nonzero even when all the leads areatequilibrium, i.e., fα(E,µα) =f(E,µ),∀α, whereµ is equilibrium chemical potential. For sake of complete- ness we mention that equilibrium chargecurrentvanishes as is evident from Eq. (9). The preceding discussion im- plies that linear response for spin currents is not appli- cable in an electrical circuit where external perturbation is applied voltages which is conjugate to charge currents and not to the spin currents as discussed in introduction. Therefore,the most widely used equation for spin cur- rent,see Ref.[4, 6] obtained by a generalization of charge current Eq. (9) has to regarded as incorrect. In view of this the theoretical study of spin dependent phenomena in mesoscopic systems needs to be re-investigated. We can gain further insight into spin current by con- sidering non-equilibrium situation such that the chemi- cal potential at the different leads differ only by a small amount so that we can expand the Fermi distribution function around equilibrium chemical potential µas, fβ(E,µβ) =f(E,µ)+(−df/dE)(µβ−µ). In this case we can immediately notice from Eq. (8) that total spin cur- rent in non-equilibrium situation will have equilibrium as well non equilibrium parts of spin current. For ESC the full Fermi sea of occupied levels will contribute. There- fore even in non-equilibrium situation ESC cannot be neglected. Equilibrium spin currents in time reversal symmetric two terminal system: In time reversalsymmetric systems spin resolved transmission and reflection probabilities in Eq. (8) obey following relations i.e.,Rσσ′ αα=R−σ′−σ ααand Tσσ′ αβ=T−σ′−σ βα[7]. In this case the spin currents Eq. (8) further simplifies to (here we denote left and right termi-nals byLandRrespectively), Is,eq L,ˆu= (/planckover2pi1 2h)/integraldisplay∞ 0dE2f(E,µ)/bracketleftbig (R−σσ LL−Rσ−σ LL) +(Tσ−σ RL−T−σσ RL)+(T−σ−σ RL−Tσσ RL)/bracketrightbig ,(10) above equation gives spin current in Left terminal. Spin current in right terminal are obtained from the same equation by interchanging L↔R. On right hand side in Eq. (10), σand−σrefers to up and down spin states alongˆu. From the above equation and previous discus- sion it is evident that even in time reversal symmetric two terminal systems ESC are non zero. Incase SU(2) symmetry in spin space is preserved, the spin resolved transmission and reflection probabilities obey a further rotational symmetry in spin space, i.e, Tσσ′ αβ=T−σ−σ′ αβ, Rσσ′ αα=R−σ−σ′ ααand spin flip components are zero, which implies that spin currents are identically zero for all ter- minals as is evident from Eq. (10). This conclusion re- mains valid even for systems without time reversal sym- metry as can be seen easily from Eq. (8). The expression in Eq. (10) can be cast in a more useful form as (the details will be provided in Ref.[14]), Is,eq αˆu=1 2π/integraldisplay Trσ[{ΓαGrΓβGa+ (ΓαgrΓαga)}(σ·u)]f(E,µ)dEdk/bardbl.(11) In the above equation all symbols represents 2 ×2 matri- ces in spin space( in σ·ubasis) and trace is taken over spin space. Where Γ α,βrepresents broadening matrices due to contacts, Gr(a)are retarded and advanced Green function and gr,ais a off diagonal matrix in spin space defined as gr,a= [{0,Gr,a σ−σ},{Gr,a −σσ,0}]. First term in Eq. (11) corresponds to spin resolved transmission while the second and third term give spin resolved reflection probabilities as required by Eq. (10). Notice that the above formula can not be simply written in terms of transmission matrix and it is reminiscent of the charge currentformulaforinteractingsystemderivedinRef.[12]. In our case this happens for spin current because in pres- ence of SO interaction spin can not be described as a non-interacting object. Equilibrium spin currents in two terminal Rashba sys- tem:We now apply Eq.(11) to study ESC(zero tempera- ture)inafinite sizeRashbasampleoflength L, contacted by two ideal and identical unpolarized leads. The Hamil- tonian for two-dimensional electron system with Rashba SO interaction and short-range spin independent disor- der isH=/planckover2pi12k2/(2m∗)I+λso(σxky−σykx) +U(x,y), whereλsois Rashba SO coupling strength, U(x,y) is the random disorder potential and Iis 2×2 identity ma- trix. Neglecting weak localization effects, the disorder averaged retarded Green function including the effect of leads is given by[13], Gr,a(E,k) =E−/planckover2pi12k2 2m+iη(k)+λso(σxky−σykx) [(E−/planckover2pi12k2 2m∗+iη(k))2−(λsok)2] (12)4 withη(k) = (2γ(k) +/planckover2pi1 τ(k)) where τ(k) is momentum relaxation time due to elastic scattering caused by im- purities and γ(k) broadening due to leads. The contact broadeningmatricesinEq.(11)arediagonalinspinspace and defined as Γ 1,2= [{γ(k),0},{0,γ(k)}]. Physically significance of γ(k) is that it represents /planckover2pi1/2 times the rate at which an electron placed in a momentum state kwill escape into left lead or right lead, hence as a first approximation we can write, γ(k) =/planckover2pi1vx(k) L≡/planckover2pi12kcos(φ mL, whereφis angle with respect to xaxis. The impurity scattering time can be approximated as1 τ(k)≈/planckover2pi1k lel, where lelis elasticmean free path. With these inputs we can in- tegrate Eq. (11) over transverse momentum (multichan- nel case) and energy to obtain an analytical expression for equilibrium spin current. We find that ESC with spin parallel(antiparallel) to the ˆzorˆzaxis and flowing to thexdirection vanishes in both leads( Iˆxs ˆz≡σzvx=0, Iˆxs ˆx≡σxvx=0). The ESC with spin parallel(antiparallel) to theˆyaxis are nonzero and given by, I−ˆx,s L,+ˆy=I+ˆx,s R,−ˆy≅m∗λsoEFL 32π/planckover2pi12(2+(L lel)2).(13) In left lead ESC is polarizedalong+ ˆydirectionand flows outwards from sample to lead,i.e.,along −ˆxdirection, while in the right lead ESC is polarized along −ˆyand flows outwards from sample to lead,i.e.,along + ˆxdirec- tion. Physically this implies that spin angular momen- tum is generated in sample with SO coupling which then flows outwards in the regions where SO coupling is zero, i.e., the left and right leads. This implies a spin rectifi- cation effect which can only occur if the transport is non linear and we see that this consistent with nonlinear na- ture of spin currents as remarked earlier. It is important to note that due to ESC there is no net magnetization in the total system (sample+leads) which is consistent with the Kramer’s degeneracy. We can gain a deeper understanding of the above ex- pression if we analyze the systems using additional sym- metries. The disorder averaging establishes reflection symmetry with respect to x(y)axis and the system has a symmetry related with the operator σyRy. (σxRx). As a result the total symmetry operator(time rever- sal+reflection) for the system is UT R=ItσxRxσyRy= It(iσz)RxRy, whereItis time reversal operator. Under this symmetry operation the disorder averaged system is invariant and the spin current operators σxvx,σyvx are even while σzvxis odd. Therefore the spin current alongˆzdirectionvanishes while in-planespin currentcanbe non zero. However as we have seen above that only theycomponent of spin current survives after integrat- ing over momentum. Fig.(1) illustrate the conservation ofσyvxunder these symmetry operation. The depen- dence of ESC on λsocan also be inferred from symmetry consideration. As we can check easily that the Rashba SO interaction changes sign under reflection along ˆz axis(λso(σxky−σykx)/mapsto→ −λso(σxky−σykx)). Physi- cally it corresponds to reversing the asymmetry of con- fining potential along ˆzaxis. Therefore the spin currents xy zspin parallel to + y direction flow parallel to −ve x directionspin parallel to −ve y directionflow parallel to +ve x direction− FIG. 1: Fig 1: The figure illustrate the conservation of spin currentσyvxif th system is rotated by πalongˆzaxis, which represents reflection along and y axis respectively. Under t his transformation configuration in left and right goes over int o each other hence the spin current remains invariant. (equilibrium as wellnon-equilibrium) can only depend on the odd powers of spin orbit coupling constants λso. Ac- cording to Eq. (13) ESC are proportional to λsowhich is consistent with these symmetry consideration and ESC vanishes if λsozero as expected on physical grounds. It is worth noting that even the local ESC in macroscopic Rashba medium, discussed in Ref.[2] are proportional to λ3 soandaswehaveseenthisisaconsequenceofsymmetry consideration. Moreover ESC are proportional to length of SO region because SO region acts as source of these currents. Note that the right hand-side of Eq. (13) has dimension of angular momentum per unit time signifying that these currents are truly transport current. To conclude, we have derived spin current formula for multiterminal spin transport for system with broken SU(2) symmetry in spin space. We have demonstrated that spin currents are fundamentally different from the charge currents and the ESC are generically non zero. In view of this the spin transport phenomena in mesoscopic system needs a fresh look. Further it will also be inter- esting and desirable to study different magnetoresistance phenomena from the perspective of spin currents. I acknowledge helpful discussion with M. B¨ uttiker and A. M. Jayannavar. [1] I. Zutic, J. Fabian, and S. Das Sarma, Rev. Mod. Phys. 76, 323 (2004). [2] E. I. Rashba, Phys. Rev. B 68, 241315(R) (2003). J. Shi, P. Zhang,D. Xiao, and Q. Niu, Phys. Rev. Lett. 96, 076604 (2004). [3] M. B¨ uttiker, Phys. Rev. B. 46, 12485 (1992).,M. B¨ uttiker, Phys. Rev. Lett. 57, 1761 (1986). [4] J. H. Bardarson, I. Adagideli and P. Jacquod, Phy. Rev. Lett.98196601, (2006). W. Ren,Z. Qiao,S. Q. Wang and H. Guo,Phy. Rev. Lett. 97066603, (2006). E. M. Hankiewicz, L. W. Molenkamp,T. Jungwirth and J. Sinova,Phy. Rev. B. 70241301 (2004). B. K. Nikolic,5 L. P. Zrbo and S. Souma, Phys. Rev. B 72075361(2005). [5] Y. Jiang and L. Hu, Phys. Rev. B 75, 195343 (2007). [6] M. Scheid, D. Bercioux, and K. Ritcher, New. J. Phys. 9, 401 (2007). [7] T. P. Pareek, Phys. Rev. Lett 92, 076601 (2004). [8] T. P. Pareek, Phys. Rev. B. 66, 193301 (2002). T. P. Pa- reek, Phys. Rev. B. 70, 033310 (2004). [9] R. H. Silsbee, J. Phys:Cond. Matter 16, R179-R207 (2004).[10] A. Brataas,Yu. V. Nazarov, and G. E. W. Bauer, Phys. Rev. Lett. 842481 (2000). [11] A. A. Kiselev and K. W. Kim, Phys. Rev. B 71, 153315 (2005). [12] Y. Meir and N. S. Wingreen, Phys. Rev. Lett. 682512 (1992). [13] M. A. Skvortsov, JETP Lett. 67, 133 (1998). [14] T. P. Pareek manuscripr under preparation.
1705.04427v1.Analytical_slave_spin_mean_field_approach_to_orbital_selective_Mott_insulators.pdf
Analytical slave-spin mean- eld approach to orbital selective Mott insulators Yashar Komijani1;, and Gabriel Kotliar1, 1Department of Physics and Astronomy, Rutgers University, Piscataway, New Jersey, 08854, USA (Dated: October 29, 2021) We use the slave-spin mean- eld approach to study particle-hole symmetric one- and two-band Hubbard models in presence of Hund's coupling interaction. By analytical analysis of Hamiltonian, we show that the locking of the two orbitals vs. orbital-selective Mott transition can be formulated within a Landau-Ginzburg framework. By applying the slave-spin mean- eld to impurity problem, we are able to make a correspondence between impurity and lattice. We also consider the stability of the orbital selective Mott phase to the hybridization between the orbitals and study the limitations of the slave-spin method for treating inter-orbital tunnellings in the case of multi-orbital Bethe lattices with particle-hole symmetry. INTRODUCTION Iron-based superconductors are the subject of inten- sive study in the pursuit of high-temperature super- conductivity [1{7] . These systems are interacting via Coulomb repulsion and Hund's rule coupling and they require the consideration of multiple bands with crys- tal eld and inter-orbital tunnelling [8, 9]. Early DMFT studies, pointed out the importance of the corrlations [10] and Hund's rule coupling [11], and reported a notice- able tendency towards orbital di erentiation, with the dxyorbital more localized than the rest [12]. They also demonstrated orbital-spin separation [13{15]. Note that the orbital di erentiations has been recently observed in experiments [16]. Another perspective on the electron correlations in these materials is that the combination of Hubbard in- teraction and Hunds coupling place them in proximity to a Mott insulator [17] and, correspondingly, the role of the orbital physics is provided by the orbital selective Mott picture [18, 19]. Ref. [18] demonstrated an orbital selec- tive Mott phase in the multi-orbital Hubbard models for such materials, in the presence of the inter-orbital kinetic tunneling. In such a phase, the wavefunction renormal- ization for some of the orbitals vanishes. Such a phase has been observed in angle-resolved photo-emission spec- troscopy (ARPES) experiments [20, 21]. Although desir- able, these e ects have not been understood analytically in the past, partly due to the fact that an analytical study is dicult for realistic models. However, there are simpler models, capable of capturing part of the relevant physics, which are amenable to such analytical under- standing, and this is what we study in this paper. The mean- eld approaches to study these problems rely on various parton constructions or slave-particle techniques. The latter include slave-bosons [22, 23], Kotliar-Ruckenstein four-boson method [24] and its rota- tionally invariant version [25], slave-rotor [26], Z2slave- spin [27{30] and its U(1) version [18, 31], slave spin-1 method [32] and the Z2mod-2 slave-spin method [33, 34]. For a comparison of some of these methods see Appendix A. While these methods are all equivalent in the sensethat they are exact representation of the partition func- tion if the degrees of freedom are taken into account ex- actly, di erent approximation schemes required for an- alytical tractability, lead to di erent nal results and therefore they have to be tested against an unbiased method like the dynamical mean- eld theory (DMFT) [36{44] in large dimensions or density function renormal- ization group (DMRG) [45] in one dimension. We use the Z2slave-spin [27{30] in the following to study the orbital selectivity with and without Hund's coupling. We brie y go through the method for the sake of completeness and setting the notations. By study- ing the free energy analytically we develop a Landau- Ginzburg theory for the orbital selectivity. A Landau-like picture has been useful in understanding the Mott tran- sition in in nite dimensions. Using a Landau-Ginzburg approach, we show how the interaction in the slave-spin sector tend to lock the two bands together in absence of Hund's coupling and that the Hund's coupling promotes orbital selectivity. We also apply the method to an im- purity problem ( nite- UAnderson impurity) and its use as an impurity show that the slave-spin mean- eld re- sult can be understood as the DMFT solution with an slave-spin impurity solver. This puts the method in per- spective by showing that the mean- eld result is a subset of DMFT. Additionally, we study the e ect on the orbital selective Mott phase produced by inter-orbital kinetic tunnelling and point out to some of the limitations of the slave-spin for treating such inter-orbital tunnelling in particle-hole symmetric Bethe lattices. Finally, we study study the instability of the orbital selective Mott phase by including hybridization between the two orbitals. Z2Slave-spin method We consider the Hamiltonian H=H0+Hint, where H0=X hiji t ijdy i dj (1) We must demand t ij= [t ji]for this Hamiltonian to be Hermitian. Unless mentioned explicitly, is a super-arXiv:1705.04427v1 [cond-mat.str-el] 12 May 20172 index that contains both spin and orbital degrees of free- dom. We replace the d-fermions with the parton con- struction [27] dy i = ^zi fy i ; ^zi =x i : (2)  i ,=x;y;z are SU(2) Pauli matrices acting on an slave-spin subspace per site/spin/ avour, that is intro- duced to capture the occupancy of the levels. Slave-spin statesj*i iandj+i icorrespond to occupied/unoccupied states of orbital/spin at sitei, respectively. Away from half- lling, [28] has shown that x i has to be replaced with+ i =2+c  i =2 wherecis a gauge degree of freedom and is determined to give the correct non-interacting re- sult. Here, for simplicity we assume ph(particle-hole) symmetry and thus maintain the form of Eq. (2). Note that this parton construction has a Z2gauge degree of freedomx;y!x;yandf!f, thus the name Z2 slave-spin. The representation (2) increases the size of the Hilbert space. Therefore, the constraint 2fy i fi =z i + 1; (3) to imposed to remove the redundancy and restrict the evolution to the physical subspace. Using Eqs. (2,3) it can be shown that the standard anti-commutation rela- tions ofd-electron are preserved. Plugging Eq. (2) in H0, and imposing the constraint (on average) via a Lagrange multiplier, we have H0=X hiji t ijfy i fj ^zy i ^zj i [fy i fi (z i + 1)=2] On a mean- eld level, the transverse Ising model of slave- spins can be decoupled from fermions. The decoupling is harmless in large dimensions [46] as the leading op- erator introduced by integrating over the fermions be- comes irrelevant at the critical point of the transverse Ising model. Therefore, writing H0Hf+H0S, we have H0S=X hiji J ijh ^zy i ^zj Q iji +X i z i =2; Hf=X hiji ~t ijfy i fj i (fy i fi 1=2) (4) where ~t ij=t ijQ ijwithQ ij=h^zy i ^zj iis the renor- malized tunnelling and J ij=t ijhfy i fj iis an Ising coupling between slave-spins. The advantage of the par- ton construction (2) is that the interaction Hintfgcan be often written only in terms of the slave-spin variables, so thatH=Hf+HSandHS=H0S+Hint. Particle-hole symmetry - phsymmetry on the orig- inal Hamiltonian is de ned as ( nis a site index) dn !(1)ndy n ; dy n !(1)ndn (5) On a bipartite lattice, the nearest neighbor tunnelling term preserves phsymmetry, even in presence of inter- orbital tunnelling. So, if the system is at half- lling theHamitonian is invariant under phsymmetry. We have to decide what phsymmetry does to our slave-spin elds. We choose fn !(1)nfy n ; x n !x n ; z n !z n (6) So, we see that if the original Hamiltonian had ph symmetry, we necessarily have i = 0. Single-site approximation - The Hamiltonian HSis a multi- avour transverse Ising model which is non-trivial in general. Following [27{34] we do a further single-site mean- eld for the Ising model, exact in the limit of large dimensions: ^zy i ^zj h^zy i i^zy j + ^zy i h^zj ih^zy i ih^zj i;(7) The last term together with the second term of Eq. (8) contributes a2P hiji J ijQ ij. We de ne zi =h^zi i andZi =jzi j2as the wavefunction renormalization of orbital at sitei. The slave-spin Hamiltonian becomes (using the symmetry of J ij) H0S=X i (h i ^zi +h:c:); hi =X j J ijzj (8) In translationally invariant cases hi andzi become in- dependent of the site index and J ijdepends on the dis- tance between sites iandj. Therefore, we can simply writeh =P J z where J X (ij)J (ij)=X jt ijD fy i fj E : In absence of inter-orbital tunnelling, Jis a diagonal ma- trix, corresponding to individual orbitals, where for each orbitalJ =RD D d ()f()is the average kinetic en- ergy and depends only on bare parameters, una ected by the renormalization factor z. For semicircular band (Bethe lattice),J=0:2122D, while for a 1 Dtight- binding modelJ1D=0:318DwithD= 2t. Since the operator ^z =x is Hermitian, we can write the slave- spin Hamiltonian (for each site) as [47] HS=X a x +Hint (9) wherea = 2P J z (at half- lling). The only non- trivial part of computation is the diagonalization of HS. This is a 4Mdimensional matrix where Mis the number of orbitals. The free energy (per site) is F=1 X nkTr log[G1 f(k;i!n)]2X nJ z z 1 logn Trh e HSio : (10) Here = 1=Tis the inverse temperature and the second part comes from two constants introduced in Eqs. (4) and3 (7). At zero temperature, the rst term is just J z z and the last term is ESwhich depends on zviaa. Hence, F=X J z z +ES(fag): (11) ONE-BAND MODEL In the one-band case the interaction is Hint= UP i~ni"~ni#where ~nini1=2. Representing the latter with z i=2 and using translational symmetry we obtainHint!(U=4)z "z #. Since we are in the para- magnetic phase ( a"=a#), only sum of the two spins 2~T=~ "+~ #enter (the singlet decouples) and the Hamil- tonain can be written as HS= 2aTx+U 2(Tz)2U=4, creating a connection to the spin-1 representation of [32]. Furthermore, we can form even and odd linear combina- tions of the empty and lled states and at the half- lling, only the even linear super-positions enters the the Hamil- tonian. Thus, choosing atomic states of HSas j 0i=j*ij+ip 2;j 1i=j*+ijOip 2(12) withE0=U=4 andE1=U=4, the Hamiltonian can be written as HS= 2ax+ (U=4)zwhere~ are Pauli matrices acting between j +0iandj +1i, i.e. it reduces to the Z2mod-2 slave-spin method [33, 34]. In writing the states in Eq. (12) we have used a short-hand notation (also used in the next section) j*"+#i ! j*i andj+"*#i!j+i ,j*"*#i!j*+i and so on. The in- set of Fig. (1b) shows a diagrammatic representation of the slave-spin Hamiltonian and two states decouple. The ground state of HSis that of a two-level system ES=U 4p 1 + (4 =U)2 (13) with the level-repulsion = 2aand the zero-temperature (free) energy is given by [factor of 2 sdue to spin] F= 2sjJjz2+ES(z) (14) The free energy is plotted in Fig. (1a) and it shows a second-order phase transition as Uis varied. Close to the the transition !0, we can approximate ES 2 2=U+ 8 4=U3. Writing the rst term of the free en- ergy as + 2=8jJj, we can read o the critical interaction UC= 16jJj. Minimization of the free energy gives the Gutzwiller projecion fomrula of Brinkman and Rice [48] Z=1u2u<1 0u>1(15) withu=U=UCand is plotted in Fig. (1b). At nite temperature this procedure gives a rst order transition terminating at a critical point [34]. FIG. 1: (color online) (a) Free energy (at T= 0) as a function of zshowing a second-order phase transition as U=UCis varied. (b) Wavefunction renormalization Z=jzj2as a function of Uhas the Brinkman-Rice form. Inset: Diagrammatic representation of the slave-spin Hamiltonian. Each dot denotes on atomic state. Two states decouple and HSis equivalent to that ofZ2mod-2 slave-spin. Spectral function - The Green's functions of the d- fermionsGd() Td()dy (0) factorizes Gd;() Tf()fy (0) hTx ()x (0)i (16) to thef-electron and the slave-spin susceptibility and thus the spectral function is obtained from a convolu- tion with the slave-spin function Ad(!) =Af(!)AS(!), in whichAfis a semicircular density of states with the widthZand within single-site approximation ASis AS(!) =Z(!) +1Z 2[(!+ 2ES) +(!2ES)] (17) The spectral density has the correct sum-rule (in contrast to the usual slave-bosons [22, 23]) since the commuta- tion relations of the slave-spins are preserved. However, the single-site approximation does not capture incoherent processes, and this re ects in sharp Hubbard peaks in the Mott phase ( Z= 0) where Af=(!). Also, the spatial independence of the self-energy implies that the inverse e ective-mass of \spinons" m=~m=Z[1 + (m=kF)@k] is zero in the Mott phase. This is again an artifact of the single-site approximation. Both of these problems are remedied, e.g. by doing a cluster mean- eld calcu- lation [28, 33] or including quantum uctuations around the mean- eld value within a spin-wave approximation to the slave-spins [33]. The fact that (beyond single-site approximation) spinons disperse in spite of hxi! 0 and they carry a U(1) charge as seen by Eq. (2), implies that vanishing of hxidoes not generally correspond to the Mott phase in nite dimensions. However, in large dimensions, this is correct [34] and that is what we refer to in the following.4 TWO-BAND MODEL In absence of inter-orbital tunnellings, the free-energy is F=a2 1=2jJ1j+a2 2=2jJ2j+ES(a1;a2) (18) whereESis the ground state of the slave-spin Hamilto- nian. For two bands we have the interaction Hint=U(~n1"~n1#+ ~n2"~n2#) +U0(~n1"~n2#+ ~n1#~n2") + (U0J)(~n1"~n2"+ ~n1#~n2#) +HXP (19) where ~n nf 1=2 =z =2. The spin- ip and pair- tunnelling terms are HXP=JX[dy 1"d1#dy 2#d2"+dy 1#d1"dy 2"d2#] +JP[dy 1"dy 1#d2#d2"+dy 2"dy 2#d1#d1"]:(20) This term mixes the Hilbert space of f-electron with that of slave-spins. Following [27{29] we include this term ap- proximately by dy !+ andd ! substitution so that it acts only in the slave-spin sector. The justi cation is that such a term captures the physics of spin- ip and pair-hopping. Using the spherical symmetry U0=UJ this can be written as Hint=U 2(~n1"+ ~n1#+ ~n2"+ ~n2#)2U 2+HXP J[~n1"~n2#+ ~n1#~n2"+ 2~n1"~n2"+ 2~n1#~n2#] (21) ForJX=JandJP= 0 it has a rotational symmetry [49]. Alternatively, U0=U2JandJX=JP=J has rotational symmetry. The choice does not a ect the discussion qualitatively. We keep the former values in the following. Atomic orbitals - We start by diagonalizing the atomic Hamiltonian in absence of the hybridizations. Close to half- lling the doubly-occupied states have the lowest en- ergy and are given by j 0i=j*1*2ij+ 1+2ip 2; E0=UJ=2; j 1i=j*1+2ij+ 1*2ip 2; E1=U+J=2JX; j 2i=j*+ 1;O2ijO1*+2ip 2; E2=U+ 3J=2JP; These 3 doublets become the 6-fold degenerate ground state when J!0. The 1;3-particle states are then next j 3i=j*+ 1ij*2ij+ 2ip 2; E3=1; j 4i=jOi1j*2ij+ 2ip 2; E4=1 j 5i=j*1ij+ 1ip 2j*+ 2i; E5=2; j 6i=j*1ij+ 1ip 2jOi2; E5=2;and nally, there are two (empty and quadruple occu- pancy) states at the top of the ladder j 7i=j*+i1j*+i2; E 7=1+2+ 3U3J=2; j 8i=jOi1jOi2; E 8=12+ 3U3J=2: No Hund's rule coupling - The hybridization causes transition among atomic states. In the case of no Hund's coupling we can block diagonalize HSinto several sectors and diagrammatically represent it as shown in Fig. (2). Therefore, the calculation can be reduced from 16 16 to 55. The larger the level-repulsion, the lower the ground state energy in each sector. The fact that the slave-spins decouple into several sectors brings about the possibility of possible ground-state crossings between various sectors as the parameters a1anda2are varied. Here, however, it can be shown that the sector Chas the lowest ground state energy for arbitrary parameters. FIG. 2: Diagrammatic representation of the slave-spin Hamiltonian HSin the two-band model with J= 0 and 1=2= 0. Each dot represents an atomic state with a certain energy, denoted on the left, whereas the connecting lines represent o -diagonal elements of the Hamiltonian matrix, all assumed to be real. We have used the short-hand notationp 2 S=D a;b a b. Also note that ai= 2Jizi. The Hamiltonian factorizes into several sectors. Numerical minimization of the free-energy leads to Fig. (3) which reproduces the results of [27]. For t2=t1> 0:2 the metal-insulator transition happens at the same criticalUfor the two bands and we refer to it as the locking phase , whereas for t2=t1<0:2 the critical Ufor the bands are di erent U2< U 1and we refer to it as orbital selective Mott (OSM) phase . In order to have the result analytically tractable we do one further simpli cation and that is to project out the zero and quartic occupancies per site, by dropping the high energy site at the apex of sector C. We expect such an approximation to be valid close to the Mott transition of the wider band, but invalid at low U. As a result5 FIG. 3: (color online Wavefunction renormalizations Z1 (blue) andZ2(green) as a function of U=UC1in absence of Hund's rule coupling J= 0. The states at the bottom row correspond to doubly occupied sites. The middle-row states have occupancy of 1 or 3 and the states at the top row correspond to zero or four-electron llings. (a) Moderate bandwidth anisotropy t2=t1= 0:5 shows locking. (b) Large bandwidth anisotropy t2=t1= 0:15 can unlock the bands and cause OSM transition (OSMT). We also reproduce the kink in the wider-bandwidth (blue) band as the narrow band transitions to the Mott phase [27], marked with an arrow. In the OSM phase, the wavefunction renormalization of the wider band follows the Brinkman-Rice formula (solid line). the sectorCdecouples into two smaller sectors C, each equivalent to a two-level system with the level-repulsions =q a2 1(3=2 +p 2) +a2 2(3=2p 2) q a2 1(3=2p 2) +a2 2(3=2 +p 2):(22) The ground state energy of the slave-spin sector is deter- mined with +inserted in the ESexpression (13) (after an inertU=4 energy shift). Note that this ground state has theZ2symmetrya1$a2of the Hamiltonian HS. ES( +) as a function of ( a2 1a2 2)=(a2 1+a2 2), is mini- mized fora1=a2. Discarding empty and lled states corresponds to truncating part of the Hilbert space and thus leads to reduced wavefunction renormalization at U0. In Fig. (4) we have compared our analytical so- lution to that of the exact result. When a2= 0, Eq. (22) gives !2a1as in the single-band case and there- fore, same critical interaction UC1= 16jJ1jis obtained. But for symmetric bands a1=a2, it gives = 2p 3a. Following similar analysis as before, the free energy is a2=jJj 2 2=Uand we obtain UC= 24jJj= 1:5UC1in agreement with [27, 29]. Locking vs. OSM phase - We formulate the locking vs. OSM question as the following. Under what condition, a1>0 buta2= 0 can be a minima of the Free energy. As mentioned before, setting a2= 0, in Eq. (22) reduces to the one-band !2a1. Therefore, the Mott transition for the wide band happens at the same critical Uas before. To have a non-zero a1solution, we must have U < Uc1. The pointa2= 0 always satis es dF=da 2= 0. To ensure that it is the energy minima we need to check the second FIG. 4: (color online) A comparison of numerical minimization of the free energy vs. the analytical two-level system. Discarding the empty and full occupancy states leads to underestimation of ZasU!0 but close to the Mott transition the approximation is accurate. derivative d2F da2 2 a2=0=1 jJ2j5 jJ1j>0; (23) which gives the condition jJ2=J1j<0:2. We can better understand the transition by using an order parameter. The trouble with the expression of is that it cannot be Taylor expanded when a1anda2are both small. However, we may assume a2=ra1, with ras an order parameter replacing a2, and write down (a1;a2) =a1 (r) where (r) = +(a1!1;a2!r). A niterclose to the transition implies locking whereas r= 0 orr=1implies OSM phase. Close to the transition of both bands 0 and we can write ES2 2=U+ 8 4=U3and Eq. (18) becomes F(a1;r) =a2 1W 2jJ1j+O(a4); Wx(r;u) = 1 +xr2 2(r) 4u Here,x=jJ1=J2j, andu=U=UC1. The metal-insulator transition for a1happens when the mass coecient W changes sign. For negative W,a2 1>0 and we still have to minimize the free energy with respect to r. At smallr, we can expand (r)2 + 5r2. To zeroth order in r, theW- sign-change happen at u= 1. Another transition from r= 0 tor > 0 happens when the corresponding mass term (x5=u)r2changes sign, giving the same critical bandwidth ratio xc= 5 as we had before. So we have two equationsW(r;u) = 0 and@rW(r;u) = 0. The function6 Wis plotted in the gure and the transition from locking r>0 to OSM phase r= 0 are shown. FIG. 5: The coecient W(r;u) is shown for various uas function of r=a2=a1. Equations W= 0 and@rW= 0 are satis ed at the minimum of the red curve, which is (a) at a niter= 1 in the Locking phase, jJ1j=jJ2j. (b) and zero r= 0 in the OSM phase, jJ1j5jJ2j. Large Hund's coupling - In presence of Hund's cou- pling the slave-spin Hamiltonian is modi ed to the dia- gram shown in Fig. (6). FIG. 6: Diagrammatic representation of the slave-spin Hamiltonian HSin the two-band model at half- lling with in presence of Hund's rule coupling J. Various degeneracies are lifted by J-interaction. In the limit of large Hund's couplingJ=U!1=4 we may only keep sector Cand neglect all the gray lines. The ground state still belongs to the sector C. In the limit of large J=U!1=4, we may ignore all the gray lines on the block Cand nd that the ground state is that of a two-level system, Eq. (13) with the level-repulsion = 2q a2 1+a2 2 (24) It is remarkable that the (orbital) rotational invariance of the model (even though absent in HS) is recovered in this ground state. When the two bands have the same bandwidth, this formula predicts UC=UC1. SinceESno longer depends on a2 1a2 2, there is no more compe- tition between the two terms and an slight bandwidth asymmetry lead to OSM phase. This can be formulated again, following previous section, in terms of stability of aa16= 0 buta2= 0 solution. We can check that d2F da2 2 a2=0=1 jJ2j1 jJ1j>0; (25) which givesjJ2j<jJ1jas the sucient condition for OSMT, i.e., any di erence in bandwidth drives the sys- tem to the OSM phasse. Alternatively, by expanding the level-repulsion in this case (r)2 +r2and plugging it intoW(r;u), we nd that the critical bandwidth ratio xc=jJ1=J2jis equal to one. TUNNELLING BETWEEN THE ORBITALS A very interesting question is about the fate of or- bital selective Mott phase upon turning on an inter- orbital tunnelling. The band in Mott insulating phase has one electron per site forming localized magnetic mo- ment. There is a large entropy associated with this phase and it is natural to expect that it would be unstable to- ward possible ordering. A possible mechanism that can compete with magnetic ordering, is the Kondo screening of the insulating band by the itinerant band, leading to conduction in the former and opening a hybridization gap in the latter band (e ectively a new locking e ect com- ing from Kondo screening). Within single-site approxi- mation, however, the form of the renormalized coupling ~t ij=z i t ijzj implies that once an orbital goes to the Mott phase, it automatically shuts down its coupling to all the other orbitals. We speculate that this e ect might be responsible for the orbital selective Mott transition so- lution found in [18]. However, it is still a valid question whether or not the critical interactions UCfor a Mott transition are modi ed by inter-orbital tunnelling, which we explore in the following. Before treating inter-orbital tunnelling, we discuss how the slave-spin method can be applied to the impurity problem, and its relation to the lattice. Impurity vs. Lattice and the DMFT loop We can also apply the slave-spin method to an im- purity problem. In particular, we can use the slave-spin (as well as any other slave-particle) method as an im- purity solver for the DMFT. We show in the following that the slave-spin mean- eld result corresponds to such a DMFT solution with the corresponding slave-spin im- purity solver. This puts the method on rm ground and allows comparison between various methods. First, consider a generic phsymmetric impurity model described by the Hamiltonian H=H0+Hint7 where H0=X k t k(dy ck +h:c:) +X k  kcy k ck :(26) Again ; are superindices that include both orbital and spin. We have assumed that the bath is diagonal and discarded any local `crystal eld' dy 1d2for simplic- ity. In the simple case of single-orbital impurity Hint= U~nd"~nd#. Via a substitution of Eq. (2), the hybridiza- tion term becomes H0=P k t k(fy x ck +h:c:). This problem can be written in a similar way as be- foreHHf+HSwhereHSis exactly what we had in single-band lattice case. However, since the fy x ck interaction happens only on the impurity site, we do not need the second single-site approximation here, and ob- taina =2P k t khfy k ck i. In order to have a gen- eral formalism that applies to both impurity and lattice, as well as scenarios with inter-orbital tunnelling for which J renormalizes and is dicult to compute, we regard a andzas independent variables and write the free energy of Eq. (12) as [47] F(fz;ag) =Ff(fzg) +FS(fag)X a z : (27) The saddle-point of Fwith respect to aandzgives the correct mean- eld equations. Ffis the free energy of thef-electron given by Ff=TP nTr log[G1 f(i!n)] whereG1 f(i!n) =i!n1zy(i!n)zwith(i!n) =P ktyGc(k;i!n)t, the hybridization function. The slave- spin part is given by FS= Tr[e HS] where for a single- orbital Anderson impurity, HS= 2ax+Uz=4, as we had in the single-band case before. The mean- eld equations w.r.t zandaare, respec- tively a =1 z Zd! f(!)!Im G f(!+i) ; (28) z =dFS da : (29) The rst equation provides a relation between aand zthat generalizes a = 2P J z (see appendix D). Having expressions for ES(a) we can eliminate ain fa- vor ofz, or vice versa, which is equivalent to a Legendre transformation. In the appendix C, we apply these equa- tions to the (single-orbital) nite- UAnderson impurity problem and show the `transition' to the Kondo phase as the temperature is lowered. In a lattice, the free energy has the same form as Eq. (27) with the di erence that Ff= TP k;nTr log[G1 f(k;i!n)] where the Green's func- tion isGf(k;!+i) = [(!+i)1zyEkz]1. It can be shown that exactly same mean- eld equations are ob- tained ifGfin Eq. (28) is replaced with G f(!+i)!P kG f(k;!+i). Therefore, we conclude that the twoproblems (lattice and impurity) are equivalent provided that the hybridization function in the impurity problem is chosen such that the impurity Green's function and the local Green's function of the lattice are equal, i.e. [i!n1zy(i!n)z]1=X k[(!+i)1zyEkz]1:(30) which is the DMFT consistency equation. Therefore, slave-spin mean- eld is equivalent to a DMFT solution using the slave-spin method as the impurity solver. Also, note that a lattice problem in the OSM phase, corre- sponds to an impurity problem in which the hybridiza- tion of one of the orbitals to the bath has been turned o [50]. See also Appendix B. Inter-orbital tunnelling Slave spins have been used to study Iron-based su- perconductors [18] where the inter-orbital tunnelling are important. We study this tunnelling e ect in the speci c case withphsymmetry and without orbital-splitting (which allows for analytic calculations). The cases that go beyond such conditions, as arising in the models for the iron-based superconductors [18], remain to be ex- plored and are left for future work. A troublesome feature of the slave-spins is that they break the rotational symmetry among the orbitals. Within the phsymmetric Bethe lattices that we study here, this rotational variation leads to ambiguities in presence of inter-orbital tunnellings, as we point out here. Let us consider a 1D chain with two orbitals H0= P n(Dy nTDn+1;+h:c:), no Hund's coupling in Hint, and a dispersion Ek=2Tcosk;T=t11t12 t12t22 (31) We have chosen t12=t21and all the elements real (and positive)to preserve the phsymmetry. Strictly speak- ing, in 1D the mean- eld factorization that led to Eq. (4) and the consequent single-site approximation are both unjusti ed. The choice of dimensionality, here, is only for the ease of discussion and not essential to the con- clusions. As long as the dispersion matrix can be diago- nalized with a momentum-independent unitary transfor- mation (as well as any Bethe lattice, see the appendix D), the following discussion applies. Diagonalizing the tunnelling matrix gives E k=2tcoskwith t=t11+t22 2rt11+t22 22 detT (32) Including renormalization just changes t !~t . We can simply use the diagonalized form of the tunnelling8 matrix to calculate F0atT= 0. Assuming det T>0, Ff=X =Zdk 2E kf(E k) !(~t++~t)Z=2 =2dk cos(k) =2(~t11+~t22)= Note thatt12does not enter the free energy. Inserting this expression into Eq. (27) and setting dF=dzi= 0, we can remove aiin favor of zi. This seems to imply that there is a nite threshold (topological stability) for inter- orbital tunnelling: as long as det T>0, introducing t12 does not change anything in the problem and it simply drops out and OSM phase is stable against inter-orbtital tunnelling. For large t12eventually det T<0. So, we get t+>0 andt<0 and second band is inverted and F0 becomes Ff!2(~t+~t)==4 s ~t11~t22 22 + ~t12 2(33) Hence,t12has non-trivial e ects on renormalization. On the other hand, we could have used the rotational invariance of Hintand done a rotation in d1d2ba- sis to band-diagonalize H0with the bandwidths T! diagft+;tg, before using slave-spins to treat the inter- actions. It is clear then that t12always has non-trivial e ects by modifying t. For example we could start in the locking phase where t=t+>0:2, and by increasing t12slightly get to the OSMT phase t=t+<0:2, without changing the sign of det T. This paradox exist for any phsymmetric lattice with diagonalizable tunnelling matrix. The root of the problem is that our expression in Eq. (9) is not invariant under rotations between vari- ous orbitals. Therefore, the critical value where the OSM phase persists, is basis-dependent. This ambiguity calls for the use of unbiased techniques to understand the role of inter-orbital tunnelling on OSMT. It might be that the model we studied analytically here is a singular limit which can be avoided by breaking phsymmetry and inclusion of crystal eld in more realistic settings [18]. This remains to be explored in a future work. As discussed in [25], the way to achieve rotational- invariance is to liberate the f-electrons that describe quasi-particles from the physical d-electrons. This is achieved by a d !P ^z f representation which leads to a wavefunction-renormalization matrix z =h^z i with o -diagonal elements. So far, we have not been able to generalize the slave-spin to a rotationally invari- ant form and we leave it as a future project. ON-SITE INTER-ORBITAL HYBRIDIZATION Even though models for the Iron-based superconduc- tors have nite crystal level splitting and no on-site hy- bridization, it is interesting to introduce a hybridizationbetween the two orbitals within the current formalism [27]. This is interesting, because the on-site hybridiza- tion, does not su er from the singe-site approximation h^zi ^zi i 6=h^zi ih^zi i, as opposed to the inter-orbital tunnelling andh^zi ^zi iappears as an independent or- der parameter, which leads to the emergence of Kondo screening as we show in this section. We can include a termP n;(v12dy n;1dn;2+h:c:) to the Hamiltonian. In order to preserve the phsymmetry, v12has to be purely imaginary. The modi cations to the mean- eld Hamiltonians are Hf=X n;(~v12fy n;1fn;2+h:c:)2sA12Z12(34) HS=X A12x 1x 2 (35) where ~v12=v12Z12withZ12=hx 1x 2iandA12= v12P nhfy n;1f2i+h:c:.Z12andA12are are related to each other via the Hamiltonian above and they are inde- pendent of in the paramagnetic regime. Alternatively, we can regard them as independent and impose the mean- eld equation Z12=@FS=@A 12to eliminate A12by a Lagrange multiplier. Assuming a small A12we can com- pute the change in slave-spin energy using second-order perturbation theory. The result is of the form  ES= (A12)2where is (in absence of Hund's coupling) a positive constant which contains all the matrix elements and the inverse gaps =P j0h 0jx 1x 2j ji(Ej E0)1h jjx 10x 20j 0iwhereEjandj jiare the eigen- value/states of the HSsolved in the previous section. Eliminating A12in favor ofZ12we nd that the free en- ergy of the system is F(z1;z2;Z12) =2s X knTr log~k1i!niZ0 12 iZ0 12 ~k2i!n +E0 S(z1;z2) +(Z0 12)2 0(36) HereE0 Sis the value of ES(a1;a2)P iaiziin absence of hybridization v12in whicha1anda2are eliminated in favor ofz1andz2. Also, we have rede ned jv12jZ12! Z0 12and jv12j2! 0. Eq. (36) is nothing but the free energy of a Kondo lattice at half- lling [51] with renormalized dispersions ~k1and ~k2. In a Kondo lattice, this form of the free energy appears using Z0 12as the Hubbard-Stratonovitch eld that decouples the Kondo coupling 0~S2dy 1~ d1. Here S2=dy 2~ d2is the spin of the Mott-localized band and 0plays the role of the Kondo coupling. As a result of this coupling, a new energy scale TKDexp[1= 0] ap- pears, with D2~t11the bandwidth of the wider band, below which the Kondo screening takes place which in thephsymmetric case gaps out both bands but away fromphsymmetry mobilizes the Mott localized band. Either way, we conclude that orbital selective Mott in- sulating phase is unstable against hybridization between9 the two orbitals in agreement with [27]. However, even though a true selective Mottness is unstable, orbital dif- ferentiation, re ected as large di erence in e ective mass can exist [16]. CONCLUSION In conclusion, we have used slave-spin mean- eld method to study two-band Hubbard systems in presence of Hund's rule coupling. We have developed a Landau- Ginzburg theory of the locking vs. OSMT. We discussed the relation between slave-spins and the KR boson methods (Appendix). We have also applied the method to impurity problems and shown a correspondence between the latter and the single-site approximation of the lattice using the DMFT loop. Finally, we have discussed the limitations of the slave-spin method for multi-orbital models with both particle-hole symmetry and inter-orbital tunnelling and shown that the orbital selective Mott phase is unstable against on-site hy- bridization between the two orbitals. We appreciate valuable discussions with P. Coleman, T. Ayral, M. Metlitski, L. de'Medici, K. Haule and C.-H. Yee, and in particular, a detailed reading of the manuscript and constructive comments by Q. Si. The authors acknowledge nancial support from NSF-ONR. After completion of this manuscript, we became aware of another work [52] which contains a Landau-Ginzburg theory of OSMT in presence of the inter-orbital tun- nelling. The conclusions of the two work agrees wherever there is an overlap. APPENDIX A. Various slave-particle methods For a one band model, KR introduces four bosons and uses the representation ^ zy =P+[py e+dyp]P, where py ,eyanddyare (hardcore) bosonic creation operators for-spinon, holon and doublon, respectively and Pare projectors that depend on the occupations of the bosons and are introduced to normalize the probability ampli- tudes over the restricted set of physical states. On the other hand, a SU(2) spin-variable ~  can be represented by two Schwinger bosons a andb satisfying the con- straintay a +by b = 1 (hardcore-ness), via z =by b ay a ; x =ay b +by a (37) On an operator level, the two methods have the same Hilbert space as depicted in Table (I) for the case of oneay "a"by "b"ay #a#by #b# 1 0 1 0 0 1 1 0 1 0 0 1 0 1 0 1eyepy "p"py #p#dyd 10 0 0 01 0 0 00 1 0 00 0 1 TABLE I: Comparison of the Schwinger boson representation of the slave-spin (left) and Kotliar-Ruckenstein slave-bosons (right). orbital. Average polarization of the spin along various di- rection in the Bloch sphere corresponds to condensation ofaandbbosons. A trouble with the slave-spin representation is that thef-quasi-particles carry the charge of the d-electron and thus the disordered phase of the slave-spins (in which thef-electrons still disperse beyond single-site approxi- mation) is not a proper description of the Mott phase. As a remedy, it has been suggested [31] to replace xin Eq. (2) with +and xing the problem of non-unity Z in the non-interacting case by applying ne-tuned pro- jectors ^zy=P++P. We note that this looks quite similar to KR. ForMspinful orbitals, KR requires introducing 4M bosons (only one of them occupied at a time) whereas only 2Mslave-spins are required (each with the Hilbert space of 2). Thus the size of the two Hilbert spaces are the same 22M= 4M. B. General low-energy considerations Generally for a lattice we can expand the self-energy Gd(k;!) = [!1Ekd(k;!)]1; (38) Expanding the self-energy d;lat(k;!) =(0;0) +~k@~k(0;0) +!@!(0;0) + Within single-site approximation, the second term is zero. Denoting the third term as @!d1Z1and assuming Z=zzywe can write Gd(k;!)z[!1zyEkz]1zy; (39) which simply means Gf(!) = [!1~Ek]1and the corre- lation functions of the slave-particles are just decoupled h^zi ()^zy j i!z z within single-site approximation also discarding any time dynamics. For the tunnelling ma- trix, we simply have ~t=zytz. Similarly, for an impurity we have i!n1d;imp (i!n) =G1 d;imp(i!n); (40) Gd;loc(i!n) =X k[i!nEkd;lat(k;i!n)]1(41)10 Denoting the interaction part of the self-energy d;imp (i!n) =(i!n) +d;I(i!n), the DMFT approxi- mation identi es d;I(i!n) =d;lat(k;i!n). Again ex- pandingd;I(!)(1Z1)!we have Gd;imp (i!n) =z[i!n1~(i!n)]1zy; (42) with ~(i!n) = zy(i!n)zin agreement with Gf;imp (i!n) = [i!n1~(i!n)]1. Using the same approximation for Gd;locleads to G1 d;loc(i!n)!zX k[i!n~Ek]1z (43) the DMFT self-consistency loop equation is Gf;loc(i!n) = Gf;imp (i!n) or X k[i!n~Ek]1= [i!n1~(i!n)]1: (44) Within the slave-spin approach there are no interactions f;imp =zy(i!n)z;f;I= 0;f;lat= 0 (45) and Eq. (44) is satis ed as it does for any non-interacting problem. Rotation - Using the vector Dfor thed-electrons, in presence of inter-orbital tunnelling we may sometimes be able to eliminate such inter-orbital tunnelling by a rotation to D=UD. SinceD=zF, we assume the same rotation in the F-spaceF=UF(otherwise they would contain inter-orbital tunnelling) and the two z-s are related by z=UyzU. Assuming that Uis a SO(2) matrix, and zis diagonal, we nd z=z++z 21z+z 2 cos 2 sin 2 sin 2 cos 2 ! (46) which has o -diagonal elements. Note that if one of the zelements vanishes, e.g. z= 0, we can factorize z z=z+ cos sin ! cos sin  : (47) Then, it can be seen that Z=zzy!z+zhas the same form. This basically means one linear combination of f electrons is decoupled (localized) and the itinerant spinon band carries characters of both d1andd2bands. This basis-dependence of the orbital Mott selectivity is again an artefact due to lack of rotational invariance. C. Finite- UAnderson model The slave-spin part of the Hamiltonian is as we had in the one band case. We can use Eq. (27) to eliminate a in favour of z. In the wide band limit for the conductionband, we have Gf(i!n) = [i!niKsign(!n)]1where K=t2z2, and the free energy is F(z) =2sZD Dd! f(!)Im [log (iK!)] +E0 S(z):(48) E0 Sis obtained by eliminating afromES(a)2sazpart of the free energy in Eqs. (13) and (27) and is equal to E0 S=U 4p 1z2. Here, we have done a simpli cation to replaceFSwith its zero temperature value (ground state energy) while maintaining the temperature dependence of theFf. We expect this approximation to be valid in the large-Ulimit especially close to the transition. The mean- eld equation w.r.t zis zZD Dd!f(!)Re1 !iK +U 4t2zp 1z2= 0 (49) Close to the transition, the second term is e ectively like az=J withJ(z)(4t2=U)p 1z2. At zero tempera- ture the left-side simpli es zlogK D+z Jp 1z2=zlogK TK(z)= 0; (50) whereTK(z) =De1=J(z). So to have non-zero zwe must have  K=TKwhich determines z. Also, we can go to non-zero temperature. We just replace the log-term in above expression with its nite-temperature expression from Eq. (49) zReh ~ (iK)~ (D)i +z J(z)= 0 (51) This is solved numerically and the result shown in Fig. (7). It shows a Kondo phase z>0 forT <T0 K. FIG. 7: (color online) The order parameter zfor the Anderson model calculated from numerical evaluation of Eq.(53).T0 K=De1=J, whereJ= 4t2=U. Note that this is o with a factor of 4, an artifact of slave-spin method. We have usedD= 100Uwhereast=Uis varied.11 D. Stability of OSMT against interorbital tunnelling in a Bethe lattice Using equations of motion, the coecient J de ned Eq. (4) can be related to the correlation function of the electrons at the same site, The result is J =t X [~t1] Zd! 2f(!)!A ii(!) (52) This together with a = 2P J z leads to Eq. (27). In a Bethe lattice we can use recursive methods [53] to computeA ii. When the tunnelling matrix is hermitian and there is no chemical potential or crystal eld, the procedure is especially simple. We diagonalize the renor- malized tunnelling matrix ~t=U~tDU1. Then the re- tarded and the spectral functions are GR(!) =~t1U(!)U1;A(!) =UAD(!)U1(53) where diagonal matirix contains-elements that sat- isfyi+1 i=!=ti Dwith the retarded boundary condi- tion.ADis diagonal matrix of semicircular density states whose width are given by the eigenvalues of ~t. By plug- ging this into Eq. (53) and (52) and using Zd! 2f(!)!AD ii(!) =0:21222 ~tD we see that if the eigenvalues of the matrix ~tall have the same sign, then U(j~tDj=~tD)U1=~t. This is the generalization of the protection of OSM phase against inter-orbital tunnelling, discussed in the 1D case in the paper. For the case of two bands, det~t>0) J =0:21222t  ; det~t<0) J =0:21222t R (54) i.e. for det ~t>0, theJ-matrix does not have any o - diagonal elements and the diagonal elements are propor- tional to the bare diagonal hoppings (as before), but but if det ~t<0, there is a matrix R=UzU1multiplying element-by-elements of the J-matrix which does depend on renormalization. Again in this problem, one could have done the ro- tation ind -sector before using the slave-spins, in which case, inter-orbital tunnelling would have an e ect and could cause OSM transition. Therefore, the stability found above is basis-dependent. This ambiguity is ab- sent whenphsymmetry is broken and the tunnelling matrix cannot be diagonalized independent of the mo- mentum [18]. [1] Y. Kamihara, T. Watanabe, M. Hirano, H. Hosono, J. Am. Chem. Soc. 130, 3296 (2008).[2] D. C. Johnston, Adv. Phys. 59, 803-1061 (2010). [3] G. R. Stewart, Rev. Mod. Phys. 83, 1589 (2011). [4] F. Wang, D.-H. Lee, Science 332, 200-204 (2011). [5] P. Dai, Rev. Mod. Phys. 87, 855-896 (2015). [6] Q. Si, R. Yu, E. Abrahams, Nature Reviews 1, 1 (2016). [7] P. J. Hirschfeld, Comptes Rendus Physique 17, 197 (2016). [8] M. Daghofer, A. Moreo, J. A. Riera, E. Arrigoni, D. J. Scalapino, E. Dagotto, Phys. Rev. Lett. 101, 237004 (2008). [9] M. J. Calderon, B. Valenzuela, E. Bascones, Phys. Rev. B80, 094531 (2009). [10] K. Haule, J. H. Shim, and G. Kotliar, Phys. Rev. Lett. 100, 226402 (2008). [11] K. Haule, G. Kotliar, New J. Phys. 11, 025021 (2009). [12] Z. P. Yin, K. Haule, G. Kotliar, Nature Materials 10, 932-935 (2011). [13] Z. P. Yin, K. Haule, G. Kotliar, Phys. Rev. B 86, 195141 (2012). [14] C. Aron, G. Kotliar Phys. Rev. B 91, 041110(R) (2015). [15] K. Stadler, Z. P. Yin, J. von Delft, G. Kotliar, A. Weich- selbaum, Phys. Rev. Lett. 115, 136401 (2015) [16] H. Miao, Z. P. Yin, F. Wu, J. M. Li, J. Ma, B.- Q. Lv, X. P. Wang, T. Qian, P. Richard, L.-Y. Xing, X.-C. Wang, C. Q. Jin, K. Haule, G. Kotliar, H. Ding, Phys. Rev. B 94, 201109(R) (2016). [17] Q. Si, E. Abrahams, Phys. Rev. Lett. 101, 076401 (2008). [18] R. Yu, Q. Si, Phys. Rev. Lett. 110, 146402 (2013). [19] L. de'Medici, G. Giovannetti, M. Capone, Phys. Rev. Lett. 112, 177001 (2014). [20] M. Yi, D. H. Lu, R. Yu, S. C. Riggs, J.-H. Chu, B. Lv, Z. K Liu, M. Lu, Y.-T. Cui, M. Hashimoto, S. -K. Mo, Z. Hussain, C. W. Chu, I. R. Fisher, Q. Si, Z.-X. Shen, Phys. Rev. Lett. 110, 067003 (2013). [21] M. Yi, Z-K Liu, Y. Zhang, R. Yu, J.-X. Zhu, J. J. Lee, R. G. Moore, F. T. Schmitt, W. Li, S. C. Riggs, J.- H. Chu, B. Lv, J. Hu, M. Hashimoto, S.-K. Mo, Z. Hus- sain, Z. Q. Mao, C. W. Chu, I. R. Fisher, Q. Si, Z.- X. Shen, D. H. Lu, Nature Commun. 6, 7777 (2015). [22] S. E. Barnes, J. Phys. F 6, 1375 (1976) ibid.7, 2637 (1977). [23] P. Coleman, Phys. Rev. B 29, 3035 (1984). [24] G. Kotliar, A. E. Ruckenstein, Phys. Rev. Lett. 57, 1362 (1986). [25] F. Lechermann, A. Goerges, G. Kotliar, O. Parcollet, Phys. Rev. B 76, 155102 (2007). [26] S. Florens, A. Georges, Phys. Rev. B 70, 035114 (2004). [27] L. de'Medici, A. Georges, S. Biermann, Phys. Rev. B 72, 205124 (2005). [28] S. R. Hassan, L. de' Medici, Phys. Rev. B 81, 035106 (2010). [29] L. de'Medici, M. Capone, arXiv:1607.08468v1 (2016). [30] A. B. Georgescu, S. Ismail-Beigi, Phys. Rev. B 92, 235117 (2015). [31] R. Yu, Q. Si, Phys. Rev. B 86, 085104 (2012). [32] M. Mardani, M.-S. Vaezi, A. Vaezi, arXiv:1111.5980 (2011). [33] A. R uegg, S. D. Huber, M. Sigrist, Phys. Rev. B 81, 155118 (2010). [34] R. Zitko, M. Fabrizio, Phys. Rev. B 91, 245130 (2015). [35] TheZ2mod-2 slave-spin introduces a spin variable that corresponds to charge per site moduli 2 (appli- cable to single-orbital systems), whereas the slave-spin12 method introduces separate slave-spin variables for each spin/orbital. To distinguish the two, we refer to the for- mer as mod-2 slave-spin method. [36] A. Georges, G. Kotliar, W. Krauth, M. J. Rozenberg, Rev. Mod. Phys. 68, 13 (1996). [37] A. Koga, N. Kawakami, T. M. Rice, M. Sigrist, Phys. Rev. B 72, 045128 (2005). [38] Y. Song, L.-J. Zou, Phys. Rev. B 72, 085114 (2005). [39] J. B unemann, D. Rasch, F. Gebhard, J. Phys.: Cond. Matt. 19, 436206 (2007). [40] A. I. Poteryaev, M. Ferrero, A. Georges, O. Parcollet, Phys. Rev. B 78, 045115 (2008). [41] Y. Song, L.-J. Zou, Eur. Phys. J. 72, 59 (2009). [42] K. S. D. Beach, F. F. Assaad, Phys. Rev. B 83, 045103 (2011). [43] E. A. Winograd, L. de'Medici, Phys. Rev. B 89, 085127 (2014). [44] Z.-Y. Song, H. Lee, Y.-Z. Zhang, New J. of Phys. 17, 033034 (2015).[45] S. R. White, Phys. Rev. Lett. 692863 (1992). [46] R. Nandkishore, M. A. Metlitski, T. Senthil, Phys. Rev. B86, 045128 (2012). [47] When dealing with complex z , the rst term of the Hamiltonian (9) and the last term of the free energy (27) have to be modi ed to a z !(h ^z +h:c:) where h =P J z and note that a =h c +h . Since ^zy = ^z here, it is more convenient to work with a . [48] W. F. Brinkman, T. M. Rice, Phys. Rev. B 2, 4302 (1970). [49] A. Georges, L. de' Medici, J. Mravlje, Annu. Rev. Con- dens. Matter, 4, 137 (2013). [50] L. de'Medici, A. Georges, G. Kotliar, S. Biermann, Phys. Rev. Lett. 95066402 (2005). [51] P. Coleman, Introduction to Many-Body Physics , Cam- bridge Univ. Press (2016). [52] R. Yu, Q. Si, private communication. [53] Y. Komijani, I. Aeck, Phys. Rev. B 90, 115107 (2014).
1705.05826v3.Theory_of_electron_spin_resonance_in_one_dimensional_topological_insulators_with_spin_orbit_couplings.pdf
Theory of electron spin resonance in one-dimensional topological insulators with spin-orbit couplings: Detection of edge states Yuan Yao,1,Masahiro Sato,2, 3Tetsuya Nakamura,4Nobuo Furukawa,4and Masaki Oshikawa1 1Institute for Solid State Physics, University of Tokyo, Kashiwa, Chiba 277-8581, Japan 2Department of Physics, Ibaraki University, Mito, Ibaraki 310-8512, Japan 3Spin Quantum Recti cation Project, ERATO, Japan Science and Technology Agency, Sendai 980-8577, Japan 4Department of Physics and Mathematics, Aoyama-Gakuin University, Sagamihara, Kanagawa 229-8558, Japan Edge/surface states often appear in a topologically nontrivial phase when the system has a bound- ary. The edge state of a one-dimensional topological insulator is one of the simplest examples. Electron spin resonance (ESR) is an ideal probe to detect and analyze the edge state for its high sensitivity and precision. We consider ESR of the edge state of a generalized Su-Schrie er-Heeger model with a next-nearest neighbor (NNN) hopping and a staggered spin-orbit coupling. The spin-orbit coupling is generally expected to bring about nontrivial changes on the ESR spectrum. Nevertheless, in the absence of the NNN hoppings, we nd that the ESR spectrum is una ected by the spin-orbit coupling thanks to the chiral symmetry. In the presence of both the NNN hopping and the spin-orbit coupling, on the other hand, the edge ESR spectrum exhibits a nontrivial frequency shift. We derive an explicit analytical formula for the ESR shift in the second-order perturbation theory, which agrees very well with a non-perturbative numerical calculation. I. INTRODUCTION In recent decades, topological phases have become a central issue in condensed matter physics. An important class of topological phases is topological insulators and topological superconductors1{4. In condensed matter and statistical physics, one- dimensional (1-D) systems, which are amenable to several powerful analytical and numerical methods, of- ten provide useful insights. 1-D topological phases are no exceptions. One of the simplest 1-D models possessing nontrivial topological nature is the Su-Schrie er-Heeger (SSH) model5, which has been used to describe the lattice structure of polyacetylene [C 2H2]n. The SSH model can be also applied to the 1-D charge density wave systems, such as quasi-one-dimensional conductors like TTF- TCNQ (tetrathiofulvalinium-tetracyanoquinodime- thanide) and KCP (potassium-tetracyanoplatinate)6. While the SSH model had been studied intensively much earlier than the notion of topological phases was conceived, there is a renewed interest from the viewpoint of topology. In fact, distinct phases of the SSH model are classi ed by the Zak phase7which is a topological invariant, and the bulk winding number of the momentum-space Hamiltonian8. In this sense, the SSH model can be regarded as a 1-D topological insulator. An important nontrivial signature of many topologi- cal phases is edge states. The SSH model indeed pos- sesses zero-energy edge states that are protected by a chiral symmetry8. The number of edge states at a do- main wall is equal to the bulk winding number. This is known as the bulk-boundary correspondence in the spin- less inversion-symmetric SSH model8. Experimentally, 1-D systems with boundaries or edges can be realized by adding impurities to the material so that the system is broken to many nite chains. However, the edge statesare often experimentally dicult to observe, since they are localized near the boundaries or the impurities and their contribution to bulk physical quantities is small. Given this challenge, electron spin resonance (ESR) pro- vides one of the best methods to probe the edge states, thanks to its high sensitivity. In fact, the edge states of theS= 1 Haldane chain were created by doping im- purities and then successfully detected by ESR9,10. Fur- thermore, combined with near-edge x-ray absorption ne- structure experiments, ESR was applied successfully to probe the magnetic edge state in a graphene nanoribbon sample11,12. Such a strategy could also be applied to 1- D topological insulators, which are described by the SSH model. Another intriguing nature of ESR is that it is highly sensitive to magnetic anisotropies, such as the anisotropic exchange interaction, single-spin anisotropy, and the Dzyaloshinskii-Moriya (DM) interaction. The e ect of magnetic anisotropies on ESR is well understood only in limited circumstances, and there remain many open issues13,14. These magnetic anisotropies are often con- sequences of spin-orbit (SO) coupling which generally breaks spin-rotation symmetry. The e ects of magnetic anisotropies and SO couplings also play important roles in magnetic dynamics in higher-dimensional topological phases1{4,15,16. Thus it is of great interest to study the e ect of SO coupling on ESR directly. However, this question has not been explored in much detail so far. An obstacle for the potential experimental ESR study of SO coupling is the electromagnetic screening in metallic systems. This problem does not exist in insulators. Un- fortunately, band insulators are generally non-magnetic and we cannot expect interesting ESR properties. On the other hand, Mott insulators can have interesting magnetic properties. However, strong correlation e ects, which are essential in Mott insulators, make theoretical analysis dicult. In this context, the 1-D topological insulator providesarXiv:1705.05826v3 [cond-mat.str-el] 18 Nov 20172 a unique opportunity to study the e ects of SO coupling on ESR. This would be of signi cant interest in several aspects. Experimentally, the insulating nature makes the observation of edge states by ESR easier. Theoret- ically, the interesting e ects of anisotropic SO coupling on ESR can be studied accurately for the SSH model of non-interacting electrons. Moreover, the chiral sym- metry, which is essential for the well-de ned topological insulator phase, is often broken explicitly in realistic sys- tems. When we introduce a chiral-symmetry breaking perturbation to the 1-D SSH model, the energy eigen- values of the edge states generally deviate from zero en- ergy. However, the edge states are expected to still sur- vive and be localized near the edge if the perturbation is small enough. As we will demonstrate, the ESR of the edge state can detect the breaking of the chiral symme- try. The purpose of this paper is to present a theoretical analysis on ESR of the edge states in 1-D topological insulators, based on a generalized SSH model with SO couplings. We demonstrate several interesting aspects of ESR, which will hopefully stimulate corresponding ex- perimental studies. The paper is organized as follows. In Sec. II, we present the model of interest and review the basic topological na- ture of the SSH model. The next three sections are the main part of this paper. The properties of edge states are discussed in detail in Sec. III. In Sec. IV, we obtain a compact analytical formula of the ESR frequency shift in perturbation theory with respect to SO coupling. Sec- tion V is devoted to a direct numerical calculation of the ESR frequency shift, which is independent of the pertur- bative approach in Sec. IV. We nd that the perturba- tion theory agrees with the numerical results very well. Finally, we present conclusions and future problems in Sec. VI. II. THE MODEL A. A generalized SSH model First let us consider a generalized SSH model with SO coupling H0=+1X j=1 t 1 + (1)j0 cy j+1exp i(1)j 2~ n~  cj +h.c.g (1) wherecjis the two-component electron annihilation op- eratorcj[cj";cj#]Tat thej-th site,t > 0 is the nearest-neighbor (NN) electron hopping amplitude, and 101 is the bond-alternation parameter. The angleand axis~ n(which is a unit vector) parametrizes the SO coupling on the NN bond. The angle denotes the ratio of the SO coupling to the hopping amplitude on the bond. In this paper, we assume that is suciently small (jj1), which is the case in many real materials.Expanding to rst order, we obtain a standard form with so-called intrinsic and Rashba SO couplings17. In our model Eq. (1), SO coupling is assumed to be staggered along the chain. This would be required, in the limit of 0= 0, if the system had site-centered inver- sion symmetry. In general, other forms of SO coupling including the uniform one along the chain are also pos- sible. In this paper, however, we focus on the particular case of the staggered SO coupling to demonstrate its in- teresting e ects on the ESR spectrum. B. The SSH model and its topological properties In the limit = 0, our model is reduced to the standard SSH model HSSH=Xn t 1 + (1)j0 cy j+1cj+ h.c.o :(2) Let us rst consider a system of 2 Nsites (Nunit cells) with the periodic boundary condition (PBC). It is then natural to take the momentum representation c2j;=1p NX kak;exp (ik(2j)); (3) c2j+1;=1p NX kbk;exp (ik(2j+ 1)); (4) where the summation of kis in the reduced Brillouin zone [0;) withk=n=N andn= 0;;N1. Correspond- ing to the each sublattice (even and odd), there are two avors of fermions, aandb. The Hamiltonian can then be written as HSSH=X k(ay k;by k)hSSH(k) ak bk ; (5) with hSSH(k)dx(k)x+dy(k)y; (6) wherex;y;zare Pauli matrices acting on the avor space, anddx;y(k) are real numbers dx(k) =2tcos(k);dy(k) = 2tsin(k)0: (7) The spin indices are again suppressed in the Hamiltonian. From this expression, the single-electron energy reads (k) =q dx(k)2+dy(k)2=2tq cos2k+02sin2k: (8) The gap is closed at k==2 when0= 0, while the sys- tem has a gap 4 tj0jwhenever the bond alternation does not vanish ( 06= 0). The gapless point can be regarded as a quantum critical point separating the two gapped phases,0<0 and0>0. Throughout this paper, we consider the half- lled case with 2 Nelectrons. The bulk mode near this gap-closing point has a linear dispersion relation indicated in Fig. 1, and it can be described by a one-dimensional Dirac fermion3.3 FIG. 1. Band structure of the SSH model in Eq. (2) with a periodic boundary condition.1The solid and dashed lines respectively represent band structures of the insulating case with06= 0 and the gapless point at 0= 0. The low-energy physics around k==2 can be described by one-dimensional Dirac fermion model. It is evident from the Hamiltonian that each of the gapped phases is simply a dimerized phase. Neverthe- less, we can identify them as a trivial insulator phase and a \topological insulator" phase. This can be under- stood by considering the system with the open boundary condition. Let us consider the chain of 2 Nsites labeled withj= 1;2;:::; 2N, and the open ends at sites j= 1 and 2N. For0>0 (0<0), sitesj= 2nandj= 2n+1 (j= 2n1 andj= 2n) form a dimerized pair, re- spectively. As a consequence, for 0>0 the end sites j= 1 andj= 2Nremain unpaired. The electrons at these unpaired sites give rise to S= 1=2 edge states. In contrast, for 0<0, there are no unpaired sites and thus no edge states. In this sense, 0>0 is a topo- logical insulator phase and 0<0 is a trivial insulator phase. Of course, considering the equivalence of the two phases in the bulk, such a distinction involves an arbi- trariness. That is, if we consider the an open chain of Nsites withj= 0;1;:::; 2N1, the edge states appear only for0<0. It is still useful to identify the gapless point at0= 0 as a quantum critical point separating the topological insulator and the trivial insulator phase. The particular shape of the Hamiltonian also implies the existence of a chiral symmetry: fhSSH;g= 0; (9) where z. The chiral symmetry turns out to be im- portant for the distinction of the two phases. In the con- text of the general classi cation of topological insulators, the present system corresponds to the \AIII" class with particle number conservation and the chiral symmetry in one spatial dimension18,19. In a general one dimensional free fermion system, we can de ne a topological invariant called the Zak phase7 for each band as follows: Zak=iI BZh (k)jOkj (k)i; (10) FIG. 2. Amplitude j jof the edge-state wave function of the SSH model in Eq. (2) under an open boundary condition.8 Blue and red colors respectively represent the spatial distri- bution of the existing probability for left and right localized edge states. The total site number is set to be even, the dimerization parameter 0>0, and symbols vandwdenote hopping amplitudes t(10) andt(1 +0), respectively. The wave-function amplitude decays exponentially into the bulk4. wherej (k)iis the Bloch wavefunction of the band with the momentum k. In the presence of the chiral symmetry, Zakis quantized to integral multiples of , if the band is separated from others by gaps20. For the present two-band SSH model in Eq. (5) with the chiral symmetry, we can compute the Zak phase using the explicit Bloch wavefunction. For the lower band, j (k)i=1p 2 exp (ik) 1 ; (11) withkarctan[dy(k)=dx(k)]. As a result, we nd Zak== 1 for 0 <  01 in which the system is a topological insulator. In the other case 10<0, where the system is a trivial insulator, Zak== 0. In this case, we can see that the Zak phase can be also iden- tifed20with a winding number of the Hamiltonian as Zak =i I BZdkOkln [dx(k)idy(k)]: (12) When Zak== 1, there is an edge state localized at each end of an open nite chain as shown in Fig. 2. This is the bulk-boundary correspondence21in the SSH model. In general, this topological number can take arbitrary integer values, corresponding to Zclassi cation of BDI or AIII class in d= 1 dimension. However, in the present SSH model, its value is restricted to 0 or 1. The existence of the edge states in the SSH model can be demonstrated by an explicit calculation for a nite- size chain. In Fig. 3, we can see that, when 0decreases from 1 to -1, the edge states merge into the bulk spectrum as0!0+. In addition, when the thermodynamic limit, N!+1, is taken in the open-end SSH model, the edge states are strictly at zero energy and topologically stable against any local adiabatic deformation that respects the chiral symmetry8.4 −1 −0.5 0 0.5 1−2−1012 δ0/tEnergy spectrum EEnergy spectrum with di fferentδ0 FIG. 3.0dependence of the energy spectrum of the SSH model with t= 1 andN= 40 under an open boundary con- dition. As 0!0+, two localized edge states at the ends merge into the bulk. For the limit N!+1, the edge states are strictly at zero energy as 0 < twhich are protected by the chiral symmetry.8 C. ESR of edge states Let us consider ESR of the 1-d half- lled topological in- sulator phases at the low-temperature and low-frequency limitj!j;Tj0jton which we focus in this paper. The ESR contribution from bulk excitations is negligible in this limit since there is a large bond-alternation driven band gap 4j0jt. On the other hand, spin-1/2 edge states are located at the (nearly) zero energy point in the band- gap regime. Therefore, ESR is dominated by the edge state contribution. When the chiral symmetry is preserved and SO cou- pling is absent, the edge spin is precisely equivalent to a freeS= 1=2. In this case, the edge ESR spectrum is triv- ial, which means that it just consists of the delta function at the Zeeman energy. However, breaking of the chiral symmetry and introduction of SO couplings can bring a nontrivial change on the edge ESR spectrum. In the following, we shall analyze this e ect theoretically. In ESR, absorption of an incoming electromagnetic wave is measured under a static magnetic eld. Thus we introduce the Zeeman term for the static, uniform magnetic eld HZ=H 2X j=1cy j(~ ~ nH)cj; (13) where~ nHis a unit vector representing the direction of the magnetic eld, H > 0 is its magnitude, and ~ = (x;y;z). In the paramagnetic resonance of independent electron spins, absorption occurs for the oscillating magnetic eld perpendicular to the static magnetic eld, which is mea-sured in the standard Faraday con guration. Therefore, in this paper, we assume that the oscillating magnetic eld is perpendicular to the static magnetic eld ~ nH. The frequency of the electromagnetic wave is denoted by !. In an electron system with the SO interaction, the electric current operator contains a \SO current" that involves the spin operator. Since the electric current couples to the oscillating electric eld, in the actual setting of the ESR experiment, the optical conductiv- ity due to the SO current also contributes to the ab- sorption of the electromagnetic wave with a spin ip. This e ect is called Electron Dipole Spin Resonance (EDSR)22{25. The EDSR contribution is generically larger than ESR if SO coupling is at the same order as the Zeeman splitting26,27, as their relative contribu- tions are of ( a=C)2106, wherea1010m is the lattice spacing and C1013m the Compton length of the electron25. In general, EDSR requires a sepa- rate consideration from ESR as they involve di erent operators25. Nevertheless, in the low-temperature/low- frequency regime, only the two spin states of the edge state are involved. Thus, although EDSR contributes to the absorption intensity di erently from ESR, the reso- nance frequency is identical between the ESR and EDSR. With this in mind, we do not consider EDSR explicitly in the rest of the paper. It should be noted that, for a higher temperature or a higher frequency, the EDSR con- tribution to the absorption spectrum is rather di erent from the ESR one, as the absorption spectrum involves bulk excitations. We now consider the ESR in the system with the Hamiltonian H0+Hz. Within the linear response the- ory28, the ESR spectrum is generally given by the dynam- ical susceptibility function in the limit of zero-momentum transfer 00 +(q= 0;!> 0) =ImGR +(q= 0;!);(14) where GR +(0;!) =iZ1 0dtX r;r0h[s+(r;t);s(r0;0)]iei!t =iZ+1 0dth[S+(t);S(0)]iei!t; (15) whereSmeans the ladder operator de ned with respect to the direction ~ nHof the static eld, hi denotes the quantum and ensemble average at the given temperature T, and S(t)X rexp(iHt)s(r) exp(iHt) =X rs(r;t); (16) whereHis the static Hamiltonian we consider (e.g., H=H0+Hz). In the absence of SO coupling ( = 0),H0has the exact SU(2) spin rotation symmetry which is broken only \weakly"14by the Zeeman term HZ. As a gen- eral principle of ESR, in this case, the ESR spectrum (if5 any) remains paramagnetic, namely a single -function at!=H. As we will demonstrate later in Sec. III, in the low-temperature limit, this paramagnetic ESR can be attributed to the edge states of the SSH model. Once the SO coupling is introduced ( 6= 0), the SU(2) sym- metry is broken and we would expect a nontrivial ESR lineshape. However, somewhat surprisingly, (as we will show later), the ESR spectrum attributed to the edge states remains a -function at !=Heven when 6= 0. Thus, in order to investigate possible nontrivial e ects of the SO coupling on ESR, we further consider the NNN hoppings H=X j=1tcy j+2exp[i ~ n =2]cj+ h.c.; (17) where tis the NNN electron hopping amplitude. The angle and~ n are the SO turn angle and the axis for the NNN hopping, respectively. The Hamiltonian of the system to be considered is then HESR=H0+HZ+ H: (18) Once we include the NNN hoppings, the chiral symme- try is broken and the edge states are not protected to be at zero energy. Nevertheless, when the chiral symmetry breaking perturbations are weak, we may still identify \edge states" localized near the ends although they are no longer at exact zero energy even when H= 0. Un- der a magnetic eld H, contributions from these edge states dominate the ESR in the low-energy limit. Now, the ESR spectrum can be nontrivially modi ed by the SO couplings and . It is the main purpose of the present paper to elucidate this e ect. In real materials, the NNN hoppings might be small but they are generally non-vanishing. Thus it is important to develop a theory of ESR in the presence of the NNN hoppings, especially because we can detect the NNN hoppings with ESR even when they are small. We will treat the NNN hopping  H, which would be smaller than the NN hopping H0in many experimental realizations, as a perturbation. This also turns out to be convenient for our theoretical analysis. We also as- sume thatjj;j j 1 since SO couplings are weak in most of the realistic systems, and they will formulate a perturbation expansion in  t,, and . III. EDGE STATES OF H0ANDU(1)~S~ nHSYMMETRY As we discussed earlier, ESR would be an ideal probe to detect the edge state of the 1D topological insulator and various perturbations. In this section, we discuss and explicitly solve the edge states of the unperturbed Hamiltonian H0to demonstrate the robustness of the edge states against SO coupling. As a consequence, in the modelH0only with NN hoppings, there is no nontrivial change in the edge ESR spectrum.Since our Hamiltonian H0is bilinear in fermion op- erators, we can focus on single-electron states and rep- resent them with the ket notation. For a half-in nite chain with sites j= 1;2;:::, wherej= 1 corresponds to the end of the chain, we nd a single-electron eigenstate jEdge;ilocalized near the edge in the \topological in- sulator" phase 0>0. Here=1 represents the spin component in the direction of the magnetic eld. Namely, (H0+HZ)jEdge,i=E(0) jEdge,i (19) ~S~ nHjEdge,i= 2jEdge,i; (20) with the energy eigenvalues E(0) =H=2 and~Sis the total spin of the system. The wave function of the edge states is exactly calculated as hj;jEdge,0i=8 < : 0p N 10 1+0(j1)=2 (j22N0+ 1 ); 0 (otherwise) ; (21) where 2 N0is the set of non-negative even integers, and Nis the normalization constant. The energy eigenvalue of the edge state for H0is, independently of the spin component, exactly zero, re ecting its topological nature. It is also remarkable that the edge state wavefunction is independent of and~ n. This is a consequence of a canonical \gauge transformation"29 ~c2k+1=c2k+1; ~c2k= exp (i~ n~ =2)c2k;(22) which eliminates the SO coupling from H0. In this sense, H0still has a hidden SU(2) symmetry30,31even though the SO coupling breaks the apparent spin SU(2) sym- metry. However, since the gauge transformation involves the local rotation of spins, the uniform magnetic eld HZgives rise to a staggered eld after the gauge trans- formation. This staggered eld completely breaks the SU(2) symmetry. This is similar to the situation in a spin chain with a staggered Dzyaloshinskii-Moriya inter- action13. Thus, following the general principle of ESR, we would expect a nontrivial ESR spectrum in the pres- ence of the staggered SO coupling as in H0. Nevertheless, somewhat unexpectedly, we nd that the edge ESR spectrum for the model H0remains the - function(!H), as if there is no anisotropy at all. This is due to the fact that the edge-state wavefunction Eq. (21) is non-vanishing only on the even sites. Since the gauge transformation can be de ned so that it only acts on the even sites where the edge-state wavefunction Eq. (21) vanishes, the edge state is completely insensi- tive to the SO coupling. Therefore, the spectral shape of the edge ESR remains unchanged by the staggered SO coupling. In addition, since the edge wave functions are eigenstates of the total spin component along ~ nHaccord- ing to Eq. (20), the edge states have U(1)~S~ nHsymmetry generated by ~S~ nH. In fact, the robustness of the edge ESR spectrum is valid for a wider class of models. The edge ESR only6 probes a transition between two states with opposite po- larization of the spin, which form a Kramers pair in the absence of the magnetic eld. Thus, at zero magnetic eld, the time-reversal invariance of the model requires these two states to be exactly degenerate. For a nite magnetic eld, if the system still has U(1)~S~ nHsymme- try of rotation about the magnetic eld axis, the two states can be labelled by the eigenvalues of Sz=1=2, and their energy splitting is exactly !ESR=H: (23) Thus, as far as the edge ESR involving only the Kramers pair is concerned, the U(1)~S~ nHsymmetry is sucient to protect the -function peak at !=H. We note that, more generally, when more than two states contribute to ESR, these symmetries are not sucient to protect the single-peak ESR spectrum, as there can be transitions between states not related by time reversal. In fact, this would be the case for the absorption due to bulk excita- tions which we do not discuss in this paper. IV. PERTURBATION THEORY OF THE EDGE ESR FREQUENCY SHIFT As we have shown in the previous Section, even in the presence of SO coupling, there is no frequency shift for edge ESR in the NN hopping model H0. This is a con- sequence of the chiral symmetry, which stems from the bipartite nature of the NN hopping model. As we will discuss below, the introduction of NNN hop- pings breaks the chiral symmetry and generally causes a nontrivial frequency shift of the edge ESR. In this Sec- tion, we develop a perturbation theory of ESR for the edge states, rst by regarding H0+HZas the unper- turbed Hamiltonian and  Has a perturbation. In the presence of the perturbation with SO couplings, the eigenstate of the Hamiltonian is generally no longer an eigenstate of the spin component ~S~ nHas in Eq. (20). Nevertheless, as long as the perturbation theory is valid, two edge states can still be identi ed by using spin = + and. Namely, for the full Hamiltonian (18), we can de ne the edge state labeled by =as the state adiabatically connected to jEdge;ias H!0. Here let us introduce new symbols E+()andE(n) +()as the energy eigenvalue of the almost spin up (down) edge state in Eq. (18) and its n-th order correction in the perturbation theory, respectively. With these symbols, the ESR frequency is given by !ESR=EE+=H+ !; (24) where the ESR trivial peak position is given by E(0) E(0) +=H. The ESR frequency shift  !, driven by the perturbation  H, is expanded in the perturbation the- ory as != !(1)+ !(2)+:::; (25)where then-th order term is !(n)=E(n) E(n) +; (26) forn2N. In the following parts, we perturbatively solve the single-electron problem to compute the eigen-energy di erence of two edge states E+andE. A. First order in t The NNN hopping  Hbreaks the chiral symmetry, and thus it can change the edge ESR spectrum. In fact, the energy of the edge states is already shifted in the rst order of  H. However, the energy shift is the same for the two edge states with di erent spin polarizations. This is a consequence of the time-reversal (TR) symmetry of the SO coupling, as demonstrated below: E(1) =hEdge,jHjEdge,i =h (Edge,)jH1j (Edge,)i =E(1)  =2t10 1 +0cos 2 (27) where  is the TR operator and  H1= His used. Therefore, the edge ESR spectrum remains un- changed in the rst order of  tas the frequency shift vanishes in this order: !(1)=E(1) E(1) += 0: (28) B. Second order in t We can formally write down the second-order pertur- bation correction E(2) =hEdge,jH(E(0) H0)1PHjEdge,i where the projection operator P 1 jEdge,ihEdge,j. It is rather dicult to evaluate this formula directly, since the intermediate states in the perturbation term include the bulk eigenstates of H0in which the hidden symmetry is generally broken. Assum- ing thatjj;j j1 (i.e., the SO coupling is suciently small), we can develop a perturbative expansion in  and in addition to  t. In this framework, we expand Eas a Taylor series of , and t. The quantity of interest is the energy splitting EE+, since it corresponds to the ESR frequency. The edge state energy splitting in the second order in t, and up to the second order in , is required to take the form E(2) E(2) 1 2Ht2 a(~ n ~ nH)2 2+b(~ n~ nH)22 +c(~ n~ nH)(~ n ~ nH) g; (29) based on the following symmetry considerations, and a, bandcare constants to be determined.7 First of all, ( EE) will not change under (H;~ nH;)!(H;~ nH;); (30) because this corresponds to a trivial rede nition of co- ordinate system. Therefore, we attach the factor Hin the r.h.s. of Eq. (29). Next we notice that always appears with ~ n, and with~ n , which leads to further constraints as we will see below. Let us consider the limiting case = 0 with nonzero . The splitting can only depend on the relative angle between ~ n and~ nH. Furthermore, if ~ n k~ nH, the hidden symmetry of the edge state implies that the energy splitting is exactly given by the Zeeman energy and there is no perturbative correction. Thus, for = 0, the energy splitting can only depend on ( ~ n ~ nH)2 2up toO( 2) since the energy split is a scalar and it must be written in terms of inner and vector products of ~ nH, ~ nand~ n . Then, similarly, if = 0 with nonzero , the energy splitting can only depend on ( ~ n~ nH)22. Finally, the O( ) term should be linear in ~ nand~ n , and it vanishes when ~ nk~ n k~ nHbecause of the U(1)~S~ n symmetry. These requirements uniquely determine the form of (~ n~ nH)(~ n ~ nH). Thus, the symmetries reduce the possible forms of the second-order corrections to Eq. (29) with only the three parameters a,b, andc. To obtain these parameters, we note that the expan- sion inand introduced above can be naturally done by regarding the SSH model without SO coupling ~H0=+1X j=1n t 1 + (1)j0 cy j+1cj+ h.c.o HX j=1cy j(~ ~ nH)cj=2 (31) as the unperturbed Hamiltonian, and Hpert= H0+ H; (32) as the perturbation, where H0+1X j=1n t 1 + (1)j0 cy j+1  exp (1)ji~ n~ =2 1 cj+ h.c. (33) and Has de ned in Eq. (17). As the result of the perturbation calculation given in the Appendix, we nd that the second-order term of fre- quency shift is non-positive given in the form !(2)=H 2X m3;j3(~M+~N)y(~M+~N)0;(34) where ~MX m2;j2hm3;j3jH0 0jm2;j2ihm2;j2jH00jEdgei m2Ej2Ej3 (~ n~ nH) ~Nhm3;j3jH0jEdgei Ej3(~ n ~ nH) : (35)Herejm;ji's are single-particle (bulk) energy eigenstates of~H0in Eq. (31) where m=labels the positive or neg- ative energy sector (i.e., band indices) and jlabels other possible quantum numbers, which is not the wave vector since we have the open-ended boundary condition. The energyEjstands for the energy eigenvalue of j;ji. The perturbation terms  H0, H00, and  H0 0are de- ned as H0X j=1tcy j+2cjh.c.; (36) H00X j=1tcy j+2cj+ h.c.; (37) H0 0+1X j=1n t 1 + (1)j0 cy j+1cjh.c.o :(38) We note that Eqs. (36) and (38) are anti-Hermitian. After putting the de nitions of ~Mand~Ninto Eq. (34), we nd the result consistent with the general form Eq. (29) required by symmetries. The parameters are then identi ed as a=b=t t(1 +0)2 ; c =2t t(1 +0)2 : The detailed derivation of a,bandcis given in Appendix. The nal result can be given in a compact form as !(2)=H 2 t t(1 +0)(~ n~ nH ~ n ~ nH) 2 ;(39) which is non-positive. Since the rst-order correction vanishes as we have already seen, the second-order term Eq. (39) gives the leading term for the frequency shift of the edge ESR. It also implies that, although is the SO-coupling turn angle for the NNN hopping terms, it is equally important in  !as the NN SO coupling turn angleeven if the NNN hopping itself is small ( tt). One of the most remarkable features of our result is that, the shift (up to the second order in the perturba- tion) vanishes when ~ n~ nH= ~ n ~ nH. This corre- sponds to the zeros of curves in Fig. 4 where the direction of the magnetic eld is in the plane spanned by ~ n= ^y and~ n =^z, andis the angle between ~ nHand ^yin the ^y-^zplane, as shown in Fig. 5. Therefore, we predict two directions of the magnetic eld for which the ESR shift vanishes, when the magnetic eld direction ~ nHsweeps the plane spanned by ~ nand~ n . V. NON-PERTURBATIVE CALCULATION OF THE EDGE ESR SPECTRUM Here, in order to see the validity of our perturba- tion theory in the preceding section, let us re-compute the edge ESR frequency with a more direct numerical method. As already mentioned, when the magnetic eld8 FIG. 4. Angle dependence of the ESR frequency shift  !. We set the parameters t= 1:0, t=0= 0:2, andH= 0:05. Red and blue curves are obtained by the second-order perturbation calculation in Sec. IV B. Square and circle points are the results of direct numerical diagonalization for a system (of 100 sites) with an open boundary condition. The de nition of angleis indicated in Fig. 5. The zeros of  !occur at ~ n~ nH= ~ n ~ nH. H(and thus the ESR frequency !) is much smaller com- pared to the bulk excitation gap 4 tj0j, we may ignore the e ects of the bulk excitations in the ESR spectrum. Then the edge ESR spectrum is of the -function form 00 +edge(q= 0;!)/[!(EE+)]; (40) whereE+()is the energy eigenvalue of the \almost" spin-up (spin-down) edge state. These energies can be accurately computed by numerical diagonalization of a - nite (but long) size full Hamiltonian with an open bound- ary condition. Then we obtain the spectrum peak shift !EE+Hfrom the numerical results of E. In the present work, we calculated Eusing a nite open chain of 100 sites. In Fig. 4, we compare the numerical results of the peak shift with the analytical perturbation theory of  !(2) ESRin Eq. (39) when the magnetic eld is in the plane spanned by~ nand~ n . The gure clearly shows that our perturba- tion theory agrees with the numerical results quite well. VI. CONCLUSIONS AND DISCUSSION We have analyzed ESR of edge states in a general- ized SSH model with staggered SO couplings and with an open end. In this paper, we assume that the en- ergy scales of the magnetic eld, the frequency, and the temperature are suciently small compared to the bulk gap. Then the ESR spectrum only consists of a single -function spectrum corresponding to the transition be- tween two spin states at the edge, but is expected to show a nontrivial frequency shift in general as the SO coupling FIG. 5. Geometric relation among some vectors in Fig. 4, where the direction of magnetic eld is in the plane spanned by~ n= ^yand~ n =^z. The parameter is de ned as the angle between ~ nHand ^yin the ^y-^zplane. breaks the SU(2) symmetry strongly under the applied magnetic eld. Nevertheless, there is no ESR frequency shift in the model with only NN hoppings, thanks to its chiral symmetry. The chiral symmetry is broken by NNN hoppings, which should be generally present in any realistic materi- als even if they are small. This NNN hoppings, together with the SO coupling, can induce a nontrivial frequency shift on the edge ESR. Thus we have developed a pertur- bation theory of the frequency shift, regarding the NNN hoppings and the SO couplings as perturbations. Our main result, the ESR frequency shift up to second order in the perturbation theory, is found in Eq. (29). It is non-positive in this order. (The resonance eld shift for a xed frequency, which is usually measured in experi- ments, is always positive.) In the presence of the NNN hoppings, the SO couplings in the NN hoppings, which did not cause a frequency shift by themselves, also con- tribute to the frequency shift. We nd an interesting dependence of the ESR frequency shift on the direction of the static magnetic eld, relative to the SO couplings on the NN and the NNN hoppings. In particular, the ESR frequency shift is predicted to vanish when the static magnetic eld points to a certain direction on the plane spanned by the two SO coupling axes (see Fig. 4). Fur- thermore, we performed a direct estimate of the ESR frequency shift by a numerical calculation of the edge state spectrum, without relying on the perturbation the- ory. The result agrees very well with the perturbation theory, establishing its validity. Our results indicate that, the chiral symmetry break- ing by the NNN hoppings in the \SSH"-type topological insulators in one dimension may be detected by ESR, in9 the presence of the SO couplings. If the NNN hoppings are small, which would be the case in many realistic ma- terials, it might be dicult to detect their e ects with other experimental techniques. ESR has been success- ful in detecting even very small magnetic anisotropies, thanks to its high sensitivity and accuracy. We hope that the present work will pave the way for a new application of ESR in detecting (small) chiral symmetry breaking. While we do not discuss any particular material in this paper, let us discuss here the prospect of experimentally observing the e ects we predict. The maximal frequency shift is given by the order of Ht tmaxf; g2 : (41) The ratio of NNN to NN hoppings,  t=t, of course strongly depends on each material. It is even possible thatjtj=t1, in which case the system may be re- garded as two chains coupled weakly by zigzag hopping. It should be however noted that our theory is valid only whenjtj=tis suciently small. We still expect that our theory works reasonably well for ( t=t)20:1. In carbon-based systems, such as polyacetylene, the SO in- teraction is known to be weak. For example, even with the enhanced SO interaction due to a curvature32,; is of order of 105. This would give an ESR shift that is too small to be observed in experiments. However, the SO interaction is stronger in heavier atoms. In fact, even in carbon-based systems, the SO interaction can be signi cantly enhanced by heavy adatoms. For example, placing Pb as adatoms can enhance ; up to 0:1 or more in graphene33. This would give the edge ESR shift cor- responding to the g-shift up to the order of 103(1,000 ppm), which should be observable. In particular, even if the absolute value of the shift is dicult to be deter- mined, the angular dependence of the ESR shift would be more evident in experiments. Furthermore, it is known that an SO interaction generally becomes larger when the electron system we consider is located in the vicinity of an interface between two bulk systems or is under a strong, static electric eld34{36. Therefore, if we set up an SSH chain system under such an environment, it would become easier to detect an ESR frequency shift due to a strong SO coupling. Throughout this paper, we have taken the NN SO cou- pling in the model Eq. (1) to be staggered. However, there are other possibilities. In particular, in a transla- tionally symmetric system, the NN SO coupling is uni- form. Although the only di erence is the signs in the Hamiltonian, ESR spectra should be signi cantly di er- ent between these two cases. This is clear if we consider the limit of the zero NNN hopping. In the staggered SO coupling case, there is no ESR frequency shift. This is because the NN SO coupling can be gauged out by the canonical transformation Eq. (22), without changing the odd site amplitude and thus leaving the SSH edge state wavefunction Eq. (21) unchanged. On the other hand, in the uniform SO coupling case, gauging out the NNSO coupling a ects any wavefunction including the SSH edge state wavefunction, resulting in the change of the ESR spectrum. A similar di erence has been recognized between the ESR spectum in the presence of a staggered DM interaction13and that with a uniform DM interac- tion along the chain29. The analysis of the edge ESR spectrum in the presence of a uniform SO coupling is left for future studies. ACKNOWLEDGMENTS M. S. was supported by Grant-in-Aid for Scienti c Research on Innovative Area, \Nano Spin Conversion Science" (Grant No.17H05174), and JSPS KAKENHI Grants (No. 17K05513 and No. 15H02117), and M. O. by KAKENHI Grant No. 15H02113. This work was also supported in part by U.S. National Science Foundation under Grant No. NSF PHY-1125915 through Kavli In- stitute for Theoretical Physics, UC Santa Barbara where a part of this work was performed by Y. Y. and M. O. Appendix: Perturbation theory for the frequency shift ! Here we explain how to determine the parameters a,b andcin the general form of Eq. (29), based on a pertur- bation theory. As we have discussed earlier, in order to expand the edge-state energy eigenstates with respect to the SO coupling parameters and , which are assumed to be small, we regard ~H0of Eq. (31) as the unperturbed part, and ~Hpertof Eq. (32) as the perturbation. More- over, we have already shown in Sections III and IV A that the ESR shift vanishes exactly up to O(t). Therefore, the leading terms with coecients a,bandcstem from the second-order perturbation proportional to  t2. Be- low, we will determine three parameters a,bandcfrom the perturbative expansion of ~Hpert. 1. Second order in Hpert The shift of the edge energy eigenvalue in the second order of Hpertis given by E[2] =X m=;j;~=hEdge,jHpertjm;j; ~ihm;j; ~jHpertjEdge,i H 2Em;j;~ X m=;jhEdgej(H0)yjm;jihm;jjH0jEdgei E2 jH jn nHj2 2=4; (A.1) where the bulk single-particle eigen-energy Em;j;~is given by Em;j;~mEj~H=2 (A.2)10 withmEjthe energy eigenvalue of spinless single-particle eigenstatesjm;ji, and we have used the facts that in the unperturbed sector, the orbital and spin parts of single- particle eigenstates can be decomposed, e.g. jm;j;i= jm;jijisince ~H0commutes with the spin operator of each site. Here we de ne E[n] as the energy eigenvalue of the edge state with spin in then-th order of pertur- bation in Hpert. This is to be distinguished from E(n)  introduced in Eq. (26), where nrefers to the order in  t. In the second line of Eq. (A.1), we assume jEm;j;~jH and perform the Taylor expansion of E[2] with respect toH. The matrixhm;jjH0jEdgeican be calculated by using the anti-Hermitian operator  H0of Eq. (36). From Eq. (A.1), the ESR frequency shift driven by E(2) is expressed as E[2] E[2] +=jn nHj2 2=2 X m=;jhEdgej(H0)yjm;jihm;jjH0jEdgei E2 jH:(A.3) This indeed corresponds to the parameter a. To obtain bandc, we should proceed to higher orders of Hpert. 2. Third order in Hpert In the third order perturbation theory within  t2, the energy correction takes the form as E[3] =X m;j;~hEdge,jHpertjm3;j3;~3ihm3;j3;~3jHpertjm2;j2;~2ihm2;j2;~2jHpertjEdge,i (~2)H 2m2Ej2 (~3)H 2m3Ej3 +hEdge,jHpertjEdge,ijhEdge,jHpertjm3;j3;~3ij2  (~3)H 2m3Ej32 X m;jRe" hEdgejH00jm3;j3ihm3;j3jH0 0jm2;j2ihm2;j2jH0jEdgei m3E2 j2Ej3H# (~ n~ nH)(~ n ~ nH) 2;(A.4) where  H00is de ned by Eq. (37). Therefore, E[3] E[3] + =X m;jRe" hEdgejH00jm3;j3ihm3;j3j(H0 0)yjm2;j2i m3E2 j2Ej3 hm2;j2jH0jEdgeiH] (~ n~ nH)(~ n ~ nH) : (A.5) We see that this correction term corresponds to the pa- rametercin Eq. (29). 3. Fourth order in Hpert In order to derive the leading term of the parameter b, we have to calculate the fourth-order term. For conve-nience of the fourth-order calculation, we introduce the abbreviated notation of the matrix elements: Aqrhmr;jr;~rjAjmq;jq;~qi; (A.6) EEqE[0] Emq;jq;~q (A.7) for any operator A, and denote the zeroth order edge state by \E". In this notation, the fourth order edge- energy correction up till ( t)2-order is given by E[4] =X m;j;~HE4H4;3 pertH3;2 pertH2E EE2EE3EE4E[2]  HE42 (EE4)22HEEHE4H43 pertH3E E2 E3EE4+ HEE2 HE42 (EE3)3 X m;j;hEdgejH00jm2;j2ihm2;j2j(H0 0)yjm3;j3ihm3;j3jH0 0jm4;j4ihm4;j4jH00jEdgei m2m4Ej2E2 j3Ej4Hj~ n~ nHj22 4:(A.8)11 Then we can arrive at E[4] E[4] + X m;j;hEdgejH00jm2;j2ihm2;j2j(H0 0)yjm3;j3ihm3;j3jH0 0jm4;j4ihm4;j4jH00jEdgei m2m4Ej2E2 j3Ej4Hj~ n~ nHj22 2:(A.9) This correspond to the term coeciented by the param- eterbin Eq. (29).4. Summation over the perturbation orders Through some algebra, we nd that terms of the fth and higher orders in Hpertdo not contribute to the order ofO t22 . Summing up Eqs. (28), (A.3), (A.5), and (A.9), we arrive at the energy shift of the energy di er- ence between down- and up-spin edge states in an elegant form as in Eq. (34). smartyao@issp.u-tokyo.ac.jp 1M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010). 2X.-L. Qi and S.-C. Zhang, Rev. Mod. Phys. 83, 1057 (2011). 3B. A. Bernevig and T. L. . X. Hughes, Topological insula- tors and topological superconductors (Princeton University Press, 2013). 4S.-Q. Shen, Topological Insulators: Dirac Equation in Con- densed Matters (Springer Series in Solid-State Sciences) (Springer-Verlag, Berlin, 2013). 5W. P. Su, J. R. Schrie er, and A. J. Heeger, Phys. Rev. Lett. 42, 1698 (1979). 6H. J. Schulz, Phys. Rev. B 18, 5756 (1978). 7J. Zak, Phys. Rev. Lett. 62, 2747 (1989). 8J. K. Asb oth, L. Oroszl any, and A. P alyi, Lecture Notes in Physics, Berlin Springer Verlag , Vol. 919 0075-8450 (Springer, 2016). 9M. Hagiwara, K. Katsumata, I. Aeck, B. I. Halperin, and J. P. Renard, Phys. Rev. Lett. 65, 3181 (1990). 10S. H. Glarum, S. Geschwind, K. M. Lee, M. L. Kaplan, and J. Michel, Phys. Rev. Lett. 67, 1614 (1991). 11V. L. J. Joly, M. Kiguchi, S.-J. Hao, K. Takai, T. Enoki, R. Sumii, K. Amemiya, H. Muramatsu, T. Hayashi, Y. A. Kim, M. Endo, J. Campos-Delgado, F. L opez-Ur as, A. Botello-M endez, H. Terrones, M. Terrones, and M. S. Dresselhaus, Phys. Rev. B 81, 245428 (2010). 12J. Campos-Delgado, J. M. Romo-Herrera, X. Jia, D. A. Cullen, H. Muramatsu, Y. A. Kim, T. Hayashi, Z. Ren, D. J. Smith, and Y. Okuno, Nano letters 8, 2773 (2008). 13M. Oshikawa and I. Aeck, Phys. Rev. Lett. 82, 5136 (1999). 14M. Oshikawa and I. Aeck, Phys. Rev. B 65, 134410 (2002). 15R. Shindou, A. Furusaki, and N. Nagaosa, Phys. Rev. B 82, 180505 (2010). 16M. Nakamura and A. Tokuno, Phys. Rev. B 94, 081411 (2016).17S. Konschuh, M. Gmitra, and J. Fabian, Phys. Rev. B 82, 245412 (2010). 18A. P. Schnyder, S. Ryu, A. Furusaki, and A. W. W. Lud- wig, Phys. Rev. B 78, 195125 (2008). 19C.-K. Chiu, J. C. Y. Teo, A. P. Schnyder, and S. Ryu, Rev. Mod. Phys. 88, 035005 (2016). 20P. Delplace, D. Ullmo, and G. Montambaux, Phys. Rev. B84, 195452 (2011). 21S. Ryu and Y. Hatsugai, Phys. Rev. Lett. 89, 077002 (2002). 22E. I. Rashba, Sov. Phys. Solid State 2, 1109 (1960). 23E. I. Rashba and A. L. Efros, Phys. Rev. Lett. 91, 126405 (2003). 24A. L. Efros and E. I. Rashba, Phys. Rev. B 73, 165325 (2006). 25A. Bolens, H. Katsura, M. Ogata, and S. Miyashita, Phys. Rev. B 95, 235115 (2017). 26A. Shekhter, M. Khodas, and A. M. Finkel'stein, Phys. Rev. B 71, 165329 (2005). 27S. Maiti, M. Imran, and D. L. Maslov, Phys. Rev. B 93, 045134 (2016). 28R. Kubo and K. Tomita, JPSJ 9, 888 (1954). 29S. Gangadharaiah, J. Sun, and O. A. Starykh, Phys. Rev. B78, 054436 (2008). 30T. A. Kaplan, Zeitschrift f ur Physik B Condensed Matter 49, 313 (1983). 31L. Shekhtman, O. Entin-Wohlman, and A. Aharony, Phys. Rev. Lett. 69, 836 (1992). 32D. Huertas-Hernando, F. Guinea, and A. Brataas, Phys. Rev. B 74, 155426 (2006). 33L. Brey, Phys. Rev. B 92, 235444 (2015). 34J. Nitta, T. Akazaki, H. Takayanagi, and T. Enoki, Phys. Rev. Lett. 78, 1335 (1997). 35A. D. Caviglia, M. Gabay, S. Gariglio, N. Reyren, C. Can- cellieri, and J. M. Triscone, Phys. Rev. Lett. 104, 126803 (2010). 36A. Soumyanarayanan, N. Reyren, A. Fert, and C. Panagopoulos, Nature 539, 509 (2016).
1809.10852v2.Spin_orbit_crossed_susceptibility_in_topological_Dirac_semimetals.pdf
arXiv:1809.10852v2 [cond-mat.mes-hall] 20 Feb 2019Spin-orbit crossed susceptibility in topological Dirac se mimetals Yuya Ominato1, Shuta Tatsumi1, and Kentaro Nomura1,2 1Institute for Materials Research, Tohoku University, Send ai 980-8577, Japan and 2Center for Spintronics Research Network, Tohoku Universit y, Sendai 980-8577,Japan (Dated: February 21, 2019) We theoretically study the spin-orbit crossed susceptibil ity of topological Dirac semimetals. Be- cause of strong spin-orbit coupling, theorbital motion of e lectrons is modulated byZeeman coupling, which contributes to orbital magnetization. We find that the spin-orbit crossed susceptibility is pro- portional totheseparation oftheDirac points anditis high lyanisotropic. The orbital magnetization is induced only along the rotational symmetry axis. We also s tudy the conventional spin susceptibil- ity. The spin susceptibility exhibits anisotropy and the sp in magnetization is induced only along the perpendicular to the rotational symmetry axis in contrast t o the spin-orbit crossed susceptibility. We quantitatively compare the two susceptibilities and find that they can be comparable. I. INTRODUCTION In the presence of an external magnetic field, magne- tization is induced by both the orbital motion and spin magnetic moment ofelectrons. When spin-orbit coupling isnegligible, the magnetizationiscomposedoftheorbital and spin magnetization, which are induced by the mini- mal substitution, p→p+eA, and the Zeeman coupling, respectively. Additionally, spin-orbit coupling gives rise to the spin-orbitcrossedresponse, in whichthe spinmag- netization is induced by the minimal substitution, and the orbital magnetization is induced by the Zeeman cou- pling. In the strongly spin-orbit coupled systems, the spin-orbit crossed response can give comparable contri- bution to the conventional spin and orbital magnetic re- sponses. Spin-orbit coupling plays a key role to realize a topo- logical phase of matter, such as topological insulators [1] and topological semimetals [2]. A natural question aris- ing is what kind of the spin-orbit crossed response occurs in the topological materials. Because of the topologically nontrivial electronic structure and the existence of the topological surface states, the topological materials ex- hibit the spin-orbit crossed response as a topological re- sponse [3–7]. The spin-orbit crossed response has been investigated in several systems. In the literature the con- nection between the spin-orbit crossed susceptibility and the spin Hall conductivity was pointed out [3, 4]. In recent theoretical work, the spin-orbit crossed response has been investigated also in Rashba spin-orbit coupled systems [8, 9]. The topological Dirac semimetal is one of the topo- logical semimetals [10–15] and experimentally observed in Na3Bi and Cd 3As2[16–18]. The topological Dirac semimetals have an inverted band structure originating from strong spin-orbit coupling. They are characterized by a pair of Dirac points in the bulk and Fermi arcs on the surface [10, 11]. The Dirac points are protected by rotational symmetry along the axis perpendicular to the (001) surface in the case of Na 3Bi and Cd 3As2[10, 11]. This is an important difference from the Dirac semimet- als appearing at the phase boundary of topological in- sulators and ordinary insulators [19–22], in which thereis no Fermi arc. A remarkable feature of the topological Dirac semimetals is the conservation of the spin angular momentum along the rotation axis within a low energy approximation [23]. The topological Dirac semimetals are regarded as layers of two-dimensional (2D) quantum spin Hall insulators (QSHI) stacked in momentum space and exhibit the intrinsic semi-quantized spin Hall effect. The magnetic responses of the generic Dirac electrons have been investigated in several theoretical papers. The orbital susceptibility logarithmically diverges and ex- hibits strong diamagnetism at the Dirac point [6, 24–26]. When spin-orbit coupling is not negligible, the spin sus- ceptibility becomes finite even at the Dirac point where the density of states vanishes [6, 27–29]. This is contrast to the conventional Pauli paramagnetism and known as the Van Vleck paramagnetism [29–32]. In this paper, we study the spin-orbit crossed suscep- tibility of the topological Dirac semimetals. We find that the spin-orbitcrossedsusceptibility isproportionalto the separation of the Dirac points and independent of the other microscopic parameters of the materials. We also include the spin conservation breaking term which mixes up and down spins [10, 11]. We confirm that the spin- orbit crossedsusceptibility is approximatelyproportional to the separation of the Dirac points even in the absence of the spin conservation as long as the separation is suf- ficiently small. We also calculate the spin susceptibility and quantitatively compare the two susceptibilities. Us- ing the material parameters for Na 3Bi and Cd 3As2, we show that the contribution of the spin-orbit crossed sus- ceptibilityisimportantinordertoappropriatelyestimate the total susceptibility. The paper is organized as follows. In Sec. II, we in- troduce a model Hamiltonian and define the spin-orbit crossed susceptibility. In Secs. III and IV, we calculate the spin-orbit crossed susceptibility and the spin suscep- tibility. In Secs. V and VI, the discussion and conclusion are given.2 II. MODEL HAMILTONIAN We consider a model Hamiltonian on the cubic lattice Hk=HTDS+Hxy+HZeeman, (1) which is composed of three terms. The first and sec- ond terms describe the electronic states in the topolog- ical Dirac semimetals, which reduces to the low energy effective Hamiltonian around the Γ point [10–12, 14, 15]. The first term is given by HTDS=εk+τxσztsin(kxa)−τytsin(kya)+τzmk,(2) where εk=C0−C1cos(kzc)−C2[cos(kxa)+cos(kya)], mk=m0+m1cos(kzc)+m2[cos(kxa)+cos(kya)]. (3) Pauli matrices σandτact on real and pseudo spin (or- bital) degrees of freedom. aandcare the lattice con- stants.t,C1, andC2are hopping parameters. C0gives constant energy shift. m0,m1, andm2are related to strength of spin-orbit coupling and lead band inversion. There are Dirac points at (0 ,0,±kD), kD=1 carccos/parenleftbigg −m0+2m2 m1/parenrightbigg . (4) The separation of the Dirac points is tuned by chang- ing the parameters, m0,m1, andm2. The first term, HTDS, commutes with the spin operator σz, andHTDS is regarded as the Bernevig-Hughes-Zhangmodel [12, 33] extended to three-dimension. The second term is given by Hxy=τxσxγ[cos(kya)−cos(kxa)]sin(kzc) +τxσyγsin(kxa)sin(kya)sin(kzc),(5) which mixes up and down spins. When Hxyis expanded around the Γ point, leading order terms are third or- der terms, which are related to the rotational symmetry alongtheaxisperpendiculartothe(001)surfaceinNa 3Bi and Cd 3As2. In the currentsystem, this axiscorresponds to thez-axis and we call it the rotational symmetry axis in the following. γcorresponds to the coefficient of the third order terms in the effective model [10, 11]. When γ iszero, the z-componentofspin conserves. At finite γ, on the otherhand, the z-componentofspin isnot conserved. Aswementionedintheintroduction,theexternalmag- netic field enters the Hamiltonian via the minimal substi- tution,p→p+eA, and the Zeeman coupling. We for- mally distinguish the magnetic field by the way it enters the Hamiltonianinordertoextractthespin-orbitcrossed response. BorbitandBspinrepresent the magnetic field in the minimal substitution and in the Zeeman coupling respectively. They are the same quantities so that we have to set Borbit=Bspinat the end of the calculation.In the following, the subscripts α,β,γ,δ refer tox,y,z. We define the orbital magnetization Morbit αand the spin magnetization Mspin αas follows Morbit α=−1 V∂Ω ∂Borbitα, (6) Mspin α=−1 V∂Ω ∂Bspin α, (7) where Ω is the thermodynamic potential and Vis the systemvolume. Thesequantitiesarewritten, up tolinear order in BorbitandBspin, as Morbit α=χorbit αβBorbit β+χSO αβBspin β, (8) Mspin α=χspin αβBspin β+χSO αβBorbit β, (9) where χorbit αβ=∂Morbit α ∂Borbit β, (10) χspin αβ=∂Mspin α ∂Bspin β, (11) χSO αβ=∂Morbit α ∂Bspin β=∂Mspin α ∂Borbit β. (12) Spin-orbit coupling can give the spin-orbit crossed sus- ceptibility χSO αβ, in addition to the conventional spin and orbital susceptibilities, χspin αβandχorbit αβ[6, 7]. In the rest of the paper, we focus on the Zeeman cou- pling, which can induce both of the orbital and spin mag- netization as we see in Eqs. (8) and (9). The Zeeman coupling is given by HZeeman=−µB 2/parenleftbigg gsσ0 0gpσ/parenrightbigg ·Bspin, =−g+µBτ0σ·Bspin−g−µBτzσ·Bspin,(13) whereµBis the Bohr magneton and gs,gpcorrespond to theg-factors of electrons in sandporbitals, respectively. We define g+= (gs+gp)/4 andg−= (gs−gp)/4, so that the Zeeman coupling contains two terms, the symmetric termτ0σand the antisymmetric term τzσ[7, 34, 35]. III. SPIN-ORBIT CROSSED SUSCEPTIBILITY A. Formulation The orbital magnetization is calculated by the formula [36–40], Morbit α=e 2/planckover2pi1/summationdisplay n/integraldisplay BZd3k (2π)3fnkǫαβγ ×Im/angbracketleft∂βn,k|(εnk+Hk−2µ)|∂γn,k/angbracketright,(14) wherefnk=/bracketleftbig 1+e(εnk−µ)/kBT/bracketrightbig−1is the Fermi distribu- tion function, ∂α=∂ ∂kα, and|n,k/angbracketrightis a eigenstate of Hk3 and its eigenenergy is εnk. The derivative of the eigen- states|∂αn,k/angbracketrightis expanded as [39] |∂αn,k/angbracketright=cn|n,k/angbracketright+/summationdisplay m/negationslash=n/angbracketleftm,k|/planckover2pi1vα|n,k/angbracketright εmk−εnk|m,k/angbracketright,(15) where the velocity operator vαis given by vα=∂αHk//planckover2pi1 andcnis a pure imaginary number. Using Eq. (15), the formula, Eq. (14), is written as Morbit α=e 2/planckover2pi1/summationdisplay n/integraldisplay BZd3k (2π)3fnkǫαβγ ×Im/summationdisplay m/negationslash=n/angbracketleftn,k|/planckover2pi1vβ|m,k/angbracketright/angbracketleftm,k|/planckover2pi1vγ|n,k/angbracketright (εmk−εnk)2(εnk+εmk−2µ). (16) Weusetheaboveformulainnumericalcalculation. Using the 2D orbital magnetization Morbit(2D) z(kz) at fixed kz, Morbit zis expressed as Morbit z=/integraldisplayπ/c −π/cdkz 2πMorbit(2D) z(kz).(17) The aboveexpressionis useful when wediscussnumerical results for χSO zz. We can relate χSO αβto the Kubo formula for the Hall conductivity, σαβ=e2 /planckover2pi1/summationdisplay n/integraldisplay BZd3k (2π)3fnkǫαβγ ×Im/summationdisplay m/negationslash=n/angbracketleftn,k|/planckover2pi1vβ|m,k/angbracketright/angbracketleftm,k|/planckover2pi1vγ|n,k/angbracketright (εmk−εnk)2.(18) When the density of states at the Fermi level vanishes, the intrinsic anomalous Hall conductivity is derived by the Streda formula [3, 4, 41], σαβ=−eǫαβγ∂Morbit γ ∂µ, =−eǫαβγ∂χSO γδ ∂µBspin δ. (19) The topological Dirac semimetals possess time reversal symmetry, so that the Hall conductivity is zero in the absence of the magnetic field. On the other hand, in the presence of the magnetic field, this formula suggests that the anomalous Hall conductivity at the Dirac point be- comes finite beside the ordinary Hall conductivity, if χSO γδ is not symmetric as a function of the Fermi energy εF. In the following section, we only consider χSO αα, because χSO αβ(α/negationslash=β) becomes zero from the view point of the crystalline symmetry in Na 3Bi and Cd 3As2. B. Numerical results Numerically differentiating Eq. (16) with respect to Bspin α, we obtain χSO αα. In Sec. III and IV, we omit εkin Eq. (2) for simplicity. This simplification does not change essential results in the following calculations. In Sec. V, we incorporate εkin order to compare the spin- orbit crossed susceptibility and the spin susceptibility quantitatively in Na 3Bi and Cd 3As2. Figure 1 shows the spin-orbit crossed susceptibility χSO zzatεF= 0 as a function of the separation of the Dirac points kD. In the present model, there are several parameters, such as t,a,m 0,and so on. We systematically change them and find which parameter affect the value of χSO zz. Figure 1 (a), (b), and (c) show that χSO zzincreases linearly with kDand satisfy following relation, χSO zz=g+µB2e hkD π. (20) χSO zzis proportional to the separation of the Dirac points kDand the coupling constant g+µB. Eq. (20) is given by numerical calculation. This result is understood as follows. χSO zzis obtained as χSO zz=/integraldisplayπ/c −π/cdkz 2πχSO zz(2D)(kz), (21) whereχSO zz(2D)(kz) is the 2D spin-orbit crossed suscepti- bility at fixed kz, which is defined in the same way as Eq. (12). χSO zz(2D)is quantized as 2 g+µBe/hin the 2D- QSHI and vanishes in the ordinary insulators [4, 7]. The topological Dirac semimetal is regarded as layers of the 2D-QSHI stacked in the momentum space and the spin Chern number on the kx-kyplane with fixed kzbecomes finite only between the Dirac points. As a result, we obtain Eq. (20). The sign of χSO zzdepends on the spin Chern number on the kx-kyplane with fixed kzbetween theDiracpoints. ThisisanalogoustotheanomalousHall conductivity in the Weyl semimetals [2, 23, 42]. In Fig. 1 (d),χSO zzincreases linearly at small kDbut deviates from Eq. (20) for finite γ. This is because the z-component of spin is not conserved in the presence of Hxy, Eq. (5), and the above argument for 2D-QSHI is not applicable to the present system. In the following calculation, we setm0=−2m2,m1=m2,m1/t= 1 and c/a= 1. Figure 2 shows χSO ααatεF= 0 as a function of γ. At γ= 0,χSO zzis finite as we mentioned above. On the other hand,χSO xxandχSO yyare zero. This means that the orbital magnetization is induced only along z-axis, which is the rotational symmetry axis. As a function of γ,χSO zzis an even function and χSO xx(yy)is an odd function. Figure 3 (a) shows χSO zzaround the Dirac point as a function of εF. Wheng−/g+= 0,χSO zzisan evenfunction around the Dirac point. At εF= 0,χSO zzis independent ofg−/g+as we see it in Fig. 1 (b). When g−/g+/negationslash= 0, however, χSO zzis asymmetric and the derivative of χSO zzis finite. This suggests that the Hall conductivity is finite wheng−/g+/negationslash= 0. Calculating Eq. (18) numerically, We confirm that the Hall conductivity is finite at εF= 0. Figure 3 (b) shows σxyas a function g−/g+.σxylinearly increases with g−/g+. The topological Dirac semimetal4 0.0 0.5 kD [ π/c]1.00.0 0.5 1.0(a)εF=0 0.0 0.5 kD [ π/c]1.00.0 0.5 1.0(c) 0.0 0.5 kD [ π/c]1.00.0 0.5 1.0(d)0.0 0.5 kD [ π/c]1.00.0 0.5 1.0(b) m1/t=0.5 1.0 2.0 g-/g+=0.0 0.5 1.0 c/a=1.0 2.0 3.0 γ/t=0.0 0.5 1.0 χzz [2g+μBe/(ha)]SO χzz [2g+μBe/(ha)]SO χzz [2g+μBe/(ha)]SO χzz [2g+μBe/(ha)]SO FIG. 1: The spin-orbit crossed susceptibility χSO zzatεF= 0 as a function of kD. We set the parameters m1=m2,m1/t= 1,g−/g+= 1,c/a= 1,andγ= 0, if the parameters are not indicated in each figure. The panels (a), (b), and (c) show that χSO zzis proportional to kD, which means that χSO zz reflects the topological property of the electronic structu re. From these numerical results, we obtain analytical express ion forχSO zz, Eq. (20), which is independent of model parameters except for kDandg+. The panel (d) show that Hxyreduces χSO zzbut it is negligible for sufficiently small kD. is viewed as a time reversal pair of the Weyl semimetal with up and down spin. Therefore, the Hall conductivity completely cancel with each other. Even in the presence ofg+Zeeman term (the symmetric term), the cancella- tion is retained. In the presence of g−Zeeman term (the antisymmetric term), on the other hand, the cancella- tion is broken. This is because g−Zeeman term changes the separation of the Dirac points and the direction of the change is opposite for the up and down spin Weyl semimetals. As a result, the Hall conductivity is finite in g−/g+/negationslash= 0 and given by σxy=2 πe2 hag−µBBspin t. (22) This expression is quantitatively consistent with the nu- merical result in Fig. 3 (b).γ [t] 1.5 1.0 0.5 0.0 -0.5 0.0 1.0 0.5 -0.5 -1.0χzz χyy χxx εF=0 SO SO SO χ [g+μB(2e/h)k D]SO FIG. 2: The spin-orbit crossed susceptibility as a function of γ. The solid black curve is χSO zz, the blue dashed curve is χSO xx, and the red dashed curve is χSO yy. We set the parameters m0=−2m2,m1=m2,m1/t= 1,g−/g+= 1, and c/a= 1. Breaking the conservation of σz, i.e., with the increase of γ, χSO zzis reduced, while χSO xxandχSO yybecome finite. IV. SPIN SUSCEPTIBILITY In this section, we calculate the spin susceptibility us- ing the Kubo formula, χspin αα(q,εF) =1 V/summationdisplay nmk−fnk+fmk−q εnk−εmk−q ×µ2 B|/angbracketleftn,k|g+τ0σα+g−τzσα|m,k−q/angbracketright|2, (23) whereVis the system volume, fnkis the Fermi distribu- tion function, εnkis energy of n-th band and |n,k/angbracketrightis a Bloch state of the unperturbed Hamiltonian. Taking the long wavelength limit |q| →0, we obtain lim |q|→0χspin αα(q,εF) =χintra αα(εF)+χinter αα(εF),(24) whereχintra αα(εF) is an intraband contribution, χintra αα(εF) =1 V/summationdisplay nk/parenleftbigg −∂fnk ∂εnk/parenrightbigg ×µ2 B|/angbracketleftn,k|g+τ0σα+g−τzσα|n,k/angbracketright|2,(25) andχinter αα(εF) is an interband contribution, χinter αα(εF) =1 V/summationdisplay n/negationslash=m,k−fnk+fmk εnk−εmk ×µ2 B|/angbracketleftn,k|g+τ0σα+g−τzσα|m,k/angbracketright|2.(26) At the zero temperature, only electronic states on the Fermi surface contribute to χintra αα. On the other hand, all electronic states below the Fermi energy can contribute toχinter αα[29]. From the above expression, we see that χinter ααbecomes finite, when the matrix elements of the5 -0.02-0.010.000.02 0.01 0 2 1 -1 -2 g-/g +σxy [e2/(ha)] εF [t]-1.0-0.50.0 0.0 1.0 0.5 -0.5 -1.01.5 1.0 0.5 -1.5g-/g +=1 g-/g +=-1 g-/g +=0 (a) (b) εF=0 Bspin =0.01t/(g +μB)χzz [g+μB(2e/h)k D]SO FIG.3: Thespin-orbitcrossed susceptibility χSO zzasafunction ofεFand the Hall conductivity as a function of g−/g+. We set the parameters m0=−2m2,m1=m2,m1/t= 1,c/a= 1, andγ= 0. At εF= 0, the value of χSO zzis independent of g−but itsεFdependence changes at finite g−. Consequently, the Hall conductivity becomes finite in accordance with Eq. (19). spin magnetization operatorbetween the conduction and valence bands is non-zero, i.e. the commutation relation between the Hamiltonian and the spin magnetization op- erator is non-zero. If the Hamiltonian and the spin mag- netization operator commute, /angbracketleftn,k|[Hk,g+τ0σα+g−τzσα]|m,k/angbracketright= 0,(27) the interband matrix element satisfies (εnk−εmk)/angbracketleftn,k|g+τ0σα+g−τzσα|m,k/angbracketright= 0.(28) This equation means that there is no interband matrix element and χinter αα= 0, because εnk−εmk/negationslash= 0. In the following, we set εF= 0, where the density of states vanishes. Therefore, there is no intraband contri- bution and we only consider the interband contribution. We numerically calculate Eq. (26). Figure 4 shows the spin susceptibility χspin ααas a function of (a) γand (b) g−/g+. In the following, we explain the qualitative be- havior of χspin ααusing the commutation relation between the Hamiltonianand the spin magnetizationoperator. In Fig. (4) (a), χspin zzvanishes at γ= 0, because the Hamil- tonian,HTDS, and the spin magnetization operator of z-component, g+µBτ0σz, commute, [HTDS,g+µBτ0σz] = 0. (29)For finite γ, on the other hand, χspin zzincreases with |γ|. This is because the commutation relation between Hxy andg+µBτ0σzis non-zero, [Hxy,g+µBτ0σz]/negationslash= 0, (30) andχinter zzgives finite contribution. χspin xxandχspin yyare finite even in the absence of Hxy, i.e.γ= 0, because HTDSandg+µBτ0σα(α=x,y) do not commute, [HTDS,g+µBτ0σx]/negationslash= 0, [HTDS,g+µBτ0σy]/negationslash= 0. (31) Atγ= 0,χspin xxis equal to χspin yy. For finite γ, however, they deviate from each other. This is because HTDSpos- sessesfour-foldrotationalsymmetryalong z-axisbut Hxy breaks the four-fold rotational symmetry. Figure (4) (b) shows that χSO zzbecomes finite when g−/g+/negationslash= 0. The antisymmetric term, g−µBτzσz, andHTDSdo not com- mute, [HTDS,g−µBτzσz]/negationslash= 0. (32) Consequently, χinter zzgives finite contribution, though the z-componentof spin is a good quantum number. The an- tisymmetric term does not break the four-fold rotational symmetry along z-axis, so that χspin xxis equal to χspin yyin Fig. (4) (b). The spin susceptibility χspin ααis also anisotropic but contrasts with the spin-orbit crossed susceptivity χSO αα. χspin xxandχspin yyare larger than χspin zz, in contrast χSO zzis larger than χSO xxandχSO yy. Therefore, the angle depen- dence measurement of magnetization will be useful to separate the contribution from the each susceptibility. V. DISCUSSION In this section, we quantitatively compare the spin- orbit crossed susceptibility χSO zzand the spin susceptibil- ityχspin zzat the Dirac points as a function of g−/g+. In the following calculation, we set the parameters to re- produce the energy band structure around the Γ point calculated by the first principle calculation for Cd 2As3 and Na 3Bi [10, 15]. The parameters are listed in the table and we omit Hxy, i.e.γ= 0. Figure 5 shows the two susceptibilities as a function of g−/g+. We find that the two susceptibilities are approx- imately written as χspin zz∼/parenleftbiggg− g+/parenrightbigg2 , (33) and χSO zz∼ −1 g+/parenleftbigg χ0+g− g+/parenrightbigg , (34) by numerical fitting. In the present parameters, χSO zzis negative and depends on g−/g+. The dependence on6 εF=0 χxx χyy χzz γ [t] g-/g +εF=0 0.6 0.4 0.2 0.0 0.0 1.0 0.5 -0.5 -1.00.0 1.0 0.5 -0.5 -1.00.30.4 0.2 0.00.1 (b) (a) χxx χyy χzz g-/g +=0 γ=0 spin spin spin spin spin spinχαα [(g +μB/2) 2/(ta 3)]spin χαα [(g +μB/2) 2/(ta 3)]spin FIG. 4: The spin susceptibility χspin ααatεF= 0 as a function of (a)γand (b) g−/g+. We set m0=−2m2,m1=m2, m1/t= 1, and c/a= 1. At γ= 0 and g−/g+= 0,χspin zz= 0 whileχspin xx,χspin yy>0. These behaviors are explained by the commutation relation between the Hamiltonian and the spin magnetization operators as discussed in the main text. g−/g+originates from the existence of εk, which breaks theparticle-holesymmetry. The g-factorsareexperimen- tally estimated as gs= 18.6 for Cd 2As3[43] andg−= 20 for Na 3Bi [44]. Unfortunately, there is no experimental datawhichdeterminesbothof gs,gporg+,g−. FromFig. 5, we see that χSO zzcan dominate over χspin zzifg−/g+≃0. As far as we know, there is no experimental observation of the magnetic susceptibility in these materials. We ex- pect the experimental observation in near future and our estimation of χSO zzwill be useful to appropriately analyze experimental data. Material parameters Cd3As2Na3Bi C00.306[eV] -1.183[eV] C10.033[eV] 0.188[eV] C20.144[eV] -0.654[eV] m00.376[eV] 1.754[eV] m1-0.058[eV] -0.228[eV] m2-0.169[eV] -0.806[eV] t0.070[eV] 0.485[eV] a12.64[˚A]5.07[˚A] c25.43[˚A]9.66[˚A] χ [(g +μB/2) 2/(ta 3)]0.030.04 0.02 0.000.01χ [(g +μB/2) 2/(ta 3)] 0.030.05 0.02 0.000.01 g-/g +0.0 0.4 0.2 -0.2 -0.40.04g+=5 g+=10 g+=15 g+=5 g+=10 g+=15 −χ zz SO −χ zz SO χzz spinχzz spin Na 3Bi Cd 2As 3 FIG. 5: The spin-orbit crossed susceptibility χSO zzand the spin susceptibility χspin zzat the Dirac points as a function of g−/g+. The dashed curve is χspin zzand the solid lines are χSO zz. The upper (lower) panel shows Cd 2As3(Na3Bi). When g−/g+are sufficiently small, χSO zzbecomes comparable to χspin zz. VI. CONCLUSION We theoretically study the spin-orbit crossed suscepti- bility of topological Dirac semimetals. We find that the spin-orbit crossed susceptibility along rotational symme- try axis is proportional to the separation of the Dirac points and is independent of the microscopic model pa- rameters. This means that χSO zzreflects topological prop- erty of the electronic structure. The spin-orbit crossed susceptibility is induced only along the rotational sym- metryaxis. We alsocalculatethe spinsusceptibility. The spin susceptibility is anisotropic and vanishingly small along the rotational symmetry axis, in contrast to the spin-orbit crossed susceptibility. The two susceptibilities are quantitatively compared for material parameters of Cd2As3and Na 3Bi. At the Dirac point, the orbital sus- ceptibility logarithmically diverges and gives dominant contribution to the total susceptibility. Off the Dirac point, on the other hand, the orbital susceptibility de- creases [6, 24, 25] and the contribution from the spin susceptibility and the spin-orbit crossed susceptibility is important for appropriate estimation of the total suscep- tibility.7 ACNOWLEDGEMENT This work was supported by JSPS KAKENHI Grant Numbers JP15H05854 and JP17K05485, and JSTCREST Grant Number JPMJCR18T2. [1] M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010). [2] N. P. Armitage, E. J. Mele, and A. Vishwanath, Rev. Mod. Phys. 90, 015001 (2018). [3] M.-F. Yang and M.-C. Chang, Phys. Rev. B 73, 073304 (2006). [4] S. Murakami, Phys. Rev. Lett. 97, 236805 (2006). [5] Y. Tserkovnyak, D. A. Pesin, and D. Loss, Phys. Rev. B 91, 041121 (2015). [6] M. Koshino and I. F. Hizbullah, Phys. Rev. B 93, 045201 (2016). [7] R. Nakai and K. Nomura, Phys. Rev. B 93, 214434 (2016). [8] H. Suzuura and T. Ando, Phys. Rev. B 94, 085303 (2016). [9] T. Ando and H. Suzuura, Journal of the Physical Society of Japan 86, 014701 (2017). [10] Z. Wang, Y. Sun, X.-Q. Chen, C. Franchini, G. Xu, H. Weng, X. Dai, and Z. Fang, Physical Review B 85, 195320 (2012). [11] Z. Wang, H. Weng, Q. Wu, X. Dai, and Z. Fang, Physical Review B 88, 125427 (2013). [12] T. Morimoto and A. Furusaki, Phys. Rev. B 89, 235127 (2014). [13] B.-J. Yang and N. Nagaosa, Nature Communications 5, 4898 EP (2014), article. [14] D. Pikulin, A. Chen, and M. Franz, Physical Review X 6, 041021 (2016). [15] J. Cano, B. Bradlyn, Z. Wang, M. Hirschberger, N. P. Ong, and B. A. Bernevig, Phys. Rev. B 95, 161306 (2017). [16] Z. Liu, B. Zhou, Y. Zhang, Z. Wang, H. Weng, D. Prab- hakaran, S.-K. Mo, Z. Shen, Z. Fang, X. Dai, et al., Sci- ence343, 864 (2014). [17] M. Neupane, S.-Y. Xu, R. Sankar, N. Alidoust, G. Bian, C. Liu, I. Belopolski, T.-R. Chang, H.-T. Jeng, H. Lin, et al., Nature communications 5(2014). [18] S. Borisenko, Q. Gibson, D. Evtushinsky, V. Zabolotnyy , B. B¨ uchner, and R. J. Cava, Physical review letters 113, 027603 (2014). [19] S. Murakami, New Journal of Physics 9, 356 (2007). [20] H. Guo, K. Sugawara, A. Takayama, S. Souma, T. Sato, N. Satoh, A. Ohnishi, M. Kitaura, M. Sasaki, Q.-K. Xue, et al., Physical Review B 83, 201104 (2011). [21] S.-Y. Xu, Y. Xia, L. Wray, S. Jia, F. Meier, J. Dil, J. Os- terwalder, B. Slomski, A. Bansil, H. Lin, et al., Science 332, 560 (2011). [22] T. Sato, K. Segawa, K. Kosaka, S. Souma, K. Nakayama, K. Eto, T. Minami, Y. Ando, and T. Takahashi, NaturePhysics7, 840 (2011). [23] A. A. Burkov and Y. B. Kim, Phys. Rev. Lett. 117, 136602 (2016). [24] H. Fukuyama and R. Kubo, Journal of the Physical So- ciety of Japan 28, 570 (1970). [25] M. Koshino and T. Ando, Phys. Rev. B 81, 195431 (2010). [26] G. P. Mikitik and Y. V. Sharlai, Phys. Rev. B 94, 195123 (2016). [27] A. Thakur, K. Sadhukhan, and A. Agarwal, Phys. Rev. B97, 035403 (2018). [28] J. Zhou and H.-R. Chang, Phys. Rev. B 97, 075202 (2018). [29] Y. Ominato and K. Nomura, Phys. Rev. B 97, 245207 (2018). [30] J. Van Vleck, The theory of electronic and magnetic sus- ceptibility (1932). [31] R. Yu, W. Zhang, H.-J. Zhang, S.-C. Zhang, X. Dai, and Z. Fang, Science 329, 61 (2010). [32] J. Zhang, C.-Z.Chang, P.Tang, Z.Zhang, X.Feng, K.Li, L.-l. Wang, X. Chen, C. Liu, W. Duan, et al., Science 339, 1582 (2013). [33] B. A. Bernevig, T. L. Hughes, and S.-C. Zhang, Science 314, 1757 (2006). [34] C.-X. Liu, X.-L. Qi, H. Zhang, X. Dai, Z. Fang, and S.-C. Zhang, Physical Review B 82, 045122 (2010). [35] R. Wakatsuki, M. Ezawa, and N. Nagaosa, Scientific Re- ports5, 13638 EP (2015). [36] G. Sundaramand Q.Niu, Phys.Rev. B 59, 14915 (1999). [37] D. Xiao, J. Shi, and Q. Niu, Phys. Rev. Lett. 95, 137204 (2005). [38] T. Thonhauser, D. Ceresoli, D. Vanderbilt, and R. Resta , Phys. Rev. Lett. 95, 137205 (2005). [39] D. Ceresoli, T. Thonhauser, D. Vanderbilt, and R. Resta , Phys. Rev. B 74, 024408 (2006). [40] J. Shi, G. Vignale, D. Xiao, and Q. Niu, Phys. Rev. Lett. 99, 197202 (2007). [41] P. Streda, Journal of Physics C: Solid State Physics 15, L717 (1982). [42] A. A. Burkov and L. Balents, Phys. Rev. Lett. 107, 127205 (2011). [43] S. Jeon, B. B. Zhou, A.Gyenis, B. E. Feldman, I. Kimchi, A. C. Potter, Q. D. Gibson, R. J. Cava, A. Vishwanath, and A. Yazdani, Nature Materials 13, 851 EP (2014). [44] J. Xiong, S. K. Kushwaha, T. Liang, J. W. Krizan, M. Hirschberger, W.Wang, R.Cava, andN.Ong, Science 350, 413 (2015).
2003.12221v1.Generalized_magnetoelectronic_circuit_theory_and_spin_relaxation_at_interfaces_in_magnetic_multilayers.pdf
Generalized magnetoelectronic circuit theory and spin relaxation at interfaces in magnetic multilayers G. G. Baez Flores,1Alexey A. Kovalev,1M. van Schilfgaarde,2and K. D. Belashchenko1 1Department of Physics and Astronomy and Nebraska Center for Materials and Nanoscience, University of Nebraska-Lincoln, Lincoln, Nebraska 68588, USA 2Department of Physics, King’s College London, Strand, London WC2R 2LS, United Kingdom (Dated: July 6, 2021) Spin transport at metallic interfaces is an essential ingredient of various spintronic device con- cepts, such as giant magnetoresistance, spin-transfer torque, and spin pumping. Spin-orbit coupling plays an important role in many such devices. In particular, spin current is partially absorbed at the interface due to spin-orbit coupling. We develop a general magnetoelectronic circuit theory and generalize the concept of the spin mixing conductance, accounting for various mechanisms respon- sible for spin-flip scattering. For the special case when exchange interactions dominate, we give a simple expression for the spin mixing conductance in terms of the contributions responsible for spin relaxation (i.e., spin memory loss), spin torque, and spin precession. The spin-memory loss param- eteris related to spin-flip transmission and reflection probabilities. There is no straightforward relation between spin torque and spin memory loss. We calculate the spin-flip scattering rates for N|N, F|N, F|F interfaces using the Landauer-Büttiker method within the linear muffin-tin orbital method and determine the values of using circuit theory. I. INTRODUCTION Spin-orbit coupling (SOC) plays an essential role at metallicinterfaces, especiallyinthecontextofspintrans- port related phenomena such as giant magnetoresistance (GMR),1,2spin injection and spin accumulation,3spin transfer torque,4spin pumping,5–7spin-orbit torque,8,9 spin Hall magnetoresistance (SMR),10and spin Seebeck effect (SSE).11–13The concept of the spin mixing conduc- tance,originallyintroducedwithinthemagnetoelectronic circuit theory,14plays a very important role in describing the spin transport at magnetic interfaces.15 Nevertheless, the spin mixing conductance in its orig- inal form cannot account for various important con- tributions associated with spin-flip processes,16–22cou- pling to the lattice,23,24and other effects associated with magnons.25–27One can generalize the concept of spin mixing conductance by considering spin pumping in the presence of spin-flip processes28or by considering the magnetoelectronic circuit theory in the presence of spin- flipscattering.29Sofarsuchgeneralizationswerenotable to clarify the role of interfacial spin relaxation (usually referred to as spin memory loss or spin loss) in processes responsible for spin pumping and spin-transfer torque. Recent progress in first-principles calculations of interfa- cial spin loss29suggests that an approach fully account- ing for spin-nonconserving processes can be developed. Experimentally, a great deal of data is available on the relationbetweenspin-orbitinteractionsandtheefficiency of spin-orbit torque.30–33This data is often interpreted intuitively in terms of the spin memory loss parameter,1 while lacking careful theoretical justification. In this work, we develop the most general form of the magnetoelectronic circuit theory and apply it to studies of spin transport, concentrating on such phenomena as spin-orbit torque and interfacial spin relaxation in mul-tilayers. We introduce a tensor form for the generalized spin mixing conductance describing spin-nonconserving processes, such as spin dephasing, spin memory loss, and spin precession. We numerically calculate parts of the spinmixingconductanceresponsibleforthespinmemory loss in N|N, F|N, F|F interfaces in the presence of spin- orbit interactions using the Landauer-Büttiker method based on linear muffin-tin orbital (LMTO) method. We show that the generalized spin mixing conductance can be also used to describe spin-orbit torque when exchange interactions dominate and the torque on the lattice can be disregarded. Our results for the generalized spin mix- ing conductance suggest that two distinct combinations of scattering amplitudes are responsible for spin mem- ory loss and torque, and in general there is no simple connection between the two. The paper is organized as follows. In Sec. II, we de- velop a general formulation of the magnetoelectronic cir- cuit theory in the presence of spin-flip scattering. In Sec. III, we apply the magnetoelectronic circuit theory to cal- culations of spin loss in (N 1N2)N, (N 1F2)N, or (F 1F2)N multilayers connected to ferromagnetic leads. In Sec. IV, we apply the magnetoelectronic circuit theory to spin- orbit torque calculations. Computational details are de- scribed in Sec. V, and the technicalities of the adiabatic embedding approach are detailed in Sec. VI. Section VII presents numerical results for the spin-flip transmission and reflection rates and area-resistance products for N|N, F|N, F|F interfaces. Section VIII concludes the paper.arXiv:2003.12221v1 [cond-mat.mes-hall] 27 Mar 20202 II. GENERALIZED CIRCUIT THEORY A. Formalism The magnetoelectronic circuit theory follows from the boundary conditions linking pairs of nodes in a circuit.14Here we consider the general case, allowing spin-nonconserving scattering at interfaces between mag- netic or non-magnetic metals due to the presence of spin- orbit interaction or non-uniform magnetization. The boundary condition at an interface between nodes 1 and 2, witharbitrarydistributionfunctions ^fa(a= 1;2labels the node), is: ^I2=G0X nmh ^t0 mn^f1(^t0 mn)y M2^f2^rmn^f2(^rmn)yi ; (1) whereG0=e2=h,^rmnis the spin-dependent reflection amplitude for electrons reflected from channel ninto channelmin node 2, ^t0 mnis the spin-dependent transmis- sion amplitude for electrons transmitted from channel n in node 1 into channel min node 2, and the Hermitian conjugate is taken only in spin space. Equation (1) can be easily rewritten for the current ^I1in node 1. For a ferromagnetic node, the spin accumulation is taken to be parallel to its magnetization. The matrices ^rmnand^t0 mn are generally off-diagonal in spin space. It is customary to assume that the distribution func- tions in the nodes, ^fa= ^0f0 a+^fs a, are isotropic, i.e., independent of k. In this case Eq. (1) reduces to gener- alized Kirchhoff relations:29 I0 2=Gcc 2f0+Gcs 2fsGm 2fs 2;(2) Is 2=Gsc 2f0+^Gss 2fs^Gm 2fs 2;(3) where f0=f0 1f0 2andfs=fs 1fs 2are interfacial drops of charge and spin components of the distribution function, and ^Ia= (^0I0 a+^Is a)=2. The conductances in Eq.(2-3)carryasubscript2emphasizingthattheygener- ally differ from their counterparts describing the currents in node 1; this subscript will be dropped where it doesn’t lead to confusion. The conductances are related through Gcs=GscGt,^Gss=Gcc^0^Gt,Gm=Gt+Gr, ^Gm=^Gt+^Grto the following scalar, vector, and tensorquantities: Gcc= 2G0X mnT mn; (4) Gt i= 4G0X mni"ijkTjk mn; (5) Gr i= 4G0X mni"ijkRjk mn; (6) Gsc i= 2G0X mn(Ti0 mn+T0i mn+i"ijkTjk mn); (7) Gt ij= 2G0kl ijX mn(Tkl mn+Tlk mn+i"klp[T0p mnTp0 mn]);(8) Gr ij= 2G0kl ijX mn(Rkl mn+Rlk mn+i"klp[R0p mnRp0 mn]); (9) wherekl ij=ijklikjl, Latin indices i;:::;ldenote Cartesian coordinates and m,nthe conduction channels, andrepeatedCartesianindicesaresummedoverhereand below. In the above expressions, we defined the following combinations of scattering matrix elements: R mn=Tr[(^rmn ^r mn)(^ ^)]=4;(10) T mn=Tr[(^t0 mn ^t0 mn)(^ ^)]=4;(11) where Greek indices can take values from 0 to 3. In order to obtain the circuit theory equations (2) and (3) from Eq. (1), we used the trace relations for Pauli matrices, Tr (^i^j) = 2ij, Tr(^i^j^k) = 2i"ijk, and Tr(^i^j^k^l) = 2(ijkl+iljkikjl). The unitar- ity condition gives the following identities: X mn^rmn^ry mn+^t0 mn(^t0 mn)y=M2^0; (12) X mn^r0 mn(^r0 mn)y+^tmn(^tmn)y=M1^0;(13) X mn^rmn(^rmn)y+^tmn(^tmn)y=M2^0;(14) X mn^r0 mn(^r0 mn)y+^t0 mn(^t0 mn)y=M1^0;(15) which relate the conductances defined for the two nodes separated by the interface as Gcc 1=Gcc 2,Gcs 1=Gcs 2+ Gm 2, and Gcs 2=Gcs 1+Gm 1. The interface conductances in the magnetoelectronic circuit theory have to be renormalized by the Sharvin re- sistancefortransparentOhmiccontacts34,35whichallows comparison between ab initio studies and experiment.36 The circuit theory in Eqs. (2) and (3) can be general- ized to account for the drift contributions in the nodes by renormalizing the conductances Gcc,Gcs,Gsc,Gm, ^Gss, and ^Gm. This can be done by connecting nodes 1 and2to proper reservoirs with spin-dependent distribu- tion functions ^fLand ^fRvia transparent contacts. The currents in the nodes then become ^I1= 2G0^M1(^fL^f1) and^I2= 2G0^M2(^f2^fR), where ^M1(2)describethenum- ber of channels (in general spin-dependent) in the nodes.3 Effectively, this leads to substitutions f"(#) 1!f"(#) 1+ I"(#) 1=(2G0M"(#) 1)andf"(#) 2!f"(#) 2I"(#) 2=(2G0M"(#) 2) in Eqs. (2) and (3). Finally, we note that the conductance ^Gmdescribes various spin-nonconserving processes, such as spin de- phasing, spin loss, and spin precession. Therefore, it can be interpreted as a tensor generalization of the spin mix- ing conductance14,37,38to systems with spin-flip scatter- ing. In the limiting case described in Ref. 28, our defi- nition reduces to the generalized tensor expression sug- gested there. However, our definition is more general as it can account for processes corresponding to spin pre- cession and spin memory loss. Spin-nonconserving pro- cesses can also result in spin-charge conversion (i.e., spin galvanic effect), which is described by GmandGcscon- ductances. Furthermore, Gscdescribes the conversion of charge imbalance into spin current (inverse spin galvanic effect), and ^Gssis the tensor spin conductance. B. Spin-conserving F|N interface We now apply the generalized circuit theory to an F|N interface. In the special case of a spin-conserving interface, Eqs. (2) and (3) should be invariant under SO(3)rotationsinspinspace, whichreproducesthespin- conserving circuit theory:14,37,38 Gm= 0; (16) Gcs=Gsc=Gscm; (17) ^Gss=Gccm m; (18) ^Gm= 2G"# r(^1m m) + 2G"# im;(19) where the tensor m mimplements a projection onto the magnetization direction, and G"# randG"# iare the real and imaginary parts of the spin-mixing conductance G"#=G0P mn(nmr"" mnr## mnt"" mnt## mn). C. General F|N interface To understand further the structure of current re- sponses, we expand the vector and tensor conductances in powers of magnetization: G i=G (0) i+G (1) i;kmk+G (2) i;klmkml+;(20) G ij=G (0) ij+G (1) ij;kmk+G (2) ij;klmkml+;(21) where stands forsc,cs,t,r, orm, stands forss, t,r, orm, and the tensors G (0) i,G (1) i;k,G (2) i;kl,G (0) ij, G (1) ij;k,G (2) ij;kl, etc. are invariant under the nonmagnetic point group of the system. The circuit theory substantially simplifies for axially symmetric interfaces, which are common in polycrys- talline heterostructures. Choosing the zaxis to be nor- mal to the interface and applying the constraints corre- sponding to the C1vsymmetry, we obtain the expansionof vector conductances Gsc,GscandGmto second order inm: ~G =0 B@mxx (1) 1+mymzx (2) 1 myx (1) 1mxmzx (2) 1 mzx (1) 21 CA;(22) wherex (1) 1,x (1) 2, andx (2) 1are arbitrary coefficients. For the tensor conductances ^Gssand ^Gmwe obtain ^G =0 B@x (0) 1 0 0 0x (0) 1 0 0 0x (0) 21 CA (23) +0 B@0mzx (1) 1myx (1) 2 mzx (1) 1 0mxx (1) 2 myx (1) 3mxx (1) 3 01 CA (24) +0 B@m2 xx (2) 1+m2 zx (2) 2mxmyx (2) 1mxmzx (2) 4 mxmyx (2) 1m2 yx (2) 1+m2 zx (2) 2mymzx (2) 4 mxmzx (2) 5 mymzx (2) 5m2 zx (2) 31 CA (25) wherex (0) 1,x (0) 2,x (1) 1,x (1) 2,x (1) 3,x (2) 1,x (2) 2,x (2) 3, x (2) 4, andx (2) 5are arbitrary coefficients. Theroleofspin-flipscatteringbecomesthemosttrans- parent if both the magnetization and the spin accumu- lation are either parallel or perpendicular to the inter- face. In this case, the tensor and vector conductances in Eqs. (2) and (3) can be simplified, and we arrive at thefollowingrelationsforrelevantcomponentsassociated with the in-plane and perpendicular directions: Gcc=G0(T""+T##+T"#+T#"); (26) Gsc=G0(T""T##+T"#T#"); (27) Gt= 2G0(T"#T#"); Gr= 2G0(R"#R#");(28) Gt= 2G0(T"#+T#");Gr= 2G0(R"#+R#");(29) alongwithGcs=GscGt,Gss=GccGt,Gm=Gt+Gr, andGm=Gt+Gr. Of course, all quantities in these ex- pressions are different for the in-plane and perpendicular orientations of the magnetization; the corresponding in- dex has been dropped to avoid clutter. The spin-resolved dimensionless transmittances and reflectances T0=X mnt0 mn(t0 mn); (30) R0=X mnr0 mn(r0 mn)(31) are defined in the reference frame with the spin quanti- zation axis aligned with the magnetization. Eqs. (26)-(29), together with Eqs. (2) and (3), are also valid for axially symmetric F|F interfaces, as long as the magnetizations of the two ferromagnets are collinear. These expressions generalize the result given in Ref. 29 for axially symmetric N|N junctions to include F|N and F|F interfaces.4 D. Relation to Valet-Fert theory The Valet-Fert model39incorporates spin relaxation in diffusive bulk regions but makes restrictive approx- imations for the interfaces, treating them as transpar- ent, spin-conserving, and prohibiting transverse spin accumulation.2,16,40–43When spin relaxation at inter- faces is of interest, the treatment based on the Valet-Fert model is forced to replace the interfaces by fictitious bulk regions,1,2which is restrictive even for N|N interfaces.29 Here we show how diffusive bulk regions can be incor- porated in the generalized circuit theory. By introduc- ing nodes near the interfaces and treating both interfaces and bulk regions as junctions, the generalized Kirchhoff’s rules2,16,40–43can be used to analyze entire devices with spin relaxation in the diffusive bulk regions and arbitrary spin-nonconserving scattering at interfaces. The Valet-Fert model employs the following equations to describe spin and charge diffusion in a normal metal: @2 x(DfN 0) = 0; (32) @2 @x2(DfN s) =fN s N sf; (33) and in a ferromagnet: @2 @x2(D"f"+D#f#) = 0; (34) @2 @x2(D"f"D#f#) =f"f# F sf: (35) Here fF s=m(f"f#)=2is the spin accumulation in the ferromagnet, and the spin-flip relaxation times N sf= (lN sf)2=DandF sf= (lF sf)2(1=D"+ 1=D#)=2are given in terms of the spin-diffusion lengths lN sf,lF sfand diffusion coefficients D,D. We now consider three basic circuit elements. 1. Diffusive N region For a diffusive N layer, the solution of Eqs. (32) and (33) leads to a simplified version of Eqs. (2) and (3) with vanishing vector conductances Gsc,Gsc,Gm, and all tensor conductances reduced to scalars: Gcc N=2D tN; (36) Gss N=Gcc NN sinhN; (37) Gm N=Gcc NNtanhN 2; (38) wheretNisthethicknessoftheNlayer, and N=tN=lN sf.2. Diffusive F region For a diffusive F layer with spin accumulation that is parallel to the magnetization, the solution of Eqs. (34) and (35) leads to vanishing Gmand the other conduc- tances defined as follows: Gcc F= (D"+D#)=tF; (39) Gsc F=Gcs F=m(D"D#)=tF; (40) Gss F=G FF sinhF+(Gsc F)2 Gcc F; (41) Gm F=G FFtanhF 2; (42) whereG F= [(Gcc F)2(Gsc F)2]=Gcc Fis the effective con- ductance and tFthe thickness of the F layer, and F= tF=lF sf. 3. Diffusive F|N junction As a simple application, consider a composite junc- tion consisting of F and N diffusive layers separated by a transparent interface. Such an idealized junction can be used to model an interface with spin-flip scattering between F and N layers.2,16,40–43Combining the results for F and N regions with boundary conditions, we find Gm= 0and the following effective conductances: Gcc= (1=Gcc F+ 1=Gcc N)1; (43) Gcs=Gsc=Gsc F; (44) ^Gss=Gss NGss F Gss N+Gss F+Gm N+Gm Fm m; (45) ^Gm=Gss N+Gm NGss N(Gss N+Gss F) Gss N+Gss F+Gm N+Gm Fm m;(46) where the conductances for the F and N layers should be taken from the previous subsections. If spin-flip scattering is negligible, we recover the known result:16 ^Gm=Gss N(1m m). III. SPIN LOSS AT INTERFACES The experimental data on interfacial spin relaxation comes primarily from the measurements of magnetore- sistance in (N 1N2)N, (N 1F2)N, or (F 1F2)Nmultilayers connected to ferromagnetic leads,1,2whereNis the num- ber of repetitions. The results have been reported1,2in terms of the effective spin memory loss parameter Nor Fobtained by treating the interface as a fictitious bulk layer and fitting the data to the Valet-Fert model. Here we relate the experimentally measured parameter Nor Fto the generalized conductances appearing in Eqs. (2) and (3). We assume that the interfaces are axially sym- metricandthatthemagnetizationandspinaccumulation are either parallel or perpendicular to the interface.5 A. N|N multilayer We first consider a multilayer with repeated interfaces between normal metals N 1and N 2. We would like to as- sess the decay of spin current which may include the spin relaxationbothatinterfacesandinthebulk. Tothisend, we place nodes in both N 1and N 2layers and consider the case of axially symmetric interfaces corresponding to relations, Gsc=Gsc=Gm= 0. The relevant conduc- tancesGcc,Gss,Gm 1, andGm 2account for the scattering in the bulk and/or at the interfaces. Using Eq. (3), we arrive at the following equations for the spin current in some arbitrary node iin the superlattice: Is i=Gss(fs i1fs i)Gm ifs i; (47) Is i=Gss(fs ifs i+1) +Gm ifs i; (48) which results in the recursive formula: 2Gm i Gssfs i=fs i12fs i+fs i+1: (49) This equation has analytical solutions: fi s=C1ei+C2ei; (50) where the constants C1andC2are determined by the boundary conditions. In the limit of weak spin-flip scat- tering, we obtain the leading term for the decay rate: 2Gm 1+Gm 2 Gcc; (51) where the constants C1andC2are defined by the bound- ary conditions. Note that to the lowest order in the spin-flip processes, only denominator in Eq. (51) needs to be renormalized by the Sharvin resistance for trans- parent Ohmic contacts, i.e., 1=~Gcc= 1=Gcc(1=M 1+ 1=M 2)=(4G0). It is clear that the constant describes how the spin current decays as we increase the num- ber of layers in the superlattice. The conductances in Eq. (51) may also include scattering in the bulk where the total conductances can be calculated by concatenat- ing the corresponding bulk and interface conductances using Eqs. (2) and (3). When obtaining from experi- mental data, one typically considers only interfacial con- tributions in Eq. (51), while the bulk contributions are simply removed.1This does not cause any problem when spin-orbit interaction is weak as in this limit the total Gmis a simple sum of contributions from interface and bulk. B. F|N and F|F multilayers By considering F|N and F|F multilayers connected to ferromagnetic leads one can also quantify spin relaxation at magnetic interfaces.1In this case, a parameter de- scribing the decay of spin current can also be relatedto the scattering matrix elements and to the general- ized conductances in Eq. (2) and (3). We assume that we have a superlattice with repeated interfaces between normal (N 1) and ferromagnetic (F 2) layers. Normal can be considered a special case of F in this section, equa- tions derived below also apply to F|F multilayers with- out any modifications. We would like to assess the decay of spin current due to spin relaxation at interfaces and in the bulk. We take nodes in F and N layers and con- sider the case of axially symmetric interfaces. We also assume collinear spin transport with the magnetization being in-plane or perpendicular to interfaces. The gen- eralized conductances may include scattering both in the bulk and at the interfaces. Using Eqs. (2) and (3), we arrive at the following equations for the spin and charge currents in node i: I0 i=Gcc(f0 i1f0 i) +Gcs i1(fs i1fs i)Gm ifs i;(52) I0 i=Gcc(f0 if0 i+1) +Gcs i+1(fs ifs i+1) +Gm ifs i;(53) Is i=Gsc i1(f0 i1f0 i) +Gss(fs i1fs i)Gm ifs i;(54) Is i=Gsc i+1(f0 if0 i+1) +Gss(fs ifs i+1) +Gm ifs i;(55) which results in the recursive formula: 2Gm i=Gsc i12Gm i=Gcc Gss=Gsc i1Gcs i1=Gccfs i=fs i12fs i+fs i+1;(56) Similar to non-magnetic case, the above equation has an- alytical solutions: fi s=C1ei+C2ei: (57) In the limit of weak spin-flip scattering, we obtain the leading term for the decay rate: 2Gm F+Gm N G; (58) whereG= [(Gcc)2(Gsc)2]=Gccis the effective con- ductance of the scattering region. Note that to the low- est order in the spin-flip processes, only denominator in Eq. (58) needs to be renormalized by the Sharvin re- sistance for transparent Ohmic contacts, i.e., 1=~G= 1=G(1=M" 1+ 1=M# 1+ 1=M" 2+ 1=M# 2)=(8G0). The constantdescribes how the spin current decays as we increase the number of layers in the multilayers. The conductances in Eq. (58) may also include scattering in the bulk. The bulk and interface conductances can be concatenated using Eqs. (2) and (3). IV. SPIN-ORBIT TORQUE The discontinuity of spin-current at the interface fol- lowing from the circuit theory in Eqs. (2) and (3) can be used to calculate the total torque transferred to both the magnetization and the lattice. In general, separating these two contributions is not possible without consider- ations beyond the circuit theory. When exchange inter- actions dominate and the torque on the lattice can be6 disregarded, we can use the circuit theory to calculate the spin torque on magnetization. Note that spin-flip scattering and spin memory loss can still be present even in the absence of the lattice torque, e.g., due to magnetic disorder at the interface. In the absence of angular momentum transfer to the lattice, it is natural to assume axial symmetry with re- spect to magnetization direction which results in simpli- fications in Eqs. (22), (23), (24), and (25), i.e., x (2) 1= 0, x (0) 1=x (0) 2,x (1) 1=x (1) 2=x (1) 3,x (2) 2= 0,x (2) 3= x (2) 4=x (2) 5=x (2) 1. This leads to the following gener- alization of Eq. (19) for the spin mixing conductance: ^Gm= 2G"# r(^1m m) + 2Gm km m+ 2G"# im;(59) whereG"# r=G0P mnRe(nmr"" mnr## mnt"" mnt## mn)de- scribes the absorption of transverse spin current and Gm k=G0(T"#+T#"+R"#+R#")the absorption of lon- gitudinal spin current (i.e., spin memory loss); G"# i= G0P mnIm(nmr"" mnr## mnt"" mnt## mn)describes the pre- cession of spins. Even though the formal expressions for G"# randG"# idid not change compared to Eq. (19), their values can still be affected by the presence of spin-flip scattering due to unitarity of the scattering matrix. The effect of the unitarity constraint, however, does not have a direct relation to the spin memory loss parameter .29 Using a typical spin-orbit torque geometry10and Eq. (3), we can write a boundary condition determining the torque: 2e2 ~~ F=e(^1m m)js= (^1m m)^Gms;(60) where sis the spin accumulation and ~ Fis the magneti- zation torque. The spin current can be further calculated from the diffusion equation: r2s=s=l2 sf; (61) and js= 2e@zs+jSH^y; (62) where the interface is orthogonal to zaxis andjSHis the spin Hall current. We recover conventional antidamping and field like torques: ~ F= (~jSH=2e)g"# rtanh=2 1 + 2g"# rcothm(m^y)(63) +g"# itanh=2 1 + 2g"# icothm^y# ; whereg"# r(i)= (lsf=)G"# r(i)andistheconductivityofthe normal metal. The results of this section are inconsistent with the notion that spin memory loss should directly af- fect spin-orbit torque.30–33As can be seen from Eq. (59), two separate parameters are responsible for spin mem- ory loss and spin-orbit torque, and in general there is nodirect connection between the two. In the presence of spin-orbit interactions, only the total torque acting on the lattice and magnetization can be obtained from the circuit theory. However, it seems that a similar conclu- sion can be reached about the absence of direct relation between spin memory loss and torque. V. COMPUTATIONAL DETAILS AND INTERFACE GEOMETRY The transmittances and reflectances (30)-(31) were calculated using the Landauer-Büttiker approach im- plemented in the tight-binding linear muffin-tin orbital (LMTO) method.44Spin-orbit coupling (SOC) was in- troduced as a perturbation to the LMTO potential parameters.44,45Local density approximation (LDA) was used for exchange and correlation.46 We have considered a number of interfaces between metals with the face-centered cubic lattice. The inter- faceswereassumedtobeepitaxialwiththe(111)or(001) crystallographic orientation. Lattice relaxations were ne- glected, and the average lattice parameter for the two lead metals was used for the given interface. The po- larization of the spin current and the magnetization (in F|N and F|F systems) were taken to be either parallel or perpendicular to the interface. Self-consistent charge and spin densities were obtained using periodic supercells with at least 12 monolayers of each metal. The surface Brillouin zone integration in transport calculations was performed with a 512512 mesh for magnetic and 128128for non-magnetic sys- tems. We also studied the influence of interfacial intermix- ing on spin-memory loss at Pt jPd and AujPd interfaces. One layer on each side of the interface was intermixed with the metal on the other side. The mixing concentra- tions were varied from 11% to 50%. For example, an A|B interface with 25% intermixing had two disordered lay- ers with compositions A 0:75B0:25and A 0:25B0:75between pure A and pure B leads. The transverse size of the su- percell was 22for 25% and 50% intermixing and 33 for 11% intermixing. The conductances were averaged over all possible configurations in the 22supercell and over 18 randomly generated configurations in 33. In addition, a model with long-range intermixing (LRI) was considered where the transition from pure A to pure B occurs over 8 intermixed monolayers with compositions A8=9B1=9, A 7=9B2=9,..., A 1=9B8=9. This model was im- plemented using 3 3 supercells. VI. ADIABATIC EMBEDDING In the Landauer-Büttiker approach, the active region where scattering takes place is embedded between ideal semi-infinite leads. In the circuit theory, the leads are imagined to be built into the nodes of the circuit on7 both sides of the given interface. In order to define spin- dependent scattering matrices with respect to the well- defined spin bases, we turn off SOC in the leads. Toavoidspuriousscatteringattheboundarieswiththe SOC-free leads, we introduce “ramp-up” regions between the interface and the leads, wherein the SOC is gradually increased from zero at the edges of the active region to its actual magnitude near the interface. Specifically, for an atom at a distance xfrom the interface ( jxj>l0), the SOC parameters are scaled by (L2jxj)=(L2l0), where Lis the total length of the active region and l0the length of the region on each side of the interface where SOC is retained at full strength. In our calculations we set l0to 2 monolayers. Because a slowly varying potential only allows scat- tering with a correspondingly small momentum transfer, such adiabatic embedding29allows a generic pure spin state from the lead to evolve without scattering into the bulk eigenstate of the metal before being scattered at the interface. In a non-magnetic metal, as explained in Ref. 29, adia- batic embedding leads to strong reflection near the lines on the Fermi surface where the group velocity is paral- lel to the interface. Geometrically, when projected or- thographically onto the plane of the interface, these lines formtheboundariesoftheprojectedFermisurface. Elec- trons with such wave vectors can backscatter from the SOC ramp-up region both with and without a spin flip. Thecontributionofthisbackscatteringtothespin-flipre- flectance is an artefact of adiabatic embedding and needs to be subtracted out.29In a magnetic lead such backscat- teringconservesspinandis,therefore,inconsequentialfor spin-memory loss calculations. Adiabatic embedding can also produce strong scatter- ing near the intersections of different sheets of the Fermi surface, where an electron can scatter from one sheet to another with a small momentum transfer. Such intersec- tions do not exist in non-magnetic metals considered in this paper (Cu, Ag, Au, Pd, Pt), but they are present in all ferromagnetic transition metals. When the two inter- secting sheets correspond to states of opposite spin, scat- tering from one sheet to the other is a spin-flip process. Depending on the signs of the normal (to the interface) components v?of the group velocities at the intersection, this scattering may or may not change the propagation direction with respect to the interface and thereby show up in spin-flip reflection or transmission. These two sit- uations are illustrated in Fig. 1. If v?has opposite signs on the two intersecting sheets [see Fig. 1(a-b)], then SOC opens a gap at the avoided crossing, and incident elec- trons with quasi-momenta close to the intersection are fully reflected from the ramp-up region with a spin flip. On the other hand, if v?has the same sign on the two sheets [see Fig 1(c-d)], then, instead of backscattering, there is a large probability of forward spin-flip scattering as the electron passes through the ramp-up region. Because we are interested in the spin-flip scattering processes introduced by the interface, the contribution FIG. 1. Crossing of the electronic bands in a ferromagnetic lead near an intersection of two Fermi surface sheets of op- posite spin. The parallel component of the quasi-momentum, kk, is fixed. (a-b) and (c-d): Cases where the normal compo- nent of the group velocity v?has the same or opposite sign on the two sheets, resulting in resonant spin-flip reflection or transmission, respectively. (a) and (c): no SOC; (b) and (d): avoided crossings induced by SOC. of spin-flip scattering due to the presence of the ramp- up regions in the leads should be subtracted out. Un- fortunately, this can only be done approximately. The approach used for N 1|N2interfaces in Ref. 29 was to sub- tract the spin-flip reflectances of auxiliary systems N 1|N1 and N 2|N2where the same lead material is used on both sidesofanimaginaryinterfacewithadiabaticembedding. This method is reasonable because the electrons incident from one of the leads and backscattered by the ramp-up region never reach the interface in the real N 1|N2system. In an F|N system, the same is true for the backscattering on Fermi sheet crossings in F [the case of Fig. 1(a-b)], but not for the forward scattering [the case of Fig. 1(c-d)]. Nevertheless, as a simple approximation, we extend the approach of Ref. 29 to the F|N interfaces, subtract- ing both the spin-flip reflectances in auxiliary F|F and N|N systems and the spin-flip transmittance in auxiliary F|F.Likewise,foranF 1|F2interface,wesubtractbothre- flectances and transmittances in F 1|F1and F 2|F2. Thus, for any kind of interface, we define T0 "#=T1j2 "#T1j1 "#T2j2 "#(64) R0 a;"#=R1j2 a;"#Raja "#; (65) wherea=Lora=Rdenotes one of the leads, and the primed quantities are used in Eq. (58). In the follow- ing, we refer to this as the subtraction method, and the parametercalculated in this way is denoted s.8 A.k-point filtering A more fine-grained approach is to identify the loca- tions in the surface Brillouin zone where spurious reflec- tion or transmission occurs and filter out the contribu- tions to spin-flip scattering probabilities from those lo- cations. This filtering requires care, because some spin- flip scattering processes near the Fermi surface crossings are, in fact, physical, rather than merely being artefacts of adiabatic embedding. This can be seen from Fig. 2, which shows possible spin-flip scattering processes facili- tated by the crossing of the Fermi sheets of opposite spin. Figure 2(a) shows a spin-flip backscattering process in the left lead, which can occur near a Fermi projection boundary in a normal metal or near a Fermi crossing of the type shown in Fig. 1(b). The processes shown in Figs. 2(b) and 2(c) result from the forward scattering near a Fermi crossing of the type shown in Fig. 1(d) in the left lead, where the electron is then either transmit- ted through or reflected from the interface, respectively. Each process has a reciprocal version. The three pro- cesses shown in Figs. 2(a-c) exist solely due to the pres- ence of a ramp-up region, which provides the small mo- mentum transfer needed to scatter from one Fermi sheet to another. In contrast, Figs. 2(d) and 2(e) show physical scatter- ing processes. Here, the momentum of an electron inci- dent from the left lead lies inside the spin-orbit gap of the type shown in Fig. 1(b) in the right lead. As a re- sult, the electron experiences a resonant spin-flip trans- mission [Fig. 2(d)] or reflection [Fig. 2(e)] at the inter- face. Resonant spin-flip transmission shown in Fig. 2(d) is possible because an electron can scatter to a different Fermi sheet with a large momentum transfer acquired from the interface. Illustrations in Fig. 2(d-e) are highly schematic because the wavefunction inside the spin-orbit gap is evanescent in the right lead. Letusfirstexaminethespin-flipscatteringprocessesin systems without a physical interface, where all scattering is due to adiabatic embedding alone. Spin-flip reflection at the Fermi projection boundaries can be seen in Figs. 3(a) and 3(d) for adiabatically embedded Pt and Pd, re- spectively, denoted in the figure caption as a fictitious “interface” of a material with itself (e.g., Pd|Pd).29The areas with strong spin-flip reflection are notably broader in Pt, which has a larger spin-orbit constant compared to Pd. Spin-flip reflection at Fermi crossings can be seen in Figs. 4(a) and 4(b) for adiabatically embedded Ni and Co, respectively. These two cases correspond to the dia- gram in Fig. 2(a). Spin-flip transmission at Fermi cross- ings in Ni and Co is seen, in turn, in Figs. 4(c) and 4(d); thisistheprocessshowninFig.2(b)withoutthephysical interface. Nowconsiderphysicalinterfaces. Contourswithstrong spin-flip reflection in, say, Fig. 3(d) for Pd|Pd are also seen in Fig. 3(c) for electrons incident from the Pd lead in Pt|Pd; the same comparison can be made for con- tours with strong spin-flip reflection in, say, Fig. 4(a) forNi|Ni and 4(g) for Ni|Co. These processes correspond to Fig. 1(a). Furthermore, the contours with strong spin- flip transmission in Fig. 4(c) for Ni|Ni show up in both Fig. 4(e) and 4(g) for spin-flip transmission and reflection in Ni|Co, respectively. These processes correspond to Fig. 2(b) and 2(c). The contours with resonant spin-flip transmission in Co|Co [Fig. 4(d)] also show up in spin- flip transmission for Ni|Co [Fig. 4(e)]; this corresponds to Fig. 2(b) with the two leads interchanged. All of the spin-flip scattering processes mentioned so far and corresponding to Fig. 2(a-c) are artefacts of adia- batic embedding and need to be filtered out in the calcu- lation of the interfacial spin loss parameter. On the other hand, the spin-flip transmission [Fig. 4(e)] and reflection [Fig. 4(g)] functions for the Ni|Co interface also show the spin-flip resonances of the types shown in Fig. 2(d-e). Consider the spin-flip reflection function for electrons in- cident from the Ni lead for the Ni|Co interface, which is shown in Fig. 4(g). Apart from the resonant contours appearing in Fig. 4(a) and 4(c) for spin-flip reflection and transmission in Ni|Ni, there are also resonant contours in Fig. 4(g) that correspond to the spin-flip reflection reso- nances in Co|Co, which are seen in Fig. 4(b). The same resonant contours appearing in Fig. 4(e) for the spin-flip transmission in Ni|Co correspond to the process shown in Fig. 2(d). These resonances correspond to the physical process depicted in Fig. 2(e) and should notbe filtered out in the calculation of the spin loss parameter. This analysis shows that both artefacts of adiabatic embedding [Fig. 2(a-c)] and physical resonant spin-flip scattering processes [Fig. 2(d-e)] can be located in k- spaceusingspin-fliptransmissionfunctionscalculatedfor auxiliarysystems. Thus, asanalternativetothesubtrac- tion method discussed above, the artefacts of adiabatic embedding can be removed using k-point filtering. For nonmagnetic (N 1|N2) interfaces, we first identify thek-points where the spin-flip reflectance in an auxil- iary system (N 1|N1or N 2|N2) exceeds a certain threshold value, which is chosen so that the spin-flip reflectance in the auxiliary system becomes less than 0:001G0if the contributions from the identified k-points are excluded. Then the contributions from those k-points are excluded in the calculation of the spin-flip reflectance for electrons incident from the corresponding lead. To ensure that the artefacts are fully removed, the excluded regions are slightly enlarged. Ferromagnetic leads induce resonant scattering near the crossings of the Fermi surfaces for opposite spins. Processes of the types shown in Fig. 2(a-c) should be filtered out, as explained above. We found that the spin- flip reflectances and transmittances for all ferromagnetic interfacesconsideredherearedominatedbyresonantpro- cessesdepictedinFig.2(d-e)ratherthanbycontributions from generic k-points. Indeed, the spin-loss parameters obtained by excluding the processes of Fig. 2(a-c) or by including only those in Fig. 2(d-e) are almost identical. Figures 4(i-l) show the spin-flip scattering functions ob- tained by starting from Figs. 4(e-h) and filtering out ev-9 FIG. 2. Spin-flip scattering mechanisms induced by a crossing of two Fermi sheets of opposite spin in an adiabatically embedded interface with no disorder. Dashed vertical lines show the interface; the label Fspecifies that the given metal must be ferromagnetic. Blue and red lines schematically show the trajectory of an electron before and after the spin flip. Crosses show physical spin-flip scattering processes, while circles denote those that occurs solely due to adiabatic embedding. FIG. 3.k-resolved spin-flip reflection functions for adiabatically embedded Pt|Pt, Pd|Pd, and Pt|Pd interfaces with and without k-point filtering. (a) R#"in PtjPt; (b)RL#"in PtjPd; (c)RR#"in PtjPd; (d)R#"in PdjPd; (e)R#"in PtjPt, filtered; (f) RL#" in PtjPd, filtered; (g) RR#"in PtjPd, filtered; (h) R#"in PdjPd, filtered. erything other than the processes of Fig. 2(d-e). By per- formingk-point filtering in this way we obtain a lower bound on the spin-flip scattering functions and the spin- loss parameter, ensuring that the artefacts of adiabatic embedding are completely removed. The values flisted in Table III were obtained in this way. VII. RESULTS A. Non-magnetic interfaces Table I lists the area-resistance products ARand the spin-loss parameters for nonmagnetic interfaces. The subtraction and k-point filtering methods result in simi- lar values of . For all material combinations, is quite similar for (001) and (111) interfaces, suggesting that the crystallographic structure of the interface does not havea strong effect on interfacial spin relaxation. In all cases, the spin-loss parameter is slightly lower for the parallel orientation of the spin accumulation relative to the inter- face. The calculated ARproducts and parameters are in goodagreementwithexperimentalmeasurements1insys- tems without Pd, but both are strongly overestimated for (Au,Ag,Cu,Pd)|Pd interfaces. However, the results for the Au|Pd (111) interface with the spin accumulation parallel to the interface are in good agreement with re- cent calculations of Gupta et al.47(AR= 0:81f m2and = 0:43) based on the analysis of the local spin currents near the interface. The large discrepancy in ARfor interfaces with Pd suggests that the idealized interface model is inadequate for these interfaces. Therefore, Pt jPd and AujPd with interfacial intermixing were also constructed as described in Section V. The results for intermixed interfaces are listed in Table II. It is notable that intermixing increases10 FIG. 4.k-resolved spin-flip transmission and reflection functions for Ni|Ni, Co|Co, and Ni|Co, and an illustration of k-point filtering. (a) R#"in NijNi; (b)R#"in CojCo; (c)T#"in NijNi; (d)T#"in Co|Co; (e) T#"in Ni|Co; (f) T"#in Ni|Co; (g) RL #"in NijCo; (h)RR #"in NijCo; (i)T#"in NijCo, filtered; (j) T"#in NijCo, filtered; (k) RL #"in NijCo, filtered; (l) RR #"in NijCo, filtered. theARproduct, while its values for ideal interfaces with Pd are already too large compared with experimental reports. The spin-loss parameter is also significantly increased by intermixing, which moves it further away from experimental data. The disagreement with experiment in the values of AR andforinterfaceswithPdislikelyduetothelackofun- derstanding of the interfacial structure in the sputtered multilayers, for which no structural characterization is available, to out knowledge. It seems somewhat implau- siblethattherealsputteredinterfacesaremuchlessresis- tive compared to both ideal or intermixed interfaces con- sidered here. It is possible that nominally bulk regions in sputtered multilayers containing Pd are more disor- dered and thereby have a higher resistivity and shorter spin-diffusion length compared to pure Pd films. The fit- ting procedure used to extract the ARandparameters for the interface1would then ascribe this additional bulk resistance and spin relaxation to the interfaces.B. Ferromagnetic interfaces Table III lists the results for interfaces with one or two ferromagnetic leads. The ARproducts for all interfaces are in excellent agreement with experimental data.1The values of the spin-loss parameter obtained using the sub- traction method ( s) tend to be larger, by up to a factor of2, comparedtothe k-pointfilteringmethod( f), which isexpectedtobemoreaccurate. ForPt|Cotheresultsfor ARandare in good agreement both with experiment and with calculations using the discontinuity of the spin current.47In other systems ARagrees very well with ex- periment but is underestimated, which may be due to the neglect of interfacial disorder and to the limitations of the adiabatic embedding method. VIII. CONCLUSIONS We have developed a general formalism for analyzing magnetoelectronic circuits with spin-nonconserving N|N, F|N, or F|F interfaces between diffusive bulk regions. A tensor generalization of the spin mixing conductance en-11 TABLE I. Area-resistance products AR(f m2) and spin- loss parameters obtained using the subtraction method ( s) and the filtering method ( f) for nonmagnetic interfaces. M denotes the orientation of the spin accumulation relative to the interface. NjNPlaneMARARexpsfexp PtjPd001k0.42 0.140.030.600.57 0.130.08?0.44 0.710.65 111k0.28 0.410.36 ?0.29 0.450.38 AujPd001k0.96 0.230.080.710.68 0.080.08?0.96 0.860.82 111k0.83 0.530.54 ?0.87 0.730.69 AgjPd001k0.92 0.350.080.410.47 0.150.08?1.12 0.500.54 111k0.89 0.410.47 ?0.92 0.500.55 CujPd001k0.81 0.450.0050.410.47 0.240.05?0.81 0.470.52 111k0.80 0.430.40 ?0.81 0.530.48 CujAu001k0.13 0.150.0050.080.08 0.130.07?0.13 0.110.11 111k0.11 0.080.07 ?0.12 0.110.10 CujPt001k0.90 0.750.051.000.87 0.90.1?0.89 1.070.9 111k0.75 0.880.72 ?0.82 1.110.83 CujAg001k0.03 0.0450.0050.020.2 0?0.03 0.030.02 111k0.13 0.030.03 ?0.13 0.040.04 codes all possible spin-nonconserving processes, such as spin dephasing, spin loss, and spin precession. In the special case when exchange interactions dominate, those contributions can be clearly separated into terms respon- sible for spin memory loss, spin-orbit torque, and spin precession. Surprisingly, there is no direct relation be- tween spin-orbit torque and spin memory loss; the two effects are described by different combinations of scat- tering amplitudes responsible for the absorption of the transverse and longitudinal components of spin current at the interface. The spin relaxation (i.e., spin memory loss) param- eterhas been numerically calculated using Eqs. (51) and (58) for a number of N|N, F|N, and F|F interfaces. First-principles calculations, aided by adiabatic embed- ding, show reasonable agreement with experiment for  and the area-resistance products with the exception of N|N interfaces including a Pd lead. For such interfacesTABLE II. Same as in Table I but for non-magnetic interfaces with intermixing. The percentage indicates the composition in the two intermixed layers. LRI refers to the long-range intermixing model; see Section V for details. NjN (mix %) PlaneMARARexpsfexp PtjPd (11%) 111k0.29 0.140.030.450.38 0.130.08?0.30 0.560.40 PtjPd (25%) 111k0.32 0.520.46 ?0.34 0.650.52 PtjPd (50%) 111k0.36 0.580.51 ?0.38 0.720.57 PtjPd (LRI) 111k0.82 1.200.91 ?0.85 1.340.96 AujPd (11%) 111k0.86 0.230.080.560.46 0.080.08?0.90 0.760.58 AujPd (25%) 111k0.96 0.600.58 ?1.01 0.810.73 AujPd (50%) 111k0.95 0.600.58 ?0.99 0.820.73 AujPd (LRI) 111k1.24 0.790.65 ?1.29 0.980.76 TABLEIII.SameasinTableIbutforF|NandF|Finterfaces. F(N)jFPlaneMAR"AR#ARARexpsfexp CujCo001k0.292.060.59 0.51 0.050.220.12 0.33 0.05?0.312.050.59 0.240.14 111k0.361.540.48 0.180.11 ?0.361.520.47 0.190.12 PtjCo001k0.464.671.28 0.85 0.121.120.91 0.9 0.4?0.444.601.26 1.170.96 111k1.701.360.76 0.810.72 ?1.821.380.80 0.910.80 AgjCo001k0.401.870.57 0.56 0.060.330.21 0.33 0.1?0.431.840.57 0.380.29 111k0.221.580.45 0.200.12 ?0.221.570.45 0.210.13 NijCo001k0.221.040.32 0.255 0.0250.320.15 0.35 0.05?0.241.020.32 0.340.16 111k0.210.730.23 0.270.17 ?0.250.720.24 0.290.16 bothandARare strongly overestimated, which can not be explained by short or long-range interfacial in- termixing. The analysis of spin-flip scattering probabil- ities for F|N and F|F interfaces suggests that interfacial spin relaxation is dominated by electronic states near the crossings of the Fermi surfaces for opposite spins in fer- romagnets. The generalized magnetoelectronic circuit theory pro- vides a convenient framework for analyzing spin trans- port in magnetic nanostructures with strong spin-orbit12 coupling at interfaces. ACKNOWLEDGMENTS A. K. is much indebted to Gerrit Bauer for stimulat- ing discussions on circuit theory with spin-flip scatter- ing. This work was supported by the National ScienceFoundation through Grant No. DMR-1609776 and the Nebraska MRSEC, Grant No. DMR-1420645, as well as by the DOE Early Career Award DE-SC0014189 (AK) and the EPSRC CCP9 Flagship project, EP/M011631/1 (MvS). Computations were performed utilizing the Hol- land Computing Center of the University of Nebraska, which receives support from the Nebraska Research Ini- tiative. 1J. Bass, J. Magn. Magn. Mater. 408, 244 (2016). 2J. Bass and W. P. Pratt, Jr., J. Phys.: Condens. Matter 19, 183201 (2007). 3M. Johnson and R. H. Silsbee, Phys. Rev. Lett. 55, 1790 (1985). 4D. Ralph and M. Stiles, J. Magn. Magn. Mater. 320, 1190 (2008). 5O. Mosendz, J. E. Pearson, F. Y. Fradin, G. E. W. Bauer, S. D. Bader, and A. Hoffmann, Phys. Rev. Lett. 104, 046601 (2010). 6B.Heinrich,C.Burrowes,E.Montoya,B.Kardasz,E.Girt, Y.-Y. Song, Y. Sun, and M. Wu, Physical Review Letters 107, 066604 (2011). 7Y. Tserkovnyak, A. Brataas, G. E. W. Bauer, and B. I. Halperin, Rev. Mod. Phys. 77, 1375 (2005). 8L. Liu, C.-F. Pai, Y. Li, H. W. Tseng, D. C. Ralph, and R. A. Buhrman, Science 336, 555 (2012). 9V. E. Demidov, S. Urazhdin, H. Ulrichs, V. Tiberkevich, A. Slavin, D. Baither, G. Schmitz, and S. O. Demokritov, Nature Materials 11, 1028 (2012). 10Y.-T. Chen, S. Takahashi, H. Nakayama, M. Althammer, S. T. B. Goennenwein, E. Saitoh, and G. E. W. Bauer, Phys. Rev. B 87, 144411 (2013). 11G.E.W.Bauer, E.Saitoh, andB.J.vanWees, Nat.Mater. 11, 391 (2012). 12H. Adachi, K.-I. Uchida, E. Saitoh, and S. Maekawa, Rep. Prog. Phys. 76, 036501 (2013). 13D. Meier, D. Reinhardt, M. van Straaten, C. Klewe, M. Al- thammer, M. Schreier, S. T. B. Goennenwein, A. Gupta, M. Schmid, C. H. Back, et al., Nat. Commun. 6, 8211 (2015). 14A. Brataas, G. E. W. Bauer, and P. J. Kelly, Phys. Rep. 427, 157 (2006). 15M. Weiler, M. Althammer, M. Schreier, J. Lotze, M. Pern- peintner, S.Meyer, H.Huebl, R.Gross, A.Kamra, J.Xiao, et al., Phys. Rev. Lett. 111, 176601 (2013). 16A. A. Kovalev, A. Brataas, and G. E. Bauer, Phys. Rev. B66, 224424 (2002). 17P.M.Haney, H.-W.Lee, K.-J.Lee, A.Manchon, andM.D. Stiles, Phys. Rev. B 87, 174411 (2013). 18J.-C.Rojas-Sánchez, N.Reyren, P.Laczkowski, W.Savero, J.-P.Attané, C.Deranlot,M.Jamet, J.-M.George,L.Vila, and H. Jaffrès, Physical Review Letters 112, 106602 (2014). 19K.ChenandS.Zhang, Phy.Rev.Lett. 114, 126602(2015). 20V. P. Amin and M. D. Stiles, Phys. Rev. B 94, 104420 (2016). 21V. P. Amin and M. D. Stiles, Phys. Rev. B 94, 104419 (2016). 22X. Tao, Q. Liu, B. Miao, R. Yu, Z. Feng, L. Sun, B. You,J. Du, K. Chen, S. Zhang, et al., Sci. Adv. 4, eaat1670 (2018). 23A. A. Kovalev, G. E. W. Bauer, and A. Brataas, Phys. Rev. B75, 014430 (2007). 24P. M. Haney and M. D. Stiles, Phys. Rev. Lett. 105, 126602 (2010). 25R. Arias and D. L. Mills, Phys. Rev. B 60, 7395 (1999). 26A. Azevedo, A. B. Oliveira, F. M. de Aguiar, and S. M. Rezende, Phys. Rev. B 62, 5331 (2000). 27M. Beens, J. P. Heremans, Y. Tserkovnyak, and R. A. Duine, J. Phys. D: Appl. Phys. 51, 394002 (2018), 1804.02172. 28Y. Tserkovnyak and H. Ochoa, Phys. Rev. B 96, 100402 (2017). 29K. D. Belashchenko, A. A. Kovalev, and M. van Schilf- gaarde, Phys. Rev. Lett. 117, 207204 (2016). 30M.-H. Nguyen, D. C. Ralph, and R. A. Buhrman, Phys. Rev. Lett. 116, 126601 (2016). 31C.-F.Pai, Y.Ou, L.H.Vilela-Leão, D.C.Ralph, andR.A. Buhrman, Phys. Rev. B 92, 064426 (2015). 32J.-C.Rojas-Sánchez, N.Reyren, P.Laczkowski, W.Savero, J.-P.Attané,C.Deranlot,M.Jamet, J.-M.George,L.Vila, and H. Jaffrès, Phys. Rev. Lett. 112, 106602 (2014). 33L. Zhu, D. C. Ralph, and R. A. Buhrman, Phys. Rev. Lett. 122, 077201 (2019). 34K. M. Schep, J. B. A. N. van Hoof, P. J. Kelly, G. E. W. Bauer, andJ.E.Inglesfield, Phys.Rev.B 56, 10805(1997). 35G. E. W. Bauer, Y. Tserkovnyak, D. Huertas-Hernando, and A. Brataas, Phys. Rev. B 67, 094421 (2003). 36A. A. Kovalev, G. E. W. Bauer, and A. Brataas, Phys. Rev. B73, 054407 (2006). 37A. Brataas, Y. V. Nazarov, and G. E. W. Bauer, Phys. Rev. Lett. 84, 2481 (2000). 38A. Brataas, Y. V. Nazarov, and G. E. Bauer, Eur. Phys. J. B22, 99 (2001). 39T. Valet and A. Fert, Phys. Rev. B 48, 7099 (1993). 40K. Eid, D. Portner, J. A. Borchers, R. Loloee, M. Al- Haj Darwish, M. Tsoi, R. D. Slater, K. V. O’Donovan, H. Kurt, W. P. Pratt, et al., Phys. Rev. B 65, 054424 (2002). 41J. Barnaś, A. Fert, M. Gmitra, I. Weymann, and V. K. Dugaev, Phys. Rev. B 72, 024426 (2005). 42S. Urazhdin, R. Loloee, and W. P. Pratt, Jr., Phys. Rev. B71, 100401 (2005). 43Y. Liu, Z. Yuan, R. J. H. Wesselink, A. A. Starikov, and P. J. Kelly, Phys. Rev. Lett. 113, 207202 (2014). 44D.Pashov, S.Acharya, W.R.Lambrecht, J.Jackson, K.D. Belashchenko, A. Chantis, F. Jamet, and M. van Schilf- gaarde, Comput. Phys. Commun. 249, 107065 (2020). 45I. Turek, V. Drchal, and J. Kudrnovská, Philos. Mag. 88,13 2787 (2008). 46U. von Barth and L. Hedin, J. Phys. C: Solid State Physics 5, 1629 (1972).47K. Gupta, R. J. H. Wesselink, R. Liu, Z. Yuan, and P. J. Kelly, arXiv:2001.11520 (2020).
1708.07247v2.Strong_influence_of_spin_orbit_coupling_on_magnetotransport_in_two_dimensional_hole_systems.pdf
arXiv:1708.07247v2 [cond-mat.mes-hall] 25 Aug 2017Strong influence of spin-orbit coupling on magnetotranspor t in two-dimensional hole systems Hong Liu, E. Marcellina, A. R. Hamilton and Dimitrie Culcer School of Physics and Australian Research Council Centre of Excellence in Low-Energy Electronics Technologies, UNSW Node, The University of New South Wales, Sydney 2052, Au stralia (Dated: August 28, 2017) With a view to electrical spin manipulation and quantum comp uting applications, recent signifi- cant attention has been devoted to semiconductor hole syste ms, which have very strong spin-orbit interactions. However, experimentally measuring, identi fying, and quantifying spin-orbit coupling effects in transport, such as electrically-induced spin pol arizations and spin-Hall currents, are chal- lenging. Here we show that the magnetotransport properties of two-dimensional (2D) hole systems display strong signatures of the spin-orbit interaction. S pecifically, the low-magnetic field Hall co- efficient and longitudinal conductivity contain a contribut ion that is second order in the spin-orbit interaction coefficient and is non-linear in the carrier numb er density. We propose an appropriate experimental setup to probe these spin-orbit dependent mag netotransport properties, which will permit one to extract the spin-orbit coefficient directly fro m the magnetotransport. Low-dimensional hole systems have attracted consid- erable recent attention in the context of nanoelectron- ics and quantum information [ 1–9]. They exhibit strong spin-orbit coupling but a weak hyperfine interaction, which allows fast, low-power electrical spin manipula- tion [10,11] and potentially increased coherence times [12–15] while their effective spin-3/2 is responsible for physics inaccessible in electron systems [ 16–20]. Struc- tures with strong spin-orbit interactions coupled to su- perconductors may enable topological superconductivity hosting Majorana bound states and non-Abelian particle statistics relevant for topological quantum computation [21–24]. In the past fabricating high-quality hole struc- tures was challenging, but recent years have witnessed extraordinary experimental progress [ 12,25–43]. A full quantitative understanding of spin-orbit cou- pling mechanisms is vital for the realization of spin- tronics devices and quantum computation architectures [44,45]. At the same time experimental measurement of spin-orbit parameters is difficult [ 46]. Spin-orbit con- stants can be estimated from weak antilocalization [ 47– 50], Shubnikov-de Haas oscillations and spin precession in large magnetic fields (up to 2 T) [ 51–53], and state-of- the-art optical measurements [ 54,55]. Many techniques yield only the ratio between the Rashba and Dresselhaus terms or allow the determination of only one type of spin splitting. Likewise,experimentallyquantifyingspin-orbit induced effects, such as via spin-to-charge conversion or vice versa, is difficult. For instance, current-induced spin polarizations in spin-orbit coupled systems are small and their relationship to theoretical estimates is ambiguous [56–58], while spin-Hall currents [ 59] can only be identi- fied via an edge spin accumulation [ 60–62]. Here we show that the spin-orbit interaction can have a sizeable effect on low magnetic-field Hall transport in a 2D hole system, which is density-dependent and experi- mentally visible. Our central result, shown in Fig. 1, is a(a) 1.2 1.1 1.0RH/R0 10 8 6 4 2 0 Fz(MV/m)p = 1 x 1011cm-2 p = 1.5 x 1011cm-2 p = 2 x 1011cm-2GaAs QW1.4 1.3 1.2 1.1 1.0RH/R0GaAs InAs InSbp = 2 x 1011cm-2 (b) Figure 1. Spin-orbit correction to the Hall coefficient RHof 2D holes in various 15 nm quantum wells as a function of the electric field Fzacross the well, where R0≡1 peis the bare Hall coefficient. Panel shows results for (a) different quantu m well materials at p= 1×1011cm−2and (b) GaAs quantum wells at different densities. correction to the low-field Hall coefficient RH=1 pe/bracketleftbigg 1+/parenleftbigg64πm∗2α2 /planckover2pi14/parenrightbigg p/bracketrightbigg , (1) whereαis the coefficient of the cubic Rashba spin-orbit term, which arisesfromthe applicationofan electricfield Fzacrossthequantumwell, m∗istheheavy-holeeffective mass atα= 0,pis the hole density, and eis the elemen-2 tary charge. Note that here we have chosen the z−axis as the quantization direction. In hole systems, where the spin-orbit coupling can account for as much as 40% of the Fermi energy [ 63], effects of second-order in the spin- orbit strength can be sizable in charge transport. These reflect spin-orbit corrections to the occupation probabili- ties, densityofstates, andscatteringprobabilities, aswell as the feedback of the current-induced spin polarization on the charge current. Quantitative evaluation shows that the spin-orbit corrections can reach more than 10% in GaAs quantum wells, and are of the order ∼20−30% in InAs and InSb quantum wells (Fig. 1a). The magni- tude of the spin-orbit corrections also increase with den- sity, which is consistent with the expectation that the strength of spin-orbit interaction increases with density (Fig.1b). It is worth noting that the correction due to spin-orbit coupling has already taken into account the fact that the spin-split subbands may have different hole mobilities. In the following we derive the formalism and show how spin-orbit coupling can give rise to corrections in the magnetotransport. We consider a 2D hole system in the presence of a constant electric field Fand a perpen- dicular magnetic field B=Bzˆz. The full Hamiltonian is ˆH=ˆH0+ˆHE+ˆU+ˆHZ, where the band Hamiltonian ˆH0 is defined below in Eq. ( 2),ˆHE=−eF·ˆrrepresents the interactionwith the externalelectricfield ofholes ˆristhe position operator, and ˆUis the impurity potential, dis- cussed below. The Zeeman term HZ= 3κµBσ·Bwith κis a material-specific parameter [ 16],µBthe Bohr mag- neton and σthe vector of Pauli spin matrices. Rashba spin-orbit coupling is expected to dominate greatly over the Dresselhaus term in 2D hole gases, even in materi- als such as InSb in which the bulk Dresselhaus term is very large [ 63]. With this in mind, the band Hamiltonian used in our analysis in the absence of a magnetic field is written as [ 64] H0k=/planckover2pi12k2 2m∗+iα(k3 −σ+−k3 +σ+)≡/planckover2pi12k2 2m∗+σ·Ωk,(2) wherem∗=m0 γ1+γ2, the Pauli matrix σ±=1 2(σx±iσy), k±=kx±iky. ForB= 0 the eigenvalues of the band Hamiltonian are εk±=/planckover2pi12k2/(2m∗)±αk3. In an exter- nal magnetic field we replace kby the gauge-invariant crystal momentum ˜k=k−eAwith the vector potential A=1 2(−y,x,0). The magnetic field is assumed small enough that Landau quantization can be neglected, in other words ωcτp≪1, where ωc=eBz/m∗is the cy- clotron frequency and τpthe momentum relaxation time. To set up our transport formalism, in the spirit of Ref. [65], we begin with a set of time-independent states {ks}, where srepresents the twofold heavy-hole pseu- dospin. We work in terms of the canonical momentum /planckover2pi1k. The terms ˆH0,ˆHEandˆHZare diagonal in wave vec- tor but off-diagonal in band index while for elastic scat- tering in the first Born approximation Uss′ kk′=Ukk′δss′. Without loss of generality, here we consider short-range impurity scattering. The impurities are assumed uncor-related and the averageof ∝an}bracketle{tks|ˆU|k′s′∝an}bracketri}ht∝an}bracketle{tk′s′ˆU|ks∝an}bracketri}htoverim- purityconfigurationsis( ni|¯Uk′k|2δss′)/V, whereniis the impurity density, Vthe crystalvolume, and ¯Uk′kthe ma- trix element of the potential of a single impurity. The central quantity in our theory is the density oper- ator ˆρ, which satisfies the quantum Liouville equation, dˆρ dt+i /planckover2pi1[ˆH,ˆρ] = 0. (3) The matrix elements of ˆ ρare ˆρkk′≡ˆρss′ kk′=∝an}bracketle{tks|ˆρ|k′s′∝an}bracketri}ht with understanding that ˆ ρkk′is a matrix in heavy hole subspace. The density matrix ρkk′is written as ρkk′= fkδkk′+gkk′, wherefkis diagonal in wave vector, while gkk′is off-diagonal in wave vector. The quantity of in- terest in determining the charge current is fksince the current operator is diagonal in wave vector. We there- fore derive an effective equation for this quantity by first breaking down the quantum Liouville equation into the kinetic equations of fkandgkk′separately, and fkobeys dfk dt+i /planckover2pi1[H0k+HZ,fk]+ˆJ(fk) =DE,k+DL,k,(4) where the scattering term in the Born approximation ˆJ(fk)=1 /planckover2pi12/integraldisplay∞ 0dt′[ˆU,e−iH0t′ /planckover2pi1[ˆU,ˆf(t)]eiH0t′ /planckover2pi1]kk,(5) and the driving terms DE,k=−eE /planckover2pi1·∂fk ∂k, (6a) DL,k=1 2e /planckover2pi1{ˆv×B,∂fk ∂k}, (6b) stem from the applied electric field and Lorentz force re- spectively [ 65]. In external electric and magnetic fields one may decompose fk=f0k+fEk+fEBk, wheref0kis theequilibriumdensitymatrix, fEkisacorrectiontofirst orderin the electricfield (but atzeromagnetic field), and fEBkis an additional correction that isfirst order in the electricand magnetic fields. The equilibrium density ma- trix is written as f0k= (1/2)[(fk++fk−)11+σ·ˆΩ(fk+− fk−)], where ˆΩis a unit vector and Ωwas defined in Eq. (2), andfk±represent the Fermi-Dirac distribution functions corresponding to the two band energies εk±. In linear response one may replace fk→f0kin Eq. (6a). On the other hand it is trivial to check that the driving termDL,kvanishes when the equilibrium density matrix is substituted, so in Eq. ( 6b) one may replace fk→fEk. Hence, in this work we perform a perturbation expansion up to first order in the electric and magnetic fields, and up to second orderin the spin-orbit interaction, retaining termsuptoorder α2. ThedetailedsolutionofEq.( 4)and the explicit evaluation of the scattering term Eq. ( 5) are given in the Supplement. We briefly summarize the pro- cedure here. Firstly, with f0kknown and only DE,kon the right-hand side of Eq. ( 4), we obtain fEk. Next, with3 onlyDL,kon the right-hand side of Eq. ( 4), we obtain fEBk. By taking the trace with current operator the lon- gitudinal and transverse components of the current are found as jx,y=eTr/bracketleftbig ˆvx,yfk/bracketrightbig , withvi= (1//planckover2pi1)∂H0k/∂k. Finally, with σxxandσxythe longitudinal and Hall con- ductivities respectively, the Hall coefficient appearing in Eq. (1) is found through RH=σxy Bz(σ2xx+σ2xy). For the Hall conductivity on the other hand one needs fEBk. We note that the topological Berrycurvature terms that give contributions analogous to the anomalous Hall effect in Rashba systems (with the magnetization replaced by the magnetic field Bz) vanish identically when both the band structure and the disorder terms are taken into account. Table I. The maximal hole densities for which the current theory is applicable for 15 nm-wide GaAs, InAs, and InSb quantum wells. Densities in units of 1011cm−2. GaAs InAs InSb 6.55 8.08 8.60 The limits of applicability of our approach are as fol- lows. We assume that the magnetotransport considered here occurs in the weak disorder regime, i.e. εFτp//planckover2pi1≫1, whereεFis Fermi energy. Furthermore, we assume that the scatteringdoes not changeappreciablywhen the gate field is changed at low density [ 40], so the condition εFτp//planckover2pi1≫1 is still valid when the gate field is changed. We assume αk3 F/ǫkin≪1 where ǫkin=/planckover2pi12k2 F 2m∗is kinetic energy, for example in Ref. [ 48], the spin-orbit-induced splitting of the heavy hole sub-band at the Fermi level is determined to be around 30% of the total Fermi energy. In addition, Eq. ( 2) withαindependent of wave vector is a result of the Schrieffer-Wolff transformation applied to the Luttinger Hamiltonian, and its use requires the Schrieffer-Wolff method to be applicable. Furthermore, throughout this paper we consider cases where only the HH1 band is occupied. We have calculated the exact window of applicability of our theory in Table I. Physically, the terms ∝α2entering the Hall coefficient aretracedbacktoseveralmechanisms. Firstly, spin-orbit coupling gives rise to corrections to: (i) the occupation probabilities, through fk±; (ii) the band energies and density of states, through dεk±/dk; and (iii) the scatter- ing term, whichincludes intra-andinter-bandscattering, as well as scattering between the charge and spin distri- butions. Secondly, Rashba spin-orbit coupling gives rise to a current-induced spin polarization [ 56], which is of first order in α, and this in turn gives rise to a feedback effect on the charge current, which is thenresponsible for approximately a quarter of the overall spin-orbit contri- bution to the Hall coefficient. As a concrete example, a 2D hole system confined to GaAs/AlGaAs heterostructures is particularly promising since it has not only a very high mobility, but also a spin splitting that hasbeen shownto be electricallytunable in both square and triangular wells [ 66]. The spin splitting can be tuned from large values to nearly zero in a square3500 3000 2500 2000 1500 1000 500 0α (meV nm3) 108 6 4 2 0 Fz (MV/m)360 340 320 300α (meV nm3) 10864 Fz (MV/m) GaAs InAs InSb GaAs Figure 2. The Rashba coefficient αof as a function of the net perpendicular electric field Fzfor 15 nm GaAs, InAs, and InSb quantum wells. The inset shows that αfor GaAs decreases by ∼20% asFzis increased from 4 MV/m to 10 MV/m, due to the fact the well becomes quasi-triangular at Fz/greaterorsimilar4 MV/m. quantumwellwhosechargedistributioncanbecontrolled from being asymmetric to symmetric via the application of a surface-gate bias. Whereas thus far the theoretical formalismhas been general, to makeconcrete experimen- tal predictions we first specialize to a two-dimensional hole gas (2DHG) in a 15 nm-wide GaAs quantum well subjected to an electric field in the ˆ zdirection, so that the symmetry ofthe quantum well can be tuned arbitrar- ily. In the simplest approximation, taking into account only the lowest heavy-hole and light-hole sub-bands, in a 2DHG the Rashba coefficient αmay be estimated as α=3/planckover2pi14 m2 0∆Eγ2∝an}bracketle{tφL|φH∝an}bracketri}ht∝an}bracketle{tφH|(−id/dz)|φL∝an}bracketri}ht.(7) where∆ Eis energysplitting ofthe lowestheavy-holeand light-hole sub-bands and γ=γ2+γ3 2, andφH,L≡φH,L(z) represents the orbitalcomponent of the heavy-hole and light-hole wave functions respectively in the direction ˆz perpendiculartotheinterface. Forasystemwith topand back gates, where the electric field Fzacross the well can be turned on or off, we use a modified infinite square well wave function in which Fzis already encoded [ 67]. TheRashbacoefficient α, asafunctionof Fz, for15nm hole quantum wells is shown Fig. 2. For GaAs, at low Fz (Fz≪4 MV/m), the Rashba coefficient increases with F, whichisinaccordancewiththetrendsreportedinRef. [68]. AsFzis increased, αthen saturates, and, at larger electricfields( Fz>4MV/m), thequantumwellbecomes quasi-triangular and the Rashba coefficient αdecreases with increasing electric field Fz. The decrease of αas a function of Fzin quasi-triangular wells is consistent with the experimental findings of Ref. [ 69]. Note that for different materials, αsaturates at different values of Fz,4 (a) (b)1.00 0.98 0.96 0.94 0.92 0.90 0.88σxx/σ0 10 8 6 4 2 0 Fz(MV/m)p = 1 x 1011cm-2 p = 1.5 x 1011cm-2 p = 2 x 1011cm-2GaAs QW1.00 0.95 0.90 0.85 0.80 0.75 0.70σxx/σ0GaAs InAs InSbp = 2 x 1011cm-2 Figure 3. Ratio of Drude conductivity at finite electric field s to its zero electric field value, with the bare Drude conduc- tivityσ0≡peµ, for (a) different quantum well materials at p= 1×1011cm−2and (b) GaAs quantum wells at different densities. Here, the well width is 15 nm. and that the αis larger in materials with a higher atomic number [ 63]. Giventhe dependence of α(Fig.2), andhence theHall coefficient RH(Fig.1), onFz, we now outline how αcan be deduced experimentally. Using a top- and backgated quantum well, the quantum well is initially tuned to be symmetric so that αwill be zero and the hole density can be measured accurately. One subsequently increases Fz, for example to ∼4 MV/m for the GaAs quantum welldiscussedabove,whilst keepingthe densityconstant. This in turn results in an appreciable increase in α, and hence a large change in RHas a function of Fz. The non-monotonic change in αas a function of Fz likewiseaffectsthelongitudinalconductivity σxx(Fig.3), which reads σxx=σ0/bracketleftbigg 1−/parenleftbigg60πm∗2α2 /planckover2pi14/parenrightbigg p/bracketrightbigg . (8) The spin-orbit corrections are larger in InAs and InSb (Fig.3a) rather than GaAs. Furthermore, as the density increases, σxxdecreases faster with Fz(Fig.3b). How- ever, although the spin-orbit corrections to σxxhave a similar functional form as and a similar magnitude to the corrections to RH, it is difficult to single out the de- pendence of σxxonαexperimentally. As the shape of the wave functions changes with Fz, the spin-orbit in-dependent scattering properties are also altered, which may then introduce a larger correction to σxxthan the spin-orbitinduced corrections[ 70]. In fact, the spin-orbit independent corrections can alter the carrier mobility by ∼20% even in electron quantum wells [ 40]. Various possibilities exist to extend the scope of the calculations presented in this paper. Here we have re- stricted ourselves, for the sake of gaining physical insight and without loss of generality, to hole systems in which the Schrieffer-Wolff approximation is applicable so that αcan be approximated as constant. In a general 2D hole systemα(k) is a function of wave vector, and decreases withkatlargerwavevectors. Itsbehaviourisinprinciple nottractableanalyticallythoughitcanstraightforwardly be calculated numerically. The results we have found re- main true in their general closed form for hole systems at arbitrary densities provided αis replaced by α(k). An alternative approach would be to start directly with the 4×4 Luttinger Hamiltonian and determine the charge conductivity using a spin-3/2 model. However, calculat- ing the conductivity as a function of Fzcan quickly be- comeverycomplicated analytically, limiting the utility of such an approach. Finally, the kinetic equation approach wehavediscussedcanstraightforwardlybe generalizedto arbitrary band structures in a way that makes it suitable for fully numerical approaches relying on maximally lo- calized Wannier functions [ 71]. It is worth mentioning how the corrections in the mag- netotransport properties of 2D electrons will differ from those of 2D holes. In 2D electrons, to lowest order the spin-orbit coupling stems from k.pcoupling with the topmost valence band, and the leading contribution to spin-orbit interaction in 2D electrons is linear in k[16]. As a result, the spin-orbit dependent corrections to the magnetotransport in 2D electrons will be much smaller compared to 2D holes, and thus may not be detectable within experimental resolution. In summary, we have presented a quantum kinetic the- ory of magneto-transport in 2D heavy-hole systems in a weak perpendicular magnetic field and demonstrated that the Hall coefficient, as well as the longitudinal con- ductivity, display strong signatures of the spin-orbit in- teraction. We have also shown that our theory provides an excellent qualitative agreementto existing experimen- tal trends for α, although to the best of our knowledge, there has not been a demonstration of RHchanging as a function of α. An appropriate experimental setup with top and back gates can lead to a direct electrical mea- surement of the Rashba spin-orbit constant via the Hall coefficient. ACKNOWLEDGMENTS This research was supported by the Australian Re- search Council Centre of Excellence in Future Low- Energy Electronics Technologies (project CE170100039) and funded by the Australian Government.5 [1] R. Cuan and L. Diago-Cisneros, Europhys. Lett. 110, 67001 (2015) . [2] T. Biswas, S. Chowdhury, and T. K. Ghosh, Eur. Phys. J. B 88, 220 (2015) . [3] F. A. Zwanenburg, A. S. Dzurak, A. Morello, M. Y. Simmons, L. C. L. Hollenberg, G. Klimeck, S. Rogge, S. N. Coppersmith, and M. A. Eriksson, Rev. Mod. Phys. 85, 961 (2013) . [4] S. Conesa-Boj, A. Li, S. Koelling, M. Brauns, J. Rid- derbos, T. T. Nguyen, M. A. Verheijen, P. M. Koen- raad, F. A. Zwanenburg, and E. P. A. M. Bakkers, Nano Letters 17, 2259 (2017) . [5] M. Brauns, J. Ridderbos, A. Li, W. G. van der Wiel, E. P. A. M. Bakkers, and F. A. Zwanenburg, Appl. Phys. Lett. 109, 143113 (2016) . [6] M. Brauns, J. Ridderbos, A. Li, E. P. A. M. Bakkers, W. G. van der Wiel, and F. A. Zwanenburg, Phys. Rev. B 94, 041411 (2016) . [7] F. Mueller, G. Konstantaras, P. C. Spruijtenburg, W. G. van der Wiel, and F. A. Zwanenburg, Nano Lett. 15, 5336 (2015) . [8] F. Qu, J. van Veen, F. K. de Vries, A. J. A. Beuk- man, M. Wimmer, W. Yi, A. A. Kiselev, B.-M. Nguyen, M. Sokolich, M. J. Manfra, F. Nichele, C. M. Marcus, and L. P. Kouwenhoven, Nano Lett. 16, 7509 (2016) . [9] J.-T. Hung, E. Marcellina, B. Wang, A. R. Hamilton, and D. Culcer, Phys. Rev. B 95, 195316 (2017) . [10] D. V. Bulaev and D. Loss, Phys. Rev. Lett. 95, 076805 (2005) . [11] F. Nichele, M. Kjaergaard, H. J. Suominen, R. Sko- lasinski, M. Wimmer, B.-M. Nguyen, A. A. Kiselev, W. Yi, M. Sokolich, M. J. Manfra, F. Qu, A. J. A. Beukman, L. P. Kouwenhoven, and C. M. Marcus, Phys. Rev. Lett. 118, 016801 (2017) . [12] T. Korn, M. Kugler, M. Griesbeck, R.Schulz, A. Wagner, M. Hirmer, C. Gerl, D. Schuh, W. Wegscheider, and C. Schller, New J. Phys. 12, 043003 (2010) . [13] J. Salfi, J. A. Mol, D. Culcer, and S. Rogge, Phys. Rev. Lett. 116, 246801 (2016) . [14] J. Salfi, M. Tong, S. Rogge, and D. Culcer, Nanotechnology 27, 244001 (2016) . [15] J. R. Petta, A. C. Johnson, J. M. Taylor, E. A. Laird, A. Yacoby, M. D. Lukin, C. M. Marcus, M. P. Hanson, and A. C. Gossard, Science309, 2180 (2005) . [16] R. Winkler, Spin-Orbit Coupling Effects in Two- Dimensional Electron and Hole systems (Springer, Berlin, 2003). [17] R. Moriya, K. Sawano, Y. Hoshi, S. Masubuchi, Y. Shiraki, A. Wild, C. Neumann, G. Abstre- iter, D. Bougeard, T. Koga, and T. Machida, Phys. Rev. Lett. 113, 086601 (2014) . [18] T. Biswas and T. K. Ghosh, J. Appl. Phys. 115, 213701 (2014) . [19] K. V. Shanavas, Phys. Rev. B 93, 045108 (2016) . [20] G. Akhgar, O. Klochan, L. H. Willems van Beveren, M. T. Edmonds, F. Maier, B. J. Spencer, J. C. Mc- Callum, L. Ley, A. R. Hamilton, and C. I. Pakes, Nano Lett. 16, 3768 (2016) . [21] R. M. Lutchyn, J. D. Sau, and S. Das Sarma, Phys. Rev. Lett. 105, 077001 (2010) .[22] J.Alicea, Y.Oreg, G.Refael, F.vonOppen, andM.P.A. Fisher,Nat. Phys. 7, 412 (2011) . [23] S. T. Gill, J. Damasco, D. Car, E. P. A. M. Bakkers, and N. Mason, Appl. Phys. Lett. 109, 233502 (2016) . [24] S. V. Alestin Mawrie and T. K. Ghosh, arXiv:1705.02483v1 (unpublished). [25] M. J. Manfra, L. N. Pfeiffer, K. W. West, R. de Picciotto, and K. W. Baldwin, Appl. Phys. Lett. 86, 162106 (2005) . [26] B. Habib, M. Shayegan, and R. Winkler, Semicond. Sci. Technol. 23, 064002 (2009) . [27] X.-J. Hao, T. Tu, G. Cao, C. Zhou, H.-O. Li, G.- C. Guo, W. Y. Fung, Z. Ji, G.-P. Guo, and W. Lu, Nano Letters 10, 2956 (2010) , pMID: 20698609. [28] S. Chesi, G. F. Giuliani, L. P. Rokhinson, L. N. Pfeiffer, and K. W. West, Phys. Rev. Lett. 106, 236601 (2011) . [29] D. Q. Wang, O. Klochan, J.-T. Hung, D. Culcer, I. Farrer, D. A. Ritchie, and A. R. Hamilton, Nano Lett. 16, 7685 (2016) . [30] A. Srinivasan, D. S. Miserev, K. L. Hudson, O. Klochan, K. Muraki, Y. Hirayama, D. Reuter, A. D. Wieck, O. P. Sushkov, and A. R. Hamilton, Phys. Rev. Lett. 118, 146801 (2017) . [31] J. D. Watson, S. Mondal, G. A. Cs´ athy, M. J. Manfra, E. H. Hwang, S. Das Sarma, L. N. Pfeiffer, and K. W. West,Phys. Rev. B 83, 241305 (2011) . [32] A. Srinivasan, K. L. Hudson, D. Miserev, L. A. Yeoh, O. Klochan, K. Muraki, Y. Hirayama, O. P. Sushkov, and A. R. Hamilton, Phys. Rev. B 94, 041406 (2016) . [33] S. J. Papadakis, E. P. De Poortere, M. Shayegan, and R. Winkler, Phys. Rev. Lett. 84, 5592 (2000) . [34] R. Winkler, H. Noh, E. Tutuc, and M. Shayegan, Phys. Rev. B 65, 155303 (2002) . [35] J. van der Heijden, J. Salfi, J. A. Mol, J. Verduijn, G. C. Tettamanzi, A. R. Hamilton, N. Collaert, and S. Rogge, Nano Letters 14, 1492 (2014) . [36] F. Nichele, A. N. Pal, R. Winkler, C. Gerl, W. Wegscheider, T. Ihn, and K. Ensslin, Phys. Rev. B 89, 081306 (2014) . [37] T. Ota, K. Ono, M. Stopa, T. Hatano, S. Tarucha, H. Z. Song, Y. Nakata, T. Miyazawa, T. Ohshima, and N. Yokoyama, Phys. Rev. Lett. 93, 066801 (2004) . [38] K. Ono, D. G. Austing, Y. Tokura, and S. Tarucha, Science297, 1313 (2002) . [39] R. Hanson, L. P. Kouwenhoven, J. R. Petta, S. Tarucha, and L. M. K. Vandersypen, Rev. Mod. Phys. 79, 1217 (2007) . [40] A. F. Croxall, B. Zheng, F. Sfigakis, K. Das Gupta, I. Farrer, C. A. Nicoll, H. E. Beere, and D. A. Ritchie, Appl. Phys. Lett. 102, 082105 (2013) . [41] H. Watzinger, C. Kloeffel, L. Vukui, M. D. Rossell, V. Sessi, J. Kukuka, R. Kirchschlager, E. Lausecker, A. Truhlar, M. Glaser, A. Rastelli, A. Fuhrer, D. Loss, and G. Katsaros, Nano Letters 16, 6879 (2016) . [42] C. Gerl, S. Schmultz, H.-P. Tranitz, C. Mitzkus, and W. Wegscheider, Appl. Phys. Lett. 86, 252105 (2005) . [43] W. R. Clarke, C. E. Yasin, A. R. Hamilton, A. P. Micol- ich, M. Y. Simmons, K. Muraki, Y. Hirayama, M. Pep- per, and D. A. Ritchie, Nat. Phys. 4, 55 (2007) . [44] I. ˇZuti´ c, J. Fabian, and S. Das Sarma, Rev. Mod. Phys. 76, 323 (2004) .6 [45] D. D. Awschalom and M. E. Flatte, Nat. Phys. 3, 153159 (2007) . [46] A. Sasaki, S. Nonaka, Y. Kunihashi, M. Kohda, T. Bauernfeind, T. Dollinger, K. Richter, and J. Nitta, Nat. Nanotechnol. 9, 703 (2014) . [47] R. L. Kallaher, J. J. Heremans, N. Goel, S. J. Chung, and M. B. Santos, Phys. Rev. B 81, 075303 (2010) . [48] B. Grbi´ c, R. Leturcq, T. Ihn, K. Ensslin, D. Reuter, and A. D. Wieck, Phys. Rev. B 77, 125312 (2008) . [49] H. Nakamura, T. Koga, and T. Kimura, Phys. Rev. Lett. 108, 206601 (2012) . [50] T. Koga, J. Nitta, T. Akazaki, and H. Takayanagi, Phys. Rev. Lett. 89, 046801 (2002) . [51] J. Nitta, T. Akazaki, H. Takayanagi, and T. Enoki, Phys. Rev. Lett. 78, 1335 (1997) . [52] Y. H. Park, H. jun Kim, J. Chang, S. H. Han, J. Eom, H.-J. Choi, and H. C. Koo, Appl. Phys. Lett. 103, 252407 (2013) . [53] T. Li, L. A. Yeoh, A. Srinivasan, O. Klochan, D. A. Ritchie, M. Y. Simmons, O. P. Sushkov, and A. R. Hamilton, Phys. Rev. B 93, 205424 (2016) . [54] P. S. Eldridge, W. J. H. Leyland, P. G. Lagoudakis, O. Z. Karimov, M. Henini, D. Taylor, R. T. Phillips, and R. T. Harley,AIP Conference Proceedings 1199, 395 (2010) . [55] G. Wang, B. L. Liu, A. Balocchi, P. Renucci, C. R. Zhu, T. Amand, C. Fontaine, and X. Marie, Nat. Commun. 4, 2372 (2013) . [56] C.-X. Liu, B. Zhou, S.-Q. Shen, and B.-f. Zhu, Phys. Rev. B 77, 125345 (2008) . [57] I. V. Tokatly, E. E. Krasovskii, and G. Vignale, Phys. Rev. B 91, 035403 (2015) .[58] C. H. Li, O. M. J. van‘t Erve, S. Rajput, L. Li, and B. T. Jonker,Nat. Commun. 7, 13518EP (2016) . [59] A. Wong and F. Mireles, Phys. Rev. B 81, 085304 (2010) . [60] E. B. Sonin, Phys. Rev. B 81, 113304 (2010) . [61] K. Nomura, J. Wunderlich, J. Sinova, B. Kaest- ner, A. H. MacDonald, and T. Jungwirth, Phys. Rev. B 72, 245330 (2005) . [62] T. Jungwirth, J. Wunderlich, and K. Olejnik, Nat. Mater. 11, 382 (2012) . [63] E. Marcellina, A. R. Hamilton, R. Winkler, and D. Cul- cer,Phys. Rev. B 95, 075305 (2017) . [64] R. Winkler, Phys. Rev. B 62, 4245 (2000) . [65] F. T. Vasko and O. E. Raichev, Quantum Kinetic Theory and Applications (Springer, New York, 2005). [66] J. P. Lu, J. B. Yau, S. P. Shukla, M. Shayegan, L. Wissinger, U. R¨ ossler, and R. Winkler, Phys. Rev. Lett. 81, 1282 (1998) . [67] G. Bastard, E. E. Mendez, L. L. Chang, and L. Esaki, Phys. Rev. B 28, 3241 (1983) . [68] S. J. Papadakis, E. P. De Poortere, H. C. Manoharan, M. Shayegan, and R. Winkler, Science 283, 2056 (1999). [69] B. Habib, E. Tutuc, S. Melinte, M. Shayegan, D. Wasserman, S. A. Lyon, and R. Winkler, Appl. Phys. Lett. 85, 3151 (2004) . [70] M. Y. Simmons, A. R. Hamilton, S. J. Stevens, D. A. Ritchie, M. Pepper, and A. Kurobe, Appl. Phys. Lett. 70, 2750 (1997) . [71] D. Culcer, A. Sekine, and A. H. MacDonald, Phys. Rev. B 96, 035106 (2017) .arXiv:1708.07247v2 [cond-mat.mes-hall] 25 Aug 2017Supplement of “Strong influence of spin-orbit coupling on ma gnetotransport in two-dimensional hole systems” Hong Liu, E. Marcellina, A. R. Hamilton and Dimitrie Culcer1 1School of Physics and Australian Research Council Centre of Excellence in Low-Energy Electronics Technologies, UNSW Node, The University of New South Wales, Sydney 2052, Au stralia I. LUTTINGER HAMILTONIAN We start from the bulk 4 ×4 Luttinger Hamiltonian [ 1]HL(k2,kz) describing holes in the uppermost valence band with an effective spin J= 3/2. So the hole system with top and back gate in z-direction can be simplified as the isotropic Luttinger-Kohn Hamiltonian plus a confining asymmetrical t riangular potential. ˆH=HL(k2,kz)−eFzz, z > 0, (1) whereFzis the gate electric field and Fz≥0. The 4 ×4 Luttinger Hamiltonian, which is expressed in the basis of Jz eigenstates {|+3 2∝an}bracketri}ht,|−3 2∝an}bracketri}ht,|+1 2∝an}bracketri}ht,|−1 2∝an}bracketri}ht}, reads HL(k2,kz)= P+Q0L M 0P+Q M∗−L∗ L∗M P−Q0 M∗−L0P−Q , (2) where P=µ 2γ1(k2+k2 z), Q=−µ 2γ2(2k2 z−k2), L=−√ 3µγ3k−kz, M=−√ 3µ 2(γk2 −+δk2 +).(3) withµ=/planckover2pi12 m0,γ1,γ2,γ3are the Luttinger parameters (Table I),γ=γ2+γ3 2,δ=γ2−γ3 2, andk2=/radicalig k2x+k2y,k±= kx±ikyandθ= arctanky kx. To obtain the spectrum of our system, we use modified infinite squa re well wave functions [3] for the heavy hole (HH) and light hole (LH) states φv=sin/bracketleftbigπ d/parenleftbig z+d 2/parenrightbig/bracketrightbig exp/bracketleftbig −βv/parenleftbigz d+1 2/parenrightbig/bracketrightbig π/radicalig e−βvdsinh(βv) 2π2βv+2β3v, (4) wherev=h,ldenote the HH and LH states and dis the width of the quantum well. The eigenvalues of the heavy hole and light hole as well as the corresponding kdependent expansion coefficients are then obtained by diagonalizing the matrix ˜H, whose elements are given as ˜H=∝an}bracketle{tν|HL(k2,ˆkz)+V(z)|ν′∝an}bracketri}ht, (5) where|ν∝an}bracketri}htdenotes the wave function Eq. ( 4) andˆkzstands for the operator −i∂ ∂z. The two lowest eigenenergies of the 4×4 matrix Eq. ( 5) correspond to the dispersion of the spin-split HH1 ±subbands. Usually, only the lowest HH-subspace is taken into account at low hole densities. Accordingly , we perform a Schrieffer-Wolff transformation on Eq.5to restrict our attention to the lowest HH subspace. Therefore, the effective Hamiltonian describing the two dimensional hole gas is [ 4] H0k=/planckover2pi12k2 2m∗+iα(k3 −σ+−k3 +σ+), (6) Table I. Luttinger parameters used in this work [ 2]. GaAs InAs InSb γ1 6.85 20.40 37.10 γ2 2.10 8.30 16.50 γ3 2.90 9.10 17.702 wherem∗≡m∗ hh=m0 γ1+γ2and the Pauli matrix σ±=1 2(σx±iσy). The eigenvalues of Eq. ( 6) areεk,±=ǫ0±αk3, whereǫ0=/planckover2pi12k2 2m∗. The Rashba coefficient αis expressed as α=3µ2 ∆Eγ2∝an}bracketle{tφL|φH∝an}bracketri}ht∝an}bracketle{tφH|kz|φL∝an}bracketri}ht. (7) where ∆ Eis energy splitting of heavy hole and light hole. II. SCATTERING TERM Thek-diagonal part of density matrix fkis a 2×2 Hermitian matrix, which is decomposed into fk=nk11+Sk, wherenkrepresents the scalar part and 11 is the identity matrix into two dimensions. The component Skis written purely in terms of the Pauli σmatrices Sk=1 2Sk·σ≡1 2Skiσi. With this notation, the scattering term is in turn decomposed into ˆJ(fk) =ni /planckover2pi12/integraldisplayd2k′ (2π)2|Ukk′|2(nk−nk′) lim η→0/integraldisplay∞ 0dt′e−ηt′e−iH0k′t′//planckover2pi1eiH0kt′//planckover2pi1+H.c. +ni 2/planckover2pi12/integraldisplayd2k′ (2π)2|Ukk′|2(Sk−Sk′)·lim η→0/integraldisplay∞ 0dt′e−ηt′e−iH0k′t′//planckover2pi1σeiH0kt′//planckover2pi1+H.c..(8) We use perturbation theory solving the kinetic equation up to α2. In the process, we decompose the matrix Sk= Sk/bardbl+Sk⊥and write those two parts as Sk/bardbl= (1/2)sk/bardblσk/bardblandSk⊥= (1/2)sk⊥σk⊥. The terms sk/bardblandsk⊥are scalars and given by sk/bardbl=Sk·ˆΩkandsk⊥=Sk·ˆΘkwithˆΩk=−sin3θˆx+cos3θˆyandˆΘk=−cos3θˆx−sin3θˆy. Withγ=θ′−θ, the scattering term becomes ˆJ(n) =πni 2/planckover2pi1/integraldisplayd2k′ (2π)2|Ukk′|2(nk−nk′)·(1+ˆΩk′·ˆΩk)/bracketleftig δ(ǫ+−ǫ′ +)+δ(ǫ−−ǫ′ −)/bracketrightig +πni 2/planckover2pi1/integraldisplayd2k′ (2π)2|Ukk′|2(nk−nk′)·σ·(ˆΩk′+ˆΩk)/bracketleftig δ(ǫ′ +−ǫ+)−δ(ǫ′ −−ǫ−)/bracketrightig +πni 2/planckover2pi1/integraldisplayd2k′ (2π)2|Ukk′|2(nk−nk′)·(1−ˆΩk′·ˆΩk)/bracketleftig δ(ǫ+−ǫ′ −)+δ(ǫ−−ǫ′ +)/bracketrightig +πni 2/planckover2pi1/integraldisplayd2k′ (2π)2|Ukk′|2(nk−nk′)·σ·/bracketleftig (ˆΩk−ˆΩk′)/bracketrightig/bracketleftig δ(ǫ′ −−ǫ+)−δ(ǫ′ +−ǫ−)/bracketrightig ,(9) and ˆJ(S) =πni 4/planckover2pi1/integraldisplayd2k′ (2π)2|Ukk′|2(Sk−Sk′)·/bracketleftig σ(1−ˆΩk·ˆΩk′)+(ˆΩk·σ)ˆΩk′+ˆΩk(ˆΩk′·σ)/bracketrightig/bracketleftig δ(ǫ+−ǫ′ +)+δ(ǫ−−ǫ′ −)/bracketrightig +πni 4/planckover2pi1/integraldisplayd2k′ (2π)2|Ukk′|2(Sk−Sk′)·(ˆΩk+ˆΩk′)/bracketleftig δ(ǫ′ +−ǫ+)−δ(ǫ′ −−ǫ−)/bracketrightig +πni 4/planckover2pi1/integraldisplayd2k′ (2π)2|Ukk′|2(Sk−Sk′)·/bracketleftig σ(1+ˆΩk·ˆΩk′)−(ˆΩk·σ)ˆΩk′−ˆΩk(ˆΩk′·σ)/bracketrightig/bracketleftig δ(ǫ+−ǫ′ −)+δ(ǫ−−ǫ′ +)/bracketrightig +πni 4/planckover2pi1/integraldisplayd2k′ (2π)2|Ukk′|2(Sk−Sk′)·/bracketleftig (ˆΩk−ˆΩk′)/bracketrightig/bracketleftig δ(ǫ+−ǫ′ −)−δ(ǫ−−ǫ′ +)/bracketrightig . (10) We now separate these terms according to the contributions from intra-band and inter-band scatterings ˆJ(n) =πni 2/planckover2pi1/integraldisplayd2k′ (2π)2|Ukk′|2(nk−nk′)(1+cos3 γ)/bracketleftig δ(ǫ+−ǫ′ +)+δ(ǫ−−ǫ′ −)/bracketrightig +πni 2/planckover2pi1/integraldisplayd2k′ (2π)2|Ukk′|2(nk−nk′)(1−cos3γ)/bracketleftig δ(ǫ+−ǫ′ −)+δ(ǫ−−ǫ′ +)/bracketrightig ,(11)3 ˆJ(S) =πni 4/planckover2pi1/integraldisplayd2k′ (2π)2|Ukk′|2/bracketleftig (sk/bardbl−sk′/bardbl)(1+cos3 γ)σk/bardbl+(sk/bardbl−sk′/bardbl)sin3γσk⊥ +(sk⊥+sk′⊥)σk/bardblsin3γ+(sk⊥+sk′⊥)(1−cos3γ)σk⊥/bracketrightig/bracketleftig δ(ǫ+−ǫ′ +)+δ(ǫ−−ǫ′ −)/bracketrightig +πni 4/planckover2pi1/integraldisplayd2k′ (2π)2|Ukk′|2/bracketleftig (sk/bardbl+sk′/bardbl)(1−cos3γ)σk/bardbl−(sk/bardbl+sk′/bardbl)sin3γσk⊥ −(sk⊥−sk′⊥)σk/bardblsin3γ+(sk⊥−sk′⊥)(1−cos3γ)σk⊥/bracketrightig/bracketleftig δ(ǫ+−ǫ′ −)+δ(ǫ−−ǫ′ +)/bracketrightig ,(12) and ˆJS→n(S) =πni 4/planckover2pi1/integraldisplayd2k′ (2π)2|Ukk′|2/bracketleftig (sk/bardbl−sk′/bardbl)(1+cos3 γ)+(sk⊥+sk′⊥)sin3γ/bracketrightig/bracketleftig δ(ǫ′ +−ǫ+)−δ(ǫ′ −−ǫ−)/bracketrightig =πni 4/planckover2pi1/integraldisplayd2k′ (2π)2|Ukk′|2/bracketleftig (sk/bardbl+sk′/bardbl)(1−cos3γ)−(sk⊥−sk′⊥)sin3γ/bracketrightig/bracketleftig δ(ǫ+−ǫ′ −)−δ(ǫ−−ǫ′ +)/bracketrightig ,(13) ˆJn→S(n) =πni 2/planckover2pi1/integraldisplayd2k′ (2π)2|Ukk′|2(nk−nk′)/bracketleftig σk/bardbl(1+cos3 γ)+σk⊥sin3γ/bracketrightig/bracketleftig δ(ǫ′ +−ǫ+)−δ(ǫ′ −−ǫ−)/bracketrightig +πni 2/planckover2pi1/integraldisplayd2k′ (2π)2|Ukk′|2(nk−nk′)/bracketleftig σk/bardbl(1−cos3γ)−σk⊥sin3γ/bracketrightig/bracketleftig δ(ǫ′ −−ǫ+)−δ(ǫ′ +−ǫ−)/bracketrightig .(14) We next decompose the kinetic equation as follows: dnk dt+ˆJn→n(nk) =Dkn, dSk/bardbl dt+P/bardblˆJS→S(Sk/bardbl) =Dk/bardbl, dSk⊥ dt+i /planckover2pi1/bracketleftbig H0k,Sk⊥/bracketrightbig =Dk⊥.(15) Note that the projection operator P/bardblabove acts on a matrix Mas Tr(Mσk/bardbl), where Tr refers to the matrix (spin) trace. III. SOLUTION FOR THE LONGITUDINAL CONDUCTIVITY Here we derive the longitudinal conductivity at zero magnetic field. E xpanding the δfunctions in Sec. IIup to ∝α2, we get the following δ(ǫ+−ǫ′ +)≈δ(ǫ0−ǫ′ 0)+α(k3−k′3)∂ ∂ǫ0δ(ǫ0−ǫ′ 0)+α2(k3−k′3)2 2∂2δ(ǫ0−ǫ′ 0) ∂ǫ2 0 δ(ǫ−−ǫ′ −)≈δ(ǫ0−ǫ′ 0)−α(k3−k′3)∂ ∂ǫ0δ(ǫ0−ǫ′ 0)+α2(k3−k′3)2 2∂2δ(ǫ0−ǫ′ 0) ∂ǫ2 0 δ(ǫ+−ǫ′ −)≈δ(ǫ0−ǫ′ 0)+α(k3+k′3)∂ ∂ǫ0δ(ǫ0−ǫ′ 0)+α2(k3+k′3)2 2∂2δ(ǫ0−ǫ′ 0) ∂ǫ2 0 δ(ǫ−−ǫ′ +)≈δ(ǫ0−ǫ′ 0)−α(k3+k′3)∂ ∂ǫ0δ(ǫ0−ǫ′ 0)+α2(k3+k′3)2 2∂2δ(ǫ0−ǫ′ 0) ∂ǫ2 0.(16) We now insert Eq. ( 16) into the electric driving term DE,kand scattering term ˆJ(fk). With ρ0k=f0++f0− 211 + f0+−f0− 2σk/bardblandf0+,f0−equilibrium Fermi distribution function, the driving term DE,kbecomes, DE,kn=−eE·ˆk 2/planckover2pi1(∂f0+ ∂k+∂f0− ∂k)≈eE·ˆk 2/planckover2pi1/bracketleftig 2/planckover2pi12k m∗δ(ǫ0−ǫF)+6α2k5∂δ(ǫ0−ǫF) ∂ǫ0/bracketrightig , DE,k/bardbl=−eE·ˆk 2/planckover2pi1(∂f0+ ∂k−∂f0− ∂k)σk/bardbl≈eE·ˆk 2/planckover2pi1/bracketleftig 6αk2δ(ǫ0−ǫF)+2/planckover2pi12k m∗αk3∂δ(ǫ0−ǫF) ∂ǫ0)/bracketrightig .(17)4 Solving Eqs. ( 15), we obtain the density matrices n(0) Ek=τpeE·ˆk /planckover2pi1/bracketleftig/planckover2pi12k m∗δ(ǫ0−ǫF)/bracketrightig , (18a) S(1) Ek/bardbl=τsαeE·ˆk /planckover2pi1/bracketleftig/planckover2pi12k4 m∗∂δ(ǫ0−ǫF) ∂ǫ0+3k2δ(ǫ0−ǫF)/bracketrightig σk/bardbl=s(1) Ek/bardblσk/bardbl, (18b) n(2) Ek=τpα2/braceleftigeE·ˆk /planckover2pi1/bracketleftig 3k5∂δ(ǫ0−ǫF) ∂ǫ0/bracketrightig −3km∗2ni 4απ/planckover2pi15s(1) Ek/bardblζ(γ)−n(0) Ek6nim∗3 π/planckover2pi17k2ξ(γ)/bracerightig . (18c) whereǫF=/planckover2pi12k2 F 2m∗,τp=2π/planckover2pi13 m∗niξ(γ),τs=4π/planckover2pi13 m∗niβ(γ), and ζ(γ)=/integraldisplay dγ|Ukk′|2(cosγ−cos3γ), ξ(γ)=/integraldisplay dγ|Ukk′|2(1−cosγ), β(γ)=/integraldisplay dγ|Ukk′|2(1−cosγcos3γ).(19) In the low temperature limit, the Thomas-Fermi wave-vector of a t wo-dimensional hole gas without spin-orbit coupling is kTF=2 aB, withaB=/planckover2pi12ǫr m∗e2. The screened Coulomb potential between plane waves is given by |Ukk′|2=Z2e4 4ǫ2 0ǫ2r/parenleftbigg1 |k−k′|+kTF/parenrightbigg2 . (20) With Eq. ( 20), we obtainζ(γ) ξ(γ)≈2 andξ(γ) β(γ)=1 3. Using the velocity operator ˆvx=/planckover2pi1kx m∗+α /planckover2pi13k2[−sin2θσx+cos2θσy],ˆvy=/planckover2pi1ky m∗+α /planckover2pi13k2[−sin2θσy−cos2θσx], (21) the longitudinal current is jx=eTr/bracketleftbig ˆvxρEk/bracketrightbig , where ρEk= (n(0) Ek+n(2) Ek)11 +S(1) Ek/bardbl. Therefore, the longitudinal conductivity with Rashba spin orbit coupling up to second order in αis σxx=e2τp 2πm∗k2 F/bracketleftig 1−15 2/parenleftbiggαk3 F ǫkin/parenrightbigg2/bracketrightig , (22) whereǫkin=/planckover2pi12k2 F 2m∗. IV. SOLUTION FOR THE HALL COEFFICIENT Now we consider the case of Bz>0. Firstly, we find that the Zeeman terms have no contribution to th e Hall coefficient. With Eqs. ( 21), the Lorentz driving term DL,kbecomes DL,k=1 2e /planckover2pi1/braceleftig ˆv×B,∂ρEk ∂k/bracerightig =1 2eBz /planckover2pi1/braceleftig/braceleftbig ˆvy,∂ρEk ∂kx/bracerightbig −/braceleftbig ˆvx,∂ρEk ∂ky/bracerightbig/bracerightig . (23) We separate DL,kinto the scalar and spin parts with DL,k=DL,n+DL,S, and, switching from the rectangular coordinates to polar coordinates with∂D ∂kx=∂D ∂kcosθ−∂D ∂θsinθ k;∂D ∂ky=∂D ∂ksinθ+∂D ∂θcosθ k, we obtain DL,n=−eBz m∗/bracketleftbig n(0) k+n(2) k/bracketrightbig (−sinθ)+eBz /planckover2pi13αk /planckover2pi1s(1) k,/bardbl(−sinθ), DL,S/bardbl=−/braceleftigeBz m∗s(1) k,/bardbl(−sinθ)+eBz /planckover2pi13αk /planckover2pi1/bracketleftbig n(0) k+n(2) k/bracketrightbig (−sinθ)/bracerightig σk/bardbl, DL,S⊥= cosθeBz /planckover2pi13αk2 /planckover2pi1∂/bracketleftbig n(0) k+n(2) k/bracketrightbig ∂kσk⊥,(24)5 withn(0) Ek=n(0) kcosθ,n(2) Ek=n(2) kcosθands(1) Ek,/bardbl=s(1) k,/bardblcosθ. Solving Eqs. ( 15), we obtain the following density matrices in presence both electric and magnetic fields nBz,k=−sinθτpeBz/braceleftign(0) k+n(2) k m∗+3αk /planckover2pi12s(1) k,/bardbl/bracerightig , SBz,k/bardbl=−sinθτseBz/braceleftigs(1) k,/bardbl m∗+3αk /planckover2pi12/bracketleftbig n(0) k+n(2) k/bracketrightbig/bracerightig σk/bardbl, SBz,k⊥= cosθ3eBz 2/planckover2pi1k∂/bracketleftbig n(0) k+n(2) k/bracketrightbig ∂kσz.(25) The Hall current is jy=eTr/bracketleftbig ˆvyρEB k/bracketrightbig , whereρEBz k=nBz,k11 +SBz,k/bardbl+SBz,k⊥. The Hall coefficient, up to the second order in α, is thus given as RH=σxy Bz(σ2xx+σ2xy)≈1 pe/bracketleftig 1+8/parenleftbiggαk3 F ǫkin/parenrightbigg2/bracketrightig , (26) whereωc=eBz m∗. [1] J. M. Luttinger, Phys. Rev. 102, 1030 (1956) . [2] R. Winkler, Spin-Orbit Coupling Effects in Two-Dimensional Electron and Hole systems (Springer, Berlin, 2003). [3] G. Bastard, E. E. Mendez, L. L. Chang, and L. Esaki, Phys. Rev. B 28, 3241 (1983) . [4] R. Winkler, Phys. Rev. B 62, 4245 (2000) .
2304.12632v1.Magnetization_Switching_in_van_der_Waals_Systems_by_Spin_Orbit_Torque.pdf
1 Magnetization Switching in van der Waals Systems by Spin-Orbit Torque Xin Lin1,2, Lijun Zhu1,2* 1. State Key Laboratory of Superlattices and Microstructures, Institute of Semiconductors, Chinese Academy of Sciences, Beijing 100083, China 2. College of Materials Science and Opto -Electronic Technology, University of Chinese Academy of Sciences, Beijing 100049, China *ljzhu@semi.ac.cn Abstract : Electrical switching of magnetization via spin -orbit torque (SOT) is of great potential in fast, dense, energy - efficient nonvolatile magnetic memory and logic technologies. Recently, enormous efforts have been stimulated to investigate switching of perpendicular magnetization in van der Waals systems that have unique, strong tunability and spin-orbit coupling effect compared to conventional metals. In this review , we first give a brief, generalized introduction to the spin -orbit torque and van der Waals materials. We will then discuss the recent advances in magnetization switching by the spin current generated from van der Waals materials and summary the progress in the switching of Van der Waals magnetization by the spin current . 1. Introduction 1.1 Spin -orbit torque Spin-orbit torque s (SOT s) are a powerful tool to manipulat e magnetization at the nanoscale for spintronic devices, such as magnetic random access memory (MRAM) and logic [1-5]. SOTs are exerted on a magnetization when angular momentum is transferred from spin accumulation or spin currents carried by a flow of electrons or magnons (Fig. 1 ). A spin current with spin polarization vector σ, can exert two types of SOTs on a magnetization M, i.e., a damping -like (DL) torque [ τDL ~ M × (M × σ)] due to the absorption of the spin current component transverse to M and a field-like (FL) torque [ τFL ~ M × σ] due to the reflection of the spin current with some spin rotation . In the simple case of the spin-current generator /magnet bilayer, the efficienc y of the damping -like SOT per unit bias current density, 𝜉DL𝑗 , can be estimated as [6] 𝜉DL𝑗 ≈ Tint θSH τM-1/(τM-1+τso-1) (1) where τM-1/(τM-1+τso-1) is the percentage of the spin current relaxed via the spin-magnetization exchange interaction (with spin relaxation rate τM-1) within the magnetic layer and is less than 1 in presence of non - negligible spin relaxation via the spin-orbit scattering (with spin relaxation rate τM-1) [6], Tint is the interfacial spin transparency which determines what fraction of the spin current enters the magnet (less than 1 in presence of spin backflow [7-11] and spin memory loss [12-15]), and θSH is the charge -to-spin conversion efficiency of the spin current generator (e.g., the spin Hall ratio in the case of spin Hall materials [2-4]). The quantitative understanding of the efficiency of the field -like torque , 𝜉FL𝑗, remains an open question. The same SOT physics can be expressed in terms of effective SOT fields: a damping -like effective SOT field (HDL) parallel to M × σ and a field -like effective SOT field (HFL) parallel to σ. The magnitudes of the damping -like and field -like SOT fields correlate to their SOT efficiencies per unit bias current density via HDL = (ℏ/2e) j𝜉DL𝑗Ms-1t -1 (2) HFL = (ℏ/2e) j𝜉FL𝑗Ms-1t -1 (3) where e is the elementary charge, ℏ reduced Plank’s constant, t the magnetic layer thickness, Ms the saturation magnetization of the magnetic layer, and j the charge current density in the spin current generating layer. The damping -like SOT is technologically more import ant because it can excite dynamics and switching of magnetization (even for low currents for which HDL is much less than the anisotropy field of a perpendicular magnetization). Field -like SOTs by themselves can destabilize magnets only if HFL is greater th an the anisotropy field, but they can still strongly affect the dynamics in combination with a damping -like SOT [16-18]. The g eneration of spin currents is central to the SOT phenomena. The spin polarization vector σ can have longitudinal, transverse, and perpendicular components, i.e., σx, σy, and σz. Transversely polarized spin current can be generated by a longitudinal charge current flow in either magnetic or non -magnetic materials via a variety of possible spin -orbit -coupling (SOC) effects. The latter includes the bulk spin Hall effect (SHE) [19-22], topological surface states [23,24], interfacial SOC effects [16,25-28], orbit -spin conversion [29], the anomalous Hall effect [30,31], the planar Hall effect [32,33], the magnetic SHE [34-36], Dresselhaus effect [37], Dirac nodal lines [38,39], etc. The bulk S HE has been widely observed in thin-film heavy metal (HM) [40-42], Bi-Sb [43], Bi xTe1-x [44], CoPt [45], FePt [46], Fe xTb1-x [47], and Co -Ni-B [48]. Fig. 1 Schematic of damping -like and field -like spin-orbit torque s exerted on a magnetic layer by an incident spin current . 2 Since the transverse spins, in principle, cannot switch a uniform perpendicular magnetization without the assistance of an in -plane longitudinal magnetic field either via coherent rotation of macros pin or via domain wall depinning [2-4,49], generation of perpendicular and longitudinal spins [50-53] are of great interests. While perpendi cular and longitudinal spins are not allowed in nonmagnetic materials that are cubic crystal or polycrystalline/amorphous, additional crystal or magnetic symmetry breaking can be introduced to make perpendicular and longitudinal spins permissive. Generat ion of perpendicular spins has been argued from low-symmetry crystals (e.g., WTe 2 [50], MoTe 2 [54,55], and CuPt [56]), non -collinear antiferromagnetic crystals with magnetic asymmetry (e.g. , IrMn 3 [57], Mn 3GaN [58,59], Mn 3Sn [34]), some collinear antiferromagnets with spin conversions (e.g. , Mn 2Au [60,61] and RuO 2 [62-64]), and also some magnetic interfaces [65]. Longitudinal spins might be generated by low-symmetry crystals ( e.g., MoTe 2 [54], (Ga, Mn)As [66], NiMnSb [67], and Fe/GaAs [68]) or by a non -zero perpendicular magnetization [69,70]. 1.2 van der Waals materials So far, the most widely used spin -source materials are heavy metals with strong spin Hall effect (e.g., Pt with a giant spin Hall conductivity of 1.6×106 (ℏ/2e) Ω-1 m-1 in the clean limit) [71], while the 3 d ferromagnets (e.g., Co, Fe, Ni81Fe19, CoFeB ) and ferrimagnets (e.g., FeTb, CoTb , and GdFeCo) are well studied a s the spin -detect ors. However, it is believed that the energy efficiency of SOT devices may be improved by developing new materials and new mechanisms that generate spin currents. Van der Waals materials have attracted enormous attention in the field of material science and condensed matter physics since the discovery of single monolayer graphene in ref. [72] and is increasingly investigated in the field of spintronics . This is particularly due to the diversity of materials, the flexibility of preparation (e.g., by mechanical exfoliation), and strong tunability by interface effects. In the field of spintronics, the van der Waals materials are also interesting for intriguing SOC effects. For example, the non-magnetic van der Waals materials of transition metal dichalcogenides (TMDs) and topological insulators (TIs) [73,74] exhibit a strong ability in generating transversely polarized spin current via the spin Hall effect and/or topological surface states [75-77], in some cases also in generating currents of perpendicular and longitudinal spins due to low crystal symmetry [50,51]. On other hand, van der Waals materials that are magnetic are also interesting for spin-orbitronics due to their highly tunable magneti sm and low magnetization at small thicknesses [78-83]. As indicated by Eq. (2) and Eq. (3), the damping -like and field -like effective SOT fields exerted on the magnetic layer by a given spin current scale inversely with the thickness and magnetization of the magnetic layer . This review is intended to focus on highlighting t he recent advances of magnetization switching in van der Waals systems by spin -orbit torque , including switching of conventional magnetic metals by spin current from non - magnetic van der Waals TMDs and TIs and switching of van der Waals magnets by incident spin currents . 2. Magnetization s witching by spin current from Transition Metal Dichalcogenide semimetals TMDs typically consist of transition metals (e.g., Mo, W, Pt, Ta, Zr) and chalcogenide elements (e.g., S, Te, Se). From point of view of SOT applications, it is most interesting to develop the TMD semi -metals with relatively high conductivity, strong spin -orbit coupling, high θSH, and reduced structural symmetry. However, TMD semiconductors (e.g., MoS 2 [84,85], WS 2 [86,87], WSe 2 [85]) are merely studied in spin -orbitronics due to their very poor conductivity that is detrimental to the energy TMD metal s (e.g., TaS 2 [88], NbSe 2 [89]) are highly conductive but typically not efficien t in generating spin currents. Therefore, b elow we mainly discuss the progress in spin-orbit torque stud ies of TMD semimetals that are interesting for SOT applications (e.g., WTe 2 [50,90,91], PtTe 2 [92], MoTe 2 [51,93], and ZrTe 2 [94]). 2.1 Exfoliated Transition Metal Dichalcogenide semimetals The pioneering s tudies of TMD semimetal s in the field of spintronics [50] mainly focused on the characterization of damping -like and field -like SOTs of transvers e, perpendicular, and longitudinal spins generated in bilayers of mechanically exfoliated TMD semimetal s and 3d ferromagnets (e.g., Ni81Fe19). These studies have opened a new subject field that unitizes van der Waals materials for the possible generation of transverse, perpendicular, and longitudinal spin polarizations. WTe 2 is a semimetal with a low inversion symmetry along the a axis of the crystal (the space group Pmn21) [Fig. 2(a)] and the Weyl points at the crossing of the oblique conduction and the valence bands only at low temperatures (typically below 100 K [95]). In a WTe 2/ferromagnet bilayer, the screw -axis and glide -plane symmetries of this space group are broken at the interface, so that WTe 2/ferromagnet bilayers have only one symmetry, a mirror symmetry relative to the bc plane (depicted in Fig. 2(a)). There is no mirror symmetry in the ac plane, and therefore no 180◦ rotational symmetry about the c axis (perpendicular to the sample plane). MacNeill et al. [50] first observe d in mechanically exfoliated WTe 2/Ni81Fe19 bilayers damping -like (≈ 8×103 (ℏ/2e) Ω-1 m-1) and field - like spin -orbit torques of transverse spins at room temperature as well as t he exotic damping -like SOT of perpendicular spins (≈ 3.6×103 (ℏ/2e) Ω-1 m-1). The damping -like SOT of perpendicular spins manifests as an additional sin2 𝜑 term in the antisymmetric signal of spin - torque ferromagnetic resonance ( ST-FMR )[Fig. 2(b)] and was attributed to the symmetry breaking at the interface of the WTe 2 crystal. The damping -like SOT of perpendicular spins is found to maximize when current is applied along the low-crystal -symmetry a axis and vanishes when current is applied along the high-crystal -symmetry b axis [50]. This is in contrast to the torques of the transverse spins that are independent of the c rystal orientation. The damping - like SOT of perpendicular spins in WTe 2/Ni 81Fe19 was also found to vary little with the WTe 2 thickness, which was suggest ed as an indication that the spin current is mainly generated near the interface of the WTe 2 [90,91,96]. Xie et al. [52] reported that in -plane direct current along the a axis of WTe 2 can induce partial switching of magnetization in absence of an external magnetic field [Fig. 2(c)] and shift 3 of the anomalous Hall resistance loop in SrRuO 3/exfoliated WTe 2 bilayers , which was speculated as an indication of damping -like torque of perpendicular spins on the perpendicular magnetization (macrospin ). However, the re can be longitudinal and perpendicular Oersted field s due to current spreading in the WTe 2 layer [90,91,96], which can also induce “field -free” switching of perpendicular magnetization and anomalous Hall loop shifts via adding to or subtracting from the domain wall depinning field (coercivity) . The SOTs of the WTe 2/Ni 81Fe19 have also been reported to switch the Ni81Fe19 layer with weak in - plane magnetic anisotropy at a current density of ≈ 3×105 A cm-2 [91]. An in-plane current along the a axis of WTe 2 has also been reported to enable partial switching of perpendicular magnetization in WTe 2/Fe2.78GeTe 2 without an external magnetic field [53]. FIG. 2. (a) Crystal structure near the surface of WTe 2, displaying a mirror symmetry relative to the bc plane but not to the ac plane. (b) Symmetric (VS) and antisymmetric (VA) components of ST -FMR signal for WTe 2 (5.5)/Py (6) device as a function of the angle of the in-plane magnetic field [50]. Reprinted with permission from Mac Neill et al., Nat. Phys. 13, 300 (2017). (c) Current -induced magnetization switching of WTe 2(15)/SrRuO3 when the current is along the low - symmetry a axis where the magnetization can be switched without an external magnetic field [52]. Reprinted with permission from Xie et al., APL Mater. 9, 051114 (2021). (d) Structure of the MoTe 2 crystal in the monoclinic ( β or 1T′) phase depicted in the a-c plane for which the mirror plane is within the page and the Mo chains run into the page. (e) Symmetric and antisymmetric ST -FMR resonance components for the MoTe 2(0.7)/Py(6) device with a current applied perpendicular to the MoTe 2 mirror plane as a function of the orientation of the in-plane magnetic field. (f) The conductivities of damping -like torque of perpendicular spins (blue) and transverse spins (red) as a function of the MoTe 2 thickness for devices with current aligned perpendicular to the MoTe 2 mirror plane [51]. Reprinted with permission from Stiehl et al ., Phys. Rev. B 100, 184402 (2019). (g) Crystal structure of the monoclinic 1T′ phase of MoTe 2. (h) Antisymmetric ST -FMR components for MoTe 2 (83.1)/Py(6) as a function of the orientation of the in-plane magnetic field. (i) MOKE images implying switching of Py by current [93]. Reprinted with permission from Liang et al., Adv. Mater. 32, 2002799 (2020). 4 FIG. 3. (a) Current -induced magnetization switching in sputter -deposited WTe 2(10)/CoTb(6)/Ta(2) Hall bar (Hx= ± 900 Oe) [97]. Reprinted with permission from Peng et al., ACS Appl. Mater. Interfaces 13, 15950 (2021). (b) Second harmonic longitudinal resistance (𝑅2𝜔𝑥𝑥) of WTe x(5)/Mo(2)/CoFeB(1) measured as a function of pulse current amplitude Ipulse under zero external field [98]. (c) Current -induced switching loops of WTe x(5)/Ti(2)/CoFeB (1.5) Hall bar under different in- plane magnetic fields at 200 K [99]. Reprinted with permission from Xie et al., Appl. Phys. Lett. 118, 042401 (2021). (d) Dependences on the WTe x thickness of damping -like SOT efficiency (𝜉DL𝑗) and the WTe x resistivity (𝜌𝑥𝑥) for WTe 2/CoFeB . (e) Apparent spin Hall conductivity as a function of the longitudinal conductivity for WTe 2/CoFeB [98]. Data in (b), (d), and (e) are r eprinted with permission from Li et al., Matter 4, 1639 (2021). FIG. 4. (a) Schematic of the CVD growth process for PtTe 2. (b) High -resolution transmission electron microscopy image of a 5 nm PtTe 2 thin film. (c) Current -induced magnetization switching in the PtTe 2(5)/Au(2.5) /CoTb (6) Hall bar under different in -plane field s [92]. Reprinted with permission from Xu et al., Adv. Mater. 32, 2000513 (2020). (d) Spin torque ferromagnetic resonance spectrum of a ZrTe 2/Py bilayer at room temperature. (e) Current -induced magnetization switching in ZrTe 2(8 u.c.)/CrTe 2(3 u.c.) Hall bar under a 700 Oe in -plane field at 50 K [94]. Reprinted with permission from Ou et al., Nat. Commun. 13, 2972 (2022). β-MoTe 2 is a semimetal that retains inversion symmetry in bulk but has a low-symmetry interface (the group space is Pmn21 in bulk but Pm11 in few -layer structures [Fig. 2(d)]). Stiehl et al. [51] observed damping - like SOT of both transverse spins (≈ 8×103 (ℏ/2e) Ω-1 m-1) and perpendicular spins (≈ 1×103 (ℏ/2e) Ω-1 m-1) in mechanically exfoliated β-MoTe 2/Ni81Fe19 bilayers [Fig. 2(e)]. This torque of perpendicular spins is one -third strong than that of WTe 2/Ni 81Fe19 and was attributed to perpendicularly polarized spin current from the surface of the low -symmetry β-MoTe 2 [Fig. 2(f)]. This appears to suggest that the breaking of bulk inversion symmetry is not an essential requirement for producing perpendicular spins. However, 1T′ -MoTe 2 [Fig. 2(g)] was reported to generate 5 no damping -like SOT of perpendicular spins in contact with Ni 81Fe19 [Fig. 2(h)][93]. Instead, 1T′ -MoTe 2 only generates a nonzero damping -like SOT of transverse spins that switches the in -plane magnetized Ni 81Fe19 layer at a current density of 6.7×105 A cm-2 [Fig. 2(i)]. NbSe 2 with resistivity anisotropy was reported to generate a perpendicular Oersted field but no perpendicular or longitudinal spins when interfaced with Ni 81Fe19 [89]. The damping -like toque of transverse spins in mechanically exfoliated NbSe 2/Ni81Fe19 is very weak and corresponds to a spin Hall conductivity of ≈ 103 (ℏ/2e) Ω-1 m-1 [89]. Here it is important to note that, while the presence of perpendicular spins has been widely concluded in the literature from a sin2 φ-dependent contribution in symmetric spin-torque ferromagnetic resonance signal of in-plane magnetization ( φ is the angle of the external magnetic field relative to the current), or a φ-independent but field -dependent contribution in the second harmonic Hall voltage of in -plane magnetization, or field -free switching, none of the three characteristics can simply “signify” the presence of a flow of perpendicular spins. This is because non -uniform current effects that can generally exist and generates out -of-plane Oersted field in nominally uniform, symmetric Hall bars a nd ST -FMR strips [90,91,96,100] also exhibit all three characteristics. As demonstrated by Liu and Zhu [100], these characteristics can be considerable especially when the devices have strong current spreading, e.g., in presence of non-symmetric electric contact s. 2.2 Large -area Transition Metal Dichalcogenides So far, most TMD studies have been based on mechanical exfoliation, which is unsuitable for the mass production of spintronic applications. Recently, efforts have been made in large -area growth of thin-film TMDs towards the goal of SOT applications [97,98]. For example, sputter -deposited WTe x has also developed to drive low - current -density switching of CoTb ( jc ≈ 7.05×105 A cm-2 under in -plane assisting field of 900 Oe) [Fig. 3(a)] [97] and in WTe x/Mo/CoFeB ( jc ≈ 7×106 A cm-2, under no in- plane assisting field, Fig. 3(b))[99] and in WTe x/Ti/CoFeB (jc ≈ 2.0×106 A cm-2 under in -plane assisting field of ± 30 Oe, Fig. 3(c))[98]. It has become a consensus that the spin - orbit torque in these sputter -deposited WTe x/FM samples arises from the bulk spin Hall effect of the WTe x [97,98]. As indicated in F igs. 3(d) and 3(e), the measured spin -orbit torque efficiency increases but the apparent spin Hall conductivity decreases as the resistivity increases in the dirty limit [41] due to increasing layer thickness. Large -area PtTe 2 films with relatively high electrical conducti vity (≈ 3.3×106 Ω-1 m−1 at room temperature) and spin Hall conductivity (2×105 ℏ/2e Ω −1m−1) have also been reported by annealing a Pt thin film in tellurium vapor at ≈ 460 °C [Figs. 4(a) and 4(b)][92]. PtTe 2 is predicted to be a type -II Dirac semimetal with spin -polarized surface states. However, there is no indication of the generation of torques of out -of-plane spins. Partial switching of magnetization by in -plane current has also been reported in a PtTe 2(10)/Au(2.5)/CoTb(6) Hall bar ( jc ≈ 9.9×106 A cm−2 under in -plane assisting field of 2 kOe) [Fig. 4(c)]. Growth of ZrTe 2 by molecular beam epitaxy (MBE) has also been reported. A ST -FMR study has measured a small damping -like torque of transverse spins for MBE -grown ZrTe 2/Ni 81Fe19 bilayer at room temperature [Fig. 4(d)][94]. This is consistent with the theoretical prediction that ZrTe 2 is a Dirac semimetal with massless Dirac fermions in its band dispersion [101] but vanishing spin Hall conductivity. Even so, a ZrTe 2(8 u.c.)/CrTe 2(3 u.c.) bilayer has been reported to be partially switched at 50 K by an in -plane current of 1.8×107 A cm-2 in density under an in-plane assisting field of 700 Oe [Fig. 4(e)]. 3. Magnetization s witching by spin current from Bi- based topological insulators Another kind of layered strong -SOC material is Bi - based topological insulators [102-104]. As displayed in Fig. 5 (a) , TIs are insulating in the bulk but conducting at the surface. The initial interest of TIs for spin-orbit torque studies is the topological surface states ( Fig. 5 (b) ). In the wavevector space, the spin and momentum of electrons are one-to-one locked to each other at the Fermi level. With a flow of charge current, the shift in the electron distribut ion in the wavevector space induces non -equilibrium spin accumulation (Fig. 5 ( c)). Fig. 5. Topological surface states and spin -accumulation in topological insulators. (a) Real -space picture of the conducting surface states in an ideal topological insulator [103]. Reprinted with permission from Han an d Liu, APL Mater. 9, 060901 (2021). (b) Angle -resolved photoemission spectrum that indicates the bulk and surface bands of a six -quintuple -layer -thick Bi 2Se3 film [102]. Reprinted with permission from Zhang et al., Nat. Phys. 6, 584 (2010). Copyright 2010 Springer Nature Limited. (c) Current -induced spin accumulation in a topological insulator [104]. The arrows denote the directions of spin magnetic moments, which are opposite to the corresponding spin angular momenta. Reprinted with permission from He et al., Nat. Mater. 21, 15 –23 (2 022). 3.1 MBE -grown and exfoliated Topological insulators Topological insulators were first introduced in the field of spin -orbit torque i n 2014 . From ST -FMR measurement , Mellnik et al . measured a giant damping - like spin-orbit torque efficiency (𝜉DL𝑗 = 3.5) at room temperature in Bi 2Se3/Py bil ayers grown by MBE [Fig. 6(a)][23]. In the same year, Fan et al . reported from harmonic Hall measurement a damping -like torque efficiency of 4 25 and the spin–orbit torque switching in the 6 (Bi 0.5Sb0.5)2Te3/(Cr 0.08Bi0.54Sb0.38)2Te3 bilayers [24] at 1.9 K [Fig. 6(b)]. As shown in Fig. 7 , room -temperature magnetization switching by spin current from TIs (e.g., Bi 2Se3, Bi 2Te3, and BiSb ) has been demonstrated in Hall -bar samples [99,105]. Han et al. [76,106] first reported magnetization switching in Hall bars of Bi 2Se3(7.4)/Co 0.77Tb0.23(4.6) bilayer ( Hx = 1000Oe, Hc ≈ 200Oe, Jc ≈ 2.8×106 A cm-2, switching ratio=85%)[Fig. 7(a)]. The damping -like SOT efficiency was determined to be 0.16 ± 0.02 for the Bi2Se3/Co 0.77Tb0.23. Similar results have been also reported by Wu et al. [77] in Bi 2Se3/Gd x(FeCo) 1−x Hall bars ( 𝜉DL𝑗= 0.13, Jc ≈ 2.2×106 A cm-2)[Fig. 7(b)]. These values of spin - orbit torque efficiency are significantly low compared to those from Bi 2Se3/Py samples, which may be understood partly by the increased spin current relaxation via spin - orbit scattering in the ferrimagnets [6]. Khang et al. have reported a spin -orbit torque efficiency of 52 (as determined from a coercivity change measurement) and resistivity of 400 μΩ cm for MBE -grown Bi 1-xSbx [107]. Switching of MBE -grown fully epitaxial Mn 0.45Ga0.55/Bi0.9Sb0.1 has also been demonstrated at a current density of 1.1×106Acm–2 (Hx = 3.5 kOe) in [Fig. 7(c)]. Non -epitaxial BiSb films (10 – 20 nm) grown by MBE were also reported to have a high spin-orbit torque efficiency of up to 3.2 and to enable magnetization switching at a current density of 2.2×106 A cm-2. There have also been reports of SOT switching of exfoliated van-der-Waals magnets at low temperatures, such as in Fe 3GeTe 2 [108,109] and Cr 2Ge2Te6 [110,111], by the spin current from TIs. Liu et al . [112] reported a strongly temperature -dependent damping -like torque efficiency of up to 70 from a field -dependent harmonic Hall response measurement[Fig. 8(a)], and current switching of MBE -grown Bi 2Te3/MnTe Hall bar at a critical current density of down to 6.6×106 Acm–2 (Hx = ± 400 Oe, T = 90 K) [Fig. 8(b)]. FIG. 6. (a) Spin torque ferromagnetic resonance spectrum for Bi 2Se3(8)/Py(16) bilayer at room temperature [23]. Reprinted with permission from Mellnik et al., Nature 511, 449 (2014). (b) Second harmonic Hall resistance for (Bi 0.5Sb0.5)2Te3(3 QL)/(Cr 0.08Bi0.54Sb0.38)2Te3(6 QL) bilayer as a function of the in -plane field angle for different applied a.c. current [24]. Reprinted with permission from Fan et al., Nat. Mater. 13, 699 (2014). 3.2 Sputter -deposited Topological Insulators Since exfoliation and molecular -beam epitaxy are less realistic methods for the preparation of large -area TI thin films for practical SOT devices, s puttering has been introduced to grow amorphous or polycrystalline “topological insulators”. The first report of sputter - deposited “topological insulators” is BixSe(1–x) [113] with relatively high electrical conductivity ( 0.78×105 Ω-1m-1 for 4 nm thickness) . Such sputter -deposited Bi xSe(1–x) exhibits a very high damping -like spin-torque efficiency of 18 and enabled magnetization switching in a BixSe(1– x)(4)/Ta(0.5)/CoFeB(0.6)/Gd(1.2)/CoFeB(1.1) at a low current density of ≈ 4.3×105 A cm-2 [Fig. 8(c)]. Wu et al. [105] also reported room -temperature witching of Bi2Te3/Ti/CoFeB at a current density of 2 .4×106Acm–2 (Hx = 100 Oe). In the Hall bar of sputter -deposited PMA Bi2Te3(8)/CoTb(6) bilayer, current -induced magnetization switching was reported at a low critical current density of 9.7×105 A cm-2 [Fig. 8(d)]. Sputter -deposited BiSb films (10 nm) were reported to provide a spin -torque efficiency of 1.2 and to drive switching of CoTb at 4×105 A/cm-2 [106]. 3.3 Practical impact As we have discussed above, some TIs and their sputter -deposited counterparts are reported to have much higher damping -like torque efficiency than heavy metals . Meanwhile, the sputter -deposited TIs are typically several times more resistive than the MBE -grown ones since disordered films typically have stronger electron scattering than crystalline films. However, for practical SOT applications, the spin -source ma terials are required to have low resistivity and large damping -like spin -orbit torque efficiency . Despite their amazingly high damping -like spin-orbit torque efficiency , most TIs are highly resistive (> 1×103 μΩ cm), much more resistive than ferromagnetic metals in metallic spintronic devices (e.g., 110 μΩ cm for CoFeB). Current shunting into the adjacent metallic layers would be considerably more than that flow s within the topological insulator layer, resulting in increases in the total switching current a nd power consumption of devices. 3.4 Mechanism of the spin current generation Despite the debate, t he two main mechanism s via which the TIs and their disordered counterparts generate spin current or spin accumulation are the spin Hall effect and the surface states. As suggested by Khang et al. [82], Chi et al. [43], Tian et al. [44], the bulk spin Hall effect is the dominant source of the spin current for the spin -orbit torque in Bi 0.9Sb0.1, Bi0.53Sb0.47, and BixTe1–x. As shown in Figs. 9(a) and 9(b), in disordered Bi0.53Sb0.47 the apparent spin Hall conductivity increases non -linearly with increasing layer thickness, which is a typical spin diffusion behavior and in good consistent with a bulk spin Hall effect being the mechanism of the spin current generation. In contrast, the surface states of the TIs have been suggested to be the main spin current source in MBE -grown (Bi 1- xSbx)2Te3 [105] and Bi 2Te3[112]. This suggestion is consistent with the strong dependence of the damping -like spin-orbit torque on the composition [105], the temperature [112], and thus the location of the Fermi level relative to the Dirac point [Figs. 9(c) - 9(e)][105,114]. In addition, DC et al . suggested that the quantum confinement effect of small grains should account for the high spin -torque efficiency in the sputter -deposited Bi xSe(1–x) [113]. 7 FIG. 7. Current -induced magnetization switching at room temperature in (a) Bi2Se3(7.4)/CoTb(4.6) bilayer ( the in-plane magnetic field is 1000 Oe) [76], (b) Bi2Se3(6)/Gd x(FeCo) 1−x(15) bilayer ( an in -plane magnetic field is 1000 Oe) [77], and (c) Mn 0.45Ga0.55(3)/Bi0.9Sb0.1(5) (3.5 kOe) [115]. Data in (a) is reprinted with permission from Han et al., Phys. Rev. Lett. 119, 077702 (2017). Data in (b) is reprinted with permission from Wu et al., Adv. Mater. 31, 1901681 (2019) ; Data in ( c) is reprinted with permission from Khang et al., Nat. Mater. 17, 808 (2018). FIG. 8. (a) Variation of the spin Hall ratio of Bi 2Te3 with temperature. ( b) Current -induced magnetization switching of Bi2Te3(8)/MnTe(20) at 90 K under different in -plane magnetic field s [112]. Reprinted with permission from Liu et al., Appl. Phys. Lett. 118, 112406 (2021). (c) Bi xSe(1–x)(4)/Ta(0.5)/CoFeB(0.6)/Gd(1.2)/CoFeB(1.1) (the in-plane magnetic field is 80 Oe) [113], reprinted with permission from DC et al., Nat. Mater. 17, 800 (2018) . (d) Current switching of sputter -deposited Bi2Te3(8)/CoTb (6) under an in -plane magnetic field of 400 Oe at room temperature [116]. Reprinted with permission from Zheng et al., Chin. Phys. B 29, 078505 (2020). 8 FIG. 9. (a) Scanning transmission electron microscopy image of the 0.5 Ta/[0.35 Bi|0.35 Sb] N/0.3 Bi/2 CoFeB/2 MgO/1 Ta structure with N = 8 . (b) thickness dependence of the apparent spin Hall conductiv ity σSH of Bi0.53Sb0.47 [43]. Reprinted with permission from Chi et al., Sci. Adv. 6, eaay2324 (2020). (c) Fermi level , (d) Resistivity ρxx and 2D carrier density |n2D| for of (Bi 1-xSbx)2Te3 with different Sb percentage s. (e) Switching current density |Jjc| and effective damping -like spin - orbit torque field vs the Sb ratio of (Bi 1-xSbx)2Te3 [105]. Reprinted with permission from Wu et al., Phys. Rev. Lett. 123, 207205 (2019). 4. Magnetization s witching of Van der Waals magnet 4.1 Van der Waals magnet The r ecent discovery of v an der Waals magnets (e.g., Cr2Ge2Te6 [82], CrI 3 [83], etc.) has attached remarkable attention in the field of magnetism and spintronics. While the origin of the long-range magnetic order is still under debate, it has been suggested to have a close correlation with the suppression of thermal fluctuations by magnetic anisotropy. Note that in absence of magnetic anisotropy, no long-range magnetic order is expected by the Mermin – Wagner theorem [117] at finite temperature in a two- dimensional system . Van-der-Waals magnets provide a unique, highly tunable platform for spintronics. Most strikingly, the properties of van der Waals FMs, such as Curie temperature (TC) [78,80], coercivity [79,80], and magnetic domain structure [81], can be tuned significantly by a variety of techniques ( e.g., layer thickness, ionic liquid gating [78], proton doping [79], strain [80,81], exchange bias [118,119], interfacial proximity -effect [120], etc.). An interesting example is CrI 3, whose magnetic ordering depends on the number of layers and can be tuned by an external magnetic field. As shown in Fig. 10(a) , the CrI 3 is ferromagnetic at 1 monolayer thickness, antiferromagnetic at 2 monolayer thickness , and ferromagnetic at 3 monolayer thickness. Ferromagnetic CrI 3 also shows a relatively square perpendicular magnetization loop [83]. Following the long -range ordering of magnetic lattices, magnetic materials can be grouped into ferromagnets, ferrimagnets, and antiferromagnets . In general, ferromagnets and ferrimagnets are considered more friendly than antiferromagnet s to be integrated into electric circuits because their magnetization states can be electrically detected by anomal ous Hall effect or tunnel magnetoresistance and efficiently switched by SOTs. In contrast, electrical detection and switching of collinear antiferromagnets [121-123] are generally much more challenging [124], despite the recent discovery of magnetoresistance and anomalous Hall in non -collinear antiferromagnets Mn 3Sn [34,125,126]. For this reason, spin-torque switching of magnetization is mostly studied and better understood in ferromagnetic and ferrimagnetic systems than in antiferromagnets. Our discussion below will be focused on van der Waals ferromagnet s [116-134]. The van-der-Waals magnet CrBr 3 (TC = 34 K)[127,128] (TC = 34 K), CrI 3 (TC = 45 K) [83], Cr 5Te8 [129] and VI 3 (TC = 60 K) [130] have perpendicular magnetic anisotropy but low Curie temperature . So far, room -temperature ferromagnetism and low-temperature perpendicular magnetic anisotropy ha ve been reported for van der Waals materials FeTe [131], Fe 4GeTe 2 (Fig. 10(b))[132], Fe 5GeTe 2 [133], CrTe [134], CrTe 2 (Fig. 10(c))[135-137], Cr 1+δTe2 [138], Cr 2Te3 [139], Cr 3Te4 [140]), CrSe [141], and Fe 3GaTe 2 (Fig. 10( d))[142]. In Fig. 11, we summar ize the representative results of the Curie temperature and magnetization of relatively thin van der Waals magnets (note that TC of van der Waals magnets is strongly thickness dependent ). While FenGeTe 2 can have good PMA at low temperatures and CrnTem and CrSe are relatively stable in air, they lose square hysteresis loops at room temperature [Fig. 10(b) and 10(c)]. The recently discovered Fe 3GaTe 2 [142] is an outstanding van der Waals ferromagnet that can have both a high Curie temperature (TC ≈ 350 - 380 K) and large PMA energy density ( Ku ≈ 4.8×105 J m-3) [Fig. 10(d)]. Search ing for Van der Waals magnets with room -temperature ferromagnetism, strong perpendicular magnetic anisotropy, a nd high stability in the air at the same is expected to be an active topic in the field. 4.2 Magnetization s witching of v an der Waals ferromagnets Spin-orbit torque switching of v an der Waals ferromagnets was first demonstrated in perpendicularly magnetized Fe3GeTe 2/Pt bilayer s [108,109], where the spin 9 current generated by the SHE in the Pt exerts a damping - like spin torque on the Fe 3GeTe 2 [Fig. 12(a)]. Interestingly , despite the small layer thicknesses and small magnetization of the Fe 3GeTe 2, the Fe 3GeTe 2/Pt samples have a high depinning field (coercivity) and strong perpendicular magnetic anisotropy such that they typically require a large current density of ~ 107 A cm-2 [108,109] as well as an in - plane magnetic field [Fig. 12(a)]. As indicated by the anomalous Hall resistance, the Fe 3GeTe 2 was also only partially switched , with the switching ratio of 20%-30% in [108] and 62% in [118], probably due to the non- uniformity of the magnetic domains with the van der Waals layer. Cr2Ge2Te6 is another well -studied van der Waals ferromagnet . Spin-orbit torque switching of Cr2Ge2Te6 has been demonstrated in Ta/Cr 2Ge2Te6 bilayers (Curie temperature < 65 K) at a low current density of 5×105 A cm-2 at 4 K, with an in -plane assisting magnetic field of 200 Oe [143]. Zhu et al . [110] reported SOT switching of Cr2Ge2Te6/W with interface -enhanced Curie temperature of up to 150 K [Fig. 12(b)]. Current -induced magnetization switching has also been realized in all van der Waals heterojunction s. Nearly full magnetization switching (88%) has been reported in MBE -grown Cr2Ge2Te6/(Bi 1-xSbx)2Te3 bilayer s [144][Fig. 13(a)]. In the (Bi0.7Sb0.3)2Te3/Fe 3GeTe 2 bilayer [145], the threshold current density for the magnetization switching is 5. 8×106 A cm-2 at 100 K [Fig. 13(b)]. Field-free switching magnetization has been reported in the exfoliation -fabricated WTe 2/Fe 2.78GeTe 2 bilayers by a current along the low symmetry axis (9.8×106 A cm-2, T =170K) [53]. In the same bilayer structure, Shin et al. also realized magnetization switching at a current density of 3.9×106 A cm-2 (T = 150 K, Hx = 300 Oe , Fig. 13(c)) [146]. FIG. 1 0. (a) Kerr rotation vs perpendicular magnetic field for monolayer (1L), bilayer (2L) , and trilayer (3L) CrI 3 flake [83]. Reprinted with permission from Huang et al., Nature 546, 270 (2017). ( b) Hall conductivity hysteresis loop of a 11 - mono layer -thick Fe 4GeTe 2 crystal at various temperature s [132]. Reprinted with permission from Seo et al., Sci. Adv. 6, eaay8912 (2019). (c) Out-of-plane magnetization hysteresis loop of 7 monolayer CrTe 2 at different temperatures along the out -of-plane direction [137]. Reprinted with permission from Zhang et al ., Nat. Commun. 12, 2492 (2021). ( d) Anomalous Hall resistance hysteresis (of Fe3GaTe 2 with different thickness es at 3 K and 300 K [142]. Reprinted with permission from Zhang et al., Nat. Commun. 13, 5067 (2022). FIG. 1 1 Saturation magnetization Ms vs Curie temperature TC of representative van der Wa als ferromagnets . 10 FIG. 1 2. (a) Current -driven perpendicular magnetization switching for Fe 3GeTe 2(4)/Pt(6) bilayer under an in-plane magnetic field of 50 0 Oe at 100 K [108]. Reprinted with permission from Wang et al., Sci. Adv. 5, eaaw8904 (2019). (b) Current -driven perpendicular magnetization switching for Cr2Ge2Te6(10)/W(7) bilayer under in -plane magnetic fields of ± 1 kOe at 150 K [110]. Reprinted with permission from Zhu et al., Adv. Funct. Mater. 32, 2108953 (2022). FIG. 1 3. Current induced switching of van der Waals magnets. (a) Normalized anomalous Hall resistance vs current density for (Bi 1-xSbx)2Te3(6)/Cr 2Ge2Te6(t) (x = 0.5) with different Cr 2Ge2Te6 thicknesses under an in-plane magnetic field of 1 kOe at 2 K [144]. Reprinted with permission from Mogi et al., Nat. Commun. 12, 1404 (2021). (b) Anomalous Hall conductivity vs current density for (Bi 1-xSbx)2Te3(8)/Fe3GeTe 2(6) at different temperatures under an in-plane magnetic field of 1 kOe [145]. Reprinted with permission from Fujimura et al., Appl. Phys. Lett. 119, 032402 (2021). (c) Hall resistance vs current density for Fe 3GeTe 2(7.3)/WTe 2(12.6) under in -plane magnetic field Hx = 300 Oe . The Hall resistance varies during three consecutive current scans due to temperature rise towards the Curie temperature [146]. Reprinted with permission from Shin et al., Adv. Mater. 34, 2101730 (2022). 5. Simplifying models of switching current density In this section, we provide a quantitative understanding of the switching current densities in the van der Waals system by considering the simplifying models. the transverse spin damping -like SOT efficiency per unit current density ( 𝜉𝐷𝐿𝑗 ) of a heterostructure with PMA inversely cor relates to the critical switching current density (jc) in the spin -current -generating layer via Eq. (4) in macrospin limit [149,150] and via Eq. (5) in the domain wall depinning limit [151-153], i.e., jc ≈ eμ0MstFM (Hk-√2|𝐻𝑥|)/ћ𝜉DL𝑗, (4) jc = (4e/πћ) μ0MstFMHc/𝜉DL𝑗, (5) where e is the elementary charge, ℏ is the reduced Planck constant, μ0 is the permeability of vacuum, Hx is the applied field along the current direction, and tFM, Ms, Hk, and Hc are the thickness, the saturation magnetization, the effective perpendicular anisotropy field, and the perpendicular coercivity of the driven magnetic layer FM, respectively. 11 However, r ecent experiments [154] on heavy metal/m agnet bilayers have shown that neither Eq. (4) nor Eq. (5) can provide a reliable prediction for the switching current and 𝜉𝐷𝐿𝑗 and that there is no simple correlation between 𝜉𝐷𝐿𝑗 and the critical switching current density of realistic perpendicularly magnetized spin -current generator/ferromagnet heterostructures. As shown in Table I, the same is true for the van der Waals systems. The macrospin analysis does not seem to apply to the switching dynamics of micrometer -scale samples so that the values of 𝜉𝐷𝐿𝑗 determined using the switching current density and Eq. ( 4) can produce overestimates by up to hundreds of times (𝜉𝐷𝐿,𝑚𝑎𝑐𝑟𝑜𝑗 and 𝜉𝐷𝐿,𝑚𝑎𝑐𝑟𝑜𝑗/𝜉𝐷𝐿𝑗 in Table I). A domain - wall depinning analysis [ Eq. ( 5)] can either under - or over -estimated 𝜉𝐷𝐿𝑗 by up to tens of times (𝜉𝐷𝐿,𝐷𝑊𝑗 and 𝜉𝐷𝐿,𝐷𝑊𝑗/𝜉𝐷𝐿𝑗 in Table I). These observations consistently suggest that the switching current or “switching efficiency” of perpendicular heterostructures in the micrometer or sub - micrometer scales cannot provide a quantitative estimation of 𝜉𝐷𝐿𝑗. While the underlying mechanis m of the failure of the simplifying models remains an open question, it is obvious that Joule heating during current switching of the resistive or low Curie -temperature van der Waals systems can have a rather significant influence on the apparent switching current density. As shown in Fig. 13(c), the anomalous Hall resistance hysteresis loop drifts for three consecutive current scans because of Joule heating [146]. TABLE I. Comparison of spin -torque efficiencies determined from the harmonic response or ST -FMR (𝜉𝐷𝐿𝑗 ) and magnetization switching ( 𝜉𝐷𝐿,𝐷𝑊𝑗, 𝜉𝐷𝐿,𝑚𝑎𝑐𝑟𝑜𝑗) of PMA samples, which is calculated from Eq. (5) and Eq. (4) using the applied external magnetic field ( Hx), saturation magnetization ( Ms), the perpendicular coercivity ( Hc), and the effective perpendicular anisotropy field ( Hk) of the driven magnetic layer , and the critical magnetization switching current density (jc). The value of 𝜉𝐷𝐿,𝑚𝑎𝑐𝑟𝑜𝑗 is estimated to be negative for Te2(10)/CoTb(6) [97] and ZrTe 2(7.2) /CrTe 2(1.8) [94] because according to th e original reports an in -plane field greater that the effective perpendicular anisotropy field was applied. Sample Technique Ms (emu cm-3) Hc (Oe) Hk (kOe) Hx (Oe) jc (MA cm-2) 𝜉𝐷𝐿𝑗 𝜉𝐷𝐿,𝐷𝑊𝑗 𝜉𝐷𝐿,𝑚𝑎𝑐𝑟𝑜𝑗 𝜉𝐷𝐿,𝐷𝑊𝑗/𝜉𝐷𝐿𝑗 𝜉𝐷𝐿,𝑚𝑎𝑐𝑟𝑜𝑗/𝜉𝐷𝐿𝑗 Te2(10)/CoTb(6) [97] sputtering 48 60 0.33 900 0.7 0.2 0.47 -5.8 2.4 -29 ZrTe 2(7.2) /CrTe 2(1.8) [94] MBE 100 -- 0.2 700 18 0.014 -- -0.12 -- -8.5 Bi2Se3(7.4)/CoTb(4.6) [76] MBE 280 300 -- 1000 2.8 0.16 2.7 -- 17 -- Bi2Se3(6)/Gd x(FeCo) 1−x(15) [77] MBE 46 160 0.35 1000 2.2 0.13 0.98 10 7.6 78 BixSe(1–x) (4)/Ta(0.5)/ CoFeB (0.6)/Gd(1.2)/CoFeB(1.1) [113] sputtering 300 30 6 80 0.43 8.67 1.2 180 0.13 21 (BiSb) 2Te3(6QL)/Ti (2)/CoFeB(1.5) [105] MBE 868 30 2.24 100 0.12 2.5 6.3 345 2.5 138 Bi2Te3/Ti(2) (6QL)/CoFeB(1.5) [105] MBE 868 27 2.06 100 2.4 0.08 0.287 16 3.5 194 Bi2Te3(8)/MnTe(20) [112] MBE 175 100 ≈50 400 6.6 10 1.0 397 0.10 40 Bi0.9Sb0.1(5) /Mn 0.45Ga0.55(3) [115] MBE 240 4500 10 3500 1.1 52 57 50 1.1 0.96 SnTe(6QL)/Ti(2)/CoFeB(1.5) [105] MBE 868 53 2.18 100 1.5 1.41 0.91 27.5 0.65 20 Fe3GeTe 2(4)/Pt(6) [108] exfoliation 16 125 11 500 11.6 0.12 0.013 0.86 0.11 7.2 Fe3GeTe 2 (15)/Pt(5) [109] exfoliation 170 750 30 3000 20 0.14 1.8 50 13 355 6. Conclusion and outlook We have reviewed recent advances in spin -orbit torque and resultant magnetization switching in van der Waals systems. Van der Waals materials provided unique opportuni ties for spintronics because of their diversity of materials, the flexibility of preparat ion (e.g., by mechanical exfoliation), and strong tunability by interface effects. Van der Waals TMDs such as WTe 2 also exhibit potential as a spin current source of both transverse spins and exotic perpendicular and lo ngitudinal spins . Bismuth -based topological insulators and their sputter -deposited disordered counterparts are shown to generate giant damping -like SOT with the efficiency of up to 1-3 orders of magnitudes greater than 5 d metals. Moreover, van der Waals materials that are m agnetic are also interesting for spin-orbitronics due to their highly tunable magnetism and low magnetization at small thicknesses since the damping -like and field -like effective SOT fields exerted on the magnetic layer by a given spin current scale invers ely with the thickness and magnetization of the magnetic layer. Efficient switching of several van der Waals magnet s by spin current has been demonstrated . Despite the se exciting progres s, essential challenges have remained to overcome for spin-orbit torque switching of van der Waals systems : (i) While the generation of different spin components has been demonstrated, the efficiencies are typically quite low. It has remained unclear as to whether and how the efficiency of generating exotic perpendicular and longitudinal spins by the low -symmetry TMDs can be improved significantly to be compelling for practical SOT applications . (ii) Some Bi -based t opological i nsulators and alloys have both giant effective spin Hall ratio and resistivity at 12 the same time, the latter is undesirable for metallic SOT devices that require energy efficiency, high endurance, and low impedance. It would be interesting if new uniform, stable spin -orbit materials can be developed to provide damping -like SOT efficiency of greater than 1 but substantially less resistive than the yet -know n topological insulators. (iii) So far, large -area growth of van der Waals magnets that have high Cu rie temperature, strong perpendicular magnetic anisotropy, and high stabilities against heating and atmosphere at the same time has remained a key obstacle that prevents van der Waals magnets from applications in spintronic technologie s. Breakthroughs in th e fabrication of such application - friendly van der Waals magnets are in urgent need. (iv) While spin -orbit torque switching of magnetization has been demonstrated in Hall -bar devices containing van der Waals magnets, TMDs, or/and topological insulators, t he simplifying models of macrospin rotation and domain wall depinning cannot provide an accurate prediction for the switching current density. So far, the quantitative understanding of the switching current remains a fundamental problem . (v) From the point of view of magnetic memory and logic application, s witching of thermally stable nanodots of van der Waals magnet s, such as the free layers of nanopillars of magnetic tunnel junctions, by current pulse s of picosecond and nanosecond durations. Efforts are a lso required on the demonstration and optimization of t he key performance , e.g., the endurance, the write error rates, the retention, and the tunnel magnetoresistance . Acknowledgments This work was supported partly by the National Key Research and Develop ment Program of China (Grant No. 2022YFA12 0400 4) and by the National Natural Science Foundation of China (Grant No. 12274405 ). Reference [1] A. Manchon, J. Železný, I. M. Miron, T. Jungwirth, J. Sinova, A. Thiaville, K. Garello, and P. Gambardella, Current -induced spin -orbit torques in ferromagnetic and antiferromagnetic systems. Rev. Mod. Phys. 91, 035004 (2019). [2] I. Mihai Miron, K. Garello, G. Gaudin, P. -J. Zermatten, M. V . Costache, S. Auffret, S. Bandiera, B. Rodmacq, A. Schuhl, a nd P. Gambardella, Perpendicular switching of a single ferromagnetic layer induced by in - plane current injection. Nature 476, 189 (2011). [3] L. Q. Liu, C. F. Pai, Y . Li, H. W. Tseng, D. C. Ralph, and R. A. Buhrman, Spin -Torque Switching with the Giant Spin Hall Effect of Tantalum. Science 336, 555 (2012). [4] L. Liu, O. J. Lee, T. J. Gudmundsen, D. C. Ralph, and R. A. Buhrman, Current -Induced Switching of Perpendicularly Magnetized Magnetic Layers Using Spin Torque from the Spin Hall Effect. Phys. Rev. Let t. 109, 096602 (2012). [5] L. Zhu, Switching of Perpendicular Magnetization by Spin-Orbit Torque. Adv. Mater. n/a, 2300853 (2023). [6] L. Zhu and D. C. Ralph, Strong variation of spin -orbit torques with relative spin relaxation rates in ferrimagnets. Nat. Commun. 14, 1778 (2023). [7] P. M. Haney, H. -W. Lee, K. -J. Lee, A. Manchon, and M. D. Stiles, Current induced torques and interfacial spin - orbit coupling: Semiclassical modeling. Phys. Rev. B 87, 174411 (2013). [8] Y.-T. Chen, S. Takahashi, H. Nakayama, M. Althammer, S. T. B. Goennenwein, E. Saitoh, and G. E. W. Bauer, Theory of spin Hall magnetoresistance. Phys. Rev. B 87 (2013). [9] V . P. Amin and M. D. Stiles, Spin transport at interfaces with spin -orbit coupling: Phenomenology. Phys. Rev. B 94, 104420 ( 2016). [10] L. J. Zhu, D. C. Ralph, and R. A. Buhrman, Effective Spin-Mixing Conductance of Heavy -Metal -Ferromagnet Interfaces. Phys. Rev. Lett. 123, 7, 057203 (2019). [11] L. Zhu, L. Zhu, and R. A. Buhrman, Fully Spin - Transparent Magnetic Interfaces Enabl ed by the Insertion of a Thin Paramagnetic NiO Layer. Phys. Rev. Lett. 126, 107204 (2021). [12] Y . Liu, Z. Yuan, R. J. H. Wesselink, A. A. Starikov, and P. J. Kelly, Interface Enhancement of Gilbert Damping from First Principles. Phys. Rev. Lett. 113, 2072 02 (2014). [13] J. C. Rojas -Sánchez, N. Reyren, P. Laczkowski, W. Savero, J. P. Attané, C. Deranlot, M. Jamet, J. M. George, L. Vila, and H. Jaffrès, Spin Pumping and Inverse Spin Hall Effect in Platinum: The Essential Role of Spin -Memory Loss at Metallic Interfaces. Phys. Rev. Lett. 112, 106602 (2014). [14] L. Zhu, D. C. Ralph, and R. A. Buhrman, Spin -Orbit Torques in Heavy -Metal –Ferromagnet Bilayers with Varying Strengths of Interfacial Spin -Orbit Coupling. Phys. Rev. Lett. 122, 077201 (2019). [15] L. J. Zhu and R. A. Buhrman, Maximizing Spin - Orbit -Torque Efficiency of Pt/Ti Multilayers: Trade -Off Between Intrinsic Spin Hall Conductivity and Carrier Lifetime. Phys. Rev. Appl. 12, 6, 051002 (2019). [16] S. Li, K. Shen, and K. Xia, Interfacial spin Hall effect and spin swapping in Fe -Au bilayers from first principles. Phys. Rev. B 99, 134427 (2019). [17] M. Baumgartner, K. Garello, J. Mendil, C. O. Avci, E. Grimaldi, C. Murer, J. Feng, M. Gabureac, C. Stamm, Y . Acremann, S. Finizio, S. Wintz, J. Raab e, and P. Gambardella, Spatially and time -resolved magnetization dynamics driven by spin –orbit torques. Nat. Nanotechnol. 12, 980 (2017). [18] S. Dutta, A. Bose, A. A. Tulapurkar, R. A. Buhrman, and D. C. Ralph, Interfacial and bulk spin Hall contributions to fieldlike spin -orbit torque generated by iridium. Phys. Rev. B 103, 184416 (2021). [19] M. I. Dyakonov and V . I. Perel, Current -induced spin orientation of electrons in semiconductors. Phys. Lett. A 35, 459 (1971). [20] J. E. Hirsch, Spin Hall Effect. Phys. Rev. Lett. 83, 1834 (1999). [21] T. Tanaka, H. Kontani, M. Naito, T. Naito, D. S. Hirashima, K. Yamada, and J. Inoue, Intrinsic spin Hall effect and orbital Hall effect in 4d and 5d transition metals. Phys. Rev. B 77, 165117 (2008). [22] G.-Y . Guo, S . Murakami, T. -W. Chen, and N. Nagaosa, Intrinsic spin Hall effect in platinum: First -principles calculations. Phys. Rev. Lett. 100, 096401 (2008). [23] A. R. Mellnik, J. S. Lee, A. Richardella, J. L. Grab, P. J. Mintun, M. H. Fischer, A. Vaezi, A. Manchon , E. A. Kim, 13 N. Samarth, and D. C. Ralph, Spin -transfer torque generated by a topological insulator. Nature 511, 449 (2014). [24] Y . Fan, P. Upadhyaya, X. Kou, M. Lang, S. Takei, Z. Wang, J. Tang, L. He, L. -T. Chang, M. Montazeri, G. Yu, W. Jiang, T. Nie, R. N. Schwartz, Y . Tserkovnyak, and K. L. Wang, Magnetization switching through giant spin –orbit torque in a magnetically doped topological insulator heterostructure. Nat. Mater. 13, 699 (2014). [25] V . P. Amin, J. Zemen, and M. D. Stiles, Interface - Genera ted Spin Currents. Phys. Rev. Lett. 121, 136805 (2018). [26] L. Wang, R. J. H. Wesselink, Y . Liu, Z. Yuan, K. Xia, and P. J. Kelly, Giant Room Temperature Interface Spin Hall and Inverse Spin Hall Effects. Phys. Rev. Lett. 116, 196602 (2016). [27] J. Borge and I. V . Tokatly, Boundary conditions for spin and charge diffusion in the presence of interfacial spin-orbit coupling. Phys. Rev. B 99, 241401 (2019). [28] S.-h. C. Baek, V . P. Amin, Y . -W. Oh, G. Go, S. -J. Lee, G.-H. Lee, K. -J. Kim, M. D. Stiles, B. -G. Park, and K. -J. Lee, Spin currents and spin –orbit torques in ferromagnetic trilayers. Nat. Mater. 17, 509 (2018). [29] S. Ding, A. Ross, D. Go, L. Baldrati, Z. Ren, F. Freimuth, S. Becker, F. Kammerbauer, J. Yang, G. Jakob, Y . Mokrousov, and M. Kläui, Harn essing Orbital -to-Spin Conversion of Interfacial Orbital Currents for Efficient Spin-Orbit Torques. Phys. Rev. Lett. 125, 177201 (2020). [30] T. Taniguchi, J. Grollier, and M. D. Stiles, Spin - Transfer Torques Generated by the Anomalous Hall Effect and Anis otropic Magnetoresistance. Phys. Rev. Appl. 3, 044001 (2015). [31] A. Bose, D. D. Lam, S. Bhuktare, S. Dutta, H. Singh, Y . Jibiki, M. Goto, S. Miwa, and A. A. Tulapurkar, Observation of Anomalous Spin Torque Generated by a Ferromagnet. Phys. Rev. Appl. 9, 064026 (2018). [32] C. Safranski, E. A. Montoya, and I. N. Krivorotov, Spin–orbit torque driven by a planar Hall current. Nat. Nanotechnol. 14, 27 (2019). [33] C. Safranski, J. Z. Sun, J. -W. Xu, and A. D. Kent, Planar Hall Driven Torque in a Ferromagnet/No nmagnet/Ferromagnet System. Phys. Rev. Lett. 124, 197204 (2020). [34] M. Kimata, H. Chen, K. Kondou, S. Sugimoto, P. K. Muduli, M. Ikhlas, Y . Omori, T. Tomita, A. H. Macdonald, S. Nakatsuji, and Y . Otani, Magnetic and magnetic inverse spin Hall effects in a non -collinear antiferromagnet. Nature 565, 627 (2019). [35] X. R. Wang, Anomalous spin Hall and inverse spin Hall effects in magnetic systems. Commun. Phys. 4, 55 (2021). [36] A. Mook, R. R. Neumann, A. Johansson, J. Henk, and I. Mertig, Origin of the ma gnetic spin Hall effect: Spin current vorticity in the Fermi sea. Physical Review Research 2, 023065 (2020). [37] M. Jiang, H. Asahara, S. Sato, T. Kanaki, H. Yamasaki, S. Ohya, and M. Tanaka, Efficient full spin –orbit torque switching in a single layer of a perpendicularly magnetized single -crystalline ferromagnet. Nat. Commun. 10, 2590 (2019). [38] T. Nan, T. J. Anderson, J. Gibbons, K. Hwang, N. Campbell, H. Zhou, Y . Q. Dong, G. Y . Kim, D. F. Shao, T. R. Paudel, N. Reynolds, X. J. Wang, N. X. Sun, E. Y . Tsymbal, S. Y . Choi, M. S. Rzchowski, Y . B. Kim, D. C. Ralph, and C. B. Eom, Anisotropic spin -orbit torque generation in epitaxial SrIrO 3 by symmetry design. Proceedings of the National Academy of Sciences 116, 16186 (2019). [39] A. Bose, J. N. Nelson, X. S. Zhang, P. Jadaun, R. Jain, D. G. Schlom, D. C. Ralph, D. A. Muller, K. M. Shen, and R. A. Buhrman, Effects of Anisotropic Strain on Spin – Orbit Torque Produced by the Dirac Nodal Line Semimetal IrO 2. ACS Appl. Mater. Interfaces 12, 55411 (2020). [40] S. V . Aradhya, G. E. Rowlands, J. Oh, D. C. Ralph, and R. A. Buhrman, Nanosecond -Timescale Low Energy Switching of In -Plane Magnetic Tunnel Junctions through Dynamic Oersted -Field -Assisted Spin Hall Effect. Nano Lett. 16, 5987 (2016). [41] L. Zhu, L. Zhu, M. Sui, D. C. Ralph, and R. A. Buhrman, Variation of the giant intrinsic spin Hall conductivity of Pt with carrier lifetime. Sci. Adv. 5, eaav8025 (2019). [42] E. Sagasta, Y . Omori, M. Isasa, M. Gradhand, L. E. Hueso, Y . Niimi, Y . Otani, and F. Casano va, Tuning the spin Hall effect of Pt from the moderately dirty to the superclean regime. Phys. Rev. B 94, 060412 (2016). [43] Z. Chi, Y . -C. Lau, X. Xu, T. Ohkubo, K. Hono, and M. Hayashi, The spin Hall effect of Bi -Sb alloys driven by thermally excited Di rac-like electrons. Sci. Adv. 6, eaay2324 (2020). [44] T.-Y . Chen, C. -W. Peng, T. -Y . Tsai, W. -B. Liao, C. -T. Wu, H. -W. Yen, and C. -F. Pai, Efficient Spin –Orbit Torque Switching with Nonepitaxial Chalcogenide Heterostructures. ACS Appl. Mater. Interfaces 12 , 7788 (2020). [45] L. Zhu, X. S. Zhang, D. A. Muller, D. C. Ralph, and R. A. Buhrman, Observation of Strong Bulk Damping -Like Spin-Orbit Torque in Chemically Disordered Ferromagnetic Single Layers. Adv. Funct. Mater. 30, 2005201 (2020). [46] L. Zhu, D. C. Ralph, and R. A. Buhrman, Unveiling the Mechanism of Bulk Spin -Orbit Torques within Chemically Disordered Fe xPt1-x Single Layers. Adv. Funct. Mater. 31, 2103898 (2021). [47] Q. Liu, L. Zhu, X. S. Zhang, D. A. Muller, and D. C. Ralph, Giant bulk spin –orbit torque and efficient electrical switching in single ferrimagnetic FeTb layers with strong perpendicular magnetic anisotropy. Appl. Phys. Rev. 9, 021402 (2022). [48] Y . Hibino, T. Taniguchi, K. Yakushiji, A. Fukushima, H. Kubota, and S. Yuasa, Large Spin -Orbit-Torque Efficiency Generated by Spin Hall Effect in Paramagnetic Co-Ni-Pt Alloys. Phys. Rev. Appl. 14, 064056 (2020). [49] C.-F. Pai, L. Liu, Y . Li, H. W. Tseng, D. C. Ralph, and R. A. Buhrman, Spin transfer torque devices utilizing the giant spin Hall effect of tungsten. Appl. Phys. Lett. 101, 122404 (2012). [50] D. Macneill, G. M. Stiehl, M. H. D. Guimaraes, R. A. Buhrman, J. Park, and D. C. Ralph, Control of spin –orbit torques through crystal symmetry in WTe 2/ferromagnet bilayers. Nat. Phys. 13, 300 (2017). [51] G. M. Stiehl, R. Li, V . Gupta, I. E. Baggari, S. Jiang, H. Xie, L. F. Kourkoutis, K. F. Mak, J. Shan, R. A. Buhrman, and D. C. Ralph, Layer -dependent spin -orbit 14 torques generated by the centrosymmetric transition metal dichalcogenide MoTe. Phy s. Rev. B 100, 184402 (2019). [52] Q. Xie, W. Lin, S. Sarkar, X. Shu, S. Chen, L. Liu, T. Zhao, C. Zhou, H. Wang, J. Zhou, S. Gradečak, and J. Chen, Field -free magnetization switching induced by the unconventional spin –orbit torque from WTe 2. APL Mater. 9, 051114 (2021). [53] I. H. Kao, R. Muzzio, H. Zhang, M. Zhu, J. Gobbo, S. Yuan, D. Weber, R. Rao, J. Li, J. H. Edgar, J. E. Goldberger, J. Yan, D. G. Mandrus, J. Hwang, R. Cheng, J. Katoch, and S. Singh, Deterministic switching of a perpendicularly polariz ed magnet using unconventional spin –orbit torques in WTe 2. Nat. Mater. 21, 1029 (2022). [54] C. K. Safeer, N. Ontoso, J. Ingla -Aynés, F. Herling, V . T. Pham, A. Kurzmann, K. Ensslin, A. Chuvilin, I. Robredo, M. G. Vergniory, F. de Juan, L. E. Hueso, M. R. Calvo, and F. Casanova, Large Multidirectional Spin -to-Charge Conversion in Low -Symmetry Semimetal MoTe 2 at Room Temperature. Nano Lett. 19, 8758 (2019). [55] P. Song, C. -H. Hsu, G. Vignale, M. Zhao, J. Liu, Y . Deng, W. Fu, Y . Liu, Y . Zhang, H. Lin, V . M. Pereira, and K. P. Loh, Coexistence of large conventional and planar spin Hall effect with long spin diffusion length in a low - symmetry semimetal at room temperature. Nat. Mater. 19, 292 (2020). [56] L. Liu, C. Zhou, X. Shu, C. Li, T. Zhao, W. Lin, J. Deng , Q. Xie, S. Chen, J. Zhou, R. Guo, H. Wang, J. Yu, S. Shi, P. Yang, S. Pennycook, A. Manchon, and J. Chen, Symmetry -dependent field -free switching of perpendicular magnetization. Nat. Nanotechnol. 16, 277 (2021). [57] Y . Liu, Y . Liu, M. Chen, S. Srivastav a, P. He, K. L. Teo, T. Phung, S. -H. Yang, and H. Yang, Current -induced Out-of-plane Spin Accumulation on the (001) Surface of the IrMn 3 Antiferromagnet. Phys. Rev. Appl. 12, 064046 (2019). [58] T. Nan, C. X. Quintela, J. Irwin, G. Gurung, D. F. Shao, J. G ibbons, N. Campbell, K. Song, S. Y . Choi, L. Guo, R. D. Johnson, P. Manuel, R. V . Chopdekar, I. Hallsteinsen, T. Tybell, P. J. Ryan, J. W. Kim, Y . Choi, P. G. Radaelli, D. C. Ralph, E. Y . Tsymbal, M. S. Rzchowski, and C. B. Eom, Controlling spin current po larization through non -collinear antiferromagnetism. Nat. Commun. 11, 4671 (2020). [59] Y . You, H. Bai, X. Feng, X. Fan, L. Han, X. Zhou, Y . Zhou, R. Zhang, T. Chen, F. Pan, and C. Song, Cluster magnetic octupole induced out -of-plane spin polarization in antiperovskite antiferromagnet. Nat. Commun. 12, 6524 (2021). [60] X. Chen, S. Shi, G. Shi, X. Fan, C. Song, X. Zhou, H. Bai, L. Liao, Y . Zhou, H. Zhang, A. Li, Y . Chen, X. Han, S. Jiang, Z. Zhu, H. Wu, X. Wang, D. Xue, H. Yang, and F. Pan, Observation of t he antiferromagnetic spin Hall effect. Nat. Mater. 20, 800 (2021). [61] R. González -Hernández, L. Šmejkal, K. Výborný, Y . Yahagi, J. Sinova, T. Jungwirth, and J. Železný, Efficient Electrical Spin Splitter Based on Nonrelativistic Collinear Antiferromagnet ism. Phys. Rev. Lett. 126, 127701 (2021). [62] A. Bose, N. J. Schreiber, R. Jain, D. -F. Shao, H. P. Nair, J. Sun, X. S. Zhang, D. A. Muller, E. Y . Tsymbal, D. G. Schlom, and D. C. Ralph, Tilted spin current generated by the collinear antiferromagnet ruthen ium dioxide. Nat. Electron. 5, 267 (2022). [63] S. Karube, T. Tanaka, D. Sugawara, N. Kadoguchi, M. Kohda, and J. Nitta, Observation of Spin -Splitter Torque in Collinear Antiferromagnetic RuO 2. Phys. Rev. Lett. 129, 137201 (2022). [64] H. Bai, L. Han, X. Y . Feng, Y . J. Zhou, R. X. Su, Q. Wang, L. Y . Liao, W. X. Zhu, X. Z. Chen, F. Pan, X. L. Fan, and C. Song, Observation of Spin Splitting Torque in a Collinear Antiferromagnet RuO 2. Phys. Rev. Lett. 128, 197202 (2022). [65] A. M. Humphries, T. Wang, E. R. J. Edwards, S. R. Allen, J. M. Shaw, H. T. Nembach, J. Q. Xiao, T. J. Silva, and X. Fan, Observation of spin -orbit effects with spin rotation symmetry. Nat. Commun. 8, 911 (2017). [66] H. Kurebayashi, J. Sinova, D. Fang, A. C. Irvine, T. D. Skinner, J. Wunde rlich, V . Novák, R. P. Campion, B. L. Gallagher, E. K. Vehstedt, L. P. Zârbo, K. Výborný, A. J. Ferguson, and T. Jungwirth, An antidamping spin –orbit torque originating from the Berry curvature. Nat. Nanotechnol. 9, 211 (2014). [67] C. Ciccarelli, L. Ander son, V . Tshitoyan, A. J. Ferguson, F. Gerhard, C. Gould, L. W. Molenkamp, J. Gayles, J. Železný, L. Šmejkal, Z. Yuan, J. Sinova, F. Freimuth, and T. Jungwirth, Room -temperature spin –orbit torque in NiMnSb. Nat. Phys. 12, 855 (2016). [68] L. Chen, M. Decker , M. Kronseder, R. Islinger, M. Gmitra, D. Schuh, D. Bougeard, J. Fabian, D. Weiss, and C. H. Back, Robust spin -orbit torque and spin -galvanic effect at the Fe/GaAs (001) interface at room temperature. Nat. Commun. 7, 13802 (2016). [69] T. Wang, S. Lendinez, M. B. Jungfleisch, J. Kolodzey, J. Q. Xiao, and X. Fan, Detection of spin -orbit torque with spin rotation symmetry. Appl. Phys. Lett. 116, 012404 (2020). [70] Y . Hibino, T. Taniguchi, K. Yakushiji, A. Fukushima, H. Kubota, and S. Yuasa, Giant cha rge-to-spin conversion in ferromagnet via spin -orbit coupling. Nat. Commun. 12, 6254 (2021). [71] L. Zhu, D. C. Ralph, and R. A. Buhrman, Maximizing spin-orbit torque generated by the spin Hall effect of Pt. Appl. Phys. Rev. 8, 031308 (2021). [72] K. S. No voselov, A. K. Geim, S. V . Morozov, D. Jiang, Y . Zhang, S. V . Dubonos, I. V . Grigorieva, and A. A. Firsov, Electric Field Effect in Atomically Thin Carbon Films. Science 306, 666 (2004). [73] Z. Y . Zhu, Y . C. Cheng, and U. Schwingenschlögl, Giant spin -orbit-induced spin splitting in two -dimensional transition -metal dichalcogenide semiconductors. Phys. Rev. B 84, 153402 (2011). [74] J. Jiang, F. Tang, X. C. Pan, H. M. Liu, X. H. Niu, Y . X. Wang, D. F. Xu, H. F. Yang, B. P. Xie, F. Q. Song, P. Dudin, T. K. Ki m, M. Hoesch, P. K. Das, I. V obornik, X. G. Wan, and D. L. Feng, Signature of Strong Spin -Orbital Coupling in the Large Nonsaturating Magnetoresistance Material WTe 2. Phys. Rev. Lett. 115, 166601 (2015). [75] Y . Wang, P. Deorani, K. Banerjee, N. Koirala, M . Brahlek, S. Oh, and H. Yang, Topological Surface States Originated Spin -Orbit Torques in Bi 2Se3. Phys. Rev. Lett. 114, 257202 (2015). [76] J. Han, A. Richardella, S. A. Siddiqui, J. Finley, N. Samarth, and L. Liu, Room -Temperature Spin -Orbit Torque Switc hing Induced by a Topological Insulator. Phys. Rev. Lett. 119, 077702 (2017). 15 [77] H. Wu, Y . Xu, P. Deng, Q. Pan, S. A. Razavi, K. Wong, L. Huang, B. Dai, Q. Shao, G. Yu, X. Han, J. -C. Rojas - Sánchez, S. Mangin, and K. L. Wang, Spin -Orbit Torque Switching o f a Nearly Compensated Ferrimagnet by Topological Surface States. Adv. Mater. 31, 1901681 (2019). [78] Y . Deng, Y . Yu, Y . Song, J. Zhang, N. Z. Wang, Z. Sun, Y . Yi, Y . Z. Wu, S. Wu, J. Zhu, J. Wang, X. H. Chen, and Y . Zhang, Gate -tunable room -temperature f erromagnetism in two-dimensional Fe 3GeTe 2. Nature 563, 94 (2018). [79] G. Zheng, W. -Q. Xie, S. Albarakati, M. Algarni, C. Tan, Y . Wang, J. Peng, J. Partridge, L. Farrar, J. Yi, Y . Xiong, M. Tian, Y . -J. Zhao, and L. Wang, Gate -Tuned Interlayer Coupling in v an der Waals Ferromagnet Fe3GeTe 2 Nanoflakes. Phys. Rev. Lett. 125, 047202 (2020). [80] Y . Wang, C. Wang, S. -J. Liang, Z. Ma, K. Xu, X. Liu, L. Zhang, A. S. Admasu, S. -W. Cheong, L. Wang, M. Chen, Z. Liu, B. Cheng, W. Ji, and F. Miao, Strain -Sensitive Magn etization Reversal of a van der Waals Magnet. Adv. Mater. 32, 2004533 (2020). [81] Q. Li, M. Yang, C. Gong, R. V . Chopdekar, A. T. N’Diaye, J. Turner, G. Chen, A. Scholl, P. Shafer, E. Arenholz, A. K. Schmid, S. Wang, K. Liu, N. Gao, A. S. Admasu, S. -W. Ch eong, C. Hwang, J. Li, F. Wang, X. Zhang, and Z. Qiu, Patterning -Induced Ferromagnetism of Fe3GeTe 2 van der Waals Materials beyond Room Temperature. Nano Lett. 18, 5974 (2018). [82] C. Gong, L. Li, Z. Li, H. Ji, A. Stern, Y . Xia, T. Cao, W. Bao, C. Wang, Y . Wang, Z. Q. Qiu, R. J. Cava, S. G. Louie, J. Xia, and X. Zhang, Discovery of intrinsic ferromagnetism in two -dimensional van der Waals crystals. Nature 546, 265 (2017). [83] B. Huang, G. Clark, E. Navarro -Moratalla, D. R. Klein, R. Cheng, K. L. Seyler, D . Zhong, E. Schmidgall, M. A. McGuire, D. H. Cobden, W. Yao, D. Xiao, P. Jarillo - Herrero, and X. Xu, Layer -dependent ferromagnetism in a van der Waals crystal down to the monolayer limit. Nature 546, 270 (2017). [84] W. Zhang, J. Sklenar, B. Hsu, W. Jiang, M. B. Jungfleisch, J. Xiao, F. Y . Fradin, Y . Liu, J. E. Pearson, J. B. Ketterson, Z. Yang, and A. Hoffmann, Research Update: Spin transfer torques in permalloy on monolayer MoS 2. APL Mater. 4, 032302 (2016). [85] Q. Shao, G. Yu, Y . -W. Lan, Y . Shi, M. -Y . Li, C. Zheng, X. Zhu, L. -J. Li, P. K. Amiri, and K. L. Wang, Strong Rashba -Edelstein Effect -Induced Spin –Orbit Torques in Monolayer Transition Metal Dichalcogenide/Ferromagnet Bilayers. Nano Lett. 16, 7514 (2016). [86] W. Lv, Z. Jia, B. Wang, Y . Lu, X. Luo, B. Zhang, Z. Zeng, and Z. Liu, Electric -Field Control of Spin –Orbit Torques in WS 2/Permalloy Bilayers. ACS Appl. Mater. Interfaces 10, 2843 (2018). [87] L. A. Benítez, W. Savero Torres, J. F. Sierra, M. Timmermans, J. H. Garcia, S. Roche, M. V . Costache, and S. O. Valenzuela, Tunable room -temperature spin galvanic and spin Hall effects in van der Waals heterostructures. Nat. Mater. 19, 170 (2020). [88] S. Husain, X. Chen, R. Gupta, N. Behera, P. Kumar, T. Edvinsson, F. García -Sánchez, R. Brucas, S. Chaudha ry, B. Sanyal, P. Svedlindh, and A. Kumar, Large Damping - Like Spin –Orbit Torque in a 2D Conductive 1T -TaS 2 Monolayer. Nano Lett. 20, 6372 (2020). [89] M. H. D. Guimarães, G. M. Stiehl, D. Macneill, N. D. Reynolds, and D. C. Ralph, Spin –Orbit Torques in NbSe2/Permalloy Bilayers. Nano Lett. 18, 1311 (2018). [90] P. Li, W. Wu, Y . Wen, C. Zhang, J. Zhang, S. Zhang, Z. Yu, S. A. Yang, A. Manchon, and X. -X. Zhang, Spin - momentum locking and spin -orbit torques in magnetic nano -heterojunctions composed of Weyl semim etal WTe 2. Nat. Commun. 9, 3990 (2018). [91] S. Shi, S. Liang, Z. Zhu, K. Cai, S. D. Pollard, Y . Wang, J. Wang, Q. Wang, P. He, J. Yu, G. Eda, G. Liang, and H. Yang, All -electric magnetization switching and Dzyaloshinskii –Moriya interaction in WTe 2/ferroma gnet heterostructures. Nat. Nanotechnol. 14, 945 (2019). [92] H. Xu, J. Wei, H. Zhou, J. Feng, T. Xu, H. Du, C. He, Y . Huang, J. Zhang, Y . Liu, H. -C. Wu, C. Guo, X. Wang, Y . Guang, H. Wei, Y . Peng, W. Jiang, G. Yu, and X. Han, High Spin Hall Conductivity i n Large -Area Type -II Dirac Semimetal PtTe 2. Adv. Mater. 32, 2000513 (2020). [93] S. Liang, S. Shi, C. -H. Hsu, K. Cai, Y . Wang, P. He, Y . Wu, V . M. Pereira, and H. Yang, Spin -Orbit Torque Magnetization Switching in MoTe 2/Permalloy Heterostructures. Adv. Mat er. 32, 2002799 (2020). [94] Y . Ou, W. Yanez, R. Xiao, M. Stanley, S. Ghosh, B. Zheng, W. Jiang, Y . -S. Huang, T. Pillsbury, A. Richardella, C. Liu, T. Low, V . H. Crespi, K. A. Mkhoyan, and N. Samarth, ZrTe 2/CrTe 2: an epitaxial van der Waals platform for sp intronics. Nat. Commun. 13, 2972 (2022). [95] Y.-Y . Lv, X. Li, B. -B. Zhang, W. Y . Deng, S. -H. Yao, Y . B. Chen, J. Zhou, S. -T. Zhang, M. -H. Lu, L. Zhang, M. Tian, L. Sheng, and Y . -F. Chen, Experimental Observation of Anisotropic Adler -Bell-Jackiw Anomaly in Type -II Weyl Semimetal WTe 1.98 Crystals at the Quasiclassical Regime. Phys. Rev. Lett. 118, 096603 (2017). [96] D. MacNeill, G. M. Stiehl, M. H. D. Guimarães, N. D. Reynolds, R. A. Buhrman, and D. C. Ralph, Thickness dependence of spin -orbit torques gener ated by WTe 2. Phys. Rev. B 96, 054450 (2017). [97] C.-W. Peng, W. -B. Liao, T. -Y . Chen, and C. -F. Pai, Efficient Spin -Orbit Torque Generation in Semiconducting WTe 2 with Hopping Transport. ACS Appl. Mater. Interfaces 13, 15950 (2021). [98] X. Li, P. Li, V . D. H. Hou, M. Dc, C. -H. Nien, F. Xue, D. Yi, C. Bi, C. -M. Lee, S. -J. Lin, W. Tsai, Y . Suzuki, and S. X. Wang, Large and robust charge -to-spin conversion in sputter ed conductive WTe x with disorder. Matter 4, 1639 (2021). [99] H. Xie, A. Talapatra, X. Chen, Z. Luo, and Y . Wu, Large damping -like spin –orbit torque and perpendicular magnetization switching in sputtered WTe x films. Appl. Phys. Lett. 118, 042401 (2021). [100] Q. Liu and L. Zhu, Current -induced perpendicular effective magnetic field in magnetic heterostructures. Appl. Phys. Rev. 9, 041401 (2022). [101] P. Tsipas, D. Tsoutsou, S. Fragkos, R. Sant, C. Alvarez, H. Okuno, G. Renaud, R. Alcotte, T. Baron, and A. Dimoulas, Massless Dirac Fermions in ZrTe 2 Semimetal Grown on InAs(111) by van der Waals Epitaxy. ACS Nano 12, 1696 (2018). [102] Y . Zhang, K. He, C. -Z. Chang, C. -L. Song, L. -L. Wang, X. Chen, J. -F. Jia, Z. Fang, X. Dai, W. -Y . Shan, S. - Q. Shen, Q. Niu, X. -L. Qi, S. -C. Zhang, X. -C. Ma, and Q. - K. Xue, Crossover of the three -dimensional topological 16 insulator Bi 2Se3 to the two -dimensional limit. Nat. Phys. 6, 584 (2010). [103] J. Han and L. Liu, Topological insulators for efficient spin–orbit torques. APL Mater . 9, 060901 (2021). [104] Q. L. He, T. L. Hughes, N. P. Armitage, Y . Tokura, and K. L. Wang, Topological spintronics and magnetoelectronics. Nat. Mater. 21, 15 (2022). [105] H. Wu, P. Zhang, P. Deng, Q. Lan, Q. Pan, S. A. Razavi, X. Che, L. Huang, B. Dai, K. Wong, X. Han, and K. L. Wang, Room -Temperature Spin -Orbit Torque from Topological Surface States. Phys. Rev. Lett. 123, 207205 (2019). [106] N. H. D. Khang, S. Nakano, T. Shirokura, Y . Miyamoto, and P. N. Hai, Ultralow power spin –orbit torque magne tization switching induced by a non -epitaxial topological insulator on Si substrates. Sci. Rep. 10, 12185 (2020). [107] J. C. Y . Teo, L. Fu, and C. L. Kane, Surface states and topological invariants in three -dimensional topological insulators: Application to Bi 1-xSbx. Phys. Rev. B 78, 045426 (2008). [108] X. Wang, J. Tang, X. Xia, C. He, J. Zhang, Y . Liu, C. Wan, C. Fang, C. Guo, W. Yang, Y . Guang, X. Zhang, H. Xu, J. Wei, M. Liao, X. Lu, J. Feng, X. Li, Y . Peng, H. Wei, R. Yang, D. Shi, X. Zhang, Z. Han, Z . Zhang, G. Zhang, G. Yu, and X. Han, Current -driven magnetization switching in a van der Waals ferromagnet Fe 3GeTe 2. Sci. Adv. 5, eaaw8904 (2019). [109] M. Alghamdi, M. Lohmann, J. Li, P. R. Jothi, Q. Shao, M. Aldosary, T. Su, B. P. T. Fokwa, and J. Shi, Highly Efficient Spin –Orbit Torque and Switching of Layered Ferromagnet Fe 3GeTe 2. Nano Lett. 19, 4400 (2019). [110] W. Zhu, C. Song, L. Han, H. Bai, Q. Wang, S. Yin, L. Huang, T. Chen, and F. Pan, Interface -Enhanced Ferromagnetism with Long -Distance Effect in van der Waals Semiconductor. Adv. Funct. Mater. 32, 2108953 (2022). [111] V . Gupta, T. M. Cham, G. M. Stiehl, A. Bose, J. A. Mittelstaedt, K. Kang, S. Jiang, K. F. Mak, J. Shan, R. A. Buhrman, and D. C. Ralph, Manipulation of the van der Waals Magnet C r2Ge2Te6 by Spin –Orbit Torques. Nano Lett. 20, 7482 (2020). [112] X. Liu, D. Wu, L. Liao, P. Chen, Y . Zhang, F. Xue, Q. Yao, C. Song, K. L. Wang, and X. Kou, Temperature dependence of spin —orbit torque -driven magnetization switching in in situ grown Bi 2Te3/MnTe heterostructures. Appl. Phys. Lett. 118, 112406 (2021). [113] M. Dc, R. Grassi, J. -Y . Chen, M. Jamali, D. Reifsnyder Hickey, D. Zhang, Z. Zhao, H. Li, P. Quarterman, Y . Lv, M. Li, A. Manchon, K. A. Mkhoyan, T. Low, and J. -P. Wang, Room -temperature hi gh spin –orbit torque due to quantum confinement in sputtered Bi xSe(1–x) films. Nat. Mater. 17, 800 (2018). [114] K. Kondou, R. Yoshimi, A. Tsukazaki, Y . Fukuma, J. Matsuno, K. S. Takahashi, M. Kawasaki, Y . Tokura, and Y . Otani, Fermi -level -dependent charge -to-spin current conversion by Dirac surface states of topological insulators. Nat. Phys. 12, 1027 (2016). [115] N. H. D. Khang, Y . Ueda, and P. N. Hai, A conductive topological insulator with large spin Hall effect for ultralow power spin –orbit torque swi tching. Nat. Mater. 17, 808 (2018). [116] Z. Zheng, Y . Zhang, D. Zhu, K. Zhang, X. Feng, Y . He, L. Chen, Z. Zhang, D. Liu, Y . Zhang, P. K. Amiri, and W. Zhao, Perpendicular magnetization switching by large spin–orbit torques from sputtered Bi 2Te3. Chin. Ph ys. B 29, 078505 (2020). [117] N. D. Mermin and H. Wagner, Absence of Ferromagnetism or Antiferromagnetism in One - or Two - Dimensional Isotropic Heisenberg Models. Phys. Rev. Lett. 17, 1133 (1966). [118] Y . Zhang, H. Xu, C. Yi, X. Wang, Y . Huang, J. Tang, J. Jiang, C. He, M. Zhao, T. Ma, J. Dong, C. Guo, J. Feng, C. Wan, H. Wei, H. Du, Y . Shi, G. Yu, G. Zhang, and X. Han, Exchange bias and spin –orbit torque in the Fe 3GeTe 2- based heterostructures prepared by vacuum exfoliation approach. Appl. Phys. Lett. 118, 262406 (2021). [119] R. Zhu, W. Zhang, W. Shen, P. K. J. Wong, Q. Wang, Q. Liang, Z. Tian, Y . Zhai, C. -w. Qiu, and A. T. S. Wee, Exchange Bias in van der Waals CrCl 3/Fe 3GeTe 2 Heterostructures. Nano Lett. 20, 5030 (2020). [120] L. Zhang, X. Huang, H. Dai, M. Wang, H. Cheng, L. Tong, Z. Li, X. Han, X. Wang, L. Ye, and J. Han, Proximity -Coupling -Induced Significant Enhancement of Coercive Field and Curie Temperature in 2D van der Waals Heterostructures. Adv. Mater. 32, 2002032 (2020). [121] C. Kittel, Theory of Antiferromagnetic Resonance. Phys. Rev. 82, 565 (1951). [122] C. D. Stanciu, A. V . Kimel, F. Hansteen, A. Tsukamoto, A. Itoh, A. Kirilyuk, and T. Rasing, Ultrafast spin dynamics across compensation points in ferrimagnetic GdFeCo: The role of angular mom entum compensation. Phys. Rev. B 73, 220402 (2006). [123] Y . Xu, M. Deb, G. Malinowski, M. Hehn, W. Zhao, and S. Mangin, Ultrafast Magnetization Manipulation Using Single Femtosecond Light and Hot -Electron Pulses. Adv. Mater. 29, 1703474 (2017). [124] V . Baltz, A. Manchon, M. Tsoi, T. Moriyama, T. Ono, and Y . Tserkovnyak, Antiferromagnetic spintronics. Rev. Mod. Phys. 90, 015005 (2018). [125] P. Qin, H. Yan, X. Wang, H. Chen, Z. Meng, J. Dong, M. Zhu, J. Cai, Z. Feng, X. Zhou, L. Liu, T. Zhang, Z. Zeng, J. Zhang, C. Jiang, and Z. Liu, Room -temperature magnetoresistance in an all -antiferromagnetic tunnel junction. Nature 613, 485 (2023). [126] X. Chen, T. Higo, K. Tanaka, T. Nomoto, H. Tsai, H. Idzuchi, M. Shiga, S. Sakamoto, R. Ando, H. Kosaki, T. Matsuo, D. Nishio -Hamane, R. Arita, S. Miwa, and S. Nakatsuji, Octupole -driven magnetoresistance in an antiferromagnetic tunnel junction. Nature 613, 490 (2023). [127] Z. Zhang, J. Shang, C. Jiang, A. Rasmita, W. Gao, and T. Yu, Direct Photoluminescence Pro bing of Ferromagnetism in Monolayer Two -Dimensional CrBr 3. Nano Lett. 19, 3138 (2019). [128] W. Chen, Z. Sun, Z. Wang, L. Gu, X. Xu, S. Wu, and C. Gao, Direct observation of van der Waals stacking – dependent interlayer magnetism. Science 366, 983 (2019). [129] C. Chen, X. Chen, C. Wu, X. Wang, Y . Ping, X. Wei, X. Zhou, J. Lu, L. Zhu, J. Zhou, T. Zhai, J. Han, and H. Xu, Air-Stable 2D Cr 5Te8 Nanosheets with Thickness -Tunable Ferromagnetism. Adv. Mater. 34, 2107512 (2022). [130] Z. Lin, B. Huang, K. Hwangbo, Q. Jiang, Q. Zhang, Z. Liu, Z. Fei, H. Lv, A. Millis, M. McGuire, D. Xiao, J. - H. Chu, and X. Xu, Magnetism and Its Structural Coupling 17 Effects in 2D Ising Ferromagnetic Insulator VI 3. Nano Lett. 21, 9180 (2021). [131] L. Kang, C. Ye, X. Zhao, X. Z hou, J. Hu, Q. Li, D. Liu, C. M. Das, J. Yang, D. Hu, J. Chen, X. Cao, Y . Zhang, M. Xu, J. Di, D. Tian, P. Song, G. Kutty, Q. Zeng, Q. Fu, Y . Deng, J. Zhou, A. Ariando, F. Miao, G. Hong, Y . Huang, S. J. Pennycook, K. -T. Yong, W. Ji, X. Renshaw Wang, and Z. Liu, Phase -controllable growth of ultrathin 2D magnetic FeTe crystals. Nat. Commun. 11, 3729 (2020). [132] J. Seo, D. Y . Kim, E. S. An, K. Kim, G. -Y . Kim, S. - Y . Hwang, D. W. Kim, B. G. Jang, H. Kim, G. Eom, S. Y . Seo, R. Stania, M. Muntwiler, J. Lee, K. W atanabe, T. Taniguchi, Y . J. Jo, J. Lee, B. I. Min, M. H. Jo, H. W. Yeom, S.-Y . Choi, J. H. Shim, and J. S. Kim, Nearly room temperature ferromagnetism in a magnetic metal -rich van der Waals metal. Sci. Adv. 6, eaay8912 (2019). [133] A. F. May, D. Ovchinni kov, Q. Zheng, R. Hermann, S. Calder, B. Huang, Z. Fei, Y . Liu, X. Xu, and M. A. McGuire, Ferromagnetism Near Room Temperature in the Cleavable van der Waals Crystal Fe 5GeTe 2. ACS Nano 13, 4436 (2019). [134] H. Wu, W. Zhang, L. Yang, J. Wang, J. Li, L. Li, Y . Gao, L. Zhang, J. Du, H. Shu, and H. Chang, Strong intrinsic room -temperature ferromagnetism in freestanding non-van der Waals ultrathin 2D crystals. Nat. Commun. 12, 5688 (2021). [135] D. C. Freitas, R. Weht, A. Sulpice, G. Remenyi, P. Strobel, F. Gay , J. Marcus, and M. Núñez -Regueiro, Ferromagnetism in layered metastable 1T -CrTe 2. J. Phys. Condens. Matter 27, 176002 (2015). [136] X. Sun, W. Li, X. Wang, Q. Sui, T. Zhang, Z. Wang, L. Liu, D. Li, S. Feng, S. Zhong, H. Wang, V . Bouchiat, M. Nunez Regueir o, N. Rougemaille, J. Coraux, A. Purbawati, A. Hadj -Azzem, Z. Wang, B. Dong, X. Wu, T. Yang, G. Yu, B. Wang, Z. Han, X. Han, and Z. Zhang, Room temperature ferromagnetism in ultra -thin van der Waals crystals of 1T - CrTe 2. Nano Res. 13, 3358 (2020). [137] X. Zhang, Q. Lu, W. Liu, W. Niu, J. Sun, J. Cook, M. Vaninger, P. F. Miceli, D. J. Singh, S. -W. Lian, T. -R. Chang, X. He, J. Du, L. He, R. Zhang, G. Bian, and Y . Xu, Room - temperature intrinsic ferromagnetism in epitaxial CrTe 2 ultrathin films. Nat. Commun. 1 2, 2492 (2021). [138] A. L. Coughlin, D. Xie, X. Zhan, Y . Yao, L. Deng, H. Hewa -Walpitage, T. Bontke, C. -W. Chu, Y . Li, J. Wang, H. A. Fertig, and S. Zhang, Van der Waals Superstructure and Twisting in Self -Intercalated Magnet with Near Room - Temperature Pe rpendicular Ferromagnetism. Nano Lett. 21, 9517 (2021). [139] Y . Wen, Z. Liu, Y . Zhang, C. Xia, B. Zhai, X. Zhang, G. Zhai, C. Shen, P. He, R. Cheng, L. Yin, Y . Yao, M. Getaye Sendeku, Z. Wang, X. Ye, C. Liu, C. Jiang, C. Shan, Y . Long, and J. He, Tunable Room -Temperature Ferromagnetism in Two -Dimensional Cr 2Te3. Nano Lett. 20, 3130 (2020). [140] R. Chua, J. Zhou, X. Yu, W. Yu, J. Gou, R. Zhu, L. Zhang, M. Liu, M. B. H. Breese, W. Chen, K. P. Loh, Y . P. Feng, M. Yang, Y . L. Huang, and A. T. S. Wee, Room Temperature Ferromagnetism of Monolayer Chromium Telluride with Perpendicular Magnetic Anisotropy. Adv. Mater. 33, 2103360 (2021). [141] Y . Zhang, J. Chu, L. Yin, T. A. Shifa, Z. Cheng, R. Cheng, F. Wang, Y . Wen, X. Zhan, Z. Wang, and J. He, Ultrathin Magneti c 2D Single -Crystal CrSe. Adv. Mater. 31, 1900056 (2019). [142] G. Zhang, F. Guo, H. Wu, X. Wen, L. Yang, W. Jin, W. Zhang, and H. Chang, Above -room -temperature strong intrinsic ferromagnetism in 2D van der Waals Fe 3GaTe 2 with large perpendicular magnetic anisotropy. Nat. Commun. 13, 5067 (2022). [143] V . Ostwal, T. Shen, and J. Appenzeller, Efficient Spin-Orbit Torque Switching of the Semiconducting Van Der Waals Ferromagnet Cr 2Ge2Te6. Adv. Mater. 32, 1906021 (2020). [144] M. Mogi, K. Yasuda, R. Fujimura, R. Yoshimi, N. Ogawa, A. Tsukazaki, M. Kawamura, K. S. Takahashi, M. Kawasaki, and Y . Tokura, Current -induced switching of proximity -induced ferromagnetic surface states in a topological insulator. Nat. Commun. 12, 1404 (2021). [145] R. Fujimura, R. Yoshim i, M. Mogi, A. Tsukazaki, M. Kawamura, K. S. Takahashi, M. Kawasaki, and Y . Tokura, Current -induced magnetization switching at charge -transferred interface between topological insulator (Bi,Sb) 2Te3 and van der Waals ferromagnet Fe 3GeTe 2. Appl. Phys. Lett. 119, 032402 (2021). [146] I. Shin, W. J. Cho, E. -S. An, S. Park, H. -W. Jeong, S. Jang, W. J. Baek, S. Y . Park, D. -H. Yang, J. H. Seo, G. -Y . Kim, M. N. Ali, S. -Y . Choi, H. -W. Lee, J. S. Kim, S. D. Kim, and G. -H. Lee, Spin –Orbit Torque Switching in an All-Van der Waals Heterostructure. Adv. Mater. 34, 2101730 (2022). [147] D. R. Klein, D. MacNeill, Q. Song, D. T. Larson, S. Fang, M. Xu, R. A. Ribeiro, P. C. Canfield, E. Kaxiras, R. Comin, and P. Jarillo -Herrero, Enhancement of interlayer exchange in an ultrat hin two -dimensional magnet. Nat. Phys. 15, 1255 (2019). [148] D. J. O’Hara, T. Zhu, A. H. Trout, A. S. Ahmed, Y . K. Luo, C. H. Lee, M. R. Brenner, S. Rajan, J. A. Gupta, D. W. McComb, and R. K. Kawakami, Room Temperature Intrinsic Ferromagnetism in Epitaxi al Manganese Selenide Films in the Monolayer Limit. Nano Lett. 18, 3125 (2018). [149] K.-S. Lee, S. -W. Lee, B. -C. Min, and K. -J. Lee, Threshold current for switching of a perpendicular magnetic layer induced by spin Hall effect. Appl. Phys. Lett. 102, 1124 10 (2013). [150] K.-S. Lee, S. -W. Lee, B. -C. Min, and K. -J. Lee, Thermally activated switching of perpendicular magnet by spin-orbit spin torque. Appl. Phys. Lett. 104, 072413 (2014). [151] A. Thiaville, S. Rohart, É. Jué, V . Cros, and A. Fert, Dynamics of Dzyaloshinskii domain walls in ultrathin magnetic films. Europhysics Letters 100, 57002 (2012). [152] D. Bhowmik, M. E. Nowakowski, L. You, O. Lee, D. Keating, M. Wong, J. Bokor, and S. Salahuddin, Deterministic Domain Wall Motion Orthogonal To Current Flow Due To Spin Orbit Torque. Sci. Rep. 5, 11823 (2015). [153] T.-Y . Chen, C. -T. Wu, H. -W. Yen, and C. -F. Pai, Tunable spin -orbit torque in Cu -Ta binary alloy heterostructures. Phys. Rev. B 96, 104434 (2017). [154] L. Zhu, D. C. Ralph, and R. A. Buhrman, Lack of Simple Correlation between Switching Current Density and Spin -Orbit -Torque Efficiency of Perpendicularly Magnetized Spin -Current -Generator -Ferromagnet Heterostructures. Phys. Rev. Appl. 15, 024059 (2021).
1606.08334v2.Nonreciprocal_Transverse_Photonic_Spin_and_Magnetization_Induced_Electromagnetic_Spin_Orbit_Coupling.pdf
1 NONRECIPROCAL TRANSVERSE PHOTONIC SPIN AND MAGNETIZATION -INDUCED ELECTROMAGNETIC SPIN -ORBIT COUPLING Miguel Levy and Dolendra Karki Physics Department, Michigan Technological University Henes Center for Quantum Phenomena ABSTRACT A study of nonreciprocal transverse -spin a ngular -momentum -density shift s for evanescent waves in magneto -optic waveguide media is presented. Their functional relation to electromagnetic spin- and orbital -momenta is presented and analyzed. It is shown that the magneto -optic gyrotropy can be re -interpreted as the nonreciprocal electromagnetic spin-density shift per unit energy flux , thus providing an interesting alternative physical picture for the magneto -optic gyrotropy . The transverse spin -density shift is fou nd to be thickness -dependent in slab optical waveguides. This dependence is traceable t o the admixture of minority helicity component s in the transverse spin angular momentum. It is also shown that the transverse spin is magnetically tunable . A formulation of electromagnetic spin-orbit coupling in magneto -optic media is presented, and an alternative source of spin -orbit coupling to non -paraxial optics vortices is proposed. It is shown that magnetization -induced electromagnetic spin -orbit coupli ng is possible, and that it leads to spin to orbit al angular momentum conversion in magneto -optic media evanescent waves . INTRODUCTION In 1939, F. J. Beli nfante introduced a spin momentum density expression for vector fields to explain the spin of quantum part icles and symmetrize the energy -momentum tensor [ 1]. For monochromatic electromagnetic waves in free -space , the corresponding time-averaged spin momentum density reads 1 2BBps , (Eq. 1) and the spin angular momentum density is * 1Im ( )2Bos E E . (Eq. 2 ) is the optical frequency and 0 the permittivity of free -space [2]. This spin angular momentum , in its transverse electromagnetic form , has merited much attention in recent years, as it can be studied in evanescent waves [ 3-7]. There are fundamental and practical reasons for this. Until recently, the quantum field theory of the electromagnetic field has lacked a description of separate local conservation laws for the spin and orbital angular momentum -generating currents [7]. Whether such spin -generating momenta , as opposed to the actual spin angular momenta they induce, are indeed observable or merely ‘virtual’ is of fundamental interest . Moreover, if the electromagnetic spin and orbital momenta are separable, the question arises as to whether there are any photonic spin -orbit interaction effects . Bliokh and co -workers give a positive answer for non - paraxial fields. [7] Using the conservation laws proposed by these authors, we show that it is possible to magnetically induce electromagnetic spin -orbit coupling in magneto -optic media. 2 We know that the transverse optical spin is a physically meaningful quantity that can be transferred to material particles [3-8]. This has potential ly appealing consequences for optical - tweezer particle manipulation , or to locate and track nanoparticles with a high degree of te mporal and spatial resolution [9 ]. Thus, developing means of control for the transverse optical spin is of practical interest. In this paper, w e address the latter question for both spin momenta Bp and angular momenta Bs . We show that the ir magnitude s and sense of circulation can be accessed and controlled in a single structure , and propose a specific configuration to this end . Explicit expressions for these physical quantities and for the spin -orbit coupling are presented. Moreover, we develop our treatment for nonreciprocal slab optical waveguides , resulting in a different response upon time reversal s. We consider the behavior of evanescent waves in transverse -magnetic (TM) modes in magnetic garnet claddings on silicon -on-insulator guides. This allows us to obtain explicit expressions for the transverse Beli nfante spin momenta and angular momenta and to propose a means for magnetically controlling these objects, with potential applicabilit y to nanoparticle manipulation. MAGNETIC -GYROTROPY -DEPENDENT EVANESCENT WAVES Consider a silicon -on-insulator slab waveguide with iron garnet top cladding , as in Fig. 1 . The off-diagonal component s of the garnet’s dielectric permittivity tensor contr ol the structure’s magneto -optic response. Infrared 1550nm wavelength light propagates in the slab, in the presence of a magnetic field transverse to the dir ection of propagation . The electromagnetic field-expressions in the top cladding for transverse magnetization (y- direction) and monochromatic light propagating in the z -direction are, () ()( , )e ( , )i z t o i z t oE E x y H H x y e (Eq. 3) Maxwell -Ampere’s and Faraday’s laws in ferrimagnetic media are 00 ˆ 0 0 0 0 00cc o o c o c ccig ig EEH i Ettig ig (Eq. 4) ooHE i Ht (Eq. 5) The off -diagonal component of the dielectric permittivity tensor ˆ is the gyrotropy parameter , parameterized by g. We examine transverse -magnetic (TM) propagation in the slab. Vertical and transverse - horizontal directions are x, and y, respectively, is the propagation constant, and the wave equation in the iron garnet is given by 3 22 22 20y o c y cgH k Hx , with 02k , for wavelength [10, 11 ]. (Eq. 6) Defining 2 eff c cg as an effective permittivity in the cover layer, we get: ,0effx ycH H e x (Top cladding ) (Eq. 7) cos( ), x 0y f x cH H k x d (Core ) (Eq. 8) exp( ( )),y s sH H x d x d , (Substrate) (Eq. 9) where 22 eff o eff k , (Eq. 10) 22 x o fkk , (Eq. 11) 22 s o s k (Eq. 12) f , and s are the silicon -slab and substrate dielectric -permittivity constants, respectively, and d is the slab thickness. Solving for the electric -field components in the top cladding layer, we get, 22 0 22 0() ()c eff zy c c eff xy cgE i Hg gEHg (Eq. 13 ) Notice that these two electric field components are /2 out of phase, hence the polarization is elliptical in the cover laye r, with optical spin transverse to the propagation direction. In addition, the polarization evinces opposite helicities for counter -propagating beam s, as /zxEE changes sign upon propagation direction reversal. This result already contains an important difference with reciprocal non-gyrotropic formulations, where //zxE E i , and the decay constant in the top cladding. E quation 13 depends on the gyrotropy parameter g, both explicitly and implicitly through , and is therefore magnetically tunable, as we shall see below. We emphasize that the magnitude and sign of the propagation constant change upon propagation direction reversal, and , separately, upon magnetization direction reversal. The difference between forward and backward propagation constants is also gyrotropy dependent. This nonreciprocal quality of magneto -optic waveguides is central to the proper functioning of certain on-chip devices, such as Mach -Zehnder -based optical isolat ors [10, 11 ]. As pointed out before, Eq. 2 applies to free-space Maxwell electromagnetism. In a dielectric medium, the momentum density expression must account for the electronic response to the optical 4 wave. Minkowski’s and Abraham’s formulations describe the canonical and the kinetic electromagnetic momenta , respectively [12]. Here we w ill focus on Minkowski’s version , p D B , as it is intimately linked to the generation of translations in the host medium, and hence to optical phase shifts , of interest in nonreciprocal phenomena. D is the dis placement vector, and B the magnetic flux density. Dual -symmetric versions of electromagnetic field theory in free space have been considered by various authors [ 2, 7, 8, 13]. However, t he interaction of light and matter at the local level often has an electric character. Dielectric probe particles will generally sense the electric part of the electromagnetic momentum and spin densities [2, 7, 8, 13]. Hence, we treat the standard (electric - biased) formulation of the electromagnetic spin and orbital angular momenta. In the presence of dielectric media , such as iron garnets in the near -infrared range , the expression for spin angular momentum become s * , Im ( )2o BMs E E . (Eq. 14) The orbital momentum is *Im ( ( ) )2Oop E E , where (Eq. 15) ()x x y y z z X Y X Y X Y X Y , and is the relative dielectric permittivity of the medium [4, a, D, E] . In magneto -optic media, the dielectric permittivity is cg , depending on the helicity of the propagating transverse circular polarization. This is usually a small correction to c , as g is two -, or three -, orders of magnitude smaller in iron garnets , in the near infrared range . For elliptical spins, where one he licity component dominates, we account for the admixture level of the minority component in through a weighted average . NONRECIPROCAL ELECTROMAGNETIC TRANSVERSE SPIN ANGULAR MOMENTUM AND SPIN -ORBIT COUPLING 1. Transverse Spin Momentum and Angular M omentum Densities in Non -Reciprocal Media In thi s section we present a formulation for the transverse -spin momentum and angular momentum densities , as well as the orbi tal angular momentum density, induced by evanescent fields in nonreciprocal magneto -optic media . The magnitude and tuning range of these objects in terms of waveguide geometry and optical gyrotropy are expounded and discussed. We detail t heir unequal response to given optical energy fluxe s in opposite propagation directions and to changes in applied magnetic fields. And we apply the recently proposed Bliokh -Dressel -Nori electromagnetic spin-orbit correction term to calculate the spin -orbit interaction for evanescent waves in gyrotropic media [7]. Equation (13), together with Eq. (14) and Eq. (15) yield the following expressions for the transv erse Belinfante -Minkowski spin angular momentum, spin momentum and the orbital momentum densities in evanescent nonreciprocal electromagnetic waves, 5 2 , 3 2 2 2 2ˆc eff c eff B M y o c cggs H y gg (Eq. 16) 2 , 3 2 2 2 2ˆeff c eff c eff B M y o c cggp H z gg (Eq. 17) 22 2 3 2 2 2 2ˆ 2c eff c eff Oy o c cggp H z gg (Eq.18 ) And the ratio , 2c eff c eff O B M c eff c effgg p s g g (Eq. 19) These expression s depend on the magneto -optic gyrotropy parameter g and the dielectric permittivity of the waveguide core channel and of its cover layer under transverse magnetization. They yield different values under magnetic field tuning, magnetization and beam propagation direction reversals, and as a function of waveguide core thickness as discussed below . The propagation constant is gyrotropy -, propagation -direction -, and waveguide -core-thickness - dependent, and this behavior strongly impacts the electromag netic spin and orbital momenta. The time -averaged electromagnetic energy flux (Poynting’s vector) in the iron garnet layer is 2* 1 2 22 01ˆ Re( )2 ()c eff y cgS E H H z g . (Eq. 20) Re-expressing the transverse Belinfante -Minkowski spin angular momentum and spin momentum densities in terms of the energy flow S , , 2 2 22ˆc eff BM cgs S y g (Eq. 21) , 2 2 22eff c eff BM cgpS g (Eq. 22) Figure 1 plots the nonreciprocal Belinfante -Minkowski transverse spin-angular -momentum - density shift per unit energy flux, as a function of silicon slab thickness in an SOI slab waveguide with Ce 1Y2Fe5O12 garnet top cladding . Calculations are performed for the same electromagnetic energy flux in opposite propagation directions, at a wavel ength of 1550 nm, 0.0086 g . The nonreciprocal shift is normalized to the average spin angular momentum, as follows, 6 ,,2f c eff b c efffb B M S f c eff b c efffbgg s gg . (Eq. 23) Subscripts f and b stand for forward, and backward propagation , respectively. This expression evinces a relatively stable value , close to 0.7% above 0 .3m thickness. What is the explanation for this? It has to do with the ellipticity of the transverse polarization in the x -z plane. Above 0.3m, the ellipticity ranges from 31.4° to 36.9°, where 45° corresponds to circular polarization. In other words, the ellipticity stays fairly constants, with a moderately small admixture of the minority circularly polarized component , ranging from 25% to 14%. Below 0.3m, the minority component admixture increases precipitously, reaching 87% at 0.13m. Magnetization reversals produce the same effect. Consider the nonreciprocal Belinfante - Minkowski transverse spin -angular -momentum -density shif t, as a function of silicon slab thickness. Figure 2 plots the normalized shift in Eq. 16 pre -factor, 2 2 2 2 2 2 2 2 , 2 2 2 2 2 21 2c eff c eff c eff c eff c c c cgg BM c eff c eff c eff c eff c c c cgg g g g g g g g s g g g g g g g 22 gg (Eq. 24) We observe the same qualitative thickness dependence as in Fig. 1 , corresponding to the moderate, and relatively stable, admixture of minority circularly -polarized component above 0.3m thickness. 7 Fig. 1. Normalized nonreciprocal Belinfante -Minkowski transverse spin -angular -momentum - density shift per unit energy flux as a function of silicon slab thickness for 0.0086 g , corresponding to Ce 1Y2Fe5O12 garnet top cladding on SOI at 1.55 m wavelength . The inset shows the slab waveguide structure. M stands for the magnetization in the garnet. The magneto -optic gyrotropy of an iron garnet can be controlled through an app lied magnetic field. These ferri magnetic materials evince a hysteretic response, suc h as the one displayed in Fig. 3 (inset) for 532nm wavelength in a sputter -deposited film. The target composition is Bi1.5Y1.5Fe5.0O12. Shown here are actual experimental data extracted from Faraday rotation measurements. Below saturation, the magneto -optic response exhibits an effective gyrotropy value that can be tuned through th e applic ation of a magnetic field. These measurements correspond to a 0.5m-thick film on a (100) -oriented terbium gallium garnet (Tb GG) substrate. The optical beam is incident normal to the surface, and the hysteresis loop probes the degree of magnetization norma l to the surface as a function of applied magnetic field. These data show that the electromagnetic spin angular momentum can b e tuned below saturation and between opposite magnetization directions. Figure 3 also revea ls an interesting feature about the magneto -optic gyrotropy. The normalized nonreciprocal Belinfante -Minkowski transverse spin -angular -momentum -density shift per unit energy flux, ,,B M Ss , linearly tracks the gyrotropy, and is of the same order of magnitude as g, although thickness -dependent . Yet, as pointed out before, this thickness dependence reflects the admixture of the minor helicity component in the spin ellipticity. At 0.4m, for example, ,, 0.0072B M Ss when 0.0086 g . However, the major polarization helicity component contribution to ,,B M Ss is 84.4% at this thickness, translating into 0.00853 at 100%. At 0.25m, 8 ,, 0.00655B M Ss , and the major polarization helicity component contribution is 76.2%, translating into 0.0086 at 100%. We, thus, re-interpret the magneto -optic al gyrotropy as the normalized Belinfante -Minkowski spin -angular -momentum density shift per unit energy flux. Fig. 2 . Normalized nonreciprocal Belinfante -Minkowski transverse spin -angular -momentum - density pre-factor shift as a function of silicon slab thickness for 0.0086 g , corresponding to Ce1Y2Fe5O12 garnet top cladding on SOI at 1.55 m wavelength. 9 Fig. 3. Normalized nonreciprocal Belinfante -Minkowski transverse spin -angular -momentum - density shift per unit energy flux as a function of magneto -optical gyrotropy. Data correspond to 0.25m silicon -slab thickness with Ce1Y2Fe5O12 garnet top cladding , 1.55 m wavelength. The inset shows the gyrotropy versus magnetic field hysteresis loop of a magnetic garnet film at 532nm , sputter -deposited using a Bi1.5Y1.5Fe5.0O12 target . 2. Magnetization -Induced Electro magnetic Spin -Orbit Coupling Bliokh and co -authors have studied the electromagnetic spin -orbit coupling in non -paraxial optical vortex beams [7, 13]. They find that there is a spin dependent term in the orbital angular momentum expression that leads to spin -to-orbit angular momentum conversion. This phenomenon occurs under tight focusing or the scattering of light [ 7, 13]. Here we consider an alternativ e source of electromagnetic spin -orbit coupling, magnetization -induced coupling in evanescent waves. The time-averaged spin- and orbital -angular momenta co nservation laws put forth in [7] are * * * * 1Im ( ) Im22o t i j ij i j j i E E H E H E H E , and (Eq. 25) 22* * * 11Im ( ) Im ( )2 2 4o t j jkl l i k j i ijk oiE r E H r E H E H E (Eq. 26) Latin indices i, j,… take on values x, y, z and ijk is the Levi -Civita symbol. Summation over repeated indices is assumed. The interesting term in these equations, responsible for spin -orbit coupling, is *Im2o jiHE . Notice that it appears with opposite signs in the above equations , signaling a transfer of angular momentum from spin to orbital motion. As it stands, so far in our treatment , this term equals zero, since the spin points in the y -direction and the electric -field components of the TM wave point in the x -, and z -directions. A way to overcome this null coupling, and enable the angular momentum transfer, is to partially rotate the applied magnetic field about the x -axis away from the y -direction , as in Fig. 4 . This action induces a Faraday rotation about the z -axis, generating a spin -orbit coupling term in the angula r momentum conservation laws. A slight rotation or directional gradient in the magnetization M will induce electromagnetic spin-orbit interaction in the magneto -optic medium . Maxwell -Ampere’s law acquires off -diagonal components ig in the dielectric permittivity tensor upon rotation of the magnetic moment in the iron garnet film away from the y -axis, as shown in Eq. 27 [14]. Hence, non -zero electromagnetic field components yE and xH , and spin -orbit coupling, are induced in the propagating wave. The spatial, non -intrinsic, component, characteristic of orbital motion, emerges in the form of a z -dependence in the angular momentum, 10 embodied in the partial or total evanescence of the major circularly -polarized component as the wave propagates along th e guide. ˆ 0 0c o o c ci g ig EH i g i Etig (Eq. 27) In what sense is there an angular momentum transfer from spin to orbital, in this case? As the polarization rotates in the x -y plane due to the Faraday Effect, there will be a spatially -dependent reduction in the circulating electric field component of the electromagnetic wave alo ng the propagation -direction. This can be seen as a negative increase in circular polarization with z, i.e. , an orbital angular momentum in the opposite direction to the electromagn etic spin. Fig. 4. Rotated magnetization M generates TM to TE waveguide mode coupling and electromagnetic spin -orbit coupling. Finally, we derive an explicit expression for the spin -orbit coupling term. The relevant term appearing in the orbital angular momentum flux in the z -direction is * 0Im2z z yHE Eq. (28) We assume t hat Faraday rotation induces the ,yzEH terms via TM to transverse -electric (TE) mode conversion, where 0zyiHEx , Eq. (29) and ,0 sinTE TEx i z y y FE E e e z . Eq. (30) ,0yE is the electric field amplitude corresponding to full TM to TE conversion, F is the specific Faraday rotation angle, TE and TE are the cover -layer decay constant and the propagation constant for the TE mode, respectively. For simplicity, we assume no linear birefringence in the wavegu ide, so TE TM . Hence, the spin to orbital angular momentum coupling term is 22 * 0 ,0 2Im sin 22 2TEx TE z z y y F FH E E e z Eq. (31) 11 Acknowledgment The authors thank Ramy El -Ganainy for suggesting this problem and for useful discussion s. 12 References 1. F. J. Beli nfante, “On the current and the density of electric charge, the energy, the linear momentum and the angular momentum of arbitrary fields,” Physica 7, 449 -474 (1940). 2. K. Y. Bliokh, A. Y. Bekshaev, and F. Nori, “Extraordinary momentum and spin in evanescent waves,” Nature Communications 5, 4300 (2014). 3. M. Neuberger an co -workers, “Experimental demonstration of the geometric so in Hall effect of light in highly focused vector beams,” Conference on Lasers and Electro -Optics,” QW1E.A (OSA 2012). 4. P. Banzer a co -workers, “The photonic wheel demonstration of a state of light with purely transverse momentum,” J. Eur. Opt. Soc. Rap. Publ. 8, 13032 (2013). 5. K. Y. Bliokh and F. Nori, “Transverse spin of a surface polariton,” Phys. Rev. A 85, 061801 (2012) . 6. K. –Y. Kim, I. –M. Lee, J. Kim, J. Jung, and B. Lee, “Time reversal and the spin angular momentum of tra nsverse electric and transverse -magnetic surface modes,” Phys. Rev. A 86, 063805 (2012). 7. K. Y. Bliokh, J. Dressel and F. Nori, “Conservation of the s pin and orbital angular momenta in electromagnetism,” New Journal of Physics 16, 093037 (2014). 8. A. Aiello, P. Banzer, M. Neugebauer and G. Leuchs, “From transverse angular momentum to photonic wheels,” Nature Photonics 9, 789 -795 (2015). 9. M.E.J. Friese, J. Enger, H. Rubinsztein -Dunlop, and N. R. Heckenberg, “Optical angular - momentum transfer to trapped absorbing particles, Phys. Rev. A 54 1593 -1596 (1996). 10. J. Fujita, M. Levy, R. M. Osgood, Jr. , L. Wilkens, and H. Dötsch, “Waveguide Optical Isolator Based on Mach -Zehnder Interferometer,” Appl. Phys. Lett. 76, 2158 (2000). 11. B. J. Stadler and T. Mizumoto, “Integrated magneto -optical materials and isolators: a review,” IEEE Photonics Journal 6, 0600215 (2014). 12. S. M. Barnett, “ Resolution of the Ab raham -Minkowski Dilemma ,” Phys. Rev. Lett. 104, 070401 (2010). 13. Y. Bliokh, A. Y. Bekshaev, and F. Nori, “Dual electromagnetism: helicity, spin, momentum and angular momentum,” New Journal of Physics 15, 033026 (2013). 14. D. C. Hutchings, B. M. Holmes, C. Zhang, P. Dulal, A. D. Block, S. -Y. Sung, N. C. A. Seaton, B. J. H. Stadler, “Quasi -phase -matched Faraday rotation in semiconductor waveguides with a magneto -optic cledding for monolithically integrated optical isolators,” IEEE Photonics Journal 5, 6602512 (2013).
1201.4842v2.Strong_Enhancement_of_Rashba_spin_orbit_coupling_with_increasing_anisotropy_in_the_Fock_Darwin_states_of_a_quantum_dot.pdf
arXiv:1201.4842v2 [cond-mat.mes-hall] 24 Jan 2012Strong Enhancement of Rashba spin-orbit coupling with incr easing anisotropy in the Fock-Darwin states of a quantum dot Siranush Avetisyan,1Pekka Pietil¨ ainen,2and Tapash Chakraborty‡1 1Department of Physics and Astronomy, University of Manitob a, Winnipeg, Canada R3T 2N2 2Department of Physics/Theoretical Physics, University of Oulu, Oulu FIN-90014, Finland We have investigated the electronic properties of elliptic al quantum dots in a perpendicular ex- ternal magnetic field, and in the presence of the Rashba spin- orbit interaction. Our work indicates that the Fock-Darwin spectra display strong signature of Ra shba spin-orbit coupling even for a low magnetic field, as the anisotropy of the quantumdot is increa sed. An explanation of this pronounced effect with respect to the anisotropy is presented. The stron g spin-orbit coupling effect manifests itself prominently in the corresponding dipole-allowed op tical transitions, and hence is susceptible to direct experimental observation. In recent years our interest in understanding the unique effects of the spin-orbit interaction (SOI) in semi- conductor nanostructures [2] has peaked, largely due to the prospect of the possible realization of coherent spin manipulation in spintronic devices [3], where the SOI is destined to play a crucial role [4]. As the SOI couples the orbital motion of the charge carriers with their spin state, an all-electrical control of spin states in nanoscale semiconductor devices could thus be a reality. In this context the Rashba SOI [5] has received particular at- tention, largelybecauseinatwo-dimensionalelectrongas thestrengthoftheRashbaSOIhasalreadybeenshownto be tuned by the application of an electric field [6]. While the earlier studies were primarily in a two-dimensional electron gas, the attention has now been focused on the role of SOI in a single InAS quantum dot [7]. The quan- tum dot (QD) [8], a system of few electrons confined in the nanometer region has the main advantage that the shape and size of the confinement can be externally con- trolled, which provides an unique opportunity to study the atomic-likepropertiesofthesesystems[8,9]. SOcou- plinginquantumdotsgeneratesanisotropicspinsplitting [10] which provides important information about the SO coupling strength. Extensivetheoreticalstudiesofthe influenceofRashba SOI in circularly symmetric parabolic confinement have already been reported earlier [11], where the SO cou- pling wasfound to manifest itself mainly in multiple level crossings and level repulsions. They were attributed to an interplay between the Zeeman and the SOI present in the system Hamiltonian. Those effects, in particular, the level repulsions were however weak and as a result, wouldrequireextraordinaryefforts todetect the strength of SO coupling [12] in those systems. Here we show that, by introducing anisotropy in the QD, i.e., by breaking the circular symmetry of the dot, we can generate a ma- jor enhancement of the Rashba SO coupling effects in a quantum dot. As shown below, this can be observed di- rectly in the Fock-Darwin states of a QD, and therefore should be experimentally observable [8, 9]. We show be- low that the Rashba SO coupling effects are manifestlystrongin an elliptical QD [13], which should providea di- rect route to unambigiously determine (and control) the SO coupling strength. It has been proposed recently that the anisotropy of a quantum dot can also be tuned by an in-plane magnetic field [14]. The Fock-Darwin energy levels in elliptical QDs sub- jected to a magnetic field was first reported almost two decadesago[13], whereit wasfound that the majoreffect of anisotropy was to lift the degeneracies of the single- particle spectrum [15]. The starting point of our present study is the stationary Hamiltonian HS=1 2m∗/parenleftBig p−e cAS/parenrightBig2 +Vconf(x,y)+HSO+Hz =H0+HSO+Hz where the confinement potential is chosen to be of the form Vconf=1 2m∗/parenleftbig ω2 xx2+ω2 yy2/parenrightbig , HSO=α /planckover2pi1/bracketleftbig σ×/parenleftbig p−e cAS/parenrightbig/bracketrightbig zis the Rashba SOI, and Hz isthe Zeemancontribution. Here m∗is the effective mass of the electron, σare the Pauli matrices, and we choose the symmetric gauge vector potential AS=1 2(−y,x,0). As in Ref. [13], we introduce the rotated coordinates and momenta x=q1cosχ−χ2p2sinχ, y=q2cosχ−χ2p1sinχ, px=p1cosχ+χ1q2sinχ, py=p2cosχ+χ1q1sinχ, where χ1=−/bracketleftbig1 2/parenleftbig Ω2 1+Ω2 2/parenrightbig/bracketrightbig1 2, χ2=χ−1 1, tan2χ=m∗ωc/bracketleftbig 2/parenleftbig Ω2 1+Ω2 2/parenrightbig/bracketrightbig1 2//parenleftbig Ω2 1−Ω2 2/parenrightbig , Ω2 1,2=m∗2/parenleftbig ω2 x,y+1 4ω2 c/parenrightbig , ωc=eB/m∗c. In terms of the rotated operators introduced above, the Hamiltonian H0is diagonal [13] H0=1 2m∗/summationdisplay ν=1,2/bracketleftbig β2 νp2 ν+γ2 νq2 ν/bracketrightbig ,2 /s48/s52/s56/s49/s50/s49/s54 /s32 /s32/s32/s69 /s32/s40/s109 /s101/s86 /s41 /s40/s97/s41 /s52/s56/s49/s50/s49/s54/s50/s48 /s32/s32 /s32/s40/s99/s41 /s48/s52/s56/s49/s50/s49/s54 /s32/s32/s69 /s32/s40/s109 /s101/s86 /s41 /s40/s98/s41 /s48 /s49 /s50 /s51 /s52/s52/s56/s49/s50/s49/s54/s50/s48 /s66/s32/s40/s84/s41 /s32/s32 /s66/s32/s40/s84/s41/s49 /s50/s51 /s52 /s48/s40/s100/s41 FIG. 1: Magnetic field dependence of the low-lying Fock- Darwin energy levels of an elliptical dot without the Rashba SO interaction ( α= 0). The results are for (a) ωx= 4 meV andωy= 4.1 meV, (b) ωx= 4 meV and ωy= 6 meV, (c) ωx= 4 meV and ωy= 8 meV, and (d) ωx= 4 meV and ωy= 10 meV. where β2 1=Ω2 1+3Ω2 2+Ω2 3 2(Ω2 1+Ω2 2), γ2 1=1 4/parenleftbig 3Ω2 1+Ω2 2+Ω2 3/parenrightbig , β2 2=3Ω2 1+Ω2 2−Ω2 3 2(Ω2 1+Ω2 2), γ2 2=1 4/parenleftbig Ω2 1+3Ω2 2−Ω2 3/parenrightbig , Ω2 3=/bracketleftBig/parenleftbig Ω2 1−Ω2 2/parenrightbig2+2m∗2ω2 c/parenleftbig Ω2 1+Ω2 2/parenrightbig/bracketrightBig1 2. Since the operator H0is obviously equivalent to the Hamiltonianoftwoindependentharmonicoscillators,the states of the electron can be described by the state vec- tors|n1,n2;sz/angbracketright. Here the oscillator quantum numbers ni= 0,1,2,...correspond to the orbital motion and sz=±1 2to the spin orientation of the electron. The Rashba Hamiltonian, in terms of the rotated op- erators is now written as, /planckover2pi1 αHSO=σx(sinχχ1−cosχω0)q1 −σy(sinχχ1+cosχω0)q2 −σy(cosχ−sinχω0χ2)p1 +σx(cosχ+sinχω0χ2)p2, whereω0=eB/2c. The effect of the SO coupling is readily handled by resortingto the standard ladder oper- ator formalism of harmonic oscillators and by diagonal- izingHSOin the complete basis formed by the vectors |n1,n2;sz/angbracketright. The Fock-Darwin states in the absence of the Rashba SOI (α= 0) are shown in Fig. 1, for ωx= 4 meV and/s48/s52/s56/s49/s50/s49/s54/s32 /s69 /s32/s40/s109 /s101/s86 /s41 /s32 /s32/s32 /s40/s97/s41 /s52/s56/s49/s50/s49/s54/s50/s48 /s32/s32/s32 /s50/s40/s100/s41 /s48/s52/s56/s49/s50/s49/s54 /s32/s32/s69 /s32/s40/s109 /s101/s86 /s41 /s40/s98/s41/s52/s56/s49/s50/s49/s54/s50/s48 /s66/s32/s40/s84/s41 /s66/s32/s40/s84/s41 /s32/s32 /s40/s99/s41 /s48 /s49 /s50 /s51 /s52/s48 /s49 /s51 /s52 FIG. 2: Same as in Fig. 1, but for α= 20. ωy= 4.1,6,8,10 meV in (a)-(d) respectively. We have considered the parameters of an InAs QD [11] through- out, because in such a narrow-gap semiconductor sys- tem, the dominant source of the SO interaction is the structural inversion asymmetry [16], which leads to the RashbaSO interaction. As expected, breakingof circular symmetry in the dot results in lifting of degeneracies at B= 0, that is otherwise present in a circular dot [13, 15]. In Fig. 1 (a), the QD is very close to being circularly symmetric, and as a consequence, the splittings of the zero-field levels are vanishingly small. As the anisotropy of the QD is increased [(b) – (d)], splitting of the levels becomes more appreciable. As the SO term is linear in the position and momen- tum operators it is also linear in the raising and lowering ladder operators. It is also off-diagonal in the quantum number sz. As a consequence, the SOI can mix only states which differ in the spin orientation, and differ by 1 either in the quntum number n1or inn2but not in both. In the case of rotationally symmetric confinements these selection rules translate to the conservation of the total angular momentum j=m+szin the planar motion of the electron. At the field B= 0 the ground states |0,0;±1 2/angbracketrightare two- fold degenerate. Due to the selection rules, this degener- acy cannot be lifted either by the eccentricity of the dot or by the Rashba coupling. Many of the excited states, such as |n1,n2;±1 2/angbracketrightretain their degeneracy no matter how strong the SO coupling is or how eccentric the dot is, as we can see in the Figs. 1-3. At the same time, many other degeneracies are removed by squeezing or streching the dot. At non-zero magnetic fields some of the cross- ings of the energy spectra are turned to anti-crossings by the Rashba term in the Hamiltonian. For example, the3 /s48/s52/s56/s49/s50/s49/s54/s69 /s32/s40/s109 /s101/s86 /s41 /s32/s32 /s32/s69 /s32/s40/s109 /s101/s86 /s41 /s40/s97/s41 /s48 /s49 /s50 /s51 /s52/s48/s52/s56/s49/s50/s49/s54 /s32/s32 /s40/s98/s41 /s48 /s49 /s50 /s51 /s52/s52/s56/s49/s50/s49/s54/s50/s48 /s32/s32 /s40/s100/s41/s52/s56/s49/s50/s49/s54/s50/s48 /s66/s32/s40/s84/s41/s66/s32/s40/s84/s41 /s32/s32 /s40/s99/s41 FIG. 3: Same as in Fig. 1, but for α= 40. second and third excited states in Fig. 2 (a) – Fig. 2 (d) are composed mainly of the states |0,0;1 2/angbracketrightand|1,0;−1 2/angbracketright which are mixed by the HSOaroundB= 3T causing a level repulsion. We can also see that the squeezing of the dot enhances the SO coupling. This can be thought of as a consequence of pushing some states out of the way, just as in our example of the state |1,1;1 2/angbracketright. SOI mixes it with the state |1,0;−1 2/angbracketrightcausing the latter state to shift down- ward in energy thereby reducing the anti-crossing gap. Squeezing the dot, however moves the state energetically farther away from |1,0;−1 2/angbracketrightand so weakens this gap re- duction effect. It is abundantly clear from the features revealed in the energy spectra that for a combination of strong anisotropy of the dot and higher values of the SO coupling strength, large anti-crossing gaps would appear even for relatively low magnetic fields. The effects of anisotropy and spin-orbit interaction on the energy spectra above are also reflected in the optical absorption spectra. Let us turn our attention on the absorption spectra for transitions from the ground state to the excited states. For that purpose we subject the dot to the radiation field AR=A0ˆǫ/parenleftBig ei(ω/c)ˆn·r−iωt+e−i(ω/c)ˆn·r+iωt/parenrightBig , whereˆǫ,ωandˆnare the polarization, frequency and the direction of propagation of the incident light, respec- tively. We let the radiation enter the dot along the direc- tion perpendicular to the motion of the electron, that is parallel to the z-axis. Due to the transversalitycondition the polarization vector will then lie in the xy-plane. As usual, we shall make two approximations. First we assume the intensity of the field be so weak that only the terms linear in ARhas to be taken into account. Then the effect of the radiative magnetic field on the spin can/s48/s51/s54/s57/s69/s32/s40/s109/s101/s86/s41 /s69/s32/s40/s109/s101/s86/s41 /s32/s32/s69/s32/s40/s109/s101/s86/s41 /s40/s97/s41 /s48/s51/s54/s57/s49/s50 /s32/s32/s32 /s40/s98/s41 /s48 /s49 /s50/s48/s51/s54/s57/s49/s50 /s40/s99/s41/s48/s51/s54/s57 /s32/s32 /s40/s100/s41 /s48/s51/s54/s57/s49/s50 /s32/s32 /s40/s101/s41 /s48 /s49 /s50/s48/s51/s54/s57/s49/s50 /s66/s40/s84/s41 /s32/s32 /s66/s40/s84/s41/s40/s102/s41 FIG.4: Opticalabsorption(dipoleallowed) spectaofellip tical QDs for various choice of parameters: (a) i α= 0,ωx= 4 meV,ωy= 6, (b) α= 20,ωx= 4 meV, ωy= 8 meV, and (c)α= 40,ωx= 4,ωy= 6. The polarization of the incident radiation is along the x-axis. The parameters for (d)-(f) are the same, except that the incident radiation is polarized al ong they-axis. The areas of the filled circles are proportional to the calculated absorption cross-section. be neglected as well. So we can simply replace in the stationary Hamiltonian HSthe vector potential ASwith the field A=AS+AR. Discarding terms higher than linear order in ARleads to the total Hamiltonian H=HS+HR, where the radiative part HRis given by HR=−e mecAR·/parenleftBig p−e cAS/parenrightBig −αe /planckover2pi1c[σ×AR]z. The radiative Hamiltonian, even in the presence of the Rashba SO coupling can be expressed in the well-known form HR= ie c/planckover2pi1AR·[x,HS], xbeing the position operator in the xy-plane. Our second approximation is the familiar dipole ap- proximation. We assume that the amplitude of radiation can be taken as constantwithin the quantum dot, so that we are allowed to write the field as AR≈A0ˆǫ/parenleftbig e−iωt+eiωt/parenrightbig .4 Since the transition energies expressed in terms of radia- tion frequences are of the order of THz, the correspond- ing wavelengths are much larger than the typical size of a dot, thus justifying our approximation. Applying now the Fermi Golden Rule leads to the dipole approximation form σabs(ω) = 4π2αfωni|/angbracketleftn|ˆǫ·x|i/angbracketright|2δ(ωni−ω) of the absorption cross section for transitions from the inital state |i/angbracketrightto the final state |n/angbracketright. Hereαfis the fine structure constant and ωniis the frequency correspond- ing to the transition energy /planckover2pi1ω. The familiar dipole selection rules for oscillator states dictate largely the features seen in Fig. 4. In the absence of the SOI, these rules – the spin state is preserved and eithern1orn2is changed by unity – completely deter- mine the allowed two transitions/vextendsingle/vextendsingle0,0;−1 2/angbracketrightbig →/vextendsingle/vextendsingle1,0;−1 2/angbracketrightbig and/vextendsingle/vextendsingle0,0;−1 2/angbracketrightbig →/vextendsingle/vextendsingle0,1;−1 2/angbracketrightbig . In contrast to the case of circular dots the absorption in the elliptical dot depends strongly on the polarization. This is explained by noting that the oscillator strengths fni=2m∗ωni /planckover2pi1|/angbracketleftn|ˆǫ·x|i/angbracketright|2. actually probe the occupations of quantum states related to oscillations in the direction of the polarization ˆǫ. In a circular dot all oscillation directions are equally probable at all energies implying that the oscillator strengths are independent of the polarization and depend only slightly on the transitionenergyvia ωni, and the final state quan- tum numbers n1,2. When the dot is squeezed in the y- direction,say,theoscillatorstatesrelatedtothe y-motion are pushed up in energy. This means that the polariza- tion being along x-axis most of the oscillator strength comes from transitions to allowed states with lowest en- ergies. Similarly, when the incident radiation is polarized along the y-axis most of the contribution is due to the transitions to the oscillator states pushed up in the en- ergy. Inellipticaldotstheoscillatorstatesarenotpure x- andy-oscillators but their superpositions. Therefore in addition to the main absorption lines, other allowed final states have also non-vanishing oscillator strength. Fur- thermore, as one can see by looking at the phase space rotation formulas the external magnetic field tends to ro- tate directions of the oscillator motion causing a shift of the oscillator strength from an allowed transition to an- other. This is exactly what we see in Fig. 4 (a) and Fig. 4 (d). Even in the presence of the SOI the two allowed final oscillator states provide major contributions to the cor- responding corrected states. Hence we still see two dom- inant absorption lines. However, now many forbidden transitions have become allowed. The lowest absorption linecorrespondingtothetransitionbetweenZeemansplit states with main components/vextendsingle/vextendsingle0,0;−1 2/angbracketrightbig and/vextendsingle/vextendsingle0,0;1 2/angbracketrightbig pro- vides a typical example. The transition involves a spinflip and is therefore strongly forbidden without the SOI. Because the SOI mixes the state/vextendsingle/vextendsingle1,0;1 2/angbracketrightbig into the for- mer one and the/vextendsingle/vextendsingle0,1;−1 2/angbracketrightbig into the latter one, the tran- sition becomes possible. The appearance of other new lines can be explained by analogous arguments. There are also additional features involving discontinuities and anti-crossings in Fig. 4. A comparision with the energy spectra indicates that these are the consequences of the anti-crossings present in the energy spectra. It is also readily verified that the oscillator strengths satisfy the Thomas-Reiche-Kuhn sum rule [17] /summationdisplay nfni= 1. In terms of the cross section this translates to the condi- tion /integraldisplay∞ −∞σabs(ω)dω=2π2/planckover2pi1αf m∗. The absorptions visible in Fig. 4 practically saturate the sum rule, the saturation being, of course complete in the absence of the SOI in panels (a) and (d). The largest fraction (of the order of 1/10) of the cross section either falling outside of the displayed energy scale or having too low intensity to be discernible in our pictures is found at the strongest Rashba coupling in the panels (c) and (f) for large magnetic fields, as expected. The results presented here clearly indicate that, the anisotropyofaQDalonecausesliftingofthedegeneracies oftheFock-DarwinlevelsatB=0,asreportedearlier[13]. However, for large SO coupling strengths α, the effects of the Rashba SOI, mainly the level repulsions at finite magnetic fields, are maginified rather significantly as one introduces anisotropy in the QD. This is reflected also in the corresponding dipole-allowed optical transitions where the distinct anti-crossingbehavioris observedthat is a direct manifestation of the anti-crossings in the en- ergy spectra. This prominent effect of the Rashba SOI predicted here could be confirmed experimentally in op- tical spectroscopy and the Fock-darwin spectra of few- electron QDs [9, 18, 19]. It would also provide a very useful step to control the SO coupling in nanostructures, en route to semiconductor spintronics [3]. The work was supported by the Canada Research Chairs Program of the Government of Canada. [‡] Electronic address: tapash@physics.umanitoba.ca [2] Y. Oreg, P.W. Brouwer, X.Waintal, andB.I. Halperin, in Nano-Physics & Bio-Electronics: A New Odyssey , edited by, T. Chakraborty, F. Peeters, and U. Sivan (Elsevier, Amsterdam, 2002). [3] For recent comprehensive reviews, see, T. Dietl, D.D. Awschalom, M. Kaminska, and H. Ono, (Eds.) Spintron- ics(Elsevier, Amsterdam, 2008); I. Zutic, J. Fabian, and5 S.DasSarma, Rev.Mod.Phys. 76, 323(2004); J. Fabian, A. Matos-Abiague, C. Ertler, P. Stano, and I. Zutic, Acta Physica Slovaca 57, 565 (2007); M.W. Wu, J.H. Jiang, and M.Q. Weng, Phys. Rep. 493, 61 (2010). [4] H.-A. Engel, B.I. Halperin, and E.I. Rashba, Phys. Rev. Lett.95, 166605 (2005). [5] Y.A. Bychkov and E.I. Rashba, J. Phys. C 17, 6039 (1984). [6] J. Nitta, T. Akazaki, H. Takayanagi, and T. Enoki, Phys. Rev. Lett. 78, 1335 (1997); M. Studer, G. Salis, K. En- sslin, D.C. Driscoll, and A.C. Gossard, ibid.103, 027201 (2009); D. Grundler, ibid.84, 6074 (2000). [7] H. Sanada, T. Sogawa, H. Gotoh, K. Onomitshu, M. Ko- hda, J. Nitta, and P.V. Santos, Phys. Rev. Lett. 106, 216602 (2011); S. Takahashi, R.S. Deacon, K. Yoshida, A. Oiwa, K. Shibata, K. Hirakawa, Y. Tokura, and S. Tarucha, ibid.104, 246801 (2010). [8] T. Chakraborty, Quantum Dots (North-Holland, Am- sterdam, 1999); T. Chakraborty, Comments Condens. Matter Phys. 16, 35 (1992); P.A. Maksym and T. Chakraborty, Phys. Rev. Lett. 65, 108 (1990). [9] D. Heitmann (Ed.), Quantum Materials (Springer, Hei- delberg, 2010). [10] J. K¨ onemann, R.J. Haug, D.K. Maude, V. Fal’ko, and B.L. Altshuler, Phys. Rev. Lett. 94, 226404 (2005). [11] T. Chakraborty and P. Pietil¨ ainen, Phys. Rev. Lett. 95, 136603 (2005); P. Pietil¨ ainen and T. Chakraborty, Phys. Rev. B 73, 155315 (2006); T. Chakraborty and P. Pietil¨ ainen, ibid.71, 113305 (2005); A. Manaselyan andT. Chakraborty, Europhys. Lett. 88, 17003 (2009); and the references therein. [12] H.-Y. Chen, V. Apalkov, andT. Chakraborty, Phys. Rev. B75, 193303 (2007). [13] A.V.MadhavandT. Chakraborty, Phys.Rev.B 49, 8163 (1994); See also, P.A. Maksym, Physica B 249-251 , 233 (1998). [14] M.P. Nowak, B. Szafran, F.M. Peeters, B. Partoens, and W.J. Pasek, Phys. Rev. B 83, 245324 (2011). [15] A. Singha, V. Pellegrini, S. Kalliakos, B. Karmakar, A. Pinczuk, L.N. Pfeiffer, and K.W. West, Appl. Phys. Lett. 94, 073114 (2009); D.G. Austing, S. Sasaki, S. Tarucha, S.M. Reimann, M. Koskinen, M. Manninen, Phys. Rev. B60, 11514 (1999). [16] W. Zawadzki and P. Pfeffer, Semicond. Sci. Technol. 19, R1 (2004). [17] J.J. Sakurai and J. Napolitano, Modern Quantum Me- chanics, second edition, (Addison-Wesley, New York, 1994), p. 368; W. Thomas, Naturwissenschaften 13, 627 (1925): W. Kuhn, Z. Phys. 33, 408 (1925); F. Reiche and W. Thomas, Z. Phys. 34, 510 (1925). [18] L.P. Kouwenhoven, D.G Austing, and S. Tarucha, Rep. Prog. Phys. 64, 701 (2001); A. Babinski, M. Potemski, S. Raymond, J. Lapointe, and Z.R. Wasilewski, phys. stat. sol. (c)3, 3748 (2006). [19] V. Pellegrini, and A. Pinczuk, phys. stat. sol. (b) 243, 3617 (2006).
1505.04301v1.Dynamics_of_a_macroscopic_spin_qubit_in_spin_orbit_coupled_Bose_Einstein_condensates.pdf
arXiv:1505.04301v1 [quant-ph] 16 May 2015Dynamics of a macroscopic spin qubit in spin-orbit coupled Bose-Einstein condensates Sh Mardonov1,2,3, M Modugno4,5and E Ya Sherman1,4 1Department of Physical Chemistry, The University of the Basque C ountry, 48080 Bilbao, Spain 2The Samarkand Agriculture Institute, 140103 Samarkand, Uzbek istan 3The Samarkand State University, 140104 Samarkand, Uzbekistan 4IKERBASQUE Basque Foundation for Science, Bilbao, Spain 5Department of Theoretical Physics and History of Science, Univer sity of the Basque Country UPV/EHU, 48080 Bilbao, Spain E-mail:evgeny.sherman@ehu.eus Abstract. We consider a macroscopic spin qubit based on spin-orbit coupled Bos e- Einstein condensates, where, in addition to the spin-orbit coupling, spin dynamics strongly depends on the interaction between particles. The evolut ion of the spin for freely expanding, trapped, and externally driven condensate s is investigated. For condensates oscillating at the frequency corresponding to the Ze eman splitting in the synthetic magnetic field, the spin Rabi frequency does not depend on the interaction between the atoms since it produces only internal forces and does not change the total momentum. However, interactions and spin-orbit coupling br ing the system into a mixed spin state, where the total spin is inside rather than on t he Bloch sphere. This greatly extends the available spin space making it three -dimensional, but imposes limitations on the reliable spin manipulation ofsuch a macroscop icqubit. The spin dynamics can be modified by introducing suitable spin-dependent initial phases, determined by the spin-orbit coupling, in the spinor wave function. PACS numbers: 03.75.Mn, 67.85.-d Keywords : Two-component Bose-Einstein condensate, spin-orbit coupling, spin dynamics Submitted to: J. Phys. B: At. Mol. Opt. Phys.Dynamics of a macroscopic spin qubit in spin-orbit coupled B EC 2 1. Introduction The experimental realization of synthetic magnetic fields and spin-o rbit coupling (SOC) [1, 2] in Bose-Einstein condensates (BECs) of pseudospin-1/2 par ticles has provided novel opportunities for visualizing unconventional phenomena in qu antum condensed matter [3, 4]. More recently, also ultracold Fermi gases with synthe tic SOC have been produced and studied [5, 6]. These achievements have motivated an d intense activity, and a rich variety of new phases and phenomena induced by the SOC h as been discussed both theoretically and experimentally [7, 8, 9, 10, 11, 12, 13, 14, 15 , 16, 17, 18, 19, 20]. Recently, it has also been experimentally demonstrated [3, 21] the a bility of a reliable measurement of coupled spin-coordinate motion. One of the prospective applications of spin-orbit coupled Bose-Eins tein condensates consists in the realization of macroscopic spin qubits [8]. A more detaile d analysis of quantum computation based on a two-component BEC was propose d in [22]. The gates for performing these operations can be produced by means of the SOC and of an external synthetic magnetic field. Due to the SOC, a periodic mecha nical motion of the condensate drives the spin dynamics and can cause spin-flip tra nsitions at the Rabi frequency depending on the SOC strength. This technique, known in semiconductor physics as the electric dipole spin resonance, is well suitable for the m anipulation of qubits based on the spin of a single electron [23, 24, 25]. For the macr oscopic spin qubits based onBose-Einstein condensate, the physics is different in at lea st two aspects. First, a relative effect of the SOC compared to the kinetic energy can be mu ch stronger here than in semiconductors. Second, the interaction between the bos ons can have a strong effect on the entire spin dynamics. Here we study how the spin evolution of a quasi one-dimensional Bos e-Einstein condensate depends on the repulsive interaction between the par ticles and on the SOC strength. The paper is organized as follows. In Section 2 we remind t he reader the ground state properties of a quasi-one dimensional condensate a nd consider simple spin- dipole oscillations. In Section 3 we analyze, by means of the Gross-Pit aevskii approach, thedynamics of free, harmonically trapped, andmechanically driven macroscopic qubits based on such a condensate. We assume that the periodic mechanic al driving resonates with the Zeeman transition in the synthetic magnetic field and find diffe rent regimes of the spin qubit operation in terms of the interaction between the ato ms, the driving frequency and amplitude. We show that some control of the spin qu bit state can be achieved by introducing phase factors, dependent on the SOC, in the spinor wave function. Conclusions will be given in Section 4. 2. Ground state and spin-dipole oscillations 2.1. Ground state energy and wave function Before analyzing the spin qubit dynamics, we remind the reader how t o obtain the ground state of an interacting BEC. In particular, we consider a ha rmonically trappedDynamics of a macroscopic spin qubit in spin-orbit coupled B EC 3 quasi one-dimensional condensate, tightly bounded in the transv erse directions. The system can be described by the following effective Hamiltonian, where the interactions between the atoms are taken into account in the Gross-Pitaevskii form: /hatwideH0=/hatwidep2 2M+Mω2 0 2x2+g1|ψ(x)|2. (1) Hereψ(x) is the condensate wave function, Mis the particle mass, ω0is the frequency of the trap (with the corresponding oscillator length aho=/radicalbig /planckover2pi1/Mω0), andg1= 2as/planckover2pi1ω⊥ is the effective one-dimensional interaction constant, with asbeing the scattering length of interacting particles, and ω⊥≫ω0being the transverse confinement frequency. For further calculations we put /planckover2pi1≡M≡1, and measure energy in units of ω0and length in units of aho, respectively. All the effects of the interaction are determined by the dimensionless parameter /tildewideg1N, where/tildewideg1= 2/tildewideas/tildewideω⊥, where/tildewideasis the scattering length in the units of aho,/tildewideω⊥is the transverse confinement frequency in the units of ω0, andN is the number of particles. In physical units, for a condensate of87Rb and an axial trapping frequency ω0= 2π×10 Hz, the unit of time corresponds to 0 .016 s, the unit of lengthahocorresponds to 3 .4µm, and the unit of speed ahoω0becomes 0.021 cm/s, respectively. In addition, considering that as= 100aB,aBbeing the Bohr radius, in the presence of a transverse confinement with frequency ω⊥= 2π×100 Hz the dimensionless coupling constant /tildewideg1turns out to be of the order of 10−3. In order to find the BEC ground state we minimize the total energy in a properly truncated harmonic oscillator basis. To design the wave function we take the real sum of even-order eigenfunctions ψ0(x) =N1/2nmax/summationdisplay n=0C2nϕ2n(x). (2) Here ϕ2n(x) =1/radicalbig π1/2(2n)!22nH2n(x)exp/bracketleftbigg −x2 2/bracketrightbigg , (3) whereH2n(x) are the Hermite polynomials, and the normalization is fixed by requirin g that nmax/summationdisplay n=0C2 2n= 1. (4) The coefficients C2nare determined by minimizing the total energy Etot, such that Emin= min C2n{Etot}, (5) where Etot=1 2/integraldisplay/bracketleftBig (ψ′(x))2+x2ψ2(x)+/tildewideg1ψ4(x)/bracketrightBig dx, (6) and|C2nmax| ≪1. Formulas (5) and (6) yield the ground state energy, while the width o f the condensate is defined as: wgs=/bracketleftbigg2 N/integraldisplay x2|ψ0(x)|2dx/bracketrightbigg1/2 . (7)Dynamics of a macroscopic spin qubit in spin-orbit coupled B EC 4 Figure 1. (Color online) Ground-state probability density of the condensate obtained from (2)-(6) (blue solid line), compared with the Thomas-Fermi app roximation in (8) (red dashed line) for /tildewideg1N= 40. In the non interacting limit, /tildewideg1= 0,ψ0(x) is the ground state of the harmonic oscillator (nmax= 0), that is a Gaussian function with wgs= 1. In the opposite, strong coupling limit /tildewideg1N≫1, the exact wave function (2) is well reproduced (see Figure 1) by the Thomas-Fermi expression ψTF(x) =√ 3 2√ N w3/2 TF/parenleftbig w2 TF−x2/parenrightbig1/2;|x| ≤wTF, (8) wherewTF= (3/tildewideg1N/2)1/3. In general, for a qualitative description of the ground state one ca n use instead of the exact wave function (2), the Gaussian ansatz ψG(x) =/parenleftbiggN π1/2w/parenrightbigg1/2 exp/bracketleftbigg −x2 2w2/bracketrightbigg , (9) where the width wis single variational parameter for the energy minimization. Then the total energy (6) becomes: Etot=N/bracketleftbigg1 4/parenleftbigg w2+1 w2/parenrightbigg +/tildewideg1N 2(2π)1/2w/bracketrightbigg . (10) The latter is minimized with respect to wby solving the equation dEtot dw=N/bracketleftbigg1 2/parenleftbigg w−1 w3/parenrightbigg −/tildewideg1N 2(2π)1/2w2/bracketrightbigg = 0. (11) Inthestrongcouplingregime, /tildewideg1N≫1, wehavew≫1sothat-toafirstapproximation - the kinetic term ∝1/w3in (11) can be neglected, yielding the following value for the width of the ground state: /tildewidewG=/parenleftbigg/tildewideg1N√ 2π/parenrightbigg1/3 . (12) The first order correction can be obtained by writing w=/tildewidew+ǫ(ǫ≪1), so that from (11) it follows: w=/parenleftbigg/tildewideg1N√ 2π/parenrightbigg1/3 +√ 2π 3/tildewideg1N. (13)Dynamics of a macroscopic spin qubit in spin-orbit coupled B EC 5 0 10 20 30 40012345 g/OverTilde l1NEmin,wgs Figure 2. (Color online) Ground state energy (black solid line) and condensate width (red dashed line) vs. the interaction parameter /tildewideg1N. By substituting (13) in (10) we obtain that the leading term in the gro und state energy for/tildewideg1N≫1 is: E[G] min=3 4N/parenleftbigg/tildewideg1N√ 2π/parenrightbigg2/3 . (14) In figure 2 we plot the ground state energy and the condensate wid th as a function of the interaction, as obtained numerically from (5) and (7), respe ctively. As expected, in the strong coupling regime /tildewideg1N≫1 both quantities nicely follow the behavior (not shown in the Figure) predicted both by the Gaussian approximation a nd by the exact solution, namely Emin∝(/tildewideg1N)2/3andwgs∝(/tildewideg1N)1/3. 2.2. Simple spin-dipole oscillations Let us now turn to the case of a condensate of pseudospin 1/2 ato ms. Here the system is described by a two-component spinor wave function Ψ = [ ψ↑(x,t),ψ↓(x,t)]T, still normalized to the total number of particles N.The interaction energy (third term in the functional (6)) now acquires the form (see, e.g. [9]) Eint=1 2/tildewideg1/integraldisplay/bracketleftbig |ψ↑(x,t)|2+|ψ↓(x,t)|2/bracketrightbig2dx, (15) where, for simplicity and qualitative analysis, we neglect the depende nce of interatomic interaction on the spin component ↑or↓and characterize all interactions by a single constant/tildewideg1. Here we consider spin dipole oscillations, induced by a given small initial s ymmetric displacement of the two spin components ±ξ. For a qualitative understanding, we assume a negligible spin-orbit coupling and a Gaussian form of the wave function presented as ΨG(x) =1√ 2/bracketleftBigg ψG(x−ξ) ψG(x+ξ)/bracketrightBigg , (16) whereψGis given by (9), and ξ≪w. The corresponding energy is given by: E=E[G] min+Esh, (17)Dynamics of a macroscopic spin qubit in spin-orbit coupled B EC 6 whereE[G] minis defined by (14) and Eshis the shift-dependent contribution: Esh=N 2ξ2/parenleftbigg 1−/tildewideg1N√ 2πw3/parenrightbigg . (18) Then, it follows that the corresponding oscillation frequency is ωsh=/radicalBigg 1−/tildewideg1N√ 2πw3. (19) For strong interaction ( /tildewideg1N≫1) by substituting (13) in (19) we obtain: ωsh≈/parenleftBigg√ 2π /tildewideg1N/parenrightBigg2/3 . (20) Therefore, for strong interaction the frequency of the spin dipo le oscillations falls as (/tildewideg1N)−2/3, and this result is common for the Gaussian ansatz and for the exac t solution; it will be useful in the following section. 3. Spin evolution and particles interaction 3.1. Hamiltonian, spin density matrix, and purity To consider the evolution of the driven quasi one dimensional pseud ospin-1/2 SOC condensate we begin with the effective Hamiltonian /hatwideH=α/hatwideσz/hatwidep+/hatwidep2 2+∆ 2/hatwideσx+1 2(x−d(t))2+/tildewideg1|Ψ|2. (21) Hereαis the SOC constant (see [11] and [12] for comprehensive review on t he SOC in coldatomicgases), /hatwideσxand/hatwideσzarethePauli matrices, ∆isthesynthetic Zeemansplitting, andd(t) is the driven displacement of the harmonic trap center as can be ob tained by a slow motion of the intersection region of laser beams trapping the c ondensate. Thetwo-componentspinorwavefunctionΨisobtainedasasolutiono fthenonlinear Schr¨ odinger equation i∂Ψ ∂t=/hatwideHΨ. (22) To describe spin evolution we introduce the reduced density matrix ρ(t)≡ |Ψ∝an}bracketri}ht∝an}bracketle{tΨ|=/bracketleftBigg ρ11(t)ρ12(t) ρ21(t)ρ22(t)/bracketrightBigg , (23) where we trace out the x−dependence by calculating integrals ρ11(t) =/integraldisplay |ψ↑(x,t)|2dx, ρ22(t) =/integraldisplay |ψ↓(x,t)|2dx, ρ12(t) =/integraldisplay ψ∗ ↑(x,t)ψ↓(x,t)dx, ρ21(t) =ρ∗ 12(t), (24) and, as a result, tr(ρ(t))≡ρ11(t)+ρ22(t) =N. (25)Dynamics of a macroscopic spin qubit in spin-orbit coupled B EC 7 Figure 3. (Color online) (a) Separation of a freely expanding condensate in tw o spin- up and spin-down components with opposite anomalous velocities. (b ) Oscillation of the spin-up and spin-down components in the harmonic trap. Although the |Ψ∝an}bracketri}htstate is pure, integration in (24) produces ρ(t) formally describing a mixed state in the spin subspace. One can characterize the resultin g spin state purity by a parameter Pdefined as P=2 N2/parenleftbigg tr/parenleftbig ρ2/parenrightbig −N2 2/parenrightbigg , (26) where 0≤P≤1, tr/parenleftbig ρ2/parenrightbig =N2+2(|ρ12|2−ρ11ρ22), (27) and we omitted the explicit t−dependence for brevity. The system is in the pure state whenP= 1,that is tr(ρ2) =N2withρ11ρ22=|ρ12|2. In the fully mixed state, where tr(ρ2) =N2/2, we have P= 0 with ρ11=ρ22=N 2, ρ 12= 0. (28) The spin components defined by ∝an}bracketle{t/hatwideσi∝an}bracketri}ht ≡tr(/hatwideσiρ)/Nbecome ∝an}bracketle{t/hatwideσx∝an}bracketri}ht=2 NRe(ρ12),∝an}bracketle{t/hatwideσy∝an}bracketri}ht=−2 NIm(ρ12), ∝an}bracketle{t/hatwideσz∝an}bracketri}ht=2 Nρ11−1, (29) and the purity P=/summationtext3 i=1∝an}bracketle{t/hatwideσi∝an}bracketri}ht2, which allows one to match the value of Pand the length of the spin vector inside the Bloch sphere. For a pure state/summationtext3 i=1∝an}bracketle{t/hatwideσi∝an}bracketri}ht2= 1, and the total spin is on the Bloch sphere. Instead, for a fully mixed state/summationtext3 i=1∝an}bracketle{t/hatwideσi∝an}bracketri}ht2= 0, and the spin null. 3.2. A simple condensate motion Let us suppose that a condensate of interacting spin-orbit couple d particles is located in a harmonic trap and characterized by an initial wave function Ψ0(x,0) =1√ 2ψin(x)/bracketleftBigg 1 1/bracketrightBigg , (30)Dynamics of a macroscopic spin qubit in spin-orbit coupled B EC 8 (a) 024680.00.20.40.60.81.0 tP/LParen1t/RParen1(b) 0246810120.00.20.40.60.81.0 t/LAngleBracket1Σ/Hat x/LParen1t/RParen1/RAngleBracket1 Figure 4. (Color online) ( a) Purity and ( b) spin component as a function of time for a condensate released from the trap, for α= 0.2. The lines correspond to /tildewideg1N= 0 (black solid line; for the purity cf. (32)), /tildewideg1N= 10 (red dashed line), and /tildewideg1N= 20 (blue dot-dashed line). with the spin parallel to the x−axis. The spin-orbit coupling modifies the commutator corresponding to t he velocity operator by introducing the spin-dependent contribution as: /hatwidev≡i/bracketleftbigg/hatwidep2 2+α/hatwideσz/hatwidep,/hatwidex/bracketrightbigg =/hatwidep+α/hatwideσz. (31) The effect of the spin-dependent anomalous velocity term on the co ndensate motion was clearly observed experimentally in [3] as the spin-induced dipole os cillations and in [21] as the Zitterbewegung . Since the initial spin in (30) is parallel to the x-axis, the expectation value of the velocity vanishes, ∝an}bracketle{t/hatwidev∝an}bracketri}ht= 0. Free and oscillating motion of the BEC is shown in figure 3(a) and figure 3(b), respectively. When one switches off the trap, the condensate is se t free, and the two spin components start to move apart and the condensate splits up , see figure 3(a). Each spin-projected component broadens due to the Heisenberg momentum-coordinate uncertainty and interaction. The former effect is characterized b y a rate proportional to 1/wgs.At large/tildewideg1N,the width wgs∼(/tildewideg1N)1/3,and, as a result, the quantum mechanical broadening rate decreases as ( /tildewideg1N)−1/3.At the same time, the repulsion between the spin-polarized components accelerates the peak sep aration [26] and leads to the asymptotic separation velocity ∼(/tildewideg1N)1/2. This acceleration by repulsion leads to opposite time-dependent phase factors in ψ↑(x,t) andψ↓(x,t) in (24) and, therefore, results in decreasing in |ρ12(t)|and in the purity. Thus, with the increase in the interaction, the purity and the x−spin component asymptotically tend to zero faster, as demonstrated in figure 4. For a noninteracting condensate with the initial Gaussian wave function ψin∼exp(−x2/2w2) the purity can be written analytically as P0(t) = exp/bracketleftBigg −2/parenleftbiggαt w/parenrightbigg2/bracketrightBigg . (32) In the presence of the trap (figure 3(b)), the anomalous velocity in (31) causes spin components (spin-dipole) oscillations with a characteristic frequen cy of the oder of ωshDynamics of a macroscopic spin qubit in spin-orbit coupled B EC 9 (a) 0102030405060700.00.20.40.60.81.0 tP/LParen1t/RParen1 (b) 0102030405060700.00.20.40.60.81.0 t/LAngleBracket1Σ/Hat x/LParen1t/RParen1/RAngleBracket1 (c) 010203040506070/MinuΣ3/MinuΣ2/MinuΣ10123 t/LAngleBracket1xΣ/Hat z/RAngleBracket1 Figure 5. (Color online) (a) Purity, (b) spin component, and (c) spin dipole mom ent as a function of time for the system in the harmonic trap with α= 0.2,∆ = 0, d0= 0. The different lines correspond to /tildewideg1N= 0 (black solid line), /tildewideg1N= 10 (red dashed line),/tildewideg1N= 20 (blue dot-dashed line), and /tildewideg1N= 60 (green dotted line). in (20). With the increase in the interatomic interaction, the freque ncyωshdecreases and, therefore, the amplitude of the oscillations arising due to the a nomalous velocity (∼α/ωsh) increases. As a result, the acceleration and separation of the sp in-projected components increase, the off-diagonal components of the densit y matrix in (24) became smaller, and the minimum in P(t) rapidly decreases to P(t)≪1 as shown by the exact numerical results presented in figures 5(a) and (b) [27]. In figure 5 (c) we show the corresponding evolution of spin density dipole moment ∝an}bracketle{tx/hatwideσz∝an}bracketri}ht=1 N/integraldisplay Ψ†x/hatwideσzΨdx. (33) Here the oscillation frequency is a factor of two larger than that of the spin density oscillation. 3.3. Spin-qubit dynamics and the Rabi frequency To manipulate the macroscopic spin qubit, the center of the trap is d riven harmonically at the frequency corresponding to the Zeeman splitting ∆ as d(t) =d0sin(t∆), (34) whered0is an arbitrary amplitude and the corresponding spin rotation Rabi f requency ΩRis defined as αd0∆.At ∆≪1,as will be considered here, for a noninteractingDynamics of a macroscopic spin qubit in spin-orbit coupled B EC 10 (a) 0 50 100 150 2000.00.20.40.60.81.0 tP/LParen1t/RParen1(b) 050100 150 200/MinuΣ1.0/MinuΣ0.50.00.51.0 t/LAngleBracket1Σ/Hat x/LParen1t/RParen1/RAngleBracket1 Figure 6. (Color online) ( a) Purity and ( b) spin component as a function of time for a driven condensate with α= 0.1,∆ = 0.1, d0= 2. The lines correspond to /tildewideg1N= 0 (black solid line), /tildewideg1N= 10 (red dashed line), and /tildewideg1N= 20 (blue dot-dashed line). condensate and a very weak spin-orbit coupling, the spin componen t∝an}bracketle{t/hatwideσx(t)∝an}bracketri}htis expected to oscillate approximately as ∝an}bracketle{t/hatwideσx(t)∝an}bracketri}ht= cos(Ω Rt). (35) The corresponding spin-flip time Tsfis Tsf=π ΩR. (36) Figure 6 shows the time dependence of the purity and the spin of the condensate for givenα,d0, and ∆ at different interatomic interactions. In figure 6(a) one can see that the increase of /tildewideg1Nenhances the variation of the purity (cf. Fig 5(a)). This variation prevents a precise manipulation of the spin-qubit state in the conde nsate [28]. It follows from figure 6(b) that although increasing the interaction strongly modifies the spin dynamics, it roughly conserves the spin-flip time Tsf= 50π, see (36). To demonstrate the role of the SOC coupling strength αand interatomic interaction at nominally the same Rabi frequency Ω R,we calculated the spin dynamics presented in Figure 7. By comparing Figures 6 and 7(a),(b) we conclude that the increase in th e SOC, at the same Rabi frequency, causes an increase in the variation of the pu rity and of the spin component. These results show that to achieve a required Rabi fr equency and a reliable control of the spin, it is better to increase the driving amplitude d0rather than the spin- orbit couping α.The increase in the SOC strength can result in losing the spin state purity and decreasing the spin length. Figure 7(c) shows the irregu lar spin evolution of the condensate inside the Bloch sphere. In figure 7(a), for α= 0.2 and/tildewideg1N= 20,the purity decreases almost to zero, placing the spin close to the cente r of the Bloch sphere, as can be seen in figure 7(c). It follows from Figures 6(b) and 7(b) t hat in order to protect pure macroscopic spin-qubit states, the Rabi frequenc y should be small. Then, taking into account that the displacement of the spin-projected w ave packet is of the order ofα(/tildewideg1N)2/3and the packet width is of the order of ( /tildewideg1N)1/3, we conclude that forα/greaterorsimilar(/tildewideg1N)−1/3, the purity of the driven state tends to zero. As a result, the Rab i frequency for the pure state evolution is strongly limited by the inte raction between the particles and cannot greatly exceed d0∆/(/tildewideg1N)1/3.Dynamics of a macroscopic spin qubit in spin-orbit coupled B EC 11 (a) 0 50 100 150 2000.00.20.40.60.81.0 tP/LParen1t/RParen1(b) 050100 150 200/MinuΣ1.0/MinuΣ0.50.00.51.0 t/LAngleBracket1Σ/Hat x/LParen1t/RParen1/RAngleBracket1 Figure 7. (Color online) ( a) Purity, ( b) spin component, and ( c) spatial trajectory of the spin inside the Bloch sphere for the driven BEC with α= 0.2,∆ = 0.1, d0= 1 resulting in the same Rabi frequency as in Figure (6). In Figures ( a) and (b) the lines correspond to: /tildewideg1N= 0 (black solid line), /tildewideg1N= 10 (red dashed line), and /tildewideg1N= 20 (blue dotted line). At /tildewideg1N= 0,the time dependence of ∝an}bracketle{tσx∝an}bracketri}htis rather accurately described by cos(Ω Rt) formula, corresponding to a relatively small variation in the purity, 1 −P(t)≪1.With the increase in /tildewideg1N,the purity variation increases and the behavior of ∝an}bracketle{tσx∝an}bracketri}htdeviates stronger from the conventional cos(Ω Rt) dependence. ( c) Here the interaction is fixed to /tildewideg1N= 20. The green and red vectors correspond to the initial and final states of the spin, respectively. Here the final time is fixed to tfin=Tsf, see (36). In addition, it is interesting to note that for /tildewideg1N≫1,where the spin dipole oscillates at the frequency of the order of ( /tildewideg1N)−2/3(as given by (20)), the perturbation due to the trap motion is in the high-frequency limit already at ∆ ≥(/tildewideg1N)−2/3, having a qualitative influence on the spin dynamics [29, 30, 31]. 3.4. Phase factors due to spin-orbit coupling The above results show that the spin-dependent velocity in (31), a long with the interatomic repulsion, results in decreasing the spin state purity an d produces irregular spin motion inside the Bloch sphere. To reduce the effect of these an omalous velocities and to prevent the resulting fast separation (with the relative velo city of 2α) of the spin components, we compensate them by introducing coordinate-dep endent phases (similar to the Bragg factors) in the wave function [32]. To demonstrate th e effect of theseDynamics of a macroscopic spin qubit in spin-orbit coupled B EC 12 (a) 0 50 100 150 2000.00.20.40.60.81.0 tP/LParen1t/RParen1(b) 050100 150 200/MinuΣ1.0/MinuΣ0.50.00.51.0 t/LAngleBracket1Σ/Hat x/LParen1t/RParen1/RAngleBracket1 Figure 8. (Color online) ( a) Purity, ( b) spin component, and ( c) trajectory of the spin inside the Bloch sphere for a driven BEC with initial phases as in (37 ) and α= 0.2,∆ = 0.1, d0= 1.In Figures ( a) and (b) the black solid line is for /tildewideg1N= 0, the red dashed line is for /tildewideg1N= 10, and the blue dotted line is for /tildewideg1N= 20. In Figure (c) the interaction is /tildewideg1N= 20. The green and red vectors correspond to the initial and final states of the spin, respectively ( tfin=Tsf). The initial spin state (a solid-line circle with white filling) is inside the Bloch sphere since ∝an}bracketle{tσx(t= 0)∝an}bracketri}ht=/radicalbig P(0), and P(0)<1 due to the mixed character in the spin subspace of the state in (37 ). phase factors, we construct the initial spinor Ψ α(x,0) by a coordinate-dependent SU(2) rotation [33] of the state with ∝an}bracketle{tσx∝an}bracketri}ht= 1 in (30) as Ψα(x,0) =e−iαx/hatwideσzΨ0(x,0) =ψin(x)√ 2/bracketleftBigg e−iαx eiαx/bracketrightBigg . (37) The expectation value of the velocity (31) at each component ψin(x)exp(±iαx) is zero, and, as a result, the α-induced separation of spin components vanishes, making, as can be easily seen [33], the spinor (37) the stationary state of the spin- orbit coupled BEC in the Gross-Pitaevskii approximation. In terms of the spin density matrix (24), the state (37) is mixed. Fo r a Gaussian condensate with the width w, we get the following expression for the purity at t= 0 P[G] α(0) = exp/bracketleftbig −2(αw)2/bracketrightbig . (38)Dynamics of a macroscopic spin qubit in spin-orbit coupled B EC 13 Instead, in the case of a Thomas-Fermi wave function as in (8), in t he limitαwTF≫1 the initial purity behaves as P[TF] α(0)∼cos2(2αwTF) (αwTF)4. (39) Both cases are characterized by a rapid decrease as αwTFis increased [25]. In the absence of external driving, the spin components and purit y of (37) state remain constant. With the driving, spinor components evolve with tim e and the observables show evolution quantitatively different from that pres ented in Figure 7. In figure 8 we show the analog of figure 7 for the initial state in (37), wit hψin(x) =ψ0(x) given by (2)-(5). By comparing these Figures one can see that the inclusion of the spin-dependent phases in (37) strongly reduces the oscillations in t hex−component of the spin, making the spin trajectory more regular and decreasing t he variations in the purityP(t) compared to the initial state without these phase factors. A general effect of the interatomic interaction can be seen in both fi gures 7 and 8. Namely, for smaller interactions /tildewideg1N, the destructive role of the interatomic repulsion on the spin state purity is reduced and the spin dynamics becomes mo re regular. As a result, at smaller /tildewideg1Nthe spin trajectory is located closer to the Bloch sphere. 4. Conclusions We have considered the dynamics of freely expanding and harmonica lly driven macroscopic spin qubits based on quasi one-dimensional spin-orbit coupled Bose- Einstein condensates in a synthetic Zeeman field. The resulting evolu tion strongly depends in a nontrivial way on the spin-orbit coupling and interaction between the bosons. On one hand, spin-orbit coupling leads to the driven spin qub it dynamics. On the other hand, it leads to a spin-dependent anomalous velocity cau sing spin splitting of the initial wave packet and reducing the purity of the spin state b y decreasing the off-diagonal components of the spin density matrix. This destruct ive influence of spin- orbit coupling is enhanced by interatomic repulsion. The effects of th e repulsion can be interpreted in terms of the increase in the spatial width of the cond ensate and the corresponding decrease in the spin dipole oscillation frequency with t he interaction strength. The joint influence of the repulsion and spin-orbit couplin g can spatially separate and modify the spin components stronger than just the spin-orbit coupling and result in stronger irregularities in the spin dynamics. The spin-flip Rabi frequency remains, however, almost unchangedinthepresence oftheintera tomicinteractions since they lead to only internal forces and do not change the condensat e momentum. As a result, to preserve the evolution within a high-purity spin-qubit sta te, with the spin being always close to the Bloch sphere, the spin-orbit coupling should be weak and, due to this weakness, the spin-rotation Rabi frequency should be sma ll and spin rotation should take a long time. The destructive effect of both the spin-orb it coupling and interatomicrepulsiononthepurityofthespinstatecanbecontrolla blyandconsiderablyDynamics of a macroscopic spin qubit in spin-orbit coupled B EC 14 reduced, although not completely removed, by introducing spin-de pendent Bragg-like phase factors in the initial spinor wave function. Acknowledgments This work was supported by the University of Basque Country UPV/ EHU under program UFI 11/55, Spanish MEC (FIS2012-36673-C03-01 and FI S2012-36673-C03- 03), and Grupos Consolidados UPV/EHU del Gobierno Vasco (IT-47 2-10). S.M. acknowledges EU-funded Erasmus Mundus Action 2 eASTANA, “evr oAsian Starter for the Technical Academic Programme” (Agreement No. 2001-2571/ 001-001-EMA2). References [1] Lin Y-J, Compton R L, K Jim´ enez-Garcia, Porto J V and Spielman I B 2009Nature462628 [2] Lin Y-J, Jim´ enez-Garc´ ıa K and Spielman I B 2011 Nature47183 [3] Zhang J-Y, Ji S-C, Z Chen, Zhang L, Z-D Du, Yan B, G-S Pan, Zha o B, Deng Y-J, Zhai H, Chen S and Pan J-W 2012 Phys. Rev. Lett. 109115301 [4] Ji S-C, Zhang J-Y, L Zhang, Du Z-D, Zheng W, Deng Y-J, Zhai H, Chen Sh and Pan J-W 2014 Nature Physics 10314 [5] Wang P, Yu Z-Q, Fu Z, Miao J, Huang L, Chai Sh, Zhai H and Zhang J 2012Phys. Rev. Lett. 109 095301 [6] Cheuk L W, Sommer A T, Hadzibabic Z, Yefsah T, Bakr W S and Zwierle in M W 2012 Phys. Rev. Lett. 109095302 [7] Liu X-J, Borunda M F, Liu X and Sinova J 2009 Phys. Rev. Lett. 102046402 [8] Stanescu T D, Anderson B and Galitski V 2008 Phys. Rev. A78023616 [9] Li Y, Martone G I and Stringari S 2012 EPL9956008 Martone G I, Li Y, Pitaevskii L P and Stringari S 2012 Phys. Rev. A86063621 [10] Zhang Y, Mao L and Zhang Ch 2012 Phys. Rev. Lett. 108035302 [11] Zhai H 2012 Int. J. Mod. Phys. B261230001 [12] Galitski V and Spielman I B 2013 Nature49449 [13] Zhang Y, Chen G and Zhang Ch 2013 Scientific Reports 31937 [14] Achilleos V, Frantzeskakis D J, Kevrekidis P G and Pelinovsky D E 20 13Phys. Rev. Lett. 110 264101 [15] Ozawa T, Pitaevskii L P and Stringari S 2013 Phys. Rev. A87063610 [16] Wilson R M, Anderson B M and Clark Ch W 2013 Phys. Rev. Lett. 111185303 [17] L¨ u Q-Q and Sheehy D E 2013 Phys. Rev. A88043645 [18] Xianlong G 2013 Phys. Rev. A87023628 [19] Anderson B M, Spielman I B and Juzeliu˜ nas G 2013 Phys. Rev. Lett. 111125301 [20] Dong L, Zhou L, Wu B, Ramachandhran B and Pu H 2014 Phys. Rev. A89011602 [21] Qu Ch, Hamner Ch, Gong M, Zhang Ch and Engels P 2013 Phys. Rev. A88021604 [22] Byrnes T, Wen K and Yamamoto Y 2012 Phys. Rev. A85040306 [23] Nowack K C, Koppens F H L, Nazarov Yu V and Vandersypen L M K 2 007Science3181430 [24] Rashba E I and Efros Al L 2003 Phys. Rev. Lett. 91126405 [25] It is interesting to mention that an increase in spin-orbit coupling strength after a certain value leads to a less efficient spin driving as shown with perturbation theory approach by Li R, You J Q, Sun C P and Nori F 2013 Phys. Rev. Lett. 111086805 [26] This procedure models the von Neumann quantum spin measurem ent, see Sherman E Ya and Sokolovski D 2014 New J. Phys. 16015013 [27] An analysis of the spin-dipole oscillations based on the sum rules wa s presented in [9]Dynamics of a macroscopic spin qubit in spin-orbit coupled B EC 15 [28] This low precision of the spin control can be seen as a general fe ature of the systems where external perturbation strongly drives the orbital motion. See, e .g. Khomitsky D V, Gulyaev L V and Sherman E Ya 2012 Phys. Rev. B85125312 [29] A technique to study and engineer the high-frequency behavio r has been introduced in Bukov M, D’Alessio L and Polkovnikov A 2014 arXiv:1407.4803. Its application to t he spin dynamics of the spin-orbit coupled BEC goes far beyond the scope of this paper , however. [30] Optimal control theory proposed by Budagosky J A and Castr o A 2015 The European Physical Journal B8815 can potentially be applied to the engineering of a d(t)−dependence more complicated than a harmonic oscillation. [31] Xiong B, Zheng J-H and Wang D-W 2014 arXiv:1410.8444 analyzed t he condensate driving in terms of the multichannel quantum interference. [32] These phase-related Josephson effect in the spin-orbit couple d BEC in a double-well potential has been recently analyzed in Garcia-March M A, Mazzarella G, Dell’Ann a L, Juli´ a-D´ ıaz B, Salasnich L and Polls A 2014 Phys. Rev. A89063607 Citro R and Naddeo A 2014 arXiv:1405.5356 [33] Tokatly I V and Sherman E Ya 2010 Phys. Rev. B82161305
1203.2795v1.Impact_of_Dresselhaus_vs__Rashba_spin_orbit_coupling_on_the_Holstein_polaron.pdf
arXiv:1203.2795v1 [cond-mat.str-el] 13 Mar 2012Impact of Dresselhaus vs. Rashba spin-orbit coupling on the Holstein polaron Zhou Li1, L. Covaci2, and F. Marsiglio1 1Department of Physics, University of Alberta, Edmonton, Al berta, Canada, T6G 2J1 2Departement Fysica, Universiteit Antwerpen, Groenenborg erlaan 171, B-2020 Antwerpen, Belgium (Dated: November 6, 2018) We utilize an exact variational numerical procedure to calc ulate the ground state properties of a polaron in the presence of Rashba and linear Dresselhaus spi n-orbit coupling. We find that when the linear Dresselhaus spin-orbit coupling approaches the Rashba spin-orbit coupling, the Van-Hove singularity in the density of states will be shifted away fro m the bottom of the band and finally disappear when the two spin-orbit couplings are tuned to be e qual. The effective mass will be suppressed; the trend will become more significant for low ph onon frequency. The presence of two dominant spin-orbit couplings will make it possible to tune the effective mass with more varied observables. I. INTRODUCTION One of the end goals in condensed matter physics is to achieve a sufficient understanding of materials fabrica- tion and design so as to ‘tailor-engineer’ specific desired properties into a material. Arguably pn-junctions long ago represented some of the first steps in this direction; nowadays, heterostructures1and mesoscopic geometries2 represent further progress towards this goal. Inthefieldof spintronics , wherethespindegreeoffree- dom is specifically exploited for potential applications,3,4 spin-orbit coupling5plays a critical role because con- trol of spin will require coupling to the orbital mo- tion. Spin orbit coupling, as described by Rashba6and Dresselhaus,7is expected to be prominent in two dimen- sional systems that lack inversion symmetry, including surface states. These different kinds of coupling are in principle independently controlled.8,9 The coexistence of Rashba and Dresselhaus spin-orbit coupling has now been realized in both semiconductor quantum wells4,9and more recently in a neutral atomic Bose-Einstein condensate.10When the Rashba and (lin- ear) Dresselhaus spin-orbit coupling strengths are tuned to be equal, SU(2) symmetry is predicted to be recov- ered and the persistent spin helix state will emerge.4,10,11 This symmetry is expected to be robust against spin- independent scattering but is broken by the cubic Dres- selhaus spin-orbit coupling and other spin-dependent scattering which may be tuned to be negligible.4 While we focus on the spin-orbit interaction, other in- teractionsarepresent. In particular, the electron-phonon interaction will be present and may be strong in semi- conductor heterostructures. Moreover, optical lattices12 with cold polar molecules may be able to realize a tune- able Holstein model.13The primary purpose of this work is to investigate the impact of electron-phonon coupling (as modelled by the Holstein model14) on the proper- ties of the spin-orbit coupled system. We will utilize a tight-binding framework; previously it was noted that in the presence of Rashba spin-orbit coupling the vicinity of a van Hove singularity near the bottom of the elec- tron band15–17had a significant impact on the polaronic propertiesofanelectron; withadditional(linear)Dressel-haus spin-orbit coupling the van Hove singularity shifts well away from the band bottom, as the two spin-orbit couplings acquire equal strength. As we will illustrate below, the presence of two separately tunable spin-orbit couplings will result in significant controllability of the electron effective mass. II. MODEL AND METHODOLOGIES We use a tight-binding model with dimensionless Hol- stein electron-phonon coupling of strength g, and with linear Rashba ( VR) and Dresselhaus ( VD) spin-orbit cou- pling: H=−t/summationdisplay <i,j>,s =↑↓(c† i,scj,s+c† j,sci,s) +i/summationdisplay j,α,β(c† j,αˆV1cj+ˆy,β−c† j,αˆV2cj+ˆx,β−h.c.) −gωE/summationdisplay i,s=↑↓c† i,sci,s(ai+a† i)+ωE/summationdisplay ia† iai(1) wherec† i,s(ci,s) creates (annihilates) an electron at site iwith spin index s, anda† i(ai) creates (annihilates) a phononatsite i. The operators ˆVj,j= 1,2arewritten in terms of the spin-orbit coupling strengths and the Pauli matrices as ˆV1=VRˆσx−VDˆσy, andˆV2=VRˆσy−VDˆσx, The sum over iis over all sites in the lattice, whereas < i,j > signifies that only nearest neighbour hopping is included. Other parameters in the problem are the phonon frequency, ωE, and the hopping parameter t, which hereafter is set equal to unity. Without the electron-phononinteractionthe electronic structure is readily obtained by diagonalizing the Hamil- tonian in momentum space. With the definitions S1≡VRsin(ky)+VDsin(kx), S2≡VRsin(kx)+VDsin(ky), (2) we obtain the eigenvalues εk,±=−2t[cos(kx)+cos(ky)]±2/radicalBig S2 1+S2 2(3)2 -5-4-3-2-1 0 1 2 3 4 -3-2-1 0 1 2 3-3-2-1 0 1 2 3VR/t = 0.5 VD/t = 0.5 -5-4-3-2-1 0 1 2 3 4 -3-2-1 0 1 2 3-3-2-1 0 1 2 3VR/t = 0.8 VD/t = 0.2 -5-4-3-2-1 0 1 2 3 4 -3-2-1 0 1 2 3-3-2-1 0 1 2 3VR/t = 0.9 VD/t = 0.1 -5-4-3-2-1 0 1 2 3 4 -3-2-1 0 1 2 3-3-2-1 0 1 2 3VR/t = 0.99 VD/t = 0.01 FIG. 1. Contour plots for the bare energy bands with Rashba- Dresselhaus spin-orbit coupling, for different values of VRand VDwhile the sum is kept constant: VR+VD=tfor these cases. (a) VR=VD= 0.5t, (b)VR= 0.8t,VD= 0.2t, (c) VR= 0.9t,VD= 0.1t, and (d) VR= 0.99t,VD= 0.01t. Note the clear progression from a two-fold degenerate ground sta te to a four-fold degenerate one.and eigenvectors Ψk±=1√ 2/bracketleftBigg c† k↑±S1−iS2/radicalbig S2 1+S2 2c† k↓/bracketrightBigg |0/angbracketright.(4) The ground state energy is E0=−4t/radicalbig 1+(VR+VD)2/(2t2). (5) Without loss of generality we can consider only VR≥0 andVD≥0. Either Rashba and Dresselhaus spin-orbit coupling independently behave in the same manner, and give rise to a four-fold degenerate ground state with wave vectors, ( kx,ky) = (±arctan(VR√ 2t),±arctan(VR√ 2t), (VD= 0), and similarly for VD/negationslash= 0 and VR= 0. With both couplings non-zero, however, the degeneracy be- comes two-fold, with the ground state wave vectors, (kx0,ky0) =±(k0,k0); where k0= tan−1(VR+VD√ 2t). (6) It is clear that the sum of the coupling strengths replaces the strength of either in these expressions, so that hence- forth in most plots we will vary one of the spin-orbit interaction strengths while maintaining their sum to be fixed. Similarly, the effective mass, taken along the diag- onal, is mSO m0=1/radicalbig 1+(VR+VD)2/(2t2), (7) wherem0≡1/(2t) (lattice spacing, a≡1, and/planckover2pi1≡1) is the bare mass in the absence of spin-orbit interaction, andmSOis the effective mass due solely to the spin-orbit interaction. As detailed in the Appendix, the effective mass becomes isotropic when the Rashba and Dressel- haus spin-orbit coupling strengths are equal. The non-interacting electron density of states (DOS) is defined for each band, as Ds(ǫ) =/summationdisplay kδ(ǫ−ǫks) (8) withs=±1. In Fig.2(a) we show the low energy DOS for various values of the spin-orbit coupling strengths, VRandVD, while keeping their sum constant; the low energy van Hove singularity disappears for VR=VD. Note that only D−(ǫ) is shown, as the upper band, with DOS D+(ǫ), exists only at higher energies. Furthermore, informa- tion concerning the upper band can always be obtained through the symmetry D+(ǫ) =D−(−ǫ). (9) In Fig.2(b) we show the value of the density of states at the bottom of the band vs. VD; as derived in the Appendix, theDOSvalueattheminimumenergyisgiven by D−(E0) =1 2πt1/radicalBig 1+(VR+VD)2 2t2−(VR−VD)2 (VR+VD)2.(10)3 /s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54 /s48/s46/s56 /s49/s46/s48 /s49/s46/s50/s48/s46/s48/s48/s46/s49/s48/s46/s50/s48/s46/s51/s48/s46/s52/s48/s46/s53/s48/s46/s54 /s32/s86 /s82/s61/s48/s46/s57/s57/s44/s32/s86 /s68/s61/s48/s46/s48/s49 /s32/s86 /s82/s61/s48/s46/s57/s44/s32/s32/s32/s86 /s68/s61/s48/s46/s49 /s32/s86 /s82/s61/s48/s46/s56/s44/s32/s32/s32/s86 /s68/s61/s48/s46/s50 /s32/s86 /s82/s61/s48/s46/s53/s44/s32/s32/s32/s86 /s68/s61/s48/s46/s53/s68/s95/s40/s69/s41 /s69/s40/s97/s41 /s32 /s74/s117/s109/s112/s32/s111/s102/s32/s100/s101/s110/s115/s105/s116/s121/s32/s111/s102/s32/s115/s116/s97/s116/s101/s115 /s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54 /s48/s46/s56 /s49/s46/s48/s48/s46/s48/s48/s46/s49/s48/s46/s50/s48/s46/s51/s48/s46/s52/s48/s46/s53/s48/s46/s54 /s32/s86 /s82/s61/s48/s44/s32/s86 /s68/s61/s49/s32/s111/s114/s32/s32/s86 /s68/s61/s48/s44/s32/s86 /s82/s61/s49 /s32/s86 /s82/s43/s86 /s68/s61/s48/s46/s52 /s32/s86 /s82/s43/s86 /s68/s61/s49/s68/s95/s40/s69 /s48/s41 /s86 /s82/s47/s40/s86 /s82/s43/s86 /s68/s41 /s32/s40/s98/s41 FIG. 2. (a)The non-interacting density of states D−(E) near the bottom of the band for four values of the spin- orbit coupling strengths: ( VR,VD)/t= (0.5,0.5) (dot-dashed curve), (0 .8,0.2) (dotted curve), (0 .9,0.1) (dashed curve), and (0.99,0.01) (solid curve). Note that for equal coupling strengths there is no van Hove singularity at low energies. (b) The value of the density of states at the bottom of the band (ground state) as a function of VD(while the total cou- pling strength, VR+VD, is held constant. The value of the density of states achieves a minimum value when VR=VD. ForVR= 0 orVD= 0 there is a discontinuity, caused by the transition from a doubly degenerate ground state to a four-fold degenerate ground state.Note that when the coupling strengths are equal, the density of states has a minimum. Also note that when one kind of spin-orbit coupling vanishes, e.g. VR= 0, orVD= 0, there will be a discontinuity for the density of states (the density of states jumps to twice its value). This is caused by a transition from a doubly degenerate groundstate to a four-fold degenerateground state. This discontinuity will also appear for VD≃0 orVR≃0 near the bottom of the band as can be seen from Fig.2(a) for VR= 0.99,VD= 0.01. III. RESULTS WITH THE ELECTRON-PHONON INTERACTION As the electron phonon interaction is turned on, the ground state energy (effective mass) will decrease (in- crease) due to polaron effects. To study the polaron problem numerically, we adopt the variational method outlined by Trugman and coworkers,18,19which is a con- trolled numerical technique to determine polaron prop- erties in the thermodynamic limit exactly. This method was recently further developed20,21to study the polaron problem near the adiabatic limit with Rashba spin-orbit coupling.17This case was also studied in Ref. [16] using the Momentum Average Approximation.22 In Fig. 3, we show the ground state energy and the effective mass correction as a function of the elec- tron phonon coupling λ≡2g2ωE/(4πt),20for various spin-orbit coupling strengths, but with the sum fixed: VR+VD=t. These are compared with the results from the Rashba-Holstein model with VD= 0. Here the phonon frequency is set to be ωE/t= 1.0, which is the typical value used in Ref.[16], and for each value of VR, the ground state energy is compared to the correspond- ing result for λ= 0. The numerical results are compared with results from the MA method and from Lang-Firsov strong coupling theory23,24(see Appendix). In Fig. 3(a), the ground state energycrossesoversmoothly (at around λ≈0.8)fromthe delocalizedelectronregimetothe small polaron regime. In the whole regime, the ground state energy is shifted up slightly as the Dresselhaus spin-orbit coupling, VD, is increased in lieu of the Rashba spin- orbit coupling. We show results for VD≤VR, as the complementary regime is completely symmetric. The MA results agree very well with the exact results and the Lang-Firsov strong coupling results agree well in the λ≥1 regime. Similarly, weak coupling perturbation theory17agrees with the exact results for λ≤1 (not shown). Fig. 3(b) shows the effective mass as a func- tion of coupling strength; it decreases slightly, for a given value ofλ, by increasing VDin lieu of VR. All these results are plotted as a function of the elec- tron phonon coupling strength, λ, as defined above; this definition requires the value of the electron density of states at the bottom of the band, and we have elected to use, for any value of spin-orbit coupling, the value 1 /(4πt appropriate to nospin-orbit coupling. If the actual DOS4 /s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54 /s48/s46/s56 /s49/s46/s48 /s49/s46/s50/s45/s53/s45/s52/s45/s51/s45/s50/s45/s49/s48 /s69/s47/s116/s61/s49/s46/s48 /s32/s76/s70/s44/s32 /s86 /s82/s47/s116/s61/s48/s46/s53 /s32/s76/s70/s44/s32 /s86 /s82/s47/s116/s61/s48/s46/s56 /s32/s76/s70/s44/s32 /s86 /s82/s47/s116/s61/s49/s46/s48/s86 /s82/s47/s116/s43/s86 /s68/s47/s116/s61/s49/s46/s48/s32/s69/s120/s97/s99/s116/s44/s32/s86 /s82/s47/s116/s61/s48/s46/s53 /s32/s69/s120/s97/s99/s116/s44/s32/s86 /s82/s47/s116/s61/s48/s46/s56 /s32/s69/s120/s97/s99/s116/s44/s32/s86 /s82/s47/s116/s61/s49/s46/s48 /s32/s77/s46/s65/s46/s32/s86 /s82/s47/s116/s61/s48/s46/s53 /s32/s77/s46/s65/s46/s32/s86 /s82/s47/s116/s61/s48/s46/s56 /s32/s77/s46/s65/s46/s32/s86 /s82/s47/s116/s61/s49/s46/s48/s40/s69 /s71/s83/s45/s69 /s48/s41/s47/s116 /s32/s40/s97/s41 /s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54 /s48/s46/s56 /s49/s46/s48/s49/s46/s48/s49/s46/s53/s50/s46/s48/s50/s46/s53/s51/s46/s48/s51/s46/s53/s52/s46/s48/s52/s46/s53/s53/s46/s48 /s86 /s82/s47/s116/s43/s86 /s68/s47/s116/s61/s49/s46/s48/s32/s77/s46/s65/s46/s32/s86 /s82/s47/s116/s61/s48/s46/s53 /s32/s77/s46/s65/s46/s32/s86 /s82/s47/s116/s61/s48/s46/s56 /s32/s77/s46/s65/s46/s32/s86 /s82/s47/s116/s61/s49/s46/s48 /s32/s32/s69/s120/s97/s99/s116/s44/s32/s86 /s82/s47/s116/s61/s48/s46/s53 /s32/s32/s69/s120/s97/s99/s116/s44/s32/s86 /s82/s47/s116/s61/s48/s46/s56 /s32/s32/s69/s120/s97/s99/s116/s44/s32/s86 /s82/s47/s116/s61/s49/s46/s48/s109/s42/s47/s109 /s83/s79 /s32/s69/s47/s116/s61/s49/s46/s48/s40/s98/s41 FIG. 3. (a) Ground state energy difference EGS−E0vs.λfor VR/t= 0.5,0.8,1.0 andωE/t= 1.0 while the total coupling strength is kept fixed: VR+VD=t. Exact numerical results are compared with those from the Momentum Average (MA) method. Agreement is excellent. Strong coupling results ar e also plotted (in red) by utilizing the Lang-Firsov (LF) stro ng coupling approximation. Agreement in the strong coupling regime ( λ≥1) is excellent. (b) Effective mass m∗/mSOvs. λ. MA results are plotted (symbols) with the exact numerical results, and again, agreement is excellent. In both (a) and ( b) the polaronic effects are minimized for VR=VD./s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54 /s48/s46/s56 /s49/s46/s48/s45/s48/s46/s53/s45/s48/s46/s52/s45/s48/s46/s51/s45/s48/s46/s50/s45/s48/s46/s49 /s86 /s82/s47/s116/s43/s86 /s68/s47/s116/s61/s49/s46/s48/s32/s77/s46/s65/s46 /s69/s47/s116/s61/s48/s46/s49 /s32/s77/s46/s65/s46 /s69/s47/s116/s61/s48/s46/s50 /s32/s77/s46/s65/s46 /s69/s47/s116/s61/s49/s46/s48 /s32/s69/s120/s97/s99/s116 /s69/s47/s116/s61/s48/s46/s49 /s32/s69/s120/s97/s99/s116 /s69/s47/s116/s61/s48/s46/s50 /s32/s69/s120/s97/s99/s116 /s69/s47/s116/s61/s49/s46/s48/s40/s69 /s71/s83/s45/s69 /s48/s41/s47/s116 /s86 /s68/s47/s116 /s32/s61/s48/s46/s51/s50/s40/s97/s41 /s48/s46/s48 /s48/s46/s50 /s48/s46/s52 /s48/s46/s54 /s48/s46/s56 /s49/s46/s48/s49/s46/s49/s48/s49/s46/s49/s53/s49/s46/s50/s48/s49/s46/s50/s53/s49/s46/s51/s48/s49/s46/s51/s53/s49/s46/s52/s48/s49/s46/s52/s53/s49/s46/s53/s48/s49/s46/s53/s53 /s32/s77/s46/s65/s46 /s69/s47/s116/s61/s48/s46/s49 /s32/s77/s46/s65/s46 /s69/s47/s116/s61/s48/s46/s50 /s32/s77/s46/s65/s46 /s69/s47/s116/s61/s49/s46/s48 /s32/s69/s120/s97/s99/s116 /s69/s47/s116/s61/s48/s46/s49 /s32/s69/s120/s97/s99/s116 /s69/s47/s116/s61/s48/s46/s50 /s32/s69/s120/s97/s99/s116/s44 /s69/s47/s116/s61/s49/s46/s48/s86 /s82/s47/s116/s43/s86 /s68/s47/s116/s61/s49/s46/s48/s109/s42/s47/s109 /s83/s79 /s86 /s68/s47/s116 /s32/s61/s48/s46/s51/s50/s40/s98/s41 FIG.4. (a)Groundstateenergy EGS−E0asafunctionofspin orbit coupling VD/tforωE/t= 0.1,0.2,1.0 with weakelectron phonon coupling, λ= 0.32, and moderate spin-orbit coupling, VR+VD=t. (b) Effective mass m∗/mSOas a function of spin orbit coupling VD/tfor the same parameters. MA results are again compared with the exact numerical results, and are reasonably accurate for these parameters.5 appropriate to the value of spin-orbit coupling were used in the definition of λ, then the effective mass, for ex- ample, would vary even more with varying VDvs.VR (see Fig. 2(b)). Moreover, this variation would be more pronounced for lower values of ωE. In Fig. 4, we show results for the ground state energy and effective mass for different values of the Einstein phonon frequency, ωE; MA results are also shown for comparison. In these plots the electron phonon coupling strength is kept fixed and VDis varied while maintaining the total spin-orbit coupling constant. The ground state energy has a maximum when the two spin-orbit coupling strengths, VDandVR, are tuned to be equal; similarly, theeffectivemasshasaminimumwhenthetwoareequal. As the phonon frequency is reduced the minimum in the effective mass becomes more pronounced. The MA re- sults track the exact results, and, as found previously,17 are slightly less accurate as the phonon frequency be- comes much lower than the hopping matrix element, t. IV. SUMMARY Linear spin-orbit coupling can arise in two varieties; taken on their own, they are essentially equivalent, and their impact on a single electron, even in the presence of electron phonon interactions, will be identical. However,with the ability to tune either coupling constant, in both solid state and cold atom experiments, one can probe the degree of Dresselhaus vs. Rashba spin-orbit coupling throughtheimpactonpolaronicproperties. Theprimary effect of this variation is the electron density of states, where the van Hove singularity can be moved as a func- tion of chemical potential (i.e. doping) through tuning of the spin-orbit parameters. These conclusions are based on exact methods (the so-called Trugman method), and are not subject to approximations. These results have been further corroborated and understood through the Momentum Average approximation, and through weak and strong coupling perturbation theory. The effect is expected to be experimentally relevant since in typi- cal materials with large spin-orbit couplings the phonon frequency is small when compared to the bandwidth, ωE/t≪1. ACKNOWLEDGMENTS This work was supported in part by the Natural Sciences and Engineering Research Council of Canada (NSERC), by ICORE (Alberta), by the Flemish Science Foundation (FWO-Vl) and by the Canadian Institute for Advanced Research (CIfAR). 1See, for example, Quantum Dot Heterostructures , byDieter Bimberg, Marius Grundmann, and Nikolai N. Ledentsov (John Wiley and Sons, Toronto, 1999). 2For example, Mesoscopic Systems , by Y. Murayama (Wiley-VCH, Toronto, 2001). 3S. A. Wolf, D. D. Awschalom, R. A. Buhrman, J. M. Daughton, S. von Molnar, M. L. Roukes, A. Y. Chtchelka- nova and D. M. Treger, Science 294, 1488, (2001). 4J. D. Koralek, C. P. Weber, J. Orenstein, B. A. Bernevig, Shou-Cheng Zhang, S. Mack and D. D. Awschalom, Na- ture,458, 610-613(2009). 5R. Winkler, Spin-Orbit Coupling Effects in Two- Dimensional Electron and Hole Systems (Springer, Berlin, 2003). 6E.I. Rashba, Sov. Phys. Solid State 2, 1109 (1960). 7Dresselhaus, G. Phys. Rev. 100, 580–586 (1955). 8S.K. Maiti, S. Sil, and A. Chakrabarti, arXiv:1109.5842v2. 9L. Meier, G. Salis, I. Shorubalko, E. Gini, S. Sch¨ on, and K. Ensslin, Nature Physics 3, 650 (2007). 10Y. J. Lin, K. Jimenez-Garcia and I. B. Spielman, Nature, 471, 83 (2011). See alsoT. Ozawa and G. Baym, Phys. Rev. A85, 013612 (2012). 11B.A. Bernevig, J. Orenstein, and S.-C. Zhang, Phys. Rev. Lett.97, 236601 (2006). 12I. Bloch, J. Dalibard, and W. Zwerger, Rev. Mod. Phys. 80, 885 (2008). 13F. Herrera and R.V. Krems, Phys. Rev. A 84, 051401(R), (2011). 14T. Holstein, Ann. Phys. (New York) 8, 325 (1959).15E. Cappelluti, C. Grimaldi and F. Marsiglio, Phys. Rev. Lett98, 167002 (2007); Phys. Rev. B 76, 085334 (2007). See also, C. Grimaldi, E. Cappelluti, and F. Marsiglio, Phys. Rev. Lett. 97, 066601 (2006); Phys. Rev. B 73, 081303(R) (2006). 16L. Covaci and M. Berciu, Phys. Rev. Lett 102, 186403 (2009). 17Zhou Li, L.Covaci, M. Berciu, D. Baillie and F. Marsiglio, Phys. Rev. B 83, 195104, (2011). 18S.A. Trugman, in Applications of Statistical and Field The- ory Methods to Condensed Matter , edited by D. Baeriswyl, A.R. Bishop, and J. Carmelo (Plenum Press, New York, 1990). 19J. Bonˇ ca, S.A. Trugman, and I. Batist´ ıc, Phys. Rev. B60,1633 (1999). 20Zhou Li, D. Baillie, C. Blois, and F. Marsiglio, Phys. Rev. B81, 115114, (2010). 21A. Alvermann, H. Fehske, and S.A. Trugman, Phys. Rev. B81, 165113 (2010). 22M. Berciu, Phys. Rev. Lett 97, 036402 (2006). 23I.G. Lang and Yu. A. Firsov, Sov. Phys. JETP16, 1301 (1963); Sov. Phys. Solid State 52049 (1964). 24F. Marsiglio, Physica C 24421, (1995). Appendix A: Density of States and effective mass Expanding εk,−around the minimum energy E0, by defining k′ x=kx±arctan(VR+VD√ 2t),k′ y=ky±6 arctan(VR+VD√ 2t),we have εk,−=E0+˜t1/braceleftbig k′2 x+k′2 y/bracerightbig ±˜t2k′ xk′ y,(A1) where ˜t1=t/braceleftBigg 1+(VR+VD)2 2t2−(VR−VD)2 2(VR+VD)2/radicalbig 1+(VR+VD)2/(2t2)/bracerightBigg ,(A2) and ˜t2=t/braceleftBigg(VR−VD)2 (VR+VD)2/radicalbig 1+(VR+VD)2/(2t2)/bracerightBigg ,(A3) Note that, with generic spin-orbit coupling, the effective mass is in general anisotropic, but when VD=VR, it becomes isotropic. To calculate the density of states at the bottom of the band, from the definition, we have D−(E0+δE) =1 4π2/integraldisplayπ −πdkx/integraldisplayπ −πdkyδ(E0+δE−εk,−), (A4) whereδEis a small amount of energy above the bottom of the band, E0. Around the two energy minimum points there are two small regions which will contribute to this integral. We choose one of them (and then multiply our resultbyafactorof2),thenusethedefinitionsof k′above instead of k, and introduce a small cutoff kc, which is the radius of a small circle around kmin. Thus the integral becomes D−(E0+δE) = 2×1 4π2/integraldisplaykc 0k′dk′/integraldisplayπ −πdθ δ/bracketleftbigg δE−/braceleftbig˜t1+1 2˜t2sin2θ/bracerightbig k′2/bracketrightbigg =1 2πt1/radicalBig 1+(VR+VD)2 2t2−(VR−VD)2 (VR+VD)2(A5) In the weak electron-phonon coupling regime, perturba- tion theory can be applied to evaluate the effective mass; the self energy to first order in λis given by Σweak(ω+iδ) =πλtωE/summationdisplay k,s=±1 ω+iδ−ωE−εk,s.(A6)Theeffectivemasscanbeobtainedthroughthederivative of the self energy m∗ weak mSO= 1−∂ ∂ωΣweak(ω+iδ)|ω=E0.(A7) By inserting the expansion of εk,−around the minimum energyE0into Eqn.[A6] and Eqn.[A7], we obtain the effective mass near the adiabatic limit as m∗ weak mSO= 1+λ 21/radicalBig 1+(VR+VD)2 2t2−(VR−VD)2 (VR+VD)2.(A8) The effective mass has a minimum for VR=VDwhile VR+VDis a constant. Appendix B: Strong coupling theory To investigate the strong coupling regime of the Rashba-Dresselhaus-Holstein model for a single polaron, we use the Lang-Firsov2324unitary transformation H= eSHe−S, whereS=g/summationtext i,σni,σ(ai−a† i). Following pro- cedures similar to those in Ref. (17), we obtain the first order perturbation correction to the energy as E(1) k±=e−g2εk±−g2ωE, (B1) wheregis the band narrowing factor, as used in the Hol- stein model. To find the second order correction to the ground state energy, we proceed as in Ref. (17), and find E(2) k−=−4e−2g2t2+(VR)2+(VD)2 ωE ×/bracketleftbig f(2g2)−f(g2)/bracketrightbig −e−2g2f(g2)ǫ2 k− ωE,(B2) wheref(x)≡∞/summationtext n=11 nxn n!≈ex/x/bracketleftbig 1 + 1/x+ 2/x2+.../bracketrightbig . Thus the ground state energy, excluding exponentially suppressed corrections, is EGS=−2πtλ/parenleftbig 1+2t2+(VR)2+(VD)2 (2πtλ)2/parenrightbig ,(B3) and there is a correction of order 1 /λ2compared to the zeroth order result. Corrections in the dispersion enter in strong coupling only with an exponential suppression. The ground state energy predicted by strong coupling theory has a maximum for VR=VDwhileVR+VDis a constant.
1606.05758v1.Spin_Transport_at_Interfaces_with_Spin_Orbit_Coupling__Formalism.pdf
Spin Transport at Interfaces with Spin-Orbit Coupling: Formalism V. P. Amin1, 2,and M. D. Stiles2 1Maryland NanoCenter, University of Maryland, College Park, MD 20742 2Center for Nanoscale Science and Technology, National Institute of Standards and Technology, Gaithersburg, Maryland 20899, USA (Dated: June 21, 2016) We generalize magnetoelectronic circuit theory to account for spin transfer to and from the atomic lattice via interfacial spin-orbit coupling. This enables a proper treatment of spin transport at inter- faces between a ferromagnet and a heavy-metal non-magnet. This generalized approach describes spin transport in terms of drops in spin and charge accumulations across the interface (as in the standard approach), but additionally includes the responses from in-plane electric elds and o sets in spin accumulations. A key nding is that in-plane electric elds give rise to spin accumulations and spin currents that can be polarized in any direction, generalizing the Rashba-Edelstein and spin Hall e ects. The spin accumulations exert torques on the magnetization at the interface when they are misaligned from the magnetization. The additional out-of-plane spin currents exert torques via the spin-transfer mechanism on the ferromagnetic layer. To account for these phenomena we also describe spin torques within the generalized circuit theory. The additional e ects included in this generalized circuit theory suggest modi cations in the interpretations of experiments involving spin orbit torques, spin pumping, spin memory loss, the Rashba-Edelstein e ect, and the spin Hall magnetoresistance. I. INTRODUCTION The spin-orbit interaction couples the spin and mo- mentum of carriers, leading to a variety of important e ects in spintronic devices. It enables the conversion between charge and spin currents [1, 2], allows for the transfer of angular momentum between populations of spins [3{9], couples charge transport and thermal trans- port with magnetization orientation [10{17], and results in magnetocrystalline anisotropy [18{20]. Many of these e ects already facilitate technological applications. The development of such applications can be assisted by both predictive (yet complicated) rst-principles calculations and clear phenomenological models, which would aid the interpretation of experiments and help to predict device behavior. In multilayer systems, bulk spin-orbit coupling plays a crucial role in spin transport but the role of interfacial spin-orbit coupling remains largely unknown. This un- certainty derives from the uncharacterized transfer of an- gular momentum between carriers and the atomic lattice while scattering from interfaces with spin-orbit coupling. This transfer of angular momentum occurs because a car- rier's spin is coupled via spin-orbit coupling to its orbital moment, which is coupled via the Coulomb interaction to the crystal lattice. Such interfaces behave as either a sink or a source of spin polarization for carriers in a way that does not yet have an accurate phenomenological descrip- tion. In this paper we develop a formal generalization of magnetoelectronic circuit theory to treat interfaces with spin-orbit coupling. In a companion paper, we extract the most important consequences of this generalization vivek.amin@nist.govand show that they capture the dominant e ects found in more complicated Boltzmann equation calculations. To understand the impact of interfacial spin-orbit cou- pling we consider a heavy metal/ferromagnet bilayer, where in-plane currents generate torques on the magne- tization through various mechanisms that involve spin- orbit coupling [6, 7, 21{24]. For example, bulk spin-orbit coupling converts charge currents in the heavy metal into orthogonally- owing spin currents, through a pro- cess known as the spin Hall e ect [25{31]. Upon entering the ferromagnetic layer these spin currents transfer angu- lar momentum to the magnetization through spin trans- fer torques [32{36]. Both the spin Hall e ect and spin- transfer torques have been extensively studied, but addi- tional sources contribute to the total spin torque. These remaining contributions arise from interfacial spin-orbit coupling, which enables carriers of the in-plane charge current to develop a net spin polarization at the inter- face [37{41]. In systems with broken inversion symmetry (such as interfaces) the generation of such spin polariza- tion is known as the Rashba-Edelstein e ect. This spin polarization can exert a torque on any magnetization at the interface via the exchange interaction [7, 42]. A re- cent experiment suggests that this mechanism can induce magnetization switching alone, without relying on the bulk spin Hall e ect [24]. The spin torque driven by the Rashba-Edelstein e ect is typically studied by con ning transport to the two- dimensional interface. Semiclassical models can capture the direct and inverse Rashba-Edelstein e ects [43{46] in this scenario. However, such models are not realistic de- scriptions of bilayers, in which carriers scatter both along and across the interface. Since spin transport across the interface is a ected by the transfer of angular momentum to the atomic lattice, the resulting spin torques are modi- ed in ways that two-dimensional models cannot capture.arXiv:1606.05758v1 [cond-mat.mes-hall] 18 Jun 20162 FIG. 1. (Color online) (a) A heavy metal/ferromagnet bilayer subject to an in-plane electric eld. The axes directly below the bilayer is used to describe electron ow, where the z-axis points normal to the interface plane. The other axes is used to describe spin orientation, where the direction `points along the magnetization while the directions dandfspan the plane transverse to `. (b) Depiction of the physics described by the spin mixing conductance. Spins incident from the heavy metal brie y precess around the magnetization when re ecting o of the interface. The imaginary part of the spin mixing conductance describes the extent of this precession. Interfacial spin-orbit coupling changes the e ective magnetic eld seen by carriers during this process in a momentum-dependent way; this alters the precession axis for each carrier and thus modi es the spin mixing conductance. (c) Depiction of the loss of spin polarization that carriers experience while crossing interfaces with spin-orbit coupling. Without interfacial spin-orbit coupling, carriers retain the portion of their spin polarization aligned with the magnetization, but lose the portion polarized transversely to the magnetization due to dephasing processes just within the ferromagnet. With interfacial spin-orbit coupling, carriers trade angular momentum with the atomic lattice; this leads to changes in all components of the spin polarization. This phenomenon, known as spin memory loss, a ects each component di erently. The panel illustrates only the loss in spin polarization aligned with magnetization. (d) Depiction of interfacial spin-orbit scattering in the presence of an in-plane electric eld. Interfacial spin-orbit coupling allows for spins aligned with the magnetization to become misaligned upon re ection and transmission. For the scattering potential discussed in Sec. IV, the spin of a single re ected carrier cancels the spin of a single carrier transmitted from the other side of the interface. However, a net cancellation of spin is prevented if the total number of incoming carriers di ers between sides, as can happen in the presence of in-plane current ow. This occurs because an in-plane electric eld drives two di erent charge currents within each layer; this forces the number of carriers with a given in-plane momentum to di er on each side of the interface. The scattered carriers then carry a net spin polarization and a net spin current. The various contributions to spin torques in bilayers re- main dicult to distinguish experimentally [22, 23] in part because of the lack of models that accurately cap- ture interfacial spin-orbit coupling [42]. Interfacial spin-orbit coupling may play an impor- tant role in other phenomena. Spin pumping is one example; it describes the process in which a precess- ing magnetization generates a spin current [8]. In heavy metal/ferromagnet bilayers, the pumped spin cur- rent ows from the ferromagnet into the heavy metal, where the inverse spin Hall e ect generates an orthog- onal charge current [47{51]. However, because inter- facial spin-orbit coupling transfers spin polarization to the atomic lattice, it modi es the pumped spin current as it ows across the interface. This transfer of spin polarization remains uncharacterized in many systems, thus contributing to inconsistencies in the quantitative interpretation of experiments [9, 52{54]. Another exam- ple, known as the spin Hall magnetoresistance, describes the magnetization-dependent in-plane resistance of heavy metal/ferromagnet bilayers [55{61]. Currently this e ect is attributed to magnetization-dependent scattering at the interface, but may also contain a contribution frominterfacial spin-orbit scattering. The impact of interfa- cial spin-orbit coupling on these e ects remains unclear due to the absence of appropriate models with which to analyze the data. Magnetoelectronic circuit theory is the most frequently used approach to model spin currents at the interface between a non-magnet and a ferromagnet. It describes spin transport in terms of four conductance parameters, where drops in spin-dependent electrochemical poten- tials across the interface play the role of traditional volt- ages. However, the theory cannot describe interfaces with spin-orbit coupling because it does not consider spin- ip processes due to spin-orbit coupling at the interface. Fig. 1(a) depicts a typical scattering process described by one of these conductance parameters. Given its success in describing spin transport in normal metal/ferromagnet bilayers, generalizing magnetoelectronic circuit theory to include interfacial spin-orbit coupling would make it a valuable tool for describing heavy metal/ferromagnet bi- layers. To generalize magnetoelectronic circuit theory one must consider all the ways that interfacial spin-orbit cou- pling potentially a ects spin transport. One such e ect,3 known as spin memory loss, describes a loss of spin cur- rent across interfaces due to spin-orbit coupling. We il- lustrate a process that contributes to spin memory loss in Fig. 1(b). This loss occurs when the atomic lattice at the interface behaves as a sink of angular momen- tum. Recent work [62] incorporates this behavior into a theory for spin pumping, but descriptions of this e ect date back to over a decade ago [63{66]. Thus generaliz- ing magnetoelectronic circuit theory for interfaces with spin-orbit coupling requires accounting for spin memory loss. By incorporating spin- ip processes at the inter- face into magnetoelectronic circuit theory, one can treat this aspect of the phenomenology of interfacial spin-orbit coupling. Another important consequence of interfacial spin- orbit coupling is that in-plane electric elds can create spin currents that ow away from the interface. First principles calculations of Pt/Py bilayers suggest that a greatly enhanced spin Hall e ect occurs at the inter- face (as compared to the bulk) that could generate such spin currents [67]. This suggests that in-plane electric elds (and not just drops in spin and charge accumula- tions across the interface) must play a role in generaliza- tions of magnetoelectronic circuit theory. It also suggests that one cannot con ne transport to the two-dimensional interface when describing the e ect of in-plane electric elds. Instead, one must consider transport both along and across the interface. Some of the consequences of this three-dimensional picture have been investigated in mul- tilayer systems containing an insulator [68, 69]. The only semiclassical calculations of three-dimensional metallic bilayers are based on the Boltzmann equation [42]. Like spin memory loss, these spin currents must be included in generalizations of magnetoelectronic circuit theory to fully capture the e ect of interfacial spin-orbit coupling. In the following we give a semiclassical picture of how such spin currents arise, and how they exert magnetic torques that are typically not considered in bilayers. Fig. 1(c) depicts how spins aligned with the magneti- zation scatter from an interface with spin-orbit coupling. For the scattering potential discussed in Sec. IV, single re ected and transmitted spins cancel on each side of the interface. However, the netcancellation of spin is avoided if the number of incoming carriers di ers between sides. In the simplest scenario, this occurs if the in-plane elec- tric eld drives di erent currents within each layer, so that the occupancy of carriers di ers on either side for a given in-plane momentum. We nd that through this mechanism, carriers subject to interfacial spin-orbit scat- tering can carry a net spin current in addition to exhibit- ing a net spin polarization. If the net spin polarization is misaligned with the magnetization, it can exert a torque on the magnetization at the interface. This describes the contribution to the spin torque normally associated with the Rashba-Edelstein e ect (discussed earlier). However, the spin currents created by interfacial spin-orbit scatter- ing can ow away from the interface, and those that ow into the ferromagnet exert additional torques. Althoughthese spin currents generate torques via the spin-transfer mechanism, they arise from interfacial spin-orbit scat- tering instead of the spin Hall e ect. This mechanism, which cannot be captured by con ning transport to the two-dimensional interface, is not usually considered when analyzing spin torques in bilayers. However, it can con- tribute to the total spin torque in important ways. For instance, it allows for spin torques generated by inter- facial spin-orbit coupling to point in directions typically associated with the spin Hall e ect. The spin polariza- tion and ow directions of these spin currents are not required to be orthogonal to each other or the electric eld, unlike the spin currents generated by the spin Hall e ect in in nite bulk systems. More work is needed to determine how this semiclassical description of interfacial spin current generation compares with the rst principles description of an enhanced interfacial spin Hall e ect [67]. In this paper, we generalize magnetoelectronic circuit theory to include interfacial spin-orbit coupling. Not only does interfacial spin-orbit coupling modify the conduc- tance parameters introduced by magnetoelectronic cir- cuit theory, it requires additional conductivity parame- ters to capture the spin currents that arise from in-plane electric elds and spin-orbit scattering. Furthermore, the transfer of angular momentum between carriers and the atomic lattice at the interface alters the spin torque that carriers can exert on the magnetization; this introduces additional parameters that are needed to distinguish spin torques from spin currents. However, we nd that many of the parameters in this generalized circuit theory may be neglected when modeling spin-orbit torques in bilayer systems, and that including the conductivity and spin torque parameters is more important than modifying the conductance parameters. As with magnetoelectronic cir- cuit theory, we provide microscopic expressions for most parameters. In a companion paper, to highlight the utility of the proposed theory, we produce an analytical model describ- ing spin-orbit torques caused by the spin-Hall and inter- facial Rashba-Edelstein e ects. We achieve this by solv- ing the drift-di usion equations with this generalization of magnetoelectronic circuit theory. In that paper, we focus on only the parameters that describe the response of in-plane electric elds, and neglect all other changes to magnetoelectronic circuit theory. We show that this simpli ed approach captures the most important e ects found in Boltzmann equation calculations of a model sys- tem. In this paper, we discuss the complete generaliza- tion of magnetoelectronic circuit theory in the presence of interfacial spin-orbit coupling. In Sec. II of this paper we describe spin transport at in- terfaces with and without interfacial spin-orbit coupling. In Sec. III we motivate the derivation of all parame- ters, leaving some details for appendices A and B. In Sec. IV we perform a numerical analysis of each bound- ary parameter for a scattering potential relevant to heavy metal/ferromagnet bilayers. This analysis allows us to determine which parameters matter the most in these4 systems. Finally, in Sec. V we discuss implications of our theory on experiments involving spin orbit torque, spin pumping, the Rashba-Edelstein e ect, and the spin Hall magnetoresistance. II. SPIN AND CHARGE TRANSPORT AT INTERFACES In the following we discuss the general phenomenology of spin transport at interfaces with and without spin or- bit coupling. We rst describe some conventional spin transport models to build up to the proposed model, and refrain from presenting explicit expressions of any param- eters until later sections. A. Collinear spin transport In the absence of spin- ip processes one often assigns separate current densities for majority ( j") and minority (j#) carriers, i.e. j"=G""j#=G##: (1) HereG"=#denotes the spin-dependent interfacial conduc- tance, while  "=#refers to the drop in quasichemical potential for each carrier population across the interface. We may then de ne charge ( c) and spin ( s) components for the drop in quasichemical potential c= "+ # (2) s= "#; (3) and for the current densities jc=j"+j# (4) js=j"j#: (5) across the interface. Using the following modi ed con- ductance parameters G=1 2 G"G# ; (6) we may rewrite Eq. (1) as js jc! = G+G GG+! s c! (7) instead. In this case both spin and charge currents are continuous across the interface. B. Magnetoelectronic Circuit Theory When describing spin orientation in bulk ferromagnetic systems, the magnetization direction provides a naturalspin quantization axes. However, at the interface be- tween a non-magnet and a ferromagnet, the net spin po- larizations of each region need not align. To account for this, one must consider spins in the non-magnet that point in any direction. In the ferromagnet, spins are mis- aligned with the magnetization near the interface but be- come aligned in the bulk. This occurs because spins pre- cess incoherently around the magnetization; eventually the net spin polarization transverse to the magnetization vanishes. In transition metal ferromagnets and their al- loys, this dephasing happens over distances smaller than the spin di usion length. To describe electron ow and spin orientation in non- magnet/ferromagnet bilayers, we use two separate coor- dinate systems. For electron ow, we choose the x=y plane to lie along the interface and the z-axis to point perpendicular to it. The interface is located at the z-axis origin, and z= 0andz= 0+describe the regions just within the non-magnet and ferromagnet respectively. To describe spin orientation, we choose the direction `to be along the magnetization ( ^`=^m) and the directions dandfto be perpendicular to ^`. The damping-like (d) and eld-like ( f) directions point along the vectors ^d/^m[^m(E^z)] and ^f/^m(E^z) re- spectively. This provides a convenient coordinate system for describing spin-orbit torques, because torques with a damping-like component push the magnetization towards theE^zdirection, while those with a eld-like com- ponent force the magnetization to precess about E^z. We rst de ne the spin and charge accumulations at the interface (  ), where the index 2[d;f;`;c;`;c] describes the type of accumulation. The rst four indices denote the spin ( d,f,l) and charge ( c) accumulations in the non-magnet at z= 0. The last two indices describe the spin (`) and charge ( c) accumulations in the fer- romagnet at z= 0+. In the ferromagnet we omit spin accumulations aligned transversely to the magnetization, due to the dephasing processes discussed above. Note that the charge and spin components of  have units of voltage. We then de ne the spin and charge current densities owing out-of-plane ( jz ) in an identical fash- ion. The charge and spin components of jz have the units of number current density [70]. We refer to as the spin/charge index. One may rede ne any tensor that contains spin/charge indices in another basis when useful. For instance, we may write the spin accumulations and spin current den- sities with longitudinal spin polarization in terms of av- erages and di erences across the interface: `=1 2 `` ; `=1 2 `+` ; (8) jz`=1 2 jz`jz` ; jz`=1 2 jz`+jz` :(9) We may de ne similar expressions for the charge accumu- lations and charge current densities. As we shall see, this basis ( 2[d;f;`;c;`;c]) provides a more physically transparent representation of all quantities.5 In the absence of interfacial spin-orbit coupling, the spin current polarized along the magnetization direction remains conserved. However, the spin current with po- larization transverse to the magnetization dissipates en- tirely upon leaving the normal metal. The interface ab- sorbs part of this spin current, while the remaining por- tion quickly dissipates within the ferromagnet due to a precession-induced dephasing of spins. The total loss of spin current then results in a spin transfer torque. Figure 2 depicts this process by use of solutions to the drift-di usion equations. In this situation, one may show [71, 72] that the spin and charge current densities at z= 0become jz =GMCT  (10) for a conductance tensor GMCT given by GMCT=0 BBBBBBBB@d f lclc d Re[G"#]Im[G"#] 0 0 0 0 f Im[G"#] Re[G"#] 0 0 0 0 l 0 0 G+G0 0 c 0 0 GG+0 0 l 0 0 0 0 0 0 c 0 0 0 0 0 01 CCCCCCCCA: (11) This formalism|known as magnetoelectronic circuit theory|disregards spin currents and accumulations in the ferromagnet with polarization transverse to the mag- netization (due to the precession-induced dephasing de- scribed above). This amounts to assuming that the pro- cesses occurring in the shaded regions of Fig. 2(b) hap- pen entirely at the interface instead. While this restric- tion helps to reduce the number of required parameters, it need not apply to non-ferromagnetic systems or ex- tremely thin ferromagnetic layers. Note that the rows corresponding to average and discontinuous quantities are switched from the columns corresponding to those quantities. This is done to emphasize that drops in ac- cumulations cause average currents in magnetoelectronic circuit theory. Equation (11) implies that spin populations polarized transverse to the magnetization decouple from those po- larized longitudinal to it. The charge and longitudinal spin current densities still obey Eq. (7), whereas the transverse (non-collinear) spin current densities experi- ences a nite rotation in polarization about the magneti- zation axis. Note that the spin mixing conductance G"# governs the latter phenomenon. In general, one obtains all parameters via integrals of the transmission and/or re ection amplitudes over the relevant Fermi surfaces. FIG. 2. (Color online) Spin current densities plotted ver- sus distance from the interface, calculated using the drift- di usion equations. Panel (a) treats the case without in- terfacial spin-orbit coupling using magnetoelectronic circuit theory as boundary conditions, whereas panel (b) treats the case with interfacial spin-orbit coupling by using Eq. (12) as boundary conditions instead. Due to precession-induced de- phasing,jzdandjzfdissipate entirely within the ferromag- net some distance from the interface (denoted by the purple dashed line). With no interfacial spin-orbit coupling, the spin current density polarized along the magnetization ( jzl) is con- served, while the spin current densities polarized transversely (jzdandjzf) exhibit discontinuities at the interface. With interfacial spin-orbit coupling, all spin currents are discon- tinuous at the interface. Furthermore, interfacial spin-orbit coupling introduces additional sources of spin current via the conductivity i and torkivity FM tensors (when an in-plane electric eld is present). These sources may oppose the spin currents that develop in the bulk. For example, the inclusion of interfacial spin-orbit coupling leads jzfto switch signs near to the interface, as seen by comparing panels (a) and (b). C. Spin transport with interfacial spin orbit coupling To generalize magnetoelectronic circuit theory, i.e. Eq. (10), to account for interfacial spin orbit coupling and in-plane electric elds, we introduce the following expression for the spin and charge current densities at the interface: ji =Gi  +i ~E: (12) Here we use a scaled electric eld de ned by ~EE=e so that the elements of the tensor i have units of con- ductivity. Without loss of generality, we assume that the electric eld points along the xaxis. The explosion of new parameters (relative to magneto-6 electronic circuit theory) is an unfortunate consequence of spin- ip scattering at the interface. Like magnetoelec- tronic circuit theory, one may express each parameter as an integral of scattering amplitudes over the relevant Fermi surfaces; to discover which parameters may be ne- glected we numerically study these integrals in Sec. IV. Here, we discuss the overarching implications of this model. In particular, three new concepts emerge from the above expression: First of all, the current density ji now includes an index describing its direction of ow ( i2[x;y;z ]), which was previously assumed to be out-of-plane. In this gen- eralization, a buildup of spin and charge accumulation at interfaces may lead to spin and charge currents that ow both in-plane and out-of-plane. The treatment of in-plane currents close to the interface requires not only the evaluation of Eq. (12), but also an extension of the drift-di usion equations themselves. Secondly, Eq. (12) depends on values of the spin and charge accumulations from each side of the interface, rather than di erences in those values across the inter- face. This suggests that currents result from both drops in accumulations andnon-zero averages of spin accumu- lation at the interface [73]. Finally, interfacial spin-orbit scattering results in a conductivity tensor ( i ) that drives spin currents in the presence of an in-plane electric eld. This feature rep- resents the greatest conceptual departure from previous theories describing spin transport and is motivated by results from the Boltzmann equation. Figure 2 describes how some of these properties alter solutions of the drift- di usion equations, as compared with magnetoelectronic circuit theory. Without interfacial spin-orbit coupling the in-plane conductance tensors ( Gx andGy ) vanish, implying that accumulations do not create in-plane currents in this scenario. The conductivity tensor vanishes as well. Spin transport transverse to the magnetization still decouples from that longitudinal to it, and magnetoelectronic cir- cuit theory is recovered. In the presence of interfacial spin-orbit coupling, none of the tensors elements intro- duced in Eq. (12) necessarily vanish, and spin transport in all polarization directions becomes coupled. However, for the interfacial scattering potential studied in Sec. IV, many parameters di er by orders of magnitude; thus cer- tain parameters may be neglected on a situational basis. D. Spin-orbit torques Without interfacial spin-orbit coupling, spin and charge accumulations at an interface create both a spin polarization and spin currents. The spin polarization de- velops atz= 0 and exerts a torque on any magnetization at the interface via the exchange interaction. The spin current that develops at z= 0+exerts an additional torque by transferring angular momentum to the ferro- magnetic region via dephasing processes. For simplicity,we assume that this spin current transfers all of its angu- lar momentum to the magnetization rather than the bulk atomic lattice. We do so under the assumption that the dephasing processes within the ferromagnet diminish spin currents faster than the spin di usive processes caused by bulk spin-orbit coupling. All of the incident transverse spin current is then lost at the interface ( z= 0) or in the bulk of the ferromagnet ( z > 0), and carriers can only exchange angular momentum with the magnetiza- tion. Thus the spin current at z= 0, which represents the incident ux of angular momentum on the magne- tized part of the bilayer, equals the total spin torque on the system. Furthermore, the spin torques at z= 0 and z>0 add up to equal the spin current at z= 0. However, at interfaces with spin-orbit coupling, the atomic lattice behaves as a reservoir that carriers may transfer angular momentum to. In this scenario, carri- ers exert spin torques on both the magnetization and the lattice. We cannot compute spin torques solely from the spin currents described by Eq. (12) if we are to account for the losses to this additional reservoir of angular mo- mentum. Thus, we introduce a separate expression for the total spin torque on the bilayer: =   + ~E; (13) Note that the index 2[d;f] describes the directions transverse to the magnetization, since spin torques only point in those directions. The tensor , known as the torkance, describes contributions to the spin torque from the buildup of spin and charge accumulation at an in- terface. The tensor , which we call the torkivity , cap- tures the corresponding contributions from an external, in-plane electric eld. The torkivity tensor originates from interfacial spin-orbit scattering, much like the con- ductivity tensor introduced earlier. We may separate the total spin torque into two contri- butions:  = mag  + FM  (14) = mag + FM : (15) The rst tensors on the right hand side of Eqs. (14) and (15) describe torques exerted by the spin polarization atz= 0. The second tensors describe the spin torque exerted in the bulk of the ferromagnet ( z > 0). Both torques are exerted on the magnetization rather than on the atomic lattice. Here we assume that the torque at z > 0 equals the transverse spin current at z= 0+as before. Thus, the spin torques exerted at z= 0 and z > 0 are both included in the torkance and torkivity tensors. Without interfacial spin-orbit coupling, the torkivity tensor vanishes and the torkance tensor  becomes identical to Gz . This indicates that the transverse spin current at z= 0equals the total spin torque, as ex- pected. In the presence of interfacial spin-orbit coupling, the lattice also receives angular momentum from carriers;7 in this case  6=Gz and 6= 0. Thus, by comput- ing the tensors introduced in Eq. (13), one may calculate spin-orbit torques such that the lattice torques are ac- counted for. Furthermore, Eqs. (14) and (15) allow one to separate the total spin torque into its interfacial and bulk ferromagnet contributions. III. DERIVATION OF BOUNDARY PARAMETERS Interfacial spin-orbit coupling causes both momen- tum and spin-dependent scattering at interfaces. If the incident distribution of carriers depends on mo- mentum and/or spin, outgoing carriers may become spin-polarized via interfacial spin-orbit scattering. This gives rise to non-vanishing accumulations, currents, and torques, which are related by Eqs. (12) and (13). We now motivate these relationships, which can be expressed in terms of scattering amplitudes. We do so by approx- imating the non-equilibrium distribution function near the interface. We rst consider the total distribution function f (k), which gives the momentum-dependent occupancy of car- riers described by the spin/charge index . In equilib- rium, this distribution function equals the Fermi-Dirac distribution feq (" k). Just out of equilibrium, f (k) is perturbed as follows f (k) = feq (" k) +@feq @" kg (k); (16) whereg (k) denotes the non-equilibrium distribution function. The equilibrium distribution functions vanish for 2[d;f;` ] since the non-magnet exhibits no equi- librium spin polarization. However, the non-equilibrium distribution functions for all spin/charge indices are gen- erally non-zero. To obtain the tensors introduced in Eqs. (12) and (13), we must evaluate g (k) near the interface. One could evaluateg (k) by solving the spin-dependent Boltzmann equation for the bilayer system. In this approach one cap- tures spin transport both in the bulk and at the interface. A more practical approach is to assume some generic form forg near the interface that is physically plau- sible. This approach yields boundary conditions suit- able for simpler bulk models of spin transport such as the drift-di usion equations. In the companion paper, we show that solving the drift-di usion equations using these boundary conditions produces quantitatively simi- lar results to solving the Boltzmann equation. For simplicity, we assume that spherical Fermi surfaces describe carriers in both layers. Later we generalize this formalism to describe non-trivial electronic structures. In the non-magnet, all carriers belong to the same Fermi surface. In the ferromagnet, majority ( ") and minor- ity (#) carriers belong to di erent Fermi surfaces. Thus we use the spin/charge basis 2[d;f;`;c;";#], since in this model carriers belonging to those populationshave well-de ned Fermi surfaces and velocities. The ten- sors derived in this section may be expressed in other spin/charge bases by straightforward linear transforma- tions. To approximate g at the interface we use the following expression: gin (kjj) =e q +~EfE (kjj) ; (17) Equation (17) represents the portion of g incident on the interface, where kjjdenotes the in-plane momentum vector and eequals the elementary charge. The right hand side of Eq. (17) describes two pieces of the incom- ing distribution function; Fig. 3 depicts both pieces over k-space for each side of the interface. The rst term captures spin/charge currents incident on the interface. They may arise, for example, from the bulk spin Hall e ect or ferromagnetic leads. The quantities q denote the isotropic spin/charge polarization of those currents. The second term represents the anisotropic contribution to the distribution function caused by an external electric eld. We remind the reader that the scaled electric eld ~Epoints along the xaxis. The simplest approximation forfE (kjj) is to use the particular solution of the Boltz- mann equation in the relaxation time approximation: fE (kjj) =evx (kjj)8 >>>>>< >>>>>:0 2[d;f;` ]  =c " =" # =#(18) This term describes the in-plane charge current caused by the external electric eld, but also describes an in- plane spin current polarized opposite to the magnetiza- tion in the ferromagnet. The momentum relaxation times in the ferromagnet di er between majority ( ") and mi- nority (#) carriers. In the non-magnet, the momentum- relaxation time ( ) is renormalized by bulk spin- ip pro- cesses (see appendix A). The outgoing distribution function gout (kjj) =S (kjj)gin (kjj); (19) is speci ed by the incoming distribution function and the unitary scattering coecients S , given by S jvz (kjj)j jvz (kjj)jS0 (kjj); (20) where S0 =8 >>>>>>>< >>>>>>>:1 2tr ry r  ; 2[d;f;`;c ] 1 2tr ty t  2[d;f;`;c ]; 2[";#] 1 2tr (t)y t  2[";#]; 2[d;f;`;c ] 1 2tr (r)y r  ; 2[";#] (21)8 FIG. 3. (Color online) Non-equilibrium distribution functions g (k) in the presence of interfacial spin-orbit scattering, resulting from an (a) incident spin and charge accumulation and an (b) in-plane external electric eld. The images depict g (k) on each side of the interface plotted over k-space. The gray spheres represent the equilibrium Fermi surface. The colored surfaces represent the non-equilibrium perturbation to the Fermi surface, given by the charge distribution gc(k) (not to scale). The arrows denote the spin distribution g(k). The blue and red regions represent the wavevectors pointing incident and away from the interface respectively. (a) Scenario in which the incident carriers exhibit a net spin and charge accumulation. The spin-polarization of the outgoing carriers di ers from the incident carriers due to interfacial spin-orbit scattering. The total spin/charge current density ( ji ) and the resulting spin torques ( ) are related to the total spin/charge accumulation (  ) by the tensors Gi and  respectively. (b) Scenario in which the incident carriers are subject to an in-plane electric eld. The in-plane electric eld drives two di erent charge currents on each side of the interface, since each layer possesses a di erent bulk conductivity. This shifts the occupancy of carriers (i.e. the charge distribution) di erently on each side of the interface. When spin- unpolarized carriers scatter o of an interface with spin-orbit coupling they become spin-polarized. Because the occupancy of incident carriers was asymmetrically perturbed at the interface, a net cancellation of spin is avoided in even the simplest scattering model. The resulting spin/charge currents and spin torques are captured by the tensors i and respectively. Note that for a ferromagnetic layer, in-plane electric elds also create incident in-plane spin currents as well (suppressed for clarity in this gure). Here we de ne the Pauli vector  such thatd=x, f=y,`=z, and c= 1 0 0 1! ; "= 1 0 0 0! ; #= 0 0 0 1! :(22) The coecient S0 gives the strength of scattering for carriers with spin/charge index into those with spin/charge index . The scattering coecients depend on the 22 re ection and transmission matrices for spins pointing along the magnetization axis. In particular, the matricesrandtdescribe re ection and transmission respectively into the ferromagnet. The matrices randt describe re ection and transmission into the non-magnet. Note that the density of states and Fermi surface area element di er between incoming and outgoing carriers. Thus to conserve particle number one must include the ratio of velocities within the scattering coecients, as done in Eq. (20). We obtain all non-equilibrium quantities near the in- terface by integrating g over the relevant Fermi surfaces. We note that the outgoing part of g includes the con- sequences of interfacial scattering, since it depends on the scattering coecients. For example, the interfacial exchange interaction leads to spin-dependent scattering, which is captured by the di erence in the diagonal ele-ments of the 22 re ection and transmission matrices. On the other hand, the interfacial spin-orbit interaction introduces spin- ip scattering, which is captured by the o -diagonal elements within these matrices. Thus, to describe the consequences of interfacial spin-orbit scat- tering we must not limit the form of the re ection and transmission matrices as was often done in the past. We write the current density ji for carriers with spin/charge index owing in direction i2[x;y;z ] as follows: ji =1 ~(2)31 vF Z FS d2kvi (k)g (k) (23) Note that all integrals run over the Fermi surface cor- responding to the population with spin/charge index . The quantity vF denotes the Fermi velocity for that population. To de ne the accumulations  we follow the example of magnetoelectronic circuit theory [71, 72] and assume that the incoming currents behave as if they originate from spin-dependent reservoirs. This implies that the incoming polarization q approximately equals the accumulation  at the interface. We have now discussed the requirements for deriv- ing the conductance and conductivity tensors found in Eq. (12). We obtain these tensors by plugging Eqs. (17) and (19) into Eq. (23) and noting that q  . In doing9 so we write the currents ji in terms of the accumulations  and the in-plane electric eld ~E. From the resulting expressions one then obtains formulas for the conduc- tance and conductivity tensors in terms of the interfacial scattering coecients. We outline this remaining process in appendix A. In appendix C we generalize those ex- pressions for the case of non-trivial electronic structures, which allows one to compute the conductance and con- ductivity tensors for realistic systems. Having discussed the currents that arise from interfa- cial spin-orbit scattering, we now discuss the spin torques caused by the same phenomenon. The transverse spin po- larization at z= 0 exerts a torque on any magnetization at the interface via the exchange interaction. The trans- verse spin current at z= 0+exerts a torque by transfer- ring angular momentum to the ferromagnet. The total spin torque then equals the sum of these two torques. To describe the spin torque at z= 0, we must compute the spin polarization at the interface. To accomplish this we de ne the following matrix T =8 < :1 2tr (t)yt  2[d;f;`;c ] 1 2tr tyt  2[";#](24) which describes phase-coherent transmission from all populations into transverse spin states at the interface. We may then compute the ensemble average of spin den- sityhsiatz= 0 as follows: hsi=1 ~(2)3X 1 vF Z FS 2ind2kT (kjj)gin (kjj): (25) The torque at z= 0 is then given by mag =0+Z 0dzJex ~ hsi^m ; (26) whereJexequals the exchange energy at the interface. We evaluate this integral over the region that describes the interface, where the exchange interaction and strong spin-orbit coupling overlap. Note that the cross prod- uct hsi^m =0hs0iis evaluated by computing Eq. (25). To describe the spin torque at z= 0+, we introduce an additional scattering matrix: S =8 < :1 2tr (t)yt  2[d;f;`;c ] 1 2tr (r)yr  2[";#](27) This scattering matrix is used to calculate the transverse spin current at z= 0+. Since this spin current rapidly de- phases, it contributes entirely to the spin torque exerted on the ferromagnet. Note that the currents discussedpreviously corresponded to carriers with well-de ned ve- locities. However, transverse spin states in the ferromag- net consist of linear combinations of majority and minor- ity spin states. Since these spin states possess di erent phase velocities, the velocities of transverse spin states oscillate over position. These states also posses di erent group velocities, and wave packets with transverse spin travel with the average group velocity. The transverse spin current at z= 0+then equals FM =1 ~(2)3X Z 2DBZdkjjvz(kjj) vz (kjj)S (kjj)gin (kjj); (28) where vz(kjj)1 2 vz"(kjj) +vz#(kjj) (29) gives the average group velocity of carriers in the ferro- magnet. Note that we write this integral over the max- imal two-dimensional Brillouin zone common to all car- riers (see appendices A and B). The total torque then equals the sum of torques at the interface and in the bulk ferromagnet: =mag +FM : (30) As before we assume that the incoming polarizations approximately equal the accumulations at the interface. Thus we obtain mag  andFM in terms of  and ~E by plugging Eqs. (17) and (19) into Eqs. (25), (26), and (28). From the resulting expressions we may de ne the torkance and torkivity tensors introduced in Eq. (13). In appendix B we discuss this process, and in appendix C we present generalized expressions for non-trivial electronic structures. We note that the conductance and conductivity tensors describe the charge current and longitudinal spin current in the ferromagnet, but not the transverse spin currents. In the ferromagnet, the transverse spin currents dissipate not far from the interface, while the charge current and longitudinal spin current can propagate across the entire layer. Thus the transverse spin currents in the ferromag- net are best described as spin torques given by FM ; this explains why we include them in the torkance and torkiv- ity tensors instead of the conductance and conductivity tensors. If we derive a similar formalism to describe a non-magnetic bilayer, spin currents polarized in all di- rections should be included in the conductance and con- ductivity tensors. With no magnetism, no spin torques are exerted at or near the interface and the torkance and torkivity tensors are not meaningful. IV. NUMERICAL ANALYSIS OF BOUNDARY PARAMETERS In the following we numerically analyze the boundary parameters introduced in Eqs. (12) and (13) in the pres- ence of an interfacial exchange interaction and spin-orbit10 FIG. 4. (Color online) Contour plots of various boundary parameters versus the interfacial exchange ( uex) and Rashba ( uR) strengths. The magnetization points away from the electric eld 45oin-plane and 22 :5oout-of-plane. Note that the parameters plotted in panels (a)-(c) describe the scattering processes illustrated in Figs. 1(a)-(c). (a) Plot of Gzdf, which generalizes Im[G"#] in the presence of interfacial spin-orbit coupling. It describes a rotation of spin currents polarized transversely to the magnetization. (b) Plot of Gzll, which contributes to spin memory loss longitudinal to the magnetization. It varies mostly withuR, since interfacial spin-orbit coupling provides a sink for angular momentum. (c) Plot of FM d, which describes the out-of-plane, damping-like spin current created by an in-plane electric eld and spin-orbit scattering. It exceeds its eld-like counterpart ( FM f); thus, the resulting spin current exerts a (mostly) damping-like spin torque upon entering the ferromagnet. (d) An array of contour plots, with each plot shown over an identical range as those in (a)-(c). The plot in row and column corresponds to the parameter Gz . From this one may visualize the coupling between spin/charge indices for this tensor, shown across the parameter space of the scattering potential given by Eq. (31). The overall structure of Gz resembles that of magnetoelectronic circuit theory, given by Eq. (11). The corresponding gures for (e) z , (f) FM , and (g) mag are also shown. scattering. We do so to provide intuition as to the rel- ative strengths of each boundary parameter. We use a scattering potential localized at the interface [42] that is based on the Rashba model of spin orbit coupling V(r) =~2kF m(z) u0+uex^m+uR(^k^z) (31) whereu0represents a spin-independent barrier, uexgov- erns the interfacial exchange interaction, and uRdenotesthe Rashba interaction strength. Plane waves comprise the scattering wavefunctions in both regions. In Fig. 4 we plot various boundary parameters versus the exchange interaction strength ( uex) and the Rashba interaction strength ( uR). Figures 4(a)-(c) display indi- vidual boundary parameters, while Figs. 4(d)-(g) display multiple boundary parameters for a given tensor. The plots in Figs. 4(d)-(g) are arranged as arrays to help visu-11 alize the coupling between spin/charge components. The spin-orbit interaction misaligns the preferred direction of spins from the magnetization axis. Thus, no two tensor elements are identical, though many remain similar. As expected, the coupling between the transverse spin com- ponents and the charge and longitudinal spin components does not vanish. The conductance tensor Gz generalizes GMCT in the presence of interfacial spin-orbit coupling. Comparison to Eq. (11) suggests that the parameters GzddandGzdf represent the real and imaginary parts of a generalized mixing conductance ( ~G"#). Each element of the conduc- tance tensor experiences a similar perturbation due to spin-orbit coupling. However, the tensor elements from the 22 o -diagonal blocks in Fig. 4(d) either vanish or remain two orders of magnitude smaller than those from the diagonal blocks. This remains true even for values ofuRapproaching the spin-independent barrier strength u0. While these blocks are small for the simple model treated here, they may become important for particular realistic electronic structures. The fact that the elements Gzc andGz cvanish for all ensures the conserva- tion of charge current and guarantees no dependence on an o set to the charge accumulations. Note that four additional parameters vanish in the conductance tensor shown in Fig. 4(d); this occurs because identical scatter- ing wavefunctions were used for both sides of the inter- face when computing the scattering coecients. These parameters do not vanish in general. The results shown in Fig. 4 were computed for a mag- netization with out-of-plane components. In magneto- electronic circuit theory, the parameters are independent of the magnetization direction. With interfacial spin- orbit coupling, this is no longer the case. In general all of the parameters in Eqs. (12) and (13) depend on the magnetization direction. However, we nd that this de- pendence is weak for the model we consider here. For in- plane magnetizations (not shown) the 2 2 o -diagonal blocks vanish, but spin-orbit coupling still modi es the diagonal blocks in the manner described above. In the presence of interfacial spin-orbit coupling the lattice also receives angular momentum from carriers. This results in a loss of spin current across the inter- face, or spin memory loss, which the elements Gzl partly characterize. The computation of these parame- ters for realistic electronic structures should help predict spin memory loss in experimentally-relevant bilayers. In particular, spin memory loss might play a crucial role when measuring the spin Hall angle of heavy metals via spin-pumping from an adjacent ferromagnet [9]. Here Gzllprovides the strongest contribution to spin mem- ory loss that is caused by accumulations, and approaches the imaginary part of the generalized mixing conductance in magnitude. Until now, we have discussed the tensors that describe how accumulations a ect transport. However, in-plane electric elds and spin-orbit scattering create additional currents that form near the interface. In particular, theconductivity parameters i describe the currents that can propagate into either layer without signi cant de- phasing. For instance, the element zldescribes an out- of-plane longitudinal spin current driven by an in-plane electric eld. The element zlthen gives the disconti- nuity in this spin current across the interface. This dis- continuity arises because of coupling to the lattice, and thus contributes to spin memory loss. Likewise, the torkivity tensors describe contributions to the total spin torque that arise from in-plane electric elds and spin-orbit scattering. This includes the torques exerted by the spin polarization at z= 0 and by the transverse spin currents at z= 0+. The tensors mag  and FM describe these torques respectively. Since the transverse spin currents at z= 0+quickly dephase in the ferromagnet, we treat them as spin torques and do not include them in the conductivity tensor. To understand how the boundary parameters con- tribute to spin-orbit torques, we note that mag f> mag d over the swept parameter space. This implies that the torque exerted at z= 0 is primarily eld-like, which agrees with previous studies of interfacial Rashba spin or- bit torques [42]. However, we also nd that FM d> FM f for stronguR; in this case the resulting spin current exerts a damping-like torque by owing into the ferromagnet. Both spin torque contributions result from the interfacial Rashba interaction. This implies that interfacial spin- orbit scattering provides a crucial mechanism to the cre- ation of damping-like Rashba spin torques. In the com- panion paper we support this claim by comparing spin- orbit torques computed using both the drift-di usion and Boltzmann equations. V. OUTLOOK In the previous section we demonstrated that only cer- tain boundary parameters remain important when mod- eling spin orbit torques. The interfacial conductivity and torkivity parameters capture physics due to in-plane ex- ternal electric elds. They depend on the di erence in bulk conductivities, which are typically easier to estimate than interfacial spin/charge accumulations. For this rea- son, calculating the conductivity and torkivity tensors for a realistically-modeled system should provide direct insight into its spin transport behavior. In particular, we showed that conductivity and torkivity parameters strongly indicate the potential to produce damping-like and eld-like torques. Further studies may yield signi - cant insight into the underlying causes of these and other phenomena for speci c material systems. Even so, treat- ing the elements of these tensors as phenomenological parameters should bene t the analysis of a variety of ex- periments, which we discuss now. (1) Spin pumping/memory loss | Spin pumping ex- periments in Pt-based multilayers suggest that the mea- sured interfacial spin current di ers from the actual spin current in Pt, leading to inconsistent predictions of the12 spin Hall angle [9, 53]. Rojas-Sanchez et al. [9] ex- plain this discrepancy in terms of spin memory loss while Zhang et al. [53] attribute it to interface transparency. The latter characterizes the actual spin current generated at an interface when backscattering is accounted for; it depends on G"#and does not require interfacial spin- orbit coupling. Though further experimental evidence is needed to resolve these claims, the elements of Gz char- acterize both spin memory loss and transparency. Fig- ure 4 implies that transparency depends on interfacial spin-orbit coupling, while spin memory loss also depends on the interfacial exchange interaction. Thus, the gen- eralized boundary conditions introduced here unify these two interpretations and allow for further investigation us- ing a single theory. (2) Rashba-Edelstein e ect | Sanchez et al. [41] mea- sure the inverse Rashba-Edelstein e ect in an Ag/Bi in- terface, in which interfacial spin orbit coupling converts a pumped spin current into a charge current. The the- oretical methods that describe this phenomena to date [37, 44{46, 68] assume orthogonality between the direc- tional and spin components of the spin current tensor. However, the conductivity tensor introduced here is ro- bust in general; this implies that interfacial spin-orbit scattering converts charge currents into spin currents with polarization and ow directions not orthogonal to the charge current. Onsager reciprocity implies that spin currents should give rise to charge currents at the inter- face that ow in all directions as well. Thus, the con- ductivity tensor describes a generalization of the direct and inverse Rashba-Edelstein e ects as they pertain to interfaces with spin-orbit coupling. (3) Spin Hall magnetoresistance | The conductivity tensor also leads to in-plane charge currents. These cur- rents depend on magnetization direction via the scat- tering amplitudes, and thus suggest a new contribution to the spin Hall magnetoresistance based on the Rashba e ect in addition to that from the spin Hall e ect. Pre- liminary calculations of this mechanism suggest a mag- netoresistance in Pt/Co of a few percent, which is com- parable or greater than experimentally measured values in various systems [58{61]. We expect that the most useful approach for inter- preting experiments as above is to treat the new trans- port parameters as tting parameters. In the future, this approach can be checked by calculating the parameters from rst principles [74, 75] as has been done for magne- toelectronic circuit theory. This requires computing the boundary parameters for realistic systems using the ex- pressions given in appendix C. Such calculations would provide a useful bridge between direct rst-principles cal- culations of spin torques [76{79] and drift-di usion cal- culations done to analyze experiments. To conclude, we present a theory of spin transport at interfaces with spin-orbit coupling. The theory describes spin/charge transport in terms of resistive elements, which ultimately describe measurable consequences of interfacial spin-orbit scattering. In particular, the pro-posed conductivity and torkivity tensors model the phe- nomenology of in-plane electric elds in the presence of interfacial spin-orbit coupling, which was previously inac- cessible to the drift-di usion equations. We calculate all parameters in a simple model, but also provide general expressions in the case of realistic electronic structure. We found that elements of the conductivity and torkiv- ity tensors are more important than the modi cations of other transport parameters (such as the mixing conduc- tance) in many experimentally-relevant phenomena, such as spin orbit torque, spin pumping, the Rashba-Edelstein e ect, and the spin Hall magnetoresistance. ACKNOWLEDGMENTS The authors thank Kyoung-Whan Kim, Paul Haney, Guru Khalsa, Kyung-Jin Lee, and Hyun-Woo Lee for useful conversations and Robert McMichael and Thomas Silva for critical readings of the manuscript. VA ac- knowledges support under the Cooperative Research Agreement between the University of Maryland and the National Institute of Standards and Technology, Cen- ter for Nanoscale Science and Technology, Grant No. 70NANB10H193, through the University of Maryland. Appendix A: Derivation of the conductance and conductivity tensors To derive the conductance and conductivity tensors we must approximate the distribution function f (k) at the interface. The distribution function gives the momentum-dependent occupancy of carriers described by the spin/charge index . Just out of equilibrium, it is perturbed by the linearized non-equilibrium distribution functiong (k), as seen in Eq. (16). In the following we complete the derivation begun in Sec. III. We write the portion of g (kjj) incident on the inter- face as done in Eq. (17). The rst term on the right hand side of Eq. (17) captures the spin and charge cur- rents incident on the interface, while the second term gives an anisotropic contribution caused by an external electric eld. As discussed in Sec. III, the simplest ap- proximation for g (kjj) is to use the particular solution of the Boltzmann equation in the relaxation time approx- imation, given by Eq. (18). The momentum relaxation times that we use account for di ering majority ( ") and minority (#) relaxation times in the ferromagnet, and are renormalized by bulk spin- ip scattering in the non- magnet: ()1= (mf)1+ (sf)1: (A1) We may better approximate Eq. (18) by forcing the dis- tribution function to obey outer boundary conditions as well. In the companion paper we present a more sophisti- cated approximation for Eq. (18) that accomplishes this13 by using solutions to the homogeneous Boltzmann equa- tion. The outgoing distribution, given by Eq. (19), is speci- ed by the incoming distribution and the scattering co- ecientsS . The scattering coecients are given by Eq. (20) and Eq. (21). Here we compute non-equilibrium accumulations analogously to the currents de ned by Eq. (23),  =1 e1 AFS Z FS d2kg (k); (A2) where denotes the accumulation. Furthermore, AFS gives the Fermi surface area while vF gives the Fermi velocity. The quantities just de ned apply to the popula- tion with spin/charge index . Likewise, all integrals are evaluated over the Fermi surface that corresponds to the spin/charge index . Note that we express the accumula- tions in units of voltage and the current densities in units of number current density. Using Eqs. (19) and (20) we may rewrite these expressions as integrals over the max- imal two-dimensional Brillouin zone common (2DBZ) to all carriers  =c eX Z 2DBZdkjj1 vz (kjj)  +S  gin (kjj) (A3) ji =cj eX Z 2DBZdkjjvi (kjj) vz (kjj)  iz+S  gin (kjj); (A4) where cvF AFS ; cje ~(2)3: (A5) Note that the velocities correspond to outgoing carriers. The factor iz(12iz) accounts for the fact that incoming and outgoing currents have the opposite sign fori=zbut the same sign for i2[x;y]. By integrat- ing over the maximal two-dimensional Brillouin zone we encounter evanescent states, since kjjvectors not corre- sponding to real Fermi surfaces have imaginary kzvalues. Here we neglect the contributions to the currents and ac- cumulations due to evanescent states. Such contributions vanish very close to the interface. We must now express the accumulations and currents in terms of the incoming polarizations and the in-plane electric eld. Plugging Eqs. (17) and (19) into Eqs. (A3) and (A4), we obtain the following  =A q +a ~E (A6) ji =Bi q +bi ~E (A7) where the tensors that contract with the incident spin/charge polarization are given by A =cZ 2DBZdkjj1 vz  +S  (A8) Bi =cjZ 2DBZdkjjvi vz  iz+S  (A9)while the tensors that multiply the in-plane electric eld become a =cX Z 2DBZdkjj1 vz  +S  fE (A10) bi =cjX Z 2DBZdkjjvi vz  iz+S  fE (A11) In the same spirit as magnetoelectronic circuit theory, these tensors represent moments of the scattering coe- cients weighted by velocities. To determine exactly how the currents depend on the accumulations, we solve for ji in terms of  . Doing so yields the following conductance and conductivity tensors Gi =Bi [A1] i =bi Gi a : To further simplify these expressions, we follow the ex- ample of magnetoelectronic circuit theory [71, 72] and assume that the incoming spin-currents behave as if they originate from spin-dependent reservoirs. This implies that the incoming spin polarization q equals the qua- sichemical potential  at the interface. For this to be true, we must nd that A / anda = 0 by in- spection of Eq. (A6). These relationships hold if one evaluates Eqs. (A8) and (A10) over the incoming portion of the Fermi surface only. We nd that the contributions from the outgoing portion of the Fermi surface cancel to a good approximation, which suggests that: Gi =Bi (A12) i =bi : (A13) The above equations give simpler expressions for the con- ductance and conductivity tensors in terms of interfacial scattering coecients. Appendix B: Derivation of the torkance and torkivity tensors To describe the spin torque at z= 0, we must compute the ensemble average of spin density hsiusing Eq. (25). The resulting torque is given by Eq. (26). To describe the spin torque at z= 0+, we must calculate the transverse spin current in the ferromagnet using Eqs. (28) and (29). We then express the spin torque in terms of the incoming polarizations and the in-plane electric eld by plugging Eqs. (17) and (19) into Eqs. (25), (26), and (28). In doing so we obtain =C q +c~E; (B1) where C =CFM  +Cmag  (B2) c=cFM +cmag  (B3)14 describes the separation of the spin torque into its bulk ferromagnet and interface contributions. The tensors that contract with the incident spin/charge polarization are given by CFM  =cjZ 2DBZdkjjvz vz S ; (B4) Cmag  =Jex ~cjX 0Z 2DBZdkjj1 vz 0T0 ; (B5) while the tensors that multiply the in-plane electric eld become cFM =cjX Z 2DBZdkjjvz vz S fE : (B6) cmag =Jex ~cjX Z 2DBZdkjj1 vz 0T0 fE (B7) where the velocity  vz(kjj) corresponds to the outgoing portion of the Fermi surface in the ferromagnet. As we did for the currents, we solve for in terms of  . Doing so yields the following torkance and torkivity tensors  =C [A1] =c a : The torkance tensor describes the contribution to the to- tal spin torque that arises from the accumulations at the interface. The torkivity tensor describes the subsequent contribution from interfacial spin-orbit scattering when driven by an in-plane electric eld. Following the argu- ments made for Eq. (A12):  =C (B8) =c: (B9) As seen in the companion paper, this approximation pro- duces good agreement with the interfacial charge cur- rents, spin currents, and spin torques computed via the Boltzmann equation. Appendix C: Boundary Parameters for Realistic Interfaces To generalize the expressions from the previous section to include electronic structure, we must consider the non- equilibrium distribution function for all bands relevant to transport: fm (k) = feq m ("m k) +@feq m @"m kgm (k): (C1) Heremdescribes the spin-independent band number and denotes the spin/charge index. If the case of a non- magnet, for each spin-independent band there are two degenerate states. Linear combinations of these statescan produce phase coherent spin states that point in any direction. Thus, for the non-magnet, the spin/charge in- dex should span 2[d;f;`;c ], where the `direction is aligned with the magnetization in the neighboring ferro- magnet for convenience. In the ferromagnet all bands are non-degenerate, so each state possesses a di erent phase velocity. As a result, linear combinations of these states have spin expectation values that oscillate over position, complicating the description presented above. There is no natural pairing of non-degenerate spin states. How- ever, if states are quantized along a particular axis, the spin accumulations and spin currents with polarization along that axis are well-de ned regardless of the choice of pairing. Thus for each spin-independent band in the ferromagnet, the spin/charge index spans the states de- scribing majority and minority carriers, i.e. 2[";#]. We generalize the approximate distribution function fE m (kjj) caused by an external electric eld to allow for a band dependence. We do so because the velocities now depend on band number and the scattering times may as well. However, we assume that the incoming polarization q does not depend on band number; thus we treat in- cident currents as if they originate from spin-dependent (but not band-dependent) reservoirs. The momentum relaxation times for each spin-independent band in the non-magnet are renormalized using Eq. (A1). To account for coherence between bands, we begin with a more general expression for the ensemble average of the outgoing current: hhjout i ii=1 ~X mnn0 Z 2DBZdkjjgin m jvmz j tr (sn0m)yJout n0n;i snm  : (C2) Heresstands for re ection or transmission, depending on what region(s) incoming and outgoing carriers are from. The indices mand correspond to incoming carriers, whilen,n0, and describe the outgoing carriers. The current operator Jout n0n;i is given by Jout n0n;i =i~ 2mZ 2DPCdrjj( n0k)y( @i  !@i) nk (C3) where the integral runs over a two-dimensional slice of the primitive cell (aligned parallel to the interface). The 22 matrix nkis de ned for outgoing modes in the ferromagnet as nk=eikjjrjj u" nk(r)eik" nzz0 0u# nk(r)eik# nzz! ;(C4) whereu"=# nk(r) andk"=# nzdenote the Bloch wavefunction and out-of-plane wavevector for majority/minority car- riers. Both are de ned at kjjon the Fermi surface cor-15 responding to band n. For outgoing modes in the non- magnet, nksimpli es to: nk=eikjjrjjeiknzzunk(r)I22: (C5) The incoming current is de ned as follows hhjin i ii=1 ~X m Z 2DBZdkjjgin m jvmz jtr Jin m;i   (C6) where Jin m;i =izi~ 2mZ 2DPCdrjj( mk)y( @i  !@i) mk (C7) gives the current operator for the incoming current. The total current is then hhji ii=hhjin i ii+hhjout i ii =1 ~X mnn0 Z 2DBZdkjjgin m jvmz j trh izJin m;i + (sn0m)yJout n0n;i snm  i (C8) where the choice of scattering matrix depends on the in- coming spin/charge index and outgoing spin/charge index as follows: snm=8 >>>>>< >>>>>:rnm ; 2[d;f;`;c ] tnm 2[d;f;`;c ]; 2[";#] t nm 2[";#]; 2[d;f;`;c ] r nm ; 2[";#] (C9) By plugging in the generalizations of Eqs. (17) and (18) into Eq. (C8), we nd that Eq. (A9) generalizes to the following Bi =e ~X mnn0Z 2DBZdkjj1 jvmz j trh izJin m;i + (sn0m)yJout n0n;i snm  i ; (C10) while Eq. (A11) now becomes: bi =e ~X mnn0 Z 2DBZdkjjfE m jvmz j trh izJin m;i + (sn0m)yJout n0n;i snm  i : (C11)Assuming as before that the incoming spin-currents be- have as if they originate from spin-dependent reservoirs ( q ), we have: Gi =Bi (C12) i =bi : (C13) Thus, Eqs. (C10) and (C11) generalize the conductance and conductivity tensors respectively to include non- trivial electronic structure. The transverse spin current that develops in the ferro- magnet atz= 0+may be obtained by using similar ex- pressions. The tensor CFM  , originally given by Eq. (B4), now becomes CFM  =e ~X mnn0Z 2DBZdkjj1 jvmz j 8 < :tr (t n0m)yJout n0n;zt nm  2[d;f;`;c ] tr (r n0m)yJout n0n;zr nm  2[";#]: (C14) Likewise, the tensor cFM , rst described by Eq. (B6), generalizes to the following: cFM =e ~X mnn0 Z 2DBZdkjjfE m jvmz j 8 < :tr (t n0m)yJout n0n;zt nm  2[d;f;`;c ] tr (r n0m)yJout n0n;zr nm  2[";#]: (C15) Evaluating the trace in Eq. (C15) gives the ensemble av- erage of velocity for the transverse spin states in the fer- romagnet. Here we do not assume that the velocity of these states equals the average velocity of majority and minority carriers. However, for the simple model dis- cussed in the previous section, one can show that the current operator Jout n0n;zsimpli es to the following: Jout n0n;z!Jout z/1 2 vz"+vz#  (C16) In this scenario, Eqs. (C14) and (C15) reduce to Eqs. (B4) and (B6) as expected. This justi es the use of the average velocity to describe transverse spin states in the simple model. For  q we have: FM  =CFM  (C17) FM =cFM : (C18) Thus we have generalized the torkance and torkivity tensors that describe bulk ferromagnet torques for non- trivial electronic structures. For realistic systems, the interface should be modeled over a few atomic layers so that an exchange potential16 Parameter Value E ective mixing conductance Re[~G"#]GzddorGzff Im[~G"#]GzdforGzfd Spin current due to interfacial spin-orbit scattering jE d(0)zd~E jE f(0)zf~E Spin torque on the lattice at the interface latt d zd d~E latt f zf f~E TABLE I. Table of phenomenological parameters relevant to the drift-di usion model of spin-orbit torque developed in the companion paper, chosen by the numerical study performed in Sec.IV. All other boundary parameters are discarded in that model. As can be seen in section IIIA of the companion paper, the rst four parameters govern the total spin torque thickness dependence, while the last two parameters describe the spin torque's zero-thickness intercept. Note that here all boundary parameters obey the sign convention that positive currents ow towards from the ferromagnet. and spin-orbit coupling may simultaneously exist. If these atomic layers make up the scattering region used to obtain the scattering coecients, then the expressions presented here describe the currents on either side of the interface as intended. However, in order to describe the interfacial torque, the tensors Cmag  andcmag must be written as sums of the layer-resolved torques within the interfacial scattering region. We save the generalization of Eqs. (B5) and (B7) for future work, since in this paper we treat the interface as a plane rather than a region of nite thickness. Appendix D: Boundary parameters relevant to bilayer spin-orbit torques In Sec. IV, we numerically analyze each boundary pa- rameter for an interfacial scattering potential that in- cludes the exchange interaction and spin-orbit coupling. We nd that many parameters di er by several orders of magnitude. In the companion paper, we use this in- formation to derive an analytical drift-di usion model of spin-orbit torques in heavy metal/ferromagnet bilayers. In the following we discuss the minimal set of parameterscrucial to that solution. Table I includes six parameters important to the inter- face of heavy metal/ferromagnet bilayers. Along with the spin di usion length ( lsf), the bulk conductivity ( NM bulk), and the spin Hall current density ( jsH d) in the non- magnet, they describe all of the phenomenological pa- rameters used by the analytical drift-di usion model in the companion paper. The rst two parameters are the real and imaginary parts of the spin mixing conductance. The generalized version of these parameters may be ex- tracted from the conductance tensor Gz . Numerical studies show that these parameters depend weakly on magnetization direction. In the companion paper, the ungeneralized spin mixing conductance is used. The pa- rametersjE d(0) andjE f(0) denote the interfacial spin currents just within the non-magnet that arise due to in-plane electric elds and spin-orbit scattering. In anal- ogy to the bulk spin Hall current, these parameters act as sources of spin current for the drift-di usion equa- tions. Thus, in the absence of jE d(0),jE f(0), andjsH d, all bulk currents and accumulations vanish. In addition to the spin mixing conductance, these parameters deter- mine the non-magnet thickness-dependence of spin-orbit torques. The nal two parameters give the approximate loss of angular momentum to the interface. They equal the damping-like and eld-like spin-orbit torques in the limit of vanishing non-magnet thickness. They are de- rived by subtracting the interfacial torque from the loss in out-of-plane spin current density across the interface. Our numerical analysis suggests that spin and charge accumulations cause negligible di erences in these two quantities. Thus, we assume that latt dandlatt fstem primarily from spin-orbit scattering at the interface. The treatment of the lattice torque presented in the compan- ion paper begins from this assumption. The model introduced in the companion paper general- izes the drift-di usion model used in Ref. [42] to include interfacial spin-orbit e ects. Only two additional phe- nomenological parameters ( jE d(0) andjE f(0)) are re- quired to capture the non-magnet thickness dependence, while an additional two parameters ( latt dandlatt f) de- scribe the corresponding zero-thickness intercept. Table I provides formulas for these phenomenological parameters in terms of the boundary parameters contained within Eqs. (12) and (13). We note that in magnetoelectronic circuit theory, the conductance parameters are given by sums of interfacial scattering coecients over the avail- able scattering states. All of the boundary parameters introduced here possess a similar form, as discussed in appendices A, B, and C. [1] T. Jungwirth, J. Wunderlich, and K. Olejnk, Nature Materials 11, 382 (2012). [2] E. M. Hankiewicz and G. Vignale, Journal of Physics: Condensed Matter 21, 253202 (2009). [3] A. Manchon and S. Zhang, Phys. Rev. B 78, 212405(2008). [4] A. Manchon and S. Zhang, Phys. Rev. B 79, 094422 (2009). [5] A. Matos-Abiague and R. L. Rodringuez-Su arez, Phys. Rev. B 80, 094424 (2009).17 [6] K. Ando, S. Takahashi, K. Harii, K. Sasage, J. Ieda, S. Maekawa, and E. Saitoh, Phys. Rev. Lett. 101, 036601 (2008). [7] I. M. Miron, K. Garello, G. Gaudin, P.-J. Zermatten, M. V. Costache, S. Au ret, S. Bandiera, B. Rodmacq, A. Schuhl, and P. Gambardella, Nature 476, 189 (2011). [8] Y. Tserkovnyak, A. Brataas, and G. E. W. Bauer, Phys. Rev. Lett. 88, 117601 (2002). [9] J.-C. Rojas-S anchez, N. Reyren, P. Laczkowski, W. Savero, J.-P. Attan e, C. Deranlot, M. Jamet, J.-M. George, L. Vila, and H. Ja r es, Phys. Rev. Lett. 112, 106602 (2014). [10] W. Thomson, Proceedings of the Royal Society of London 8, 546 (1856). [11] T. McGuire and R. Potter, IEEE Transactions on Mag- netics 11, 1018 (1975). [12] C. Gould, C. Ruster, T. Jungwirth, E. Girgis, G. M. Schott, R. Giraud, K. Brunner, G. Schmidt, and L. W. Molenkamp, Phys. Rev. Lett. 93, 117203 (2004). [13] A. N. Chantis, K. D. Belashchenko, E. Y. Tsymbal, and M. van Schilfgaarde, Phys. Rev. Lett. 98, 046601 (2007). [14] N. Nagaosa, J. Sinova, S. Onoda, A. H. MacDonald, and N. P. Ong, Rev. Mod. Phys. 82, 1539 (2010). [15] G. E. W. Bauer, E. Saitoh, and B. J. van Wees, Nature Materials 11, 391 (2012). [16] V. P. Amin, J. Zemen, J. Zelezn y, T. Jungwirth, and J. Sinova, Phys. Rev. B 90, 140406 (2014). [17] C. L opez-Monins, A. Matos-Abiague, and J. Fabian, Phys. Rev. B 90, 174426 (2014). [18] L. N eel, J. Physique Rad. 15, 225 (1954). [19] G. H. O. Daalderop, P. J. Kelly, and M. F. H. Schuur- mans, Phys. Rev. B 41, 11919 (1990). [20] M. T. Johnson, P. J. H. Bloemen, F. J. A. den Broeder, and J. J. de Vries, Reports on Progress in Physics 59, 1409 (1996). [21] K. Garello, I. M. Miron, C. O. Avci, F. Freimuth, Y. Mokrousov, S. Blugel, S. Au ret, O. Boulle, G. Gaudin, and P. Gambardella, Nature Nanotechnology 8, 587593 (2013). [22] X. Fan, H. Celik, J. Wu, C. Ni, K.-J. Lee, V. O. Lorenz, and J. Q. Xiao, Nature Communications 5, (2014). [23] G. Allen, S. Manipatruni, D. E. Nikonov, M. Doczy, and I. A. Young, Phys. Rev. B 91, 144412 (2015). [24] S. Emori, T. Nan, A. M. Belkessam, X. Wang, A. D. Matyushov, C. J. Babroski, Y. Gao, H. Lin, and N. X. Sun, Phys. Rev. B 93, 180402 (2016). [25] M. I. D'yakonov and V. I. Perel, 13, 467 (1971). [26] J. E. Hirsch, Phys. Rev. Lett. 83, 1834 (1999). [27] S. Zhang, Phys. Rev. Lett. 85, 393 (2000). [28] S. Murakami, N. Nagaosa, and S.-C. Zhang, Science 301, 1348 (2003). [29] J. Sinova, D. Culcer, Q. Niu, N. A. Sinitsyn, T. Jung- wirth, and A. H. MacDonald, Phys. Rev. Lett. 92, 126603 (2004). [30] Y. K. Kato, R. C. Myers, A. C. Gossard, and D. D. Awschalom, Science 306, 1910 (2004). [31] J. Wunderlich, B. Kaestner, J. Sinova, and T. Jungwirth, Phys. Rev. Lett. 94, 047204 (2005). [32] J. C. Slonczewski, Phys. Rev. B 39, 6995 (1989). [33] J. Slonczewski, Journal of Magnetism and Magnetic Ma- terials 159, L1 (1996). [34] L. Berger, Phys. Rev. B 54, 9353 (1996). [35] D. Ralph and M. Stiles, Journal of Magnetism and Mag- netic Materials 320, 1190 (2008).[36] M. D. Stiles and A. Zangwill, Phys. Rev. B 66, 014407 (2002). [37] V. Edelstein, Solid State Communications 73, 233 (1990). [38] S. D. Ganichev, E. L. Ivchenko, V. V. Bel'kov, S. A. Tarasenko, M. Sollinger, D. Weiss, W. Wegscheider, and W. Prettl, Nature 417, 153156 (2002). [39] Y. K. Kato, R. C. Myers, A. C. Gossard, and D. D. Awschalom, Phys. Rev. Lett. 93, 176601 (2004). [40] A. Y. Silov, P. A. Blajnov, J. H. Wolter, R. Hey, K. H. Ploog, and N. S. Averkiev, Applied Physics Letters 85, 5929 (2004). [41] J. C. R. S anchez, L. Vila, G. Desfonds, S. Gambarelli, J. P. Attan e, J. M. D. Teresa, C. Mag en, and A. Fert, Nature Communications 4, (2013). [42] P. M. Haney, H.-W. Lee, K.-J. Lee, A. Manchon, and M. D. Stiles, Phys. Rev. B 87, 174411 (2013). [43] C. Gorini, P. Schwab, R. Raimondi, and A. L. Shelankov, Phys. Rev. B 82, 195316 (2010). [44] C. Gorini, R. Raimondi, and P. Schwab, Phys. Rev. Lett. 109, 246604 (2012). [45] R. Raimondi, P. Schwab, C. Gorini, and G. Vignale, Annalen der Physik 524, n/a (2012). [46] K. Shen, G. Vignale, and R. Raimondi, Phys. Rev. Lett. 112, 096601 (2014). [47] O. Mosendz, J. E. Pearson, F. Y. Fradin, G. E. W. Bauer, S. D. Bader, and A. Ho mann, Phys. Rev. Lett. 104, 046601 (2010). [48] H. Nakayama, K. Ando, K. Harii, T. Yoshino, R. Taka- hashi, Y. Kajiwara, K. Uchida, Y. Fujikawa, and E. Saitoh, Phys. Rev. B 85, 144408 (2012). [49] Z. Feng, J. Hu, L. Sun, B. You, D. Wu, J. Du, W. Zhang, A. Hu, Y. Yang, D. M. Tang, B. S. Zhang, and H. F. Ding, Phys. Rev. B 85, 214423 (2012). [50] V. Vlaminck, J. E. Pearson, S. D. Bader, and A. Ho - mann, Phys. Rev. B 88, 064414 (2013). [51] W. Zhang, V. Vlaminck, J. E. Pearson, R. Divan, S. D. Bader, and A. Ho mann, Applied Physics Letters 103, 242414 (2013). [52] M. Obstbaum, M. H artinger, H. G. Bauer, T. Meier, F. Swientek, C. H. Back, and G. Woltersdorf, Phys. Rev. B89, 060407 (2014). [53] W. Zhang, W. Han, X. Jiang, S.-H. Yang, and S. S. P. Parkin, Nature Physics 11, 496 (2015). [54] Y. Liu, Z. Yuan, R. J. H. Wesselink, A. A. Starikov, and P. J. Kelly, Phys. Rev. Lett. 113, 207202 (2014). [55] Y.-T. Chen, S. Takahashi, H. Nakayama, M. Althammer, S. T. B. Goennenwein, E. Saitoh, and G. E. W. Bauer, Phys. Rev. B 87, 144411 (2013). [56] M. Weiler, M. Althammer, F. D. Czeschka, H. Huebl, M. S. Wagner, M. Opel, I.-M. Imort, G. Reiss, A. Thomas, R. Gross, and S. T. B. Goennenwein, Phys. Rev. Lett. 108, 106602 (2012). [57] S. Y. Huang, X. Fan, D. Qu, Y. P. Chen, W. G. Wang, J. Wu, T. Y. Chen, J. Q. Xiao, and C. L. Chien, Phys. Rev. Lett. 109, 107204 (2012). [58] H. Nakayama, M. Althammer, Y.-T. Chen, K. Uchida, Y. Kajiwara, D. Kikuchi, T. Ohtani, S. Gepr ags, M. Opel, S. Takahashi, R. Gross, G. E. W. Bauer, S. T. B. Goennenwein, and E. Saitoh, Phys. Rev. Lett. 110, 206601 (2013). [59] C. Hahn, G. de Loubens, O. Klein, M. Viret, V. V. Nale- tov, and J. Ben Youssef, Phys. Rev. B 87, 174417 (2013). [60] N. Vlietstra, J. Shan, V. Castel, B. J. van Wees, and18 J. Ben Youssef, Phys. Rev. B 87, 184421 (2013). [61] M. Althammer, S. Meyer, H. Nakayama, M. Schreier, S. Altmannshofer, M. Weiler, H. Huebl, S. Gepr ags, M. Opel, R. Gross, D. Meier, C. Klewe, T. Kuschel, J.-M. Schmalhorst, G. Reiss, L. Shen, A. Gupta, Y.-T. Chen, G. E. W. Bauer, E. Saitoh, and S. T. B. Goennenwein, Phys. Rev. B 87, 224401 (2013). [62] K. Chen and S. Zhang, Phys. Rev. Lett. 114, 126602 (2015). [63] K. Eid, D. Portner, J. A. Borchers, R. Loloee, M. Al- Haj Darwish, M. Tsoi, R. D. Slater, K. V. O'Donovan, H. Kurt, W. P. Pratt, and J. Bass, Phys. Rev. B 65, 054424 (2002). [64] H. Kurt, R. Loloee, K. Eid, W. P. Pratt, and J. Bass, Applied Physics Letters 81, 4787 (2002). [65] H. Nguyen, W. P. Jr., and J. Bass, Journal of Magnetism and Magnetic Materials 361, 30 (2014). [66] J. Bass and W. P. P. Jr, Journal of Physics: Condensed Matter 19, 183201 (2007). [67] L. Wang, R. J. H. Wesselink, Y. Liu, Z. Yuan, K. Xia, and P. J. Kelly, Phys. Rev. Lett. 116, 196602 (2016). [68] J. Borge, C. Gorini, G. Vignale, and R. Raimondi, Phys. Rev. B 89, 245443 (2014). [69] S. S.-L. Zhang, G. Vignale, and S. Zhang, Phys. Rev. B 92, 024412 (2015).[70] To convert these quantities back into their traditional units, one must multiply the components describing charge current densities by eand those describing spin current densities by ~=2. [71] A. Brataas, Y. V. Nazarov, and G. E. W. Bauer, Phys. Rev. Lett. 84, 2481 (2000). [72] A. Brataas, Y. Nazarov, and G. Bauer, The European Physical Journal B - Condensed Matter and Complex Systems 22, 99 (2001). [73] Note, however, that an average charge accumulation amounts to an o set in electric potential; thus it can- not a ect the interfacial spin currents. [74] M. D. Stiles and D. R. Penn, Phys. Rev. B 61, 3200 (2000). [75] K. Xia, P. Kelly, G. Bauer, A. Brataas, and I. Turek, Physical Review B: Condensed matter and materials physics 65, 220401 (2002). [76] P. M. Haney, H.-W. Lee, K.-J. Lee, A. Manchon, and M. D. Stiles, Phys. Rev. B 88, 214417 (2013). [77] F. Freimuth, S. Bl ugel, and Y. Mokrousov, Phys. Rev. B90, 174423 (2014). [78] F. Freimuth, S. Bl ugel, and Y. Mokrousov, Phys. Rev. B92, 064415 (2015). [79] G. G eranton, F. Freimuth, S. Bl ugel, and Y. Mokrousov, Phys. Rev. B 91, 014417 (2015).
1002.0441v2.Spin_resolved_scattering_through_spin_orbit_nanostructures_in_graphene.pdf
arXiv:1002.0441v2 [cond-mat.mes-hall] 26 Apr 2010Spin-resolved scattering through spin-orbit nanostructu res in graphene D. Bercioux1,2,∗and A. De Martino3,† 1Freiburg Institute for Advanced Studies, Albert-Ludwigs- Universit¨ at, D-79104 Freiburg, Germany 2Physikalisches Institut, Albert-Ludwigs-Universit¨ at, D-79104 Freiburg, Germany 3Institut f¨ ur Theoretische Physik, Universit¨ at zu K¨ oln, Z¨ ulpicher Straße 77, D-50937 K¨ oln, Germany (Dated: November 10, 2018) We address the problem of spin-resolved scattering through spin-orbit nanostructures in graphene, i.e., regions of inhomogeneous spin-orbit coupling on the nanom eter scale. We discuss the phe- nomenon of spin-double refraction and its consequences on t he spin polarization. Specifically, we study the transmission properties of a single and a double in terface between a normal region and a region with finite spin-orbit coupling, and analyze the pol arization properties of these systems. Moreover, for the case of a single interface, we determine th e spectrum of edge states localized at the boundary between the two regions and study their propert ies. PACS numbers: 72.80.Vp, 73.23.Ad, 72.25.-b, 72.25.Mk, 71. 70.Ej I. INTRODUCTION Graphene1,2— a singlelayerofcarbonatoms arranged in a honeycomb lattice — has attracted huge atten- tion in the physics community because of many unusual electronic, thermal and nanomechanical properties.3,4In graphene the Fermi surface, at the charge neutrality point, reduces to two isolated points, the two inequiv- alent corners KandK′of the hexagonal Brillouin zone of the honeycomb lattice. In their vicinity the charge carriers form a gas of chiral massless quasiparticles with a characteristic conical spectrum. The low-energy dy- namics is governed by the Dirac-Weyl (DW) equation5,6 in which the role of speed of light is played by the elec- tron Fermi velocity. The chiral nature of the quasipar- ticles and their linear spectrum lead to remarkable con- sequences for a variety of electronic properties as weak localization, shot noise, Andreev reflection, and many others. Also the behavior in a perpendicular magnetic field discloses new physics. Graphene exhibits a zero- energy Landau level, whose existence gives rise to an un- conventional half-integer quantum Hall effect, one of the peculiar hallmarks of the DW physics. Driven by the prospects of using this material in spin- tronic applications,7,8the study of spin transport is one of the most active field in graphene research.9–14Sev- eral experiments have recently demonstrated spin injec- tion, spin-valve effect, and spin-coherent transport in graphene, with spin relaxation length of the order of few micrometers.10,14In this context a crucial role is playedbythe spin-orbitinteraction. Ingraphenesymme- tries allow for two kinds of spin-orbit coupling (SOC).15 Theintrinsic SOC originates from carbon intra-atomic SOC. It opens a gap in the energy spectrum and con- verts graphene into a topological insulator with a quan- tized spin-Hall effect.15This term has been estimated to be rather weak in clean flat graphene.16–19Theex- trinsicRashba-like SOC originates instead from inter- actions with the substrate, presence of a perpendic- ular external electric field, or curvature of graphene membrane.16–18,20This term is believed to be responsiblefor spin polarization21and spin relaxation22,23physics in graphene. Optical-conductivitymeasurementscouldpro- vide a way to determine the respective strength of both SOCs.24 In this article we address the problem of ballistic spin- dependent scattering in the presence of inhomogeneous spin-orbit couplings. Our main motivation stems from a recent experiment that reported a large enhancement of Rashba SOC splitting in single-layer graphene grown on Ni(111) intercalated with a Au monolayer.25Further experimental results show that the intercalation of Au atoms between graphene and the Ni substrate is essen- tial in order to observe sizable Rashba effect.26,27The preparation technique of Ref. 25seems to provide a sys- tem with properties very close to those of freestanding graphene in spite of the fact that graphene is grown on a solid substrate. The presence of the substrate does not seem to fundamentally alter the electronic properties ob- served in suspended systems, i.e., the existence of Dirac points at the Fermi energy and the gapless conical dis- persion in their vicinity. These results suggest that a certain degree of control on the SOC can be achieved by appropriate substrate engineering, with variations of the SOC strength on sub- micrometer scales, without spoiling the relativistic gap- less nature of quasiparticles. This could pave the way for the realization of spin-optics devices for spin filtration and spin control for DW fermions in graphene. An opti- mal design would require a detailed understanding of the spin-resolved ballistic scattering through such spin-orbit nanostructures , which is the aim of this paper. The problem of spin transport through nanostructures with inhomogeneous SOC has already been thoroughly studied in the case of two-dimensional electron gas in semiconductor heterostructures with Rashba SOC.28–30 Here the Rashba SOC31— arising from the inversion asymmetry of the confinement potential — couples the electron momentum to the spin degree of freedom and thereby lifts the spin degeneracy. In this case, a region with finite SOC between two normal regions has prop- erties similar to biaxial crystals: an electron wave inci-2 N region SO region ky kxky kxφE k+k- ξ+ξ- Figure 1: (Color online) Illustration of the kinematics of t he scattering at a N-SO interface in graphene. The circles rep- resent constant energy contours. dent from the normal region splits at the interface and the two resulting waves propagate in the SO region with different Fermi velocities and momenta.28This effect — analogousto the opticaldouble-refraction— producesan interference pattern when the electron waves emerge in the second normal region. Moreover, electrons that are injected in an spin unpolarized state emerge from the SO region in a partially polarized state. Herewe shallfocus on the twosimplest examplesofSO nanostructuresingraphene: (i)asingleinterfacebetween two regions with different strengths of SOC; (ii) a SOC barrier, or double interface, i.e., a region of finite SOC in between two regions with vanishing SOC. Our analysis shows — in analogy to the case of 2DEG — that the ballistic propagation of carriers is governed bythespin-doublerefraction. Wefind thatthescattering properties of the structure strongly depend on the injec- tion angle. As a consequence, an initially unpolarized DW quasiparticle emerges from the SOC barrier with a finite spin polarization. In analogy to the edge states in the quantum spin-Hall effect,15we also consider the pos- sibility of edge states localized at the interface between regions with and without SOC. This paper is organized as follows. In Sec. IIwe intro- duce the model and the transfer matrix formalism used in the rest of the paper. In Sec. IIIwe discuss the scat- tering problem at a single interface and the spectrum of edge states. In Sec. IVwe address the case of a dou- ble interface — a SOC barrier — and the final Sec. Vis devoted to the discussion of results and conclusions. II. MODEL AND FORMALISM We consider a clean graphene sheet in the xy- plane with SOCs15,16,21,32,33inhomogeneous along the x-direction. We shall restrict ourselves to a single- particle picture and neglect electron-electron interaction effects. The length scale over which the SOCs vary is as- sumed to be much largerthan graphene’slattice constant (a= 0.246 nm) but much smaller than the typical Fermiwavelength of quasiparticles λF. Since close to the Dirac pointsλF∼1/|E|, at low energy Ethis approximation is justified. This assumption ensures that we can use the continuum DW description, in which the two valleys are not coupled. Yet close to a Dirac point we can approxi- mate the variation of SOCs as a sharp change. Focusing on a single valley, the single-particle Hamiltonian reads H=vFσ·p+HSO, (1) HSO=λ(x) 2(σ×s)z+∆(x)σzsz, (2) wherevF≈106m/s is the Fermi velocity in graphene. In the following we set /planckover2pi1=vF= 1. The vector of Pauli matrices σ= (σx,σy) [resp.s= (sx,sy)] acts in sublat- tice space [resp. spin space]. The term HSOcontains the extrinsic or Rashba SOC of strength λand the intrinsic SOC of strength ∆. While experimentally the Rashba SOC can be enhanced by appropriate optimization of the substrate up to values of the order of 14 meV,25the intrinsic SOC seems at least two orders of magnitude smaller. Yet, the limit of large intrinsic SOC is of con- siderable interest since in this regime graphene becomes a topological insulator.15Thus in this paper we shall not restrict ourself to the experimentally relevant regime λ≫∆ but consider also the complementary regime. The wave function Ψ is expressed as ΨT= (ΨA↑,ΨB↑,ΨA↓,ΨB↓), where the superscriptTdenotes transposition. Spectrum and eigenspinors of the Hamiltonian ( 1) with uniform SOCs are briefly reviewed in Appendix A. The spec- trum consists of four branches Eα,ǫ(k) labelled by the two quantum numbers ǫ=±1 andα=±1. Here, the first distinguishes particle and hole branches, the second gives the sign of the expectation value of the spin pro- jection along the in-plane direction perpendicular to the propagation direction k. The spectrum strongly depends on the ratio η=∆ λ. (3) Forη >1/2 a gap separates particle and hole branches. The gap closes at η= 1/2 and forη <1/2 one particle branch and one hole branch are degenerate at k= 0 (see Fig.8in App.A). We now briefly summarize the transfer matrix ap- proach employed in this paper to solve the DW scat- tering problem.35–38We assume translational invariance in they-direction, thus the scattering problem for the Hamiltonian( 1)reducestoaneffectivelyone-dimensional (1D) one. The wave function factorizes as Ψ( x,y) = eikyyχ(x), wherekyis the conserved y-component of the momentum, which parameterizes the eigenfunctions of the Hamiltonian of given energy E. For simplicity we consider piecewise constant profiles of SOCs, and solve the DW equation in each region of constant couplings. Then we introduce the x-dependent3 0 !/4 !/2 0 !/4 !/2 Incident angle !Refraction angles "#"-"+ 0 !/4 !/2 !-!+(a) (b) Figure 2: (Color online) Refraction angles as function of th e incidence angle for fixed energy and fixed SOCs. Panel (a): E= 5,λ= 0.5, ∆ = 2; panel (b): E= 5,λ= 2, ∆ = 0 .5. 4×4 matrix Ω( x), whose columns are given by the com- ponents of the independent, normalized eigenspinor of the 1D DW Hamiltonian at fixed energy.39Due to the continuity of the wave function at each interface between regions of different SOC, the wave function on the left of the interface can be expressed in terms of the wave function on the right via the transfer matrix M=/bracketleftbig Ω(x− 0)/bracketrightbig−1Ω(x+ 0), (4) wherex0is the position of the interface and x± 0=x0±δ with infinitesimal positive δ. The condition det M= 1 guarantees conservation of the probability current across theinterface. Thegeneralizationtothecaseofasequence ofNinterfaces at positions xi,i= 1,...,N, is straight- forward since the transfer matrices relative to individual interfaces combine via matrix multiplication: M=N/productdisplay i=1/bracketleftbig Ω(x− i)/bracketrightbig−1Ω(x+ i). (5) From the transfer matrix it is straightforward to deter- mine transmission and reflection matrices, which encode all the relevant information on the scattering properties. III. THE N-SO INTERFACE First we concentrate on the elastic scattering problem at the interface separating a normal region N ( x <0), where SOCs vanish, and a SO region ( x >0), where SOCs are finite and uniform. We consider a quasiparticle of energy E, withEas- sumed positive for definiteness and outside the gap pos- sibly opened by SOCs. This quasiparticle incident from the normal region is characterized by the y-component of the momentum, or equivalently, the incidence angle φ measured with respect to the normal at the interface, seeFig.1. Conservation of kyimplies that kN y=Esinφ=Eαsinξα=kSO y (6a) kN x=Ecosφ (6b) kSO xα=Eαcosξα (6c) whereα=±1 andEα=/radicalbig (E−∆)(E+∆−αλ). The refraction angles ξαare fixed by momentum conservation along the interface ( 6a) and read ξα= arcsin/parenleftbiggE Eαsinφ/parenrightbigg . (7) Figure1illustrates the refraction process at the N- SO interface. The incident wave function, assumed to have fixed spin projection in the z-direction, in the SO region splits in a superposition of eigenstates of the SOCs Hamiltonian corresponding to states in the differ- ent branches of the spectrum. These eigenstates prop- agate along two distinct directions characterized by the anglesξα, whose difference depends on SOC and is an in- creasing function of the incidence angle, see Fig. 2. The anglesξαcoincide only for normal incidence or for λ= 0. Equation ( 7) implies that there exists a critical angle for each of the two modes given by ˜φα= arcsin/parenleftbiggEα E/parenrightbigg . (8) Forφlarger than both critical angles ˜φα, the quasipar- ticle is fully reflected, since there are no available trans- mission channels in the SO region. For φin between the two critical angles the quasiparticletransmits only in one channel.40 After this qualitative discussion of the kinematics, we now present the exact solution of the scattering problem. In the N region x <0 a normalized scattering state of energyE >0, incident from the left on the interface with incidence angle φand spin projection sis given by χN(x) =[δ↑,s|↑/angb∇acket∇ight+δ↓,s|↓/angb∇acket∇ight]/parenleftbigg1 eiφ/parenrightbiggeikxx /radicalbig2vx F +[r↑s|↑/angb∇acket∇ight+r↓s|↓/angb∇acket∇ight]/parenleftbigg1 −e−iφ/parenrightbigge−ikxx /radicalbig2vx F,(9) wherekx≡kN x(cf. Eq. 6b). Here the index s=↑,↓ specifies the spin projection of the incoming quasipar- ticle with |↑/angb∇acket∇ightand|↓/angb∇acket∇ighteigenstates of szandδi,jis the Kronecker delta. The velocity vx F= cosφis included to ensure proper normalization of the scattering state. The complex coefficients rs′sare reflection probability ampli- tudes for a quasiparticle with spin sto be reflected with spins′. The associated matrix Ω N(x) reads ΩN(x) =1/radicalbig2vx F eikxxe−ikxx0 0 ei(kxx+φ)−e−i(kxx+φ)0 0 0 0 eikxxe−ikxx 0 0 ei(kxx+φ)−e−i(kxx+φ) .4 00.1 0.2 0.3 0.4 0.5 -90 -75 -60 -45 -30 -15 015 30 45 60 75 90 (a) 00.1 0.2 0.3 0.4 0.5 -90 -75 -60 -45 -30 -15 015 30 45 60 75 90 (b) 00.1 0.2 0.3 0.4 0.5 0.6 -90 -75 -60 -45 -30 -15 015 30 45 60 75 90 (c) Figure 3: (Color online) Angular dependence of the transmis - sion probabilities T+↑(blue dashed line) and T−↑(red solid line) at energy E= 2.5. The SOC are fixed as follows: (a) λ= 0.1 and ∆ = 0, (b) λ= 0 and ∆ = 0 .1, and (c) λ= 0.5 and ∆ = 0 .1. Similarly the wave function in the SO region ( x>0) can be expressed in general form as χSO(x) =1/radicalbigvx ++/bracketleftbig t+ψ++(x)+r+¯ψ++(x)/bracketrightbig +1/radicalbigvx −+/bracketleftbig t−ψ−+(x)+r−¯ψ−+(x)/bracketrightbig (10) wheret±(resp.r±) are complex amplitudes for right- moving (resp. left-moving) states. The coefficient tα represents the transmission amplitude into mode α. The wave functions ψα+and the Fermi velocities vx α+in the SO region are obtained from the expressions given in App.Awith the replacement kx→kSO xα, where for no- tational simplicity the label SO will be understood. The wave functions ¯ψα+are in turn obtained from ψα+by replacingkxα→ −kxα. The matrix Ω SO(x) then reads ΩSO(x) = (11) e−iξ+−θ+ 2−eiξ+−θ+ 2e−iξ−−θ− 2−eiξ−−θ− 2 eθ+ 2 eθ+ 2 eθ− 2 eθ− 2 ieθ+ 2 ieθ+ 2 −ieθ− 2−ieθ− 2 ieiξ+−θ+ 2−ie−iξ+−θ+ 2−ieiξ−−θ− 2ie−iξ−−θ− 2 N+eikx++x0 0 0 0N+e−ikx+x0 0 0 0 N−eikx−x0 0 0 0 N−e−ikx−x where in the second matrix the normalization factors are defined as Nα= 1/(2/radicalbig vα+sinhθα). According to Eq. ( 4) the transfer matrix for the sin- gle interface problem is given by the matrix product M= [ΩN(0−)]−1ΩSO(0+). FromMwe obtain the trans- mission and the reflection probabilities for a spin-up orspin-down incident quasiparticle: T+s=/vextendsingle/vextendsingle/vextendsingle/vextendsingleM33δs,↑+M13δs,↓ M13M31−M11M33/vextendsingle/vextendsingle/vextendsingle/vextendsingle2 Υ+(φ), (12) T−s=/vextendsingle/vextendsingle/vextendsingle/vextendsingleM31δs,↑+M11δs,↓ M13M31−M11M33/vextendsingle/vextendsingle/vextendsingle/vextendsingle2 Υ−(φ), (13) R↑s=/vextendsingle/vextendsingle/vextendsingle/vextendsingleM31M23−M33M21 M13M31−M11M33/vextendsingle/vextendsingle/vextendsingle/vextendsingle2 δs,↑ +/vextendsingle/vextendsingle/vextendsingle/vextendsingleM13M21−M11M23 M13M31−M11M33/vextendsingle/vextendsingle/vextendsingle/vextendsingle2 δs,↓,(14) R↓s=/vextendsingle/vextendsingle/vextendsingle/vextendsingleM31M43−M33M41 M13M31−M11M33/vextendsingle/vextendsingle/vextendsingle/vextendsingle2 δs,↑ +/vextendsingle/vextendsingle/vextendsingle/vextendsingleM13M41−M11M43 M13M31−M11M33/vextendsingle/vextendsingle/vextendsingle/vextendsingle2 δs,↓,(15) where Υ α(φ) =θ(˜φα−φ)θ(˜φα+φ) withθ(x) the Heav- iside step function. Here, Tαsis the probability for an incident quasiparticle with spin projection sto be trans- mittedinmode αintheSOregion. Ofcourse,probability current conservation enforces T+s+T−s+R↑s+R↓s= 1. Figures3(a)–(c) show the angular dependence of the transmission probabilities for an incident spin-up quasi- particle into the (+) and ( −) modes of the SO region for different values of the SOCs. Panel (a) refers to the case of vanishing intrinsic SOC (∆ = 0). Here the (+) and the (−) energy bands are separated by a SOC-induced splitting ∆Eext=λ. Therefore at fixed energy the two propagating modes in the SO region have two different momenta, which gives rise to the two different critical angles (cf. Eq. ( 8) with ∆ = 0). Panel (b) refers to the caseλ= 0, where the SOC opens a gap ∆ Eint= 2∆ between the particle- and the hole-branches, however the (+)/(−)-modes remain degenerate. Therefore at fixed energy these modes have the same momentum and, as a consequence, the same critical angles (cf. Eq. ( 8) for λ= 0 and ∆ /negationslash= 0). When both SOCs are finite — the situation illustrated in panel (c) — the transmission probabilitiesexhibit morestructure. Forincidenceangles smaller than ˜φ+no particular differences with the cases of panels (a) and (b) are visible. When the (+) mode is closed, an increase (resp. decrease) of the ( −) mode transmission is observed for positive (resp. negative) an- gles, before the transmission drops to zero for incidence angles approaching ˜φ−. The asymmetry between posi- tive and negative angles is reversed if the spin state of the incident quasiparticle is reversed. These symmetryproperties can be rationalizedby con- sidering the operator of mirror symmetry through the x-axes.41This consists of the transformation y→ −y and at the same time the inversion of the spin and the pseudo-spin states. It reads Sy= (σx⊗sy)Ry, (16) whereRytransforms y→ −y. The operator Sycom- mutes with the total Hamiltonian of the system [ Sy,H0+5 HSO] = 0, therefore allows for a common basis of eigen- states. For the scattering states in the SO region ( 10) we haveSyχ+(ξ+) =χ+(ξ+) andSyχ−(ξ−) =−χ−(ξ−). Instead, it induces the following transformation on the scattering states ( 9) in the normal region: Syχs(φ) = iχ−s(−φ). By comparing the original scattering matrix with the Sy-transformed one we find that Tα,s(φ) =Tα,−s(−φ) (17) withα=±ands=↑,↓, which is indeed the symmetry observedintheplots. Theasymmetryofthetransmission coefficients occurs only when both SOCs are finite. A. Edge states at the interface In addition to scatteringsolutions ofthe DW equation, it is interesting to study the possibility that edge states exist at the N-SO interface, which propagate alongthe interface but decay exponentially away from it. The in- terestinthesetypesofsolutionsisconnectedtothestudy of topological insulators. It has been shown — first by Kane and Mele15— that a zig-zag graphene nanoribbon with intrinsic SOC supports dissipationless edge states within the SOC gap. In fact, similar states are always expected to exist at the interface between a topologically trivial and a topologically non-trivial insulator. In our case, the latter is represented by graphene with intrinsic SOC. Of course SOC-free graphene is not an insulator, howeverit is topologicallytrivial and edgestate solutions do arise for |ky|>|E|. WhenEis within the gap in the SO region the corresponding mode is evanescent along thexdirection on both sides of the interface. Note that 012345012345 kyE(ky) Figure 4: (Color online). Energy dispersion of the edge stat e at the N-SO interface as a function of the momentum along the interface kyfor different values of SOCs. Solution of the transcendental equation is allowed only for |ky|>|E|(white area). In all three cases shown η >1/2: ∆ = 1 and λ= 0.4 (lower-red line), ∆ = 1 .5 andλ= 0.7 (middle-blue line), and ∆ = 2 and λ= 0.9 (upper-green line).the edge state we find is different from the one discussed in Refs.15,32where zig-zag or hard-wall boundary con- ditions at the edge of the SOC region were imposed. The wave function on the N side then reads χN(x) =/parenleftbigg1 −iq+iky E/parenrightbigg (A|↓/angb∇acket∇ight+B|↑/angb∇acket∇ight)eqx(18) withq=/radicalbig |ky|2−E2. In the SO region the wave func- tion can be written as χSO(x) =C (−q++ky) i(E−∆) E−∆ i(q++ky) e−q+x+D (q−−ky) −i(E−∆) E−∆ (q−+iky) e−q−x withqα=/radicalig k2y−(E−∆)(E+∆−αλ). The continu- ity of the wave function at the N-SO interface leads to a linear system of equations for the amplitudes AtoD. The matrix of coefficients must have a vanishing deter- minant for a non-trivial solution to exist. This condition provides a transcendental equation for the energy of pos- sible edge states, whose solutions are illustrated in Fig. 4 for different values of the intrinsic and extrinsic SOCs. The condition |ky|>|E|implies that solutions only exist outside the shadowed area. In addition, they are allowed onlyin the caseSOCsopena gapinthe energyspectrum, which occurs when η>1/2 (see App. Aand Eq. ( 3)). As can be seen in Fig. 4the result is quite insensitive to the precise value of the extrinsic SOC. Edge states exist only for values of the momentum along the interface larger than the intrinsic SOC, i.e., ky>kmin y= ∆. The apparent breaking of time-reversal invariance (the dispersion is not even in ky) is due to the fact that we are considering a single-valley theory. The full two-valley SOC Hamiltonian is invariant under time- reversal symmetry, that interchanges the valley quantum number. This invariance implies that solutions for neg- ative values of kycan be obtained by considering the Dirac-Weyl Hamiltonian relative to the other valley. The twocounter-propagatingedge states live then at opposite valleysand haveoppositespinstateandrealizeapeculiar 1D electronic system. As mentioned in the Introduction, the intrinsic SOC is estimated to be much smaller than the extrinsic one, therefore in a realistic situation one would not expect the opening of a significant energy gap and the presence of edge states. It would be interesting to explore the pos- sibility to artificially enhance the intrinsic SOC, thereby realizing the condition for the occurrence of edge states. IV. THE N-SO-N INTERFACE The analysis of the scattering problem on a N-SO in- terface of the previous section can be straightforwardly generalized to the case of a double N-SO-N interface (SO barrier). Here the transmission matrix Dis given by6 00.2 0.4 0.6 0.8 1 -90 -75 -60 -45 -30 -15 015 30 45 60 75 90 (a) 00.2 0.4 0.6 0.8 1 -90 -75 -60 -45 -30 -15 015 30 45 60 75 90 (b) 00.2 0.4 0.6 0.8 1 -90 -75 -60 -45 -30 -15 015 30 45 60 75 90 (c) Figure 5: (Color online). Panel (a): Angular plots for T↑↑as a function of the injection angle for E= 2, ∆ = 1 and λ= 0. The three lines correspond to different distance between the interfaces: d=π/2 (dashed black), d=π(dotted red), and d= 2π(solid blue). The spin-precession length is ℓSO=π. Whenλ= 0 the transmission probability in the spin state opposed to the injected spin is always zero. Panel (b) and (c): angular plots of T↑↑(solid-blue) and T↓↑(dashed red) as a function of the injection angle for E= 2,λ= 1 and ∆ = 0. The distance between the two interfaces is d=πin panel (a) and d= 2πin panel (b). The spin-precession length is ℓSO= 2π. Eq. (5) withN= 2. The transmission and the reflection probabilities in the case of a spin-up or -down incident quasiparticle read T↑s=/vextendsingle/vextendsingle/vextendsingle/vextendsingleD33δs,↑+D13δs,↓ D13D31−D11D33/vextendsingle/vextendsingle/vextendsingle/vextendsingle2 , (19) T↓s=/vextendsingle/vextendsingle/vextendsingle/vextendsingleD31δs,↑+D11δs,↓ D13D31−D11D33/vextendsingle/vextendsingle/vextendsingle/vextendsingle2 , (20) R↑s=/vextendsingle/vextendsingle/vextendsingle/vextendsingleD31D23−D33D21 D13D31−D11D33/vextendsingle/vextendsingle/vextendsingle/vextendsingle2 δs,↑ +/vextendsingle/vextendsingle/vextendsingle/vextendsingleD13D21−D11D23 D13D31−D11D33/vextendsingle/vextendsingle/vextendsingle/vextendsingle2 δs,↓,(21) R↓s=/vextendsingle/vextendsingle/vextendsingle/vextendsingleD31D43−D33D41 D13D31−D11D33/vextendsingle/vextendsingle/vextendsingle/vextendsingle2 δs,↑ +/vextendsingle/vextendsingle/vextendsingle/vextendsingleD13D41−D11D43 D13D31−D11D33/vextendsingle/vextendsingle/vextendsingle/vextendsingle2 δs,↓.(22) Inthiscasethereisanadditionalparameterwhichcon- trols the scattering properties of the structure, namely the widthdof the SO region. In order to compare this length to a characteristic length scale of the system, we introduce the spin-precession length defined as ℓSO= 2π/planckover2pi1vF λ+2∆. (23) The intrinsic SOC alone cannot induce a spin preces- sion on the carriers injected into the SO barrier — an injected spin state, say up, is obviously never converted into a spin-down state. Figure 5(a) shows the angular00.2 0.4 0.6 0.8 1 -90 -75 -60 -45 -30 -15 015 30 45 60 75 90 (a) 00.2 0.4 0.6 0.8 1 -90 -75 -60 -45 -30 -15 015 30 45 60 75 90 (b) 00.2 0.4 0.6 0.8 1 -90 -75 -60 -45 -30 -15 015 30 45 60 75 90 (c) Figure 6: (Color online). Angular plot of T↑↑(solid-blue) andT↓↑(dashed-red) as a function of the injection angle for E= 2,λ= 1 and (a) ∆ = λ/4, (b) ∆ = λ/2, and ∆ = λ. The distance between the two interfaces is kept fixed to d=ℓSO. dependence of the transmission in the case of injection of spin-up. The behavior of the transmission as a function of the injection angle depends sensitively on the width d compared to the spin-precession length. For small width d < ℓSO(dashed line) the transmission is a smooth de- creasing function of φand stays finite also for φlarger than the critical angle. In the case d≥ℓSO(dotted- and solid-lines) instead the transmission probability exhibits a resonant behavior and drops to zero as soon as the injection angle equals the critical angle. When only the extrinsic SOC is finite, the transmis- sion behavior changes drastically. Two different critical angles appear — the biggest coincides usually with π/2. The extrinsic SOC induces spin precession because of the coupling between the pseudo- and the real-spin. This is illustrated in Fig. 5(b)-(c). In Panel (b) we consider the case of spin-up injection with d=ℓSO/2. At normal inci- dence the transmission is entirely in the spin-down chan- nel (dashed line). Moving away from normal incidence, the transmission in the spin-up channel (solid line) in- creases from zero and, after the first critical angle, the transmissions in spin-up and spin-down channels tend to coincide. In panel (c) the width of the barrier is set to d=ℓSO. Here, the width of the SO region permits to an injected carrier at normal incidence to perform a com- plete precession of its spin state — the transmission is in the spin-up channel. For finite injection angles the spin- down transmission (dashed line) also becomes finite. For φ/lessorsimilar˜φ+the transmission in the spin-up channel is almost fully suppressed while that in the spin-down channel is large. Finally, for φ >˜φ+the two transmission coeffi- cients do not show appreciable difference. In the case where both extrinsic and intrinsic SOC are finite, the transmission probability exhibits a richer structure. We focus again on the case of injection of spin-up quasiparticles. Moreover we fix the width of the SO regionsothat it is alwaysequalto the spin-precession lengthd=ℓSO. Fig.6illustrates the transmission proba- bilitiesTs↑forthree values ofthe ratio∆ /λ= 1/4,1/2,1. Notice that from panel (a) to (c) the width of SO region7 00.2 0.4 0.6 0.8 1-90 -75 -60 -45 -30 -15 015 30 45 60 75 90 (a) -90 -45 0 45 90 -0.4 -0.2 00.2 0.4 ϕ Injection angle Pz Polarization (b) Figure 7: (Color online). Panel (a): total transmission T as a function of the injection angle for E= 2,d= 2πand several values of SOCs: λ= 1 and ∆ = 0 (blue-solid line), λ= 0 and ∆ = 0 .5 (red-dotted line), λ= 1 and ∆ = λ/4 (yellow-dashed line), ∆ = λ/2 (orange-dashed-dotted line), andλ= ∆ (black-dotted-dotted-dashed line). Panel (b): z- component of the spin polarization Pzas a function on the injection angle for E= 2 and d= 2πand the following values of the SOCs: λ= 1, ∆ = 0 and λ= 0, ∆ = 1 (same black- dashed line), λ= 1 and ∆ = λ/4 (red-dotted), ∆ = λ/2 (blue-dotted-dashed line), and ∆ = λ(green-solid line). decreases. The symmetry properties of the transmission func- tion can be rationalized by using the symmetry opera- tion (16). Proceeding in a similar manner as in the case of the single interface, for the SO barrier we find the following symmetry relations Ts,s(φ) =Ts,s(−φ), (24a) Ts,−s(φ) =T−s,s(−φ), (24b) which are confirmed by the explicit calculations. So far we have considered the injection of a pure spin state — the injected carrier was either in a spin-up state or a spin-down state. Following Ref. 30we now address the transmission of an unpolarized statistical mixture of spin-up and spin-down carriers. This will characterize the spin-filtering properties of the SO barrier. In the injection N region, an unpolarized statistical mixture of spins is defined by the density matrix ρin=1 2|χ↑/angb∇acket∇ight/angb∇acketleftχ↑|+1 2|χ↓/angb∇acket∇ight/angb∇acketleftχ↓|, (25) where|χs/angb∇acket∇ight ≡ |s/angb∇acket∇ight ⊗ |σ/angb∇acket∇ightwith|σ/angb∇acket∇ight= (1/√ 2)(1,eiφ) cor- responds to a pure spin state. When traveling throughthe SO region, the injected spin-unpolarized state is sub- jected to spin-precession. The density matrix in the out- put N region can be expressed in terms of the transmis- sion functions ( 19) as ρout=1 2T↑|ζ↑/angb∇acket∇ight/angb∇acketleftζ↑|+1 2T↓|ζ↓/angb∇acket∇ight/angb∇acketleftζ↓|,(26) where the coefficients Ts=T↑s+T↓sare the total trans- missions for fixed injection state. The spinor part is de- fined as |ζs/angb∇acket∇ight=1√Ts/parenleftbigg t↑s t↓s/parenrightbigg ⊗|σ/angb∇acket∇ight, (27) wherets′,sare the transmission amplitudes for incoming (resp. outgoing) spin s(resp.s′). The output density matrix is used to define the total transmission T=T↑+T↓ 2(28) andthe expectation valueofthe zcomponent ofthe spin- polarization Pz=1 2(T↑↑+T↑↓−T↓↑−T↓↓).(29) In Fig.7we report the total transmission (panel (a)) and thez-component of the spin-polarization (panel (b)) as a function of the injection angle for fixed energy and width of the SO region. We observe that for an un- polarized injected state the transmission probability is an even function of the injection angle T(φ) =T(−φ). Moreover, for injection angles larger than the first criti- cal angleφ >˜φ+, the transmission has an upper bound atT= 1/2. On the contrary Pzis an odd function of the injection angle Pz(φ) =−Pz(−φ). It is zero when at least one SOC is zero. When both SOC parameters are finite Pzis finite and reaches the largest values for φ>˜φ+. The maxima in this case increase as a function of the intrinsic SOC. To experimentally observe this polarization effect the measurement should not involve an average over the an- gleφ, which, otherwise — due to the antisymmetry of Pz — would wash out the effect. To achieve this, one could use,e.g., magnetic barriers,37,42which are known to act as wave vector filters. V. CONCLUSIONS In this paper we have studied the spin-resolved trans- missionthroughSOnanostructuresin graphene, i.e., sys- tems where the strength of SOCs — both intrinsic and extrinsic — is spatially modulated. We have considered the case ofan interface separatinga normal regionfrom a SO region, and a barrier geometry with a region of finite SOC sandwiched between two normal regions. We have shown that — because of the lift of spin degeneracy due to the SOCs — the scatteringat the single interface gives8 rise to spin-double refraction: a carrier injected from the normal region propagates into the SO region along two different directions as a superposition of the two avail- able channels. The transmission into each of the two channels depends sensitively on the injection angle and on the values of SOC parameters. In the case of a SO barrier, this result can be used to select preferential di- rections along which the spin polarization of an initially unpolarized carrier is strongly enhanced. We have also analyzed the edge states occurring in the single interface problem in an appropriate range of pa- rameters. These states exist when the SOCs open a gap in the energy spectrum and correspond to the gapless edge states supported by the boundary of topological in- sulators. A natural follow-up to this work would be the detailed analysis of transport properties of such SO nanostruc- tures. From our results for the transmission probabil- ities, spin-resolved conductance and noise could easily be calculated by means of the Landauer-B¨ uttiker formal- ism. Moreoverweplantostudyothergeometries,as, e.g., nanostructureswith aperiodic modulation ofSOCs. The effects of various types of impurities on the properties discussed here is yet another interesting issue to address. We hope that our work will stimulate further theo- retical and experimental investigations on spin transport properties in graphene nanostructures. Acknowledgments We gratefully acknowledge helpful discussions with L. Dell’Anna, R. Egger, H. Grabert, M. Grifoni, W. H¨ ausler, V. M. Ramaglia, P. Recher and D. F. Urban. The work of DB is supported by the Excellence Initia- tive of the German Federal and State Governments. The work of ADM is supported by the SFB/TR 12 of the DFG. Appendix A: Graphene with uniform spin-orbit interactions. In this appendix we briefly review the basic proper- ties of DW fermions in graphene with homogeneous SO interactions.21The energy eigenstates are plane waves ψ∼Φ(k)eik·rwith Φ a four-componentspinorand eigen- values given by ( vF=/planckover2pi1= 1) Eα,ǫ(k) =αλ 2+ǫ/radicaligg k2x+k2y+/parenleftbigg ∆−αλ 2/parenrightbigg2 ,(A1) whereα=±andǫ=±. The energy dispersion as a function of kxat fixedky= 0 is illustrated in Fig. 8for several values of ∆ and λ. The index ǫ=±specifies the particle/holebranchesofthe spectrum. The eigenspinorsΦα,ǫ(k) read ΦT α,ǫ(k) =1 2√coshθα× (A2) (e−iφ−ǫθα/2,ǫeǫθα/2,iαǫeǫθα/2,iαeiφ−ǫθα/2), whereTdenotes transposition and sinhθα=αλ/2−∆ k, (A3) eiφ=kx+iky k, (A4) withk=/radicalig k2x+k2y. The spin operator components are expressed as Sj=1 2sj⊗σ0. Their expectation values in the eigenstate Φ α,ǫread /angb∇acketleftSx/angb∇acket∇ight=−ǫαsinφ 2coshθα, (A5a) /angb∇acketleftSy/angb∇acket∇ight=ǫαcosφ 2coshθα, (A5b) /angb∇acketleftSz/angb∇acket∇ight= 0, (A5c) which shows that the product ǫαcoincides with the sign of the expectation value of the spin projection along the inplanedirectionperpendiculartothedirectionofpropa- gation. For vanishing extrinsic SOC, the eigenstates Φ α,ǫ reduce to linear combinations of eigenstates of Sz. Similarly, the expectation value of the pseudo-spin op- (b) -4 -2 024 Momentum (d)-4 -2 024Energy (a) -4 -2 024 Momentum -4 -2 024Energy (c) Figure 8: Spectrum of the DW Hamiltonian with intrinsic and Rashba SOC as a function of kxforky= 0 . Panel (a): dashed lines refer to ∆ = 0 .5 andλ= 0; solid and dotted lines refer to ∆ = 0 and λ= 1. Panel (b): ∆ = 0 .4 andλ= 1. Panel (c): ∆ = 0 .5 andλ= 1. Panel (d): ∆ = 0 .8 andλ= 1.9 eratorσis given by /angb∇acketleftσx/angb∇acket∇ight=ǫcosφ coshθα, (A6a) /angb∇acketleftσy/angb∇acket∇ight=ǫsinφ coshθα. (A6b) Since the SOCs in graphene do not depend on momen- tum, thevelocityoperatorstillcoincideswiththepseudo- spin operator: v=˙r= i[H,r] =σ. Thus the velocityexpectationvalueinthestateΦ α,ǫisgivenbyEqs. ( A6a– A6b). Alternatively, it can be obtained from the energy dispersion as vα,ǫ=∇kEα,ǫ=ǫk/radicalig k2+/parenleftbig ∆−αλ 2/parenrightbig2.(A7) The groupvelocity is then independent of the modulus of the wave vector if either the SOCs vanish or ∆ = αλ/2. ∗Electronic address: dario.bercioux@frias.uni-freiburg.de †Electronic address: ademarti@thp.uni-koeln.de 1K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Griegorieva, and A. A. Firsov, Science306, 666(2004); Nature(London) 438, 197(2005). 2Y. Zhang, Y. W. Tan, H. L. Stormer, and P. Kim, Nature (London) 438, 201 (2005). 3For recent reviews, see A. K. Geim and K. S. Novoselov, Nature Mat. 6, 183 (2007); A.H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, and A. K. Geim, Rev. Mod. Phys. 81, 109 (2009); A. K. Geim, Science 324, 1530 (2009). 4C. W. J. Beenakker, Rev. Mod. Phys. 80, 1337 (2008). 5G. W. Semenoff, Phys. Rev. Lett. 53, 2449 (1984). 6D. P. Di Vincenzo and E.J. Mele, Phys. Rev. B 29, 1685 (1984). 7I.ˇZuti´ c, J. Fabian, S. Das Sarma, Rev. Mod. Phys. 76, 323 (2004). 8B. Trauzettel, D.V. Bulaev, D. Loss, and G. Burkard, Na- ture Phys. 3, 192 (2007). 9E. W. Hill, A.K.Geim, K. S. Novoselov, F. Schedin, and P. Blake, IEEE Trans. Magn. 42(10),2694 (2006). 10N. Tombros, C. Jozsa, M. Popinciuc, H. T. Jonkman, and B. J. van Wees, Nature 448, 571 (2007). 11S. Cho, Y.F. Chen, and M. S. Fuhrer, Appl. Phys. Lett. 91123105 (2007). 12M. Nishioka and A. M. Goldman, Appl. Phys. Lett. 90 252505 (2007). 13C. J´ ozsa, M. Popinciuc, N. Tombros, H. T. Jonkman, and B. J. van Wees, Phys. Rev. Lett. 100, 236603 (2008) 14N. Tombros, S. Tanabe, A. Veligura, C. Jozsa, M. Popin- ciuc, H. T. Jonkman, and B. J. van Wees, Phys. Rev. Lett. 101, 046601 (2008). 15C. L. Kane and E. J. Mele, Phys. Rev. Lett. 95, 226801 (2005). 16D. Huertas-Hernando, F. Guinea, and A. Brataas, Phys. Rev B74155426 (2006). 17Hongki Min, J.E. Hill, N. A. Sinitsyn, B. R. Sahu, L. Kleinman, and A.H. MacDonald, Phys. Rev. B 74, 165310 (2006). 18Y. Yao, F. Ye, X. L. Qi, S. C. Zhang, and Z. Fang, Phys. Rev. B75, 041401(R) (2007). 19J. C. Boettger and S. B. Trickey, Phys. Rev. B 75, 121402(R) (2007); Phys. Rev. B 75, 199903(E) (2007). 20M. Zarea and N. Sandler, Phys. Rev. B 79, 165442 (2009). 21E. I. Rashba, Phys. Rev. B 79, 161409(R) (2009). 22D. Huertas-Hernando, F. Guinea, and A. Brataas, Phys. Rev. Lett. 103, 146801 (2009). 23C. Ertler, S. Konschuh, M. Gmitra, and J. Fabian, Phys.Rev. B80, 041405(R) (2009). 24P. Ingenhoven, J. Z. Bern´ ad, U. Z¨ ulicke, and R. Egger Phys. Rev. B 81, 035421 (2010) . 25A. Varykhalov, J. S´ anchez-Barriga, A. M. Shikin, C. Biswas, E. Vescovo, A. Rybkin, D. Marchenko, and O. Rader, Phys. Rev. Lett. 101, 157601 (2008). 26O. Rader, A. Varykhalov, J. S´ anchez-Barriga, D. Marchenko, A. Rybkin, and A. M. Shikin, Phys. Rev. Lett. 102, 057602 (2009). 27A. Varykhalov and O. Rader, Phys. Rev. B 80, 035437 (2009). 28V. M. Ramaglia, D. Bercioux, V. Cataudella, G. De Filip- pis, C. A. Perroni, and F. Ventriglia, Eur. Phys. J. B 36, 365 (2003). 29M. Khodas, A. Shekhter, and A. M. Finkel’stein, Phys. Rev. Lett. 92, 086602 (2004). 30V. M. Ramaglia, D. Bercioux, V. Cataudella, G. De Fil- ippis, and C. A. Perroni, J. Phys.: Condens. Matter 16, 9143 (2004). 31E. I. Rashba, Fiz. Tverd. Tela (Leningrad) 2, 1224 (1960) [Sov. Phys. Solid State 2, 1109 (1960)]. 32T. Stauber and J. Schliemann, New J. Phys. 11, 115003 (2009). 33A. Yamakage, K.-I. Imura, J. Cayssol, and Y. Kuramoto, EPL87, 47005 (2009). 34F.Guinea, M.I.Katsnelson, andA.G.Geim, NaturePhys. 6, 30 (2009). 35B. H. J. McKellar and G. J. Stephenson, Jr., Phys. Rev. C35, 2262 (1987). 36M. Barbier, F. M. Peeters, P. Vasilopoulos, and J. M. Pereira , Phys. Rev. B 77, 115446. 37L. Dell’Anna and A. De Martino, Phys. Rev. B 79, 045420 (2009); Phys. Rev. B 80, 089901(E) (2009). 38L. Dell’Anna and A. De Martino, Phys. Rev. B 80, 155416 (2009). 39Eigenspinors are normalized in order to ensure probability flux conservation across the interface. 40For incidence angle larger than one of the two critical an- gles, the respective refraction angle becomes complex: ξα=π 2−iξ′ α, where the correct determination of the imaginary part is obtained for ξ′>0 by the relations sin ξα= coshξ′ αand cosξα= isinhξ′ α. 41F. Zhai and H. Q. Xu, Phys. Rev. Lett. 94, 246601 (2005). 42A. De Martino, L. Dell’Anna, and R. Egger, Phys. Rev. Lett.98, 066802 (2007); Sol. St. Comm. 144, 547 (2007).
1509.06118v1.Nature_of_Valance_Band_Splitting_on_Multilayer_MoS2.pdf
1 Nature of Valance Band Splitting on M ultilayer MoS 2 Xiaofeng Fana, *, W.T. Zhenga, and David J. Singha,b, † a. College of Materials Science and Engineering, Jilin University, Changchun 1300 12, China b. Department of Physics and Astronomy, University of Missouri, Columbia, Missouri 65211 -7010, USA *E-mail: xffan@jlu.edu.cn ; † E-mail: singhdj@ missouri.edu Abstract Understanding the origin of splitting of valance band is important since it govern s the unique spin and valley ph ysics in few -layer MoS 2. With first principle methods, we explore the effects of spin -orbit coupling and layer ’s coupling on few -layer MoS 2. It is found that intra-layer spin -orbit coupling has a major contribution to t he splitting of valance band at K. In double -layer MoS 2, the layer ’s coupling result s in the widen ing of energy gap of splitted states induced by intra -layer spin -orbit coupling. The valance band splitting of bulk MoS 2 in K can follow this model. We also f ind the effect of inter -layer spin -orbit coupling in triple -layer MoS 2. In addition , the inter -layer spin -orbit coupling is found to become to be stronger under the pressure and results in the decrease of main energy gap in the splitting valance bands at K. . Introduction A new class of 2D materials, the single -layer and/or few -layer of hexagonal transition metal dichalcogenides (h-TMDs) , have attracted broad attentions due to the extraordinary physic al properties and promising applications in electric and optoelectronic devices1-5. As the prototypical 2D materials, single -layer h -TMDs are direct band gap semiconductors with spin -splitting at valance band maximum which is much different from graphene6-9. This promises a chance to manipulate the spin degree of freedom and valley polarization10-12. In addition , with extreme dimensional confinement , tightly -bound excitons and strong electron -electron interactions due to weak screening, h -TMDs have been ideal low -dimensional compounds to explore many interesting quantum phenomena11, 13-16, such as spin- and valley - Hall effects and superconductivity17-19. There are also a lot of fascinating optical properties in single -layer h -TMDs, such as the strong band gap photoluminescence at edge5, surface sensitive luminescence20, 21 and strain -controlled optical band gap22-25, and so on. Among the se h-TMDs, MoS 2 is a representative . Bulk MoS 2 is a layered compound stacked with the weak van der Waals interaction26. Due to the highly anisotropic mechanic al property , it is used in dry lubrication . It has also made the interest due to the special catalytic activity from its edge27. In each layer of MoS 2, there are three atomic layers with a center layer of Mo around S layers in both sides. The state s near band gap are 2 well-known to be mainly from the d -orbitals of Mo28. There is a priori proposal that the layer ’s coupling is possible to have a very weak effect to the states near band gap. Bulk MoS 2 is an indirect -gap semiconductor with a band gap of 1.29 eV. However, following the reduction of layers to sinlge -layer, there is a transition between indirect band gap and direct gap3. The single -layer MoS 2 is found to have a direct band gap of about 1.8 eV5, 29. Therefore, the layer’s coupling has a strong effect to the states near band gap with recent reports30. Especially, the states of valance band top a t point (VB-) and cond uction band bottom along Λ (CB-Λ) is much sensitive to the layer ’s coupling (LC) . Compared with the states of VB - and CB - Λ, the LC effects on the states of valance band top and conduction band bottom at K point (VB -K, CB -K) are very weak. Therefore, there remains an open question about the origin of the splitting at valance band of K point which govern s the unique spin and valley ph ysics. In the single -layer limit, the splitting can be attributed entirely to spin -orbit coupling (SOC). In bulk limit, it is considered to be a result of combination of SOC and LC. However, there is disagreement about the relative strength of both mechanisms31-37. In the work, we explore the effect of SOC on few -layer MoS 2 with th e rule of LC by first principle methods in details. We analyze the splitting of states at VB -, VB -K CB -Λ, and CB -K and explore the change of splitting by following the increase of distance of both layers for double -layer MoS 2. It is found that intra-layer SOC (intra -SOC) has a major contribution to the splitting at VB -K, while LC can open effectively the degeneracy of states at VB -K. With the analysis of charge distribution in real space, the double - degeneracy of states at the valance band maximum of K po int, which isn't broken due to the inter-layer inverse symmetry for both layers which result s in the forbidding of inter-layer SOC, are mainly from the spin -up state of first -layer and spin -down state of second layer. For triple -layer MoS 2, the LC with int er-layer SOC due to the absence of inter-layer inverse symmetry in t hree-layer system makes the splitting complicated. The intra-layer SOC results in two main bands splitting, while in each main band, the triple -degeneracy is broken mainly due to the inter -layer SOC. With the pressure, it is found in double -layer MoS 2 that the double -degeneracy of states in each main band isn ’t broken when the splitting of both main bands is increased due to the strengthening of LC. For triple -layer Mo S2 under large pressure, the splitting of triple -degeneracy in each main band is very obvious. Computational Method The present calculations are performed within density functional theory using accurate frozen -core full -potential projector augmented -wave (P AW) pseudopotentials , as implemented in the VASP code38-40. The generalized gradient approximation (GGA) with the parametrization of Perdew -Burke -Ernzerhof (PBE) and with added van der Waals corrections is used41. The k-space integrals and the plane -wave basis sets are chosen to ensure that the total energy is converged at the 1 meV/atom level. A kinetic energy cutoff of 500 eV for the plane wave expansion is found to be sufficient . The effect of dispersion interaction is included by the empirical correction scheme of Grimme (DFT +D/PBE)42. This approach has been successful in describing layered structures43, 44. The lattice constants a and c of bulk MoS 2 are 3.191 Å and 12.374 Å which is similar 3 to that from the experiments (3.160 Å and 12.295 Å). For the different layered MoS 2, the supercells are constructed with a vacuum space of 20 Å along z direction. The Brillouin zones are sampled with the Γ -centered k -point grid of 18 181. With the state -of-the-art method of adding the stress to stress tensor in V ASP code39, 40, the structure of bulk MoS 2 is optimized under a specified hydrostatic pressure of 15GPa . With these structural parameters from bulk MoS 2, the double - and triple -layer MoS 2 structures under the pressures are constructed. The electronic properties can analyzed with and/or without spin-orbit coupling to explore the band splitting near band gap. The calculated band gap of single -layer MoS 2 without the consideration of spin -orbit interaction is 1.66 eV and less than the experimental report of about 1.8 eV . Obviously , the band gap from PBE is underestimated as in common in usual density functional calculations . Though the band gap is underestimated b y PBE, the band structure near Fermi level doesn ’t have obvious difference from that from other many body method s. Results and discussion The structure of single -layer MoS 2 has the hexagonal symmetry with space group P-6m2. The six sulfur atoms near each Mo atom form a trigonal prismatic structure with the mirror symmetry in c direction. Obviously, the reversal symmetry is absent, and the intra-SOC in the band structure becomes to be free. An obvious band splitting on the valance band maximum around K (K’) point has been observed and is contributed to the SOC. In addition , the SOC also results in the band splitting on conduction band minimum around Λ point, while the splitting at VB - and CB -K is not opened. The states of VB- and CB -K are contribut ed mostly from d z2 orbital of Mo and the effect of the spin-orbit effect is very weak . At the same time, the states of VB-K and CB -Λ are mainly from d x2 -y2 and d xy orbitals of Mo and the spin -orbit effect on Mo can be revealed in the case of the absence of reversal symmetry . The band splitting at VB -K (149 meV) is larger than that at CB-Λ (about 79 meV) . It may be that the distribution of change (or wave function) at CB-Λ around Mo atoms in xy plane is more localized than that at VB -K30. We also notice that the charge distribution of spin -up state is much different from that of spin -down state at VB -K. The spin -down state around Mo is more localized than the spin -up state. For double -layer MoS 2, the interaction between two layers become s important to the states near Fermi level. One of much evident effect s is the direct band gap (K-K) of single -layer becomes to be indirect band gap (-K) due to the uplift of state at VB -. It can be ascribe d to large band splitting (0.618 eV) at VB -. It is also observed that the band splitting at the conduction band bottom around is about 0.352 eV . However, without the consideration of SOC effect , the band splitting at VB -K is just 73.8 meV . The difference of LC’s strength of different states at VB-, VB -K and CB - is ascribed to the charge distribution near sulfur atoms. The large contribution of charge on sulfur atoms for the states at VB - makes the LC to become easier30. The weak LC at VB-K may make the SOC important. In order to explore the rules of SOC and LC in double -layer MoS 2, we calculat ed the change of band structures by following the change of distance between both layers with and without the consideration of spin -orbit effect. As shown in Fig. 1, the band splitting including that at VB -, VB -K and CB - approach es zero quickly following the increase of distance, especially that of VB -K, if the spin -orbit effect is not considered. With 4 SOC, the band splitting at VB-K and CB - converge towards some constants (about 149 meV and 79 meV), while the band splitting at VB - approaches zero. Obviously, there is no spin -orbit effect at VB-. With the large distance between both layers, the effect of LC can be ignored and the splitting is from SOC. It has been well-known that the spin -up and spin -down states at VB -K’ are reversed by compared with that at VB -K in single -layer MoS 2. For double -layer MoS 2, both splitting bands at VB -K are two -degeneracy . As shown in Fig. 2c, the upper band of both bands is composed by the spin -up state of first layer and spin -down state of second layer (~|1, ~|2). The low er band is with the spin -down state of first layer and spin-up state of second layer (~|1, ~|2). Obviously, the energies of spin-up and spin -down of second layer at VB -K are reversed by compared with that of first layer. Since there is reversal symmetry for double -layer system, the energ ies of states |1 and |2 with same energy cannot be split due to the absence of inter-layer SOC (inter -SOC). Therefore, we can understand the splitting at VB -K based on the intra -SOC and LC with the theoretical model shown in Fig. 2a. Because of the splitting of intra -SOC, the energies of |1 and |2 are very different . This reduces largely the coupling of both states due to layer ’s interaction. From the band splitting value (166 meV) of double layer at VB -K, the increased splitting from LC effect is about 17 meV and is much less than that (73.8 meV) from LC without the consideration of spin -orbit effect. For the spin -down channel , the mechanism of band splitting is same to that of spin -up channel. Therefore , the contribution of intra -SOC (149 meV) to the band splitting at VB -K is much larger than that of LC. The same mechanism about the splitting at VB -K (shown in Fig. 2a) can be used for bulk MoS 2. The contribution of LC is increased to about 59 meV since the band splitting at VB -K is about 0.208 eV . If no considering the spin -orbit effect, the band splitting due to LC is abou t 145.7 meV which much similar to the value from SOC(149 meV) in single -layer. This may be the reason that there is disagreement about the relative strength of both effects in bulk limit. Based on the model mentioned above and analysis, the intra -SOC effec t is the main mechanism for the splitting at VB -K in bulk limit . For triple -layer MoS 2, the band splitting near band gap is complicated , since there are three states from three layers which are coupling with each other and hybridized with possible int er-SOC . For the states at VB -, there is no SOC effect and the three degenerate states will be splitting due to LC. It is found that the two splitting values ( 1 and 2 in Fig. 3a) which control the relative energy difference of three states after the hybridization are 0.293 eV and 0.502 eV , respectively . The much different value of both splitting implies that there is strong coupling between first layer and third layer since both splitting value s should be equivalent if the nearest -neighbor interaction is just considered for the three degenerate states . Without the consideration of spin -orbit effect, the splitting values at CB- (1 and 2) are 0.241 eV and 0.225 eV and that at VB -K (K1 and K2) are 49 meV and 55 meV , respectively . Based on the nearest -neighbor LC strength (73.8/2 meV) at VB-K from double layer, the LC strength between first-layer and third layer of triple layer is about 2 meV at VB -K and may be ignored. Therefore, we propose a coupling model based on the in tra-SOC a nd nearest -neighbor LC, as shown in Fig. 3c. With this model, the spin-up and spin -down bands of each layer are splitted by the intra -SOC. Then the LC will perturb these states for each spin channel . For example , the spin -up states are composed 5 with two degenerate upper states (|1 , |3) and one lower state (|2 ) in Fig. 3c and LC will result in the splitting of two degenerate upper states with the increase of energy gap between |2 and |3. With the spin -up and spin -down channels together after LC, there should two main bands and each main band is composed with two degenerate states and one single states. The LC doesn't change the energy gap between the main bands. It is found that the energy gap ’SOC is 148.7 meV and similar to the splitting from i ntra-SOC (SOC =149 meV). However, it is interesting that the two degenerate states in each main band, such as the upper states ~|1 and ~|2 and the lower states ~|2 and ~|3, are splitted, as shown in the inset of Fig. 3c. In addition , it is found that the splitting values are so large (such as, 11.3 meV between ~|1 and ~|2) that the contribution of LC between first layer and third layer is not enough. We propose that the splitting of degenerate states in each main band is from the inter-SOC. While the small pressure doesn ’t induce the obvious splitting from intra -SOC in single -layer MoS 2, it is possible there is strong effect to the splitting from inter -SOC in triple -layer MoS 2. For double -layer MoS 2, the inter -SOC is forbidden and the de generate state in each main band isn ’t opened and the energy gap between both main bands is increased with the strengthening of LC under the pressure in Fig. 4a. In triple -layer, the strengthening of LC under pressure should have the obvious effect, such as the increase of energy gap between ~|1 and ~|3 in Fig. 4c. Besides the enhanced LC effect, an apparent observation is the energy gap between main bands ’SOC has been decreased to 135.8 meV under 15 GPa in Fig. 4b. This should be the typical evidence for the inter -SOC. Conclusions We study the band splitting at valance band maximum of multi -layer MoS 2 by first principle methods in details. We propose a model based on the int ra-layer spin -orbit coupling to solve t he valance band splitting at K point of multi -layer MoS 2 and bulk MoS 2 with the perturbation of layer ’s coupling and inter -layer spin -orbit coupling. It is also found that the direct interaction between second near -neighbor layers is weak at VB -K. While the inter-layer spin -orbit coupling is forbidden in double -layer MoS 2, this effect appear s in triple -layer MoS 2. Especially, under the pressure, the inter -layer spin -orbit coupling is raised with the decrease of energy gap between main bands from intr a-layer spin -orbital coupling. References 1. Wang, Q. H.; Kalantar -Zadeh, K.; Kis, A.; Coleman, J. N.; Strano, M. S., Electronics and Optoelectronics of Two -Dimensional Transition Metal Dichalcogenides. Nat. Nanotech. 2012, 7, 699 -712. 2. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V .; Kis, A., Single -Layer MoS 2 Transistors. Nat Nanotech. 2011, 6, 147 -150. 3. Mak, K. F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T. F., Atomically Thin MoS 2: A New Direct -Gap Semiconductor. Phys. Rev. Lett. 2010, 105, 136805. 4. Lee, C.; Yan, H.; Brus, L. E.; Heinz, T. F.; Hone, J.; Ryu, S., Anomalous Lattice Vibrations of Single - and Few -Layer MoS 2. ACS Nano 2010, 4, 2695 -2700. 6 5. Splendiani, A.; Sun, L.; Zhang, Y .; Li, T.; Kim, J.; Chim, C.-Y .; Galli, G.; Wang, F., Emerging Photoluminescence in Monolayer MoS 2. Nano Lett. 2010, 10, 1271 -1275. 6. Zhang, Y .; Chang, T. -R.; Zhou, B.; Cui, Y . -T.; Yan, H.; Liu, Z.; Schmitt, F.; Lee, J.; Moore, R.; Chen, Y .; Lin, H.; Jeng, H. -T.; Mo, S. -K.; Hussai n, Z.; Bansil, A.; Shen, Z. -X., Direct observation of the transition from indirect to direct bandgap in atomically thin epitaxial MoSe2. Nat Nanotech 2014, 9, 111 -115. 7. Yeh, P. -C.; Jin, W.; Zaki, N.; Zhang, D.; Liou, J. T.; Sadowski, J. T.; Al -Mahboob, A .; Dadap, J. I.; Herman, I. P.; Sutter, P.; Osgood, R. M., Layer -dependent electronic structure of an atomically heavy two-dimensional dichalcogenide. Phys. Rev. B 2015, 91, 041407. 8. Zhu, Z. Y .; Cheng, Y . C.; Schwingenschl 枚gl, U., Giant spin -orbit -induce d spin splitting in two-dimensional transition -metal dichalcogeni de semiconductors. Phys. Rev. B 2011, 84, 153402. 9. Zeng, H.; Dai, J.; Yao, W.; Xiao, D.; Cui, X., Valley polarization in MoS 2 monolayers by optical pumping. Nat. Nanotech. 2012, 7, 490 -493. 10. Cao, T.; Wang, G.; Han, W.; Ye, H.; Zhu, C.; Shi, J.; Niu, Q.; Tan, P.; Wang, E.; Liu, B.; Feng, J., Valley -selective circular dichroism of monolayer molybdenum disulphide. Nat Commun 2012, 3, 887. 11. Xiao, D.; Liu, G. -B.; Feng, W.; Xu, X.; Yao, W., Coupled Spin and Valley Physics in Monolayers of MoS2 and Other Group -VI Dichalcogenides. Phys. Rev. Lett. 2012, 108, 196802. 12. Yao, W.; Xiao, D.; Niu, Q., Valley -Dependent Optoelectronics from Inversion Symmetry Breaking. Phys. Rev. B 2008, 77, 235406. 13. Klots, A. R.; Newaz, A. K. M.; Wang, B.; Prasai, D.; Krzyzanowska, H.; Lin, J.; Caudel, D.; Ghimire, N. J.; Yan, J.; Ivanov, B. L.; Velizhanin, K. A.; Burger, A.; Mandrus, D. G.; Tolk, N. H.; Pantelides, S. T.; Bolotin, K. I., Probing excitonic states in suspended two -dimensional semiconductors by photocurrent spectroscopy. Scientific Reports 2014, 4, 6608. 14. Li, X.; Zhang, F.; Niu, Q., Unconventional Quantum Hall Effect and Tunable Spin Hall Effect in Dirac Materials: Application to an Isolated MoS 2 Trilayer. Phys. Rev. Lett. 2013, 110, 066803. 15. Ross, J. S.; Wu, S.; Yu, H.; Ghimire, N. J.; Jones, A. M.; Aivazian, G.; Yan, J.; Mandrus, D. G.; Xiao, D.; Yao, W.; Xu, X., Electrical control of neutral and charged excitons in a monolayer semiconductor. Nat Commun 2013, 4, 1474. 16. Mak, K. F.; He, K.; Lee, C.; Lee, G. H.; Hone, J.; Heinz, T. F.; Shan, J., Tightly bound trions in monolayer MoS 2. Nat Mater 2013, 12, 207 -211. 17. Mak, K. F.; McGill, K. L.; Park, J.; McEuen, P. L., The valley Hall effect in MoS2 transistors. Science 2014, 344, 1489 -1492. 18. Roldá n, R.; Cappelluti, E.; Guinea, F., Interactions and superconductivity in heavily doped MoS${}_{2}$. Phys. Rev. B 2013, 88, 054515. 19. Geim, A. K.; Grigorieva, I. V ., Van der Waals heterostructures. Nature 2013, 499, 419 -425. 20. Mouri, S.; Miyauchi, Y .; Matsuda, K., Tunable Photoluminescence of Monolayer MoS 2 via Chemical Doping. Nano Letters 2013 , (12), 5944 -5948. 21. Tongay, S.; Suh, J.; Ataca, C.; Fan, W.; Luce, A.; Kang, J. S.; Liu, J.; Ko, C.; R aghunathanan, R.; Zhou, J.; Ogletree, F.; Li, J.; Grossman, J. C.; Wu, J., Defects activated photoluminescence in two -dimensional semiconductors: interplay between bound, charged, and free excitons. Scientific Reports 2013, 3, 2657. 22. Feng, J.; Qian, X.; Huang, C. -W.; Li, J., Strain -Engineered Artificial Atom as a Broad -Spectrum Solar Energy Funnel. Nat. Photon. 2012, 6, 866 -872. 23. Fan, X.; Zheng, W.; Kuo, J. -L.; Singh, D. J., Structural Stability of Single -layer MoS 2 under Large Strain. J. Phys.: Conde ns. Matter 2015, 27, 105401. 24. Scalise, E.; Houssa, M.; Pourtois, G.; Afanasev, V .; Stesmans, A., Strain -Induced Semiconductor to 7 Metal Transition in the Two -Dimensional Honeycomb Structure of MoS 2. Nano Research 2012, 5, 43 -48. 25. Conley, H. J.; Wang, B.; Ziegler, J. I.; Haglund, R. F.; Pantelides, S. T.; Bolotin, K. I., Bandgap Engineering of Strained Monolayer and Bilayer MoS2. Nano Lett. 2013, 13, 3626 -3630. 26. Mattheiss, L. F., Band Structures of Transition -Metal -Dichalcogenide Layer Compounds. Phys. Rev. B 1973, 8, 3719 -3740. 27. Kibsgaard, J.; Chen, Z.; Reinecke, B. N.; Jaramillo, T. F., Engineering the Surface Structure of MoS 2 to Preferentially Expose Active Edge Sites for Electrocatalysis. Nat. Mater. 2012, 11, 963 -969. 28. Chang, C. -H.; Fan , X.; Lin, S. -H.; Kuo, J. -L., Orbital analysis of electronic structure and phonon dispersion in MoS 2, MoSe 2, WS 2, and WSe 2 monolayers under strain. Phys. Rev. B 2013, 88, 195420. 29. Kuc, A.; Zibouche, N.; Heine, T., Influence of Quantum Confinement on the Electronic Structure of the Transition Metal Sulfide TS 2. Phys. Rev. B 2011, 83, 245213. 30. Fan, X.; Chang, C. H.; Zheng, W. T.; Kuo, J. -L.; Singh, D. J., The Electronic Properties of Single -Layer and Multilayer MoS2 under High Pressure. J. Phys. Chem. C 2015, 119, 10189 -10196. 31. Latzke, D. W.; Zhang, W.; Suslu, A.; Chang, T. -R.; Lin, H.; Jeng, H. -T.; Tongay, S.; Wu, J.; Bansil, A.; Lanzara, A., Electronic structure, spin -orbit coupling, and interlayer interaction in bulk MoS 2 and WS 2. Phys. Rev. B 2015 , 91, 235202. 32. Klein, A.; Tiefenbacher, S.; Eyert, V .; Pettenkofer, C.; Jaegermann, W., Electronic band structure of single -crystal and single -layer WS 2: Influence of interlayer van der Waals interactions. Phys. Rev. B 2001, 64, 205416. 33. Molina -Sanchez, A.; Sangalli, D.; Hummer, K.; Marini, A.; Wirtz, L., Effect of Spin -Orbit Interaction on the Optical Spectra of Single -Layer, Double -Layer, and Bulk MoS 2. Phys. Rev. B 2013, 88, 045412. 34. Alidoust, N.; Bian, G.; Xu, S. -Y .; Sankar, R.; Neu pane, M.; Liu, C.; Belopolski, I.; Qu, D. -X.; Denlinger, J. D.; Chou, F. -C.; Hasan, M. Z., Observation of monolayer valence band spin -orbit effect and induced quantum well states in MoX 2. Nat Commun 2014, 5, 1312. 35. Eknapakul, T.; King, P. D. C.; Asakawa , M.; Buaphet, P.; He, R. H.; Mo, S. K.; Takagi, H.; Shen, K. M.; Baumberger, F.; Sasagawa, T.; Jungthawan, S.; Meevasana, W., Electronic Structure of a Quasi -Freestanding MoS 2 Monolayer. Nano Letters 2014, 14, 1312 -1316. 36. Jin, W.; Yeh, P. -C.; Zaki, N.; Zhang, D.; Sadowski, J. T.; Al -Mahboob, A.; van der Zande, A. M.; Chenet, D. A.; Dadap, J. I.; Herman, I. P.; Sutter, P.; Hone, J.; Osgood, R. M., Direct Measurement of the Thickness -Dependent Electronic Band Structure of MoS 2 Using Angle -Resolved Photoem ission Spectroscopy. Phys. Rev. Lett. 2013, 111, 106801. 37. Suzuki, R.; Sakano, M.; Zhang, Y . J.; Akashi, R.; Morikawa, D.; Harasawa, A.; Yaji, K.; Kuroda, K.; Miyamoto, K.; Okuda, T.; Ishizaka, K.; Arita, R.; Iwasa, Y ., Valley -dependent spin polarization in bulk MoS 2 with broken inversion symmetry. Nat Nano 2014, 9, 611 -617. 38. Hohenberg, P.; Kohn, W., Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864. 39. Kresse, G.; Furthmü ller, J., Efficient Iterative Schemes for Ab Initio Total -Energy Calculatio ns Using a Plane -wave Basis Set. Phys. Rev. B 1996, 54, 11169 –11186 40. Kresse, G.; Furthmü ller, J., Efficiency of Ab -Initio Total Energy Calculations for Metals and Semiconductors Using a Plane -wave Basis Set Computat. Mater. Sci. 1996, 6, 15 -50. 41. Perdew, J. P.; Burke, K.; Ernzerhof, M., Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865 -3868. 42. Grimme, S., Semiempirical GGA -Type Density Functional Constructed with a Long -Range Dispersion Correction. J. Comput. Chem. 2006, 27, 1787. 43. Fan, X. F.; Zheng, W. T.; Chihaia, V .; Shen, Z. X.; Kuo, J. -L., Interaction Between Graphene and the 8 Surface of SiO 2. J. Phys.: Condens. Matter 2012, 24, 305004. 44. Mercurio, G.; McNellis, E. R.; Martin, I.; Hagen, S.; Leyssner, F.; So ubatch, S.; Meyer, J.; Wolf, M.; Tegeder, P.; Tautz, F. S.; Reuter, K., Structure and Energetics of Azobenzene on Ag(111): Benchmarking Semiempirical Dispersion Correction Approaches. Phys. Rev. Lett. 2010, 104, 036102. 9 Fig. 1. Fig.1 Band structure of double -layer MoS 2 calculated without spin -orbit coupling (a) and with spin -orbit coupling, and the changes of conduction band splitting ( ) at point and valance band splitting at point () and K point (K) following the distance between two layers of double -layer MoS 2, calculated without spin -orbit coupling (c) and with spin -orbit coupling (d). Note the red circle s in Fig. 1c and d represent the data from the equilibrium (or stable) state (ES) and the dot and dash dot lines presen t the conduction band splitting ( ) at point and valance band splitting ( K) at K point of single -layer MoS 2, respectively. 10 Fig.2 Fig. 2 Schematic of valance band splitting of valance band maximum at K point due to the spin -orbit coupling in the each layer (intra -SOC) and layer ’s coupling (LC) in band structure of double -layer MoS 2 (a), schematic structure of double -layer MoS 2 (b), the isosurface of band -decomposed charge density of four states at valance band maximum of K point including the states ~|1 , ~|2, ~|1 and ~|2 shown in Fig. 2a after considering the effects of intra -SOC and LC. 11 Fig. 3 Fig. 3 Band structure of tri ple-layer MoS 2 calculated without spin -orbit coupling (a) and with spin -orbit coupling (b), S chematic of valance band splitting of valance band maximum at K point due to the spin-orbit (SOC) and layer ’s coupling (LC) in band structure of triple -layer MoS 2 (c). Note that in the inset of Fig. 3c, the band structure is plotted with two directions K and KM and the lengths for K and KM are the 1/10 of total lengths in the two directions, respectively. 12 Fig. 4 Fig. 4 Band structure s of double -layer MoS 2 (a) and triple -layer MoS 2 (b) under the pressure of 15 GPa calculated with spin -orbit coupling, and valance bands of triple -layer MoS 2 under 15 GPa near K Point and the related s chematic about band splitting due to the spin -orbit (SOC) and layer ’s couplin g (LC).
0805.1028v1.Frustration_and_entanglement_in_the__t__2g___spin__orbital_model_on_a_triangular_lattice__valence__bond_and_generalized_liquid_states.pdf
arXiv:0805.1028v1 [cond-mat.str-el] 7 May 2008Frustration and entanglement in the t2gspin–orbital model on a triangular lattice: valence–bond and generalized liquid states Bruce Normand D´ epartement de Physique, Universit´ e de Fribourg, CH–170 0 Fribourg, Switzerland Theoretische Physik, ETH–H¨ onggerberg, CH–8093 Z¨ urich, Switzerland Andrzej M. Ole´ s Marian Smoluchowski Institute of Physics, Jagellonian Uni versity, Reymonta 4, PL–30059 Krak´ ow, Poland Max-Planck-Institut f¨ ur Festk¨ orperforschung, Heisenb ergstrasse 1, D-70569 Stuttgart, Germany (Dated: November 1, 2018) We consider the spin–orbital model for a magnetic system wit h singly occupied but triply degen- eratet2gorbitals coupled into a planar, triangular lattice, as woul d be exemplified by NaTiO 2. We investigate the ground states of the model for interactions which interpolate between the limits of pure superexchange and purely direct exchange interaction s. By considering ordered and dimerized states at the mean–field level, and by interpreting the resul ts from exact diagonalization calculations on selected finite systems, we demonstrate that orbital inte ractions are always frustrated, and that orbital correlations are dictated by the spin state, manife sting an intrinsic entanglement of these degrees of freedom. In the absence of Hund coupling, the grou nd state changes from a highly res- onating, dimer–based, symmetry–restored spin and orbital liquid phase, to one based on completely static, spin–singlet valence bonds. The generic propertie s of frustration and entanglement survive even when spins and orbitals are nominally decoupled in the f erromagnetic phases stabilized by a strong Hund coupling. By considering the same model on other lattices, we discuss the extent to which frustration is attributable separately to geometry a nd to interaction effects. PACS numbers: 71.10.Fd, 74.25.Ha, 74.72.-h, 75.30.Et I. INTRODUCTION Frustration in magnetic systems may be of geometri- cal origin, or may arise due to competing exchange in- teractions, or indeed both.1For quantum spins, frustra- tion acts to enhance the effects of quantum fluctuations, leading to a number of different types of magnetically disordered state, among which some of the more familiar are static and resonating valence–bond (VB) phases. A further form of solution in systems with frustrated spin interactionsis the emergenceof novelorderedstates from ahighlydegeneratemanifoldofdisorderedstates,andthe mechanism for their stabilizationhas become known sim- ply as “order–by–disorder”.1,2Many materials are now known whose physical properties could be understood only by employing microscopic models with frustrated spin interactions in which some of these theoretical con- cepts operate. A different and still richer situation occurs in the class of transition–metal oxides or fluorides with partly filled 3dorbitals and near–degeneracy of active orbital degrees of freedom. In undoped systems, large Coulomb interac- tions on the transition–metal ions localize the electrons, and the low–energy physics is that of a Mott (or charge– transfer3) insulator. Their magnetic properties are de- scribed by superexchange spin–orbital models, derived directly from the real electronic structure and contain- ing linearly independent but strongly coupled spin and orbital operators.4Such models emerge from the charge excitations which involve various multiplet states,5,6in which ferromagnetic (FM) and antiferromagnetic (AF)interactions, as well the tendencies towards ferro–orbital (FO) and alternating orbital (AO) order, compete with eachother. This leadsto a profound, intrinsic frustration of spin–orbital exchange interactions, which occurs even in case of only nearest–neighbor interactions for lattices with unfrustrated geometry, such as the square and cu- bic lattices.7The underlying physics is formulated in the Goodenough–Kanamorirules,8which imply that the two types of order are complementary in typical situations: AO order favorsa FM state while FO order coexists with AF spin order. Only recently have exceptions to these rules been noticed,9and the search for such exceptions, and thus for more complex types of spin–orbital order or disorder, have become the topic of much active research. A case study for frustration in coupled spin–orbital systems is provided by the one–dimensional (1D) SU(4) model.10One expects a priori no frustration in one di- mension and with only nearest–neighbor interactions. However, spin and orbital interactions, the latter for- mulated in terms of pseudospin operators, appear on a completelysymmetricalfooting foreverybond, and favor respectively AF and AO ordering tendencies, which com- pete with each other. In fact a low–energy but magnet- ically disordered spin state also frustrates the analogous pseudospin–disordered state, and conversely. This com- petition results in strong, combined spin–orbital quan- tum fluctuations which make it impossible to separate the two subsystems, and it is necessary to treat explicitly entangled spin–pseudospin states.9,11While in one sense this may be considered as a textbook example of frustra- tion and entanglement, the symmetry of the entangled2 sectorsissohighthatjointspin–pseudospinoperatorsare asfundamentalasthe separatespinandpseudospinoper- ators, forming parts of a larger group of elementary (and disentangled) generators. The fact that the 1D SU(4) modelisexactlysolvablealsoresultsinfundamentalsym- metriesbetween the intersitecorrelationfunctions forthe spin and orbital (and spin–orbital) sectors.12We return below to a more detailed discussion of entanglement and its consequences. Although indicative of the rich under- lying physics (indeed, unconventional behavior has been identified for the SU(4) Hamiltonian on the triangular lattice,10,13)theimplicationsofthismodelareratherlim- ited because it does not correspond to the structure of superexchange interactions in real correlated materials. Realistic superexchange models for perovskite transition–metal oxides with orbital degrees of freedom have been known for more than three decades,5,6but the intrinsic frustrating effects of spin–orbital interac- tions have been investigated only in recent years.7,14 A primary reason for this delay was the complexity of the models and the related quantum phenomena, which require advanced theoretical methods beyond a straightforward mean–field theory. The structure of spin–orbital superexchangeinvolves interactionsbetween SU(2)–symmetric spins {/vectorSi,/vectorSj}on two nearest–neighbor transition–metal ions {i,j}, each coupled to orbital op- erators{/vectorTi,/vectorTj}which obey only much lower symmetry (at most cubic for a cubic lattice), and its general form is4 HJ=J/summationdisplay /angbracketleftij/angbracketright/bardblγ/braceleftBig ˆJ(γ) ij/parenleftBig /vectorSi·/vectorSj/parenrightBig +ˆK(γ) ij/bracerightBig .(1.1) The energy scale Jis determined (Sec. II) by the in- teraction terms and effective hopping matrix elements between pairs of directional egorbitals [(ddσ) element] ort2gorbitals [(ddπ) element] The orbital operators ˆJ(γ) ij andˆK(γ) ijspecifytheorbitalsoneachbond /an}bracketle{tij/an}bracketri}ht /bardblγ,which participate in dn idn j⇀↽dn+1 idn−1 jvirtual excitations, and thus have the symmetry of the lattice. The form of the orbital operators depends on the valence n, on the type (egort2g) of the orbitals and, crucially, on the bond di- rectionin realspace.15It is clearfrom Eq.(1.1) that indi- vidualtermsintheHamiltonian HJcanbeminimized for particularly chosen spin and orbital configurations,4but in general the structure of the orbital operators ensures a competition between the different bonds. This directional nature is the microscopic origin of the intrinsic frustration mentioned above, which is present even in the absence of geometrical frustration. Both spin andorbitalinteractionsarefrustrated, makinglong– range order more difficult to realize in either sector, and enhancing the effects of quantum fluctuations. Quite generally, because insufficient (potential) energy is avail- able from spin or orbital order, instead the system is driven to gain (kinetic) energy from resonance processes, promoting phases with short–range dynamical correla- tions and leading naturally to spin and/or orbital dis-order. Disordering tendencies are particularly strong in highlysymmetricsystems, whichforcrystallinematerials means cubic and hexagonal structures. Among possible magnetically disordered phases for spin systems, tenden- cies towards dimer formation are common in the regime of predominantly AF spin interactions, and new phases with VB correlations occur. This type of physics was discussed first for egorbitals on the cubic lattice,7and, in the context of BaVS 3, for one version of the problem oft2gorbitals on a triangular lattice.16The same generic behaviorhassincebeenfoundfor t2gorbitalsonthecubic lattice,17eg–orbitalsystemsonthetriangularlattice,18,19 and fort2gorbitals in the pyrochlore geometry.20,21By analogy with spin liquids, the orbital–liquid phase1has been introduced for systems with both eg7,22andt2g14,23 orbital degrees of freedom. The orbital liquid is a phase in which strongorbital fluctuations restorethe symmetry of the orbital sector, in the sense that the instantaneous orbital state of any site is pure, but the time average is a uniform occupation of all available orbital states. We note that in the discussion of orbital liquids in t2g systems,14,23it was argued that the spin sector would be ordered. To date little is known concerning the behavior oforbitalcorrelationsin an orbitalliquid, the possible in- stabilities of the orbital liquid towards dimerized or VB phases, or its interplay with lattice degrees of freedom. Onepossiblemechanismforthe formationofanorbital liquid state is the positional resonanceof VBs. There has been considerable recent discussion of spin–orbital mod- els in the continuing search for a realistic system real- izing such a resonating VB (RVB) state,19including in a number of the references cited in the previous para- graph. While the RVB state was first proposed for the S=1 2Heisenberg model on a triangular lattice,24exten- sive analysis of spin–only models has not yet revealed a convincing candidate system, although the nearest– neighbor dimer basis has been shown to deliver a very good description of the low–energy sector for the S=1 2 Heisenberg model on a kagome lattice.25To date, the only rigorous proof for RVB states has been obtained in rather idealized quantum dimer models (QDMs),26most notably on the triangular lattice.27The insight gained from this type of study can, however, be used19to for- mulate some qualitative criteria for the emergence of an RVB ground state. These combine energetic and topo- logical requirements, both of which are essential: the energetics of the system must establish a proclivity for dimer formation, a high quasi–degeneracy of basis states in the candidate ground manifold, and additional energy gains from dimer resonance; exact degeneracy between topological sectors (determined by a non–local order pa- rameter related to winding of wave functions around the system) is a prerequisite to remove the competing possi- bility of a “solid” phase with dimer, plaquette or other “crystalline” order.28 We comment here that the “problem” of frustration, and the resulting highly degenerate manifolds of states whichmaypromoteresonancephenomena,isoftensolved3 Na-ionO-ion Ti-ionYZ XYZX (a) (b) FIG. 1: (Color online) Structure of the transition–metal ox ide with edge–sharing octahedra realized for NaTiO 2: (a) frag- ment of crystal structure, with Ti and Na ions shown respec- tively by black and green (grey) circles separated by O ions (open circles); (b) titanium /an}bracketle{t111/an}bracketri}htplane with adjacent oxy- gen layers, showing each Ti3+ion coordinated by six oxygen atoms (open circles). The directions of the Ti–Ti bonds are labeled as XY,YZ, andZX, corresponding to the plane spanned by the connecting Ti–O bonds. This figure is repro- duced from Ref. 33, where it served to explain the structure of LiNiO 2. by interactions with the lattice. Lattice deformations act to lift degeneracies and to stabilize particular patterns of spin and orbital order, the most familiar situation being that in colossal–magnetoresistance manganites.29 The samephysicsis alsodominantin anumberofspinels, where electron–lattice interactions are responsible both for the Verwey transition in magnetite30and fort2gor- bital order below it, as well as for inducing the Peierls state in CuIr 2S4and MgTi 2O4.31Similar phenomena are also expected31to play a role in NaTiO 2. Here, however, we will not introduce a coupling to phonon degrees of freedom, and focus only on purely electronic interactions whose frustration is not quenched by the lattice. The spin–orbital interactions on a triangular lattice are particularly intriguing. This lattice occurs for edge– sharing MO 6octahedra in structures such as NaNiO 2or LiNiO 2, where the consecutive /an}bracketle{t111/an}bracketri}htplanes of Ni3+ions are well separated. These two eg–electron systems be- havequitedifferently: while NaNiO 2undergoesacooper- ativeJahn–Tellerstructuraltransitionfollowedbyamag- netic transition at low temperatures ( TN= 20 K), both transitions are absent in LiNiO 2.32Possible reasons for this remarkable contrast were discussed in Ref. 33, where the authors noted in particular that realistic spin–orbital superexchange neither has an SU(2) ⊗SU(2) structure,18 nor can it ever be reduced only to the consideration of FM spin terms.34These studies showed in addition that LiNiO 2is not a spin–orbital liquid, and that the rea- sons for the observed disordered state are subtle, as spins and orbitals are thought likely to order in a strictly two– dimensional (2D) spin–orbital model.33 Thepossibilitiesofferedforexoticphasesinthistypeof model and geometry motivate the investigation of a real- istic spin–orbital model with active t2gorbitals, focusing first on 3d1electronic configurations. The threefold de-generacyofthe orbitalsismaintained, although, asnoted above, this condition may be hard to maintain in real materials at low temperatures. A material which should exemplify this system is NaTiO 2(Fig. 1), which is com- posed of Ti3+ions int1 2gconfiguration, but has to date had rather limited experimental35,36and theoretical37 attention. Considerably more familiar is the set of tri- angular cobaltates best known for superconductivity in NaxCoO2: here the Co4+ions havet5 2gconfiguration and are expected to be analogous to the d1case by particle– hole symmetry. The effects of doping have recently been removed by the synthesis of the insulating end–member CoO2.38Another system for which the same spin–orbital model could be applied is Sr 2VO4, where the V4+ions occupy the sites of a square lattice.39 The model with hopping processes of pure superex- changetypewasconsideredinthecontextofdopedcobal- tates by Koshibae and Maekawa.40These authors noted that, like the cubic system, two t2gorbitals are active for each bond direction in the triangular lattice, but that thesuperexchangeinteractionsareverydifferentfromthe cubic case because the effective hopping interchanges the activeorbitals. Here we focus onlyon insulating systems, whose entire low–energy physics is described by a spin– orbital model. In addition to superexchange processes mediated by the oxygen ions, on the triangular lattice it is possible to have direct–exchange interactions, which result from charge excitations due to direct d−dhop- ping between those t2gorbitals which do not participate in the superexchange. The ratio of these two types of interaction ( α, defined in Sec. II) is a key parameter of the model. Further, in transition–metal ions4the coef- ficients of the different microscopic processes depend on the Hund exchange JHarising from the multiplet struc- ture of the excited intermediate d2state,41and we intro- duce η=JH U, (1.2) as the second parameter of the model. The aim of this investigation is to establish the general properties of the phase diagram in the ( α,η) plane. We conclude our introductory remarks by returning to the question of entanglement. In the analysis to follow wewillshowthat thepresenceofconflictingorderingten- dencies driven by different components of the frustrated intersite interactions can be related to the entanglement of spin and orbital interactions. By “entanglement” we mean that the correlations in the ground state involve simultaneous fluctuations of the spin and orbital com- ponents of the wave function which cannot be factor- ized. We will introduce an intersite spin–orbital corre- lation function to identify and quantify this type of en- tanglement in different regimes of the phase diagram. It has been shown9that such spin–orbital entanglement is present in cubic titanates or vanadates for small values of the Hund exchange η. Here we will find entanglement to be a generic feature of the model for all exchange in-4 teractions, even in the absence of dimer resonance, and that only the FM regime at sufficiently high η, which is fully factorizable, provides a counterpoint where the entanglement vanishes. The paper is organized as follows. In Sec. II we derive the spin–orbitalmodel for magneticions with the d1elec- tronicconfiguration(Ti3+orV4+)onatriangularlattice. The derivation proceeds from the degenerate Hubbard model, and the resulting Hamiltonian contains both su- perexchange and direct exchange interactions. We begin our analysis of the model, which covers the full range of physical parameters, in Sec. III by considering pat- terns of long–ranged spin and orbital order representa- tive of all competitive possibilities. These states compete with magnetically or orbitally disordered phases domi- nated by VB correlations on the bonds, which are in- vestigated in Sec. IV. The analysis suggests strongly that all long–range order is indeed destabilized by quan- tum fluctuations, leading over much of the phase dia- gram to liquid phases based on fluctuating dimers, with spin correlations of only the shortest range. In Sec. V we present the results of exact diagonalization calcula- tions performed for small clusters with three, four, and six bonds, which reinforce these conclusions and provide detailed information about the local physical processes leading to the dominance of resonating dimer phases. In each of Secs. III, IV, and V, we conclude with a short summary of the primary results, and the reader who is more interested in an overview, rather than in detailed energetic comparisons and actual correlation functions for the different phases, may wish to read only these. Some insight into the competition and collaboration be- tween frustration effects of different origin can be ob- tainedbyvaryingthegeometryofthesystem,andSec.VI discusses the properties of the model on related lattices. A discussion and concluding summary are presented in Sec. VII. II. SPIN–ORBITAL MODEL A. Hubbard model for t2gelectrons We consider the spin–orbital model on the triangular lattice which follows from the degenerate Hubbard–like model fort2gelectrons. It contains the electron kinetic energy and electronic interactions for transition–metal ions arranged on the /an}bracketle{t111/an}bracketri}htplanes of a compound with localcubic symmetryandwith the d1ionicconfiguration, and as such is applicable to Ti3+or V4+[Fig. 1(a)]. The kinetic energy is given by Ht=−/summationdisplay /angbracketleftij/angbracketright/bardblγ,µν,σt(γ) µν/parenleftBig d† iµσdjνσ+d† jνσdiµσ/parenrightBig ,(2.1) whered† iµσare creation operators for an electron with spinσ=↑,↓andorbital“color” µat sitei, and the sum is /0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0 /1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1 /0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0 /1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0 /1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1 /0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0 /1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1 /0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0 /1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1 /0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0 /1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1 /0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0 /1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0 /1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0/0 /1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1/1 (a) c2 (c)3 23 1 1(bc) (ca) (ab)ba (b) FIG. 2: (Color online) (a) Schematic representation of the hopping processes in Eq. (2.1) which contribute to magnetic interactions on a representative bond /an}bracketle{tij/an}bracketri}htalong the c–axis in the triangular lattice. The t2gorbitals are represented by different colors (greyscale intensities). Superexchange p ro- cesses involve O 2 pzorbitals (violet), and couple pairs of a andborbitals (red, green) with effective hopping elements t, interchanging their orbital color. Direct exchange couple sc orbitals (blue) with hopping strength t′. (b) Pairs of t2gor- bitals active in superexchange and (c) single orbitals acti ve in direct exchange; horizontal bonds correspond to the situat ion depicted in panel (a). made over all the bonds /an}bracketle{tij/an}bracketri}ht/bardblγspanning the three direc- tions,γ=a,b,c, of the triangular lattice. This notation isadoptedfromthesituationencounteredinacubicarray of magnetic ions, where only two of the three t2gorbitals areactiveonanyonebond /an}bracketle{tij/an}bracketri}ht, andcontribute t(γ) µνtothe kineticenergy, whilethe third liesinthe planeperpendic- ular to the γaxis and thus hopping processes involving the 2pπoxygen orbitals is forbidden by symmetry.42,43 We introduce the labels a≡yz,b≡xz, andc≡xyalso for the three orbital colors, and in the figures to follow their respective spectral colors will be red, green, and blue. For the triangular lattice formed by the ions on the /an}bracketle{t111/an}bracketri}htplanes of transition–metal oxides (Fig. 1) it is also thecasethatonlytwo t2gorbitalsparticipatein(superex- change)hoppingprocessesviatheoxygensites. However, unlike the cubic lattice, where the orbital color is con- served, here any one active orbital color is exchanged for the other one [Fig. 2(a)]. Using the same convention, that each direction in the triangular lattice is labeled by its inactive orbital color44γ=a,b,c, the hopping ele-5 ments for a bond oriented (for example) along the c–axis in Eq. (2.1) are t(c) ab=t(c) ba=t, whilet(c) aa=t(c) bb= 0. In addition, and also in contrastto the cubic system, for the triangular geometry a direct hopping from one corbital tothe other, i.e.withoutinvolvingtheoxygenorbitals, is also permitted on this bond (Fig. 2), and this element is denoted by t′=t(c) cc. We will also refer to these hopping processes as off–diagonal and diagonal. We stress that while the lattice structure of magnetic ions is triangular, the system under consideration retains local cubic sym- metry of the metal–oxygen octahedra, which is crucial to ensure that the degeneracy of the three t2gorbitals is preserved. The electron–electroninteractionsaredescribedby the on–site terms45 Hint=U/summationdisplay iµniµ↑niµ↓+/parenleftbigg U−5 2JH/parenrightbigg/summationdisplay i,µ<ν,σσ′niµσniνσ′ −2JH/summationdisplay i,µ<ν/vectorSiµ·/vectorSiν+JH/summationdisplay i,µ/negationslash=νd† iµ↑d† iµ↓diν↓diν↑,(2.2) whereUandJHrepresent respectively the intraorbital Coulomb and on–site Hund exchange interactions. Each pair of orbitals {µ,ν}is included only once in the inter- action terms. The Hamiltonian (2.2) describes rigorously the multiplet structureof d2ions within the t2gsubspace, and is rotationally invariant in the orbital space.45 When the Coulomb interaction is large compared with the hopping elements ( U≫t,t′), the system is a Mott insulator with one delectron per site in the t2gorbitals, whence the local constraint in the strongly correlated regime is nia+nib+nic= 1, (2.3) whereniγ=niγ↑+niγ↓. The operators act in the re- stricted space niγ= 0,1. The low–energy Hamiltonian may be obtained by second–order perturbation theory, andconsistsofasuperpositionoftermswhichfollowfrom virtuald1 id1 j⇀↽d2 id0 jexcitations. Because each hopping process may be of either off–diagonal ( t) [Fig. 2(b)] or diagonal (t′) type [Fig. 2(c)], the Hamiltonian consists of several contributions which are proportional to three coupling constants, Js=4t2 U, Jd=4t′2 U, Jm=4tt′ U.(2.4) These represent in turn the superexchange term, the di- rect exchange term, and mixed interactions which arise from one diagonal and one off–diagonal hopping process. We chooseto parameterizethe Hamiltonian by the sin- gle variable α= sin2θ, (2.5) with tanθ=t′ t, (2.6)which gives Js=Jcos2θ,Jm=Jsinθcosθ, andJd= Jsin2θ;Jis the energy unit, which specifies respectively the superexchange ( J=Js) and direct–exchange ( J= Jd) constants in the two limits α= 0 andα= 1. The Hamiltonian H=J/braceleftBig (1−α)Hs+/radicalbig (1−α)αHm+αHd/bracerightBig (2.7) consists of three terms which follow from the processes described by the exchange elements in Eqs. (2.4), each of which contains contributions from both high– and low– spin excitations. B. Superexchange Superexchange contributions to Hcan be expressed in the form Hs=1 2/summationdisplay /angbracketleftij/angbracketright/bardblγ/braceleftBig r1/parenleftBig /vectorSi·/vectorSj+3 4/parenrightBig/bracketleftBig A(γ) ij+1 2(niγ+njγ)−1/bracketrightBig +r2/parenleftBig /vectorSi·/vectorSj−1 4/parenrightBig/bracketleftBig A(γ) ij−1 2(niγ+njγ)+1/bracketrightBig −2 3(r2−r3)/parenleftBig /vectorSi·/vectorSj−1 4/parenrightBig B(γ) ij/bracerightBig , (2.8) where one recognizes a structure similar to that for su- perexchange in cubic vanadates,4,14with separation into aspinprojectionoperatoronthetripletstate,( /vectorSi·/vectorSj+3 4), and an operator ( /vectorSi·/vectorSj−1 4) which is finite only for low–spin excitations. These operators are accompanied by coefficients ( r1,r2,r3) which depend on the Hund ex- change parameter (1.2), and are given from the multiplet structure of d2ions41by r1=1 1−3η, r2=1 1−η, r3=1 1+2η.(2.9) TheCoulombandHundexchangeelementsdeducedfrom the spectroscopic data of Zaanen and Sawatzky46are U= 4.35 eV and JH= 0.59 eV, giving a realistic value ofη≃0.136 for Ti2+ions. For V2+one finds46U= 4.98 eV andJH= 0.64 eV, whence η≃0.13, and the values for V3+ions are expected to be very similar. Finally, for Co3+ions,47U= 6.4 eV andJH= 0.84 eV, giving againη≃0.13. The value η= 0.13 therefore appears to be quite representative for transition–metal oxides with partly filled t2gorbitals, whereas somewhat larger values have been found for systems with active egorbitals due to a stronger Hund exchange.4 The orbital operators AijandBijin Eq. (2.8) depend on the bond direction γand involve two active orbital colors, A(γ) ij=/parenleftBig T+ iγT+ jγ+T− iγT− jγ/parenrightBig −2Tz iγTz jγ+1 2n(γ) in(γ) j,(2.10) B(γ) ij=/parenleftBig T+ iγT− jγ+T− iγT+ jγ/parenrightBig −2Tz iγTz jγ+1 2n(γ) in(γ) j.(2.11)6 For illustration, in the case γ=c(/an}bracketle{tij/an}bracketri}ht /bardblc), the orbitals aandbat siteiare interchanged (off–diagonal hopping) at sitej, and the electron number operator is n(γ) i= nia+nib. The quantity niγin Eq. (2.8) is the number operator for electrons on the site in orbitals inactive for hopping on bond γ,niγ= 1−n(γ) i, ornicin this example. For a single bond, the orbital operators in Eq. (2.10) may be written in a very suggestive form by performing a local transformation in which the active orbitals are exchanged on one bond site, specifically |a/an}bracketri}ht → |b/an}bracketri}htand |b/an}bracketri}ht → |a/an}bracketri}hton bondγ=c.40Then A(γ) ij= 2/parenleftBig /vectorTiγ·/vectorTjγ+1 4n(γ) in(γ) j/parenrightBig , (2.12) B(γ) ij= 2/parenleftBig /vectorTiγ×/vectorTjγ+1 4n(γ) in(γ) j/parenrightBig ,(2.13) where the scalar product in Aijis the conventional expression for pseudospin–1/2 variables, and the cross product in Bijis defined as /vectorTiγ×/vectorTjγ=1 2(T+ iγT+ jγ+T− iγT− jγ)+Tz iγTz jγ.(2.14) Equations (2.8) and (2.12) make it clear that for a sin- gle superexchange bond, the minimal energy is obtained either by forming an orbital singlet, in which case the op- timal spin state is a triplet, or by forming a spin singlet, in which case the preferred orbital state is a triplet; we refer to these bond wavefunctions respectively as (os/st) and (ss/ot). The two states are degenerate for η= 0, while for finite Hund exchange E(os/st)=−Jr1, (2.15) E(ss/ot)=−1 3J(2r2+r3), (2.16) and the (os/st) state is favored. This propensity for sin- glet formation in the α= 0 limit will drive much of the physics to be analyzed in what follows. Becauseoftheoff–diagonalnatureofthehoppingterm, in the original electronic basis (before the local transfor- mation) the orbital singlet is the state |ψos/an}bracketri}ht=1√ 2(|aa/an}bracketri}ht−|bb/an}bracketri}ht), (2.17) while the orbital triplet states are |ψot+/an}bracketri}ht=|ab/an}bracketri}ht, (2.18) |ψot0/an}bracketri}ht=1√ 2(|aa/an}bracketri}ht+|bb/an}bracketri}ht), (2.19) |ψot−/an}bracketri}ht=|ba/an}bracketri}ht. (2.20) The locally transformed basis then gives a clear analogy which can be used for single bonds and dimer phases in combination with all of the understanding gained for the Heisenberg model. However, we stress here that the local transformationfailsforsystemswithmorethan1bondinthe absence of static dimer formation. This arises due to frustration, and can be shown explicitly in numerical cal- culations, but we will not enter into this point in more detail here. However, we take the liberty of retaining the notation of the local transformation, particularly in Sec. IV when considering dimers. Because the transfor- mation interchanges the definitions of FO and AO con- figurations, we will state clearly in each section the basis in which the notation is chosen. C. Direct Exchange The direct exchange part is obtained by considering virtual excitations of active γorbitals on a bond /an}bracketle{tij/an}bracketri}ht /bardblγ, which yield Hd=1 4/summationdisplay /angbracketleftij/angbracketright/bardblγ/braceleftBig/bracketleftBig −r1/parenleftBig /vectorSi·/vectorSj+3 4/parenrightBig +r2/parenleftBig /vectorSi·/vectorSj−1 4/parenrightBig/bracketrightBig ×/bracketleftBig niγ(1−njγ)+(1−niγ)njγ/bracketrightBig +1 3(2r2+r3)/parenleftBig /vectorSi·/vectorSj−1 4/parenrightBig 4niγnjγ/bracerightBig .(2.21) Here there are no orbital operators, but only number operators which select electrons of color γon bonds ori- entedalongthe γ–axis. Whenonlyonlyoneactiveorbital is occupied [ niγ(1−njγ)], this electron can gain energy −1 4Jfrom virtual hopping at η= 0, a number which has only a weak dependence on the bond spin state at η>0. When both active orbitals are occupied ( niγnjγ), placing the two electrons in a spin singlet yields the far lower bond energy −J, and thus again one may expect muchofthe discussionto followtocenterondimer–based states of the extended system. Again the triplet d2spin excitation corresponds to the lowest energy, ( U−3JH), and only the lower two excitations involve spin singlets which could minimize the bond energy. The structure of these terms is the same as in the 1D egspin–orbital model,48or the case of the spinel MgTi 2O4.20A simpli- fied model for the triangular–lattice model in this limit, using a lowest–order expansion in ηfor the spin but not for orbital interactions, was introduced in Ref. 49. D. Mixed Exchange Finally, the twodifferent types ofhopping channelmay also contribute to two–step, virtual d1 id1 j⇀↽d2 id0 jexci- tations with one off–diagonal ( t) and one diagonal ( t′) process. The occupied orbitals are changed at both sites (Fig. 2), and as for the superexchange term the result- ing effective interaction may be expressed in terms of or- bital fluctuation operators. To avoid a more general but complicated notation, we write this term only for c–axis bonds, H(c) m=−1 4/summationdisplay /angbracketleftij/angbracketright/bardblc/bracketleftBig r1/parenleftBig /vectorSi·/vectorSj+3 4/parenrightBig −r2/parenleftBig /vectorSi·/vectorSj−1 4/parenrightBig/bracketrightBig7 ×/parenleftBig T+ iaT+ jb+T− ibT− ja+T+ ibT+ ja+T− iaT− jb/parenrightBig ,(2.22) where the orbital operators are T+ ia=b† ici, T+ ib=c† iai, T− ia=c† ibi, T− ib=a† ici. (2.23) These definitions are selected to correspond to the ↑– pseudospin components of both operators being |bi/an}bracketri}htfor Tz iaand|ci/an}bracketri}htforTz ib. The form of the H(a) mandH(b) mterms is obtained from Eq. (2.22) by a cyclic permutation of the orbital indices. By inspection, this type of term is finite only for bonds whose sites are occupied by linearsuperpositions of different orbital colors, and creates no strong preference for the spin configuration at small η. E. Limit of vanishing Hund exchange In the subsequent sections we will give extensive con- sideration to the model of Eq. (2.7) at η= 0. In this special case the multiplet structure collapses (spin sin- glet and triplet excitations are degenerate), one finds a singlechargeexcitationofenergy U, andtheHamiltonian reduces to the form H(η= 0) =J/summationdisplay /angbracketleftij/angbracketright/bardblγ/braceleftBig (1−α)/bracketleftBig 2/parenleftBig /vectorSi·/vectorSj+1 4/parenrightBig/parenleftBig /vectorTiγ·/vectorTjγ+1 4n(γ) in(γ) j/parenrightBig +1 2(niγ+njγ)−1/bracketrightBig +α/bracketleftBig/parenleftBig /vectorSi·/vectorSj−1 4/parenrightBig niγnjγ−1 4/parenleftBig niγ(1−njγ)+(1−niγ)njγ/parenrightBig/bracketrightBig −1 4/radicalbig α(1−α)/parenleftBig T+ i¯γT+ j˜γ+T− i˜γT− j¯γ+T+ i˜γT+ j¯γ+T− i¯γT− j˜γ/parenrightBig/bracerightBig , (2.24) which depends only on the ratio of superexchange to di- rect exchange (0 ≤α≤1). The first line of Eq. (2.24) makes explicit the fact that the spin and orbital sectors are completely equivalent and symmetrical at α= 0, at least at the level of a single bond. However, we will showthatthisequivalenceisbrokenwhenmorebondsare considered, and no higher symmetry emerges because of the color changes involved for different bond directions, which change the SU(2) orbital subsector. The second line of Eq. (2.24) emphasizes the importance of bond oc- cupation and singlet formation at α= 1 (Sec. IIC). In the third line of Eq. (2.24), the labels ¯ γ/ne}ationslash= ˜γre- fer to the two mixed orbital operators on each bond [Eq. (2.23)]. Orbital fluctuations are the only processes contributing to the mixed terms in this limit, where the spin state of the bond has no effect. We draw the at- tention of the reader to the fact that for the parameter choiceα= 0.5, anelectronofanycoloratanysitehasthe same matrix element to hop in any direction. However, because of the different color changes involved in these processes, again the spin–orbital Hamiltonian does not exhibit ahigher symmetry at this point, a result reflected in the different operator structures of superexchange and direct exchange components. III. LONG–RANGE–ORDERED STATES In this Section we study possible ordered or partially ordered states for the Hamiltonian of Eq. (2.7). As ex- plained in Sec. II, the parameters of the problem arethe ratio of the direct and superexchange interactions, α(2.5), and the strength of the Hund exchange interac- tion,η(1.2). Regarding the latter, we will discuss briefly the transition to ferromagnetic (FM) spin order for in- creasingηin this framework. The first necessary step in any analysis of such an interacting system is to establish the energies of differ- ent (magnetically and orbitally) ordered states. The high connectivity of the triangular–lattice system sug- gests that ordered states will dominate, and claims of more exotic ground states are justifiable only when these are shown to be uncompetitive. The calculations in this Section will be performed for static orbital and spin con- figurations, with the virtualprocessesresponsiblefor(su- per)exchange as the only fluctuations. In the language of the discussion in Sec. I, fully ordered states gain only potential energy at the cost of sacrificing the kinetic (res- onance) energy from fluctuation processes, which we will show in Secs. IV and V is of crucial importance here. A. Possible orbital configurations The results to follow will be obtained by first fixing the orbital configuration, either on every site or on par- ticular bonds, and then computing the spin interaction and optimizing the spin state accordingly. While this is equivalent to the converse, the procedure is more trans- parent and offers more insight into the candidate phases. We limit the number of states to ordered phases with small unit cells, and the orbital states to be considered8 (a) (b) (c) (d) (e) (f) FIG. 3: (Color online) Schematic representation of possibl e orbital states with a single color on each site of the triangu - lar lattice: (a) one–color state; (b) and (c) two inequivale nt two–color states; (d) three–sublattice three–color state ; (e) and (f) two inequivalent three–color states. The latter two configurations are degenerate with similar states where the lines of occupied aandborbitals repeat rather than being staggered along the direction perpendicular to the lines of occupied corbitals. The three–sublattice state (3d) is nonde- generate ( d= 1), states (3a), (3b), and (3e) have degeneracy d= 3, and states (3c) and (3f) have degeneracy d= 6. are enumerated in this subsection. For clarity we adopt the conventionofFig. 2(c) that horizontal( c) bonds have diagonal (direct exchange) hopping of corbitals, which areshowninblue, andoff–diagonal(superexchange)hop- pingprocessesfor aandborbitals[Fig.2(b)], respectively red and green; up–slanting ( a) bonds have diagonal hop- ping foraorbitals and off–diagonal hopping between b andcorbitals; down–slanting ( b) bonds have diagonal hopping for borbitals and off–diagonal hopping between aandcorbitals. All Hamiltonians and energies are func- tions ofαandη, as given by Eqs. (2.7), (2.8), (2.21), and (2.22). To minimize additional notation, they will be quoted in this and in the next section as functions of the single argument α, with implicit η–dependence con- tained in the parameters ( r1,r2,r3). The orbital bond indexγwill also be suppressed here and in Sec. IV. We continue to refer to the orbital type as a “color”, and begin by listing symmetry–inequivalent states where(a) (b) (c) (d) (e) FIG. 4: (Color online) Schematic representation of possibl e orbital configurations with superpositions of (a) two orbit als in a two–color state, (b) three orbitals, (c) two orbitals wi th equal net weight, and (d) and (e) two orbitals with differing net weights of all three orbitals. State (a) has degeneracy d= 3, states (b) and (c) have d= 1, and the degeneracies of states (d) and (e) are d= 6 and d= 3. each site has a unique color. If the same orbital is occu- pied at every site [Fig. 3(a)], the three states with a,b, orcorbitals occupied are physically equivalent (degener- acy isd= 3). When lines of the same occupied orbitals alternate along the perpendicular direction there are two basicpossibilities, whichareshowninFigs.3(b) and3(c). These two–color states differ in their numbers of active superexchange or direct–exchange bonds, which depend on how the monocolored lines are oriented relative to the active hopping direction(s) of the orbital color. There is only one three–color configuration with equal occupa- tions, which is shown in Fig. 3(d). Turning to orbital states with unequal occupations, motivated by the tendency of Hto favor dimer forma- tion in certain limits we extend our considerations to the possibility of a four–site unit cell [Figs. 3(e) and 3(f)]. More elaborate three–color unit cells are not considered. In this case the same state is obtained when the fourth site is occupied by electrons whose orbital color is any of the other three. Again this state, which breaksrotational symmetry, differs depending on its orientation relative to the active hopping axes.9 States involving a superposition of either two or three orbitals at each site can be expected to allow a signif- icantly greater variety of hopping processes. When ei- ther two or three orbital states are partially occupied at each site (we stress that the condition of Eq. (2.3) is always obeyed rigorously), one finds the two uniform states represented in Figs. 4(a) and 4(b). These denote the symmetric wavefunctions |ψ2/an}bracketri}ht= (|φa/an}bracketri}ht+|φb/an}bracketri}ht)/√ 2 and|ψ3/an}bracketri}ht= (|φa/an}bracketri}ht+|φb/an}bracketri}ht+|φc/an}bracketri}ht)/√ 3 at every site, where |φγ/an}bracketri}ht=γ†|0/an}bracketri}ht. The remaining states shown in Fig. 4 in- volveonlytwoorbitalspersite, but with allthreeorbitals partly occupied in the lattice. The average electron den- sitypersiteandperorbitalis1 /3in thestateofFig.4(c), while in Figs. 4(d) and 4(e) it is nc=1 2,na=nb=1 4. The latter two states are neither unique nor (for general interactions) equivalent to each other, and represent two classes of states with respective degeneracies 3 and 6. B. Ordered–state energies: superexchange Before analyzing the different possible ordered states for any of the model parameters, we stress that the spin interactions on a given bond depend strongly on the or- bital occupation of that bond. We begin with the pure superexchange model Hs(2.8), meaning α= 0, for which the question of spin and orbital singlets was addressed in Sec. IIB. Here the spin and orbital scalar products /an}bracketle{t/vectorSi·/vectorSj/an}bracketri}htand/an}bracketle{t/vectorTi·/vectorTj/an}bracketri}htmay take only values consistent with long–range order throughout the system and thus vary between −1/4 and +1/4. For a bond on which both electrons occupy active or- bitals, one has the possibility of either FO or AO states. For the FO state, /an}bracketle{t/vectorTi·/vectorTj/an}bracketri}ht= 1/4 =/an}bracketle{t/vectorTi×/vectorTj/an}bracketri}htand /an}bracketle{tAij/an}bracketri}ht=/an}bracketle{tBij/an}bracketri}ht= 1, whence the terms of Hscan be sepa- rated into the physically transparent form H(FO) 1(0) =1 2Jr1/parenleftbigg1 2/an}bracketle{tniγ+njγ/an}bracketri}ht/parenrightbigg/parenleftbigg /vectorSi·/vectorSj+3 4/parenrightbigg = 0, H(FO) 2(0) =1 2Jr2/parenleftbigg 2−1 2/an}bracketle{tniγ+njγ/an}bracketri}ht/parenrightbigg/parenleftbigg /vectorSi·/vectorSj−1 4/parenrightbigg =Jr2/parenleftbigg /vectorSi·/vectorSj−1 4/parenrightbigg , (3.1) H(FO) 3=1 3J(r3−r2)/parenleftbigg /vectorSi·/vectorSj−1 4/parenrightbigg , specifying a net spin interaction which, because niγ= 0, must be AF if any hopping processes are to occur. In the AO case, /an}bracketle{t/vectorTi·/vectorTj/an}bracketri}ht=−1/4 =/an}bracketle{t/vectorTi×/vectorTj/an}bracketri}htand/an}bracketle{tAij/an}bracketri}ht= /an}bracketle{tBij/an}bracketri}ht= 0, giving H(AO) 1(0) =−1 2Jr1/parenleftbigg /vectorSi·/vectorSj+3 4/parenrightbigg , H(AO) 2(0) =1 2Jr2/parenleftbigg /vectorSi·/vectorSj−1 4/parenrightbigg ,(3.2) H(AO) 3(0) = 0,and the spin interaction is constant at η= 0, with only a weak FM preference emerging at finite η. We remind the reader here that the designations FO and AO continue to be based on the conventional notation22obtained by a local transformation on one bond site, and in the basis of the original orbitals correspond respectively to opposite active orbitals and to equal active orbitals. Cases where only one orbital is active on a bond are by definition AO, but do contribute a finite spin interaction H1 1(0) =−1 4Jr1/parenleftbigg /vectorSi·/vectorSj+3 4/parenrightbigg , H1 2(0) =1 4Jr2/parenleftbigg /vectorSi·/vectorSj−1 4/parenrightbigg , (3.3) H1 3(0) = 0, which again has only a weak FM tendency at η >0. Clearly, when neither electron may hop, the bond does not contribute a finite energy. We begin with the uniform, one–color orbital state of Fig. 3(a), meaning that all bonds are AO by the defini- tion of the previous paragraph. In two directions both electrons are active, while in the third none are. The energy per bond is E(3a) FM(0) =−1 3Jr1. (3.4) and the spin configuration is FM. However, an antifer- romagnetic (AF) spin configuration on the square lattice defined by the active hopping directions has energy E(3a) AF(0) =−1 6J(r1+r2), (3.5) from which one observes that all spin states are degen- erate atη= 0. The ordered spin state spin is then FM for any finite η. We note in passing that the energy per bond for a square lattice would have the significantly lower value −1 2Jfor the same Hsconvention, by which is meant the presence of the constants +3 4and−1 4in Eq. (2.8). This result is a direct reflection of the geo- metrical frustration of the triangular lattice, an issue to which we return in Sec. VI. The state of Fig. 3(b) involves one set of (alternating) AO lines with two active orbitals and two sets of (AO) lines each with one active orbital. All sets of lines favor FM order at finite η, with E(3b) FM(0) =−1 3Jr1. (3.6) Here the square–lattice state which becomes degenerate atη= 0, with E(3b) AF(0) =−1 6J(r1+r2), (3.7) is more accurately described as one with two lines of AF spins andone ofFM spins [Fig. 5(a)], and will be denoted henceforth as AFF.10 (a) (b) FIG. 5: (Color online) Spin configurations minimizing the total energy of the superexchange Hamiltonian Hs(α= 0) for given fixed patterns of orbital order: (a) AFF state for the orbital ordering pattern of Fig. 3(c), showing how the FM lineisselected bythedirection(here b)givingzerofrustration; (b) 60–120◦ordered spin configuration minimizing the total energy for the orbital ordering pattern of Fig. 3(d). The state of Fig. 3(c) involves one set of FO lines with twoactiveorbitals,onesetoflineswithoneactiveorbital, one half set of AO lines with two active orbitals and one half set of inactive lines. The two–active FO lines will favor AF order, while the AO and the one–active lines will favor FM order only at η>0, giving E(3c) AFF(0) =−1 72J(9r1+11r2+4r3) (3.8) from the AFF configuration, but with 2 equivalent di- rections for the FM line. At η= 0 the energy is again −1 3J. BothE(3b) AF(0) andE(3c) AFF(0) can be regarded as the energy of an unfrustrated system, in the sense that the spin order enforced in any one direction by the orbital configuration at no time denies the system the ability to adopt the energy–minimizing configuration in other di- rections. However, at finite ηthe configurations shown in Figs. 3(b) and 3(c) will be penalized relative to the uniform (AO) order of Fig. 3(a) due to the presence of AF bonds. We insert here an important observation: the orbital state of Fig. 3(c) also admits the formation of 1D AF Heisenberg spin chains on the FO ( b–axis) lines. The energy per bond of such a state includes constant inter- chain contributions which are independent of the spin state (/an}bracketle{t/vectorSi·/vectorSj/an}bracketri}ht= 0) on these bonds. Of these interchain bonds, 1/4 are FO with two active orbitals and 1/2 have one active orbital. One finds E(3c) 1D(0) =−1 9Jln2 (2r2+r3)−1 24J(3r1+r2),(3.9) which gives E(3c) 1D(0) =−0.3977Jatη= 0. This energy is significantly lower than that of an ordered magnetic state, a result showing that the kinetic energy gained from resonance processes on the chains is far more signif- icantthan minimalpotential energygainobtainablefrom an ordering of the magnetic moments on the active inter- chain bonds which are active, and thus provides strong evidence in favorof the hypothesis that any orderedstatewill “melt” to a quantum disordered one in this system. We will return to this issue below. For the two–color superposition [Fig. 4(a)], one set of bonds always has two active orbitals, but with equal probability of being FO or AO, while the other two sets of bonds have a 1/4 probability of having two active or- bitals, which are FO, or a 1/2 probability of having one active orbital (and a 1/4 probability of having none). Under these circumstances, the net system Hamiltonian can be expressedby summing overall the possible orbital states, although this is not necessarily a useful exercise when the spin state may not be isotropic. By insert- ing the three most obvious ordered spin states, FM, AF (meaning here the AF state of the triangular lattice with 120◦bond angles and /an}bracketle{t/vectorSi·/vectorSj/an}bracketri}ht=−1 8) and AFF, the can- didate energies are E(4a) FM(0) =−1 6Jr1, (3.10) E(4a) AF(0) =E(4a) AFF(0) =−1 48J(5r1+7r2+2r3). The coincidence for the results for the AF and AFF or- dered states in this case is an accidental degeneracy. The final energy E(4a) AF(F)=−7 24Jatη= 0 shows that both states are compromises, and it is not possible to put all bonds in their optimal spin state simultaneously. This arises because of the presence of two–active FO compo- nents in all three lattice directions, and will emerge as a quite generic feature of superposition states, albeit not one without exceptions. In general there is no compelling reason (given by H for any value of α) to expect that two–color superposi- tions of this type may be favorable. While the 120◦state ofatriangular–latticeantiferromagnetisonecompromise withinaspaceofSU(2)operators,thistypeofsymmetry– breaking is not relevant within the orbital sector, where there are three colors and the two–color subsector of ac- tive orbitals in the α= 0 limit changes as a function of the bond orientation. In the equally weighted three–color state [Fig. 3(d)], all bonds are FO and it is easy to show that 1/3 of them (arranged as isolated triangles) have two active orbitals while the other 2/3 have one active orbital. The two– active bonds favor AF order while the one–active bonds have only a weak preference for FM order at finite η. In this case the problem becomes frustrated and is best resolved by a kind of AF state on the triangular lattice where the strong triangles have 120◦angles and alternat- ing triangles have spins either all pointing in or all point- ing out [Fig. 5(b)]; then 2/3 of the intertriangle bonds have 60◦angles while the other 1/3 have 120◦angles. The energy of this state is E(3d)(0) =−1 144J(19r1+17r2+6r3),(3.11) andE(3d)(0) =−7 24Jatη= 0, a value again inferior to the optimal energy due to the manifest spin frustration.11 In the state of Fig. 3(e), the only AO bonds (1/6 of the total) contain inactive orbitals. Of the remaining bonds, 3/6 have two active FO orbitals (in all three directions) and 2/6 have one active orbital. Once again the system is composed of strongly coupled triangles, but this time in a square array and with strong coupling in their basal direction by one set of two–active FO bonds. Possible competitive spin–ordered states would be AF or AFF, with energies E(3e) AF(0) =−1 96J(5r1+15r2+6r3),(3.12) E(3e) AFF(0) =−1 288J(15r1+35r2+16r3). The lowest energy is obtained for 120◦AF order, with the frustrated value E(3e) AF(0) =−13 48Jforη= 0. For the state in Fig. 3(f) the FO bonds (1/6) and only 1/6 of the AO bonds have two active orbitals, while the other 2/3 of the bonds have one active orbital. In this case E(3f) FM(0) =−1 4Jr1, E(3f) AF(0) =−1 96J(15r1+13r2+2r3),(3.13) E(3f) AFF(0) =−1 192J(36r1+17r2+5r3), leading again to an AF spin state. At η= 0 one has E(3f) AF(0) =−5 16J,i.e.relatively weaker frustration. Turning now to three–color superpositions, the “uni- form” orbital state [Fig. 4(b)] is one in which on every bond there is a probability 2/9 of having two active FO orbitals, 2/9 for two active AO orbitals, 4/9 of one active orbital and 1/9 of no active orbitals. The appropriately weighted bond interaction strengths may be summed to give the net interaction, which for the three spin states considered results in the energies E(4b) FM(0) =−2 9Jr1, E(4b) AF(0) =−1 36J(5r1+5r2+r3),(3.14) E(4b) AFF(0) =−1 81J(12r1+10r2+2r3), andthusthe AFstateislowest, with thevalue E(4b) AF(0) = −11 36Jatη= 0. While this orbital configuration does not attain the minimal energy of −1 3J, it is a close competi- tor: although it involves every bond, the fractional prob- abilities of each being in a two–active state mean that it cannot maximize individual bond contributions. How- ever, we will see in Sec. IIID that state (4b) lies lowest over much of the phase diagram (0 <α<1) as a result of the contributions from mixed terms. For states with unequal site occupations, in Fig. 4(c) one has a situation where on 1/3 of the bonds (arranged in separate triangles) there is a 1/4 probability of twoactive FO orbitals and a 1/2 probability of one active orbital, while on the remaining 2/3 of the bonds there is a 1/4 probability of two active AO orbitals, 1/4 of two active FO orbitals and 1/2 of having one active orbital. On computing the net energies for the three standard spin configurations, one obtains E(4c) FM(0) =−5 24Jr1, E(4c) AF(0) =−1 192J(25r1+25r2+8r3),(3.15) E(4c) AFF(0) =−1 216J(30r1+25r2+8r3), wherethe AF statewith E(4c) AF(0) =−29 96Jisthe lowestat η= 0. However, this state is also manifestly frustrated. In the unequally weighted state of Fig. 4(d), the prob- lemisbestconsideredonceagainaslinesofdifferentbond types. Here 1/6 of the lines have two active orbitals (1/2 FO and 1/2 AO), 1/6 of the lines have probability 1/4 of two active orbitals (AO) and 1/2 of one active orbital, 1/3 of the lines have probability 1/4 of two active FO or- bitals, 1/4oftwoactiveAOorbitalsand1/2ofoneactive orbital, and the remaining 1/3 of the lines have probabil- ity 1/4 of two active orbitals (FO) and 1/2 of one active orbital. The ordered spin states yield the energies E(4d) FM(0) =−5 24Jr1, E(4d) AF(0) =−1 192J(25r1+27r2+6r3),(3.16) E(4d) AFF(0) =−1 144J(21r1+17r2+4r3), whenceitisagaintheAFstate, withasmalldegreeofun- relieved frustration in its energy E(4d) AF(0) =−29 96J, which lies lowest at η= 0. Finally, the state of Fig. 4(e) has the orbital pattern of Fig. 4(d) rotated in such a way that the number of active orbitals in different bond directions is changed. Now 1/3 of the bonds have probabilities 1/4 of two active orbitals (AO) and 1/2 of one active orbital, while the remaining 2/3 have probabilities 1/4 of two active orbitals (FO), 1/4 of two active orbitals (AO) and 1/2 of one active orbital. The ordered–state energies are E(4e) FM(0) =−1 4Jr1, E(4e) AF(0) =−1 96J(15r1+13r2+2r3),(3.17) E(4e) AFF(0) =−1 36J(6r1+5r2+r3), of which the AFF states lies lowest at η= 0, achieving the unfrustrated value E(4e) AFF(0) =−1 3J. That it is pos- sible to obtain this energy in an orbital superposition is because of the absence of FO bond contributions in one direction, which can then be chosen to be FM. The results of this section and the conclusions one may draw from them are summarized in Subsec. IIIE below.12 C. Ordered–state energies: direct exchange In the limit of only direct exchange, the analysis is somewhat simpler. The Hamiltonian is Hdof Eq. (2.21), and in this case a particle on any site is active in only one direction, which leads to the immediate observation thatin astaticorbitalconfigurationit isneverpossibleto have, on average,activeexchangeprocessesonmorethan 2/3 of the bonds. For simplicity we repeat the Hamilto- nianforthe twocasesofAO orderbetweensites, inwhich case by definition at most one of the orbitals is active, and FO order between sites, which is restricted to the case where neighboring sites have the same orbital color and the correct bond orientation. We stress that in this subsection the definitions FO and AO are entirely con- ventional, as the local transformation of Sec. IIB is not relevant at α= 1, and thus the designation FO implies orbitals of the same color, and AO orbitals of different colors. One obtains the expressions H(AO)(1) =1 4J/bracketleftbigg −r1/parenleftbigg /vectorSi·/vectorSj+3 4/parenrightbigg +r2/parenleftbigg /vectorSi·/vectorSj−1 4/parenrightbigg/bracketrightbigg ,(3.18) H(FO)(1) =1 3J(2r2+r3)/parenleftbigg /vectorSi·/vectorSj−1 4/parenrightbigg ,(3.19) which in the η= 0 limit reduce to the forms H(AO)(1) =−1 4J, (3.20) H(FO)(1) =J/parenleftbigg /vectorSi·/vectorSj−1 4/parenrightbigg .(3.21) It is clear (Sec. II) that for a single bond, the most fa- vorablestate is aspin singlet, which would contribute en- ergy−J, but at the possible expense of placing all of the neighboring bonds in suboptimal states. The very strong preference for such singlet bonds means that any mean– field study of the minimal energy is incomplete without the consideration of dimerized (or valence–bond) states (Sec. IV). The analysis of this section can be considered as elucidating the optimal energies to be gained from long–ranged magnetic and orbital order on these bonds, where the optimal energy of any one is −1 2J. Also as noted in Sec. II, any active AO bond gains an exchange energy (−1 4J) simply because it does not prevent one of the two particles from performing virtual hopping pro- cesses, and this we term “avoided blocking”. In the limit of zero Hund exchange, these will give a highly degener- ate manifold of all possible spin states, from which FM states are selected at finite η. We beginagainwith one–colorstateofFig. 3(a), which we denote henceforth as (3a). Only one set of lattice bonds has finite interactions, all FO, and therefore the system behaves as a set of AF Heisenberg spin chains with energy per bond E(3a) AF1D(1) =−1 9Jln2 (2r2+r3),(3.22)whenceE(3a) AF1D(1) =−0.2310Jatη= 0. In state (3b), the FO lines do not correspond to active hopping directions. The remaining two directions then form an AO square lattice with E(3b) FM(1) =−1 6Jr1. (3.23) This can be called a “pure avoided–blocking” energy. The spins are unpolarized at η= 0, where all bond spin states are equivalent, but any finite ηwill select FM or- der (hence the notation). We will see in the remainder of this section that E=−1 6Jis the optimal energy obtain- able by a 2D ordered state in the direct–exchange limit (α= 1), where the net energy is generically higher than atα= 0quitesimplybecausetherearehalfasmanyhop- ping channels. Thus the “melting” of such ordered states into quasi–1D states becomes clear from the outset, and can be understood due to the very low connectivity of the active hopping network on the triangular lattice. In state (3c), one of the FO lines is active, and forms AF Heisenberg spin chains. Electrons in the other FO line areactiveonly in across–chaindirection, wheretheir bonds areAO,and gainavoided–blockingenergy, whence E(3c) AF(1) =−1 12J(2ln2+1) = −0.1988J(3.24) atη= 0. As in the preceding subsection, the coherent state of each Heisenberg chain is not altered by the pres- ence of additional electrons from other chains executing virtual hopping processes onto empty orbitals of individ- ual sites. The spin chains remain uncorrelated and only quasi–long–range–ordereduntil a finite value of η, where FM spin polarization and a long–range–orderedstate are favored. In the two–color superposition (4a), 1/3 of the bonds are inactive, while on the other 2/3 one has probabil- ity 1/4 of two active electrons (FO), 1/2 of one active (AO) and 1/4 of two inactive electrons. In this case, one obtains an effective square lattice on which an AF spin configuration is favored by the FO processes, with E(4a) AF(1) =−1 72J(3r1+7r2+2r3),(3.25) so againE(4a) AF(1) =−1 6Jatη= 0. The uniform three–color state (3d) maximizes AO bonds, but 1/3 of the bonds on the lattice remain in- active. Thus E(3d) FM(1) =−1 6Jr1, (3.26) and Hund exchange will select the FM spin state. The three–color state (3e) has FO lines oriented in their active direction and will, as in state (3c), form Heisenberg chains linked by bonds with AO order. While the geometry of the interchain coupling can differ de- pending on the orbital alignment in the inactive chains, it does not create a frustrated spin configuration and13 the net energy is E(3e) AF(1) =E(3c) AF(1). The state (3f) has only inactive FO lines and so gains only avoided– blocking energy, from 2/3 of the bonds in the system, whenceE(3f) FM(1) =E(3d) FM(1). In the uniform three–color superposition (4b), every bond has probability 1/9 of containing two active elec- trons (FO), 4/9 of one active electron and 4/9 of remain- ing inactive. For the three different ordered spin config- urations considered in Subsec. IIIB the energies are E(4b) FM(1) =−1 9Jr1, E(4b) AF(1) =−1 72J(5r1+5r2+r3),(3.27) E(4b) AFF(1) =−1 81J(6r1+5r2+r3), and one finds the energy E(4b) AF(1) =−11 72Jfor the 1200 AF state at η= 0. The three–color state (4c) is one in which 1/3 of the bonds (arranged on isolated triangles) have probability 1/4 of being in a state with two active electrons and 1/2 of containing one active electron, while on the other 2/3 of the bonds there is simply a 1/2 probability of one active orbital. The respective energies are E(4c) FM(1) =−1 8Jr1, E(4c) AF(1) =−1 192J(15r1+13r2+2r3),(3.28) E(4c) AFF(1) =−1 216J(18r1+13r2+2r3). Atη= 0, the energy E(4c) FM(1) =−5 32Jis minimized by a 120◦state on the triangles, which are also isolated magnetically in this limit. Finite values of ηresult in FM interactions between the triangles, and a frustrated problem in the spin sectorwhich by inspection is resolved in favor of a net FM configuration only at large η(η > 0.23). Finally, the three–color states (4d) and (4e) yield two possibilities in the α= 1 limit, namely where one of the minority colors is aligned with its active direction and where neither is. In the former case, E(4d) FM(1) =−5 48Jr1, E(4d) AF(1) =−1 384J(25r1+27r2+6r3),(3.29) E(4d) AFF(1) =−1 96J(7r1+7r2+2r3), and the lowest energy E(4d) AFF(1) =−1 6Jatη= 0 is given by the directionally anisotropic AFF spin configuration. This is because 1/2 of the lines, in two of the three di- rections, have some AF preference from their 1/4 prob- ability of containing two active orbitals, while the third direction has no preference at η= 0, and in any case favors FM spins at η >0. In the latter case, the onlyAF tendencies arise along lines in a single direction, but avoided–blocking energy is sufficient to exclude the pos- sibility of a Heisenberg chain state. Here E(4e) FM(1) =−1 8Jr1, E(4e) AF(1) =−1 192J(15r1+13r2+2r3),(3.30) E(4e) AFF(1) =−1 72J(6r1+5r2+r3), whenceE(4e) AFF(1) =−1 6Jatη= 0, in fact with two de- generate possibilities for the orientation of the FM line. D. Ordered–state energies: α= 0.5 Toillustratethepropertiesofthemodelinthepresence of finite direct and superexchange contributions, i.e.at intermediate values of α, we consider the point α= 0.5. As shown in Sec. II, there is no special symmetry at this point, because the contributions from diagonal and off– diagonal hopping remain intrinsically different. States with long–ranged orbital (and spin) order at α= 0.5 are mostly very easy to characterize, because all virtual processes, of both types, allowed by the given configura- tion are able to contribute in full to the net energy. For the many of the states considered in this section, the en- ergetic calculation for α= 0.5 is merely an exercise in adding the α= 0 andα= 1 results with equal weight. Exceptions occur for superposition states gaining energy from processes contained in Hm[Eq. (2.22)], and are in fact decisive here. Because these terms involve explicitly a finite density of orbitals of all three colors on the bond in question, with the active diagonalcolor representedon both sites, only for states (4b), (4c), and (4d), but not (4e) [Figs. 4(b–e)], will it be necessary to consider this contribution. For state (3a), in two directions both electrons are active by off–diagonal hopping, while in the third both may hop diagonally. Diagonal hopping favorsan AF spin configuration, while the off–diagonalhopping bonds have only a weak preference (by Hund exchange) for FM or- der. The ordered–state spin solution is then a doubly degenerate AFF state with energy per bond E(3a)(0.5) =−1 72J(9r1+7r2+2r3),(3.31) givingE(3a)(0.5) =−1 4Jatη= 0. We remind the reader that the prefactor of the superexchange and di- rect exchange contributions is only half as large as in Subsecs. IIIB and IIIC [Eq. (2.7)], so the overall effect of additional hopping processes in this state is in fact an unfrustrated energy summation. We also comment that, exactly at η= 0, there is no obvious preference for any magnetic order between the diagonal–hopping chains. Only at unrealistically large values of ηwould14 the system sacrifice this diagonal–hopping energy to es- tablish a square–lattice FM state. At finite η, the one– colororbitalstate representsa compromisebetween com- peting spin states preferred by the two types of hopping contribution. State (3b) has no diagonal–hopping chains, and these processes therefore enforce only a weak preference for a FMsquarelattice. Becausetheoff–diagonalhoppingpro- cesses also favor FM order at finite η(Subsec. IIIB), the two types of contribution cooperate and one obtains E(3b)(0.5) =−1 4Jr1. (3.32) State (3c) contains one half set of diagonal–hopping chains, which fall along one of the directions which in the spin state favored by the off–diagonal hopping pro- cesses could be FM or AF; this degeneracy will therefore be broken. The other half set of chains will gain only avoided–blocking energy from diagonal processes, which will take place in the FM direction and thus cause no frustration even at finite η. One obtains E(3c)(0.5) =−1 144J(3r1+7r2+2r3),(3.33) and thusE(3c)(0.5) =−1 4Jatη= 0 from this AFF configuration. The additive contributions from superex- change and direct exchange remove the possibility that Heisenberg–chain states in either of the directions fa- vored separately by off–diagonal (Sec. IIIB) or diagonal (Sec. IIIC) hopping could result in an overall lowering of energy. As in Subsec. IIIC, in the two–colorsuperposition (4a) the diagonal hopping processes are optimized by an AFF spin configuration. Although this is one of the degener- ate states minimizing the off–diagonal Hamiltonian, the directions of the FM lines do not match. Insertion of the four possible spin states yields E(4a) FM(0.5) =−1 8Jr1, E(4a) AF(0.5) =−1 96J(8r1+10r2+3r3), E(4a) AFF(0)(0.5) =−1 72J(6r1+7r2+r3),(3.34) E(4a) AFF(1)(0.5) =−1 144J(12r1+14r2+3r3), whence the lowest final energy is E(4a) AF(0.5) =−7 32Jat η= 0. As noted in the previous sections for this spin configuration, the optimal energy for all bonds is not attainable within the off–diagonal hopping sector, and the addition of the (small) diagonal–hopping contribu- tion causes little overall change. The equally weighted three–color state (3d) has no lines of diagonal–hopping bonds, and in fact these con- tribute only avoided–blocking energy on the bonds be- tween the strong triangles defined by the off–diagonalproblem, adding to the weak propensity for FM intertri- angle bonds arising only from the Hund exchange. The diagonal processes can be taken only to alter this energy, and not to promote any tendency towards an alteration of the spin state, whose energy is then E(3d)(0.5) =−1 144J(19r1+11r2+3r3),(3.35) withE(3d)(0.5) =−11 48Jatη= 0. State (3e) is already frustrated in the off–diagonal sec- tor, and diagonal–hoppingprocessescontributeprimarily on otherwise inactive bonds without changing the frus- tration conditions. For the two candidate spin configu- rations E(3e) AF(0.5) =−1 96J(5r1+12r2+4r3), E(3e) AFF(0.5) =−1 192J(11r1+19r2+8r3),(3.36) a competition won by the 120◦AF–ordered state with E(3e) AF(0.5) =−7 32Jatη= 0. State (3f) lacks active lines of diagonal–hopping pro- cesses, and thus the avoided–blocking energy may be added simply to the results for the off–diagonal sector, giving E(3f) FM(0.5) =−5 24Jr1, E(3f) AF(0.5) =−1 192J(25r1+19r2+2r3),(3.37) E(3f) AFF(0.5) =−5 384J(12r1+5r2+r3), or a minimum of E(3f) AF(0.5) =−23 96Jatη= 0. In the uniform three–color superposition (4b), on ev- ery bond there is a probability 4/9 of having only off– diagonal hopping processes, 2/9 for 2 active FO orbitals and 2/9 for two active AO orbitals, a probability 1/9 of having only diagonal hopping processes, and a probabil- ity 4/9 of other processes. These last include the contri- butions fromoneactivediagonaloroff–diagonalelectron, and mixed processes contained in the Hamiltonian Hm (2.22); none of these three possibilities favors any given bond spin configuration other than a FM orientation at finiteη. The net energy contributions are E(4b) FM(0.5) =−2 9Jr1, E(4b) AF(0.5) =−1 144J(20r1+18r2+3r3),(3.38) E(4b) AFF(0.5) =−1 54J(8r1+6r2+r3), and thus the AF state is lowest, with E(4b) AF(0.5) =−41 144J atη= 0. While this energy differs from that for the AFF spin configuration by only1 144J, its crucial property is that it lies below the value −1 4Jobtained by direct sum- mation of the superexchangeand direct–exchangecontri- butions.15 For this orbital configuration, all three spin states gain a net energy of −1 18Jatη= 0 from mixed processes, and these are sufficient, as we shall see, to reduce the otherwisepartiallyfrustratedordered–stateenergytothe global minimum for this value of α. By a small extension of the calculation, the energy of the 1200AF spin state may be deduced at η= 0 for all values of α, and is given by E(4b) AF(α) =−1 72J/parenleftbig 22−11α+8/radicalbig α(1−α)/parenrightbig .(3.39) Comparison with the value obtained by direct summa- tion,E=−1 6(2−α), revealsthat state(4b) isthe lowest– lying fully spin and orbitally ordered configuration in the region0.063<α<0.983. Thatthis statedominatesover the majorityofthe phasediagramis a directconsequence of its ability to gain energy from mixed processes. The non–uniform three–colorstate (4c) also presents a delicate competition between spin configurations of very similar energies. From the preceding subsections, it is clearthat inthis casediagonalandoff–diagonalprocesses favor different ground states, while there will also be a mixed contribution from 1/3 of the bonds. The energies of the three standard spin configurations are E(4c) FM(0.5) =−1 12Jr1, E(4c) AF(0.5) =−1 384J(45r1+41r2+10r3),(3.40) E(4c) AFF(0.5) =−1 432J(54r1+41r2+10r3), where the AF state, obtaining E(4c) AF(0.5) =−1 4Jis the lowest atη= 0. Finally, in the three–color states (4d) and (4e), which are composed of lines of two–colorsites, this delicate bal- ance between different spin configurations persists. For configuration (4d), an AFF state with the same orien- tation of the FM line (along the b–axis) is both favored by diagonal hopping processes and competitive for off– diagonal processes. With inclusion of a small contribu- tion due to mixed processes, the three orderedspin states have energies E(4d) FM(0.5) =−51 288Jr1, E(4d) AF(0.5) =−1 768J(85r1+84r2+18r3),(3.41) E(4d) AFF(0.5) =−1 576J(71r1+59r2+14r3), from which the AFF state minimizes the energy at η= 0 withE(4d) AFF(0.5) =−1 4J. For state (4e), which has no mixed contribution, the orientationsofthe FM lines in the optimal AFF states do not match, and it is necessary, as above, to consider both possibilities when performing a full comparison. These four ordered spin states yield the energies E(4e) FM(0.5) =−3 16Jr1,E(4e) AF(0.5) =−1 128J(15r1+13r2+2r3), E(4e) AFF(0)(0.5) =−1 144J(18r1+13r2+2r3), E(4e) AFF(1)(0.5) =−1 48J(6r1+4r2+r3),(3.42) among which the AF state in fact lies lowest at η= 0, achievingthe weaklyfrustratedvalue E(4e) AF(0.5) =−15 64J. E. Summary Here we summarize the results of this section in a con- cise form. For the superexchange model ( α= 0), a con- siderable number of 2D ordered orbital and spin states exist which return the energy −1 3Jatη= 0. This de- generacy is lifted at any finite Hund exchange in favor of orbital states [(3a), (3b)] permitting a fully FM spin alignment. Most other orbital configurations introduce a frustration in the spin sector at small η, while some offer the possibility of a change of ground–state spin configu- ration at finite η, wherer1exceeds the r2andr3contri- butions and begins to favor states with more FM bonds. However, the value E=−1 3Jper bond remains a rather poor minimum for a system as highly connected as the triangular lattice, even if, as in the superexchange limit, active hopping channels exist only in two of the three lattice directions for each orbital color. Indeed, the limitations of the available ordering (potential) en- ergy are clearly visible from the fact that a significantly lower overall energy is attained in systems which aban- don spin order in favor of the resonance (kinetic) en- ergy gains available in one lattice direction. The result E(3c) 1D(0) =−0.3977Jis the single most important ob- tained in this section, and in a sense obviates all of the considerations made here for fully ordered states, man- dating the full consideration of 2D magnetically and or- bitally disordered phases. In the study of ordered states, it becomes clear that the Hund exchange acts to favor FM spin alignments at highη. Because the “low–spin” states of minimal energy are in fact stabilized by quantum corrections due to AF spin fluctuations, the lowest energies at η= 0 are never obtained for FM states, and therefore increasing ηdrives a phase transition between states of differing spin and orbital order. We show in Fig. 6 the transitions from quasi–1D AF–correlated states at low η, for bothα= 0 andα= 1, to FM states of fixed orbital and spin order (3b). The transitions occur at the values ηc(0) = 0.085 andηc(1) = 0.097, indicating that FM ordered states may well compete in the physical parameter regime. We note again that the energies in the superexchange limit are lower by approximately a factor of two compared to the direct–exchange limit simply because of the number of available hopping channels. Wenotealsothatthereisneverasituationinwhichthe spin Hamiltonian becomes that of a Heisenberg model on16 0 0.05 0.1 0.15 0.2 η−0.8−0.6−0.4−0.2EAF/J, EFM/J FIG. 6: (Color online) Minimum energies per bond obtained for orbitally ordered phases, showing a transition as a func - tion of Hund exchange ηfrom quasi–1D, AF–correlated to FM ordered spin states. For the superexchange Hamiltonian H∫of Sec. II ( α= 0), the transition is from the quasi–1D spin state on orbital configuration (3c) [black, dashed line from Eq. (3.9)] to the one–color orbital state (3a) [red, sol id line from Eq. (3.4)]. For the direct–exchange Hamiltonian Hd(α= 1), the transition is from the purely 1d spin state on the one–color orbital state (3a) [green, dot–dashed line from Eq. (3.22)] to the two–colour, avoided–blocking state (3b) [blue, dotted line from Eq. (3.23)]. The transitions to FM order as obtained from the mean–field considerations of this section are marked by arrows. a triangular lattice. This demonstrates again the inher- entfrustrationintroducedbytheorbitalsector. However, the fact that the ordered–state energy can never be low- ered to the value EHAF=−3 8J, which might be expected for a two–active FO situation on every bond, far less the value−1 2Jwhich could be achieved if it were possible to optimize every bond in some ordered configuration, can be taken as a qualitative reflection of the fact that on the triangular lattice the orbital degeneracy “enhances” rather than relieves the (geometrical) frustration of su- perexchange interactions (Sec. VI). The limit of direct exchange ( α= 1) is found to be quite different: the very strong tendency to favor spin singlet states, and the inherent one–dimensionalityof the model in this limit (one active hopping direction per or- bital color), combine to yield no competitive states with long–ranged magnetic order. Their optimal energy is very poor because of the restricted number of hopping channels, and coincides with the (“avoided–blocking”) value for the model with only AO bonds, E=−1 6J. Thus these states form part of a manifold with very high degeneracy. However, even at this level it is clear that more energy, meaning kinetic (from resonance pro- cesses) rather than potential, may be gained by forming quasi–1D Heisenberg–chain states with little or no inter- chain coupling and only quasi–long–ranged magnetic or- der. Studies of orbital configurations permitting dimer- ized states are clearly required (Sec. IV). Finite Hund exchange acts to favor ordered FM configurations, whichwill take over from chain–like states at sufficiently high values ofη(Fig. 6). Finally, orderedstatesofthemixedmodelshowanum- ber of compromises. At α= 0.5, where the coefficients of superexchange and direct–exchange are equal, some con- figurations are able to return the unfrustrated sum of the optimal states in each sector when considered separately, namely−1 4J. However, superposition states, which are not optimal in either limit, can redeem enough energy from mixed processes to surpass this value, and in fact the maximally superposed configuration (4b) is found to minimize the energy over the bulk of the phase di- agram. Still, the net energy of such states remains small compared to expectations for a highly connected state with three available hopping channels per orbital color. Because of the directional mismatch between the diago- nal and off–diagonal hopping sectors, no quasi–1D states with only chain–like correlations are able to lower the ordered–state energy in the intermediate regime. IV. DIMER STATES As shown in Sec. II, the spin–orbital model on a sin- gle bond favors spin or orbital dimer formation in the superexchange limit, and spin dimer formation in the direct–exchange limit. The physical mechanism respon- sible for this behavior is, as always, the fluctuation en- ergy gain from the highly symmetric singlet state. On the basis of this result, combined with our failure to find any stable, energetically competitive states with long– ranged spin and orbital order in either limit of the model (Sec. III), we proceed to examine states based on dimers. Given the high connectivity of the triangular lattice, dimer–based states are not expected a priorito be capa- ble of attaining lower energies than ordered ones, and if found to be true it would be a consequence of the high frustration, which as noted in Sec. I has its ori- gin in both the interactions and the geometry. Here we consider static dimer coverings of the lattice, and com- pute the energies they gain due to inter–singlet corre- lations. The tendency towards the formation of singlet dimerstates will be supportedby the numericalresultsin Sec. V, which will also address the question of resonant dimer states. A. Superexchange model Motivated by the fact that the spin and orbital sectors inHs(2.8) are not symmetrical, we proceed with a sim- ple decoupling of spin and orbital operators. Extensive research on spin–orbital models has shown that this pro- cedure is unlikely to capture the majority of the physical processes contributing to the final energy, particularly in the vicinity of highly symmetric points of the general Hamiltonian. The results to follow are therefore to be treated as a preliminary guide, and a basis from which17 to consider a more accurate calculation of the missing energetic contributions. We remind the reader that the notation FO and AO used in this subsection is again that obtained by performing a local transformation on one site of every dimer. As noted in Sec. IIB, this procedure is valid for the discussion of states based on individual dimerized bonds, where it represents merely a notational convenience. For FO configurations, which in the origi- nal basis have different orbital colors, one might in prin- ciple expect that, because of the color degeneracy, there should be more ways to realize these without frustration than thereareto realizeAF spinconfigurations; however, because of the directional dependence of the hopping, we will find that this is not necessarily the case (below). Thebasicpremiseofthespin–orbitaldecouplingisthat if the spin (orbital) degrees of freedom on a dimer bond form a singlet state, their expectation value /an}bracketle{t/vectorSi·/vectorSj/an}bracketri}ht (/an}bracketle{t/vectorTi·/vectorTj/an}bracketri}ht) on the neighboring interdimer bonds will be precisely zero. The optimal orbital (spin) state of the interdimer bond may then be deduced from the effec- tive bond Hamiltonian obtained by decoupling. Because Hsdepends on the number of electrons on the sites of a given bond which are in active orbitals, and this num- ber is well defined only for the dimer bonds, the effective Hamiltonian will be obtained by averaging over all occu- pation probabilities. In contrast to the pure Heisenberg spin Hamiltonian, here the interdimer bonds contribute with finite energies, and the dimer distribution must be optimized. A systematic optimization will not be per- formed in this section, where we consider only represen- tative dimer coverings giving the semi–quantitative level ofinsightrequiredasapreludetoaddingdimerresonance processes (Sec. V). On the triangular lattice there are three essentially different types of interdimer bond, which are shown in Fig. 7). For a “linear” configuration [Fig. 7(a)], the num- ber of electrons in active orbitals on the interdimer bond is two; for the 8 possible configurations where one dimer bond is aligned with the interdimer bond under consider- ation [Fig. 7(b)], the number is one on the corresponding site and one or zero with equal probability on the other; for the 14 remaining configurations where neither dimer bond is aligned with the interdimer bond [Fig. 7(c)], the number is one or zero for both sites. The number of elec- trons in active orbitals is then two for type (7a), two or one, each with probability 1/2, for type (7b), and two, one or zero with probabilities 1/4, 1/2, and 1/4 for type (7c). The effective interdimer interactions for each type of bond can be deduced in a manner similar to the treat- ment of the previous section. Considering first the situa- tion for a bond of type (7a) with (os/st) dimers, setting /an}bracketle{t/vectorTi·/vectorTj/an}bracketri}ht= 0 yields one high–spin and two low–spin terms which contribute H(os,7a) 1(0) =−1 4Jr1/parenleftBig /vectorSi·/vectorSj+3 4/parenrightBig , H(os,7a) 2(0) =3 4Jr2/parenleftBig /vectorSi·/vectorSj−1 4/parenrightBig , (4.1)(a) (b) (c) FIG. 7: (Color online) Types of interdimer bond differing in effective interaction due to dimer coordination: (a) “linea r”, (b) “semi–linear”, (c) “non–linear”. H(os,7a) 3(0) =1 6J(r3−r2)/parenleftBig /vectorSi·/vectorSj−1 4/parenrightBig . ClearlyH(os,7a) 1favors FM (high–spin) interdimer spin configurations with coefficient1 4, whileH(os,7a) 2and H(os,7a) 3favor AF (low–spin) configurations with coeffi- cient3 8(both atη= 0). Because r1exceedsr2andr3 when Hund exchangeis finite, one expects a criticalvalue ofηwhere FM configurations will be favored. Simple al- gebraic manipulations using all three terms suggest that this value, which should be relevant for a linear chain of (os/st)dimers, is ηc=1 8. In the limit η→0, the effective bond Hamiltonian simplifies to H(os,7a) eff(0) =1 2J/parenleftbigg /vectorSi·/vectorSj−3 4/parenrightbigg .(4.2) For a bond of type (7a) with (ss/ot) dimers, setting /an}bracketle{t/vectorSi·/vectorSj/an}bracketri}ht= 0 on the interdimer bond yields H(ss,7a) 1(0) =3 4Jr1/parenleftbigg /vectorTi·/vectorTj−1 4/parenrightbigg , H(ss,7a) 2(0) =−1 4Jr2/parenleftbigg /vectorTi·/vectorTj+3 4/parenrightbigg , (4.3) H(ss,7a) 3(0) =−1 6J(r3−r2)/parenleftbigg /vectorTi×/vectorTj+1 4/parenrightbigg . HereH(ss,7a) 1favorsAO configurations with coefficient3 8, whileH(ss,7a) 2andH(ss,7a) 3both favor FO configurations with coefficient1 4(atη= 0). Over the relevant range of Hund exchangecoupling, 0 <η<1/3, there is no change in sign and AO configurations are always favored. The effective bond Hamiltonian for η→0 is H(ss,7a) eff(0) =1 2J/parenleftbigg /vectorTi·/vectorTj−3 4/parenrightbigg .(4.4) For bonds of type (7b), when only one electron occu- pies an active orbital the corresponding decoupled inter- dimer bond Hamiltonians are, for (os/st) dimers, H(os,1) 1(0) =−1 4Jr1/parenleftBig /vectorSi·/vectorSj+3 4/parenrightBig ,18 H(os,1) 2(0) =1 4Jr2/parenleftBig /vectorSi·/vectorSj−1 4/parenrightBig ,(4.5) H(os,1) 3(0) = 0. The final interdimer interaction is obtained by averag- ing over these expressions and those (4.1) for two active orbitalsper bond, and takes the rathercumbersome form H(os,7b) eff(0) =1 12J(r3+5r2−3r1)/vectorSi·/vectorSj −1 48J(9r1+5r2+r3),(4.6) which reduces in the limit η→0 to H(os,7b) eff(0) =1 4J/parenleftbigg /vectorSi·/vectorSj−5 4/parenrightbigg .(4.7) For (ss/ot) dimers, the situation cannot be formulated analogously, because if only one electron on the bond is active, the orbital state of the other electron has no influence on the hopping process, i.e./vectorTi·/vectorTjis not a meaningful quantity. The resulting expressions lead then to H(ss,7b) eff(0) =1 8J(3r1−r2)/vectorTi·/vectorTj−1 12J(r3−r2)/vectorTi×/vectorTj −1 48J(9r1+5r2+r3), (4.8) which has the η→0 limit H(ss,7b) eff(0) =1 4J/parenleftbigg /vectorTi·/vectorTj−5 4/parenrightbigg .(4.9) Finally, forabondoftype(7c), thereisnocontribution from interdimer bond states with no electrons in active orbitals, so the above results [(4.1, 4.5) and (4.3, 4.8)] are already sufficient to perform the necessary averaging. With (os/st) dimers H(os,7c) eff(0) =1 48J(2r3+13r2−9r1)/vectorSi·/vectorSj −1 192J(27r1+13r2+2r3),(4.10) which reduces in the limit η→0 to H(os,7c) eff(0) =1 8J/parenleftbigg /vectorSi·/vectorSj−7 4/parenrightbigg ,(4.11) while for (ss/ot) dimers, H(ss,7c) eff(0) =1 16J(3r1−r2)/vectorTi·/vectorTj−1 24J(r3−r2)/vectorTi×/vectorTj −1 192J(27r1+13r2+2r3),(4.12) which in the η→0 limit gives H(ss,7c) eff(0) =1 8J/parenleftbigg /vectorTi·/vectorTj−7 4/parenrightbigg .(4.13)These results have clear implications for the nearest– neighbor correlations in an extended system. By inspec- tion, systems composed of either type of dimer would favor AF (spin) and AO interdimer bonds, to the extent allowed by frustration, and “linear” [type (7a)] bonds over “semi–linear” [type (7b)] bonds over “non–linear” [type (7c)] bond types in Fig. 7, to the extent allowed by geometry. Discussion of this type of state requires in principletheconsiderationofallpossibledimercoverings, but will be restricted here to a small number of periodic arrays which illustrate much of the essential physics of extended dimer systems within this model. We begin by considering the periodic covering of Fig. 8(a), a fully linear conformation (of ground–state degeneracy 12) whose interdimer bond types (Table I) maximize the possible number of bonds of type (7a). The counterpoint shown in Fig. 8(b) consists of pairs of dimer bonds with alternating orientations in two of the three lattice directions, and constitutes the simplest configuration minimizing (to zero) the number of type– (7a) interdimer bonds. The coverings in Figs. 8(c) and (d) have the same property. These configurations exem- plify a quite general result, that any dimer covering in which there are no linear configurations [type (7a)] of any pair of dimers will have 1/3 type–(7b) bonds, and thus the remaining 1/2 of the bonds must be of type (7c). The coverings shown in Figs. 8(a) and (b, c, d) represent the limiting cases on numbers of each type of bond, in that any random dimer covering will have val- ues between these. Indeed, it is straightforward to argue that, in changes of position of any set of dimers within a covering, the creation of any two bonds of type (7b) will destroy one of type (7a) and one of type (7c), and conversely. Havingestablishedthiseffectivesumrule, weturnnext to the energies of the dimer configurations. First, for both types of dimer [(os/st) and (ss/ot)], all states with equal numbers of each bond type are degenerate, sub- ject to equal solutions of the frustration problem. Next, if frustration is neglected, it is clear from Eqs. (4.2,4.4), (4.7,4.9), and (4.11,4.13), that the AF and AO energy values for the three bond types (obtained by substitut- ing−1 4for/vectorSi·/vectorSjand/vectorTi·/vectorTj) are respectively −1 2J,−3 8J and−1 4J, which, when taken together with the sum rule, suggest a very large degeneracy of dimer covering ener- gies. Returning to the question of frustration, a covering of minimal energy is one which both minimizes the number TABLE I: Occurrence probabilities for bonds of each type for four simple periodic dimer coverings of the triangular latt ice. configuration dimer bond (7a) bond (7b) bond (7c) Fig. 8(a)1 61 602 3 Fig. 8(b)1 601 31 2 Fig. 8(c)1 601 31 2 Fig. 8(d)1 601 31 219 (a) (b) (c) (d) FIG. 8: (Color online) Periodic dimer coverings on the trian - gular lattice, each representative of a class of coverings: (a) linear; (b) plaquette; (c) 12–site unit cell; (d) “zig–zag” . of FM or FO bonds, and ensures that they fall on bonds of type (7c); both criteria are equally important. For the dimer covering (8a), with maximal aligned bonds, it is possible by using the spin (for (os/st) dimers) or orbital (for (ss/ot) dimers) configuration represented by the arrows in Fig. 9(a) to make the number of frustrated (FM/FO) interdimer bonds equal to 1/6 of the total. Bearing in mind that the 1/6 of bonds covered by dimers are also FM/FO, and that at least 1/3 of bonds on the triangular lattice must be frustrated for collinear spins, this number is an absolute minimum. [Here we do not considerthe possibilityof non–collinearorderof the non– singlet degree of freedom.] Further, for this configuration one observes that all of the FM/FO bonds already fall on bondsoftype(7c), providinganoptimalcasewith energy Edim(0) =−J/parenleftbigg1 6+1 6·1 2+1 2·1 4+1 6·3 16/parenrightbigg =−13 32J (4.14) atη= 0. This value constitutes a basic bound which demonstrates that a simple, static dimer covering has lowerenergythan any long–range–orderedspin ororbital state discussed in Sec. III in this limit ( α= 0) of the model. It remains to establish the degeneracy of the ground– state manifold of such coverings, and we provide only a qualitative discussion using further examples. If al- ternate four–site (dimer pair) clusters in Fig. 8(a) are rotated to give the covering of Fig. 8(b), the minimal frustration is spoiled: by analogy with Fig. 9, it is easy to show that, if only 1/6 of the bonds are to be frus-(a) (b) FIG. 9: (Color online) Spin or orbital configurations (black arrows) within (a) linear and (b) zig–zag orbital– or spin– singlet dimer coverings of the triangular lattice. The numb er of frustrated interdimer bonds is reduced to 1/6 of the total , and all are of type (7c). This figure emphasizes that for the spin–orbital model, dimer singlet formation does not exhau st the available degrees of freedom. trated, then they are of type (7b), and otherwise 1/3 of the bonds are frustrated if all are to be of type (7c). On the periodic 12–site cluster [Fig. 8(c)], one may place three four–site clusters in each of the possible orienta- tions, which as above removes all bonds of type (7a) and maximizes those of type (7b). Within this cluster it is possible to have only four frustrated interdimer bonds out of 18, while between the clusters there is again an arrangement of the spin or orbital arrows ( cf.Fig. 9) with only six FM or FO bonds out of 24, for a net total of 1/6 frustrated interdimer bonds, of which half are of type (7b). The covering of Fig. 8(d) represents an exten- sion ofthe procedureof enlargingunit cells and removing four–site plaquettes, which demonstrates that it remains possiblein the limit ofno type–(7a) bonds to reduce frus- tration to 1/6 of the bonds, and to bonds of type (7c) [Fig. 9(b)], whence the energy of the covering is again Edim(0) =−13 32(4.14). Thus it is safe to conclude that, for the static–dimer problem, the ground–state manifold forα= 0 consists of a significant number of degener- ate coverings. We do not pursue these considerations further because of degeneracy lifting by dimer resonance processes, and because the energetic differences between staticdimerconfigurationsarelikelytobedwarfedbythe contributions from dimer resonance, the topic to which we turn in Sec. V. B. Direct exchange model The very strong preference for bond spin singlets (the factor of 4 in Eq. (2.21)] suggests that dimer states will also be competitive in this limit, even though only 1/6 of the bonds may redeem an energy of −J. Following the considerations and terminology of the previous subsec- tion, we note (i) that /an}bracketle{t/vectorSi·/vectorSj/an}bracketri}ht= 0 on interdimer bonds and (ii) that in this case, interdimer bonds have energy −1 4Jatη= 0 for types (7a) and (7b), and 0 for type (7c). Because any state with a maximal number (1/6) of20 type–(7a) bonds must have only bonds of type (7c) for the other 2/3 [states (8a)], such a state is manifestly less favorable at α= 1 than those of type (8b)–(8d), where there are no aligned pairs of dimers. In this latter case, the full calculation gives Edim(1) =−1 144J(9r1+19r2+8r3),(4.15) andEdim(1) =−1 4Jforη= 0. This energy does now ex- ceed that availablefrom the formationof Heisenbergspin chains in one of the three lattice directions (Sec. IIIC), which gave the value E1DAF(1) =−0.231J. At the level of these calculations, the manifold of de- generate states with this energy is very large, and its counting is a problem which will not be undertaken here. We will show in Sec. V that, precisely in this limit, no dimerresonanceprocessesoccurandthestaticdimercov- erings do already constitute a basis for the description of the ground state. The question of fluctuations leading to the selection of a particular linear combination of these states which is of lowest energy, i.e.of a type of order– by–disorder mechanism, is addressed in Ref. 49. At finite values of the Hund exchange, this type of statewillcomeintocompetitionwiththesimpleavoided– blocking states which gain, with a FM spin state, an energy EFM(1) =−1 6Jr1, (4.16) as 2/3 of the bonds contribute with an energy of −1 4Jr1. The critical value of ηrequired to drive the transition from the low–spin dimerized state to the FM state is found to be ηc= 0.1589. (4.17) C. Mixed model Because both of the endpoints, α= 0 andα= 1, favor dimerized states over states of long–ranged order, it is natural to expect that a dimer state will provide a lower energy also at α= 0.5. However, we remind the reader that there are no intermediate dimer bases, and caution there is no strong reason to expect one or other of the limiting dimer states to be favored close to α= 0.5. By inspection, the energy of an α>0 state can be obtained by direct addition of the diagonal interdimer bond con- tributions in an (ss/ot) or (os/st) dimer state, which is established by pure off–diagonalhopping, because no site occupancies arise which allow mixed processes. For the same reason, no interdimer terms impede a calculation of the energy of an α <1 state by summing the off– diagonal interdimer bond contributions in a spin–singlet dimer state stabilized by purely diagonal processes. We will not analyze the static dimer solutions for the inter- mediate regime in great detail, and provide only a crudeestimate of the α= 0.5 energy by averagingoverboth re- sults at the limits of their applicability. We will make no attempt here to exclude other forms of disordered state atα= 0.5, and return to this question in Sec. V. For each type of bond it is straightforward to compute the energy gained from interdimer hopping processes of the type not constituting the dimer state, and the re- sults are shown in Table II. The first four lines give the energies per bond from diagonal hopping processes oc- curring on the bonds of the different α= 0 dimer states, and conversely for the final two lines. It is clear that the occupations of type (7a) bonds preclude any hopping of the opposite type. For α= 0 dimer configurations, the interdimer diagonal hopping on (7b) bonds is always of avoided–blocking type, while on (7c) bonds a blocking can occur, and like the other terms is evaluated using /an}bracketle{t/vectorSi·/vectorSj/an}bracketri}ht. Forα= 1, off–diagonal hopping on the inter- dimer bonds is evaluated with /an}bracketle{t/vectorSi·/vectorSj/an}bracketri}ht= 0 between the spinsinglets: allprocesseson(7b)bondsarethoseforone active orbital; complications arise only for (7c) bonds, where an interdimer bond between parallel dimers has two active AO orbitals, while one between dimers which are not parallel has two active FO orbitals. Atα= 0.5, the energy of an (os/st) or (ss/ot) dimer state augmented by diagonal hopping processes is mini- mized by states (8a) and (8d): the interdimer bond con- tributions of all coverings in Fig. 8 are equal, despite the differenttypecounts, soonlythe α= 0energyisdecisive. Atη= 0, E(8a) o(0.5) =−1 2/parenleftbigg13 32+2 3·1 4/parenrightbigg J=−55 192J,(4.18) E(8d) o(0.5) =−1 2/parenleftbigg13 32+1 3·1 8+1 2·1 4/parenrightbigg J =−55 192J. (4.19) The energy of a spin–singlet dimer state augmented by off–diagonal hopping is minimal in states (8b) and, cu- riously, (8a): although the latter has explicitly a worse ground–state energy than the other states shown, the effect of the additional hopping is strong, not least be- cause all interdimer type–(8c) bonds arebetween parallel TABLE II: Additional interdimer bond energies at α= 0.5 due respectively to (i) diagonal hopping occurring in a stat e (designated by α= 0) stabilized by off–diagonal processes and (ii) off–diagonal hopping in a state ( α= 1) stabilized by diagonal processes. bond (7a) (7b) (7c) α= 0, (os/st), AF 0 −1 16(r1+r2)−1 8(r1+r2) α= 0, (os/st), FM 0 −1 8r1 −1 4r1 α= 0, (ss/ot), AO 0 −1 32(3r1+r2)−1 16(3r1+r2) α= 0, (ss/ot), FO 0 −1 32(3r1+r2)−1 16(3r1+r2) α= 1,/bardbldimers 0 −1 16(3r1+r2)−1 8(3r1+r2) α= 1, non– /bardbldimers 0 −1 16(3r1+r2)−1 12(2r2+r3)21 dimers. Thus at η= 0, E(8a) d(0.5) =−1 2/parenleftbigg5 24+1 2·2 3/parenrightbigg J=−13 48J, E(8b) d(0.5) =−1 2/parenleftbigg1 4+1 3·1 4+1 2·2 3·1 2+1 2·1 3·1 4/parenrightbigg J =−13 48J. (4.20) Despite the fact that these are two completely differ- ent expansions, it is worth noting that the two sets of numbers are rather similar, which occurs because the significantly inferior energy of the α= 1 ground state is compensated by the significantly greater interdimer bond energies available from off–diagonal hopping pro- cesses. However, this result also implies that no special combinations of diagonal and off–diagonal dimers can be expected to yield additional interdimer energies beyond this value. Takingthe covering(8a)asrepresentativeofthelowest available energy, but bearing in mind that many other states lie very close to this value, an average over the two approaches yields E(8a) dim(0.5) =−107 384J (4.21) atη= 0. This number is no longer lower than the value obtained in Sec. IIID for fully ordered states gaining en- ergy from mixed processes, raising the possibility that non–dimer–based phases may be competitive in the in- termediate regime, where neither of the limiting types of dimer state alone is expected to be particularly suit- able. However, we will not investigate this question more systematicallyhere, and cautionthat the approximations made both in Sec. IIID and here make it difficult to draw a definitive conclusion. D. Summary The results of this section make it clear that static dimer states, while showing the same energetic trend, are considerably more favorable than any long–range– ordered states (Sec. III) over most of the phase diagram. As a function of α, the dimer energy increases mono- tonically from −13 32Jto−1 4J, and both end–point values also lie below the results obtained for quasi–1D spin– disordered states in Sec. III. We stress that the results of this section are provisional in the sense that we have not performed a systematic exploration of all possible dimer coverings, but rather have focused on a small number of examples illustrative of the limiting cases in terms of interdimer bond types. More importantly, we have con- sidered only static dimer coverings with effective inter- dimer interactions: the kinetic energy contributions due to dimer resonance processes for all values of α <1 are missing in this type of calculation. For this reason, wehave also refrained from investigating higher–order pro- cesses, which may select particular dimer states from a manifold of static coverings degenerate at the level of the current considerations. Gaining some insight into the magnitude and effects of resonance contributions is the subject of the following section. V. EXACT DIAGONALIZATION A. Clusters and correlation functions In this Section we present results obtained for small systems by full exact diagonalization (ED). Because each site has two spin and three orbital states, the dimension of the Hilbert space increases with cluster size as 6N, whereNis the number of sites. As a consequence, we focus here only on systems with N= 2, 3, and 4 sites: all three clusters can be considered as two–, three– or four–site segments of an extended triangular lattice, con- nected with periodic boundary conditions. For the single bond and triangle this only alters the bond energies by a factor of two, a rescaling not performed here, but for the four–site system it is easy to see that the intercluster bonds ensure that the system connectivity is tetrahedral. We will also compare some of the single–bond and tetra- hedron results with those for a four–site chain. Other ac- cessible cluster sizes ( N= 5 and 6) yield awkwardshapes which disguise the intrinsic system properties. Indeed we will emphasize throughout this Section those features of our very small clusters which can be taken to be generic, and those which are shape–specific. Given the clear tendency to dimerization illustrated in Secs. III and IV, it is to be expected that spin correla- tion lengths in all regimes of αare very small. To the extent that the behavior of the model for any parameter setis drivenby localphysics, the clusterresultsshouldbe highly instructivefor such trends asdimer formation, rel- ative roles of diagonal and off–diagonal hopping, dimer resonance processes, lifting of degeneracies both in the orbital sector and between states of (os/st) and (ss/ot) dimers, and the importance of joint spin–orbital corre- lations. However, generic features of extended systems which cannot be accessed in small clusters are those con- cerning questions of high system degeneracy and subtle selection effects favoring specific states. We will compute and discuss the cluster energies, de- generacies, site occupations, bond hopping probabili- ties in diagonal and off–diagonal channels (discussed in Sec. VC), and the spin, orbital, and spin–orbital (four– operator) correlation functions. All of these quantities will be calculated for representative values of αandη coveringthefullphasediagram,andeachcontainsimpor- tant information of direct relevance to the local physics properties listed in the previous paragraph. Although the systems we study are perforce rather small, we will showthat onemayrecognizein them anumber ofgeneral trends valid also in the thermodynamic limit.22 We introduce here the three correlation functions, which for a bond /an}bracketle{tij/an}bracketri}htoriented along axis γare given respectively by Sij≡1 d/summationdisplay n/angbracketleftbig n/vextendsingle/vextendsingle/vectorSi·/vectorSj/vextendsingle/vextendsinglen/angbracketrightbig , (5.1) Tij≡1 d/summationdisplay n/angbracketleftbig n/vextendsingle/vextendsingle/vectorTiγ·/vectorTjγ/vextendsingle/vextendsinglen/angbracketrightbig , (5.2) Cij≡1 d/summationdisplay n/angbracketleftbig n/vextendsingle/vextendsingle(/vectorSi·/vectorSj)(/vectorTiγ·/vectorTjγ)/vextendsingle/vextendsinglen/angbracketrightbig −1 d2/summationdisplay n/angbracketleftbig n/vextendsingle/vextendsingle/vectorSi·/vectorSj/vextendsingle/vextendsinglen/angbracketrightbig/summationdisplay m/angbracketleftbig m/vextendsingle/vextendsingle/vectorTiγ·/vectorTjγ/vextendsingle/vextendsinglem/angbracketrightbig ,(5.3) wheredis the degeneracy of the ground state. The definitions of the spin ( Sij) and orbital ( Tij) correla- tion functions are standard, and we have included ex- plicitly all of the quantum states {|n/an}bracketri}ht}which belong to the ground–state manifold. The correlation function Cij (5.3) contains information about spin–orbital entangle- ment, as defined in Sec. I: it represents the difference between the average over the complete spin–orbital op- erators and the product of the averagesover the spin and orbital parts taken separately. It is formulated in such a way that Cij= 0 means the mean–field decoupling of spin and orbital operators on every bonds is exact, and bothsubsystemsmaybetreatedindependentlyfromeach other. Such exact factorizability is found9in the high– spin states at large η; its breaking, and hence the need to handle coupled spin and orbital correlations in a sig- nificantly more sophisticated manner, is what is meant by “entanglement” in this context. B. Single bond We consider first a single bond oriented along the c– axis (Fig. 10). In the superexchange limit the active or- bitals areaandb, while for direct exchange only the cor- bitalscontributeinEq.(2.7). AsdiscussedinSec.IVA, a single bond gives energy −Jin the superexchange model (α= 0) [Fig. 10(a)], where the ground state has degener- acyd= 6 atη= 0, from the two triply degenerate wave functions (ss/ot) and (os/st). At finite η, the latter is favored as it permits a greater energy gain from excita- tions to the lowest triplet state in the d2configuration [Eqs. (2.15) and (2.16)]. Although orbital fluctuations which appear in the mixed exchange terms in Eq. (2.22) may in principle contribute at α >0, one finds that the wave function remains precisely that for α= 0,i.e.(ss/ot) degenerate with (os/st), all the way to α= 0.5. Thus for the param- eter choice specified in Sec. II, the ground–state energy increases to a maximum of E0=−0.5Jhere [Fig. 10(a)]. The degeneracy d= 6 is retained throughout the regime α<0.5, and only at α= 0.5 do several additional states join the manifold, causing the degeneracy to increase to d= 15. For the entire regime α∈(0.5,1], the ground0 0.2 0.4 0.6 0.8 1 α−0.8−0.6−0.4−0.20.0Sij , Tij , Cij0 0.2 0.4 0.6 0.8 1−1.0−0.8−0.6−0.4−0.20.0En/J {a,b} orbitals c orbitals(ss/ot) (st/os)(ss/cc)(a) (b) resonating static13 18 86 6 FIG. 10: (Color online) Evolution of the properties of a sing le bondγ≡cas a function of αatη= 0: (a) energy spectrum (solid lines) with degeneracies as shown; (b) spin ( Sij, filled circles), orbital ( Tij, empty circles), and spin–orbital ( Cij,×) correlations: Sij=Tij=Cij=−0.25 forα <0.5, while Tij=Cij= 0 for α >0.5. The ground–state energy E0is −Jfor both the superexchange ( α= 0) and direct–exchange (α= 1) limits, and its increase between these is a result of the scaling convention. The transition between the two regimes occurs by a level crossing at α= 0.5. For α <0.5, the two types of dimer wave function [(ss/ot) and (os/st)] are degenerate ( d= 6) for resonating orbital configurations {ab}, while at α >0.5, the nondegenerate spin singlet is supported by occupation of corbitals at both sites [(ss/cc)]. state is a static orbitalconfigurationwith corbitals occu- pied at both sites to support the spin singlet, and d= 1. The evolution of the spectrum with αdemonstrates not only that superexchange and direct exchange are physi- cally distinct, unable to contribute at the same time, but that the two limiting wave functions are extremely ro- bust, their stability quenching all mixed fluctuations for a single bond. In this situation it is not the ground–state energy but the higher first excitation energy which re- veals the additional quantum mechanical degrees of free- dom active at α= 0 compared to α= 1 [Fig. 10(a)]. The spin, orbital, and composite spin–orbital corre- lation functions defined in Eqs. (5.1)–(5.3) give more insight into the nature of the single–bond correlations. The degeneracy of wavefunctions (ss/ot) and (os/st) for 0≤α <0.5 leads to equal spin and orbital correlation functions, as shown in Fig. 10(b), and averaging over the different states gives Sij=Tij=−1 4. As a singlet for onequantityismatchedbyatripletfortheother, the two23 sectors are strongly correlated, and indeed Cij=−1 4, in- dicating an entangled ground state. However, a consider- ablymoredetailedanalysisispossible. Eachofthe sixin- dividual states {|n/an}bracketri}ht}within the ground manifold has the expectation value/angbracketleftbig n/vextendsingle/vextendsingle(/vectorSi·/vectorSj)(/vectorTiγ·/vectorTjγ)/vextendsingle/vextendsinglen/angbracketrightbig =−3 16, which we assert is the minimum possible when the spin and pseudospin are the quantum numbers of only two elec- trons. It is clear that if the operator in Cijis evaluated for any one of these states alone, the result is zero. En- tanglement arisesmathematically because of the product ofaveragesinthesecondtermofEq.(5.3), andphysically because the ground state is a resonant superposition of a number of degenerate states. We emphasize that the resulting value, Cij=−1 4, is the minimum obtainable in this type of model, reflecting the maximum possible entanglement. We will show in Sec. VE that this value is also reproduced for the Hamiltonian of Eq. (2.7) on a linear four–site cluster, whose geometry ensures that the system is at the SU(4) point of the 1D SU(2) ⊗SU(2) model.9 By contrast, for α >0.5 those states favored by su- perexchangebecomeexcited, andthespin–singletground state hasSij=−3 4. The orbital configuration is charac- terized by /an}bracketle{tnicnjc/an}bracketri}ht= 1, a rigid order which quenches all orbital fluctuations (indeed, the orbital pseudospin vari- ables/vectorTiγare zero). Thus the spin and orbital parts are trivially decoupled, giving Cij= 0. Finally, at the tran- sition point α= 0.5, averaging over all 15 degenerate states yields Sij=−0.15,Tij=−0.10, andCij=−0.09. In summary, the very strong tendency to dimer forma- tion in the two limits α= 0 andα= 1 precludes any contribution from mixed terms on a single bond, lead- ing to a very simple interpretation of the ground–state properties for all parameters. C. Triangular cluster We turn next to the triangle, which has one bond in each of the lattice directions a,b, andc. Unlike the case of the single bond, here the spin–orbital interactions are strongly frustrated, in a manner deeper than and quali- tatively different from the Heisenberg spin Hamiltonian. Not only can interactions on all three bonds not be sat- isfied at the same time, but also the actual form of these interactions changes as a function of the occupied or- bitals. The triangle is sufficient to prove (numerically and analytically) the inequivalence in general of the orig- inal model and the model after local transformation, for frustration reasons discussed in Sec. IIB. Webeginwiththeobservationthattheresultstofollow are interpreted most directly in terms of resonant dimer states on the triangle. This fact is potentially surprising, given that the number of sites is odd and dimer forma- tion must always exclude one of them, but emphasizes the strong tendencies to dimer formation in all param- eter regimes of the model. For their interpretation we use a VB ansatz where it is assumed that one bond isoccupied by an optimal dimer state, minimizing its en- ergy, and the final state of the system is determined by the contributions of the other two bonds. This ansatz is perforce only static, and breaks the symmetry at a crude level, but enables one to understand clearly the effects of the resonance processes captured by the numerical stud- ies in restoring symmetries and lowering the total energy. Considering first the VB ansatz for the superexchange model, the energy −Jmay be gained only on a single bond, in one of two ways. For the bond spin state to be a singlet ( S= 0, (ss/ot) wave function), two differ- ent active orbitals are occupied at both sites in one of the orbital triplet states. The other two bonds lower the total energy when the third site has an electron of the third orbital color, each gaining an energy of −0.25Jdue to the orbital interactions in Eq. (2.8). The energy of the triangle is then EVB(0) =−0.5Jper bond, and the cluster has a low–spin ( S=1 2) ground state with degen- eracyd= 6 from the combination of the orbital triplet and the spin state of the third electron. We stress that the location ( a,b, orcbond) of the spin singlet does not contribute to the degeneracybecause the three VB states are mixed within the ground state by the contributing off–dimer hopping processes. The same considerations applied to an (os/st) dimer on one of the bonds of the triangle shows that there is no color and spin state of the third electron which allows both non–dimer bonds to gain the energy −0.25Jsimultaneously, so the cluster has a higher energy of −5 12Jper bond. Thus the VB ansatz illustrates a lifting of the degeneracy between the two types singlet state, the physical origin of which lies in the permitted off–dimerfluctuation processes,and this will be borne out in the calculations below. However, the net spin state of the cluster has little effect on the esti- mated energy of the (os/st) case, and its high–spin ver- sion (S=3 2) will be a strong candidate for the ground state at higher values of η. In the direct–exchange limit (α= 1), the VB ansatz for spin singlets again returns an energyEVB(1) =−5 12J, alsobecauseonlyonenon–dimer bond cancontribute. Herethe off–dimerprocessesarere- stricted to the third electron, which has arbitrary color and spin, and cannot mix the three VB states, whence the degeneracy is d= 12. With this framework in mind, we turn to a description of the numerical calculations at all values of α, beginning withthemostimportantresults: at α= 0thedegeneracy isd= 6, and hence VB resonance is confirmed, yielding an energy very much lower than the static estimate, at E0=−0.75Jper bond [Fig. 11(a)]. Thus strong orbital dynamics and positional resonance effects operate in the ground–state manifold. These break the (ss/ot)/(os/st) symmetry, but act to restore other symmetries broken in the VB ansatz. At α= 1, the energy and degeneracy from the VB ansatz are exact, showing that the orbital sector is classical andintroduces no resonance effects. Figure 11(a) shows the complete spectrum of the tri- angular cluster for all ratios of superexchange to di- rect exchange, and in the absence of Hund coupling.24 0.0 0.2 0.4 0.6 0.8 1.0−0.80−0.60−0.40−0.200.000.20En/J 0.0 0.2 0.4 0.6 0.8 1.0−0.3−0.2−0.10.00.1Sij , Tij , Cij 0.0 0.2 0.4 0.6 0.8 1.0α0.00.20.40.6n1γ (ss/ot) (ss/cc) orbitalsfluctuating{a,b} orbitals c orbitals(b)(a) resonating static488 (c) FIG. 11: (Color online) (a) Energy spectrum per bond for a triangular cluster as a function of αforη= 0. Ground– state degeneracies are as indicated, with d= 6 at α= 0 andd= 12 atα= 1. The arrows mark two transitions in the nature of the (low–spin) ground state, which are further cha r- acterized in panels (b) and (c). (b) Spin ( Sij, filled circles), orbital (Tij, empty circles), and spin–orbital ( Cij,×) correla- tion functions on the cbond. (c) Average electron densities in thet2gorbitals at site 1 [Figs. 2(b,c)], showing n1b(solid line) andn1a=n1c(dashed). The orbital labels are shown for a c bond. All three panels show clearly a superexchange regime forα <0.32, a direct–exchange regime for α >0.69, and an intermediate regime (0 .32< α <0.69). A full description is presented in the text. Frustration of spin–orbital interactions is manifest in rather dense energy spectra away from the symmetric points, and in a ground–state energy per bond signifi- cantly higher than the minimal value −J. Atα= 0 the spectrum is rather broad, with a significant number of states of relatively low degeneracy due to the strong fluctuations and consequent mixing of VB states in this regime. However, even in this case the ground state is well separated from the first excited state. As empha- sized above, the ground–state energy, E0(0) =−0.75J, is quite remarkable, demonstrating a very strong energy gain from dimer resonance processes. By contrast, thevalueE0(1) =−5 12Jper bond found at α= 1 is exactly equal to that deduced from the VB ansatz, demonstrat- ing that this wave function is exact. Here the excited states have high degeneracies, mostly of orbital origin, and thus the spectrum shows wide gaps between these manifolds of states; this effect is more clearly visible in Fig. 12(c). The degeneracies shown in Fig. 11(a) are dis- cussed below. In the intermediate regime, many of the degeneracies at the end–points are lifted, leading to a very dense spectrum. The two transitions at α= 0.32 andα= 0.69 appear as clear level–crossings: the inter- mediate ground state is a highly excited state in both of the limits ( α= 0,1), reinforcing the physical picture of a very different type of wave function dominated by orbital fluctuations and, as we discuss next, with little overt dimer character. Thecorrelationfunctionsforanyonebondofthetrian- gle are shown in Fig. 11(a). That Sijis constant for all α can be understood in the dimer ansatz by averaging over the three configurations with one (ss/ot) or (ss) bond and one ’decoupled’ spin on the third site, which gives Sij=−1 4everywhere. The orbital and spin–orbital cor- relation functions show a continuous evolution accompa- nied by discontinuous changes at two transitions, where the nature of the ground state is altered. The orbital correlation function Tij=−1 12atα= 0 may be under- stood as an average over the orbital triplet (+1 4) and the two non–dimer bonds (each −1 4). Whenαincreases, this value is weakened by orbital fluctuations, and undergoes atransitionat α= 0.32toaregimewhereorbitalfluctua- tions dominate, and Tijis close to zero. Above α= 0.69, Tijbecomes positive, and approaches +1 12asα→1, indicating that the wave function changes to the static– dimer limit. While /vectorTicvanishes on the cbond here, the cluster average has a finite value due to the contribution Tij=1 4from the active non–singlet bond. The spin–orbital correlation function Cijalso marks clearly the three different regimes of α. Whenα<0.32, Cijhasasignificantnegativevalue[Fig.11(b)]whosepri- mary contributions are given by the four–operator com- ponent/an}bracketle{t(/vectorSi·/vectorSj)(/vectorTiγ·/vectorTjγ)/an}bracketri}ht. Bycontrast, Cijis closetozero in the intermediate regime, increasing again to positive values forα>0.69. For all α>0.32,Cijcan be shown to be dominated by the term −SijTijin Eq. (5.3), while the four–operator contribution is small, and vanishes as α→1. Thus entanglement, defined as the lack of fac- torizability of the spin and orbital sectors, can be finite even for vanishing joint spin–orbital dynamics. Further valuable information is contained in the or- bital occupancies at individual sites [Fig. 11(b)], which show clearly the three different regimes. Although there is always on average one electron of each orbital color on the cluster, these are not equally distributed, as each site participates only in two bonds and the symmetry is broken. A representative site, labelled 1 in Figs. 2(b,c)] has onlyaandcbonds, and hence the electron density in theborbital is expected to differ from the other two. The values nb= 2/3 andna=nc= 1/6 found in the25 regimeα<0.32 is understood readily as following from a 1/3 average occupation of ( ab) and (bc) orbital triplet states on the candabonds, respectively, and of an ( ac) orbitaltriplet state on the bbond, which ensuresthat the electron at site 1 is in orbital b[Fig. 2(b)]. By contrast, in the regime α >0.69, only the two static orbital con- figurations ( cc) and (bb) on thecandbbonds contribute, andna=nc=1 2, whilenb= 0; when the system is in the third possible spin–singlet state, with a ( bb) orbital state on thebbond, the third electron is either aorc. Between these two regimes(0 .32<α<0.69) is an extended phase with equal averageoccupancy of all three orbitals at each site, a potentially surprising result given the broken site symmetry of the cluster. While this may be interpreted as a restoration of the symmetry of the orbital sector by strong orbital fluctuations, including those due to terms inHm(2.22), it does not imply a higher symmetry of the strongly frustrated interactions at α= 0.5. ThespectraasafunctionofHundcoupling ηareshown in Fig. 12 for the α= 0 andα= 1 limits, and at α= 0.5 to represent the intermediate regime. The lifting of de- generacies as a function of ηis a generic feature. States of higher spin are identifiable by their stronger depen- dence onη, and in all three panels a transition is vis- ible from a low–spin to a high–spin state. At α= 0 [Fig. 12(a)], the large low– ηgap to the next excited state resultsinthetransitionoccurringattheratherhighvalue ofηc= 0.158. This can be taken as a further indication of the exceptional stability of the resonance–stabilized ground state in the low–spin sector. The degeneracy d= 12 of the high–spin state is discussed below. The transition to the high–spin state at α= 1 also occurs at a high critical value, ηc= 0.169 [Fig. 12(a)], due in this case quite simply to the lack of competition for the strong singlet states on individual bonds. Only in the intermediate regime, 0 .32< α <0.69, where we have shown already that the orbital state is quite differ- ent from that in either limit [Fig. 11], is the transition to the high–spin state much more sensitive to η. The or- bitalfluctuations inthis phaseoccurboth in the low–spin and the high–spin channel, making these very similar in energy, and the transition occurs for α= 0.5 at only ηc= 0.033 [Fig. 12(b)]. As expected from the α= 0 limit, where fluctuations are also strong, the characteris- tic features of this energy spectrum are low degeneracy and a semicontinuous nature. The location of the high– spin transition as a function of αmay be used to draw a phase diagram for the triangular cluster, which has the rather symmetric form shown in Fig. 13. Yet more information complementary to that in the energy spectra and correlation functions can be obtained by considering the average “occupation correlations” for a bond/an}bracketle{tij/an}bracketri}ht /bardblγ, P=/an}bracketle{tniγnjγ/an}bracketri}ht, (5.4) Q=/an}bracketle{tniγ(1−njγ)/an}bracketri}ht+/an}bracketle{t(1−niγ)njγ/an}bracketri}ht,(5.5) R=/an}bracketle{t(1−niγ)(1−njγ)/an}bracketri}ht. (5.6) These probabilities ( P+Q+R= 1) reflect directly the0 0.05 0.1 0.15 0.2−1.0−0.6−0.20.2En/J0 0.05 0.1 0.15 0.2−1.2−0.8−0.40.0En/J 0 0.05 0.1 0.15 0.2 η−0.6−0.4−0.20.0En/J12 6 128(a) (b) (c)44 FIG.12: (Color online) Energyspectrafor atriangular clus ter as a function of Hund exchange η. Energies are quoted per bond, and shown for: (a) α= 0, (b) α= 0.5, and (c) α= 1. The arrows indicate transitions at ηcfrom the low–spin ( S= 1/2) to the high–spin ( S= 3/2) ground state. The numbers in all panels give degeneracies for the two lowest states for η < ηcandη > ηc, respectively. nature of the resonance processes contributing to the en- ergy of the cluster states, in that they show the relative importance of diagonal and off–diagonal hopping in the ground states, and the evolution of these contributions withαandη. We do not present these quantities in de- tail here, but only summarize the overall picture of the ground state whose understanding they help elucidate. For this summary we return to the VB framework, which accounts for many of the basic properties illus- trated in the numerical results presented above. Con- sidering first the low–spin states ( η= 0), atα= 0 the ground state is given by one (ss/ot) dimer resonating around the three bonds of the cluster; the third site has the third color, its hopping gives a largevalue of Q= 1/3 (R= 2/3 from the pure superexchange channel) and its26 0 0.2 0.4 0.6 0.8 1 α0.000.040.080.120.160.20η S=1/2S=3/2 FIG. 13: (Color online) Phase diagram of the triangular clus - ter in the plane ( α,η). The spin states below and above the transition line ηc(α) are respectively spin doublet ( S= 1/2) and spin quartet ( S= 3/2). spin an addition twofold degeneracy ( d= 3×2 = 6); the orbital occupation of the (ss/ot) dimer is responsible for the net 1/6:1/6:2/3 occupation distribution. When α >0 the state remains essentially one with a resonat- ing spin singlet, large Qand dominant R, but the orbital triplet degeneracy is lifted to 2+1 and the ground–state degeneracy to d= 2×2. All quantities, including P,Q, andR, undergo discontinuous changes at α≃0.32, and in this regime there is no longer strong evidence for an interpretation in terms of resonating spin singlets: large Q≃2/3 and the equal site occupations suggest the dom- inance of mixed hopping processes which are not con- sistent with either mechanism of singlet formation. The retention of fourfold degeneracy across this transition is largely accidental, and stems from twofold spin and or- bital contributions. Only for α >0.69 is a spin–singlet description once again valid: here Pbecomes significant, astheresonatingsingletisstabilizedbydiagonalhopping where the orbital has the bond color. The third site now has one of two possible colors, its hopping keeps Qlarge, and its spin yields another twofold degeneracy, as do the orbital states, whence the net degeneracy is d= 2×2×2. Only atα= 1 does the spin singlet become static, while the third site still has either of the other colors, yield- ing the symmetric result P= 1/9,Q=R= 4/9, and degeneracy d= 12. A similar description is possible in the high–spin states atη >ηc. Atα= 0 the (os/st) dimer is rendered static by the fact that hopping to the third site is now excluded if it has the third color, and so instead this site takes one of the singlet colors, a twofold degree of freedom which, however, does not allow singlet motion; as a consequence the orbital occupation is uniform (1/3:1/3:1/3), the hop- pingprocessesincludecontributionsinthediagonalchan- nel (P= 1/6,Q= 1/3,R= 1/2) and the degeneracy isd= 3×4 = 12. For α >0 the orbital singlet mayagain resonate, but the third site retains one of the sin- glet colors, orbital degeneracy is broken and d= 4. Once again strong mixed processes dominate the intermedi- ate regime, in which the spin state is not an important determining factor. Above α= 0.69 the critical value ηcrequired to overcome spin singlet formation becomes largeagain,andthehigh–spinstateisonewhereavoided– blocking processes (large Q) dominate, while broken or- bital degeneracy keeps d= 4. Finally, at α= 1 one obtains a pure avoided–blocking state with orbital con- figurations acborcbafor the sites (1 ,2,3) of Fig. 2(c), and consequent degeneracy d= 4×2 = 8. Thus it is clear that the high– ηregion is also one yielding interest- ing orbital models with nontrivial ground states, some including orbital singlet states. D. Tetrahedral cluster Asinthecaseofthetriangularlattice,interpretationof thenumericalresultsforthetetrahedralcluster(four–site plaquette of the triangular lattice) is aided by considera- tion of the VB ansatz in the two limits of superexchange and direct–exchange interactions. The tetrahedral clus- ter can accommodate exactly two dimers, with all inter- dimer bonds of type (7c), and may thus be expected to favor dimer–based states by simple geometry. However, because the considerations and comparisons of this sub- sectionaregivenonlyforthis singleclustertype, anybias of this sort would not invalidate the results and trends discussed here. Because of the different forms and symmetries of the spin and orbital sectors, there is no possibility of elemen- tary spin–orbital operators, or of a ground–state wave function which is a net singlet of a higher symmetry group. The state with two orbital singlets on one pair of bonds, two spin singlets on a second pair and pure interdimer bonds on the third pair does exist, but is not competitive: the energy cost forremovingthe orbital sin- glets from the spin state maximizing their energy is by no means compensated by the energy gain from having two spin singlet bonds in an orbital state which also does not maximize their energy. This result may be taken as a further indication for the stability of dimers only in the forms (os/st) or (ss/ot) in this model, and states of sharedorbitalandspinsingletsarenotconsideredfurther here. We return to this point in the following subsection, in the context of the four–site chain. We discuss only the energies of the VB wave functions atη= 0. The minimal values obtainable for /an}bracketle{t/vectorSi·/vectorSj/an}bracketri}htand /an}bracketle{t/vectorTiγ·/vectorTjγ/an}bracketri}htonthe interdimerbondsis −1/4,corresponding to the AF/AO order. Thus at α= 0 the energy per bond is Eos/st(0) =Ess/ot(0) =−1 2J, (5.7) with the degeneracy of the (ss/ot) and (os/st) wave func- tionsrestoredasforthe singlebond. In the limit ofdirect27 exchange, the VB wave function consists of spin singlets with twoactiveorbitalsofthe bond. Thegeometryofthe cluster precludes these orbitals from being active on any of the interdimer bonds, as a result of which the energy per bond at η= 0 is E(1) =−1 3J, (5.8) and the ground state has degeneracy d= 3. The most important results for the tetrahedron, which we discuss in detail in the remainder of the subsection, are the following. At α= 0, the exact ground state energy isE0=−0.5833J: while not as large as in the case of the triangle (Sec. VC), the resonance energy contribution is very significant also for an even number of cluster sites. The degeneracy of the numerical ground state,d= 6, has its origin in only one of the (ss/ot) or (os/st) wavefunctions (below), demonstratingagainthat thereisnosensein whichthe quantumfluctuationsinthe spinandorbitalsectorsaresymmetrical,andthattheVB ansatz is capturing the essence of the local physics only at a very crude level. At α= 1, as also for the triangular cluster, the numerical results confirm not only the energy given by the VB ansatz but every detail (degeneracies, occupations, correlations) of this state. We begin the systematic presentation of results by dis- cussing the energy spectra at η= 0 [Fig. 14(a)]. As soon as the degeneracies of the superexchange limit ( α= 0) are broken, the spectrum becomes very dense, and re- mains so across almost the complete phase diagram until a level–crossing at αc= 0.92. The ground–state energy for all intermediate values of αinterpolates smoothly to- wards the transition, showing an initial decrease not ob- served in the triangle: for the tetrahedron, mixed hop- ping terms make a significant contribution, leading to an overall energy minimum around α= 0.15. The domi- nance of these terms is indicated by both the extremely high value of αcand the steepness of the low– αcurve where the transition to the static VB phase is finally reached. The bond correlation functions shown in Fig. 14(b) illustrate the effects of corrections to the VB ansatz. The spin correlations always have the constant value Sij=−1 4, which is the most important indication of the breakingofsymmetrybetween(ss/ot)and(os/st)sectors at lowα: this value is an average over the spin–singlet result−3/4 (on two bonds) and four bonds with value 0, and thus it is clear that (ss/ot) dimers afford more res- onance energy. However, the proximity of (os/st) states suggests that a low value of ηc, the critical Hund cou- pling for the transition to the high–spin state, is to be expected (below). The orbital correlations average to zero at α= 0, a non–trivial result whose origin lies in the breaking of nine–fold degeneracy within the orbital sector, and re- main close to this value until the transition at αc. It is worth noting here that Tij= 0 implies a higher frustra- tion in the orbital sector than would be obtained in the0.0 0.2 0.4 0.6 0.8 1.0 α−0.3−0.2−0.10.00.1Sij , Tij , Cij0.0 0.2 0.4 0.6 0.8 1.0−0.6−0.5−0.4−0.3−0.2En/J (a) (b)2 1 FIG. 14: (Color online) (a) Energy spectrum per bond for a tetrahedral cluster as a function of αforη= 0. Ground– state degeneracies are as indicated, with d= 6 at α= 0 andd= 150 at α= 1. The arrow marks a transition in the nature of the (low–spin) ground state. (b) Spin ( Sij, filled circles), orbital ( Tij, empty circles), and spin–orbital ( Cij,×) correlation functions on the cbond of the tetrahedral cluster as functions of αforη= 0. spin sector for an (os/st) state ( Sij=−1 12), which is due to the complex direction–dependence of the orbital de- grees of freedom. This phase is maintained across much of the phase diagram, with only small changes to the cor- relation functions, the negative value of Tijreflecting an easing of orbital frustration. The lack of a phase transi- tion throughout the region in which mixed processes are also important suggests that a dimer–based schematic picture of the ground state remains appropriate for the four–site system, with only quantitative evolution as a function of αuntilαc= 0.92. Atα= 1, the result Tij=−1 6is the consequence of c–orbital operators on the interdimer aandbbonds. Significant spin–orbital correlations, Cij≃ −0.1 at α= 0 [Fig. 14(b)], are found to be due exclusively to the four–operator term at low α. While these negative contributions drop steadily through most of the regime α < α c, signifying a gradual decoupling of orbitals and spins as the static limit ( α= 1) is approached, near αc the negative value of Cijis again enhanced by the con- tribution −SijTijdue to the interdimer bonds. Thus, as for the triangle (Sec. VC), the entanglement is finite,28 0 0.05 0.1 0.15 0.2−1.0−0.8−0.6−0.4−0.20.0En/J 0 0.05 0.1 0.15 0.2 η−0.4−0.3−0.2−0.10.0En/J0 0.05 0.1 0.15 0.2−1.0−0.8−0.6−0.4−0.20.00.2En/J(a) (b) (c) FIG. 15: (Color online) Energy spectra for a tetrahedral clu s- ter as a function of Hundexchange η. Energies are quoted per bond, and shown for: (a) α= 0, (b) α= 0.5, and (c) α= 1. The arrows indicate transitions from the low–spin ( S= 0) to the high–spin ( S= 2) ground state. complete factorization is not possible, and a finite value Cij=−1 24is found even at α= 1. We note here that on the tetrahedron there is little information in the orbital occupations, which are constant ( nγ=1 3) over the entire phase diagram, demonstrating only the symmetry of this cluster geometry, and are therefore not shown. ThespectraasafunctionofHundcoupling ηareshown for the three parameter choices α= 0, 0.5, and 1 in Fig. 15. Once again, the spectra become very dense away fromη= 0. Atα= 0 [Fig. 15(a)] high–spin states are found also in the low–energy sector, as a consequence of the near–degeneracy of (ss/ot) and (os/st) states, and the high–spin transition occurs at a very low value of ηc [Fig. 15(a)]. The direct–exchange limit is both qualita- tively and quantitatively different, because the quantum fluctuations and the corresponding energy gains are lim- ited to the spin sector, making the low–spin states con-0 0.2 0.4 0.6 0.8 1 α0.000.040.080.120.160.20η S=0S=2 FIG.16: (Color online)Phase diagram of thetetrahedral clu s- ter in the plane ( α,η). As for the triangular cluster, the spin states below and above the line ηc(α) are respectively singlet (S= 0) and quintet ( S= 2), with no intermediate triplet phase. siderably more stable and giving ηc= 0.175 [Fig. 15(c)]. The spin excitation gap decreasesgraduallywith increas- ingη, but until just below ηc, for all values of α, the spin excitation is to S= 1 states. However, these triplet states are never the ground state in the entire regime ofη, a single transition always occurring directly into anS= 2 state. In the intermediate regime represented byα= 0.5, the energy spectrum is so dense that indi- vidual states are difficult to follow (a more systematic analysis of the spectra in different subspaces of Szis not presented here). The high–spin transition occurs at the relatively high value ηc= 0.136, due mainly to the large energygains in the low–spinsector from mixed exchange. Further evidence for the importance of the orbital exci- tations in Hm(2.22) can be found in the broadening of the spectrum which leads to the occurrence of quantum states with weakly positive energies: for both superex- change and direct–exchange processes, the Hamiltonians are constructed as products of projection operators with negative coefficients, so positive energies are excluded. The low– to high–spin transition points at all values of αcan be collected to give the full phase diagram of the tetrahedron shown in Fig. 16. As shown above, in the superexchange limit the high–spin state lies very close to the low–spin ground state, and the transition to an S= 2 spin quintet occurs at ηc= 0.017. We comment here that this high–spin state is in no sense classical or trivial, being based on orbital singlets which are stabi- lized by strong orbital fluctuations, and emphasize again that the high–spin sector also contains a manifold of rich problems in orbital physics, which we will not consider further here. The near degeneracy of (ss/ot) and (os/st) states is further lifted in the presence of the mixed terms inHm, raisingηcto values on the order of 0.12 acrossthe bulk of the phase diagram. For no choice of parameters29 is a spin triplet state found at intermediate values of η. The reentrant behavior close to α= 0.5 is an indication oftheimportanceofmixedtermsinstabilizingalow–spin state, the tetrahedral geometry providing one of the few examples we have found of anything other than a direct competition, andhenceaninterpolation,betweenthetwo limiting cases. The rapid upturn in the limit of α→1 reflects the anomalous stability of the static VB states in the direct–exchange limit. The very strong asymme- try of the transition line in Fig. 16 contrasts sharply with the near–symmetryabout α= 0.5 observed for the trian- gle (Fig. 13), and shows directly the differences between those features of the phase diagram which are universal and those which are effects of even or odd cluster sizes in a dimer–based system. We close our discussion of the tetrahedral cluster with a brief discussion of degeneracies and summary of the picture provided by the VB ansatz with additional res- onance. For the orbital occupation correlations and de- generacies, we begin with the low–spin sector ( η= 0). Atα= 0 one has two (ss/ot) VBs resonating around the 6 bonds of the cluster, a state characterized by P= 1/6, Q= 1/3, andR= 1/2; however, a mixing of the orbital triplet states lowers the degeneracy from 9 to d= 6. Forα >0 the state is the same, with slow evolution of P <1/6,Q >1/2, andR >1/3, but now mixed hop- ping terms break all orbital degeneracies, giving d= 1. Only when α>0.92 is the ground state more accurately characterized as one based on spin singlets of the bond color, with significant values of Pand the restoration of anorbitaldegeneracy d= 2. Asα→1, thediagonalhop- ping component is strengthened ( P→1/3) as the pair of bond–colored spin singlets resonates, until at α= 1 they become static and the degeneracy is d= 3. For the high–spin states in the regime η>ηc, atα= 0 one has two resonating (os/st) VBs, with the hopping channels unchanged and only the spin degeneracy d= 5. This state is not altered qualitatively for any α <0.92, a transition value independent of η. For 0.92< α < 1, orbital correlations are strongly suppressed and the state is characterized by hopping processes largely of the avoided–blocking type (one active orbital, Qdominant), still withd= 5. Finally, α= 1 represents the limit of a pure avoided–blocking state ( P= 0,Q= 2/3,R= 1/3), where the degeneracy jumps to 150, a number which can be understood as 5 (spin degeneracy) ×[6 (number of two–colorstates with no bonds requiring spin singlets) + 24 (number of three–color states with no bonds requiring spin singlets)]. E. Four–site chain As a fourth and final case, we present results from a linear four–site cluster. While not directly relevant to the study of the triangular lattice, this system offers further valuable insight into the intrinsic physics of the spin–orbital model. The cluster is oriented along the c–0 0.2 0.4 0.6 0.8 1 α−0.6−0.4−0.20.0Sij , Tij , Cij0 0.2 0.4 0.6 0.8 1−1.0−0.8−0.6−0.4−0.20.0En/J {a,b} orbitals c orbitals(ss/os) (ss/cc)(a) (b) resonating static1 1h FIG. 17: (Color online) Evolution of the properties of the four–site chain as a function of αatη= 0: (a) energy spec- trum and (b) spin ( Sij, filled circles), orbital ( Tij, empty cir- cles), and spin–orbital ( Cij,×) correlation functions. Both panels show a transition occurring at a level crossing at α= 4/7. In panel (a), the labels show a nondegenerate ground state ( d= 1) in both regimes, which has predimi- nantly spin singlet character at α >0.571, but both spin and orbital singlet components at α <0.571. In panel (b), Sij=Tij=Cij=−1 4forα <0.571 due to a resonating ( ab) orbital configuration, while Tij=Cij= 0 forα >0.571 as a consequence of the static corbital configuration. axis with periodic boundary conditions. As for the single bond (Sec. VB), only the aandborbitals contribute at α= 0, where indeed one finds average electron densities per sitenia=nib=1 2, andnic= 0. Likewise, at α= 1 only thecorbitals are occupied, with nic= 1, a result dictated by the spin singlet correlations, which are fully developed only for complete orbital occupation. The energy per bond for the four–site chain in the superexchange limit is again −J, as for a single bond [Fig. 17(a)]: somewhat surprisingly, the bonds do not ”disturb” each other, and joint spin–orbital fluctuations extend over the entire chain. However, in contrast to a single bond, this behavior is due to only one quan- tum state, the SU(4) singlet. In this geometry, only one SU(2) orbital subsector is selected, and the result- ingSU(2) ⊗SU(2) systemislocatedpreciselyattheSU(4) point ofthe Hamiltonian.50Thus, exactly asin the SU(4) chain, all spin, orbital and spin–orbital correlation func- tions are equal, Sij=Tij=Cij=−0.25, as shown in Fig. 17(b). For SijandTij, this result may be under-30 stoodasan averageoverequalprobabilitiesofsinglet and triplet states on each bond. In more detail, the condition set on the correlation functions by SU(4) symmetry12is 4 3/an}bracketle{t(/vectorSi·/vectorSj)(/vectorTic·/vectorTjc)/an}bracketri}ht=Sij=Tij, an equality also obeyed by the single bond (Sec. VB). The product of Sijand Tijin its definition ensures the identity for Cij. The unique ground state is nevertheless a linear superposi- tion of states expressed in the spin and orbital bases, and has not only finite but maximal entanglement. This state persists, with a perfectly linear α–dependence, all the way to α= 1, but ceases to be the ground state atα=4 7[Fig. 17(a)], where there is a level–crossing with theα= 1 ground state (also perfectly linear). This latter state has a completely different, fluctuation–free orbital configuration, with pure c–orbital occupation at every site, and gains energy solely in the direct–exchange channel. The spins and orbitals are decoupled, Tijand Cijvanish, and the spin state has Sij=−0.50: this re- sult can be understood as an equal average over bond states with /vectorSi·/vectorSj=−3 4and−1 4, and matches that ob- tained for the four–site AF Heisenberg model with a res- onating VB (RVB) ground state.2The energy at α= 1, E0=−0.75J[Fig. 17(a)], is given directly by including the constant term, −1 4Jper bond, in the definition of the Hamiltonian (2.21). The results for the linear four–site cluster demonstrate again the competition between superexchange and direct exchange. The orbital fluctuations arising due to the mixed exchange term, Hm(2.22), are responsible for re- moving the high degeneracies of the eigenenergies in the limitsα= 0 andα= 1 [Fig. 17(a)]. In fact the spectrum of the excited states is quasi–continuous in the regime aroundα= 0.5, but has a finite spin and orbital gap ev- erywhereother than the quantum critical point at α=4 7. These chain results raise a further possibility for the spontaneous formation at α= 0 of a 1D state not dis- cussed in Sec. III. A set of (for example) c–axis chains, with onlyaandborbitalsoccupied in the pseudospinsec- tor, would createexactly the 1D SU(4) model, and would therefore redeem an energy E=−3 4Jper bond from the formation of linear, four–site spin–orbital singlets. The energy of the triangular lattice would receive a further, constant contribution from the cross–chain bonds, which wascalculatedin Eq.(3.9) forgeneral η, andhence would be given at η= 0 by ESU(4) 1D(0) =−1 3·3 4J−1 6J=−5 12J. (5.9) This energy represents a new minimum compared with all of the results in Sec. III. That it was obtained from a melting of both spin and orbital order confirms the conclusion that ordered phases are inherently unstable in this class of model, being unable to provide sufficient energy to compete with the kinetic energy gains avail- able through resonance processes. That its value is now lower than that obtained for a static, 2D dimer covering (Sec. IV) is not of any quantitative significance, given the results of Sec. V confirming the importance of the0.0 0.2 0.4 0.6 0.8 1.0 α−0.8−0.6−0.4−0.20.0E0/J FIG. 18: (Color online) Ground–state energy per bond as a functionof α, obtainedwith η= 0for atriangular clusterwith 3 bonds (blue, dashed line), and a tetrahedral cluster with 6 bonds (red, solid line). For comparison, the energies obtai ned from the VB ansatz in the limiting cases α= 0 and α= 1 are shown for the triangular cluster (blue, diamonds) and tetrahedral cluster (red, yellow–filled, open circles); at α= 0 both VB energies are the same, while at α= 1 they match the exact solutions. Green, upward–pointing triangles sho w the static–dimer results of Sec. IV for the extended system, and the black, dot–dashed line the lowest energy per bond obtained for fully spin and orbitally ordered phases in Sec. III. The violet, downward–pointing triangle shows the energy of the orbitally ordered but spin–disordered Heisenberg–cha in state at α= 0 [Eq. (3.9)] and the open, yellow–filled square that of the analogous state at α= 1 [Eq. (3.22)], while the cross shows the energy of the spin– and orbitally disordered , SU(4)–chain state [Eq. (5.9)]. positional resonance of dimers. F. Summary To summarize, we have shown in this section the re- sults of exact numerical diagonalization calculations per- formed on small clusters. Detailed analysis of ground– state energies, degeneracies, site occupancies and a num- ber of correlation functions can be used to extract valu- able information about the local physics of the model across the full regime of parameters. Essentially all of the quantities considered show strong local correlations and the dominance of quantum fluctuations of the short- est range, with ready explanations in terms of resonating dimer states. We draw particular attention to the extremely low ground–state energy of the triangular cluster, which shows large gains from dimer resonance. The tetrahedral cluster also has a very significant resonance contribution, although more of its ground–state energy is captured at31 the level of a static dimer model. Such a VB ansatz provides the essential framework for the understanding of all the results obtained, even for systems with odd site numbers. The energies and their evolution with α contain some quantitative contrasts between even– and odd–site systems, allowing further insight concerning the range over which the qualitative features of the cluster results extend. Focusing in detail upon these energies, Fig. 18 sum- marizes the exact diagonalization results at zero Hund coupling, and provides a comparison not only with the VB ansatz, but with all of the other results obtained in Secs. III–V. From bottom to top are shown: the ex- act cluster energies including all physical processes; the clusterVB ansatz, showingthe importance ofdimer reso- nance energy; the static VB ansatz for extended systems, suggestingbycomparisonwithclusterstheeffectsofreso- nance; the energies of “melted” states with 1D spin (and orbital) correlations; the optimal energy of states with full, long–ranged spin and orbital order. Returning to the cluster results, their degeneracies can be understood precisely, and demonstrate the restora- tion of various symmetries due to resonance processes. We provide a complete explanation for all the correla- tion functions computed, and use these to quantify the entanglement as a function of α,ηand the system size. There is a high–spin transition as a function of ηfor all values ofα, which sets the basic phase diagram and es- tablishes a new set of disentangled orbital models at high η. The extrapolation of the cluster results to states of extended systems, some approximations for which are shown in Fig. 18, is not straightforward, and cannot be expected to include any information relevant to subtle selection effects within highly degenerate manifolds of states. However, with the exception of the static–dimer regime around α= 1, our calculations suggest that noth- ing subtle is happening in this model over the bulk of the phase diagram, where the physics is driven by large energetic contributions from strong, local resonance pro- cesses. VI. RHOMBIC, HONEYCOMB, AND KAGOME LATTICES In Sec. I we alluded to the question of different sources offrustrationincomplexsystemssuchasthe spin–orbital model of Eq. (2.7). More specifically, this refers to the relative effects of pure geometrical frustration, as under- stood for AF spin interactions, and ofinteraction frustra- tion of the type which can arise in spin–orbital models even on bipartite lattices.7Because the interaction frus- tration depends in a complex manner on system geom- etry, no simple separation of these contributions exists. In this section we alter the lattice geometry to obtain somequalitativeresultswithabearingonthisseparation, by considering the same spin–orbital model on the three(a) (b) (c) FIG. 19: (Color online) (a) Rhombic lattice, showing a two– color orbitally ordered state. (b) Honeycomb lattice, show ing a one–color orbitally ordered state. (c) Kagome lattice, sh ow- ing a three–color orbitally ordered state. simple lattice geometries which can be obtained from the triangular lattice by the removal of active bonds or sites. The geometries we discuss are rhombic, obtained by removing all bonds in one of the three triangular lat- tice directions [Fig. 19(a)], honeycomb, or hexagonal, ob- tained by removing every third lattice site [Fig. 19(b)], and kagome, obtained by removing every fourth lattice site in a 2 ×2 pattern [Fig. 19(c)]. Simple geometrical frustration is removed in the rhombic and honeycomb cases, but for Heisenberg spin interactions the kagome geometry is generally recognized (from the ground–state degeneracy of both classical and quantum problems) to be even more frustrated than the triangular lattice. We consider only the α= 0 andα= 1 limits of the model, andη= 0. We discuss the results for long–range–ordered states (Sec. III) and for static dimer states (Sec. IV) for all three lattice geometries. Here we do not enter into nu- merical calculations on small clusters, and comment only on those systems for which exact diagonalization may be expected to yield valuable information not accessible by analytical considerations. A. Rhombic lattice While the connectivity of this geometry is precisely that of the square lattice, we refer to it here as rhombic to emphasize the importance of the bond angles of the32 (a) (b) FIG. 20: (Color online) Rhombic lattice with (a) columnar and (b) plaquette dimer coverings. chemical structure in maintaining the degeneracy of the t2gorbitalsandindeterminingthenatureoftheexchange interactions. It is worth noting that the spin–orbital model (2.7) on this lattice may be realized in Sr 2VO4 (below). In the absence of geometrical frustration, the spin problem created by imposing any fixed orbital con- figuration selected from Sec. III (Figs. 3 and 4) is gen- erally rather easy to solve. Further, at η= 0 both FM and AF, and by extension AFF, spin states have equal energies, leading to a high spin degeneracy. Following Sec. III, the α= 0 energies for the majority of the orbitally ordered states of Fig. 3 are Erh lro(0) =−1 2J (6.1) per bond at η= 0 for a number of possible spin con- figurations, whose degeneracy is lifted (in favor of FM lines or planes) at finite η. Indeed, the only exceptions to this rule occur for the three–color state [Fig. 3(d)] and for orientations of the other states which preclude hopping in one of the two lattice directions, whose tri- angular symmetry properties are broken by the missing bond. As noted in Sec. III, for superpositions is it the exception rather than the rule for all hopping processes to be maximized, but on the rhombic lattice this is pos- sible for the states in Fig. 4(a) and some orientations of those in Figs. 4(d) and 4(e). Forα= 1, the energy limit even on the triangular lat- tice wasset ratherby the number ofactive bonds than by the problem of minimizing their frustration. Similar to theα= 0 case, all states where the active hopping direc- tion is one of the two lattice directions, plus in this case state (3d), can redeem the maximum energy available, Erh lro(1) =−1 4J (6.2) atη= 0, which is simply the avoided–blocking energy, for a large number of possible spin configurations. Finite Hund exchange favors FM spin states. Turning to dimerized states, the calculation of the en- ergy of any given dimer covering proceeds as in Sec. IV, namely by counting for each the respective numbers of bonds of types (7a), (7b), and (7c) [Fig. 7]. For the rhombic lattice, lack of geometrical frustration meansthat all interdimer bonds can be chosen to be AF/AO. The two most regular dimer coveringsof the rhombic lat- tice with small unit cells may be designated as “colum- nar” [Fig. 20(a)] and “plaquette” [Fig. 20(b)]. In both cases, 1/4 of the bonds are the dimers, and by inspection 1/4 of the interdimer bonds in the columnar state are of type (7a), while the remainder are (7c); by contrast, the plaquette state has no type–(7a) bonds, 1 /2 type–(7b) bonds, and the remainder are of type (7c). For α= 0, the energies are Erh dc(0) =−1 4J−1 4·1 2J−1 2·1 4J=−1 2J, Erh dp(0) =−1 4J−1 2·3 8J−1 4·1 4J=−1 2J(6.3) atη= 0, both for (ss/ot) and for (os/st) dimers. The de- generacyofthese two limiting cases, in the sense of maxi- malandminimalnumbersoftypes–(7a)and–(7b)bonds, suggests a degeneracy of all dimer coverings at this level of analytical sophistication. Further, all of these dimer coverings are degenerate with all of the unfrustrated or- dered states at η= 0. The selection of a true ground state from this large manifold of static states (order– by–disorder) would hinge on higher–order processes, but these considerations are likely to be rendered irrelevant by dimer resonance (Sec. V). For the spin–singlet dimer states at α= 1 one finds Erh dc(1) =−1 4J−1 4·1 4J−1 2·0J=−5 16J, Erh dp(1) =−1 4J−1 2·1 4J−1 4·0J=−3 8J,(6.4) atη= 0, and thus that, as for the triangular lattice, the energy is minimized by dimer configurations excluding linear interdimer bonds. This remains a large manifold ofdimer coverings,whose energyis manifestly lowerthan any of the possible orbitally ordered states in this limit of the model, and within which order–by–disorder is ex- pected to operate (Sec. V).49 The considerations of this subsection, extended to fi- nite values of η, may be relevant in the understanding of experimentalresultsforSr 2VO4. ThesesuggestweakFM order,51accompanied by an AO order52which could be interpreted as arising from the formation of dimer pairs. When the oxygen octahedra distort, the threefold degen- eracy of the t2gorbitals is lifted, to give a model con- taining only two degenerate orbitals, dyzanddxz. This leads to a situation with Ising–like superexchange inter- actionsand quasi–1Dholepropagationin aneffective t–J model.53 B. Honeycomb lattice The situation for the honeycomb lattice is very simi- lar to that for the rhombic case. Again the absence of geometrical frustration makes it possible to obtain the33 (a) (b) FIG. 21: (Color online) Honeycomb lattice with (a) columnar and (b) three–way dimer coverings. minimal energy for a number of orbital orderings, with a high spin degeneracy at η= 0. For pure superexchange interactions, once again Eh lro(0) =−1 2J (6.5) per bond, while in the direct–exchange limit Eh lro(1) =−1 4J, (6.6) both atη= 0, for the same physical reasons as above. For dimer states, on the honeycomb lattice all inter- dimer bonds are by definition of type (7c), and again can be made AF/AO because frustration is absent, so the energies of all dimer coverings are de facto identical. By way of demonstration, the two simplest regular configu- rations, which we label “columnar”and “three–way”,are shown in Fig. 21, and, from the fact that now 1 /3 of the bonds contain dimers, their energies are Eh dc(0) =−1 3J−2 3·1 4J=−1 2J, Eh d3(0) =−1 3J−2 3·1 4J=−1 2J,(6.7) per bond at α= 0 =η. Thus static dimer states are again degenerate with unfrustrated ordered states in thesuperexchangelimit, anddetailedconsiderationofkinetic processes would be required to deduce the lowest total energy. In this context, the dimer coverings shown in Fig. 21 exemplify two limits about which little kinetic energy can be gained from resonance (Fig. 21(a), where large numbers of dimers must be involved in any given process)and in which kinetic energy gainsfrom processes involving short loops [the three dimers around 2/3 of the hexagons, Fig. 21(b)] are maximized. Atα= 1, only the dimer energy is redeemed, and this on 1/3 of the bonds, so Eh d(1) =−1 3J (6.8) atη= 0 for a large manifold of coverings. This energy is once again significantly better than any of the possi- ble ordered states, a result which can be ascribed to the low connectivity. That the ground state of the extended system in this limit for both the rhombic and honeycomb lattices involves a selection from a large number of nearly degeneratestates suggeststhat numerical calculationson small clusters would not be helpful in resolving detailed questions about its nature. The same model for the hon- eycomb geometry in the α= 1 limit has been discussed for theS= 1 compound Li 2RuO3,54where the authors invoked the lattice coupling, in the form of a structural dimerization driven by the formation of spin singlets, to select the true ground state. C. Kagome lattice The kagome lattice occupies something of a special place among frustrated spin systems1as one of the most highly degenerate and intractable problems in existence, for both classical and quantum spins, and even with only nearest–neighbor Heisenberg interactions. Inter- est in this geometry has been maintained by the dis- covery of a number of kagome spin systems, and has risen sharply with the recent synthesis of a true S= 1/2 kagome material, ZnCu 3(OH)6Cl2.55Preliminary local– probe experiments56,57show a state of no magnetic or- der and no apparent spin gap, whose low–energy spin excitations have been interpreted58as evidence for an exotic spin–liquid phase. Both experimentally and the- oretically, kagome systems of higher spins ( S= 3/2 and 5/2) are found to have flat bands of magnetic excita- tions, reflecting the very high degeneracy of the spin sector.59While no kagome materials are yet known with both spin and orbital degrees of freedom, Maekawa and coworkers40,44have considered the itinerant electron sys- tem on the triangular lattice for α= 0 (actually for the motion of holes in Na xCoO2), demonstrating that the combination of orbital, hopping selection, and geometry leads to any one hole being excluded from every fourth site, and thus moving on a system of four interpenetrat- ing kagome lattices.34 (a) (b) FIG. 22: (Color online) Kagome lattice with unequally weighted two–color states oriented (a) with and (b) against the lattice direction corresponding to the majority orbita l color. Considering first the energies per bond for states of long–ranged spin and orbital order, in a number of cases the values for the kagome lattice are identical to those of the triangular lattice. This is easy to show by inspection for the one–color state (3a), and for the superposition states (4a), (4b), and (4c), where bonds of all types are removed in equal number. However, for the less symmet- rical orbital color configurations a more detailed analysis of the type performed in Sec. III is required, and yields provocative results. The two simple possibilities for or- dered two–color states with a single color per site are shown in Fig. 22, and differ only in the orientation of the continuous lines (the majority color) relative to the active orbitals. These can be considered as the kagome– lattice analogs of states (3b) and (3c), as well as of (3e) and (3f). When the lines of c–orbitalsarealignedwith the c–axis [Fig. 22(a)], this direction is inactive at α= 0, and only the other two directions contribute, one with two active FO orbitals, mandating an AF spin state to give energy −1 2Jper bond, and the other with energy −1 4Jand no strong spin preference, whence E(k3b)(0) =−1 4J (6.9) atη= 0 for sets of unfrustrated AF chains. By contrast, when the lines of c–orbitals fall along the b–direction [Fig. 22(b)], the α= 0 problem contains one FO and one AO line each with two active orbitals, and one line with one active orbital. Only the first requires AF spin alignment, while the other two lines are not frustrating, with the result that an energy E(k3c)(0) =−5 12J (6.10) can be obtained. This value is lower than that on the triangular lattice, showing that for the class of models under consideration, where not all hopping channels are activein alldirections, asystem oflowerconnectivitycan lead to frustration relief even when its geometry remains purely that of connected triangles.(a) (b) FIG. 23: (Color online) Kagome lattice with two different, equally weighted three–color states: (a) two–color lines o ri- ented such that only one superexchange channel, plus the di- rect exchange channel, is active on every bond. (b) two–colo r lines oriented such that all superexchange channels are act ive, but no direct exchange channels. With this result in mind, we consider again the possi- bilities offered by different three–color states, specifically those shown in Fig. 23. With reference to the superex- change problem, the state in Fig. 19(c), which by anal- ogy with (3d) we denote as (k3d), contains only a small numberofremnanttrianglesand isolatedbonds still with two active orbitals. However, the state (k3d1), shown in Fig. 23(a) is that which ensures that no such bonds re- main, and every single bond of the lattice has one active superexchange channel. The state (k3d2) in Fig. 23(b) is that in which every single bond of the lattice has two active (FO) superexchange channels: this possibility can be realized for the kagome geometry, at the cost of creat- ing a frustrated magnetic problem requiring a 120◦spin state to minimize the energy, E(k3d)(0) =−5 16J, (6.11) E(k3d1)(0) =−1 4J, (6.12) E(k3d2)(0) =−3 8J. (6.13) Thus one finds that lower energies than the value −1 3J per bond, which was the lower bound for fully (or- bitally and spin–)ordered states on the triangular lattice, are again possible for three–color ordered states. How- ever, the residual spin frustration means that the lowest ordered–stateenergyonthekagomelatticeisgivenbythe unfrustrated, two–color AFF state, E(k3c)(0) =−5 12J. We present briefly the energies of the same states at α= 1,whereonlyamaximumofonehoppingchannelper bond can be active, and as noted abovethis is generallya stricter energetic limit than any frustration constraints. The results at η= 0 are E(k3b)(1) =−1 4J (6.14) for an AFF state gaining most of its energy from the35 c–axis chains, and E(k3c)(1) =−1 12J (6.15) due to the dearth of active orbitals in this orientation. Similarly, by counting active orbitals in the three–color states, E(k3d)(1) =−1 6J, (6.16) E(k3d1)(1) =−1 4J, (6.17) E(k3d2)(1) = 0, (6.18) and it is the state of Fig. 23(a) which achieves the un- frustrated value −1 4Jby permitting one active hopping channel on every bond of the kagome lattice. We will not discuss the orbital superposition states which are the analogs of (4d) and (4e), noting only that these present again two different possibilities on the kagome lattice, depending on the orientation of the ma- jority lines. Even with the frustration relief offered by this geometry for the type of model under consideration, superposition states contain too many hopping channels for all to be satisfied simultaneously, and it is not possi- ble to equal the energy values found respectively for the configurations in Figs. 23(a) and (b) at α= 1 andα= 0. It remains to consider dimer states on the kagome lat- tice, as these have been of equal or lower energy for ev- ery case analyzed so far. The set of nearest–neighbor dimer coverings of the kagome lattice is large, and for theS= 1/2 Heisenberg model in this geometry the spin singlet manifold has been proposed as the basis for an RVB description.25Two dimer coverings degenerate at the level of the current treatment are shown in Fig. 24. Dimer coverings of the kagome lattice have the prop- erty that 3 /4 of the triangles contain one dimer. In this case, the other bonds of the triangle are interdimer bonds, one of which is of type (7b) while the other is of type (7c). The other 1/4 of the triangles, known60as “defect triangles”, have no dimers, and their three bonds are either all of type (7b), with probability 1/4, or one each of types (7a), (7b), and (7c), with probability 3/4. The frustration of the system is contained in the problem of minimizing the number of FM/FO interdimer bonds; this exercise is complex and no solution is known, so only an upper bound will be estimated here. The bonds of a defect triangle connect three differ- ent dimers, and so one (or all three) must be FM/FO. A hexagonofthekagomelatticewithnodimersonitsbonds issurroundedbysixnon–defectivetriangles,onewithone dimer by one defective neighbor, with two dimers two, and a hexagon with three dimers shares its non–dimer bonds with three defect triangles. Hexagons with odd dimer numbers must create a FM/FO bond between at least one pair of dimers, and it is reasonable to place this bond on the defect triangle(s) where an energy cost is al- ready incurred. We note immediately that the cost of re- versing the type–(7a) bond,1 4J(Sec. IVA), exceeds that(a) (b) FIG. 24: (Color online) Kagome lattice with two different dimer coverings, (a) and (b). In both examples, only two of the twelve triangles shown explicitly on the cluster are “de fec- tive”(contain nodimer), butthereader maynotice thatmany of the next twelve triangles adjoining the boundary must als o be so. of reversingboth interdimer bonds of a non–defective tri- angle, which is1 8J+1 16J. As a consequence, we take this cost, which is equal to that of reversing both a non– defective triangle and the weakest bond of the defect tri- angle, to be an upper bound on the effect of frustration. The net energy of a dimer state for α= 0 =ηis then estimated to be Ekd(0) =−3 4·1 3J−3 4·1 3J/parenleftbigg3 8+1 4/parenrightbigg −1 4J/bracketleftbigg1 4/parenleftbigg2 3·3 8+1 3·1 4/parenrightbigg +3 4·1 3/parenleftbigg1 4+3 8+1 4/parenrightbigg/bracketrightbigg =−209 384J≃ −13 24J. (6.19) This is a very large number for the kagome lattice, ex- ceeding even the value −1 2Jper bond (which, however, is of no special significance here). Thus we find that dimer states in this type of model are strongly favored, gaining a very much higher energy than even the best ordered states. Qualitatively, the dimer energy shares with the ordered–state energy the feature that it is considerably36 betterthananythingobtainableforthetriangularlattice. This implies that the reduced connectivity of the lattice geometry for a model where the orbital degeneracy pro- vides a number of mutually exclusive hopping channels makesit easier to find states where every remainingbond can support a favorable hopping process without strong frustration. Applying all of the above geometrical considerations to the direct–exchange model ( α= 1), where there is no frustration problem between the spin singlets, one finds Ekd(1) =−3 4·1 3J−3 4·1 3J/bracketleftbigg1 4+0/bracketrightbigg −1 4·1 3J/bracketleftbigg1 4·3 4+1 4·1 2/bracketrightbigg =−21 64J (6.20) atη= 0. Once again this energy is significantly lower than the value Edim(1) =−1 4Jobtained for the triangu- lar lattice in Eq. (4.15), demonstrating that the multi- channel spin–orbital model of the type considered here is less frustrated in the kagome geometry. Wecommentinclosingthatthe dimerenergieswehave estimatedareonlythoseofstaticVBconfigurations,and, away from α= 1, the possibility remains of a signifi- cant resonance energy gain from quantum fluctuations between these states ( cf.Sec. V). Numerical calculations on small clusters of sufficient size (here at least 6 sites for a unit cell) would be helpful in this frustrated case. To summarize this section, the spin–orbital model on bipartite lattices appears to present competing ordered and dimerized states with the prospect of high degenera- cies. Among “frustrated” systems (in the sense of being non–bipartite), the kagome lattice provides an example where geometrical and orbital frustration effects cancel partially, affording favorable dimerized solutions. Thus, while it is possible to ascribe some of the frustration ef- fects we have studied in the triangular lattice to a purely geometrical origin, for more complex models it is in gen- eralnecessaryto extend the concept of “geometricalfrus- tration” beyond that applicable to pure spin systems. VII. DISCUSSION AND SUMMARY We have considered a spin–orbital model representa- tive of a strongly interacting 3 d1electron system with the cubic structural symmetry of edge–sharing metal– oxygen octahedra, conditions which lead to a triangular lattice of magnetic interactions between sites with un- broken, threefold orbital degeneracy. We have elucidated the qualitativephasediagram,whichturns outto be very rich, in the physicalparameterspace presented by the ra- tio (α) of superexchange to direct–exchange interactions and the Hund exchange ( η). Despite the strong changes in the fundamental nature of the model Hamiltonian as a function of αandη, anumber of generic features persist throughout the phase diagram. With the exception ofthe ferromagneticphases at highη, which effectively suppressesquantum spin fluc- tuations (below), there is no long–rangedmagnetic or or- bital order anywhere within the entire parameter regime. This shows a profound degree of frustration whose origin lies both in the geometry and in the properties of the spin–orbital coupling; a qualitative evaluation of these respective contributions is discussed below. All of the phases ofthe model show a strong preference for the formation of dimers. This can be demonstrated in a simple, static valence–bond (VB) ansatz, and is re- inforced by the results of numerical calculations. The staticansatzis alreadyanexactdescriptionofthe direct– exchange limit, α= 1, and gives the best analytic frame- work for understanding the properties of much of the re- mainder of the phase diagram. The most striking single numerical result is the prevalence of VB states even on a triangular cluster, and the underlying feature reinforced by all of the calculations is the very large additional “ki- netic” contribution to the ground–state energy arising from the resonance of VBs due to quantum fluctuations. It is this resonance which drives symmetry restoration in some or all of the spin, orbital, and translational sectors over large regions of the parameter space. The sole ex- ception to dimerization is found at high ηand around α= 1, where the only mechanism for virtual hopping is the adoption of orbital configurations which permit one orbital to be active (“avoided blocking”). The “most exotic” region of the phase diagram is that at smallαandη, and this we have assigned tentatively as an orbital liquid. In this regime, quantum fluctua- tions are at their strongest and most symmetrical, and every indication obtained from energetic considerations of extended systems, and from microscopic calculations of a range of local quantities on small clusters, suggests a highly resonant, symmetry–restored phase. While this orbital liquid is in all probability (again from the same indicators) based on resonating dimers, an issue we dis- cuss in full below, we cannot exclude fully the possibility of a type of one–dimensional physics: short, fluctuating segments of frustration–decoupled spin or orbital chains, whose character persists despite the high site coordina- tion. It should be stressed here that the point ( α,η) = (0,0)isnotinanysenseaparentphaseforexoticstatesin the restofthe phasediagram: mixed and direct exchange processes are qualitatively different elements, which in- troduce different classes of frustrated model at finite α. While the matter is somewhat semantic, we comment only that one cannot argue for the point α= 0.5 being “more exotic” than α= 0 despite having the maximal number of equally weighted hopping channels, because it does not possess any additional symmetries which man- date qualitative changes to the general picture. In this sense, the limit α= 1 serves as a valuable fixed point which is understood completely, and yet is still domi- nated by the purely quantum mechanical concept of sin- glet formation.37 One indicatorwhich can be employed to quantify “how exotic” a phase may be is the entanglement of spin and orbital degrees of freedom. We define entanglement as the deviation of the spin and orbital sectors from the factorized limit in which their fluctuations can be treated separately. We compute a spin–orbital correlation func- tion and use it to measure entanglement, finding that this is significant over the whole phase diagram. Qual- itatively, entanglement is maximal around the superex- change limit, which is dominated by dimers where sin- glet formation forces the other sector to adopt a local triplet state. However, for particular clusters and dimer configurations, the high symmetry may allow less entan- gled possibilities to intervene exactly at α= 0. The direct–exchange limit, α= 1, provides additional in- sight into the entanglement definition: the four–operator spin–orbital correlation function vanishes, reflecting the clear decoupling of the two sets of degrees of freedom at this point, but the finite product of separate spin and orbital correlation functions violates the factorizability condition. This preponderance of evidence for quantum states based on robust, strongly resonating dimers implies fur- ther that the (spin and orbital) liquid phase is gapped. Such a state would have only short–ranged correlation functions. However, these gapped states are part of a low–energy manifold, and for the extended system we have shown that this consists quite generally of large numbers of (nearly) degenerate states. The availability of arbitrary dimer rearrangements at no energy cost has been suggested to be sufficient for the deconfinement of elementary S= 1/2 (and by analogy T= 1/2) excita- tions with fractional statistics.61However, the spinons (orbitons) are massive in such a model, in contrast to the properties of algebraic liquid phases.62 Alow–spintohigh–spintransition,occurringasafunc- tion ofη, is present for all values of α. The quantitative estimation of ηcin the extended system remains a prob- lem for a more sophisticated analysis. At the qualitative level, largeηcan be considered to suppressquantum spin fluctuations by promoting parallel–spin (ferromagnetic) intermediate states on the magnetic ions. However, even when this sector is quenched, the orbital degrees of free- dom remain frustrated, and contain non–trivialproblems in orbital dynamics. In the superexchange (low– α, high– η) region, frustration is resolved by the formation of or- bital singlet (spin triplet) dimers, whose resonance min- imizes the ground–state energy. The frustration in the direct–exchange (high– α, high–η) region is resolved by avoided–blocking orbital configurations, and order–by– disorder effects are responsible for the selection of the true ground state from a degenerate manifold of possi- bilities; this is the only part of the phase diagram not displaying dimer physics. Thus the ferromagnetic orbital models in both limits exhibit a behavior quite different from that of systems with only S= 1/2 spin degrees of freedom on the triangular lattice. We have commented on both geometry and spin–orbital interactions as the origin of frustration in the models under consideration. However, a statement such as “on the triangular lattice, geometrical frustration en- hances interaction frustration for spin–orbital models” must be qualified carefully. We have obtained anecdotal evidence concerning such an assertion in Sec. VI by con- sideringotherlatticegeometries,andfind thatindeed the same model on an unfrustrated geometry appears capa- ble of supporting ordered states; however, the interplay of the two effects is far from direct, as the kagome lattice presents a case where dimer formation acts to reduce the net frustation. Quite generally, spin–orbital models con- tain in principle more channels which can be used for re- lievingfrustration,buttheexactnatureofthecouplingof spin and orbital sectors may result in the opposite effect. Specificdatacharacterizingmutualfrustrationcanbeob- tained from the spin and orbital correlations computed on small clusters: as shown in Sec. V, for the triangular lattice there are indeed regimes where, for example, the effective orbital interactions enforced by the spin sector make the orbital sector more frustrated (higher Tij) than would be the analogous pure spin problem (measured by Sij), and conversely. We comment briefly on other approaches which might be employed to obtain more insight into the states of the extended system, with a view to establishing more defini- tively the nature and properties of the candidate orbital liquid phase. More advanced numerical techniques could be used to analyze larger unit cells, but while Lanczos diagonalization, contractor renormalization63or other truncation schemes might afford access to systems two, or even four, times larger, it seems unlikely that these clusters could provide the qualitatively different type of data required to resolve the questions left outstanding in Sec. V. An alternative, but still non–perturbative and predominantly unbiased, approach would be the use of variational wave functions, either formulated generally or in the more specific projected wave function technique which leads to different types of flux phase.64,65Adapt- ing this type of treatment to the coupled spin and orbital sectorswithout undue approximationremainsa technical challenge. Within the realm of effective models which could be obtained by simplification of the ground–state manifold, we cite only the possibility motivated by the current re- sults of constructing dimer models based on (ss/ot) and (os/st) dimers. Dimer models26are in general highly simplified, and there is no systematic procedure for their derivation from a realistic Hamiltonian, but they are thought to capture the essential physics of certain classes of dimerized systems. Because QDM Hamiltonians pro- vide exact solutions, and in some cases genuine exam- ples of exotica long sought in spin systems, including the RVB phase and deconfined spinon excitations, they rep- resent a valuable intermediate step in understanding how such phenomena may emerge in real systems. Here we have found (i) a very strong tendency to dimer forma- tion, (ii) a large semi–classical degeneracy of basis states38 formed from these dimers, and (iii) that resonance pro- cesses even at the four–site plaquette scale providea very significant energetic contribution. From the final obser- vation alone, a minimal QDM, meaning only exchange of parallel dimers of all three directions and on all possible plaquette units, would already be expected to contain the most significant corrections to the VB energy. At this point we emphasize that, because of the change of SU(2) orbitalsectorwith lattice direction, our2Dmodels are not close to the SU(4) point where four–site plaque- tte formation, and hence very probably a crystallization, would be expected.13From the results of Secs. IV and V, a rather more likely phase of the QDM would be one with complete plaquette resonance through all three col- ors, and without breaking of translational symmetry. Rigorous proof of a liquid phase, such as that repre- sented by an RVB state, is more complex, and as noted in Sec. I it requires satisfying both energetic and topo- logical criteria. Following the prescription in Ref. 19, three conditions must be obeyed: (i) a propensity for dimer formation, (ii) a highly degenerate manifold of ba- sis states from which the RVB ground state may be con- structed, and (iii) a mapping of the system to a liquid phase of a QDM. Criteria (i) and (ii) match closely the labels in the previous paragraph, and both dimer forma- tion and high degeneracy have been demonstrated ex- tensively here. The energetic part of criterion (iii) also appears to be obeyed here: static dimers have an en- ergy (V), and allowing their location and orientation to change gains more ( t). The regime V/t <1 of the triangular–lattice QDM is the RVB phase demonstrated in Ref. 27, whose properties include short–range corre- lation functions and gapped, deconfined spinons. This mapping also contains the criterion of togological degen- eracy, andcouldinprinciplebe partiallycircumventedby a direct demonstration. However, no suitable numericalstudies are available of non–simply connected systems, and so here we can present only plausibility arguments based on the high degeneracy and spatial topology of the dimer systems analyzed in Secs. IV and V. It is safe to concludethatthe threefold–degenerate t2gorbitalsystem on the triangular lattice is one of best candidates yet for a true spin–orbital RVB phase. In closing, spin–orbital models have become a frontier of intense current interest for both experimental and the- oretical studies of novel magnetic and electronic states emerging as a consequence of intrinsic frustration. Our model has close parallels to, and yet crucial differences from, similar studies of manganites (cubic systems of eg orbitals), LiNiO 2(triangular, eg), YTiO 3and CaVO 3 (cubic,t2g), and many other transition–metal oxides, ap- pearing in some respects to be the most frustrated yet discussed. One of its key properties, arising from the ex- treme(geometricalandinteraction–driven)frustration,is that ordered states become entirely uncompetitive com- pared to the resonance energy gained by maximizing quantum (spin and orbital) fluctuations. In the orbital sector, the restoration of symmetry by orbital fluctua- tions makes the model a strong candidate to display an orbital liquid phase. Because this liquid is based on ro- bustdimerstates, the mechanismforitsformationisvery likely to be spin–orbital RVB physics. Acknowledgments We thank G. Khaliullin and K. Penc for helpful discussions, and J. Chaloupka for technical assistance. A. M. Ole´ s acknowledges support by the Foundation for Polish Science (FNP) and by the Polish Ministry of Sci- enceandEducationunderProjectNo.N20206832/1481. 1Frustrated Spin Systems , edited by H. T. Diep (World Sci- entific, Singapore, 2004). 2P. Fazekas, Lectures on Electron Correlation and Mag- netism(World Scientific, Singapore, 1999). 3J. Zaanen, G. A. Sawatzky, and J. W. Allen, Phys. Rev. Lett.55, 418 (1985). 4A. M. Ole´ s, P. Horsch, G. Khaliullin, and L. F. Feiner, Phys. Rev. B 72, 214431 (2005). 5K. I. Kugel and D. I. Khomskii, Usp. Fiz. Nauk 136, 621 (1982) [Sov. Phys. Usp. 25, 231 (1982)]. 6C. Castellani, C. R. Natoli, and J. Ranninger, Phys. Rev. B18, 4945 (1985); Phys. Rev. B 18, 4967 (1985); Phys. Rev. B18, 5001 (1985). 7L. F. Feiner, A. M. Ole´ s, and J. Zaanen, Phys. Rev. Lett. 78, 2799 (1997). 8J. B. Goodenough, Magnetism and the Chemical Bond (In- terscience, New York, 1963); J. Kanamori, J. Phys. Chem. Solids10, 87 (1959). 9A. M. Ole´ s, P. Horsch, L. F. Feiner, and G. Khaliullin, Phys. Rev. Lett. 96, 147205 (2006).10Y. Q. Li, M. Ma, D. N. Shi, and F. C. Zhang, Phys. Rev. Lett.81, 3527 (1998). 11A. M. Ole´ s, P. Horsch, and G. Khaliullin, Phys.Stat. Solidi B244, 3478 (2007). 12B. Frischmuth, F. Mila, and M. Troyer, Phys. Rev. Lett. 82, 835 (1999); F. Mila, B. Frischmuth, A. Deppeler, and M. Troyer, ibid.82, 3697 (1999). 13K.Penc, M. Mambrini, P.Fazekas, andF. Mila, Phys. Rev. B68, 012408 (2003). 14G. Khaliullin and S. Maekawa, Phys. Rev. Lett. 85, 3950 (2000); G. Khaliullin, Phys. Rev. B 64, 212405 (2001). 15J. van den Brink, New J. Phys. 6, 201 (2004). 16G. Mihaly, I. Kezsmarki, F. Zamborszky, M. Miljak, K. Penc, P. Fazekas, H. Berger, and L. Forro, Phys. Rev. B61, R7831 (2000); P. Fazekas, K. Penc, K. Radnoczi, N. Barisic, H. Berger, L. Forro, S. Mitrovic, A. Gauzzi, L. Demko, I. Kezsmarki, and G. Mihaly, J. Magn. Magn. Mater.310, 928 (2007). 17P. Horsch, G. Khaliullin, and A. M. Ole´ s, Phys. Rev. Lett. 91, 257203 (2003).39 18F. Vernay, K. Penc, P. Fazekas, and F. Mila, Phys. Rev. B70, 014428 (2004); F. Vernay, A. Ralko, F. Becca, and F. Mila, ibid.74, 054402 (2006). 19F. Mila, F. Vernay, A. Ralko, F. Becca, P. Fazekas, and K. Penc, J. Phys.: Condens. Matter 19, 145201 (2007). 20S. Di Matteo, G. Jackeli, C. Lacroix, and N. B. Perkins, Phys. Rev. Lett. 93, 077208 (2004); S. Di Matteo, G. Jack- eli, and N. B. Perkins, Phys. Rev. B 72, 024431 (2005). 21S. Di Matteo, G. Jackeli, and N. B. Perkins, Phys. Rev. B 72, 020408(R) (2005). 22L. F. Feiner and A. M. Ole´ s, Phys. Rev. B 71, 144422 (2005). 23G. Khaliullin, Prog. Theor. Phys. Suppl. 160, 155 (2005). 24P. Fazekas and P. W. Anderson, Philos. Mag. 30, 423 (1974). 25F. Mila, Phys. Rev. Lett. 81, 2356, (1988); M. Mambrini and F. Mila, Eur. Phys. J. B 17, 651 (2000). 26D. S. Rokhsar and S. A. Kivelson, Phys. Rev. Lett. 61, 2376 (1988). 27R. Moessner and S. L. Sondhi, Phys. Rev. Lett. 86, 1881 (2001). 28A. Ralko, M. Ferrero, F. Becca, D. Ivanov, and F. Mila Phys. Rev. B 71, 224109 (2005). 29A. Weisse and H. Fehske, New J. Phys. 6, 158 (2004); E. Dagotto, New J. Phys. 7, 67 (2005); K. Ro´ sciszewski and A. M. Ole´ s, J. Phys.: Cond. Matter 19, 186223 (2007). 30P. Piekarz, K. Parlinski, and A. M. Ole´ s, Phys. Rev. Lett. 97, 156402 (2006); Phys. Rev. B 76, 165124 (2007). 31D. I. Khomskii and T. Mizokawa, Phys. Rev. Lett. 94, 156402 (2005). 32M. Holzapfel, S. de Brion, C. Darie, P. Bordet, E. Chappel, G. Chouteau, P. Strobel, A. Sulpice, and M. D. N´ u˜ nez– Regueiro, Phys. Rev. B 70, 132410 (2004). 33A. Reitsma, L. F. Feiner, and A. M. Ole´ s, New J. Phys. 7, 121 (2005). 34M. V. Mostovoy and D. I. Khomskii, Phys. Rev. Lett. 89, 227203 (2002). 35K. Hirakawa, H.Kadowaki, andK. Ubukoshi, J. Phys. Soc. Jpn.54, 3526 (1985). 36K. Takeda, K. Miyake, K. Takeda, and K. Hirakawa, J. Phys. Soc. Jpn. 61, 2156 (1992). 37H. F. Pen, J. van den Brink, D. I. Khomskii, and G. A. Sawatzky, Phys. Rev. Lett. 78, 1323 (1997). 38T. Motohashi, Y.Katsumata, T. Ono, R.Kanno, M. Karp- pinen, and H. Yamauchi, J. Appl. Phys. 103, 07C902 (2008). 39M. Itoh, M. Shikano, H. Kawaji, and T. Nakamura, Solid State Commun. 80, 545 (1991); Y. Imai, I. Solovyev, and M. Imada, Phys. Rev. Lett. 95, 176405 (2005). 40W. Koshibae and S. Maekawa, Phys. Rev. Lett. 91, 257003 (2003). 41J. S. Griffith, The Theory of Transition Metal Ions (Cam- bridge University Press, Cambridge, 1971). 42G. Khaliullin, P. Horsch, and A. M. Ole´ s, Phys. Rev. Lett. 86, 3879 (2001); A. M. Ole´ s, P. Horsch, and G. Khaliullin, Phys. Rev. B 75, 184434 (2007).43A. B. Harris, T. Yildirim, A. Aharony, O. Entin-Wohlman, and I. Ya. Korenblit, Phys. Rev. Lett. 91, 087206 (2003). 44This convenient notation for the triangular lattice was in- troduced by G. Khaliullin, W. Koshibae, and S. Maekawa, Phys. Rev. Lett. 93, 176401 (2004). 45A. M. Ole´ s, Phys. Rev. B 28, 327 (1983). 46J. Zaanen and G. A. Sawatzky, J. Solid State Chem. 88, 8 (1990). 47T. Mizokawa and A. Fujimori, Phys. Rev. B 54, 5368 (1996). 48M. Daghofer, A. M. Ole´ s, and W. von der Linden, Phys. Rev. B70, 184430 (2004); Phys. Stat. Solidi B 242, 311 (2005). 49G. Jackeli and D. A. Ivanov, Phys. Rev. B 76, 132407 (2007). 50C. Itoi, S. Qin, and I. Affleck, Phys. Rev. B 61, 6747 (2000). 51A. Nozaki, H. Yoshikawa, T. Wada, H. Yamauchi, and S. Tanaka, Phys. Rev. B 43, 181 (1991). 52J. Matsuno, Y. Okimoto, M. Kawasaki, and Y. Tokura, Phys. Rev. Lett. 95, 176404 (2005). 53M. Daghofer, K. Wohlfeld, A. M. Ole´ s, E. Arrigoni, and P. Horsch, Phys. Rev. Lett. 100, 066403 (2008). 54G. Jackeli and D. I. Khomskii, Phys. Rev. Lett. 100, 147203 (2008). 55M. P. Shores, E. A. Nytko, B. M. Barlett, and D. G. No- cera, J. Am. Chem. Soc. 127, 13462 (2005). 56T. Imai, E. A. Nytko, B. M. Bartlett, M. P. Shores, and D. G. Nocera, Phys. Rev. Lett. 100, 077203 (2008). 57P. Mendels, F. Bert, M. A. de Vries, A. Olariu, A. Har- rison, F. Duc, J. C. Trombe, J. S. Lord, A. Amato, and C. Baines, Phys. Rev. Lett. 98, 077204 (2007); F. Bert, S. Nakamae, F. Ladieu, D. L’Hˆ ote, P. Bonville, F. Duc, J.–C. Trombe, and P. Mendels, Phys. Rev. B 76, 132411 (2007); A. Olariu, P. Mendels, F. Bert, F. Duc, J.–C. Trombe, M. A. de Vries, and A. Harrison, Phys. Rev. Lett. 100, 087202 (2008). 58J. S. Helton, K. Matan, M. P. Shores, E. A. Nytko, B. M. Barlett, Y. Yoshida, Y. Takano, A. Suslov, Y. Qiu, J.–H. Chung, D. G. Nocera, and Y. S. Lee, Phys. Rev. Lett. 98, 107204 (2007). 59K. Matan, D. Grohol, D. G. Nocera, T. Yildirim, A. B. Harris, S. H. Lee, S. Nagler, and Y. S. Lee, Phys. Rev. Lett.96, 247201 (2006). 60V. Elser, Phys. Rev. Lett. 62, 2405 (1989). 61S. Dommange, M. Mambrini, B. Normand, and F. Mila, Phys. Rev. B 68, 224416 (2003). 62M. Hermele, T. Senthil, and M. P. A. Fisher, Phys. Rev. B72, 104404 (2005). 63S. Capponi, A. L¨ auchli, and M. Mambrini, Phys. Rev. B 70, 104424 (2004). 64M. B. Hastings, Phys. Rev. B 63, 014413 (2000). 65Y. Ran, M. Hermele, P. A. Lee, and X.–G. Wen, Phys. Rev. Lett. 99, 117205 (2007).
1604.04623v3.Determining_the_spin_orbit_coupling_via_spin_polarized_spectroscopy_of_magnetic_impurities.pdf
arXiv:1604.04623v3 [cond-mat.mes-hall] 15 Oct 2016Determining the spin-orbit coupling via spin-polarized sp ectroscopy of magnetic impurities V. Kaladzhyan,1,2,∗P. Simon,2and C. Bena1,2 1Institut de Physique Th´ eorique, CEA/Saclay, Orme des Meri siers, 91190 Gif-sur-Yvette Cedex, France 2Laboratoire de Physique des Solides, CNRS, Univ. Paris-Sud , Universit´ e Paris-Saclay, 91405 Orsay Cedex, France (Dated: October 1, 2018) We study the spin-resolved spectral properties of the impur ity states associated to the presence of magnetic impurities in two-dimensional, as well as one-dim ensional systems with Rashba spin-orbit coupling. We focus on Shiba bound states in superconducting materials, as well as on impurity states in metallic systems. Using a combination of a numeric al T-matrix approximation and a direct analytical calculation of the bound state wave funct ion, we compute the local density of states (LDOS) together with its Fourier transform (FT). We fi nd that the FT of the spin-polarized LDOS, a quantity accessible via spin-polarized STM, allows to accurately extract the strength of the spin-orbit coupling. Also we confirm that the presence of mag netic impurities is strictly necessary for such measurement, and that non-spin-polarized experim ents cannot have access to the value of the spin-orbit coupling. I. INTRODUCTION The electronic bands of materials that lack an inver- sion center are split by the spin-orbit (SO) coupling. A strong SO coupling implies that the spin of the electron is tied to to the direction of its momentum. Materials with strong SO coupling have been receiving a consider- able attention in the past decade partly because SO is playing an important role for the discovery of new topo- logical classes of materials.1,2Two-dimensional topolog- ical insulators, first predicted in graphene,3have been discovered in HgTe/CdTe heterostructures4following a theoretical prediction by Bernevig et al..5They are char- acterized by one-dimensional helical edge states where the spin is locked to the direction of propagation due to the strong SO coupling. Similar features occur for the surface states of 3D topological insulators which also haveastrongbulk SO coupling.1The spin-to-momentum locking was directly observed by angle-resolved photoe- mission spectroscopy (ARPES) experiments.6,7 Topological superconductors share many properties with topological insulators. They possess exotic edge statescalled Majoranafermions, particleswhich aretheir own antiparticles.1Topological superconductivity can be either induced by the proximity with a standard s-wave superconductor or be intrinsic. In the former case, Ma- jorana states have thus been proposed to form in one- dimensional8,9and two-dimensional semiconductors10,11 withstrongSOcouplingwhenproximitizedwithas-wave superconductor, and in the presence of a Zeeman field. Following this strategy, many experiments have reported signatures of Majorana fermions through transport spec- troscopy in one dimensional topological wires.12–16How- ever, there are presently only a few material candidates such as strontium ruthenate,17certain heavy fermion superconductors,18or some doped topological insulators such as Cu xBi2Se3,19that may host intrinsic topological superconductivity.Although SO coupling has been playing an essential role in the discovery of new topological materials, it is also of crucial importance in the physics of spin Hall effect,20in spintronics21and quantum (spin) computa- tion since it allows to electrically detect and manipulate spin currents in confined nanostructures (see Ref. 22 for a recent review). Based on the prominent role played by SO in the past decades, it is thus of great interest to be able to evaluate the SO coupling value in a given mate- rial accurately, though in general this is a very dif- ficult task. Inferences can be made from ARPES measurements;23–25in particular spin-polarized ARPES measurements have been used to evaluate the SO cou- pling in variousmaterials.26–32Other possibilities involve magneto-transport measurements in confined nanostruc- tures: this technique has been used to measure the SO coupling in clean carbon nanotubes33or in InAs nanowires.34 Here we propose a method to measure the magnitude of the SO coupling directly using spin-polarized scan- ning tunnelling microscopy (STM),35and the Fourier transform (FT) of the local density of states (LDOS) near magnetic impurities (FT-STM). The FT-STM tech- nique has been used in the past in metals, where it helped in mapping the band structure and the shape and the properties of the Fermi surface,36–43as well as in extracting information about the spin properties of the quasiparticles.44More spectacularly, it was used successfully in high-temperature SCs to map with high resolution the particular d-wave structure of the Fermi surface, as well as to investigate the properties of the pseudogap.45–47 In this paper, on one hand, we calculate the Fourier transform of the spin-polarized local density of states (SP LDOS) of the so-called Shiba bound state48–51as- sociated with a magnetic impurity in a superconductor. Shiba bound states have been measured experimentally by STM52–54and it has actually been shown that the ex-2 tentoftheShibawavefunction canreachtensofnanome- ters in 2D superconductors, which allows one to mea- sure the spatial dependence of the LDOS of such states withhighresolution.55Weconsiderbothone-dimensional and two-dimensional superconductors with SO coupling. While two-dimensional systems such as e.g. Sr 2RuO4,17 or NbSe 255,56become superconducting when brought at lowtemperature, one-dimensionalwiressuchasInAs and InSb are not superconducting at low temperature. In or- der to see the formation of Shiba states one would need to proximitize them by a SC substrate. The formation of Shiba states in such systems,57,58as well as in p-wave superconductors,59,60has been recently touched upon, but the effect of the SO coupling on the FT of the SP LDOS in the presence of magnetic impurities has not previously been analyzed. On the other hand we focus on the effects of the spin- orbit coupling on the impurity states of a classical mag- netic impurity in one-dimensional and two-dimensional metallic systems such as Pb61and Bi, as well as InAs and InSb semiconducting wires that can be also mod- eled as metals in the energy range that we consider. We should note that for these systems no bound state forms at a specific energy, but the impurity is affecting equally the entire energy spectrum. By studying the two classesofsystems described above we show that the SO coupling can directly be read-off from the FT features of the SP LDOS in the vicinity of the magnetic impurity. We note that such a signa- ture appearsonly formagneticimpurities, and onlywhen thesystemisinvestigatedusingspin-polarizedSTMmea- surements, the non-spin-polarized measurements do not provide information on the SO, as it has also been previ- ously noted.62The main difference between the SC and metallic systems, beyond the existence of a bound state in the former case, is that the spin-polarized Friedel os- cillations around the impurity haveadditional features in theSCphase, themostimportantonebeingtheexistence of oscillations with a wavelength exactly equal to the SO coupling length scale; such oscillations are not present in the metallic phase. Another difference is the broadening of the FT features in the superconducting phase com- pared to the non-SC phase in which the sole broadening is due to the quasiparticle lifetime. We focus on Rashba SO coupling as assumed to be the most relevant for the systems considered, but we have checked that our conclusion holds for other types of SO. To obtain the SP LDOS we use a T-matrix approximation,43,63,64and we present both numerical and analytical results which allow us to obtain a full un- derstandingofthe observedfeatures, ofthe splittings due to the SO, as well as of the spin-polarization of the im- purity states and of the symmetry of the FT features. In Sec. II, we present the general model for two- dimensional and one-dimensional cases and the basics of the T-matrix technique. In Sec. III we show our results for the SP LDOS, calculated both numerically and an- alytically, for 2D systems, both in the SC and metallicphase. Sec. IV is devoted to SP LDOS of impurity in one-dimensional systems. Our Conclusions are presented in section V. Details of the analytical calculations are given in the Appendices. II. MODEL We consider an s-wave superconductor with a SC paring ∆ s, and Rashba SO coupling λ, for which the Hamiltonian, written in the Nambu basis Ψ p= (ψ↑p,ψ↓p,ψ† ↓−p,−ψ† ↑−p)T, is given by: H0=/parenleftbigg ξpσ0∆sσ0 ∆sσ0−ξpσ0./parenrightbigg +HSO. (1) The energy spectrum is ξp≡p2 2m−εF, whereεFis the Fermi energy. The operator ψ† σpcreates a particle of spin σ=↑,↓of momentum p≡(px,py) in 2D and p≡pxin 1D. Below we set /planckover2pi1to unity. The system is considered to lay in the (x,y) plane in 2D case, whereas in 1D case we setpyto zero in the expressions above, and we consider a system lying along the x-axis. The metallic limit is recovered by setting ∆ s= 0. The Rashba Hamiltonian can be written as HSO=λ(pyσx−pxσy)⊗τz, (2) in 2D and simply as HSO=λpxσy⊗τzin 1D. We have introduced σandτ, the Pauli matrices acting re- spectively in the spin and the particle-hole subspaces. The unperturbed retarded Green’s function can be ob- tained from the above Hamiltonian via G0(E,p) = [(E+iδ)I4−H0(p)]−1, whereδis the inverse quasipar- ticle lifetime. In what follows we study what happens when a sin- gle localized impurity is introduced in this system. We consider magnetic impurities of spin J= (Jx,Jy,Jz) de- scribed by the following Hamiltonian: Himp=J·σ⊗τ0·δ(r)≡V·δ(r),(3) whereJis the magnetic strength. We only consider here classical impurities oriented either along the z-axis,J= (0,0,Jz), or along the x-axis,J= (Jx,0,0). This is justifiedprovidedtheKondotemperatureismuchsmaller than the superconducting gap.51 To find the impurity states in the model described above we use the T-matrix approximation described in [51, 63, and 64] and [43]. We also neglect the renormal- ization of the superconducting gap because it is mainly local51,65and therefore only introduces minor effects for our purposes. Since the impurity is localized, the T- matrix is given by: T(E) =/bracketleftbigg 1−V/integraldisplayd2p (2π)2G0(E,p)/bracketrightbigg−1 V.(4)3 The real-space dependence of the non-polarized, δρ(r,E), and SP LDOS, Sˆn(r,E), with ˆn=x,y,z, can be found as Sx(r,E) =−1 πℑ[∆G12+∆G21], Sy(r,E) =−1 πℜ[∆G12−∆G21], Sz(r,E) =−1 πℑ[∆G11−∆G22], δρ(r,E) =−1 πℑ[∆G11+∆G22], with ∆G(E,r)≡G0(E,−r)T(E)G0(E,r), where ∆Gijdenotes the ij-th component of the matrix ∆G, andG0(E,r) is the unperturbed retarded Green’s function in real space, given by the Fourier transform G0(E,r) =/integraldisplaydp (2π)2G0(E,p)eipr. (5) The FT of the SP LDOS components in momentum space,Sˆn(p,E) =/integraltextdrSˆn(r,E)e−ipr, with ˆn=x,y,z, as well as the FT of the non-polarized LDOS, δρ(p,E) =/integraltext drδρ(r,E)e−iprare given by Sx(p,E) =i 2π/integraldisplaydq (2π)2[˜g12(E,q,p)+ ˜g21(E,q,p)],(6) Sy(p,E) =1 2π/integraldisplaydq (2π)2[g21(E,q,p)−g12(E,q,p)],(7) Sz(p,E) =i 2π/integraldisplaydq (2π)2[˜g11(E,q,p)−˜g22(E,q,p)],(8) δρ(p,E) =i 2π/integraldisplaydq (2π)2[˜g11(E,q,p)+ ˜g22(E,q,p)],(9) wheredq≡dqxdqy, g(E,q,p) =G0(E,q)T(E)G0(E,p+q) +G∗ 0(E,p+q)T∗(E)G∗ 0(E,q), ˜g(E,q,p) =G0(E,q)T(E)G0(E,p+q) −G∗ 0(E,p+q)T∗(E)G∗ 0(E,q), andgij, ˜gijdenote the corresponding components of the matricesgand ˜g. Note that while the non-polarized and the SP LDOS are of course real functions when evalu- ated in position space, their Fourier transforms need not be. Sometimes we get either or both real and imaginary components for the FT, depending on their correspond- ing symmetries. In the figureswe shallindicate eachtime if we plot the real or the imaginarycomponent of the FT. To obtain the FT of the non-polarized and the SP LDOS, we first evaluate the momentum integrals in Eqs. (4-9) numerically. For this we use a square lattice version of the Hamiltonians (1) and (2), where we take the tight-binding spectrum Ξ p≡µ−2t(cospx+cospy) with chemical potential µand hopping parameter t. We set the lattice constant to unity. It is also worth noting thatallthe numericalintegrationsareperformedoverthe first Brillouin zone and that we use dimensionless units by settingt= 1.Alternatively, as detailed in the appendices, we find the exact form for the non-polarized and SP LDOS in the continuum limit by performing the integrals in the FTofthe Green’sfunctions analytically. Moreover,when considering the SC systems, the energies Eof the Shiba statestogetherwith the correspondingeigenstatesforthe ShibawavefunctionsΦattheorigincanbeobtainedfrom the corresponding eigenvalue equation66 [I4−VG0(E,r=0)]Φ(0) = 0. (10) The spatial dependence of the Shiba state wave function is determined using Φ(r) =G0(E,r)VΦ(0). (11) The real-space Green’s function is obtained simply by a Fourier transformof the unperturbed Green’s function in momentum space, G0(E,p). The non-polarized and the SP LDOS are given by ρ(E,r) = Φ†(r)/parenleftbigg 0 0 0σ0/parenrightbigg Φ(r), (12) and S(E,r) = Φ†(r)/parenleftbigg 0 0 0σ/parenrightbigg Φ(r), (13) where we take into account only the hole components of the wave function, and not the electron ones. This is because the physical observables are related to only one of the two components, for example in a STM measure- ment one injects an electron at a given energy and thus have access to the allowed number of electronic states, not to both the electronic and hole states simultaneously. The Bogoliubov-de Gennes Hamiltonian contains the so- called particle-hole redundancy, and the electron and the holecomponentscanbesimplyrecoveredfromeachother by overall changes of sign, and/or changing the sign of the energy. Belowwecompute only the hole components, but there would have been no qualitative differenced had we computed the electron component. III. RESULTS FOR TWO DIMENSIONAL SYSTEMS A. Real and momentum space dependence of the 2D Shiba bound states For a 2D superconductorwith SO coupling in the pres- ence of a magnetic impurity one expects the formation of a single pair of Shiba states.57,58The energies of the particle-hole symmetric Shiba states67are given by (in- dependent of the direction of the impurity): E1,¯1=±1−α2 1+α2∆s, whereα=πνJandν=m 2π. (See Appendix A for details of how the energies of the Shiba states are calculated.)4 Up to the critical value αc= 1 these energies are ordered the following way: E1>E¯1. As soon as α>α c, energy levelsE1andE¯1exchange places, making the order the following:E¯1> E1. This corresponds to a change of the ground state parity.51,68,69Forα≫1 the subgap states approach the gap edge and eventually merge with the continuum. For the type of impurities considered here, there is no dependence of these energies on the SO couplingin thelow-energyapproximation,thoughaweak dependence is introduced when one takes into account the non-linear form of the spectrum. The dependence of energy of the Shiba states on the impurity strength Jis depicted in Fig. 1 where we plot the total spin of the impurity state S(p= 0) as a function of energy and impurity strength. Note that the two opposite-energy Shiba states have opposite spins. FIG. 1. (Color) The averaged SP LDOS induced by an impu- rity as a function of the impurity strength for an in-plane magnetic impurity. The dashed line shows the supercon- ducting gap. A similar result is obtained when the impu- rity spin is perpendicular to the plane. Note that the two Shiba states with opposite energies have opposite spin. We sett= 1,µ= 3,δ= 0.01,λ= 0.5,∆s= 0.2. We are interested in studying the spatial structure of the Shiba states in the presence of magnetic impurities oriented both perpendicular to the plane, and in plane. Thiscanbedonebothinrealspaceandmomentumspace bycalculatingthe Fouriertransformofthe spin-polarized LDOSusingtheT-matrixtechniquedetailedintheprevi- ous section. We focus on the positive-energy Shiba state, noting that its negative energy counterpart exhibits a qualitativelysimilarbehavior. InFig. 2weshowthereal- space dependence of the non-polarized and SP LDOS. Each of the panels corresponds to the interference pat- terns originating from different types of scattering. Note that the spin-orbit value cannot be accurately extracted from these type of measures, since the system contains oscillationswith manydifferentsuperposingwavevectors. ToovercomethisproblemwefocusontheFTofthesefea- tures, as it is oftentimes done in spatially resolved STM experiments, which allow for a more accurate separation of the different wavevectors.36–43Thus in Fig. 3 we focus on the FT of the SP LDOS for two types of impuritieswith spin oriented along zandxaxes respectively. z-impurity x-impurity Jz(Jx=Jy= 0) Jx(Jy=Jz= 0) FIG. 2. (Color) The real-space dependence of the non- polarized as well as of the SP LDOS components for the positive energy Shiba state, for a magnetic impurity with Jz= 2 (left column), and Jx= 2 (right column). We take t= 1,µ= 3,δ= 0.01,λ= 0.5,∆s= 0.2. Note that the SO introducesnon-zerospin components in the directions different from that of the impurity spin. These components exhibit either two-fold or four-fold symmetric patterns. Also the SO is affecting strongly the spin component parallelto the impurity, in particular when the impurity is in-plane, in which case the struc- ture of the SP LDOS around the impurity is no longer radially symmetric. However, as can be seen in the bot- tom panel of Fig. 3, the non-spin-polarized LDOS is5 not affected by the presence of SO, preserving a radially symmetric shape quasi-identical to that obtained in the absence of SO. Thus the SO coupling can be measured onlyviathe spin-polarizedcomponentsofthe LDOS, and not the non-polarized LDOS. These results, which are obtained using a numerical integration of the T-matrix equations, are also supported by analytical calculations which help to understand the fine structure of the FT of the SP LDOS (see Appendices for details). These calculations yield for the SP LDOS generated by a magnetic impurity perpendicular to the plane Sx(r) = +J2 z/parenleftbigg 1+1 α2/parenrightbigge−2psr rcosφr× ×/braceleftigg/summationdisplay σσν2 σ pσ Fcos(2pσ Fr−θ)+2ν2v2 F v2pFsinpλr/bracerightigg , Sy(r) = +J2 z/parenleftbigg 1+1 α2/parenrightbigge−2psr rsinφr× ×/braceleftigg/summationdisplay σσν2 σ pσ Fcos(2pσ Fr−θ)+2ν2v2 F v2pFsinpλr/bracerightigg , Sz(r) =−J2 z/parenleftbigg 1+1 α2/parenrightbigge−2psr r× ×/braceleftigg/summationdisplay σν2 σ pσ Fsin(2pσ Fr−θ)−2ν2v2 F v2pFcospλr/bracerightigg , ρ(r) = +J2 z/parenleftbigg 1+1 α2/parenrightbigge−2psr r× ×/braceleftbigg 2ν2 mv+2ν2v2 F v2pFsin(2mvr−θ)/bracerightbigg ,(14) with tanθ=/braceleftigg 2α 1−α2,ifα/ne}ationslash= 1 +∞,ifα= 1. (15) We have introduced eiφr=x+iy r, and pσ F=−σmλ+mv, (16) pλ= 2mλ, (17) ps=/radicalig ∆2s−E2 1/v, (18) withv=/radicalbig v2 F+λ2, andvF=/radicalbig 2εF/m. Herepσ F, pλ andpsare the different momenta which can be read off from the SP LDOS. For an in-plane magnetic impuritywe have Ss x(r) =−J2 x/parenleftbigg 1+1 α2/parenrightbigg/braceleftbigg/summationdisplay σν2 σ pσ F[1+sin(2pσ Fr−2β)], +2ν2v2 F v2pF[cospλr+sin(2mvr−2β)]/bracerightbigge−2psr r Sa x(r) = +J2 x/parenleftbigg 1+1 α2/parenrightbigg/braceleftbigg/summationdisplay σν2 σ pσ F[1−sin(2pσ Fr−2β)] −2ν2v2 F v2pF[cospλr+sin(2mvr−2β)]/bracerightbigge−2psr rcos2φr, Sy(r) = +J2 x/parenleftbigg 1+1 α2/parenrightbigg/braceleftbigg/summationdisplay σν2 σ pσ F[1−sin(2pσ Fr−2β)] −2ν2v2 F v2pF[cospλr−sin(2mvr−θ)]/bracerightbigge−2psr rsin2φr, Sz(r) =−J2 x/parenleftbigg 1+1 α2/parenrightbigg/braceleftbigg 2/summationdisplay σσν2 σ pσ Fcos(2pσ Fr−θ) +4ν2v2 F v2pFsinpλr/bracerightbigge−2psr rcosφr, ρ(r) = +J2 x/parenleftbigg 1+1 α2/parenrightbigg/braceleftbigg 4ν2 mv+4ν2v2 F v2pF× ×sin(2mvr−θ)/bracerightbigge−2psr r, (19) with tanβ=α. TheSxcomponent is the sum of symmetric part a Ss x and an asymmetric part Sa x. Note that the features ob- served in the FT of the SP LDOS plots are well captured by the analytical calculations. In particular we note that the oscillations in the SP LDOS are dominated by the following four wavevectors: 2p± F, p+ F+p− F= 2mv,andp− F−p+ F=pλ≡2mλ, which should give rise in the FT to high-intensity fea- tures at these wavevectors (the red arrows in Fig. 3). Indeed, we note in the numerical results for the FT of the SP LDOS the existence of four rings, correspond- ing to 2p± F,p+ F+p− F= 2mvandp− F−p+ F=pλ, hav- ing the proper two-fold or fold-fold symmetries, consis- tent with the cos /sinφrand cos/sin2φrdependence of the SP LDOS obtained analytically. For example, in the xcomponent of the SP LDOS induced by an ximpu- rity, the 2p+ F, 2p− Fandpλrings have a maximum along x and a minimum along y, while the 2 mvring has a sym- metry corresponding to a rotation by 90 degrees. The ycomponent of the FT of the SP LDOS has a four- fold symmetry in which we can again identify the same wavevectors,while the Szcomponent has a two-foldsym- metry, and the 2 mvvector is absent. Similarly, for the Sxand theSycomponents of the SP LDOS induced by azimpurity (these components should be zero in the absence of the SO coupling) only the 2 p± Fandpλwave vectors are present, with similar symmetries, while the Szcomponent is symmetric. Note also the central peak6 z-impurity x-impurity Jz(Jx=Jy= 0) Jx(Jy=Jz= 0) FIG. 3. (Color) The FT of the non-polarized as well as of the SP LDOS components for the positive energy Shiba state as a function of momentum, for a magnetic impurity with Jz= 2 (left column), and Jx= 2 (right column). We take t= 1,µ= 3,δ= 0.01,λ= 0.5,∆s= 0.2. For a z-impurity we depict the real part of the FT for δρand forSz, and the imaginary part for and SxandSy, whereas for an x-impurity we take the imaginary part only for the Szcomponent. Black two- headed arrows correspond to the value of 2 pλ≡4mλ(see the analytical results) and thus allow to extract the SO couplin g constant directly from these strong features in momentum space. The other arrows correspond to the other important wavevectors that can be observed in these FTs, as identified with the help of the analytical results. atpx=py= 0 which is due to the terms independent of FIG. 4. (Color) The FT of various SP LDOS component for a Shiba state as a function of the SO coupling λand ofpy(for px= 0 - vertical cut). We take t= 1,µ= 3,δ= 0.01,∆s= 0.2,Jz= 2. position in the SP LDOS. The most important observation is that all the compo- nents of the FT of the SP LDOS exhibit a strong feature at wave vector pλ. Thus an experimental observation of this feature via spin-polarized STM would allow one to read-off directly the value of the SO coupling. The spin orbit can also be read-off from the distance between the 2p+and 2p−peaks, though the intensity of these fea- tures is not as strong. This appears clearly in Fig. 4, in which we plot a horizontal cut though two of the FT – SP LDOS above as a function of the SO coupling λ. Note that the only wave vector present in the non- polarized LDOS is 2 mv, which has only a very weak de- pendence on λfor not too large values of the SO with re- spect to the Fermi velocity, thus it is quasi-impossible to determine the SO coupling from a measurement without spin resolution. Note also the typical two-dimensional 1/rdecay of the Friedel oscillations is overlapping in this case with an exponential decay with wave vector ps. B. Comparison to the metallic phase Asimilaranalysiscanbe performedforimpurity states forming in the vicinity of a magnetic impurity in a metal- lic system. Here the classical magnetic impurity does not lead to any localized bound states at a specific energy, and the intensity of the impurity contribution is roughly independent of energy. Thus in Fig. 5 we plot the FT of the impurity con- tribution to the LDOS and SP LDOS at a fixed energy E= 0.1. We note that we have similar features to those observed in the SC regime, with the main differences be- ing that the long-wavelengthcentral features are now ab- sent, andthat the FT peaksaremuch sharperthan in the SC regime. This behavior can be explained from the an- alytical expressions of the non-polarized and SP LDOS, whose derivation is presented in Appendix B. The results7 z-impurity x-impurity Jz(Jx=Jy= 0) Jx(Jy=Jz= 0) FIG. 5. (Color) The FT of the impurity contributions to the non-polarized and SP LDOS for an energy E= 0.1 and for a magnetic impurity with Jz= 2 (left column), and Jx= 2 (right column). We take the inverse quasiparticle lifetime δ= 0.03 and we set t= 1,µ= 3,λ= 0.5,∆s= 0. For az-impurity we depict the real part of the FT for δρand forSz, and the imaginary part for SxandSy, whereas for anx-impurity we take the imaginary part only for the Sz component. UnlikeintheSCcase, thestrongpeaksappearing in the center and at pλare absent here. The arrows denote the wavevectors of the observed features as identified from t he analytical calculations.are presented below for an out-of-plane spin impurity: Sx(r)∼J 1+α2cosφr r/summationdisplay σσν2 σ pσsin2pσr, Sy(r)∼J 1+α2sinφr r/summationdisplay σσν2 σ pσsin2pσr, Sz(r)∼ −J 1+α22 r/summationdisplay σν2 σ pσcos2pσr, ρ(r)∼ −J 1+α24αν2v2 F v21/radicalbig p2 F+2mE+E2/v2× ×sinpεr r, (20) while for an xdirected impurity (in-plane): Sx(r)∼ −J 1+α2/braceleftbigg 2ν2v2 F v21−cos2φr rcospεr/radicalbig p2 F+2mE+E2/v2 +/summationdisplay σν2 σ pσ1+cos2φr rcos2pσr, Sy(r)∼ −J 1+α2sin2φr r/bracketleftbigg −2ν2v2 F v2cospεr/radicalbig p2 F+2mE+E2/v2 +/summationdisplay σν2 σ pσcos2pσr/bracketrightbigg , Sz(r)∼ −J 1+α2cosφr r/summationdisplay σσν2 σ pσsin2pσr, ρ(r)∼ −J 1+α2·α r·4ν2v2 F v2sinpεr/radicalbig p2 F+2mE+E2/v2,(21) withpF=mvF,pσ=pσ F+E/v/ne}ationslash= 0,pε≡p++p−= 2(mv+E/v) andνσ=ν/bracketleftbig 1−σλ v/bracketrightbig . Note that these expressions are very similar to those obtained in the SC regime, except that the wave vec- tors of the oscillations now do not include pλ. However, this could still be read-off experimentally from the differ- ence between p−andp+. Another important difference between the SC and non-SC regimes is the presence of the exponentially decaying term in the expressions de- scribing the LDOS dependence for the Shiba states in the SC regime. The Shiba states have an exponential decay for distances larger than the superconducting co- herence length, while the impurity states in the non-SC regime only decay algebraically as 1 /r. In the Fourier space this is translated into a much larger broadening of the features corresponding to the Shiba states in the SC regime with respect to that of the features corresponding to the impurity contributions in metals. The width of the peaks in the latter is solely controlled by the inverse quasiparticle lifetime δand is generally quite small. Note also that in both regimes one needs to use the spin-polarized LDOS and magnetic impurities to be able to extract the value of the SO coupling, while the non- polarized LDOS is not sensitive to this wavevector. Last but notleast, asdescribedin Appendix B,theLDOS per- turbations induced by a non-magnetic impurity do not8 show any direct signature of the SO coupling (the only contributing wavevector is 2 mvin the metallic regime, while in the SC regime no Shiba state form for a non- magnetic impurity), thus the only manner to have access to the SO coupling is via spin-polarized STM in the pres- ence of magnetic impurities. IV. ONE-DIMENSIONAL SYSTEMS While in one-dimensional systems superconductivity is not intrinsic, a superconducting gap can be opened via proximitizing them with a superconducting substrate. For such systems it is thus particularly interesting to study the FT of the SP LDOS for both the supercon- ducting and non-superconducting regimes, as both these regimes can be achieved experimentally at low tempera- ture for the same materials. WeconsidertheHamiltoniangivenbyEqs.(1-3), where we setpy→0, and we perform a T-matrix analysis sim- ilar to that described in the previous section for both the SC and non-SC phases, for different directions of the magnetic impurity. The wire is considered to be oriented along thexdirection, and the SO coupling is oriented alongy.8,9We thus expect a similar and more exotic be- havior for impurities directed along xandz, and a more classical behavior for impurities with the spin parallel to the direction of the SO, thus oriented along y. The energies and wave functions of the Shiba states can be found using the same procedure as for the two- dimensional systems (see Appendix C). This yields for the energies of the states: E1,¯1=±1−α2 1+α2∆s,whereα=J/v. The FT of the positive energy state as a function of momentum and the SO coupling is presented in Fig. 6 for a SC (left column) and non-SC state (right column), for an impurity directed along z. For this situation the spin of the Shiba state has two non-zero components, one parallel to the wire, and one parallel to the impurity spin, and these two components are depicted in Fig. 6. Note that, similar to the two-dimensional case, there is a split of the FT features increasing linearly with the SO coupling strength. Also note that in the non-SC phase the central feature, whose wave vector is given by pλ, is absent, and that the FT features are broadened in the SC regime with respect to the non-SC one. Also, same as in the two-dimensional case, the SO affects the spin-polarized components but almost do not change the non-polarized LDOS, as it can be seen in Fig. 6 where it appears that the non-polarized LDOS FT features do not evolve with the SO coupling. These results are confirmed by analytical calculations. Below we give the spin components and the LDOS in the SC state for an impurity directed along zobtained analytically (see Appendix C), for the positive energy Shiba state:SC case Non-SC case FIG. 6. (Color) The FT of various SP LDOS component for a Shiba state (left column), and for an impurity state at E= 0.1 (right column), as a function of the SO coupling λ and of momentum p, for an impurity perpendicular to the wire and directed along z. We set t= 1,µ= 1. We take ∆s= 0.2,Jz= 4,δ= 0.01 in the SC case and ∆ s= 0,Jz= 2,δ= 0.05 in the non-SC case. Sx(x) =1+α2 4[2sinpλx+sin(2mv|x|+pλx−2θ) −sin(2mv|x|−pλx−2θ)]·e−2ω|x|/v Sy(x) = 0 Sz(x) =−1+α2 4[2cospλx+cos(2mv|x|+pλx−2θ) +cos(2mv|x|−pλx−2θ)]·e−2ω|x|/v ρ(x) =1+α2 2[1+cos(2mv|x|−2θ)]·e−2ω|x|/v(22) where tanθ=α. We also present the FT of the SP LDOSforthenon-SCphasefortheimpuritycontribution9 corresponding to the energy E(see Appendix D): Sx(x) = +α 1+α2·1 πv[cos(pε|x|−pλx)−cos(pε|x|+pλx)] Sy(x) = 0 Sz(x) = +α 1+α2·1 πv[sin(pε|x|−pλx)+sin(pε|x|+pλx)] ρ(x) =−2α2 1+α2·1 πvcospεx As before, in the expressions above pε= 2(mv+ E/v),pλ= 2mλ. Indeed these calculations confirm our observations, in the SC state the dominantwavevectorsare2 p± F= 2mv± pλ, 2mvandpλ, while in the non-SC phase only pǫ±pλ, and 2mv. Similar results are obtained if the impurity is oriented alongx, with the only difference that the xandzcompo- nents will be interchanged, up to on overall sign change (see Appendices C and D). For impurities parallel to y, and thus to the SO vector, we expect the SP LDOS to be less exotic, and indeed in this case the only non-zero component of the impurity SP LDOS is Sy. In the SC regime we thus find Sx(x) = 0 Sy(x) =−(1+α2)[1+cos(2mv|x|−2θ)]·e−2ω|x|/v Sz(x) = 0 ρ(x) = +(1+α2)[1+cos(2mv|x|−2θ)]·e−2ω|x|/v while in the non-SC regime we have Sx(x) = 0 Sy(x) = +2α 1+α2·1 πvsinpε|x| Sz(x) = 0 ρ(x) =−2α2 1+α2·1 πvcospεx We see that Syexhibits features only at the 2 mvand correspondingly at the pǫwave vectors, same as the non- polarized LDOS, thus not allowing for the detection of the SO coupling. For intermediate directions of the impurity spin, all three components will be present, with the xandzex- hibiting all the wave vectors, while the ycomponent solely the 2 mv, and with relative intensities given by the relative components of the impurity spin. Thus, we conclude that, same as in the 2D case, the SO can be measured using spin-polarized STM and mag- netic impurities; moreover, in the 1D case one needs to consider impurities that have a non-zero component per- pendicular to the direction of the SO. V. CONCLUSIONS We have analyzed the formation of Shiba states and impuritystatesin1Dand2Dsuperconductingandmetal-lic systems with Rashba SO coupling. In particular we have studied the Fourier transform of the local density of states of Shiba states in SCs and of the impurity states in metals, both non-polarized and spin-polarized. We have shown that the spin-polarized density of states con- tains information that allows one to extract experimen- tally the strength of the SO coupling. In particular the features observed in the FT of the SP LDOS split with a magnitude proportional to the SO coupling strength. Moreover, the Friedel oscillations in the SP LDOS in the SC regime show a combination of wavelengths, out of which the SO length can be read off directly and non- ambiguously. We note that these signatures are only vis- ible in the spin-polarized quantities and in the presence of magnetic impurities. For non-spin-polarized measure- ments, no such splitting is present and the wave vectors observed in the FT of the SP LDOS basically do not depend on the SO coupling. When comparing the re- sults for the SC Shiba states to the impurity contribu- tion in the metallic state and we find a few interesting differences, such as a broadening of the FT features cor- responding to a spatial exponential decay of the Shiba states compared to the non-SC case. Moreover, the FT of the SP LDOS in the SC regime exhibits extra fea- tures with a wavelengthequal to the SO length which are not present in the non-SC phase. It would be interest- ing to generalize our results to more realistic calculations which may include some specific lattice characteristics, more realistic material-dependent tight-binding parame- ters for the band structure and the SO coupling values. However, we should note that our results have a fully general characteristic, independent of the band structure or other material characteristics, and that the features in the FT of the non-polarized LDOS will correspond to split features in the spin-polarized LDOS, and thus the spin-orbit can be measured unequivocally from the split obtained from the comparison between the non-polarized and spin-polarized measurements. We have checked that up to a rotation in the spin space our results hold also for other types of SO coupling such as Dresselhaus. According to our knowledge, the FT-STM is a well- established experimental technique which does not deal with large systematic errors.36–43The experimental data presented e.g. in Ref. 43 shows that the resolution in the Fourier space (momentum space) reaches 0 .05˚A−1, whereas a typical value of spin-orbit coupling wave vec- torpλ∼0.15˚A−1(see e.g. Ref. 22), and thus it is suf- ficient to resolve the features originating from the spin- orbit coupling. Moreover, we would like to point that the exponent e−2psrdefines in the real space how far the impurity-induced states areextended, and it manifests in the momentum space as the widening of the ring-like fea- tures appearing at particular momenta. The condition of resolving the spin-orbit is thus 2 ps< pλ, otherwise the widening is large enough to blur the spin-orbit feature. This condition can be rewritten in a more explicit way,10 namely 1/radicalbig 1+(λ/vF)2·α 1+α2·∆s εF<λ vF For any realistic parameters the first two factors on the left side are of the order of unity, and ∆ s/εF∼10−3 for superconductors. However, for realistic values of the spin-orbit coupling λ, this inequality holds and therefore there should not be any technical problem with resolving those features. Our results can be tested using for example materials such as Pb, Bi, NbSe 2or InAs and InSb wires, which are known to have a strong SO coupling, using spin- polarized STM which is nowadays becoming more andmore available.35 While finalizing this manuscript we became aware of a recent work70focusing on issues similar to some of the subjects (in particular the real space Friedel oscillations in the metallic regime) addressed in our work. ACKNOWLEDGMENTS This work is supported by the ERC Starting Indepen- dent Researcher Grant NANOGRAPHENE 256965. PS would like to acknowledge interesting discussions with T. Cren and financial support from the French Agence Na- tionale de la Recherche through the contract Mistral. ∗vardan.kaladzhyan@cea.fr 1M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010). 2X.-L. Qi and S.-C. Zhang, Rev. Mod. Phys. 83, 1057 (2011). 3C. L. Kane and E. J. Mele, Phys. Rev. Lett. 95, 226801 (2005). 4M. Konig and et al., Science 318, 766 (2007). 5B. A. Bernevig, T. L. Hughes, and S.-C. Zhang, Science314, 1757 (2006). 6D. Hsieh, Y. Xia, D. Qian, L. Wray, F. Meier, J. H. Dil, J. Osterwalder, L. Patthey, A. V. Fedorov, H. Lin, A. Ban- sil, D. Grauer, Y. S. Hor, R. J. Cava, and M. Z. Hasan, Phys. Rev. Lett. 103, 146401 (2009). 7D. Hsieh, , Y. Xia, D. Qian, L. Wray, J. H. Dil, F. Meier, J. Osterwalder, L. Patthey, J. G. Checkelsky, N. P. Ong, A. V. Fedorov, H. Lin, A. Bansil, D. Grauer, Y. S. Hor, R. J. Cava, and M. Z. Hasan, Nature 460, 1101 (2009). 8Y. Oreg, G. Refael, and F. von Oppen, Phys. Rev. Lett. 105, 177002 (2010). 9R. M. Lutchyn, J. D. Sau, and S. Das Sarma, Phys. Rev. Lett. 105, 077001 (2010). 10J. D. Sau, R. M. Lutchyn, S. Tewari, and S. Das Sarma, Phys. Rev. Lett. 104, 040502 (2010). 11J. Alicea, Phys. Rev. B 81, 125318 (2010). 12V. Mourik, K. Zuo, S. M. Frolov, P. S. M., E. P. A. M. Bakkers, and L. P. Kouwenhoven, Science 336, 1003 (2012). 13M. T. Deng, C. L. Yu, G. Y. Huang, M. Larsson, P. Caroff, and H. Q. Xu, Nano lett. 12, 64146419 (2012). 14A. Das, Y. Ronen, Y. Most, Y. Oreg, M. Heiblum, and H. Shtrikman, Nature Physics 8, 887 (2012). 15H. O. H. Churchill, V. Fatemi, K. Grove-Rasmussen, M. T. Deng, P. Caroff, H. Q. Xu, and C. M. Marcus, Phys. Rev. B 87, 241401 (2013). 16S.M. Albrecht, A.P.Higginbotham, M.Madsen, F.Kuem- meth, T. S. Jespersen, J. Nygard, P. Krogstrup, and C. M. Marcus, Nature 531, 206 (2016). 17A. P. Mackenzie and Y. Maeno, Rev. Mod. Phys. 75, 657 (2003). 18R.Joynt andL.Taillefer, Rev. Mod. Phys. 74, 235 (2002). 19S. Sasaki, M. Kriener, K. Segawa, K. Yada, Y. Tanaka, M. Sato, and Y. Ando,Phys. Rev. Lett. 107, 217001 (2011). 20J. Sinova, S. O. Valenzuela, J. Wunderlich, C. H. Back, and T. Jungwirth, Rev. Mod. Phys. 87, 1213 (2015). 21I.ˇZuti´ c, J. Fabian, and S. Das Sarma, Rev. Mod. Phys. 76, 323 (2004). 22A. Manchon, A. H. C. Koo, J. Nitta, S. M. Frolov, and R. A. Duine, Nature Materials 14, 871 (2015). 23S. LaShell, B. A. McDougall, and E. Jensen, Phys. Rev. Lett. 77, 3419 (1996). 24C. R. Ast, J. Henk, A. Ernst, L. Moreschini, M. C. Falub, D. Pacil´ e, P. Bruno, K. Kern, and M. Grioni, Phys. Rev. Lett. 98(2007), 10.1103/physrevlett.98.186807. 25A. Varykhalov, D. Marchenko, M. R. Scholz, E. D. L. Rienks, T. K. Kim, G. Bihlmayer, J. S´ anchez-Barriga, and O. Rader, Phys. Rev. Lett. 108(2012), 10.1103/physrevlett.108.066804. 26M. Hoesch, T. Greber, V. Petrov, M. Muntwiler, M. Hengsberger, W. Auw¨ arter, and J. Osterwalder, Journal of Electron Spectroscopy and Related Phenomena 124, 263 (2002). 27M. Hochstrasser, J. G. Tobin, E. Rotenberg, and S. D. Kevan, Phys. Rev. Lett. 89(2002), 10.1103/physrevlett.89.216802. 28M. Hoesch, M. Muntwiler, V. N. Petrov, M. Hengsberger, L. Patthey, M. Shi, M. Falub, T. Greber, and J. Osterwalder, Phys. Rev. B 69(2004), 10.1103/physrevb.69.241401. 29T. Hirahara, K. Miyamoto, I. Matsuda, T. Kadono, A. Kimura, T. Nagao, G. Bihlmayer, E. V. Chulkov, S. Qiao, K. Shimada, H. Na- matame, M. Taniguchi, and S. Hasegawa, Phys. Rev. B 76(2007), 10.1103/physrevb.76.153305. 30D.Hsieh, Y.Xia, L.Wray, D.Qian, A.Pal, J. H.Dil, J. Os- terwalder, F. Meier, G. Bihlmayer, C. L. Kane, Y. S. Hor, R. J. Cava, and M. Z. Hasan, Science 323, 919 (2009). 31K. Yaji, Y.Ohtsubo, S.Hatta, H.Okuyama, K. Miyamoto, T. Okuda, A. Kimura, H. Namatame, M. Taniguchi, and T. Aruga, Nature Communications 1, 1 (2010). 32C. Jozwiak, Y. L. Chen, A. V. Fedorov, J. G. An- alytis, C. R. Rotundu, A. K. Schmid, J. D. Den- linger, Y.-D. Chuang, D.-H. Lee, I. R. Fisher, R. J. Birgeneau, Z.-X. Shen, Z. Hussain, and A. Lanzara, Phys. Rev. B 84(2011), 10.1103/physrevb.84.165113.11 33F. Kuemmeth, S. Ilani, D. C. Ralph, and P. L. McEuen, Nature , 448 (2004). 34C. Fasth, A. Fuhrer, L. Samuelson, V. N. Golovach, and D. Loss, Phys. Rev. Lett. 98, 266801 (2007). 35R.Wiesendanger,Reviews of Modern Physics 81, 1495 (2009). 36Y. Hasegawa and P. Avouris, Phys. Rev. Lett. 71, 1071 (1993). 37M. F. Crommie, C. P. Lutz, and D. M. Eigler, Nature363, 524 (1993). 38P. T. Sprunger, L. Petersen, E. W. Plummer, E. Lægs- gaard, and F. Besenbacher, Science 275, 1764 (1997). 39L. Petersen, P. Hofmann, E. Plum- mer, and F. Besenbacher, Journal of Electron Spectroscopy and Related Phenomena 109, 97 (2000). 40P. Hofmann, B. G. Briner, M. Doering, H.-P. Rust, E. W. Plummer, and A. M. Bradshaw, Phys. Rev. Lett. 79, 265 (1997). 41F. Vonau, D. Aubel, G. Gewinner, C. Pirri, J. C. Peruchetti, D. Bolmont, and L. Simon, Phys. Rev. B 69, 081305 (2004). 42F. Vonau, D. Aubel, G. Gewinner, S. Zabrocki, J. C. Peruchetti, D. Bolmont, and L. Simon, Phys. Rev. Lett. 95, 176803 (2005). 43L. Simon, C. Bena, F. Vonau, M. Cranney, and D. Aubel, Journal of Physics D: Applied Physics 44, 464010 (2011). 44J. I. Pascual, G. Bihlmayer, Y. M. Koroteev, H.-P. Rust, G. Ceballos, M. Hansmann, K. Horn, E. V. Chulkov, S. Bl¨ ugel, P. M. Echenique, and P. Hofmann, Phys. Rev. Lett. 93, 196802 (2004). 45J. E. Hoffman, K. McElroy, D.-H. Lee, K. M. Lang, H. Eisaki, S. Uchida, and J. C. Davis, Science297, 1148 (2002). 46K. McElroy, R. W. Simmonds, J. E. Hoffman, D.-H. Lee, J. Orenstein, H. Eisaki, S. Uchida, and J. C. Davis, Nature422, 592 (2003). 47M. Vershinin, S. Misra, S. Ono, Y. Abe, Y. Ando, and A. Yazdani, Science 303, 1995 (2004). 48L. Yu, Acta Physica Sinica 21, 75 (1965). 49H. Shiba, Progress of Theoretical Physics 40, 435 (1968). 50A. I. Rusinov, Soviet Journal of Experimental and Theo- retical Physics Letters 9, 85 (1969). 51A. V. Balatsky, I. Vekhter, and J.-X. Zhu, Rev. Mod. Phys. 78, 373 (2006). 52A. Yazdani, B. A. Jones, C. P. Lutz, M. F. Crommie, and D. M. Eigler, Science 275, 1767 (1997).53S.-H. Ji, T. Zhang, Y.-S. Fu, X. Chen, X.-C. Ma, J. Li, W.-H. Duan, J.-F. Jia, and Q.-K. Xue, Phys. Rev. Lett. 100, 226801 (2008). 54N. Hatter, B. W. Heinrich, M. Ruby, J. I. Pascual, and K. J. Franke, Nat. Comm. 6, 8988 (2015). 55G. M´ enard and et al., Nature Physics 11, 1013 (2015). 56M. M. Ugeda, A. J. Bradley, Y. Zhang, S. Onishi, Y. Chen, W. Ruan, C. Ojeda-Aristizabal, H. Ryu, M. T. Edmonds, H.-Z. Tsai, A. Riss, S.-K. Mo, D. Lee, A. Zettl, Z. Hussain, Z.-X. Shen, and M. F. Crommie, Nat Phys 12, 92 (2015). 57P. M. R. Brydon, S. Das Sarma, H.-Y. Hui, and J. D. Sau, Phys. Rev. B 91, 064505 (2015). 58Y. Kim, J. Zhang, E. Rossi, and R. M. Lutchyn, Phys. Rev. Lett. 114, 236804 (2015). 59V. Kaladzhyan, C. Bena, and P. Simon, Phys. Rev. B 93, 214514 (2016). 60V. Kaladzhyan, C. Bena, and P. Simon, arXiv:1606.06338 (2016). 61C. Brun and et al., Nature Physics 10, 444 (2014). 62L. Petersen, L. Brgi, H. Brune, F. Besenbacher, and K. Kern, Surface Science 443, 154 (1999). 63G. Mahan, Many-Particle Physics , Physics of Solids and Liquids (Springer, 2000). 64H. Bruus and K. Flensberg, Many-Body Quantum Theory in Condensed Matter Physics: An In troduction , Oxford Graduate Texts (OUP Oxford, 2004). 65T. Meng, J. Klinovaja, S. Hoffman, P. Simon, and D. Loss, Phys. Rev. B 92, 064503 (2015). 66F. Pientka, L. I. Glazman, and F. von Oppen, Phys. Rev. B 88, 155420 (2013). 67In this paper, we use the plural when refering to Shiba states in order to facilitate the discussion. However, it should be kept in mind that for a given localized mag- netic impurity, there is a single Shiba state with particle and hole components whose real space wave function can actually behave differently. 68A.Sakurai,Progress of Theoretical Physics 44, 1472 (1970). 69M. I. Salkola, A. V. Balatsky, and J. R. Schrieffer, Phys. Rev. B 55, 12648 (1997). 70N. Khotkevych, Y. Kolesnichenko, and J. van Ruitenbeek, arXiv:1601.03154v1 (2016). Appendix A: Analytical calculation of the Shiba states wave functions for a 2D system Wecancalculateanalyticallythenon-polarizedandtheSPLDOSforth eShibastatesexploitingthemodeldescribed by the Hamiltonians in Eqs. (1-3). All the integrations below are perf ormed using a linearization around the Fermi energy. The energies of the Shiba states can be found by solving th e corresponding eigenvalue equation66 [I4−VG0(E,r=0)]Φ(0) = 0 (A1) whereG0(E,r) is the retarded Green’s function in real space obtained by a Fourie r transform from the retarded Green’s function in momentum space G0(E,p) = [(E+iδ)I4−H0(p)]−1, whereδis the inverse quasiparticle lifetime. In all the calculations below we take the limit of δ→+0, and we specify + i0 only in the cases when it affects the results. The wave functions of the Shiba states at r= 0 are given by the eigenfunctions obtained from the equation above. Their spatial dependence is determined using Φ(r) =G0(E,r)VΦ(0) (A2)12 Consequently, the non-polarized and the SP LDOS are given by ρ(E,r) = Φ†(r)/parenleftbigg 0 0 0σ0/parenrightbigg Φ(r), (A3) S(E,r) = Φ†(r)/parenleftbigg 0 0 0σ/parenrightbigg Φ(r). (A4) Thus, in order to find the energies and the wave functions corresp onding to the Shiba states we need to find the real-space Green’s function. This is obtained simply by a Fourier tran sform of the unperturbed Green’s function in momentum space, G0(E,p). We start by writing down the unperturbed Green’s function in mom entum space, which is given by G0(E,p) =1 2/summationtext σ=±Gσ 0(E,p), where Gσ 0(E,p) =−1 ξ2σ+ω2/parenleftbigg 1iσe−iφp −iσeiφp1/parenrightbigg ⊗/parenleftbigg E+ξσ∆s ∆sE−ξσ/parenrightbigg , (A5) whereω=/radicalbig ∆2s−E2, ξσ=ξp+σλp. Toobtainitsreal-spacedependence oneneedstoperformthe Fo uriertransform: Gσ 0(E,r) =/integraldisplaydp (2π)2Gσ 0(E,p)eipr We will have four types of integrals: Xσ 0(r) =−/integraldisplaydp (2π)2eipr ξ2σ+ω2(A6) Xσ 1(r) =−/integraldisplaydp (2π)2ξσeipr ξ2σ+ω2(A7) Xσ 2(s,r) =−/integraldisplaydp (2π)2−isσeisφpeipr ξ2σ+ω2(A8) Xσ 3(s,r) =−/integraldisplaydp (2π)2−isσeisφpξσeipr ξ2σ+ω2(A9) Since the spectrum is split by SO coupling, there will be two Fermi mome nta which can be found the following way: p2 2m+σλp−εF= 0, pσ F=−σλ+/radicalbig λ2+2εF/m 1/m Forp>0 we linearize the spectrum around the Fermi momenta, thus: ξσ≈/parenleftbiggpσ F m+σλ/parenrightbigg (p−pσ F) =/radicalbig λ2+2εF/m(p−pσ F)≡v(p−pσ F), thereforep=pF+ξσ/v, wherev=/radicalbig v2 F+λ2. We rewrite: dp (2π)2=m 2π/bracketleftbigg 1−σλ v/bracketrightbigg dξσdφ 2π=νσdξσdφ 2π, whereνσ=ν/bracketleftbig 1−σλ v/bracketrightbig , withν=m/2π. Due to the symmetry all the integrals are zero at r=0except for the first one, namely, Xσ 0(0) =−νσπ ω. (A10) All the coordinate dependences can be calculated using the formalis m introduced in Ref. 60. Finally we get: Xσ 0(r) =−2νσ·1 ω·ℑK0[−i(1+iΩσ)pσ Fr] (A11) Xσ 1(r) =−2νσ·ℜK0[−i(1+iΩσ)pσ Fr] (A12) Xσ 2(s,r) = 2sσνσ·1 ω·eisφr·ℜK1[−i(1+iΩσ)pσ Fr] (A13) Xσ 3(s,r) =−2sσνσ·eisφr·ℑK1[−i(1+iΩσ)pσ Fr], (A14)13 where Ω σ=ω/pσ Fvdefines the inverse superconducting decay length, and pS=ω/v. Therefore, the Green’s function can be written as Gσ 0(E,r) = EXσ 0(r)+Xσ 1(r)EXσ 2(−,r)+Xσ 3(−,r) ∆ sXσ 0(r) ∆ sXσ 2(−,r) EXσ 2(+,r)+Xσ 3(+,r)EXσ 0(r)+Xσ 1(r) ∆ sXσ 2(+,r) ∆ sXσ 0(r) ∆sXσ 0(r) ∆ sXσ 2(−,r)EXσ 0(r)−Xσ 1(r)EXσ 2(−,r)−Xσ 3(−,r) ∆sXσ 2(+,r) ∆ sXσ 0(r)EXσ 2(+,r)−Xσ 3(+,r)EXσ 0(r)−Xσ 1(r) . (A15) Thus we have: G0(E,r=0) =−πν/radicalbig ∆2s−E2/parenleftbigg Eσ0∆sσ0 ∆sσ0Eσ0/parenrightbigg . (A16) 1. z-impurity The coordinate dependence of the eigenfunctions is given by Φ¯1(r) = +Jz 2/summationdisplay σ=± (E¯1−∆s)Xσ 0(r)+Xσ 1(r) (E¯1−∆s)Xσ 2(+,r)+Xσ 3(+,r) −(E¯1−∆s)Xσ 0(r)+Xσ 1(r) −(E¯1−∆s)Xσ 2(+,r)+Xσ 3(+,r) ,Φ1(r) =−Jz 2/summationdisplay σ=± (E1+∆s)Xσ 2(−,r)+Xσ 3(−,r) (E1+∆s)Xσ 0(r)+Xσ 1(r) (E1+∆s)Xσ 2(−,r)−Xσ 3(−,r) (E1+∆s)Xσ 0(r)−Xσ 1(r) . (A17) Using these expressions we can compute the asymptotic behavior o f the non-polarized and SP LDOS in coordinate space for the state with positive energy (thus we omit index 1 below) : Sx(r) = +J2 z/parenleftbigg 1+1 α2/parenrightbigg/braceleftigg/summationdisplay σσν2 σcos(2pσ Fr−θ) pσ F+2ν2v2 F v2·sinpλr pF/bracerightigg ·e−2psr rcosφr (A18) Sy(r) = +J2 z/parenleftbigg 1+1 α2/parenrightbigg/braceleftigg/summationdisplay σσν2 σcos(2pσ Fr−θ) pσ F+2ν2v2 F v2·sinpλr pF/bracerightigg ·e−2psr rsinφr (A19) Sz(r) =−J2 z/parenleftbigg 1+1 α2/parenrightbigg/braceleftigg/summationdisplay σν2 σsin(2pσ Fr−θ) pσ F−2ν2v2 F v2·cospλr pF/bracerightigg ·e−2psr r(A20) ρ(r) = +J2 z/parenleftbigg 1+1 α2/parenrightbigg/braceleftbigg 2ν2 mv+2ν2v2 F v2·sin(2mvr−θ) pF/bracerightbigg ·e−2psr r(A21) with tanθ=/braceleftigg 2α 1−α2,ifα/ne}ationslash= 1 +∞,ifα= 1,andpλ= 2mλ. Performing the Fourier transforms of these expressions we can obtain information about the main features and symmetries that we observe in momentum space: Sx(p) = +2πiJ2 z/parenleftbigg 1+1 α2/parenrightbigg cosφp+∞/integraldisplay 0drJ1(pr)/braceleftigg/summationdisplay σσν2 σcos(2pσ Fr−θ) pσ F+2ν2v2 F v2·sinpλr pF/bracerightigg ·e−2psr(A22) Sy(p) = +2πiJ2 z/parenleftbigg 1+1 α2/parenrightbigg sinφp+∞/integraldisplay 0drJ1(pr)/braceleftigg/summationdisplay σσν2 σcos(2pσ Fr−θ) pσ F+2ν2v2 F v2·sinpλr pF/bracerightigg ·e−2psr(A23) Sz(p) =−2πJ2 z/parenleftbigg 1+1 α2/parenrightbigg+∞/integraldisplay 0drJ0(pr)/braceleftigg/summationdisplay σν2 σsin(2pσ Fr−θ) pσ F−2ν2v2 F v2·cospλr pF/bracerightigg ·e−2psr(A24) ρ(p) = +2πJ2 z/parenleftbigg 1+1 α2/parenrightbigg+∞/integraldisplay 0drJ0(pr)/braceleftbigg 2ν2 mv+2ν2v2 F v2·sin(2mvr−θ) pF/bracerightbigg ·e−2psr(A25)14 2. x-impurity The coordinate dependence of the eigenfunctions is given by Φ¯1(r) = +Jx 2/summationdisplay σ=± +(E¯1−∆s)[Xσ 0(r)+Xσ 2(−,r)]+Xσ 1(r)+Xσ 3(−,r) +(E¯1−∆s)[Xσ 0(r)+Xσ 2(+,r)]+Xσ 1(r)+Xσ 3(+,r) −(E¯1−∆s)[Xσ 0(r)+Xσ 2(−,r)]+Xσ 1(r)+Xσ 3(−,r) −(E¯1−∆s)[Xσ 0(r)+Xσ 2(+,r)]+Xσ 1(r)+Xσ 3(+,r) , (A26) Φ1(r) =−Jx 2/summationdisplay σ=± +(E1+∆s)[Xσ 0(r)−Xσ 2(−,r)]+Xσ 1(r)−Xσ 3(−,r) −(E1+∆s)[Xσ 0(r)−Xσ 2(+,r)]−Xσ 1(r)+Xσ 3(+,r) +(E1+∆s)[Xσ 0(r)−Xσ 2(−,r)]−Xσ 1(r)+Xσ 3(−,r) −(E1+∆s)[Xσ 0(r)−Xσ 2(+,r)]+Xσ 1(r)−Xσ 3(+,r) . (A27) For the positive energy state we compute the asymptotic behavior of the non-polarized and SP LDOS in coordinate space. We write Sx(r) =Ss x(r)+Sa x(r): Ss x(r) =−J2 x/parenleftbigg 1+1 α2/parenrightbigg/braceleftigg/summationdisplay σν2 σ1+sin(2pσ Fr−2β) pσ F+2ν2v2 F v2·cospλr+sin(2mvr−2β) pF/bracerightigg ·e−2psr r(A28) Sa x(r) = +J2 x/parenleftbigg 1+1 α2/parenrightbigg/braceleftigg/summationdisplay σν2 σ1−sin(2pσ Fr−2β) pσ F−2ν2v2 F v2·cospλr+sin(2mvr−2β) pF/bracerightigg ·e−2psr rcos2φr(A29) Sy(r) = +J2 x/parenleftbigg 1+1 α2/parenrightbigg/braceleftigg/summationdisplay σν2 σ1−sin(2pσ Fr−2β) pσ F−2ν2v2 F v2·cospλr−sin(2mvr−θ) pF/bracerightigg ·e−2psr rsin2φr(A30) Sz(r) =−J2 x/parenleftbigg 1+1 α2/parenrightbigg/braceleftigg 2/summationdisplay σσν2 σcos(2pσ Fr−θ) pσ F+4ν2v2 F v2·sinpλr pF/bracerightigg ·e−2psr rcosφr (A31) ρ(r) = +J2 x/parenleftbigg 1+1 α2/parenrightbigg/braceleftbigg 4ν2 mv+4ν2v2 F v2·sin(2mvr−θ) pF/bracerightbigg ·e−2psr r(A32) with tanβ=α. Sameasbefore, performingthe Fouriertransformsoftheseex pressionsallowsus to obtaininformation about the most important features and symmetries we observe in m omentum space: Ss x(p) =−2πJ2 x/parenleftbigg 1+1 α2/parenrightbigg+∞/integraldisplay 0drJ0(pr)/braceleftigg/summationdisplay σν2 σ1+sin(2pσ Fr−2β) pσ F+ (A33) +2ν2v2 F v2·cospλr+sin(2mvr−2β) pF/bracerightbigg ·e−2psr(A34) Sa x(p) =−2πJ2 x/parenleftbigg 1+1 α2/parenrightbigg cos2φp+∞/integraldisplay 0drJ2(pr)/braceleftigg/summationdisplay σν2 σ1−sin(2pσ Fr−2β) pσ F− (A35) −2ν2v2 F v2·cospλr+sin(2mvr−2β) pF/bracerightbigg ·e−2psr(A36) Sy(p) =−2πJ2 x/parenleftbigg 1+1 α2/parenrightbigg sin2φp+∞/integraldisplay 0drJ2(pr)/braceleftigg/summationdisplay σν2 σ1−sin(2pσ Fr−2β) pσ F− (A37) −2ν2v2 F v2·cospλr−sin(2mvr−θ) pF/bracerightbigg ·e−2psr(A38) Sz(p) =−2πiJ2 x/parenleftbigg 1+1 α2/parenrightbigg cosφp+∞/integraldisplay 0drJ1(pr)/braceleftigg 2/summationdisplay σσν2 σcos(2pσ Fr−θ) pσ F+4ν2v2 F v2·sinpλr pF/bracerightigg ·e−2psr(A39) ρ(p) = +2πJ2 x/parenleftbigg 1+1 α2/parenrightbigg+∞/integraldisplay 0drJ0(pr)/braceleftbigg 4ν2 mv+4ν2v2 F v2·sin(2mvr−θ) pF/bracerightbigg ·e−2psr(A40)15 Appendix B: The SPDOS for a 2D metallic system in the presence of a magnetic impurity The low-energy Hamiltonian can be written as H0=ξpσ0+λ(pyσx−pxσy) =/parenleftbigg ξpiλp− −iλp+ξp/parenrightbigg , (B1) whereξp=p2 2m−εF. The corresponding spectrum is given by E=ξp±λp. The retarded Green’s function reads G0(E,p) =1 (E−ξp+i0)2−λ2p2/parenleftbigg E−ξp+i0iλp− −iλp+E−ξp+i0/parenrightbigg (B2) To compute the eigenvalues for a single localized impurity we calculate G0(E,r=0) =/integraldisplaydp (2π)2E−ξp+i0 (E−ξp+i0)2−λ2p2/parenleftbigg 1 0 0 1/parenrightbigg =1 2/summationdisplay σ/integraldisplaydp (2π)21 E−ξσ+i0/parenleftbigg 1 0 0 1/parenrightbigg , whereξσ=ξp+σλp. Forp>0 we linearize the spectrum around Fermi momenta, thus: ξσ≈/parenleftbiggpσ F m+σλ/parenrightbigg (p−pσ F) =/radicalbig λ2+2εF/m(p−pσ F)≡v(p−pσ F), withpσ F=m[−σλ+v], and thus we rewrite: dp (2π)2=m 2π/bracketleftbigg 1−σλ v/bracketrightbigg dξσdφ 2π=νσdξσdφ 2π, whereνσ=ν/bracketleftbig 1−σλ v/bracketrightbig , withν=m/2π. Thus we get: /integraldisplaydp (2π)21 E−ξσ+i0=νσ/integraldisplay dξσ1 E−ξσ+i0=−iπνσ, and therefore: G0(E,r=0) =1 2/summationdisplay σ(−iπνσ)/parenleftbigg 1 0 0 1/parenrightbigg =−iπν/parenleftbigg 1 0 0 1/parenrightbigg (B3) Since there is no energy dependence, there will be no impurity-induc ed states. To find the coordinate dependence of the Green’s function we calculate: Xσ 0(r) =/integraldisplaydp (2π)2eipr E−ξσ+i0(B4) Xσ 1(s,r) =/integraldisplaydp (2π)2−iseisφpeipr E−ξσ+i0(B5) Below we use the Sokhotsky formula: 1 x+i0=P1 x−iπδ(x) Xσ 0(r) =/integraldisplaydp (2π)2eipr E−ξσ+i0=νσ/integraldisplay dξσ/integraldisplaydφp 2πeiprcos(φp−φr) E−ξσ+i0=νσ/integraldisplay dξσJ0[(pσ F+ξσ/v)r] E−ξσ+i0= =νσ/braceleftbigg P/integraldisplay dξσJ0[(pσ F+ξσ/v)r] E−ξσ−iπ/integraldisplay dξσδ(E−ξσ)J0[(pσ F+ξσ/v)r]/bracerightbigg =♠ We calculate separately the first integral: P/integraldisplay dξσJ0[(pσ F+ξσ/v)r] E−ξσ=2 π+∞/integraldisplay 1du√ u2−1P/integraldisplay dξσsin[(pσ F+ξσ/v)r] E−ξσ= =2 πℑ+∞/integraldisplay 1du√ u2−1P/integraldisplay dξσei(pσ F+ξσ/v)r E−ξσ=2 πℑ+∞/integraldisplay 1du√ u2−1eipσru·P/integraldisplay dxe−ir vx x=♣16 P/integraldisplay dxe−ir vx x=P/integraldisplaycosr vx xdx−iP/integraldisplaysinr vx xdx= 0−iπ=−iπ Therefore: ♣=−2ℑ+∞/integraldisplay 1ieipσru √ u2−1du=−2+∞/integraldisplay 1cospσru√ u2−1du=πY0(pσr), pσ/ne}ationslash= 0 ♠=πνσ[Y0(pσr)−iJ0(pσr)]. The second integral is Xσ 1(s,r) =/integraldisplaydp (2π)2−iseisφpeipr E−ξσ+i0=νσ/integraldisplay dξσ/integraldisplaydφp 2π−iseisφpeiprcos(φp−φr) E−ξσ+i0=seisφrνσ/integraldisplay dξσJ1[(pσ F+ξσ/v)r] E−ξσ+i0= =seisφr·νσ/braceleftbigg P/integraldisplay dξσJ1[(pσ F+ξσ/v)r] E−ξσ−iπ/integraldisplay dξσδ(E−ξσ)J1[(pσ F+ξσ/v)r]/bracerightbigg =♥ We calculate separately the first integral: P/integraldisplay dξσJ1[(pσ F+ξσ/v)r] E−ξσ=P/integraldisplay dxJ1[(pσ−x/v)r] x=−∂ ∂rP/integraldisplay dxJ0[(pσ−x/v)r] x(pσ−x/v)= =−∂ ∂rP/integraldisplay dyJ0[(pσ−y)r] y(pσ−y)=−∂ ∂(pσr)/bracketleftbigg P/integraldisplay dyJ0[(pσ−y)r] y+P/integraldisplay dyJ0[(pσ−y)r] pσ−y/bracketrightbigg = =−∂ ∂(pσr)2 πℑ+∞/integraldisplay 1du√ u2−1/bracketleftbigg P/integraldisplayei(pσ−y)ru ydy+P/integraldisplayei(pσ−y)ru pσ−ydy/bracketrightbigg =−2∂ ∂(pσr)ℑ+∞/integraldisplay 1idu√ u2−1/bracketleftbig 1−eipσru/bracketrightbig = =−2+∞/integraldisplay 1usinpσru√ u2−1du= 2∂ ∂(pσr)+∞/integraldisplay 1cospσru√ u2−1du=−π∂ ∂(pσr)Y0(pσr) =πY1(pσr), pσ/ne}ationslash= 0 Therefore: ♥=πνσ[Y1(pσr)−iJ1(pσr)]. Finally: Xσ 0(r) =πνσ[Y0(pσr)−iJ0(pσr)] (B6) Xσ 1(s,r) =seisφr/braceleftig πνσ[Y1(pσr)−iJ1(pσr)]/bracerightig ≡seisφr˜Xσ 1(r), (B7) wherepσ=pσ F+E/v/ne}ationslash= 0. Thus the Green’s function for r/ne}ationslash=0can be written as: G0(E,r) =1 2/summationdisplay σ/parenleftbigg Xσ 0(r)−σe−iφr˜Xσ 1(r) σeiφr˜Xσ 1(r)Xσ 0(r)/parenrightbigg (B8) Below we compute the T-matrix for different types of impurities. Imp urity potentials take the following forms: Vsc=U/parenleftbigg 1 0 0 1/parenrightbigg , Vz=Jz/parenleftbigg 1 0 0−1/parenrightbigg , Vx=Jx/parenleftbigg 0 1 1 0/parenrightbigg (B9) The corresponding T-matrices are Tsc=U 1+iπνU/parenleftbigg 1 0 0 1/parenrightbigg , Tz=/parenleftbiggJ 1+iπνJ0 0−J 1−iπνJ/parenrightbigg , T x=J 1+π2ν2J2/parenleftbigg −iπνJ 1 1−iπνJ/parenrightbigg (B10) For each type of impurity we can compute the SP and non-polarized L DOS using ∆G(E,r) =G0(E,−r)T(E)G0(E,r) (B11)17 Sx(E,r) =−1 π[ℑ∆G12+ℑ∆G21] (B12) Sy(E,r) =−1 π[ℜ∆G12−ℜ∆G21] (B13) Sz(E,r) =−1 π[ℑ∆G11−ℑ∆G22] (B14) ∆ρ(E,r) =−1 π[ℑ∆G11+ℑ∆G22] (B15) Asymptotic expansions of Bessel functions Since the integrals are expressed in terms of Neumann function and Bessel function of the first kind, we give their asymptotic behavior for x→+∞: J0(x)∼+/radicalbigg 2 πxcos/parenleftig x−π 4/parenrightig , J 1(x)∼ −/radicalbigg 2 πxcos/parenleftig x+π 4/parenrightig Y0(x)∼ −/radicalbigg 2 πxcos/parenleftig x+π 4/parenrightig , Y 1(x)∼ −/radicalbigg 2 πxcos/parenleftig x−π 4/parenrightig Fourier transforms in 2D F[f(r)] = 2π+∞/integraldisplay 0rJ0(pr)f(r)dr (B16) F[cosφrf(r)] = 2πicosφp·+∞/integraldisplay 0rJ1(pr)f(r)dr,F[sinφrf(r)] = 2πisinφp·+∞/integraldisplay 0rJ1(pr)f(r)dr(B17) F[cos2φrf(r)] =−2πcos2φp+∞/integraldisplay 0rJ2(pr)f(r)dr,F[sin2φrf(r)] =−2πsin2φp+∞/integraldisplay 0rJ2(pr)f(r)dr(B18) 1. z-impurity We denote α=πνJand write the asymptotic expansions of the non-polarized and SP LD OS components in coordinate space: Sx(r)∼J 1+α2cosφr r/summationdisplay σσν2 σ pσsin2pσr (B19) Sy(r)∼J 1+α2sinφr r/summationdisplay σσν2 σ pσsin2pσr (B20) Sz(r)∼ −J 1+α22 r/summationdisplay σν2 σ pσcos2pσr (B21) ρ(r)∼ −J 1+α24αν2v2 F v21/radicalbig p2 F+2mE+E2/v2·sinpεr r, (B22)18 wherepε= 2(mv+E/v). and we get for pσ>0: Sx(p)∼+J 1+α2·2πicosφp+∞/integraldisplay 0drJ1(pr)/summationdisplay σσν2 σ pσsin2pσr (B23) Sy(p)∼+J 1+α2·2πisinφp+∞/integraldisplay 0drJ1(pr)/summationdisplay σσν2 σ pσsin2pσr (B24) Sz(p)∼ −J 1+α2·4π+∞/integraldisplay 0drJ0(pr)/summationdisplay σν2 σ pσcos2pσr (B25) ρ(p)∼ −J 1+α2·8παν2v2 F v21/radicalbig p2 F+2mE+E2/v2+∞/integraldisplay 0drJ0(pr)sinpεr (B26) 2. x-impurity Sx(r)∼ −J 1+α21 r/braceleftigg 2ν2v2 F v2cospεr/radicalbig p2 F+2mE+E2/v2+/summationdisplay σν2 σ pσcos2pσr+ (B27) +cos2φr/bracketleftigg −2ν2v2 F v2cospεr/radicalbig p2 F+2mE+E2/v2+/summationdisplay σν2 σ pσcos2pσr/bracketrightigg/bracerightigg (B28) Sy(r)∼ −J 1+α2sin2φr r/bracketleftigg −2ν2v2 F v2cospεr/radicalbig p2 F+2mE+E2/v2+/summationdisplay σν2 σ pσcos2pσr/bracketrightigg (B29) Sz(r)∼ −J 1+α2cosφr r/summationdisplay σσν2 σ pσsin2pσr (B30) ρ(r)∼ −J 1+α2·α r·4ν2v2 F v2sinpεr/radicalbig p2 F+2mE+E2/v2(B31) With the corresponding Fourier transforms: Sx(p) =Ssym x(p)+Sasym x(p) =−J 1+α2·2π+∞/integraldisplay 0drJ0(pr)/bracketleftigg 2ν2v2 F v2cospεr/radicalbig p2 F+2mE+E2/v2+/summationdisplay σν2 σ pσcos2pσr/bracketrightigg −(B32) −J 1+α2·2πcos2φp+∞/integraldisplay 0drJ2(pr)/bracketleftigg 2ν2v2 F v2cospεr/radicalbig p2 F+2mE+E2/v2−/summationdisplay σν2 σ pσcos2pσr/bracketrightigg (B33) Sy(p) =−J 1+α2·2πsin2φp+∞/integraldisplay 0drJ2(pr)/bracketleftigg 2ν2v2 F v2cospεr/radicalbig p2 F+2mE+E2/v2−/summationdisplay σν2 σ pσcos2pσr/bracketrightigg (B34) Sz(p)∼ −J 1+α2·2πicosφp+∞/integraldisplay 0drJ1(pr)/summationdisplay σσν2 σ pσsin2pσr (B35) ρ(p)∼ −J 1+α2·8παν2v2 F v21/radicalbig p2 F+2mE+E2/v2+∞/integraldisplay 0drJ0(pr)sinpεr (B36)19 Appendix C: Analytical calculation of the Shiba states wave functions for a 1D system The unperturbed Green’s function in momentum space is G0(E,p) =1 2/summationtext σ=±Gσ 0(E,p), where Gσ 0(E,p) =−1 ξ2σ+∆2s−E2/parenleftbigg 1iσ −iσ1/parenrightbigg ⊗/parenleftbigg E+ξσ∆s ∆sE−ξσ/parenrightbigg , (C1) whereξσ=ξp+σλp. To get the coordinate value one needs to perform the Fourier tra nsform: Gσ 0(E,x) =/integraldisplaydp 2πGσ 0(E,p)eipx We will have two types of integrals: Xσ 0(x) =−/integraldisplaydp 2πeipx ξ2σ+ω2, (C2) Xσ 1(x) =−/integraldisplaydp 2πξσeipx ξ2σ+ω2, (C3) whereω2= ∆2 s−E2. Since the spectrum is split by SO coupling, there will be two Fermi mom enta which can be found the following way: p2 2m+σλp−εF= 0, pσ F=−σλ+/radicalbig λ2+2εF/m 1/m≡m[−σλ+v] Forp>0 we linearize the spectrum around Fermi momenta, thus: ξσ≈/parenleftbiggpσ F m+σλ/parenrightbigg (p−pσ F) =/radicalbig λ2+2εF/m(p−pσ F)≡v(p−pσ F), thereforep=pσ F+ξσ/vand we get: Xσ 0(x) =−/integraldisplaydp 2πeipx ξ2σ+ω2=− +∞/integraldisplay 0dp 2πeipx ξ2σ+ω2++∞/integraldisplay 0dp 2πe−ipx ξ2 −σ+ω2 =♣ +∞/integraldisplay 0dp 2πeipx ξ2σ+ω2≈1 2πveipσ Fx/integraldisplay dξσeiξσx/v ξ2σ+ω2=1 2vωeipσ Fxe−ω|x|/v +∞/integraldisplay 0dp 2πe−ipx ξ2 −σ+ω2≈1 2πve−ip−σ Fx/integraldisplay dξ−σe−iξ−σx/v ξ2 −σ+ω2=1 2vωe−ip−σ Fxe−ω|x|/v ♣=−1 2vω/bracketleftig eim[−σλ+v]x+e−im[σλ+v]x/bracketrightig e−ω|x|/v=−1 v·1 ωcosmvxe−iσmλxe−ω|x|/v Xσ 1(x) =−/integraldisplaydp 2πξσeipx ξ2σ+ω2=− +∞/integraldisplay 0dp 2πξσeipx ξ2σ+ω2++∞/integraldisplay 0dp 2πξ−σe−ipx ξ2 −σ+ω2 =♠ +∞/integraldisplay 0dp 2πξσeipx ξ2σ+ω2≈1 2πveipσ Fx/integraldisplay dξσξσeiξσx/v ξ2σ+ω2=i 2vsgnxeipσ Fxe−ω|x|/v +∞/integraldisplay 0dp 2πξ−σe−ipx ξ2 −σ+ω2≈1 2πve−ip−σ Fx/integraldisplay dξ−σξ−σe−iξ−σx/v ξ2 −σ+ω2=−i 2vsgnxe−ip−σ Fxe−ω|x|/v20 ♠=−i 2vsgnx/bracketleftig eim[−σλ+v]x−e−im[σλ+v]x/bracketrightig e−ω|x|/v=1 v·sinmv|x|e−iσmλxe−ω|x|/v Finally: Xσ 0(x) =−1 v·1 ωcosmvxe−iσmλxe−ω|x|/v(C4) Xσ 1(x) = +1 v·sinmv|x|e−iσmλxe−ω|x|/v(C5) and G0(E,x) =1 2/summationdisplay σ=±/parenleftbigg 1iσ −iσ1/parenrightbigg ⊗/parenleftbigg EXσ 0(x)+Xσ 1(x) ∆ sXσ 0(x) ∆sXσ 0(x)EXσ 0(x)−Xσ 1(x)/parenrightbigg (C6) G0(ǫ,x= 0) =−1 v1√ 1−ǫ2/parenleftbigg ǫσ0σ0 σ0ǫσ0/parenrightbigg ,whereǫ=E ∆s(C7) The eigenvalues and eigenfunctions at r=0can be obtained using Eq. (10) The energy levels are E1,¯1=±1−α2 1+α2∆s,whereα=J/v. (C8) In case of an impurity along the z-axis the corresponding eigenvectors are Φ¯1(0) =/parenleftbig1 0−1 0/parenrightbigT,Φ1(0) =/parenleftbig0 1 0 1/parenrightbigT(C9) and in case of an impurity along the x-axis: Φ¯1(0) =/parenleftbig1 1−1−1/parenrightbigT,Φ1(0) =/parenleftbig1−1 1−1/parenrightbigT. (C10) 1. z-impurity Φ¯1(x) = +Jz 2/summationdisplay σ +(E¯1−∆s)Xσ 0(x)+Xσ 1(x) −iσ[(E¯1−∆s)Xσ 0(x)+Xσ 1(x)] −(E¯1−∆s)Xσ 0(x)+Xσ 1(x) +iσ[(E¯1−∆s)Xσ 0(x)−Xσ 1(x)] ,Φ1(x) =−Jz 2/summationdisplay σ +iσ[(E1+∆s)Xσ 0(x)+Xσ 1(x)] (E1+∆s)Xσ 0(x)+Xσ 1(x) +iσ[(E1+∆s)Xσ 0(x)−Xσ 1(x)] (E1+∆s)Xσ 0(x)−Xσ 1(x) . (C11) Using these expressions we can compute the non-polarized and SP L DOS in both coordinate and momentum space for the positive energy state (omitting the index 1): Sx(x) =1+α2 4[2sinpλx+sin(2mv|x|+pλx−2θ)−sin(2mv|x|−pλx−2θ)]·e−2ω|x|/v(C12) Sy(x) = 0 (C13) Sz(x) =−1+α2 4[2cospλx+cos(2mv|x|+pλx−2θ)+cos(2mv|x|−pλx−2θ)]·e−2ω|x|/v(C14) ρ(x) =1+α2 2[1+cos(2mv|x|−2θ)]·e−2ω|x|/v(C15)21 where tanθ=α. We perform the Fourier transform to get the momentum space be havior, exploiting the following ’standard’ integrals: /integraldisplay e−2ω|x|/ve−ipxdx= 22ω/v p2+(2ω/v)2(C16) /integraldisplay cospλx·e−2ω|x|/ve−ipxdx=2ω v/bracketleftbigg1 (p+pλ)2+(2ω/v)2+1 (p−pλ)2+(2ω/v)2/bracketrightbigg (C17) /integraldisplay sinpλx·e−2ω|x|/ve−ipxdx=i2ω v/bracketleftbigg1 (p+pλ)2+(2ω/v)2−1 (p−pλ)2+(2ω/v)2/bracketrightbigg (C18) /integraldisplay sin2mv|x|·e−2ω|x|/ve−ipxdx=p+2mv (p+2mv)2+(2ω/v)2−p−2mv (p−2mv)2+(2ω/v)2(C19) We rewrite these expressions using p± F, thus we get: /integraldisplay cospλx·e−2ω|x|/ve−ipxdx=2ω v/braceleftigg 1 /bracketleftbig p+(p− F−p+ F)/bracketrightbig2+(2ω/v)2+1 (/bracketleftbig p−(p− F−p+ F)/bracketrightbig2+(2ω/v)2/bracerightigg (C20) /integraldisplay sinpλx·e−2ω|x|/ve−ipxdx=i2ω v/braceleftigg 1 /bracketleftbig p+(p− F−p+ F)/bracketrightbig2+(2ω/v)2−1 /bracketleftbig p−(p− F−p+ F)/bracketrightbig2+(2ω/v)2/bracerightigg (C21) /integraldisplay sin2mv|x|·e−2ω|x|/ve−ipxdx=p+(p− F+p+ F) /bracketleftbig p+(p− F+p+ F)/bracketrightbig2+(2ω/v)2−p−(p− F+p+ F) /bracketleftbig p−(p− F+p+ F)/bracketrightbig2+(2ω/v)2(C22) For the last two integrals we introduce symbols/summationtext p′and/tildewider/summationtext p′(wide tilde signify that we take the difference, not sum), wherep′∈ {p−pλ,p+pλ}. Thus we have /integraldisplay cos(2mv|x|−2θ)cospλx·e−2ω|x|/ve−ipxdx= (C23) =1 2/summationdisplay p′/braceleftbigg1−α2 1+α2·2ω v/bracketleftbigg1 (p′+2mv)2+(2ω/v)2+1 (p′−2mv)2+(2ω/v)2/bracketrightbigg + (C24) +2α 1+α2·/bracketleftbiggp′+2mv (p′+2mv)2+(2ω/v)2+p′−2mv (p′−2mv)2+(2ω/v)2/bracketrightbigg/bracerightbigg (C25) /integraldisplay cos(2mv|x|−2θ)sinpλx·e−2ω|x|/ve−ipxdx= (C26) =1 2i/tildewidest/summationdisplay p′/braceleftbigg1−α2 1+α2·2ω v/bracketleftbigg1 (p′+2mv)2+(2ω/v)2+1 (p′−2mv)2+(2ω/v)2/bracketrightbigg + (C27) +2α 1+α2·/bracketleftbiggp′+2mv (p′+2mv)2+(2ω/v)2+p′−2mv (p′−2mv)2+(2ω/v)2/bracketrightbigg/bracerightbigg (C28) We rewrite these expressions using p± F, thus we get: /integraldisplay cos(2mv|x|−2θ)cospλx·e−2ω|x|/ve−ipxdx= (C29) =1−α2 1+α2·ω v/bracketleftbigg1 (p+2p+ F)2+(2ω/v)2+1 (p−2p− F)2+(2ω/v)2/bracketrightbigg + +α 1+α2·/bracketleftbiggp+2p+ F (p+2p+ F)2+(2ω/v)2+p−2p− F (p−2p− F)2+(2ω/v)2/bracketrightbigg + +1−α2 1+α2·ω v/bracketleftbigg1 (p+2p− F)2+(2ω/v)2+1 (p−2p+ F)2+(2ω/v)2/bracketrightbigg + +α 1+α2·/bracketleftbiggp+2p− F (p+2p− F)2+(2ω/v)2+p−2p+ F (p−2p+ F)2+(2ω/v)2/bracketrightbigg22 /integraldisplay cos(2mv|x|−2θ)sinpλx·e−2ω|x|/ve−ipxdx= (C30) =1 i/braceleftbigg1−α2 1+α2·ω v/bracketleftbigg1 (p+2p+ F)2+(2ω/v)2+1 (p−2p− F)2+(2ω/v)2/bracketrightbigg + +α 1+α2·/bracketleftbiggp+2p+ F (p+2p+ F)2+(2ω/v)2+p−2p− F (p−2p− F)2+(2ω/v)2/bracketrightbigg/bracerightbigg − −1 i/braceleftbigg1−α2 1+α2·ω v/bracketleftbigg1 (p+2p− F)2+(2ω/v)2+1 (p−2p+ F)2+(2ω/v)2/bracketrightbigg + +α 1+α2·/bracketleftbiggp+2p− F (p+2p− F)2+(2ω/v)2+p−2p+ F (p−2p+ F)2+(2ω/v)2/bracketrightbigg/bracerightbigg Using the formula cos2γ= (1+cos2 γ)/2 we can write the momentum space expressions for the non-polariz ed and SP LDOS components: Sx(p) =i(1+α2)ω v/braceleftigg 1 /bracketleftbig p+(p− F−p+ F)/bracketrightbig2+(2ω/v)2−1 /bracketleftbig p−(p− F−p+ F)/bracketrightbig2+(2ω/v)2/bracerightigg + (C31) +1 i/braceleftbigg1−α2 2·ω v/bracketleftbigg1 (p+2p+ F)2+(2ω/v)2+1 (p−2p− F)2+(2ω/v)2/bracketrightbigg + +α 2·/bracketleftbiggp+2p+ F (p+2p+ F)2+(2ω/v)2+p−2p− F (p−2p− F)2+(2ω/v)2/bracketrightbigg/bracerightbigg − −1 i/braceleftbigg1−α2 2·ω v/bracketleftbigg1 (p+2p− F)2+(2ω/v)2+1 (p−2p+ F)2+(2ω/v)2/bracketrightbigg + +α 2·/bracketleftbiggp+2p− F (p+2p− F)2+(2ω/v)2+p−2p+ F (p−2p+ F)2+(2ω/v)2/bracketrightbigg/bracerightbigg Sz(p) =−(1+α2)ω v/braceleftigg 1 /bracketleftbig p+(p− F−p+ F)/bracketrightbig2+(2ω/v)2+1 /bracketleftbig p−(p− F−p+ F)/bracketrightbig2+(2ω/v)2/bracerightigg − (C32) −1−α2 2·ω v/bracketleftbigg1 (p+2p+ F)2+(2ω/v)2+1 (p−2p− F)2+(2ω/v)2/bracketrightbigg − −α 2·/bracketleftbiggp+2p+ F (p+2p+ F)2+(2ω/v)2−p−2p− F (p−2p− F)2+(2ω/v)2/bracketrightbigg − −1−α2 2·ω v/bracketleftbigg1 (p+2p− F)2+(2ω/v)2+1 (p−2p+ F)2+(2ω/v)2/bracketrightbigg − −α 2·/bracketleftbiggp+2p− F (p+2p− F)2+(2ω/v)2+p−2p+ F (p−2p+ F)2+(2ω/v)2/bracketrightbigg ρ(p) = (1+α2)/braceleftigg 2ω/v p2+(2ω/v)2+/bracketleftigg ω/v /bracketleftbig p+(p− F+p+ F)/bracketrightbig2+(2ω/v)2+ω/v (/bracketleftbig p−(p− F+p+ F)/bracketrightbig2+(2ω/v)2/bracketrightigg/bracerightigg + (C33) +α/braceleftigg p+(p− F+p+ F) /bracketleftbig p+(p− F+p+ F)/bracketrightbig2+(2ω/v)2−p−(p− F+p+ F) /bracketleftbig p−(p− F+p+ F)/bracketrightbig2+(2ω/v)2/bracerightigg23 2. x-impurity Φ¯1(x) = +Jx 2/summationdisplay σ (1+iσ)[(E¯1−∆s)Xσ 0(x)+Xσ 1(x)] (1−iσ)[(E¯1−∆s)Xσ 0(x)+Xσ 1(x)] −(1+iσ)[(E¯1−∆s)Xσ 0(x)−Xσ 1(x)] −(1−iσ)[(E¯1−∆s)Xσ 0(x)−Xσ 1(x)] ,Φ1(x) = +Jx 2/summationdisplay σ −(1−iσ)[(E1+∆s)Xσ 0(x)+Xσ 1(x)] (1+iσ)[(E1+∆s)Xσ 0(x)+Xσ 1(x)] −(1−iσ)[(E1+∆s)Xσ 0(x)−Xσ 1(x)] (1+iσ)[(E1+∆s)Xσ 0(x)−Xσ 1(x)] . (C34) Using these expressions we can compute the non-polarized and SP L DOS in both coordinate and momentum space. We perform the calculation for the positive-energy state, and we fi nd, omitting index 1: Sx(x) =−1+α2 2[2cospλx+cos(2mv|x|+pλx−2θ)+cos(2mv|x|−pλx−2θ)]·e−2ω|x|/v(C35) Sy(x) = 0 (C36) Sz(x) =−1+α2 2[2sinpλx+sin(2mv|x|+pλx−2θ)−sin(2mv|x|−pλx−2θ)]·e−2ω|x|/v(C37) ρ(x) = (1+α2)[1+cos(2mv|x|−2θ)]·e−2ω|x|/v(C38) where tanθ=α. Momentum space dependence can be derived from the z-impurity expressions since everything coincides up to coefficients. 3. y-impurity Φ¯1(x) = +Jy 2/summationdisplay σ (1−σ)[(E¯1−∆s)Xσ 0(x)+Xσ 1(x)] i(1−σ)[(E¯1−∆s)Xσ 0(x)+Xσ 1(x)] −(1−σ)[(E¯1−∆s)Xσ 0(x)−Xσ 1(x)] −i(1−σ)[(E¯1−∆s)Xσ 0(x)−Xσ 1(x)] ,Φ1(x) = +Jy 2/summationdisplay σ −(1+σ)[(E1+∆s)Xσ 0(x)+Xσ 1(x)] i(1+σ)[(E1+∆s)Xσ 0(x)+Xσ 1(x)] −(1+σ)[(E1+∆s)Xσ 0(x)−Xσ 1(x)] i(1+σ)[(E1+∆s)Xσ 0(x)−Xσ 1(x)] , (C39) after summation over σ: Φ¯1(x) = +Jy +/bracketleftbig (E¯1−∆s)X− 0(x)+X− 1(x)/bracketrightbig i/bracketleftbig (E¯1−∆s)X− 0(x)+X− 1(x)/bracketrightbig −/bracketleftbig (E¯1−∆s)X− 0(x)−X− 1(x)/bracketrightbig −i/bracketleftbig (E¯1−∆s)X− 0(x)−X− 1(x)/bracketrightbig ,Φ1(x) = +Jy −/bracketleftbig (E1+∆s)X+ 0(x)+X+ 1(x)/bracketrightbig i/bracketleftbig (E1+∆s)X+ 0(x)+X+ 1(x)/bracketrightbig −/bracketleftbig (E1+∆s)X+ 0(x)−X+ 1(x)/bracketrightbig i/bracketleftbig (E1+∆s)X+ 0(x)−X+ 1(x)/bracketrightbig .(C40) Using these expressions we can compute the non-polarized and SP L DOS in coordinate space Sx(x) = 0, (C41) Sy(x) =−(1+α2)[1+cos(2mv|x|−2θ)]·e−2ω|x|/v, (C42) Sz(x) = 0, (C43) ρ(x) = +(1+α2)[1+cos(2mv|x|−2θ)]·e−2ω|x|/v. (C44) Appendix D: The SPDOS for a non-superconducting one-dimens ional system in the presence of a magnetic impurity The low-energy Hamiltonian in the non-SC regime can be written as H0=ξpσ0+λ(pyσx−pxσy) =/parenleftbigg ξpiλp −iλp ξ p/parenrightbigg (D1)24 whereξp=p2 2m−εF. The corresponding spectrum is given by E=ξp±λpand the retarded Green’s function reads G0(E,p) =1 (E−ξp+i0)2−λ2p2/parenleftbigg E−ξp+i0iλp −iλp E −ξp+i0/parenrightbigg . (D2) To compute the eigenvalues for a single localized impurity we calculate G0(E,x= 0) =/integraldisplaydp 2πE−ξp+i0 (E−ξp+i0)2−λ2p2/parenleftbigg 1 0 0 1/parenrightbigg =1 2/summationdisplay σ/integraldisplaydp 2π1 E−ξσ+i0/parenleftbigg 1 0 0 1/parenrightbigg , (D3) whereξσ=ξp+σλp. Forp>0 we linearize the spectrum around the Fermi momenta, thus: ξσ≈/parenleftbiggpσ F m+σλ/parenrightbigg (p−pσ F) =/radicalbig λ2+2εF/m(p−pσ F)≡v(p−pσ F), wherepσ F=m[−σλ+v], and thus we get: /integraldisplaydp 2π1 E−ξσ+i0≈1 2πv/bracketleftbigg/integraldisplaydξσ E−ξσ+i0+/integraldisplaydξ−σ E−ξ−σ+i0/bracketrightbigg =−i v This leads to: G0(E,x= 0) =1 2/summationdisplay σ/parenleftbigg −i v/parenrightbigg/parenleftbigg 1 0 0 1/parenrightbigg =−i v/parenleftbigg 1 0 0 1/parenrightbigg (D4) Since there is no energy dependence, there will be no impurity-induc ed states. The Green’s function coordinate dependence is given by the following expression: G0(E,x) =1 2/summationdisplay σ/integraldisplaydp 2πeipx E−ξσ+i0/parenleftbigg 1iσ −iσ1/parenrightbigg (D5) To find the coordinate dependence of the Green’s function we calcu late: Xσ 0(x) =/integraldisplaydp 2πeipx E−ξσ+i0(D6) Integral calculation Below we use the Sokhotsky formula1 x+i0=P1 x−iπδ(x): Xσ 0(x) =/integraldisplaydp 2πeipx E−ξσ+i0=1 2πv/bracketleftbigg eipσ Fx/integraldisplay dξσeiξσx/v E−ξσ+i0+e−ip−σ Fx/integraldisplay dξ−σe−iξ−σx/v E−ξ−σ+i0/bracketrightbigg We compute explicitly only one of the integrals in the brackets since th e other one can be computed in the similar fashion: /integraldisplay dξσeiξσx/v E−ξσ+i0=P/integraldisplay dξσeiξσx/v E−ξσ−iπ/integraldisplay dξσδ(E−ξσ)eiξσx/v=−iπ(1+sgnx)eiEx/v Finally we have: Xσ 0(x) =−i vexp/bracketleftbigg i/parenleftbigg mv+E v/parenrightbigg |x|/bracketrightbigg e−iσmλx, (D7) and the Green’s function can be written as: G0(E,x) =1 2/summationdisplay σ/parenleftbigg 1iσ −iσ1/parenrightbigg Xσ 0(x). (D8)25 Below we compute the T-matrix for different types of impurities. Imp urity potentials take the following forms: Vsc=U/parenleftbigg 1 0 0 1/parenrightbigg , Vz=Jz/parenleftbigg 1 0 0−1/parenrightbigg , Vx=Jx/parenleftbigg 0 1 1 0/parenrightbigg (D9) The corresponding T-matrices are Tsc=U 1+iU/v/parenleftbigg 1 0 0 1/parenrightbigg , Tz=/parenleftiggJ 1+iJ/v0 0−J 1−iJ/v/parenrightigg , T x=J 1+J2/v2/parenleftbigg −iJ/v1 1−iJ/v/parenrightbigg (D10) For each type of impurity we can compute the non-polarized and SP L DOS using Eq. (B11) and Eqs. (B15) where we replace rbyx. By taking the Fourier transforms of the expressions above we ge t the the momentum space dependence. Below we denote α=J/v. 1. z-impurity Sx(x) = +α 1+α2·1 πv[cos(pε|x|−pλx)−cos(pε|x|+pλx)] (D11) Sy(x) = 0 (D12) Sz(x) = +α 1+α2·1 πv[sin(pε|x|−pλx)+sin(pε|x|+pλx)] (D13) ρ(x) =−2α2 1+α2·1 πvcospεx (D14) where we denote pε= 2(mv+E/v),pλ= 2mλ. After taking the Fourier transform we get: Sx(p) = +α 1+α2·i πv/bracketleftbigg1 p+pε+pλ−1 p+pε−pλ−1 p−pε+pλ+1 p−pε−pλ/bracketrightbigg (D15) Sy(p) = 0 (D16) Sz(p) = +α 1+α2·1 πv/bracketleftbigg1 p+pε+pλ+1 p+pε−pλ−1 p−pε+pλ−1 p−pε−pλ/bracketrightbigg (D17) ρ(p) =−2α2 1+α2·1 v[δ(p−pε)+δ(p+pε)] (D18) 2. x-impurity Sx(x) = +α 1+α2·1 πv[sin(pε|x|−pλx)+sin(pε|x|+pλx)] (D19) Sy(x) = 0 (D20) Sz(x) =−α 1+α2·1 πv[cos(pε|x|−pλx)−cos(pε|x|+pλx)] (D21) ρ(x) =−2α2 1+α2·1 πvcospεx (D22) We do not give the Fourier transform for these expressions since t hey coincide with the ones for a z-impurity if we exchangeSzandSxand change the overall sign.26 3. y-impurity Sx(x) =Sz(x) = 0 (D23) Sy(x) = +2α 1+α2·1 πvsinpε|x| (D24) ρ(x) =−2α2 1+α2·1 πvcospεx (D25) The corresponding Fourier transform is: Sy(p) =2α 1+α2·1 πv/bracketleftbigg1 p+pε−1 p−pε/bracketrightbigg (D26)
1403.4728v1.Spin_orbit_coupling_effects_on_spin_dependent_inelastic_electronic_lifetimes_in_ferromagnets.pdf
arXiv:1403.4728v1 [cond-mat.mtrl-sci] 19 Mar 2014Spin-Orbit Coupling Effects on Spin Dependent Inelastic Ele ctronic Lifetimes in Ferromagnets Steffen Kaltenborn and Hans Christian Schneider∗ Physics Department and Research Center OPTIMAS, University of Kaiserslautern, 67663 Kaiserslautern, Germ any (Dated: July 4, 2021) For the 3d ferromagnets iron, cobalt and nickel we compute th e spin-dependentinelastic electronic lifetimes due to carrier-carrier Coulomb interaction incl uding spin-orbit coupling. We find that the spin-dependent density-of-states at the Fermi energy d oes not, in general, determine the spin dependence of the lifetimes because of the effective spin-fli p transitions allowed by the spin mixing. The majority and minority electron lifetimes computed incl uding spin-orbit coupling for these three 3-d ferromagnets do not differ by more than a factor of 2, and ag ree with experimental results. PACS numbers: 71.70.Ej,75.76.+j,75.78.-n,85.75.-d I. INTRODUCTION The theoretical and experimental characterization of spin dynamics in ferromagnetic materials due to the in- teraction with short optical pulses has become an impor- tant part of research in magnetism.1–6In this connec- tion, spin-dependent hot-electron transport processes in metallic heterostructures have received enormous inter- est in the past few years.7In particular, superdiffusive- transport theory has played an increasingly important role in the quantitative interpretationof experimental re- sults.4,6,8Superdiffusive transport-theory, which was in- troduced and comprehensively described in Refs. 9and 10, uses spin- and energy-dependent electron lifetimes as input,10and its quantitative results for hot-electron transportonultrashorttimescalesinferromagneticmate- rials rely heavily, to the best of our knowledge, on the re- lation between majority and minority electrons for these materials. The spin-dependent lifetimes that are used for hot- electrontransport,bothinferromagnetsandnormalmet- als, are the so-called “inelastic lifetimes.” These state (orenergy)dependentlifetimesresultfromout-scattering processes due to the Coulomb interaction between an ex- cited electron and the inhomogeneous electron gas in the system. These lifetimes can be measured by tracking op- tically excited electrons using spin- and time-resolved 2- photon photoemission (2PPE)11,12and can be calculated asthebroadeningoftheelectronicspectralfunctionusing many-body Green function techniques.13,14The problem oftheaccuratedeterminationoftheselifetimes hasfueled method development on the experimental and theoreti- cal side,15but has always suffered from the presence of interactions (electron-phonon, surface effects) that can- not be clearly identified in experiment and are difficult to include in calculations. Qualitative agreement was reached for the spin-integrated lifetimes in simple met- als and iron,16but even advanced quasiparticle calcula- tions including many-body T-matrix contributions, have yielded a ratio between majority and minority lifetimes, which is in qualitative disagreement with experiment forsome ferromagnets. A particularly important material in recent studies has been nickel,4,6,10for which the the- oretical ratio comes out between 6 and 8,16while the experimental result17is 2. Recent experimental results point toward a similar disagreement for cobalt.12 In light normal metals and ferromagnets spin-orbit coupling generally leads to very small corrections to the single-particle energies, i.e., the band structure , but it changes the single-particle states qualitatively by in- troducing a state-dependent spin mixing. With spin- orbit coupling, the average spin of an electron can be changed in transitions due to any spin-diagonal interac- tion, in particular by electron-phonon momentum scat- tering.18–21This isalsotrue forthetwo-particleCoulomb interaction,22,23as long as one monitors only the average spin of one of the scattering particles, as is done in life- time measurements by 2-photon photoemission experi- ments. While this spin mixing due to spin-orbit cou- pling has recently been included in lifetime calculations for lead,24it was not included in DFT codes used for ex- isting lifetime calculations for 3d-ferromagnets and alu- minum,16,25,26whose results are nowadays widely used. This paper presents results for electron lifetimes in metals and spin-dependent lifetimes in ferromagnets in- cluding spin-orbit coupling . We show that spin-orbit cou- pling can be important for electron lifetimes in metals in general. Moreover, the ratio between the calculated majority and minority lifetimes is, for the first time, in agreement with experiment.11,12,17We believe that our calculated electronic lifetimes should be used as an accu- rate input for calculationsof spin-dependent hot-electron dynamics in ferromagnets. II. SPIN-DEPENDENT ELECTRON AND HOLE LIFETIMES IN CO AND NI We first discuss briefly our theoretical approach to cal- culate the lifetimes. We start from the dynamical and wave-vector dependent dielectric function ε(/vector q,ω) in the random phase approximation (RPA).13,14,25,26Our ap- proach, cf. Ref. 27, evaluates the wave-vector summa-2 tions inε(/vector q,ω) without introducing an additional broad- ening of the energy-conserving δfunction. This proce- dure removes a parameter whose influence on the calcu- lation for small qis not easy to control and which would otherwise need to be separately tested over the whole energy range. The/vectork- and band-resolved electronic scattering rates, i.e., the inverse lifetimes, γν /vectork= (τν /vectork)−1, are calculated using the expression25,26 γν /vectork=2 ¯h/summationdisplay µ/vector q∆q3 (2π)3Vq/vextendsingle/vextendsingleBµν /vectork/vector q/vextendsingle/vextendsingle2fµ /vectork+/vector qℑε(/vector q,∆E) |ε(/vector q,∆E)|2.(1) Here, the band indices aredenoted by µandν, and/vectorkand /vector qdenote wave-vectors in the first Brillouin zone (1. BZ). The energies ǫµ /vectork, occupation numbers fµ /vectorkand overlap ma- trix elements Bµν /vectork/vector q=/angbracketleftψµ /vectork+/vector q|ei/vector q·/vector r|ψν /vectork/angbracketrightare extracted from the ELK DFT (density functional theory) code,28which employs a full-potential linearized augmented plane wave (FP-LAPW) basis. Last, Vq=e2/(ε0q2) denotes the FouriertransformedCoulombpotentialand∆ E=ǫµ /vectork+/vector q− ǫν /vectorkis the energy difference between initial and final state. For negative ∆ E, the distribution function has to be re- placed by −(1−fµ /vectork+/vector q). By using the overlap matrix el- ements as defined above we neglect corrections due to local field effects. In the language of many-body Green functions, this corresponds to an on-shell G0W0calcula- tion,16,26where the screened Coulomb interaction ( W0) is obtained from the full RPA dielectric function. The /vectork- and band-dependent wave-functions that result from the DFT calculations including spin-orbit coupling are of the form |ψµ /vectork/angbracketright=aµ /vectork|↑/angbracketright+bµ /vectork|↓/angbracketright,18where|σ/angbracketrightare spinors identified by the spin projection σ=↑,↓along the mag- netization direction. According to whether |aµ /vectork|2or|bµ /vectork|2 is larger, we relabel each eigenstate by its dominant spin contribution σ, so that we obtain spin-dependent life- times,τσ /vectork. Our choice of quantization axis is such that σ=↑denotes majority carriers states and σ=↓minor- ity carrier states. Due to the existence of several bands (partially with different symmetries) in the energy range of interest and the anisotropy of the DFT bands ǫ(ν)(/vectork), several lifetimes τν /vectorkcan be associated with the same spin and energy. When we plot these spin and energy depen- dent lifetimes τσ(E) in the following, in particularFigs. 1 and2, this leads to a scatter of τσ(E) values. Figures1and2displaythecalculatedenergy-andspin- resolved carrier lifetimes τσ(E) around the Fermi energy for cobalt and nickel. The spread of lifetimes at the same energy, which was mentioned above, can serve as an indi- cation for the possible range of results for measurements of energy resolved lifetimes. These “raw data” are im- portant for the interpretation of the theoretical results because they already show two important points. First, we checked that there is no good Fermi-liquid type fit to these lifetimes. Second, even if one fits the lifetimes in a restricted energy range by a smooth τ(E) curve, this−2−1 012051015202530τCo(fs) E−EF(eV) FIG. 1. (Color online) Energy-resolved majority ( τ↑, blue +) and minority ( τ↓, red◦) carrier lifetimes for cobalt. There are in general several different lifetime points at the same ener gy (see text). We used 173/vectork-points in the full BZ. −2−1 012020406080 E−EF(eV)τNi(fs) FIG. 2. (Color online) Same as Fig. 1for nickel. ignoresthe spread of lifetimes, which can be quite sizable as shown in Figs. 1and2. We believe that such a spread of electronic lifetimes, in particular in the range around 1eV above the Fermi energy should be important for the interpretation of photoemission experiments in this en- ergy range, and when these results are used as input in hot-electron transport calculations. Figure1shows the energy- and spin-resolved lifetimes in cobalt. In addition to the longer lifetimes close to the Fermi energy, hole lifetimes in excess of 5fs occur at the top of some d-bands around −1.5,−1.2, and−1eV. For electronic states with energies above 0.5eV longer life- timesoccuratsome /vectork-points. Therearealso kstateswith apronouncedspin-asymmetryinthelifetimes(seediscus- sion below). Another important property of cobalt is the existence of two different conduction bands, which inter- sect the Fermi surface with different slope. This leads to3 tworather well-defined lifetime curves, both for electrons and holes. This can be best seen between −0.6 and 0eV, where the two curves are shifted by about 0.2eV. The calculated lifetimes in nickel, see Fig. 2, do not show a pronounced influence of d-bands and/or anisotropy below the Fermi energy as in cobalt, which is due to the smaller number of bands in the vicinity of the Fermi energy. However, there is a clear spin-dependence of electronic lifetimes, which is most pronounced around 0.4eV, but persists almost up to 2eV. III. SPIN ASYMMETRY OF ELECTRON LIFETIMES IN FE, CO, AND NI In the following, we will mainly be concerned with life- times above 0.3eV above the Fermi energy, which is the interesting energy range for the interpretation of photoe- mission experiments and hot-electron transport calcula- tions, because close to the Fermi energy the influence of phonons is expected to become more pronounced and leadtosignificantlyshorterlifetimes thanthosepredicted by a calculation that includes only the Coulomb interac- tion. To facilitate comparison with experiment we aver- age the lifetimes in each spin channel in bins of 100meV and denote the result by ¯ τ(E). The standard deviation of the averaging process then yields “error bars” on the ¯τ(E) values. Note that this procedure does not corre- spond to a “random k” approximation. Figure3displays the averagedelectron lifetimes deter- mined from the data shown in Figs. 1and2. As insets we have included the ratio of majority and minority life- times,τ↑/τ↓, together with experimental data11,12,17for iron, cobalt and nickel. Figure 3(a) shows that there is only a veryweakspin dependence for iron, and the agree- ment of the ratio τ↑/τ↓with experiment17and recent in- vestigations15,16,29is quite good, but there is a slight dis- agreement with earlier, semiempirical studies.30,31How- ever, even an increase of the ratio around 0 .5eV in the experiment17is well reproduced in our results. The averaged lifetimes of cobalt, which are shown in Fig.3(b), agree quite well with the experimental life- times,11,12but the large error bars extend to a much wider energy range than in iron. This can be traced back to the scatterof lifetimes in Fig. 1. The correspond- ing figure for iron (not shown) exhibits a much smaller scatter. The ratio of majority and minority electron life- times, see inset in Fig. 3(b), is around 1 below 0.5eV and increases to τ↑/τ↓≃2 for larger energies, a trend that agrees extremely well with measurements.11,12,17To put this result into perspective we note that the experi- mental data in Ref. 17were compared with a theoretical model based on the random kapproximation.32If the random-kinteraction matrix elements are taken to be spin and energy independent, the majority and minor- ity relaxation times are determined by double convolu- tions over the spin-dependent density-of-states (DOS).16 It was found that the experimental results were not in050100150200¯τFe(fs) 01020¯τCo(fs) 00.511.522.530204060 E−EF(eV)¯τNi(fs) 0.30.50.70.91.10.511.52 E−EF(eV)τ↑ Fe/τ↓ Fe 0.30.50.70.91.1123 E−EF(eV)τ↑ Co/τ↓ Co 0.30.50.70.91.1123 E−EF(eV)τ↑ Ni/τ↓ Ni(c)(a) (b) FIG. 3. (Color online) Energetically averaged majority (bl ue up triangles) and minority (red down triangles) lifetimes f or (a) Fe, (b) Co and (c) Ni. The error bars denote the standard deviation obtained from the scatter of the lifetimes as show n in Figs.1and2. The insets show the calculated ratio of ma- jority and minority electrons (“ ◦”),τ↑/τ↓, in comparison to experimental data, where the “ •” (“×”) correspond to values extracted from Ref. 11and17(12). agreement with the ratio of the DOS at the Fermi en- ergy, which led the authors of Ref. 17to speculate that the matrix elements for parallel and antiparallel spins should be different due to the Pauli exclusion principle. In our calculations, the effective spin-dependence of the matrix elements is caused exclusively by the spin-mixing due to spin-orbit coupling, but the effect is the same: It makes the ratio of the lifetimes different from the spin- dependent DOS at the Fermi energy. In Fig.3(c) we turn to nickel. Here, as in the case of iron, the average lifetimes are slightly larger than the measured ones17(not shown), but due to the small anisotropy in the band structure, the lifetimes in nickel show the smallest error bars and thus an extremely well- defined spin dependence. Only our calculated majority electron lifetimes are similar to earlier ab-initio evalua- tions,15,16,29but there is an important discrepancy in the ratioτ↑/τ↓: The inset of Fig. 3(c) shows a ratio of about τ↑/τ↓≃2, which is independent of energy above 0.4eV.4 012345020406080100 E−EF(eV)τAl(fs) FIG. 4. (Color online) Calculated energy-resolved electro nic lifetimes for aluminum (blue “ ◦”) in comparison to earlier in- vestigations without spin-orbit interaction. The black sq uares (stars) correspond to some data extracted from Ref. 25(33). There are in general several different lifetime points at the same energy (see text). We used 173/vectork-points in the full BZ. This results compares extremely well with experiment, and should be contrasted with the calculated result of Ref.16forτ↑/τ↓≃8 around 0.5eV. These GW calcu- lations (even with a T-matrix approach) gave very simi- lar results to those of the random kapproximation16in the energy range above 0.5eV. This indicates that the resulting spin asymmetry τ↑/τ↓≃6–8 is solely deter- mined by the spin-dependent DOS Dσ(E). Indeed, one hasD↑(EF)/D↓(EF)≃8. With the inclusion of spin- orbit coupling, which gives rise to effective spin-flip tran- sitions, the spin asymmetry is no longer determined by the spin-dependent DOS alone. This interpretation is again supported by Ref. 17where a strongly enhanced spin-flip matrix element had to be introduced by hand to improve the agreement between a random- kcalculation and experiment. To conclude the discussion of the ferromagnets, we comment on the spin-integrated lifetimes which can be obtained from the spin-dependent lifetimes, but are not shown here. Compared with experimental lifetimes of Ref.17we generally find an agreement for energies above 0.5eV that is on par with earlier calculations.15–17,29For energies below 0.5eV where the error bars on the av- eraged lifetimes are largest, the calculated lifetimes are larger than the measured ones, but in this energy range a good agreement with experiments cannot be expected because of scattering processes, which appear as elastic due to the energy resolution of the photoemission exper- iments.IV. INFLUENCE OF SPIN-ORBIT COUPLING ON ELECTRON LIFETIMES IN AL To underscore the importance of spin-orbit coupling for lifetime calculations, we also briefly discuss our calcu- lated results for electronic lifetimes in aluminum in com- parison with earlier investigations25,33without spin-orbit effects. Fig. 4shows that smaller electronic lifetimes re- sult for aluminum when spin-orbit coupling is included. In particular in the energy range between 1 and 3eV the lifetimes differ by almost a factor of two. Thus the inclusion of spin-orbit coupling improves the agreement withexperiment(see,forinstance, Ref. 34), whichwasal- ready quite good for the existing calculations.33It is con- ceivable that the use of more sophisticated many-body techniques, such as the inclusion of vertex corrections or using a T-matrix approach,33might lead to further improvements. As in the case of the ferromagnets, the electronic band structure is practically unchanged by the inclusion of spin-orbit coupling, but the rather large ef- fect of the spin-orbit coupling on spin relaxation in alu- minum through spin hot-spots has already been demon- strated.18Another argument for the importance of the spin-orbit coupling is that the spin-mixing allows transi- tion between the Kramersdegeneratebands. These tran- sitionsbetweenKramersdegeneratebandsmayhaveare- markable influence even on electron-gas properties that are usually assumed to be spin-independent, such as the intraband plasma frequency.27 V. CONCLUSION In conclusion, we presented ab-initio results for spin- dependent electronic lifetimes in ferromagnets and alu- minum including spin-orbit coupling. We found that the electronic lifetimes in iron exhibit no visible spin de- pendence in the range of −2 up to 3eV in agreement with earlier results, whereas the ratio τ↑/τ↓between majority and minority lifetimes does not exceed 2 for cobalt and nickel. Our results agree well with experi- mental data, but differ from earlier calculations, which found thatτ↑/τ↓was essentially determined by the spin- dependent density-of-states. We showed that, by allow- ing for effectively spin-changing transitions as contribu- tions to the lifetime, spin-orbit coupling is the essential ingredient that can make the spin asymmetry of the elec- troniclifetimes much smallerthan the spin-asymmetryof the density-of-states. Inclusion of our calculated spin de- pendentlifetimesintransportcalculationsshouldmakeit possible to more accurately characterize the influence of spin-dependent hot-electron transport on magnetization dynamics. ACKNOWLEDGMENTS We are grateful for a CPU-time grant from the J¨ ulich Supercomputer Centre (JSC). We acknowledge helpful discussions with M. Aeschlimann, M. Cinchetti, and S. Mathias.5 ∗hcsch@physik.uni-kl.de 1B. Koopmans, G. Malinowski, F. Dalla Longa, D. Steiauf, M. F¨ ahnle, T. Roth, M. Cinchetti, and M. Aeschlimann, Nature Materials 9, 259 (2010). 2G. M. M¨ uller, J. Walowski, Marija Djordjevic1, G.-X. Miao, A. Gupta, A. V. Ramos, K. Gehrke, V. Moshnyaga, K. Samwer, J. Schmalhorst, A. Thomas, A. H¨ utten, G. Reiss, J. S. Moodera, and M. M¨ unzenberg, Nature Mate- rials8, 56 (2009). 3B. Pfau, S. Schaffert, L. M¨ uller, C. Gutt, A. Al-Shemmary, F. B¨ uttner, R. Delaunay, S. D¨ usterer, S. Flewett, R. Fr¨ omter, J. Geilhufe, E. Guehrs, C.M. G¨ unther, R. Hawal- dar, M. Hille, N. Jaouen, A. Kobs, K. Li, J. Mohanty, H. Redlin, W.F. Schlotter, D. Stickler, R. Treusch, B. Vo- dungbo, M. Kl¨ aui, H.P. Oepen, J. L¨ uning, G. Gr¨ ubel, and S. Eisebitt, Nature Comm. 3, 1100 (2012). 4D. Rudolf, C. La-O-Vorakiat, M. Battiato, R. Adam, J. M. Shaw, E. Turgut, P. Maldonado, S. Mathias, P. Grych- tol, H. T. Nembach, T. J. Silva, M. Aeschlimann, H. C. Kapteyn, M. M. Murnane, C. M. Schneider, and P. M. Oppeneer, Nature Comm. 3, 1037 (2012). 5G. P. Zhang, W. H¨ ubner, G. Lefkidis, Y. Bai and T. F. George, Nature Physics 5, 499 (2009). 6A. Eschenlohr, M. Battiato, P. Maldonado, N. Pontius, T. Kachel, K. Holldack, R. Mitzner, A. F¨ ohlisch, P. M. Oppeneer, and C. Stamm, Nature Mat. 12, 332 (2013). 7A. Melnikov, I. Razdolski, T. O. Wehling, E. T. Papaioan- nou, V. Roddatis, P. Fumagalli, O. Aktsipetrov, A. I. Lichtenstein, and U. Bovensiepen, Phys. Rev. Lett. 107, 076601 (2011). 8T. Kampfrath, M. Battiato, P. Maldonado, G. Eilers, J. N¨ otzold, S. M¨ ahrlein, V. Zbarsky, F. Freimuth, Y. Mokrousov, S. Bl¨ ugel, M. Wolf, I. Radu, P.M. Oppeneer, and M. M¨ unzenberg, Nature Nano. 8, 256 (2013). 9M. Battiato, K. Carva, and P. M. Oppeneer, Phys. Rev. Lett.105, 027203 (2010). 10M. Battiato, K. Carva, and P. M. Oppeneer, Phys. Rev. B86, 024404 (2012). 11M. Aeschlimann, M. Bauer, S. Pawlik, W. Weber, R.Burg- ermeister, D. Oberli, and H. C. Siegmann, Phys. Rev. Lett.79, 5158 (1997). 12A. Goris, K. M. D¨ obrich, I. Panzer, A. B. Schmidt, M. Do- nath, and M. Weinelt, Phys. Rev. Lett. 107, 026601 (2011). 13G. Giuliani and G. Vignale, Quantum Theory of the Elec- tron Liquid , Cambridge University Press (2005).14G. Onida, L. Reining, and A. Rubio, Rev. Mod. Phys. 74 601 (2002). 15V. P. Zhukov, E. V. Chulkov, and P. M. Echenique, Phys. Rev. Lett. 93, 096401 (2004). 16V. P. Zhukov, E. V. Chulkov, and P. M. Echenique, Phys. Rev. B73, 125105 (2006). 17R. Knorren, K. H. Bennemann, R. Burgermeister, and M. Aeschlimann, Phys. Rev. B 61, 9427 (2000). 18J. Fabian and S. Das Sarma, Phys. Rev. Lett. 81, 5624 (1998). 19S. Essert and H. C. Schneider, Phys. Rev. B 84, 224405 (2011). 20K. Carva, M. Battiato, and P. M. Oppeneer, Phys. Rev. Lett.107, 207201 (2011). 21D. Steiauf and M. F¨ ahnle, Phys. Rev. B 79, 140401(R) (2009). 22B. Y. M¨ uller, A. Baral, S. Vollmar, M. Cinchetti, M. Aeschlimann, H. C. Schneider, and B. Rethfeld Phys. Rev. Lett.111, 167204 (2013). 23M. Krauß, T. Roth, S. Alebrand, D. Steil., M. Cinchetti, M. Aeschlimann, and H. C. Schneider, Phys. Rev. B 80, 180407(R) (2009). 24X. Zubizarreta, V. M. Silkin, and E. V. Chulkov, Phys. Rev. B84, 115144 (2011). 25F. Ladst¨ adter, U. Hohenester, P. Puschnig, and C. Ambrosch-Draxl, Phys. Rev. B 70, 235125 (2004). 26V. P. Zhukov, F. Aryasetiawan, E. V. Chulkov, I. G. de Gurtubay, and P. M. Echenique, Phys. Rev. B 64, 195122 (2001). 27S. Kaltenborn and H. C. Schneider, Phys. Rev. B 88, 045124 (2013). 28The Elk FP-LAPW Code, http://elk.sourceforge.net, ver- sion 1.4.18. 29V. P. Zhukov and E. V. Chulkov, Physics Uspekhi 52, 105- 136 (2009). 30D. R. Penn, S. P. Apell, and S. M. Girvin, Phys. Rev. Lett.55, 518 (1985). 31D. R.Penn, S.P. Apell, and S.M. Girvin, Phys.Rev.B 32, 7753 (1985). 32W. F. Krolikowski and W. E. Spicer, Phys. Rev. 185, 882 (1969). 33V. P. Zhukov, E. V. Chulkov, and P. M. Echenique, Phys. Rev. B72, 155109 (2005). 34M. Bauer, S. Pawlik, and M. Aeschlimann, Proc. SPIE3272, 201 (1998).
0710.2866v1.Intersubband_spin_orbit_coupling_and_spin_splitting_in_symmetric_quantum_wells.pdf
arXiv:0710.2866v1 [cond-mat.mtrl-sci] 15 Oct 2007Intersubband spin-orbit coupling and spin splitting in sym metric quantum wells F. V. Kyrychenko and C. A. Ullrich Department of Physics and Astronomy, University of Missour i, Columbia, Missouri, 65211, USA I. D’Amico Department of Physics, University of York, York YO10 5DD, Un ited Kingdom In semiconductors with inversion asymmetry, spin-orbit co upling gives rise to the well-known Dresselhaus and Rashba effects. If one considers quantum wel ls with two or more conduction sub- bands, an additional, intersubband-induced spin-orbit te rm appears whose strength is comparable to the Rashba coupling, and which remains finite for symmetri c structures. We show that the con- duction band spin splitting due to this intersubband spin-o rbit coupling term is negligible for typical III-V quantum wells. PACS numbers: 73.50.-h, 73.40.-c, 73.20.Mf, 73.21.-b Keywords: spintronics, spin Coulomb drag, spin-orbit coup ling, quantum wells I. INTRODUCTION Research in nanoscience is crucial for its technologi- cal implications and for the fundamental exploration of the quantum properties of nanostructures such as quan- tum wells, wires and dots. Of particular interest is the study of spin dynamics, which hopes to revolutionizetra- ditional electronics using the spin properties of the carri- ers (spintronics) [1]. In this context, the theoretical pre- diction [2] and experimental confirmation [3] of the spin- Coulomb drag (SCD) effect was of great importance, as this effect results in the natural decay ofspin current and intrinsic dissipation in AC-spintronic circuits [4]. Due to Coulomb interactions between spin-up and spin-down electrons, theupanddowncomponentsofthetotallinear momentum are not separately conserved. This momen- tum exchange between the two populations represents an intrinsic source of friction for spin currents, known as spin-transresistivity [5]. In [4] we demonstrated that the SCD produces an in- trinsic linewidth in spin-dependent optical excitations, which can be as big as a fraction of a meV for intersub- band spin plasmons in parabolic semiconductor quan- tum wells (QWs). This intrinsic linewidth would be ideal to experimentally verify the behavior of the spin- transresistivity in the frequency domain. In our proposed experiment, we suggested to use sym- metricparabolic QWs to avoid an undesired splitting of the spin plasmons due to Rashba spin-orbit (SO) cou- pling. We based our discussion on earlier work [6], in which collective intersubband spin excitations in QWs weredescribedinthepresenceofDresselhausandRashba SO interaction terms [7, 8] for strictly two-dimensional (2D) systems [9]. In symmetric QWs, the Rashba term vanishes and only bulk inversion asymmetry (Dressel- haus) interaction is present. However, as shown recently by Bernardes et al. [10], the Rashba SO coupling gives finite contributions even for symmetric structures, if treated in higher order per- turbation theory. As a consequence, for QWs with morethan one subband, there appears an additional intersub- band SO interaction, whose magnitude can become com- parable to that of 2D Dresselhaus and Rashba interac- tions. This interaction gives rise to a nonzero spin-Hall conductivity and renormalizes the bulk mass by ∼5% in InSb double QWs [10]. This raises the question whether this effect must be accounted for when extracting the SCD from intersubband spin plasmon linewidths [4]. In this paper we are going to show that while intersub- band SO interaction may manifest itself in some special cases, as for example in the double well analyzed in Ref. [10], it has little to no effect on spin splitting and spin mixing in QWs once the 2D Dresselhaus and/or Rashba terms are taken into account. In Sec. II we present the general formalism of calcu- lating conduction band states in quantum structures in- cluding both 2D and intersubband SO interaction. In Sec. III we consider the specific case of symmetric single- well quantum structures, and in Sec. IV we present re- sults for a parabolic model QW. Sec. V gives a brief summary. II. GENERAL FORMALISM We consider conduction electrons in a QW described by the Hamiltonian ˆH=ˆH0+ˆHso, (1) whereˆH0is spin independent and ˆHsois the SO inter- action projected on the conduction band. For simplic- ity we will consider only spin off-diagonal (spin-mixing) terms in ˆHso. The eigenfunctions associated with ˆH0 alone can be obtained by solving a single-particle equa- tion of the Schr¨ odinger-Poisson or Kohn-Sham type, re- sulting in spin-independent subband envelope functions ψi(z,k/bardbl) and energy eigenvalues εi, whereiis the sub- band index and zis the direction of quantum confine- ment.2 Let us now consider the two lowest conduction sub- bands of the QW. In the basis of the first two subband spinors|ψ1↑/angbracketright,|ψ1↓/angbracketright,|ψ2↑/angbracketright,|ψ2↓/angbracketright, the Schr¨ odinger equation with the full Hamiltonian (1) has the form ε1α10β α∗ 1ε1β′0 0β′∗ε2α2 β∗0α∗ 2ε2 A=εA, (2) where α1=/angbracketleftψ1↑ |ˆHso|ψ1↓/angbracketright α2=/angbracketleftψ2↑ |ˆHso|ψ2↓/angbracketright, β=/angbracketleftψ1↑ |ˆHso|ψ2↓/angbracketright, β′=/angbracketleftψ1↓ |ˆHso|ψ2↑/angbracketright. (3) To remove the off-diagonal terms mixing the ↑,↓states within the same subband, we apply the unitary transfor- mationB=U·Awith U=1√ 2 1−α1 |α1|0 0 1α1 |α1|0 0 0 0 1 −α2 |α2| 0 0 1α2 |α2| . (4) Equation (2) then transforms into ε1−|α1|0 −γ1γ2 0ε1+|α1| −γ2γ1 −γ∗ 1−γ∗ 2ε2−|α2|0 γ∗ 2γ∗ 1 0ε2+|α2| B=εB, (5) where the off-diagonal matrix elements γ1,2=1 2/bracketleftig βα∗ 2 |α2|±β′α1 |α1|/bracketrightig (6) connect the first and second subbands. We treat these contributions to the conduction band Hamiltonian per- turbatively to second order, and obtain the following so- lutions of Eq. (5): ε± 1=ε1±|α1| +|γ1|2 ε1±|α1|−ε2∓|α2|+|γ2|2 ε1±|α1|−ε2±|α2|, ε± 2=ε2±|α2| +|γ1|2 ε2±|α2|−ε1∓|α1|+|γ2|2 ε2±|α2|−ε1±|α1| and B− 1= 1 0 −γ∗ 1 ε1−|α1|−ε2+|α2| γ∗ 2 ε1−|α1|−ε2−|α2| , (7)B+ 1= 0 1 −γ∗ 2 ε1+|α1|−ε2+|α2| γ∗ 1 ε1+|α1|−ε2−|α2| , (8) B− 2= −γ1 ε2−|α2|−ε1+|α1| −γ2 ε2−|α2|−ε1−|α1| 1 0 , (9) B+ 2= γ2 ε2+|α2|−ε1+|α1|γ1 ε2+|α2|−ε1−|α1| 0 1 .(10) The eigenvectors B± iare normalized up to first order in the off-diagonal perturbation. In the absence of intrasubband (2D) terms, α1=α2= 0, theintersubband SO interactiongives rise to spin mix- ing without lifting the spin degeneracy( ε+ i=ε− i); it only causes a spin-independent shift of the subband energies. By contrast, if an intrasubband interaction is present (or ifspindegeneracyislifted byothermeans, e.g.,byamag- netic field), the spin splitting is affected. For the lowest subband it is given by ε+ 1−ε− 1= ∆ε1, where ∆ε1= 2|α1|+2|γ1|2 |α2|−|α1| (ε2−ε1)2−(|α2|−|α1|)2 −2|γ2|2 |α2|+|α1| (ε2−ε1)2−(|α2|+|α1|)2.(11) To proceed further we need the explicit form of the SO Hamiltonian ˆHso. III. RASHBA AND DRESSELHAUS SO INTERACTION IN SYMMETRIC QWS By folding down the 14 ×14k·pHamiltonian for a QW grown in [001] direction in a zinc-blende crystal to a 2×2 conduction band problem [11], one obtains an effective SO Hamiltonian in the conduction band: ˆHso≈/parenleftbigg 0hso h∗ so0/parenrightbigg , (12) where hso=R(z)k−−iλk+∂2 ∂z2−iλ 4(k2 −−k2 +)k−,(13) with λ= 4√ 2 3PQP′/parenleftbigg1 (E∆−ε)(E′v−ε)−1 (Ev−ε)(E′ ∆−ε)/parenrightbigg3 and R(z) =√ 2 3P2/bracketleftbigg∂ ∂z/parenleftbigg1 Ev−ε−1 E∆−ε/parenrightbigg/bracketrightbigg +√ 2 3P′2/bracketleftbigg∂ ∂z/parenleftbigg1 E′v−ε−1 E′ ∆−ε/parenrightbigg/bracketrightbigg .(14) Here,k±=1√ 2(kx±iky),εis the electron energy, Ev(z) andE∆(z) are the position-dependent Γ 8and Γ7valence bandedges, and P=−i/planckover2pi1 m/angbracketleftS|ˆpx|X/angbracketright=/radicalig Ep/planckover2pi12 2misthe mo- mentum matrix element. Primed quantities correspond to the higher lying Γ 8−Γ7conduction band and Qis the momentum matrix element between the valenceband and the higher conduction band. Along with the Rashba and linear Dresselhaus terms in Eq. (12) we keep the cu- bic Dresselhaus term as well. During the derivation we assumed that the variation of the band edges is small compared with the energy gaps in the material. In symmetric structures, due to parity conservation theintrasubband SO interaction contains only the Dres- selhaus contribution, α1=−λ 4√ 2k3sin(2ϕ)e−iϕ+D11√ 2kei(ϕ+π 2),(15) α2=−λ 4√ 2k3sin(2ϕ)e−iϕ+D22√ 2kei(ϕ+π 2),(16) and theintersubband SO interaction(between the lowest two subbands) involves only the Rashba term β=β′∗=R12√ 2ke−iϕ, (17) whereϕis the polar angle of the in-plane vector k/bardblmea- sured from the [100] direction, and k=|k/bardbl|. Further- more, Dii=−λ/angbracketleftbigg ψi(z)/vextendsingle/vextendsingle/vextendsingle/vextendsingle∂2 ∂z2/vextendsingle/vextendsingle/vextendsingle/vextendsingleψi(z)/angbracketrightbigg (18) and R12=/angbracketleftψ1(z)|R(z)|ψ2(z)/angbracketright. (19) The quantity R12corresponds to the coupling parameter ηderived in Ref. [10] using an 8-band k·pmodel. For smallkthe linear term in Eqs. (15)-(16) dominates and we can approximate α1 |α1|≈α2 |α2|≈ei(ϕ+π 2). (20) Then, γ1=1√ 2R12kcos/parenleftig 2ϕ+π 2/parenrightig (21) γ2=−i√ 2R12ksin/parenleftig 2ϕ+π 2/parenrightig , (22)and the ground state spin splitting follows from Eq. (11) as ∆ε1≈2|α1|−R2 12D11√ 2(ε2−ε1)2k3−R2 12D22√ 2(ε2−ε1)2k3cos(4ϕ). (23) Theintersubband interaction results thus in an addi- tional spin splitting proportional to k3. Next, we expand the spin splitting that is induced by theintrasubband SO interaction. Up to order k3we ob- tain |α1| ≈D11√ 2k+λ 8√ 2k3−λ 8√ 2k3cos(4ϕ),(24) which givesthe final expressionforthe subband splitting: ∆ε1=√ 2D11k+/parenleftbiggλ 4−R2 12D11 (ε2−ε1)2/parenrightbiggk3 √ 2 −/parenleftbiggλ 4+R2 12D22 (ε2−ε1)2/parenrightbiggk3 √ 2cos(4ϕ).(25) One finds that the intersubband SO interaction produces an additional spin splitting of the same symmetry as the intrasubband cubic Dresselhaus term. We will now es- timate the magnitude of this additional contribution for GaAs parabolic QWs. IV. SUBBAND SPIN SPLITTING IN PARABOLIC WELLS Let us consider a parabolic QW with conduction band confining potential V(z) =1 2Kz2, (26) resulting in the noninteracting energy spectrum εj=/radicalbigg /planckover2pi12K m∗/parenleftbigg j−1 2/parenrightbigg , j = 1,2,...(27) The first and second subband envelope functions are ψ1(z) =4/radicalbigg 2ξ πe−ξz2, (28) ψ2(z) =4/radicalbigg 32ξ3 πze−ξz2, (29) whereξ=/radicalbig m∗K/4/planckover2pi12. Straightforward calculations give /angbracketleftbigg ψ1/vextendsingle/vextendsingle/vextendsingle/vextendsingle∂2 ∂z2/vextendsingle/vextendsingle/vextendsingle/vextendsingleψ1/angbracketrightbigg =−ξ, (30) /angbracketleftbigg ψ2/vextendsingle/vextendsingle/vextendsingle/vextendsingle∂2 ∂z2/vextendsingle/vextendsingle/vextendsingle/vextendsingleψ2/angbracketrightbigg =−3ξ, (31) /angbracketleftψ1|z|ψ2/angbracketright=−1 2√ξ. (32)4 For our parabolic well, the positional dependence of the valence band edge (the valence band potential) is Ev=−1 4Kz2, corresponding to a valence band offset VBO=0.33. For GaAs parameters ( Eg= 1.42 eV, ∆ = 0 .34 eV,Ep= 22 eV) Eq. (14) gives R(z)≈ −/parenleftbig∂Ev ∂z/parenrightbig 7˚A2. Using Eqs. (18), (19) and (30)–(32) we then get R12=−(7˚A2)K 4√ξ, D 11=λξ, D 22= 3λξ, and R2 12D22 (ε2−ε1)2= ∆ε /planckover2pi12 2m∗˚A2 2 147 64λ∼10−6λ, form∗= 0.065m0and ∆ε=ε2−ε1= 40 meV. The con- tribution of the intersubband SO interaction to the spin splitting of the lowest conduction subband is six orders of magnitude weaker than that of the cubic Dresselhaus intrasubband termand thus canbe completely neglected. The spin mixing induced by the intersubband SO in- teraction can be estimated from Eq. (7): /vextendsingle/vextendsingle/vextendsingle/vextendsingleγ2 ε2−ε1/vextendsingle/vextendsingle/vextendsingle/vextendsingle2 ≈R2 12k2 2(∆ε)2=49 32 ∆ε /planckover2pi12 2m∗˚A2 k2˚A2∼10−7, fork= 0.01˚A−1. This is seven orders of magnitude weaker than the spin mixing induced by intrasubband SO interaction and also can be completely neglected. SimilarresultswereobtainedforGaAssymmetricrect- angular QWs.V. CONCLUSIONS In this paper, we have considered the effects of SO coupling on the conduction subband states in symmetric QWs. Our work was motivated by Ref. [10], which dis- cussed a SO coupling effect specific to QWs with more than one subband and showed that it can affect the elec- tronic and spin transport properties in some systems. We found that although the magnitude of this in- tersubband SO interaction can be comparable to that of the 2D Dresselhaus and Rashba terms, its effect on the spin splitting and spin mixing of conduction band states is several orders of magnitude weaker since it con- nects states with different energies. This is due to the fact that the spin splitting and spin mixing of conduc- tion band states are renormalized by the intersubband energy difference. Therefore, ifone considerssystem with non-degenerate subbands, one can completely neglect the intersubband SO interaction compared to the usual 2D Dresselhaus and Rashba terms. These findings provide an a posteri- orijustification for the approach used to calculate sub- band splittings and spin plasmon dispersions carried out in Ref. [6]. This opens the way for a comprehensive the- ory of collective intersubband excitations in QWs in the presence of SCD and SO coupling. Acknowledgments This work was supported by DOE Grant No. DE-FG02-05ER46213, the Nuffield Foundation Grant NAL/01070/G, and by the Research Fund 10024601 of the Department of Physics of the University of York. [1]Semiconductor spintronics and quantum computation , edited by D. D. Awschalom, N. Samarth, and D. Loss (Springer, Berlin, 2002) [2] I.D’AmicoandG.Vignale, Phys.Rev.B 62, 4853(2000). [3] C. P. Weber, N. Gedik, J. E. Moore, J. Orenstein, J. Stephens, and D. D. Awschalom, Nature (London) 437, 1330 (2005). [4] I. D’Amico and C. A. Ullrich, Phys. Rev. B 74, 121303(R) (2006); I. D’Amico and C. A. Ullrich, Physica Status Solidi (b) 243, 2285 (2006); I. D’Amico and C. A. Ullrich, Journal of Magnetism and Magnetic Materials 316, 484 (2007) [5] I. D’Amico I and G. Vignale, Europhys. Lett. 55, 566 (2001); I. D’Amico and G. Vignale, Phys. Rev. B 65, 085109 (2002); I. D’Amico and G. Vignale, Phys. Rev. B 68, 045307 (2003); K. Flensberg, T. S. Jensen, and N. A.Mortensen, Phys. Rev. B 64, 245308 (2001) [6] C. A. Ullrich and M. E. Flatte, Phys. Rev. B 66, 205305 (2002); C. A. Ullrich and M. E. Flatte, Phys. Rev. B 68, 235310 (2003) [7] G. Dresselhaus, Phys. Rev. 100, 580 (1955) [8] Yu. L. Bychkov and E. I. Rashba, J. Phys. C 17, 6039 (1984) [9] R. Winkler, Spin-orbit coupling effects in two- dimensional electron and hole systems (Springer, Berlin, 2003) [10] E. Bernandes, J. Schliemann, M. Lee, J.C. Egues, and D. Loss, Phys. Rev. Lett 99, 076603 (2007) [11] P. Pfeffer and W. Zawadzki, Phys. Rev. B 41, 1561 (1990); P. Pfeffer, Phys. Rev. B 59, 15902 (1999); P. Pf- effer and W. Zawadzki, Phys. Rev. B 15, R14332 (1995)
1805.00047v1.Superconducting_tunneling_spectroscopy_of_spin_orbit_coupling_and_orbital_depairing_in_Nb_SrTiO__3_.pdf
Superconducting tunneling spectroscopy of spin-orbit coupling and orbital depairing in Nb:SrTiO 3 Adrian G. Swartz,1, 2, 3,Alfred K. C. Cheung,4Hyeok Yoon,1, 2, 3Zhuoyu Chen,1, 2, 3Yasuyuki Hikita,2Srinivas Raghu,2, 4and Harold Y. Hwang1, 2, 3 1Geballe Laboratory for Advanced Materials, Stanford University, Stanford, California 94305, USA 2Stanford Institute for Materials and Energy Sciences, SLAC National Accelerator Laboratory, Menlo Park, California 94025, USA 3Department of Applied Physics, Stanford University, Stanford, California 94305, USA 4Department of Physics, Stanford University, Stanford, California 94305, USA (Dated: May 2, 2018) We have examined the intrinsic spin-orbit coupling (SOC) and orbital depairing in thin lms of Nb-doped SrTiO 3by superconducting tunneling spectroscopy. The orbital depairing is geometrically suppressed in the two-dimensional limit, enabling a quantitative evaluation of the Fermi level spin- orbit scattering using Maki's theory. The response of the superconducting gap under in-plane magnetic elds demonstrates short spin-orbit scattering times so1:1 ps. Analysis of the orbital depairing indicates that the heavy electron band contributes signi cantly to pairing. These results suggest that the intrinsic spin-orbit scattering time in SrTiO 3is comparable to those associated with Rashba e ects in SrTiO 3interfacial conducting layers and can be considered signi cant in all forms of superconductivity in SrTiO 3. The relativistic spin-orbit interaction is fundamental in the solid state, connecting the conduction electron spin to the atomic, electronic, orbital, and structural symmetry properties of the material [1]. SrTiO 3is an oxide semiconductor with highly mobile t2gconduction electrons and exhibits superconductivity at the lowest known carrier density of any material [2{4]. The rele- vance of the intrinsic spin-orbit coupling (SOC) for su- perconductivity in the bulk material remains an open question: the atomic SOC produces a relatively small splitting (29 meV [2]) of the t2gbands, butmight be an important energy scale considering the small supercon- ducting gap in SrTiO 3. Moreover, SrTiO 3is the host material for unconventional two-dimensional (2D) super- conductors such as FeSe/SrTiO 3[5],-doped SrTiO 3[3], and LaAlO 3/SrTiO 3[6]. Spin-orbit coupling in SrTiO 3 interfacial accumulation layers has been extensively stud- ied both experimentally and theoretically [7{13]. In these systems, Rashba SOC has been suggested to give rise to many of the unusual normal- and superconducting-state properties due to the broken inversion symmetry and the highly asymmetric con nement potential. Understand- ing the competition between the intrinsic and Rashba coupling scales is critical to understanding the spin-orbit textures and superconducting phases in both bulk and 2D systems. The spin-orbit coupling strength can be quantitatively extracted from superconducting tunneling spectra of thin lms in large parallel magnetic elds [14{16]. In a con- ventionals-wave superconductor, a magnetic eld acts in two ways on the conduction electrons: by inducing cyclotron orbits and via the electron magnetic moment (spin). Both of these e ects lead to the breaking of Cooper pairs once their energy scale competes with the condensation energy. For thin lms in the 2D limit, theorbital depairing can be geometrically suppressed, lead- ing to highly anisotropic upper-critical elds with large in-planeHc2;k. In the absence of spin-orbit coupling, spin is a good quantum number and Hc2;kis determined by the Pauli paramagnetic limit ( HP=  0=p 2B, where 0is the superconducting gap at T= 0 andBis the Bohr magneton) [14, 17, 18]. The application of an in- plane magnetic eld splits the spin-up and spin-down superconducting quasiparticle density of states (DOS) through the Zeeman e ect (Fig. 1 left panel) [14]. In- creasing the spin-orbit coupling leads to a mixing of the spin-up and spin-down states and lifts Hc2;kabove the Pauli limit [14, 19, 20]. If the spin-orbit scattering rate is very fast (h=so>0, wheresois the normal-state spin- orbit scattering time), then the superconducting DOS does not exhibit measurable Zeeman splitting (Fig. 1 right panel). Fitting the tunneling spectra using Maki's theory [21{23] enables a quantitative extraction of both the orbital depairing parameter ( o) andsofrom the tunneling spectra. This approach, pioneered by Tedrow and Meservey, has been used extensively to explore de- pairing mechanisms of conventional elemental supercon- ductors [14{16, 22, 23]. Here we examine spin-orbit coupling and orbital de- pairing in thin lms of Nb-doped SrTiO 3(NSTO) using tunneling spectroscopy. Recently, we have developed an approach for realizing high-quality tunneling junctions for bulk NSTO with eV resolution of the superconduct- ing gap [24, 25]. By carefully engineering the band align- ments using polar tunneling barriers, the interfacial car- rier density probed by tunneling corresponds to the nom- inal density of dopants. We study the tunneling conduc- tance (di=dv ) of NSTO lms in the 2D limit ( d <  GL, wheredis the lm thickness and GLis the Ginzburg- Landau coherence length). We nd a single supercon-arXiv:1805.00047v1 [cond-mat.supr-con] 30 Apr 20182 -4-2024E/Δ2.52.01.51.00.50.0ρ↑,↓ /ρ0 -4-2024E/Δ !∥!#SrTiO3(001)Nb:SrTiO3LaAlO3AgdΩΩa) b) SrTiLaAlO(SrO)0(TiO2)0(LaO)+(AlO2)-(AlO2)-(LaO)+ %=0(=0.1+,-∆/=0.6%=6(=0.1+,-∆/=0.61↑+1↓ FIG. 1. (Color online) a) Schematic of the tunneling junc- tion device structure and atomic stacking of the oxide het- erostructure. b) Expected e ect of Zeeman splitting on the spin-dependent DOS for two cases: zero spin-orbit coupling (b= 0) (left panel) and large spin-orbit coupling ( b= 6) (right panel). The dimensionless SOC parameter b= h=(3so0) re- ects the strength of the SOC relative to the gap energy scale. Dashed blue (dashed grey) and solid red (solid grey) curves represent the spin-up and spin-down DOS, respectively, while the solid black curve gives the total DOS from "+#(shifted upwards by 1 for clarity). The spectra were calculated using Maki's theory (Eq. (2)) at T= 0 K, lifetime broadening parameter= 0:1, and magnetic eld BH=0= 0:6. ducting gap which closes at the superconducting transi- tion temperature ( Tc). We extract Hc2;kfrom the tun- neling spectra and nd that it greatly exceeds the Pauli limit. Under in-plane applied elds, Zeeman splitting is not observed and an apparent single gap persists at all elds until closing completely near 1.6 T, indicating that the spin-orbit coupling scale ( h=so) is larger than  0. We analyze the data using Maki's theory [21{23] and ex- amine the relative contributions from orbital depairing and spin-orbit scattering. Due to the heavy mixing of the spin states, Maki's theory provides an upper-bound for the spin-orbit scattering time of so1:1 ps and spin di usion length s32 nm. We fabricated tunneling junctions consisting of super- conducting NSTO thin lms of thickness d= 18 nm, with a 2 unit cell (u.c.) epitaxial LaAlO 3tunneling barrier, and Ag counter electrodes as described elsewhere [24, 25]. NSTO with 1 at.% Nb-doping was homoepitaxially de- posited on undoped SrTiO 3(001) by pulsed-laser deposi- tion [26]. Films grown by this technique exhibit full car- 6050403020100 Δ (µeV)0.60.40.20.0T (K)1.00.80.60.40.20.0R/RN0.60.50.40.30.20.10.0di/dv (mS)-400-2000200400V (µV)!"#= 0 T!"#= 0 Ta)b)FIG. 2. (Color online) Tunneling spectroscopy and resis- tivity in zero eld. a) Tunneling conductance ( di=dv ) of 18 nm thick Nb-doped SrTiO 3thin lm measured at the base temperature of the dilution refrigerator. b) Superconducting gap amplitude () (open circles, left axis) compared to the normalized resistance (solid blue (grey) line, right axis). The superconducting gap closes at T= 3155 mK, which is very close to the resistive transition temperature Tc= 330 mK de ned as 50% of the normal state resistivity at T= 0:6 K. rier activation and bulk-like electron mobility. The polar LaAlO 3tunnel barrier plays a crucial role in enabling access to the electronic structure of NSTO in the 2D su- perconducting limit. The LaAlO 3layer provides an in- terfacial electric dipole which shifts the band alignments between the Ag electrode and semiconducting SrTiO 3by 0:5 eV/u.c. [27{29]. Aligning the Fermi-level between the two electrodes signi cantly reduces the Schottky bar- rier and eliminates the long depletion length which pro- hibits direct tunneling. First, we report the zero- eld superconducting behav- ior of the sample. Figure 2a shows di=dv measured at base temperature ( T= 20 mK) and 0H= 0 T exhibit- ing a single superconducting gap (). Although we ob- serve high-energy coupling to longitudinal-optic phonon modes (not shown) as reported recently [24], we do not nd other strong-coupling renormalizations (i.e. McMil- lan and Rowell [30]) in the tunneling spectra. The su- perconducting gap is well t by the Bardeen-Cooper- Schrie er (BCS) equation for the density of states with 0= 471eV. Due to the nite resolution of the measurement and thermal broadening, the minimum of the superconducting gap is nite. Here, the gap mini- mum is two-orders of magnitude smaller than the normal state conductance, demonstrating the dominance of elas- tic tunneling and the high quality of the junction, even in the 2D limit. The superconducting gap closes near Tc= 330 mK as measured by four-point resistivity (Fig. 2b). Importantly, we do not observe a pseudogap as was recently observed in LaAlO 3/SrTiO 3[31], indicating the pseudogap is speci c to the LAO/STO interface and not a generic feature in the 2D limit. We now turn to the magnetic- eld response of the su- perconducting gap. Figure 3a shows the superconducting gap at several characteristic values of applied magnetic3 eld (left panel: H?, right panel: Hk). Figure 3b dis- plays the zero-bias conductance (gap minimum) normal- ized to the normal-state zero-bias conductance for both eld orientations. We nd a large anisotropy between Hc2;?andHc2;kwith a ratio Hc2;?/Hc2;k= 0.052. We extract the Ginzburg-Landau superconducting coherence lengthGL=p 0=(2Hc2;?) = 62 nm > d, con rming the superconducting state is in the 2D regime. SrTiO 3 is a type-II superconductor with large London penetra- tion depth compared to GLandd, and the quenching of superconductivity due to an out-of-plane eld can be attributed to the formation of vortices. For elds applied in-plane, the large size of a vortex core is energetically un- favorable to form in the 2D limit and the orbital depair- ing is dramatically suppressed leading to enhanced Hc2;k. We nd that the superconducting gap exhibits large Hc2;k far in excess of the Pauli limit ( HP=  0=p 2B= 0.574 T), and in agreement with a study of upper-critical elds from resistivity measurements in -doped SrTiO 3quan- tum wells [20]. Here, we can examine the spin-dependent response of the superconducting gap spectra to extract the relevant contributions to orbital and spin depairing mechanisms. The superconducting DOS has been given by Maki's theory, which takes into account orbital depairing, Zee- man splitting of the spin states, and SOC [15, 21]. The spin-dependent DOS is given by, ";#=0 2sgn(E)Re0 @uq u2 11 A; (1) where0is the normal-state DOS and uare de ned by, u=EBH 0+uq 1u2 +b0 @uuq 1u2 1 A;(2) for whichEis the energy relative to the Fermi level ( EF), b= h=(3so0) is a dimensionless quantity representing the strength of the spin-orbit scattering relative to  0, andrepresents spin-independent lifetime corrections. Maki's equation (Eq. (2)) reduces to the BCS DOS in the limit of vanishing andb. The quantity BHrepre- sents the Zeeman splitting of the spin-dependent states and observation of this splitting in the experimental data depends on the strength of b(see Fig. 1). The parame- ter= i+ oH2 kincludes eld-independent broadening ( i) and o=De2d2=(6h0) is the standard orbital de- pairing for a thin lm in a parallel magnetic eld ( Dis the di usion coecient) [14, 15, 21]. We follow the nu- merical approach of Worledge and Geballe in applying Eq. (2) to the tunneling data [21{23]. We now focus on the spectra shown in Fig. 3a (right panel) for in-plane applied elds. The magnetic elds ex- plored here ( BHk=0<2) are large enough to observe Zeeman splitting in the weak spin-orbit limit ( b<1) [32]. 1.21.00.80.60.40.20.0(σ / σN)|V=0 2.01.51.00.50.0µ0H (T)0.60.50.40.30.20.10.0di/dv (mS)-400-2000200400V (µV) 0.11 T 0.09 T 0.08 T 0.07 T 0.06 T 0.05 T 0.04 T 0.03 T 0.02 T 0.01 T 0 T-400-2000200400V (µV) 2 T 1.6 T 1.5 T 1.4 T 1.2 T 1.0 T 0.8 T 0.5 T 0.3 T 0 T!"#,%!"#,∥'(!∥='(!%= !)a) b)FIG. 3. (Color online) Tunneling spectroscopy of the super- conducting gap under applied magnetic eld. a) Raw di=dv data for several values of magnetic elds applied out-of plane (0H?, left panel) and in-plane ( 0Hk, right panel). b) Zero- bias conductivity ( =di=dv ) of the gap minimum normalized to the normal-state conductance ( N) for both eld orienta- tions. The out-of plane ( Hc2;?) and in-plane ( Hc2;k) upper critical elds are indicated. The vertical dashed blue line in- dicates the Pauli paramagnetic limiting eld ( HP). However, for all measured magnetic elds, the data does not exhibit a clear signature of Zeeman splitting indicat- ing strong spin scattering relative to the superconducting gap (compare Fig. 1 with Fig. 3a right panel) and con- sistent with the violation of the Pauli-limit. While the spin-orbit parameter bis eld-independent, the e ect of Zeeman splitting in combination with rapid spin mixing is to produce an e ective broadening of the total DOS ("+#, see Fig. 1) following an H2dependence [15]. Therefore since both orbital depairing and the large SOC produce quasiparticle broadening under an applied eld, it is a useful exercise to rst consider a reduced version of Maki's theory which ignores the spin-degree of freedom in the problem, such that, u!u=E 0+0up 1u2; (3) which in zero- eld is equivalent to the Dynes formulation were the phenomenological Dynes quasiparticle broaden- ing parameter is given by = 00[33]. We rst t the data of Fig. 3a right panel using Eq. (3) where 0is4 the only free parameter. The results for 0are shown in Fig. 4a as a function of H2 kand are well described by 0= i+H2with i= 0.056 and = 0.4 T2. The small intrinsic quasiparticle broadening ( i) gives = 2eV and identical to our previous report in the bulk limit [24]. The extracted value re ects the total contri- bution to eld-induced broadening from both spin-orbit coupling and orbital depairing. To quantify the spin-orbit and orbital depairing con- tributions, we apply Maki's full theory (Eq. (2)) to the set of tunneling data between 300 and 700 mT (0:3<  BH=0<0:86) including the spin-dependent density of states, spin-orbit parameter b, and depairing parameter= i+ oH2. The only free parameters are b and owhich must both be singly valued at all elds. We nd that the best ts are statistically equivalent for b>4 (with varying 0) [32], indicating short spin-orbit scat- tering times so<1:1 ps. In this regime ( b>4), 0and bare correlated. This can be understood as a competi- tion between the spin-orbit induced e ective broadening and orbital depairing. For instance, in the limit b!1 , the broadening from SOC vanishes and orbital depairing must asymptotically approach to account for the exper- imentally observed broadening. Fig. 4b shows a charac- teristic best t for 0Hk= 0.5 T (BH=0= 0.61) with b= 6 and 0= 0:11. Additionally, an upper-bound on the spin di usion length is given by s=q 3 4Dtrso<32 nm [34], where Dtr=v2 Ftr=30:0012 m2/s is the trans- port di usion coecient. Here we have estimated the Fermi velocity vFin a single-band approximation with e ective mass m= 1.24m0[3] and Fermi level EF= 61 meV [2, 24]. We have used the Drude scattering time tr=me=ewheree= 300 cm2/Vs is the experimen- tally measured electron mobility. The contribution from orbital depairing in the tunnel- ing data provides additional information on the super- conducting phase. The best ts from Maki's theory in the rangeb >4 correspond to 0.016 T2< o, for which oincreases commensurately with b. Thus, even though spin-orbit and orbital depairing cannot be quan- ti ed independently, there are clear experimental limits on o. We can compare the experimental owith the ex- pected orbital contribution from normal-state transport parameters with o=Dtre2d2=(6h0)2 T2, which is far in excess of the measured total broadening of = 0:4 T2. This apparent discrepancy can be resolved by con- sidering the multi-band nature of bulk SrTiO 3with three occupiedt2gorbitals comprised of two light- and one heavy-electron bands [2, 10]. Normal-state transport co- ecients are dominated by the highly mobile light elec- trons, but these carriers only make-up a fraction of the total DOS, whereas the lowest lying heavy band com- prises the majority of the electrons at EF[2, 3, 10, 35]. In other words, the experimental data cannot be explained by solely considering highly mobile, light electrons in 1.61.20.80.40.0σ / σN-400-2000200400V (µV)0.80.60.40.20.0ζ´1.61.20.80.40.0(µ0H||)2 (T)a)b) !"#∥= 0.5 T%=6("=0.11FIG. 4. (Color online) Maki analysis of the superconducting gap spectra under in-plane magnetic elds. a) Total quasi- particle broadening 0(black dots) determined by tting the tunneling data of Fig. 3a right panel using Eq. (3). The total broadening exhibits a dependence on the square of the applied magnetic eld and the solid line represents a t to 0= i+H2. b) Normalized di=dv data (solid black line) measured at 0Hk= 0.5 T and theoretical t (dashed red (grey) line) using Maki's full theory as expressed in Eq. (2) withb= 6, i= 0.056, o= 0.11, and  0= 47eV. forming the superconducting phase. We can transpose the orbital depairing extracted from the superconduct- ing tunneling data to DSCrepresenting the di usion co- ecient for electrons which contribute to pairing. We nd 0:1104m2/s< D SC<2:3104m2/s, which agrees very well with a simplistic estimate of the di u- sion constant for the heavy electron band with m6m0 [36] and momentum scattering time he100 fs, giving Dhe1104m2/s. Therefore, the robustness of super- conductivity at high magnetic elds is consistent with the established bulk band structure for which the heavy elec- tron band dominates the total DOS and results in weak orbital depairing. We note that the importance of the heavy bands for superconductivity has been suggested in LaAlO 3/SrTiO 3[37, 38] The spin-orbit scattering times observed here are com- parable to the momentum scattering time ( tr=so0:1) and signi cantly shorter than those suggested theoreti- cally in a single band limit [39]. Additionally, we can expect that Rashba and Dresselhaus elds are minimal in the current sample structure under investigation [32]. Therefore, the rapid spin mixing near the Fermi level can be understood in the context of the multiband elec- tronic structure of bulk SrTiO 3with hybridized orbital character arising from the tetragonal crystal eld split- ting and the intrinsic atomic spin-orbit interaction [2]. This picture is analogous to p-type Si where short spin relaxation times are characteristic despite the modest SOC [40, 41]. The spin-orbit scattering explored here re ects the electrons with the largest contribution to the density of states and the superconducting conden- sate, which in bulk SrTiO 3is the heavy electron band. This is in contrast to transport experiments exploring5 spin-orbit coupling in the normal state (i.e. weak (anti- )localization, Subnikov de Haas oscillations) which are most sensitive to the highly mobile subset of carriers [3, 35]. Therefore, careful analysis of the sub-band struc- ture and orbital character in con ned SrTiO 3-based het- erostructures (e.g. LaAlO 3/SrTiO 3) is critical to un- derstanding the spin-orbit properties of the normal and superconducting phases. Regardless, it is interesting to note that the scattering times found here ( so1 ps), are in the ballpark of the vast majority of experimen- tal ndings in LaAlO 3/SrTiO 3[7, 8], suggesting that the spin-orbit scattering at the Fermi level arising from the intrinsic atomic spin-orbit interaction contributes at least on equal footing with Rashba e ects. In conclusion, we have performed tunneling experi- ments on the dilute superconductor SrTiO 3doped with 1 at.% Nb in the 2D superconducting limit. These re- sults were enabled by precisely designing the tunneling junction with epitaxial dipole tunnel barriers, which shift band alignments and facilitates high-resolution tunneling spectroscopy. The data indicates a single superconduct- ing gap which closes at Tc. By geometrically suppressing the orbital depairing, we show that the large intrinsic SOC can be observed directly in the tunneling spectra by the violation of the Pauli-limit and the absence of Zeeman splitting. Surprisingly short spin-orbit scatter- ing times of order 1 ps were obtained. Examination of the orbital depairing parameter indicates that the heavy electron band, which is dicult to explore in transport experiments, plays an important role in the formation of the superconducting phase. We thank M. E. Flatt e for useful discussions. This work was supported by the Department of Energy, Oce of Basic Energy Sciences, Division of Mate- rials Sciences and Engineering, under Contract No. DE-AC02-76SF00515; and the Gordon and Betty Moore Foundation's EPiQS Initiative through Grant GBMF4415 (dilution fridge measurements). aswartz@stanford.edu [1] R. Winkler, Spin-Orbit Coupling E ects in Two- Dimensional Electron and Hole Systems . [2] D. van der Marel, J. L. M. van Mechelen, and I. I. Mazin, Phys. Rev. B 84, 205111 (2011). [3] M. Kim, C. Bell, Y. Kozuka, M. Kurita, Y. Hikita, and H. Y. Hwang, Phys. Rev. Lett. 107, 106801 (2011). [4] X. Lin, Z. Zhu, B. Fauqu e, and K. Behnia, Phys. Rev. X3, 021002 (2013). [5] W. Qing-Yan, L. Zhi, Z. Wen-Hao, Z. Zuo-Cheng, Z. Jin- Song, L. Wei, D. Hao, O. Yun-Bo, D. Peng, C. Kai, et al. , Chin. Phys. Lett. 29, 037402 (2012). [6] A. D. Caviglia, S. Gariglio, N. Reyren, D. Jac- card, T. Schneider, M. Gabay, S. Thiel, G. Hammerl, J. Mannhart, and J.-M. Triscone, Nature 456, 624(2008). [7] M. Ben Shalom, M. Sachs, D. Rakhmilevitch, A. Palevski, and Y. Dagan, Phys. Rev. Lett. 104, 126802 (2010). [8] A. D. Caviglia, M. Gabay, S. Gariglio, N. Reyren, C. Can- cellieri, and J.-M. Triscone, Phys. Rev. Lett. 104, 126803 (2010). [9] H. Nakamura, T. Koga, and T. Kimura, Phys. Rev. Lett. 108, 206601 (2012). [10] Z. Zhong, A. T oth, and K. Held, Phys. Rev. B 87, 161102 (2013). [11] G. Khalsa, B. Lee, and A. H. MacDonald, Phys. Rev. B 86, 125121 (2012). [12] P. King, S. M. Walker, A. Tamai, A. De La Torre, T. Eknapakul, P. Buaphet, S. Mo, W. Meevasana, M. Bahramy, and F. Baumberger, Nature Commun. 5, 3414 (2014). [13] P. K. Rout, E. Maniv, and Y. Dagan, Phys. Rev. Lett. 119, 237002 (2017). [14] R. Meservey and P. Tedrow, Phys. Rep. 238, 173 (1994). [15] P. Fulde, Adv. Phys. 22, 667 (1973). [16] R. Meservey, P. Tedrow, and R. C. Bruno, Phys. Rev. B17, 2915 (1978). [17] A. M. Clogston, Phys. Rev. Lett. 9, 266 (1962). [18] B. Chandrasekhar, Appl. Phys. Lett. 1, 7 (1962). [19] X. Wu, P. Adams, Y. Yang, and R. McCarley, Phys. Rev. Lett. 96, 127002 (2006). [20] M. Kim, Y. Kozuka, C. Bell, Y. Hikita, and H. Hwang, Phys. Rev. B 86, 085121 (2012). [21] K. Maki, Pog. Theor. Phys. 32, 29 (1964). [22] J. Alexander, T. Orlando, D. Rainer, and P. Tedrow, Phys. Rev. B 31, 5811 (1985). [23] D. Worledge and T. Geballe, Phys. Rev. B 62, 447 (2000). [24] A. G. Swartz, H. Inoue, T. A. Merz, Y. Hikita, S. Raghu, T. P. Devereaux, S. Johnston, and H. Y. Hwang, Proc. Natl. Acad. Sci. 115, 1475 (2018). [25] H. Inoue, A. G. Swartz, N. J. Harmon, T. Tachikawa, Y. Hikita, M. E. Flatt e, and H. Y. Hwang, Phys. Rev. X5, 041023 (2015). [26] Y. Kozuka, Y. Hikita, C. Bell, and H. Y. Hwang, Appl. Phys. Lett. 97, 012107 (2010). [27] T. Yajima, M. Minohara, C. Bell, H. Kumigashira, M. Oshima, H. Y. Hwang, and Y. Hikita, Nano Lett. 15, 1622 (2015). [28] T. Tachikawa, M. Minohara, Y. Hikita, C. Bell, and H. Y. Hwang, Adv. Mater. 27, 7458 (2015). [29] Y. Hikita, K. Nishio, L. C. Seitz, P. Chakthranont, T. Tachikawa, T. F. Jaramillo, and H. Y. Hwang, Adv. Energy Mater. 6, 1502154 (2016). [30] W. L. McMillan and J. M. Rowell, Phys. Rev. Lett. 14, 108 (1965). [31] C. Richter, H. Boschker, W. Dietsche, E. Fillis-Tsirakis, R. Jany, F. Loder, L. Kourkoutis, D. Muller, J. Kirtley, C. Schneider, et al. , Nature 502, 528 (2013). [32] See Supplementary Information. [33] R. Dynes, V. Narayanamurti, and J. P. Garno, Phys. Rev. Lett. 41, 1509 (1978). [34] I. Zuti c, J. Fabian, and S. D. Sarma, Rev. Mod. Phys. 76, 323 (2004). [35] A. D. Caviglia, S. Gariglio, C. Cancellieri, B. Sac ep e, A. Fete, N. Reyren, M. Gabay, A. F. Morpurgo, and J.-M. Triscone, Phys. Rev. Lett. 105, 236802 (2010). [36] Y. J. Chang, A. Bostwick, Y. S. Kim, K. Horn, and E. Rotenberg, Phys. Rev. B 81, 235109 (2010).6 [37] A. Joshua, S. Pecker, J. Ruhman, E. Altman, and S. Ilani, Nature Commun. 3, 1129 (2012). [38] Y. Nakamura and Y. Yanase, J. Phys. Soc. Jpn. 82, 083705 (2013).[39] C. S ahin, G. Vignale, and M. E. Flatt e, Phys. Rev. B 89, 155402 (2014). [40] K. Ando and E. Saitoh, Nature Commun. 3, 629 (2012). [41] R. Jansen, S. P. Dash, S. Sharma, and B. C. Min, Semi- cond. Sci. Technol. 27, 083001 (2012).
1803.03549v1.Spin_vorticity_coupling_in_viscous_electron_fluids.pdf
arXiv:1803.03549v1 [cond-mat.mes-hall] 9 Mar 2018Spin-vorticity coupling in viscous electron fluids Ruben J. Doornenbal,1Marco Polini,2and Rembert A. Duine1,3 1Institute for Theoretical Physics, Utrecht University, Leuvenlaan 4, 3584 CE Utrecht, The Netherlands 2Istituto Italiano di Tecnologia, Graphene Labs, Via Morego 30, I-16163 Genova, Italy 3Department of Applied Physics, Eindhoven University of Tec hnology, P.O. Box 513, 5600 MB Eindhoven, The Netherlands (Dated: November 7, 2021) We consider spin-vorticity coupling—the generation of spi n polarization by vorticity—in viscous two-dimensional electron systems with spin-orbit couplin g. We first derive hydrodynamic equations for spin and momentum densities in which their mutual coupli ng is determined by the rotational viscosity. We then calculate the rotational viscosity micr oscopically in the limits of weak and strong spin-orbit coupling. We provide estimates that show that th e spin-orbit coupling achieved in recent experiments is strong enough for the spin-vorticity coupli ng to be observed. On the one hand, this coupling provides a way to image viscous electron flows by ima ging spin densities. On the other hand, we show that the spin polarization generated by spin-v orticity coupling in the hydrodynamic regime can, in principle, be much larger than that generated , e.g. by the spin Hall effect, in the diffusive regime. PACS numbers: 85.75.-d, 75.30.Ds, 04.70.Dy Introduction. —The field of spintronics is concerned with electric control of spin currents [1]. For the de- scription of experimentally relevant systems it has, until very recently, been sufficient to consider their coupled spin-charge dynamics in the diffusive regime where the time scale for electron momentum scattering is fast com- pared to other time scales. The celebrated Valet-Fert theory for electron spin transport in magnetic multilay- ers [2] and the Dyakonov-Perel drift-diffusion theory for spin generation by the spin Hall effect [3], for example, fall within this paradigm. Very recent experimental developments have brought about solid-state systems, such as ultra-clean encapsu- lated graphene, in which the momentum scattering time can be much longer than the time scale for electron- electron interactions [4–7]. In this so-called hydrody- namic regime, the electron momentum needs to be in- cluded as a hydrodynamic variable and the viscosity of the electron system cannot be neglected [8–17]. The finite electron viscosity leads to several physical conse- quences, such as a negative nonlocal resistance [4] and super-ballistic transport through point contacts [7, 18]. These developments have spurred on a great deal of re- search, including proposals for measuring the Hall vis- cosity [19–21] and connections to strong-coupling predic- tions from string theory [22]. In a seemingly unrelated development, spin- hydrodynamic generation, i.e. the generation of voltages from vorticity, was recently experimentally observed in liquid Hg [23]. Spin-hydrodynamic generation is believed to be a consequence of spin-vorticity coupling. Phe- nomenological theories of spin-vorticity coupling were developed early on [24] and have been applied to fluids consisting of particles with internal angular momentum such as ferrofluids [25], molecular nanofluids [26], andnematic liquid crystals [27]. In these phenomenological theories, the coupling between orbital angular momen- tum, i.e. vorticity of the fluid, and internal angular is governed by a dissipative coefficient, the so-called “rotational viscosity”. This type of viscosity has been estimated microscopically for classical systems (see e.g. [27]) and Hg [23], but not for viscous electrons in a crystal. Motivated by the recent realization of solid-state sys- tems hosting viscous electron fluids, we develop in this Letter the theory for spin-vorticity coupling in such sys- tems. Wederivethephenomenologicalequationsdescrib- ing coupled spin and momentum diffusion, and compute the rotational viscosity microscopically. We apply our theory to viscous electron flow through a point contact and show that the spin densities generated hydrodynam- ically can be much larger than the ones that are gen- erated by the spin Hall effect in the diffusive transport regime. Our results may therefore stimulate experimen- tal research towards novel ways of spin detection and generation. Phenomenology. —We consider two-dimensional (2D) electron systems with approximate translation invari- ance and approximate rotation invariance around the axis perpendicular to the plane (chosen to be the ˆz- direction). The conserved quantities of this system are energy, charge, linear momentum in the plane and an- gular momentum in the ˆz-direction. For brevity, we do not consider energy conservation explicitly and focus on momentum and angular momentum conservation. In the following, we follow the discussion of Ref. [24] and gen- eralize it to include spin diffusion and lack of Galilean invariance. The momentum density is denoted by p(r,t) and is a 2D vector p= (px,py) in the ˆx-ˆy-plane with r= (x,y) = (rx,ry). The total angular momentum den-2 sity in the ˆz-direction is the sum of orbital angular mo- mentum density ǫαβrαpβand spin density s(r,t) (in the ˆz-direction). Here, ǫαβis the 2D Levi-Civita tensor and summation over repeated indices α,β,γ,δ ∈ {x,y}is im- plied. We denote with vthe conjugate variable to the momentum density, i.e., the velocity, whereas the spin chemical potential, commonly referred to as spin accu- mulation,µsis the conjugate variable to the spin density. Conservation of linear momentum yields ∂pα(r,t) ∂t=−∂Παβ(r,t) ∂rβ, (1) with Π αβ(r,t) the stress tensor. Conservation of angular momentum in the z-direction is expressed as ∂[ǫαβrαpβ(r,t)+s(r,t)] ∂t=−∂jJ α(r,t) ∂rα,(2) withjJ α(r,t) theα-th component of the angular momen- tum current and in the above equations the summation is overboth αandβ. The equation for the spin density is found by subtracting the cross-product of rwith Eq. (1) from Eq. (2) and yields ∂s(r,t) ∂t=−∂js α(r,t) ∂rα−2Πa(r,t),(3) with Πa(r,t) =ǫαβΠβα(r,t)/2 the antisymmetricpart of the stress tensor and js α(r,t) =jJ α(r,t)−ǫβγrβΠγα(r,t) the spin current. Anonzerovelocityandspindensityincreasetheenergy of the system. By symmetry, a nonzero velocity leads to a contribution ρkinv2/2 to the energy density. This ex- pression defines the kinetic mass density ρkin, such that p(r,t) =ρkinv(r,t)[28]. Forthecasethatisofinterestto us, i.e., 2D electrons with spin-orbit coupling, the kinetic mass density is not equal to the average mass density ρ because spin-orbit coupling breaks Galilean invariance. Likewise, a nonzero spin density contributes χsµ2 s/2 to the energy density, where χsis the static spin suscepti- bility, so that s(r,t) =/planckover2pi1χsµs(r,t). These terms in the energy density lead to contributions to the entropy pro- duction from which relations between the fluxes (the spin current and antisymmetric part of the pressure tensor) and the forces (spin accumulation and velocity) are de- rived phenomenologically. In terms of µs(r,t) andv(r,t) we havefor the antisymmetric part of the pressuretensor that [24] Πa(r,t) =−ηr[ω(r,t)−2µs(r,t)//planckover2pi1],(4) withω(r,t) =ǫαβ∂vβ(r,t)/∂rαthe vorticity and ηrthe rotational viscosity . The aboveexpression showsthat an- gular momentum is transferred, by spin-orbit coupling, between orbital and spin degrees of freedom until the an- tisymmetric part of the pressure tensor is zero. For the spin current we have that js α(r,t) =−σs∂µs(r,t)/∂rα=−Ds∂s(r,t)/∂rαwhich defines the spin diffusion con- stantDsand spin conductivity σs, which obey the Ein- stein relation σs=/planckover2pi1Dsχs. Note that we are omitting an advective contribution ∼vαsto the spin current as we restrict ourselves to the linear-responseregime. Inserting these results for the fluxes into Eq. (3) and using Eq. (1) leads to ∂s(r,t) ∂t=Ds∇2s(r,t) +2ηr/bracketleftbigg ω(r,t)−2s(r,t) /planckover2pi12χs/bracketrightbigg −s(r,t) τsr; ρkin∂vα(r,t) ∂t=−eρEα m+νρkin∇2vα(r,t) +ηrǫαβ∂ ∂rβ/bracketleftbigg ω(r,t)−2s(r,t) /planckover2pi12χs/bracketrightbigg −ρkinvα(r,t) τmr.(5) In the above we have assumed the linear-response regime and introduced the kinematic viscosity νusing that the symmetric part of the stress tensor is given by Π αβ= νρkin∂vα/∂rβ. Furthermore, we have added spin and momentum relaxation terms, parameterized by the phe- nomenological time scales τsrandτmr, respectively. We have also included an electric field E(the electron has charge−e). Eqs. (5) are the main phenomenological equations for spin density and velocity. The term proportional to ηrin the first equation describes generation of spin accumula- tion in response to vorticity, e.g., spin-vorticity coupling. In the steady state the hydrodynamic equations are characterized by three length scales. The first is a length scalethatresultsfromthespin-vorticitycouplingequalto ℓsv=/radicalbig Ds/planckover2pi12χs/(2ηr), which is the characteristic length over which the orbital and spin angular momentum equi- librate. Furthermore, we have the spin diffusion length ℓsr=√Dsτsrthat determines the length scales for relax- ation of spin due to impurities, and the momentum dif- fusion length ℓmr=√ντmr. The most interesting regime, whichoccursinthe limitofstrongspin-orbitcouplingrel- ative to momentum and spin relaxation, is the one where ℓsvis the shortest length scale. In this case the spin den- sity locally follows the vorticity, which is determined by the electron flow. Application. —We consider electron flow through a point contact (PC) [7, 18] driven by a voltage V. Tak- ingτmr,τsr→ ∞we have from Ref. [18] for the velocity distribution at the PC that vy(x) =−πρeV 4mνρkin/radicalbigg/parenleftBigw 2/parenrightBig2 −x2, (6) wheretheflowisinthe y-directionand wisthePCwidth. From Eq. (5), in the limit ℓsv≪wthe steady-state spin density generated at the PC by spin-vorticity coupling in the hydrodynamic regime is then s(x) /planckover2pi12χsjc=−m πewρ4x/radicalbig (w/2)2−x2, (7)3 wherejc=−eρ/integraltext dxvy(x)/(mw) is the average current density. Let us compare Eq. (7) with the spin density gener- ated by the spin Hall effect in the diffusive limit. In the latter case, the spin accumulation is determined by∂2µs/∂x2=µs/ℓ2 sr, which follows from Eqs. (5) in the limitℓsr≪ℓsv, together with the expression js y= −σs∂µx/∂x+θSH/planckover2pi1jc y/(2e) for the spin current. Here jc y=σeEyis the diffusive charge current through the PC, withσe=e2ρ2τmr/(m2ρkin) the electrical conduc- tivity andθSHthe spin Hall angle. Using the boundary conditionsjs(−w/2) =js(w/2) = 0, we find for the spin density in the diffusive limit that sdiff(x) /planckover2pi12χsjcy=θSHℓsr 2eσssech/parenleftbiggw 2ℓsr/parenrightbigg sinh/parenleftbiggx ℓsr/parenrightbigg .(8) A crucial difference is thus that for diffusive spin trans- port and when w≫ℓsr, the spin density is only nonzero within a distance ∼ℓsraway from the edges of the PC, while when w≫ℓsvand in the hydrodynamic limit, the spin density [see Eq. (7)] is nonzero everywhere (except atx= 0 where it vanishes by symmetry). In both hydrodynamic and diffusive limits, the max- imum spin density occurs at the edges. In the hy- drodynamic limit the spin density formally diverges as |x| →w/2, since the vorticity that results from the ve- locity in Eq. (6) diverges in the same limit. This diver- gence is, however, unphysical, as there will be a micro- scopic length scale ℓedgeover which the velocity goes to zero near the edge of the sample, resulting in a maxi- mum spin density of |s(±w/2)|/(/planckover2pi12χsjc)∼m/(eρℓedge) near the edges of the sample. We expect the lat- ter to be much larger than the maximum spin density |sdiff(±w/2)/(/planckover2pi12χsjc)| ∼m2θSHℓsr/(e/planckover2pi1ρτmr) generated by the spin Hall effect in the diffusive regime (where we estimatedσs∼/planckover2pi1ρτmr/m2), because /planckover2pi1τmr/(mθSHℓsr)∼ ℓmr/(θSHkFℓsr) is expected to be much larger than the microscopic length scale ℓedge. Here,kFis the Fermi wave number. Microscopic theory. —We proceed by calculating the rotational viscosity microscopically. This is most easily achieved [29] by noting that even when spin relaxation due to impurities is absent ( τsr→ ∞), the spin-vorticity coupling opens a channel for spin relaxation, with rate 4ηr//planckover2pi12χs, which microscopically stems from the com- bined effect of spin-orbit coupling and electron-electron interactions. Hence, ηrcan be extracted from the re- tarded spin-spin response function (for spin in the ˆz- direction) at zero wave vector, denoted by χ(+) s(ω), when this response function is computed for a clean system with spin-orbit coupling and interactions. From Eqs. (5) we find that for v=0this response function has the form χ(+) s(ω) =χs 1−iω/planckover2pi12χs/(4ηr). (9)Hence, we have that 1 ηr=−/parenleftbigg2 /planckover2pi1χs/parenrightbigg2 lim ω→0Im[χ(+) s(ω)] ω.(10) As a representative example, we compute the rota- tionalviscosityusingstandardlinear-responsetechniques for a 2D electron gas with Rashba spin-orbit coupling, which has the following Hamiltonian [30]: ˆH=/integraldisplay dr/summationdisplay σ∈{↑,↓}ˆψ† σ(r)/bracketleftbigg −/planckover2pi12∇2 2m+λ/planckover2pi1ˆz·/parenleftbigg∇ i×τ/parenrightbigg/bracketrightbigg ˆψσ(r), (11) whereˆψσ(r) [ˆψ† σ(r)] is an electron annihilation [creation] operator and τis a vector of Pauli matrices. The unit vector in the ˆz-direction is denoted by ˆz. The con- stantλparametrizes the strength of spin-orbit interac- tions. The spin density operator in imaginary time τis ˆs(r,τ) =/planckover2pi1[ˆψ† ↑(r,τ)ˆψ↑(r,τ)−ˆψ† ↓(r,τ)ˆψ↓(r,τ)]/2, where the dependence on τof the electron creation and annihi- lationoperatorsindicates theircorrespondingHeisenberg evolution in imaginary time. We have for the imaginary- time spin-spin response function χs(iωn) =1 /planckover2pi1/integraldisplay dr/integraldisplay/planckover2pi1β 0dτ∝an}b∇acketle{tˆs(r,τ)ˆs(r,0)∝an}b∇acket∇i}ht0eiωnτ,(12) whereiωn= 2πn/(/planckover2pi1β) is a bosonic Matsubara frequency withβ= 1/(kBT) the inverse thermal energy, and the expectation value ∝an}b∇acketle{t···∝an}b∇acket∇i}ht0is taken at equilibrium. Neglect- ing vertex corrections due to interactions, this is worked out to yield χs(iωn) =−1 4/planckover2pi1V/summationdisplay k/summationdisplay δ/ne}ationslash=δ′/integraldisplay d/planckover2pi1ωd/planckover2pi1ω′Aδ(k,ω)Aδ′(k,ω′) ×/bracketleftbiggN(/planckover2pi1ω)−N(/planckover2pi1ω′) ω−ω′+iωn/bracketrightbigg , (13) withN(/planckover2pi1ω) =/bracketleftbig eβ(/planckover2pi1ω−µ)+1/bracketrightbig−1the Fermi-Dirac distri- bution function at chemical potential µ. The spectral functionsAδ(k,ω) are labeled by the Rashba spin-orbit- split band index δ=±. We incorporateelectron-electron interactions into the spectral function by taking them equal to Lorentzians broadened by the electron collision timeτee[this corresponds to dressing bare propagator lines in the spin bubble in Eq. (12) by self-energy inser- tions], i.e., Aδ(k,ω) =/planckover2pi1 2πτee1 [/planckover2pi1ω−/planckover2pi1ωδ(k)]2+/parenleftBig /planckover2pi1 2τee/parenrightBig2,(14) where/planckover2pi1ωδ(k) =/planckover2pi12k2/2m+δ/planckover2pi1λkis the Rashba band dis- persion. Inserting Eq. (14) into Eq. (13) and performing a Wick rotation iωn→ω+i0+yields ηr=4π2/planckover2pi14χ2 s mτee/bracketleftBigg 2π+8/parenleftbigµτee /planckover2pi1/parenrightbig 1+4/parenleftbigµτee /planckover2pi1/parenrightbig2+4tan−1/parenleftbigg2µτee /planckover2pi1/parenrightbigg/bracketrightBigg , (15)4 where we took λ→0. In the limit µτee//planckover2pi1≫1, we have ηr=π/planckover2pi14χ2 s/(mτee). Since we have neglected vertex corrections, the result in Eq. (15) does not vanish in the λ→0 limit and is strictly speaking only valid when spin-orbit coupling is so strong that the spin-vorticity coupling is limited by electron-electron interactions, i.e., when λkFτee≫1. In the opposite limit, where the bottleneck for spin relax- ation is the spin-orbit coupling, we perform a Fermi’s Golden Rule calculation to determine the decay rate of a spin polarization to second order in the strength of the spin-orbit interactions. This gives at low temperatures that ηr=−π/planckover2pi1 8/integraldisplaydk (2π)2A2(k,µ)(λ/planckover2pi1k)2,(16) whereA(k,µ) is the spectral function obtained from Eq. (14) by replacing /planckover2pi1ωδ(k)→/planckover2pi12k2/2m. Carrying out the remaining integral gives ηr=mλ2 2/planckover2pi1/bracketleftbigg 1+π/parenleftBigµτee /planckover2pi1/parenrightBig +2/parenleftBigµτee /planckover2pi1/parenrightBig tan−1/parenleftbigg2µτee /planckover2pi1/parenrightbigg/bracketrightbigg , (17) which indeed vanishes as λ→0. Whenµτee//planckover2pi1≫1, we have that /planckover2pi1ηr∼(λkF)(λkFτee), showing the dependence on the small parameter λkFτee≪1 explicitly. Inter- estingly, since the kinematic viscosity ν∝τee, we have that the rotational viscosity ηr∝1/νin the limit of strongspin-orbit coupling and ηr∝νin the limit ofweak spin-orbit coupling, with a maximum rotational viscosity whenλkFτee∼1. Estimates. —Next, we estimate the spin-vorticity cou- pling for graphene with proximity-induced spin-orbit coupling. We take λ/planckover2pi1kFto be on the order of 1 meV [32]. Furthermore, we take τee∼100 fs [4]. We thus have thatλ/planckover2pi1kFis about one order of magnitude smaller than/planckover2pi1/τeeand use the weak spin-orbit coupling expres- sion in Eq. (16). Evaluating Eq. (16) for a linear disper- sion/planckover2pi1vFk, wherevF∼106m/s is the graphene Fermi velocity, we find that ηr∼(λ/planckover2pi1kF)2 /planckover2pi1v2 F/parenleftBigµτee /planckover2pi1/parenrightBig , (18) usingµτee≫/planckover2pi1. We estimate the corresponding inverse time scale as ηr /planckover2pi12χs∼(λ/planckover2pi1kF)2 /planckover2pi13χsv2 F/parenleftBigµτee /planckover2pi1/parenrightBig ∼100 GHz,(19) where we took µτee//planckover2pi1∼10, and estimated the spin sus- ceptibility as χs∼D(µ), with the density of states at the Fermi level D(µ)∼√ne/(/planckover2pi1vF), and the electron number densityne∼1012cm−2[4]. To estimate the corresponding length scale ℓsv, we as- sume that spin diffusion is in the hydrodynamic regime determined by electron-electron interactions that lead tospin drag [33]. We then have for the spin diffusion con- stant thatDs∼/planckover2pi1ρτee/(m2χs). The spin-vorticity length scale is then ℓsv∼vF/planckover2pi1/radicalbig τeeχs/ηr∼1µm. This is the same order of magnitude as the momentum relax- ation length scale ℓmr[4], so that the rotational viscosity appears to be high enough to lead to observable spin- vorticity coupling. Moreover, the limit where ℓsv< ℓmr seems to be within experimental reach. Note that in the regime of weak spin-orbit coupling we have for the spin relaxation the Dyakonov-Perel result that 1 /τsr∝τmr [36], which yields that in the hydrodynamic regime we haveℓsr∼ℓsv/radicalbig τee/τmr≫ℓsv. A simple interpretation of the spin-vorticity coupling is that the electron spins are polarized by an effective magnetic field /planckover2pi1ω(r,t)/µB, withµBthe Bohr magneton, in the frame that rotates with the electron flow vorticity. We estimate the vorticity ω∼v/ℓmrusingℓmr∼0.1- 1µm, and a drift velocity of v∼100 m/s [4], which yields a substantial effective magnetic field of 1-10 mT. Discussion and conclusions. —We have developed the theory for spin-vorticity coupling in viscous electron flu- ids, both phenomenologically and microscopically, and we have estimated that the proximity-induced spin-orbit coupling in graphene is large enough for observable ef- fects. As an example, we predict alargespin polarization induced by spin-hydrodynamic generation in a PC. This large spin density may e.g. be observed optically [37] or via nitrogen-vacancy centre magnetometry [34, 35]. The imaged spin density would provide a fingerprint of the vorticity of the electron flow. An interesting direction for future research is general- ization of the phenomenological and microscopic deriva- tion to other spin-orbit couplings, including, in particu- lar, also the effects of violation of translational and ro- tational invariance beyond the phenomenological relax- ation terms that we included here. One example would be that of Weyl semi-metals that naturally have size- able spin-orbit coupling and have also been reported to be able to reach the hydrodynamic regime [38]. Other candidates are bismuthene [39] and stanene [40] that combine strong spin-orbit coupling with high mobility. Further interesting directions of research include incor- porating effects of a magnetic field and computation of the rotational viscosity in the regime where spin-orbit interactions and electron-electron interactions are com- parable in magnitude. In this regime, the crossover from weak-to-strong spin-orbit coupling takes place, whereas inclusion of momentum-relaxing scattering would lead to a crossover from the spin-vorticity coupling to the spin Hall effect. Acknowledgements. —We thank Denis Bandurin, Eu- gene Chudnovsky, and Harold Zandvliet for useful com- ments. R.D. is member of the D-ITP consortium, a program of the Netherlands Organisation for Scientific Research (NWO) that is funded by the Dutch Min- istry of Education, Culture and Science (OCW). This5 work is in part funded by the Stichting voor Funda- menteel Onderzoek der Materie (FOM) and the Euro- pean Research Council (ERC). M.P. is supported by the European Union’s Horizon 2020 research and inno- vation programme under grant agreement No. 696656 “GrapheneCore1”. [1] S. A. Wolf, D. D. Awschalom, R. A. Buhrman, J. M. Daughton, S. von Moln´ ar, M. L. Roukes, A. Y. Chtchelkanova, and D. M. Treger, Science 294, 1488 (2001). [2] T. Valet and A. Fert, Phys. Rev. B 48, 7099 (1993). [3] M. I. Dyakonov and V. I. Perel, Sov. Phys. JETP Lett. 13, 467 (1971). [4] D. A. Bandurin, I. Torre, R. K. Kumar, M. Ben Shalom, A. Tomadin, A. Principi, G. H. Auton, E. Khestanova, K. S. Novoselov, I. V. Grigorieva, L. A. Ponomarenko, A. K. Geim, and M. Polini, Science 351, 1055 (2016). [5] J. Crossno, J. K. Shi, K. Wang, X. Liu, A. Harzheim, A. Lucas, S. Sachdev, P. Kim, T. Taniguchi, K. Watanabe, T. A. Ohki, and K. C. Fong, Science 351, 1058 (2016). [6] P. J. W. Moll, P. Kushwaha, N. Nandi, B. Schmidt, and A. P. Mackenzie, Science 351, 1061 (2016). [7] R. K. Kumar, D. A. Bandurin, F. M. D. Pellegrino, Y. Cao, A. Principi, H. Guo, G. H. Auton, M. Ben Shalom, L. A. Ponomarenko, G. Falkovich, K. Watan- abe, T. Taniguchi, I. V. Grigorieva, L. S. Levitov, M. Polini, and A. K. Geim, Nature Phys. 13, 1182 (2017). [8] R. Gurzhi, Sov. Phys. JETP 44, 771 (1963). [9] J. E. Black, Phys. Rev. B 21, 3279 (1980). [10] Z. Z. Yu, M. Haerle, J. W. Zwart, J. Bass, W. P. Pratt, and P. A. Schroeder, Phys. Rev. Lett. 52, 368 (1984). [11] M. J. M. de Jong and L. W. Molenkamp, Phys. Rev. B 51, 13389 (1995). [12] M. Dyakonov and M. Shur, Phys. Rev. Lett. 71, 2465 (1993). [13] E. Chow, H. P. Wei, S. M. Girvin, and M. Shayegan, Phys. Rev. Lett. 77, 1143 (1996). [14] A. V. Andreev, S. A. Kivelson, and B. Spivak, Phys. Rev. Lett.106, 256804 (2011). [15] I. Torre, A. Tomadin, A. K. Geim, and M. Polini, Phys. Rev. B92, 165433 (2015). [16] A. Lucas, New J. Phys. 17, 113007 (2015). [17] L. Levitov and G. Falkovich, Nature Phys. 12, 672 (2016). [18] H. Guo, E. Ilseven, G. Falkovich, and L. S. Levitov, Proc . Natl. Acad. Sci. (USA) 114, 3068 (2017). [19] T. Scaffidi, N. Nandi, B. Schmidt, A. P. Mackenzie, and J. E. Moore, Phys. Rev. Lett. 118, 226601 (2017). [20] F. M. D. Pellegrino, I. Torre, and M. Polini, Phys. Rev. B96, 195401 (2017). [21] L. V. Delacr´ etaz and A. Gromov, Phys. Rev. Lett. 119, 226602 (2017). [22] J. M. Link, B. N. Narozhny, and J. Schmalian,arXiv:1708.02759. [23] R. Takahashi, M. Matsuo, M. Ono, K. Harii, H. Chudo, S. Okayasu, J. Ieda, S. Takahashi, S. Maekawa and E. Saitoh, Nature Phys. 12, 52 (2016). [24] S. R. de Groot and P. Mazur, Non-equilibrium thermo- dynamics (North-Holland, Amsterdam, 1962). [25] B. U. Felderhof, Phys. Fluids 23, 042001 (2011). [26] J. S. Hansen, P. J. Daivis, and B. D. Todd, Microfluids Nanofluidics 6, 785 (2009). [27] D.L. Cheung, S.J. Clark, M.R. Wilson, Chemical Physics Letters356, 140 (2002). [28] J. de Boer, J. Hartong, N. A. Obers, W. Sybesma, and S. Vandoren, arXiv:1710.04708. [29] Alternatively, the rotational viscosity may be evalua ted using the linear-response relation in Eq. (4). From this relation we find that ηrcan be extracted from the corre- lation function K(r−r′;τ−τ′) =/angbracketleftˆΠa(r,τ)ˆs(r′,τ′)/angbracketright0 at zero wave vector so that ηr=K(k→0,iωn)/2, where the factor 1 /2 compensates for the factor 2 in Eq. (4). Using Eq. (3) at zero wave vector we have that K(k→0;iωn) =−i/planckover2pi12ωnχs(iωn)/2, so that insertion of Eq. (9) indeed yields that K(k→0,iωn) = 2ηrin the “fast”ωn≫ηr/(/planckover2pi12χs) limit [31]. For completeness, we also compute the kinetic mass density for the Rashba model in the noninteracting limit. We then have that the momentum density is given by p=/summationdisplay δ/integraldisplaydk (2π)2/planckover2pi1kAδ(k,ω)N(/planckover2pi1ω−µ−/planckover2pi1k·v), which yields p=ρkinvforv→0, with the kinetic mass densityρkin=ρ[1 +λ2m/2µ+O(λ4)]. The correction, i.e. the second term between the brackets, is typically small since one almost always has that λ2m/2µ≪1. [30] A. Manchon, H. C. Koo, J. Nitta, S. M. Frolov and R. A. Duine, Nature Mater. 14, 871 (2015). [31] J.M. Luttinger, Phys. Rev. 135, A1505 (1964). [32] Z. Wang et al., Nat. Commun. 6, 8339 (2015). [33] I. D’??Amico and G. Vignale, Phys. Rev. B 62, 4853 (2000). [34] J. R. Maze, P. L. Stanwix, J. S. Hodges, S. Hong, J. M. Taylor, P. Cappellaro, L. Jiang, M. V. G. Dutt, E. Togan, A. S. Zibrov, A. Yacoby, R. L. Walsworth, and M. D. Lukin, Nature 455, 644 (2008). [35] L. Rondin, J.-P. Tetienne, T. Hingant, J.-F. Roch, P. Maletinsky, and V. Jacques, Rep. Prog. Phys. 77, 56503 (2014). [36] M. I. Dyakonov and V. I. Perel, Sov. Phys. Solid State 13, 3023 (1972) [37] V. Sih, R. C. Myers, Y. K. Kato, W. H. Lau, A. C. Gos- sard and D. D. Awschalom, Nature Phys. 1, 31 (2005). [38] J. Gooth, F. Menges, C. Shekhar, V. S¨ uss, N. Kumar, Y. Sun, U. Drechsler, R. Zierold, C. Felser, B. Gotsmann, arXiv:1706.05925. [39] F. Reis et al., Science 10.1126/science.aai8142 (2017). [40] F. Zhu et al., Nature Material 14. 1020 (2015).
1402.5817v2.Low_Energy_Effective_Hamiltonian_for_Giant_Gap_Quantum_Spin_Hall_Insulators_in_Honeycomb_X_Hydride_Halide__X_N_Bi__Monolayers.pdf
Low-Energy Effective Hamiltonian for Giant-Gap Quantum Spin Hall Insulators in Honeycomb X-Hydride/Halide ( X= N-Bi) Monolayers Cheng-Cheng Liu,1Shan Guan,1Zhigang Song,2Shengyuan A. Yang,3Jinbo Yang,2,4and Yugui Yao1, 1School of Physics, Beijing Institute of Technology, Beijing 100081, China 2State Key Laboratory for Mesoscopic Physics, and School of Physics, Peking University, Beijing 100871, China 3Engineering Product Development, Singapore University of Technology and Design, Singapore 138682, Singapore 4Collaborative Innovation Center of Quantum Matter, Beijing, China Usingthetight-bindingmethodincombinationwithfirst-principlescalculations,wesystematically derive a low-energy effective Hilbert subspace and Hamiltonian with spin-orbit coupling for two- dimensional hydrogenated and halogenated group-V monolayers. These materials are proposed to be giant-gap quantum spin Hall insulators with record huge bulk band gaps opened by the spin-orbit coupling at the Dirac points, e.g., from 0.74 to 1.08 eV in Bi X(X= H, F, Cl, and Br) monolayers. We find that the low-energy Hilbert subspace mainly consists of pxandpyorbitals from the group-V elements, and the giant first-order effective intrinsic spin-orbit coupling is from the on-site spin-orbit interaction. These features are quite distinct from those of group-IV monolayers such as graphene and silicene. There, the relevant orbital is pzand the effective intrinsic spin-orbit coupling is from the next-nearest-neighbor spin-orbit interaction processes. These systems represent the first real 2D honeycomb lattice materials in which the low-energy physics is associated with pxandpyorbitals. A spinful lattice Hamiltonian with an on-site spin-orbit coupling term is also derived, which could facilitate further investigations of these intriguing topological materials. PACS numbers: 73.43.-f, 73.22.-f, 71.70.Ej, 85.75.-d I. INTRODUCTION Recent years have witnessed great interest in two- dimensional (2D) layered materials with honeycomb lat- tice structures. Especially, the 2D group-IV honey- comb lattice materials, such as successively fabricated graphene,1,2and silicene,3,4have attracted considerable attention both theoretically and experimentally due to their low-energy Dirac fermion behavior and promising applications in electronics. Recently, we have discovered stable 2D hydrogenated and halogenated group-V hon- eycomb lattices via first-principles (FP) calculations.5 Their structures are similar to that of a hydrogenated silicene (silicane), as shown in Fig. 1(a). In the absence of spin-orbit coupling (SOC), the band structures show linear energy crossing at the Fermi level around Kand K0points of the hexagonal Brillouin zone. It is quite unusual that the low-energy bands of these materials are ofpxandpyorbital character. Previous studies in the context of cold atoms systems have shown that pxand pyorbital character could lead to various charge and or- bital ordered states as well as topological effects.6,7Our proposed materials, being the first real condensed mat- ter systems in which the low-energy physics is associated withpxandpyorbitals, are therefore expected to exhibit rich and interesting physical phenomena. The quantum spin Hall (QSH) insulator state has gen- erated great interest in condensed matter physics and materialscienceduetoitsscientificimportanceasanovel quantum state and its potential technological applica- tions ranging from spintronics to topological quantum computation.8–10This novel electronic state is gaped in the bulk and conducts charge and spin in gapless edge states without dissipation protected by time-reversal (b) ΓΚ Μ (a)FIG. 1. (Color online). (a) The lattice geometry for 2D X-hydride/halide ( X= N-Bi) monolayer from the side view (top) and top view (bottom). Note that two sets of sublat- tice in the honeycomb group V element Xare not coplanar (a buckled structure). The monolayer is alternatively hydro- genated or halogenated from both sides. (b) The first Bril- louin zone of 2D X-hydride/halide monolayer and the points of high symmetry. symmetry. The concept of QSH effect was first proposed by Kane and Mele in graphene in which SOC opens a nontrivial band gap at the Dirac points.11,12Subsequent works, however, showed that the SOC for graphene is tiny, hence the effect is difficult to be detected experi- mentally.13–15So far, QSH effect has only been demon- strated in HgTe/CdTe quantum wells,16,17and experi- mental evidence for helical edge modes has been pre- sented for inverted InAs-GaSb quantum wells.18–20Nev- ertheless, these existing systems more or less have serious limitationsliketoxicity,difficultyinprocessing,andsmall bulk gap opened by SOC. Therefore, an easy and envi-arXiv:1402.5817v2 [cond-mat.mtrl-sci] 22 Sep 2014ronmental friendly realization of a QSH insulator is much desired. Extensive effort has been devoted to the search for new QSH insulators with large SOC gap.21–28For instance, new layered honeycomb lattice type materials such as silicene, germanene24or stanene25, and chemi- cally modified stanene27have been proposed. Ultrathin Bi(111) films have drawn attention as a candidate QSH insulator, whose 2D topological properties have been re- ported.29An approach to design a large-gap QSH state on a semiconductor surface by a substrate orbital filter- ing process was also proposed.30However, desirable QSH insulators preferably with huge bulk gaps are still rare. A sizable bulk band gap in QSH insulators is essential for realizing many exotic phenomena and for fabricating new quantum devices that can operate at room temperature. Using FP method, we have recently demonstrated that the QSH effect can be realized in the 2D hydrogenated and halogenated group-V honeycomb monolayers family, with a huge gap opened at the Dirac points due to SOC.5 Although the low-energy spectrum of these materials is similar to the 2D group-IV honeycomb monolayers such as graphene and silicene, the low-energy Hilbert space changes from the pzorbital to orbitals mainly consisting ofpxandpyfrom the group-V atoms (N-Bi). More- over, the nature of the effective SOC differs between the two systems. Motivated by the fundamental inter- est associated with the QSH effect and huge SOC gaps in these novel 2D materials, we develop a low-energy ef- fective model Hamiltonian that captures their essential physics. In addition, we propose a minimal four-band latticeHamiltonianwiththeon-siteSOCtermusingonly thepxandpyorbitals. From the symmetry analysis, the next-nearest- neighbor (NNN) intrinsic Rashba SOC should exist in these systems due to the low-buckled structure, similar to the case of silicene.25However, as we shall see, the dominant effect is from the much larger first-order SOC of on-site origin. Therefore, in the following discussion, we shall focus on the first-order on-site SOC and neglect the higher-order effects. This point will be further dis- cussed later in this paper. The paper is organized as follows. In Sec. II, we de- rivestepbystepthelow-energyeffectiveHilbertsubspace and Hamiltonian for honeycomb X-hydride ( X= N-Bi) monolayers, and also investigate in detail the effective SOC.SectionIIIpresentsthederivationofthelow-energy effectivemodelfor X-halide( X=N-Bi, halide=F-I)hon- eycomb monolayers. In Sec. IV, a simple spinful lattice Hamiltonian for the honeycomb X-hydride/halide mono- layers family is constructed. We conclude in Sec. V with a brief discussion of the effective SOC and present a sum- mary of our results.II. LOW-ENERGY EFFECTIVE HAMILTONIAN FOR HONEYCOMB XH(X= N-BI) MONOLAYERS A. Low-energy Hilbert subspace and effective Hamiltonian without SOC As is shown in Fig. 1(a), there are two distinct sites A and B in the unit cell of X-hydride ( X= N-Bi) honeycomb lattice with full hydrogenation from both sides of the 2D Xhoneycomb sheet. The primitive lattice vectors are chosen as ~ a1=a(1=2;p 3=2)and ~ a2=a(1=2;p 3=2), whereais the lattice constant. We consider the outer shell orbitals of textitX ( X= N-Bi), namely s,px,py,pz, and also the sorbital of H in the modeling. Therefore, in the representation fjpA yi;jpA xi;jpA zi;jsA Hi;jsAi;jpB yi;jpB xi;jpB zi;jsB Hi;jsBig (for simplicity, the Dirac ket symbol is omitted in the following), the Hamiltonian (without SOC) at Kpoint with the nearest-neighbor hopping considered in the Slater-Koster formalism31reads H0=HAA 0HAB 0 HABy 0HBB 0 ; (1) with HAA 0=2 666640 0 0 0 0 0 0 0 0 0 0 0 0VH sp 0 0 0VH sp HVH ss 0 0 0 VH ss 3 77775;(2) HAB 0=2 66664V0 1iV0 10 0V0 2 iV0 1V0 10 0iV0 2 0 0 0 0 0 0 0 0 0 0 V0 2iV0 20 0 03 77775;(3) HBB 0=2 666640 0 0 0 0 0 0 0 0 0 0 0 0VH sp 0 0 0VH sp HVH ss 0 0 0VH ss 3 77775; (4) whereVH sp(VH ss) is the hopping between the pz(s) orbital from Xatom and the sorbital from H, and V0 1(3=4) (VppVpp)andV0 2(3=2)VspwithVpp, Vsp, andVppbeing the standard Slater-Koster hopping parameters. andHare on-site energies for sorbitals of atom Xand of atom H, respectively. The on-site en- ergies forporbitals are taken to be zero. To diagonalize the Hamiltonian, we first perform the 2(c)(a) (b) (d)FIG. 2. (Color online). (a)(b) The partial band structure projection for NH and NF without SOC, respectively. Symbol size is proportional to the population in the corresponding states. The Fermi level is indicated by the dotted line. (c)(d) Band structures for BiH and BiF without (black dash lines) and with (red solid lines) SOC. The four band structures are obtained from the first-principles methods implemented in the VASP package32using projector augmented wave pseudo-potential, and the exchange-correlation is treated by PAW-GGA. The Fermi level is indicated by the solid line. following unitary transformation: 'A 1=1p 2 pA x+ipA y =jpA +i; 'B 2=1p 2 pB xipB y =jpB i; '3=1p 2 1p 2 pA xipA y 1p 2 pB x+ipB y ; '4=1p 21p 2 pA xipA y 1p 2 pB x+ipB y :(5) In the basisf'A 1;sB;sB H;pB z;'B 2;sA;sA H;pA z;'3;'4g, the Hamiltonian can be written as a block-diagonal form with three decoupled blocks H ,H , andH : H0!H1=Uy 1H0U1; (6)U1=2 666666666666664ip 20i 2i 20 0 0 0 0 0 1p 201 21 20 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0ip 2i 2i 20 0 0 0 0 0 01p 21 21 20 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 13 777777777777775;(7) H1=H H H ; (8) with H =2 6640iV20 0 iV2VH ss 0 0VH ssHVH sp 0 0VH sp 03 775; (9) 3H =2 6640iV2 0 0 iV2VH ss 0 0VH ss HVH sp 0 0VH sp 03 775;(10) H =diagfV1;V1g; (11) whereV1= 2V0 1andV2=p 2V0 2. The eigenvectors for the first diagonal block H can be easily obtained as j"ii=1 Ni2 66641 i"i V2 i"2 i"iV2 2 V2VHss iVH sp "i"2 i"iV2 2 V2VHss3 7775;(12) where"iandNi(i= 1;2;3;4)are the correspond- ing eigenvalues and normalization factors, respectively. Therefore, upon performing the unitary transforma- tionf1;2;3;4g=f'A 1;sB;sB H;pB zgU withU = fj"iigi=1;2;3;4fu jig, the above upper-left 44block H is diagonalized. For the second diagonal block H , its eigenvalues are denoted as"4+i(i= 1;2;3;4), and it can be easily shown that"4+i="i, where"iare eigenvalues of H . This is consistent with FP results, i.e., there are four two-fold degeneracy points at Kpoint as shown in Fig. 2(a). The eigenvectors of H are given by j"ii=1 Ni2 66641 i"i V2 i"2 i"iV2 2 V2VH ss iVH sp "i"2 i"iV2 2 V2VHss3 7775;(13) where"iandNi(i= 5;6;7;8)are the corresponding eigenvalues and normalization factors. Similar to the case ofH , upon performing the unitary transforma- tionf5;6;7;8g=f'B 2;sA;sA H;pA zgU withU = fj"i+4igi=1;2;3;4fu jig, the block H is diagonalized. The third block H is already diagonal with eigenvalues fV1;V1gand eigenvectors f'3;'4g  f9;10g. Therefore, in the new basisf1;2;3;4;5;6;7;8;9;10g  'A 1;sB;sB H;pB z;'B 2;sA;sA H;pA z;'3;'4 U2, where U2u u I22, the total Hamiltonian (1) takes a fully diagonlized form. The whole diagonalization process can be summarized as follows: f1;2;3;4;5;6;7;8;9;10g = pA y;pA x,pA z;sA H;sA;pB y;pB x;pB z;sB H;sB U;(14) where U=U1U2; (15) H0!H0 0=UyH0U; (16)H0 0=diagf"1;"2;"3;"4;"5;"6;"7;"8;V1;V1g:(17) From the band components projection as shown in Fig. 2(a), in the vicinity of the Dirac points (around Fermilevel), themaincomponentsofthebandcomefrom thepxandpyorbitals of group-V element textitX mixed with a small amount of sorbital of textitX. Compared with the expressions of the eigenstates obtained above, we find that the orbital features agree with that of j"1i andj"5iifwetaketheireigenenergiesastheFermienergy. Therefore the corresponding states 1and5constitute thelow-energyHilbertsubspace. Inthefollowing, wewill give the explicit forms of the low-energy states 1and5 as well as their eigenvalues. Note that, in the above 44H , the scale of the 22 non-diagonal block H 12is smaller than the difference of the typical eigenvalues between the upper 22diagonal blockH 11and the lower 22diagonal block H 22. Hence, through the downfolding procedure33, we could obtain the low-energy effective Hamiltonian as Heff 11=H 11+H 12("H 22)1H 21:(18) Up to the second order, one obtains "1=1 2 0+q 02+ 4V2 2 ; (19) with 0=  +"VH ss2 "2H"VHsp2; "=1 2  +q 2+ 4V2 2 :(20) Consequently, we can obtain the explicit expressions of j"1i u j1 j=1;4and1. In a similar way, the explicit expressions ofj"5ifu j1gj=1;4and5can also be ob- tained. So far, we have obtained the eigenvalues "1="5 [Eqs. (19) and (20)] and the corresponding low-energy Hilbert subspace consisting of 1and5, 1=u 11'A 1+u 21sB+u 31sB H+u 41pB z; 5=u 11'B 2u 21sAu 31sA H+u 41pA z:(21) The above coefficients fu j1gj=1;4are given in Eq. (12). Further simplification could be made in order to cap- ture the main physics. We can omit the second-order correction for the eigenvalues and the first-order correc- tion for the eigenvectors, i.e., the terms (u 31sB H+u 41pB z) for1and(u 31sA H+u 41pA z)for5, and only keep the zeroth-order eigenvectors and eigenvalues, 1=u 11'A 1+u 21sB; 5=u 11'B 2u 21sA; "1="=1 2  +q 2+ 4V2 2 :(22) 4This approximation is justified by our FP calculations, namely in the vicinity of the Fermi level, px,py, and sorbitals overwhelmingly dominate over the sHandpz orbitals in the band components. In the Hamiltonian (17), one can take the Fermi en- ergyEF="1="5as energy zero point. Hence, states 1and5, which constitute the low-energy Hilbert sub- space, take the following explicit forms: 1=u 11 1p 2 pA x+ipA y +u 21sB; 5=u 111p 2 pB xipB y u 21sA;(23) with u 11=  +q 2+ 18V2sp r 22+ 36V2sp2q 2+ 18V2sp; u 21=3p 2iVspr 22+ 36V2sp2q 2+ 18V2sp: Since we are interested in the low-energy physics near theDiracpoint, weperformthesmall ~kexpansionaround Kby~k!~k+Kand keep the terms that are first order in~k. We find that HK= 0vFk vFk+ 0 ; (24) withvFbeing the Fermi velocity vF=p 3a 21 2ju 11j2(VppVpp) +ju 21j2Vss ; (25) and k=kxiky: Either following similar procedures, or using the in- version symmetry (or time-reversal symmetry ) of the system, we can easily obtain the low-energy Hilbert sub- space and the low-energy effective Hamiltonian around theK0point. Finally, we can summarize the basis for the low-energy Hilbert subspace as 1=u 11 1p 2 pA x+izpA y +u 21zsB; 5=u 111p 2 pB xizpB y u 21zsA;(26) and the low-energy effective Hamiltonian without SOC reads H=vF(kxx+zkyy); (27) where Pauli matrices denote the orbital basis degree of freedom, and z=1labels the two valleys Kand K0. Note that under the space inversion operation P= xxand the time-reversal operation T=x^K(^Kis the complex conjugation operator), the above low-energy effective Hamiltonian [Eq. (27)] is invariant.B. Low-energy effective Hamiltonian involving SOC The SOC can be written as Hso=0^L^s=0 2L+s+Ls+ 2+Lzsz ;(28) wheres=sxisyandL=LxiLydenote the ladder operators for the spin and orbital angular momenta, re- spectively. Here ^s= (~=2)~ s, and in the following we shall take ~= 1.0is the magnitude of atomic SOC. Because of the presence of pxandpyorbital component in the low-energy Hilbert subspace [Eq. (26)] f1;5g f";#g, an on-site effective SOC is generated with Hso=sozzsz; (29) where so=1 2ju 11j20 =1 22 419V2 sp 2q 2+ 18V2sp+ 18V2sp3 50: (30) Again we stress that in the honeycomb textitX-hydride monolayers the dominant intrinsic effective SOC is on- site rather than from the NNN hopping processes as in the original Kane-Mele model. Consequently, from the above Hamiltonian (27) and (29), we obtain the generic low-energy effective Hamil- tonian around the Dirac points acting on the low-energy Hilbert subspace: Heff=H+Hso=vF(kxx+zkyy) +sozzsz; (31) where the analytical expressions for Fermi velocity vF and magnitude of intrinsic effective SOC soare given in Eqs. (25) and (30), whose explicit values are presented in Table I via FP calculations. Again we note that the above spinful low-energy effective Hamiltonian is invari- ant under both the space-inversion symmetry operation and time-reversal symmetry operation with T=isyx^K. The two model parameters vFandsocan be obtained by fitting the band dispersions of the FP results. Their values are listed in Table I. III. LOW-ENERGY EFFECTIVE HAMILTONIAN FOR HONEYCOMB TEXTITX-HALIDE ( X= N-BI) MONOLAYERS A. Low-energy Hilbert subspace and effective Hamiltonian without SOC For the textitX-halide ( X= N-Bi) systems, the outer shellorbitalsofXlabeledas Xs,Xpx,Xpy,Xpz, andthe 5TABLE I. Values of Fermi velocity vFand magnitude of in- trinsic SOC sofor textitX-hydride honeycomb monolayers obtained from FP calculations. Note that so=Eg=2, with Egthe gap opened by SOC at the Dirac point. system vF 105m=s so(eV) NH 6.8 6:7103 PH 8.3 18103 AsH 8.7 97103 SbH 8.6 0.21 BiH 8.9 0.62 outer shell orbitals of halogen labeled as Hs,Hpx,Hpy, Hpzwith(H=F-I)aretakenintoaccountinthefollowing derivation. As is shown in Fig. 1(a), there are also two distinct sites A and B in the honeycomb lattice unit cell of textitX-halide with full halogenation from both sides ofthe2DtextitXhoneycombsheet. Intherepresentation fXpA y,XpA x,XpA z,HpA z,HpA y,HpA x,HsA,XsA,XpB y, XpB x,XpB z,HpB z,HpB y,HpB x,HsB,XsBgand at theK point, the total Hamiltonian with the nearest-neighbor hopping considered in the Slater-Koster formalism reads Hha 0= hAA 0hAB 0 hAB 0yhBB 0! ; (32) with hAA 0= 2 66666666666640 0 0 0 Vha pp 0 0 0 0 0 0 0 0 Vha pp 0 0 0 0 0 Vha pp 0 0Vha sp 0 0 0Vha pp ha p 0 0 0 Vha sp Vha pp 0 0 0 ha p 0 0 0 0Vha pp 0 0 0 ha p 0 0 0 0Vha sp 0 0 0 ha sVha ss 0 0 0 Vha sp 0 0Vha ss 3 7777777777775; (33) hAB 0=2 6666666666664V0 1iV0 10 0 0 0 0 V0 2 iV0 1V0 10 0 0 0 0iV0 2 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 V0 2iV0 20 0 0 0 0 03 7777777777775;(34)hBB 0= 2 66666666666640 0 0 0 Vha pp 0 0 0 0 0 0 0 0 Vha pp 0 0 0 0 0 Vha pp 0 0Vha sp 0 0 0Vha pp ha p 0 0 0Vha sp Vha pp 0 0 0 ha p 0 0 0 0Vha pp 0 0 0 ha p 0 0 0 0Vha sp 0 0 0 ha sVha ss 0 0 0Vha sp 0 0Vha ss 3 7777777777775; (35) where ha pis the on site energy for the porbitals of the halogen atom, (ha s) is the on site energy for the sor- bital of textitX (halogen) atom, the on site energies for p orbitals of textitX atoms are taken to be zero. Vha pp(Vha pp )isthehoppingbetweenthe pzorbitalfromtextitXatom and thepzorbital from halogen atom in the "shoulder by shoulder" ("head to tail") type. Vha spis the hopping be- tween thepz(s) orbital from textitX atom and the s(pz) orbital from halogen atom. Vha ssis the hopping between thesorbital from textitX atom and the sorbital from halogen atom. The parameters V0 1andV0 2take the same expressions as in Sec.II A. Firstly, we perform the unitary transformation as in Eq. (5), as well as the following unitary transformation H'A 1=1p 2 HpA x+iHpA y H'B 2=1p 2 HpB xiHpB y H'A 3=1p 2 HpA xiHpA y H'B 4=1p 2 HpB x+iHpB y: (36) In the new basis fX'A 1,XsB,H'A 1,HsB,XpB z,HpB z, X'B 2,XsA,H'B 2,HsA,XpA z,HpA z,X'3,X'4,H'A 3, H'B 4g=fXpA y,XpA x,XpA z,HpA z,HpA y,HpA x,HsA, XsA,XpB y,XpB x,XpB z,HpB z,HpB y,HpB x,HsB,XsBg Uha 1, we could rewrite the Hamiltonian in the following block-diagonalformwiththreedecoupleddiagonalblocks Hha 1=Hha 1; Hha 1; Hha 1; ; (37) Hha 1; =2 6666666640iV2Vha pp 0 0 0 iV2 0 Vha ss 0Vha sp Vha pp 0 ha p 0 0 0 0Vha ss 0 ha sVha sp 0 0 0 0 Vha sp 0Vha pp 0Vha sp 0 0Vha pp ha p3 777777775; (38) 6Hha 1; =2 6666666640iV2Vha pp 0 0 0 iV2 0 Vha ss 0Vha sp Vha pp 0 ha p 0 0 0 0Vha ss 0 ha sVha sp 0 0 0 0Vha sp 0Vha pp 0Vha sp 0 0 Vha pp ha p3 777777775; (39) Hha 1; =2 6666664V1 0Vha ppp 2Vha ppp 2 0V1Vha ppp 2Vha ppp 2 Vha ppp 2Vha ppp 2ha p 0 Vha ppp 2Vha ppp 20 ha p3 7777775:(40) For the first diagonal block Hha 1; , in the presentation fX'A 1;XsB;H'A 1;HsB;XpB z;HpB zgits eigenvectors can be written as j"ha ii=1 Nha i 2 666666666641 i C Vha pp "ha ihap i[Vha pp(Vha2 sp+Vha ppVha ss)"ha iVha ss("ha iha p)] DC iVha sp[ha sVha ppha pVha ss"ha i(Vha ppVha ss)] DC iVha spha2 sp+Vha ssVha pp"ha i("ha iha s)] DC3 77777777775;(41) with D "ha i  "ha iha s Vha2 pp"ha i "ha iha p + "ha iha p Vha2 sp; (42) and CV2 "ha iha p Vha2pp"ha i "ha ihap: (43) Here,"ha iandNha i(i= 1;2;;6)are the corresponding eigenvalues and the normalization factors, respectively. Therefore, by the unitary transformation  ha 1;ha 2;ha 3;ha 4;ha 5;ha 6 = X'A 1;XsB;H'A 1;HsB;XpB z;HpB z U ;(44) withU =fj"ha iigi=1;2;;6fu jig, the above 66block Hha 1; is diagonalized. From our FP calculations [Fig. 2(b)], the main compo- nents of the band around the Dirac points and the Fermi level come from the XpxandXpyorbitals, mixed with a small amount of the HpxandHpyorbitals as well as Xsorbital. The orbital features are identical with the eigenvectors of "ha 1. When we take its eigenvalue as theFermi energy EF. Following similar procedures as in the previous section, we can obtain the eigenvalues up to the second-order correction and the eigenvectors up to the first-order correction with "ha 1=1 20 @0+vuut02+ 4V2 220Vha2pp "hap+Vha4pp "hap21 A; (45) where 0=  "ha02 1 Vha2 ss+Vha2 sp "ha0 1 ha pVha2 ss+ ha sVha2 sp D "ha0 i + + Vha2 sp+Vha ssVha pp D "ha0 i; (46) "ha0 1=1 20 @ +vuut2+ 4V2 22Vha2pp "hap+Vha4pp "hap21 A; (47) and "=1 2  +q 2+ 4V2 2 : (48) Uptothispoint, wehavefoundthelow-energyeigenvalue "ha 1and the corresponding basis ha 1. Again, in order to capture the essential physics, we simply the above ex- pressions by taking only the zeroth-order terms. So in the following, we take "ha 1="ha0 1and omit the correc- tion withfHsB;XpB z;HpB zgfor the eigenvector fj"ha 1ig. Consequently, the eigenvector has the following form in the basisfX'A 1;XsB;H'A 1g j"ha 1i=1 nha 12 6641 iV2 "ha0 1 Vha pp "ha0 1hap3 7752 64uha 11 uha 21 uha 313 75;(49) withnha 1being a normalization constant, and the eigen- value"ha0 1is given in Eqs. (47) and (48). The eigenvalues of the second diagonal block Hha 1; are denoted as "ha 6+i(i= 1;2;;6), and one finds that "ha 6+i="ha i(i= 1;2;;6), where"ha iare eigenvalues of Hha 1; . Through similar procedures, the low-energy eigen- vectorfj"ha 7ighas the following simple form in the basis fX'B 2;XsA;H'B 2g: j"ha 7i=1 nha 12 6641 iV2 "ha0 1 Vha pp "ha0 1hap3 775=2 64uha 11 uha 21 uha 313 75:(50) 7The third diagonal block Hha 1; are of high energy hence is not of interest here. Fromtheaboveanalysis, thelow-energystates ha 1and ha 7constitute the low-energy Hilbert subspace. They have the following explicit forms: ha 1=uha 11 1p 2 XpA x+iXpA y +uha 21XsB +uha 31 1p 2 HpA x+iHpA y ; ha 7=uha 111p 2 XpB xiXpB y uha 21XsA +uha 311p 2 HpB xiHpB y :(51) Againweperformthesmall ~kexpansionintheabovelow- energy Hilbert subspace around Kpoint by~k!~k+K and keep the first-order terms in ~k, HK= 0vFk vFk+ 0! ; (52) withvFthe Fermi velocity vF=p 3a 21 2juha 11j2(VppVpp) +juha 21j2Vss : (53) Note that for the textitX-halide systems, juha 11j2is much larger than juha 21j2andjuha 31j2. Either follow- ing similar procedures, or via the inversion symmetry (or time-reversal symmetry ), one can obtain the low- energy Hilbert subspace and and the low-energy effective Hamiltonian around the K0point. Finally the basis for low-energy Hilbert subspace can be summarized as ha 1=uha 11 1p 2 XpA x+izXpA y +uha 21zXsB +uha 31 1p 2 HpA x+izHpA y ; ha 7=uha 111p 2 XpB xizXpB y uha 21zXsA +uha 311p 2 HpB xizHpB y :(54) and the low-energy effective Hamiltonian without SOC reads H=vF(kxx+zkyy); (55) where Pauli matrices denote the orbital basis degree of freedom, and zlabels the two valleys KandK0. Note thatunderthespacereversaloperation P=xxandthe time-reversal operation T=x^K, the above low-energy effective Hamiltonian Eq. (55) is also invariant.B. Low-energy effective Hamiltonian involving SOC In a similar way as in Sec. II B, we obtain an on- site SOC in the spinful low-energy Hilbert subspace f1;7g f";#g, Hso=sozzsz; (56) so=1 2juha 11j2X 0+1 2juha 31j2ha 0;(57) whereuha 11anduha 31are given in Eq. (49), and X 0(ha 0) is the magnitude of atomic SOC of pnictogen (halogen). It should be noted that due to the presence of major px andpyorbital components, the first-order on-site effec- tive SOC also dominates in the textitX-halide systems. Equation (49) explains the tendency that the soin- creases with the atomic number of halogen for the same pnictogen element, as shown in Table II. From Eqs. (55) and (56), we obtain the generic low- energyeffectiveHamiltonianaroundtheDiracpointsact- ing on the low-energy Hilbert subspace f1;7g f";#g Heff=H+Hso=vF(kxx+zkyy) +sozzsz; (58) where Fermi velocity vFand magnitude of intrinsic effec- tive SOCsoare given in Eqs. (53) and (57), and their values are listed in Table II. One notes that this Hamil- tonian is also invariant under both the space-inversion symmetry and time-reversal symmetry with T=isyx^K. The two model parameters vFandsofor halides ob- tained by fitting the band dispersions of the FP results are listed in Table II. IV. A SIMPLE SPINFUL LATTICE HAMILTONIAN FOR THE HONEYCOMB TEXTITX-HYDRIDE/HALIDE ( X= N-BI) MONOLAYERS FAMILY For the purpose of studying the topological proper- ties of the honeycomb textitX-hydride/halide ( X= N- Bi) monolayers family, as well as their edge states, it is convenient to work with a lattice Hamiltonian via lat- tice regularization of the low-energy continuum models (Eq. (31) and Eq. (58)). Taking into account the main physics involving pxandpyorbitals, we construct the fol- lowing spinful lattice Hamiltonian for the 2D honeycomb textitX-hydride/halide ( X= N-Bi) monolayers H=X hi;ji; ; =px;pyt ijcy i cj +X i; ; =px;py;;0=";# ;0cy i ci 0sz ;0;(59) wherehi;jimeansiandjsites are nearest neighbors, and are the orbital indices. The first term is the hoppingtermandthesecondoneistheon-siteSOCterm. 8TABLE II. Values of two model parameters vFandsofor honeycomb textitX-halide ( X= N-Bi) monolayers obtained from FP calculations. Note that so=Eg=2, withEgthe gap opened by SOC at the Dirac point. system vF 105m=s so(eV) system vF 105m=s so(eV) NF 5.5 8:5103NBr 4.2 19103 PF 7.2 13103PBr 8.0 17103 AsF 7.3 80103AsBr 8.2 98103 SbF 6.6 0.16 SbBr 7.7 0.20 BiF 7.2 0.55 BiBr 7.3 0.65 NCl 4.3 9:7103NI 3.8 28103 PCl 7.8 17103PI 8.1 19103 AsCl 8.0 95103AsI 9.1 0:10 SbCl 7.3 0.19 SbI 7.7 0.21 BiCl 6.9 0.56 BiI 7.7 0.65 After Fourier transformation of the above lattice Hamiltonian, its energy spectrum over the entire Bril- louin zone can be obtained. Since here spin is good quan- tum number, we can divide the model Hamiltonian into two sectors for spin up and spin down separately. For each sector, the corresponding model Hamiltonian reads H"(k) =2 66640i0 2hAB xx(k)hAB xy(k) 0hAB xy(k)hAB yy(k) 0i0 2y03 7775;(60) H#(k) =2 66640i0 2hAB xx(k)hAB xy(k) 0hAB xy(k)hAB yy(k) 0i0 2y03 7775;(61) where hAB xx(k)1 2(3Vpp+Vpp) coskx 2 exp iky 2p 3 +Vppexp ikyp 3 ; hAB xy(k)ip 3 2(VppVpp) sinkx 2 exp iky 2p 3 ; and hAB yy(k)1 2(Vpp+ 3Vpp) coskx 2 exp iky 2p 3 +Vppexp ikyp 3 : For simplicity, we choose the lattice constant a= 1. The on-site energies for porbitals are taken to be zero. Near theKandK0points, the above model Hamiltonian re- duces to the low-energy effective Hamiltonian [Eq. (31) and (58)] with vF=p 3a 4(VppVpp)andso=0=2. -4-3-2-101234 Energy(eV)Κ ΜΓ Γ FIG. 3. (Color online). A comparison of the band structures for monolayer SbH calculated using FP and TB methods with SOC . The dashed green curve is the FP result. The solid red and blue curves are the TB model results. The red curve is with the NN hopping only, while the blue curve also includes the NNN hopping terms. For the NN case, the parameters are taken as Vpp= 1:68eV,Vpp=0:60eV. For the NNN case, the parameters are taken as Vpp= 1:69eV,Vpp= 0:62eV,VNNN pp = 0eV,VNNN pp =0:23eV. For both cases, so= 0:21eV. The superscript NNN means the next-nearest- neighbor hoping. The Fermi level is set to zero. Taking SbH as an example, we compare the results from FP calculations and from the lattice models. As shown in Fig. 3, there is a good agreement between the two results around the Kpoint. The fitting away from Kpoint can be improved by including hopping terms between far neighbors. In Fig. 3, we also show the result with NNN hopping, for which a fairly good agreement with the FP low-energy bands over the whole Brillouin zone can be achieved. 9V. DISCUSSION AND SUMMARY Wehaveobtainedthelow-energyeffectiveHamiltonian for the textitX-hydride and textitX-halide ( X= N-Bi) family of materials, which is analogous to the Kane-Mele model proposed for the QSH effect in graphene.11The important difference is that in Kane-Mele model the ef- fective SOC is of second-order NNN type, which is much weaker than the on-site SOC in our systems. The SOC term in our Hamiltonian opens a large nontrivial gap at the Dirac points. From KtoK0the mass term changes sign for each spin species and the band is in- verted. As a result, the QSH effect can be realized in the textitX-hydride and textitX-halide ( X= N-Bi) mono- layers. Some of these materials, such as BiH/BiF, have record huge SOC gap with magnitude around 1 eV, far higher than the room-temperature energy scale, hence making their detection much easier. On the experimental side, the buckling honeycomb Bi(111) monolayer and film have been manufactured via molecular beam epitaxy (MBE).23,29,34On the other hand, chemical functionalization of 2D materials is a powerful tool to create new materials with desirable features, such as modifying graphene into graphane, graphone, and fluorinated graphene via H and F, re- spectively.35Therefore, it is very promising that Bi- Hydride/Halidemonolayer, thehugegapQSHinsulators, may be synthesized by chemical reaction in the solvents or by the exposure of the Bi (111) monolayer and film to the atomic or molecular gases. It is noted that even though one side (full passivation) instead of both sides (alternatingpassivation)ofBi(111)bilayersispassivated, the band structure is almost unchanged and the topol- ogy properties remain nontrivial. This will provide more freedom to realize these kinds of materials. It is known that the low-energy Hilbert space for graphene consists of the pzorbital from carbon atoms. In that system, the SOC term from NNN second-order pro- cesses is vanishingly small, and the on-site SOC as well as the nearest neighbor SOC are forbidden by symme- try constraint. In contrast, for the honeycomb textitX- hydride/halide monolayers, pxandpyorbitals from the group V elements constitute the low-energy Hilbert sub- space. In fact, this represents the first class of materials for which the Dirac fermion physics is associated with pxandpyorbitals. Because of this, the effective on-site SOC can has nonzero matrix elements and results in the huge SOC gap at the Dirac points. The leading-order effective SOC processes in the textitX-hydride and textitX-halide systems, silicene, and graphene are schematically shown in Fig. 4. As shown in Figs. 44(a) and 4(b), the representative leading-order effective SOCprocesses aroundthe Kpoint inthe honey- comb textitX-hydride and textitX-halide monolayers are jpA +"iso!jpA +"i;jpA +#iso!jpA +#i; jpB "iso!jpB "i;jpB #iso!jpB #i;(62)wheresorepresents the atomic spin-orbit interaction strength, which is given in Eq. (30) for textitX-hydride systems and Eq. (57) for textitX-halide systems. In a Hilbert subspace consisting of pxandpyorbitals, such effective SOC arises in the first-order on-site processes, which leads to its huge magnitude. As for silicene, which has a low-buckled structure, the typical leading-order SOC is from the (first-order) NNN processes,25as shown in Fig. 4(c), jpA z"iV!jpB "i0 2!jpB "iV!jpA z"i; jpA z#iV!jpB #i0 2!jpB #iV!jpA z#i; jpB z"iV!jpA +"i0 2!jpA +"iV!jpB z"i; jpB z#iV!jpA +#i0 2!jpA +#iV!jpB z#i;(63) whereVis the nearest-neighbor direct hopping ampli- tudeand0representstheatomicintrinsicSOCstrength. The whole process can be divided into three steps. For example, we consider the pA zorbital. Firstly, due to the low-buckled structure, pA zcouples topB . Carriers in pA z orbital then hop to the nearest neighbor pB orbital. Sec- ondly, the atomic intrinsic SOC shifts the energy of the spin up and spin down carriers by 0 2. In the third step, carriersin the pB orbitalhop toanothernearest-neighbor pA zorbital, making the resulting effective SOC an NNN process and of first order in 0. As for graphene, around Dirac point, the leading-order effective SOC is from (second-order) NNN effective SOC process, as shown in Fig. 4(d): jpA z"i0=p 2! jpA +#iV!jsB #iV!jpA +#i0=p 2! jpA z"i; jpB z#i0=p 2! jpB "iV!jsA "iV!jpB "i0=p 2! jpB z#i:(64) During the whole NNN hopping process, the atomic SOC appears twice, making the effective SOC second order in 0and hence much weaker. In summary, using the TB method and the FP calcu- lation, we have derived the low-energy effective Hilbert subspace and Hamiltonian for the honeycomb textitX- hydride/halide monolayers materials. These 2D group-V honeycomb lattice materials have the same low-energy effective Hamiltonian due to their same D3dpoint group symmetry and the same D3small group at the Kand K0points. The low-energy model contains two key pa- rametersvFandso. We have obtained their analytic expressions and also their numerical values by fitting the FP calculations. Moreover, we have found that the low- energy Hilbert subspace consists of pxandpyorbitals from the group-V elements, which is a key reason for the huge SOC gap. This feature is distinct from the group- IV honeycomb lattice monolayers such as silicene and graphene. Finally, we construct a spinful lattice Hamil- tonian for these materials. Our results will be useful for further investigations of this intriguing class of materials. 10ππ πσ σsoc ππ πσσsocsocσ σsoc σσσ σ σsocσ AB AB BB A A(a) (b) (c) (d)FIG. 4. (Color online). The leading-order effective SOC pro- cesses in textitX-hydride or textitX-halide ( X= N-Bi), sil- icene and graphene. (a) and (b) Sketches of the huge effective on-site SOC in textitX-hydride systems and textitX-halide systems. (c) Sketch of the effective SOC from NNN hopping processes caused by the buckling in silicene. (d) Sketch of the second-order effective SOC from NNN hopping processes in graphene.ACKNOWLEDGMENTS This work was supported by the MOST Project of China (Nos. 2014CB920903, 2010CB833104, and 2011CBA00100), the National Natural Science Foun- dation of China (Grant Nos. 11225418, 51171001, and 11174337), SUTD-SRG-EPD2013062, and the Spe- cialized Research Fund for the Doctoral Program of HigherEducationofChina(GrantNo. 20121101110046). Cheng-Cheng Liu was supported Excellent young schol- ars Research Fund of Beijing Institute of Technology (Grant No. 2014CX04028). Note added Recently, we notice another relevant work36discussing effective models of a honeycomb lat- tice withpxandpyorbitals. ygyao@bit.edu.cn 1A. K. Gelm and K. S. Novoselov, Nat. Mater. 6, 183-191 (2007). 2A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, and A. K. Geim, Rev. Mod. Phys. 81, 109 (2009). 3P. Vogt, P. De Padova, C. Quaresima, J. Avila, E. Frantzeskakis, M. C. Asensio, A. Resta, B. Ealet, and G. Le Lay, Phys. Rev. Lett. 108, 155501 (2012). 4L. Chen, C.-C. Liu, B. Feng, X. He, P. Cheng, Z. Ding, S. Meng, Y. Yao, and K. Wu, Phys. Rev. Lett. 109, 056804 (2012). 5Z. Song, C.-C. Liu, J. Yang, J. Han, B. Fu, Y. Yang, Q. Niu, J. Lu, and Y.G. Yao, arXiv:cond-mat/1402.2399. 6C. Wu, D. Bergman, L. Balents, and S. Das Sarma, Phys. Rev. Lett. 99, 070401 (2007). 7C. Wu, Phys. Rev. Lett. 100, 200406 (2008). 8M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010). 9X. Qi and S. Zhang, Rev. Mod. Phys. 83, 1057 (2011). 10B. Yan and S.-C. Zhang, Rep. Prog. Phys. 75, 096501 (2012). 11C. L. Kane and E. J. Mele, Phys. Rev. Lett. 95, 226801 (2005). 12C. L. Kane and E. J. Mele, Phys. Rev. Lett. 95, 146802 (2005). 13Y. G. Yao, F. Ye, X. L. Qi, S. C. Zhang, and Z. Fang, Phys. Rev. B 75, 041401(R) (2007). 14H.Min, J.E.Hill, N.A.Sinitsyn, B.R.Sahu, L.Kleinman, and A. H. MacDonald, Phys. Rev. B 74, 165310 (2006). 15S. Konschuh, M. Gmitra, and J. Fabian, Phys. Rev. B 82, 245412 (2010).16B. A. Bernevig, T. L. Hughes, and S. C. Zhang, Science 314, 1757-1761 (2006). 17M. K onig, S. Wiedmann, C. Br une, A. Roth, H. Buhmann, L. W. Molenkamp, X. L. Qi, and S.C. Zhang, Science 318, 766-770 (2007). 18C. X. Liu, T. L. Hughes, X. L. Qi, K. Wang, and S. C. Zhang, Phys. Rev. Lett. 100, 236601 (2008). 19I. Knez, R.R. Du, and G. Sullivan, Phys. Rev. Lett. 107, 136603 (2011). 20I. Knez, R.R. Du, and G. Sullivan, Phys. Rev. Lett. 109, 186603 (2012). 21S. Murakami, Phys. Rev. Lett. 97, 236805 (2006). 22Z. Liu, C. X. Liu, Y.S. Wu, W.H. Duan, F. Liu, and Jian Wu, Phys. Rev. Lett. 107, 136805 (2011). 23T. Hirahara, G. Bihlmayer, Y. Sakamoto, M. Yamada, H. Miyazaki, S.I. Kimura, S. Bl ugel, and S. Hasegawa, Phys. Rev. Lett. 107, 166801 (2011). 24C.-C. Liu, W. Feng, and Y. Yao, Phys. Rev. Lett. 107, 076802 (2011). 25C.-C. Liu, H. Jiang, and Y. Yao, Phys. Rev. B 84, 195430 (2011). 26C. Weeks, J. Hu, J. Alicea, M. Franz, and R. Wu, Phys. Rev. X 1, 021001 (2011). 27Y. Xu, B. Yan, H.-J. Zhang, J. Wang, G. Xu, P. Tang, W. Duan, and S.-C. Zhang, Phys. Rev. Lett. 111, 136804 (2013). 28Z.F. Wang, Z. Liu, and F. Liu, Nat. Commun. 4, 1471 (2013). 29F. Yang,L. Miao, Z. F. Wang, M.-Y. Yao, F. Zhu, Y. R. Song, M.-X. Wang, J.-P. Xu, A. V. Fedorov, Z. Sun, G. B. Zhang, C. Liu, F. Liu, D. Qian, C. L. Gao, and J.-F. Jia, Phys. Rev. Lett. 109, 016801 (2012). 1130M. Zhou, W. Ming, Z. Liu, Z. Wang, Y.G. Yao, and F. Liu, arXiv:cond-mat/1401.3392. 31J. C. Slater and G. F. Koster, Phys. Rev. 94, 1498 (1954). 32G. Kresse and J. Furthm uller, Phys. Rev. B 54, 11169- 11186 (1996). 33R. Winkler, Spin-Orbit Coupling Effects in Two- Dimensional Electron and Hole Systems , 1st ed. (Springer, Berlin, 2003). 34C. Sabater, D. Gos albez-Mart inez, J. Fern andez-Rossier, J. G. Rodrigo, C. Untiedt, and J. J. Palacios, Phys. Rev. Lett.110, 176802 (2013). 35J. O. Sofo, A. S. Chaudhari, and G. D. Barber, Phys. Rev.B75, 153401 (2007); D. C. Elias, R. R. Nair, T. M. G. Mohiuddin, S. V. Morozov, P. Blake, M. P. Halsall, A. C. Ferrari, D. W. Boukhvalov, M. I. Katsnelson, A. K. Geim, K. S. Novoselov, Science 323, 610 (2009); J. Zhou, Q. Wang, Q. Sun, X. S. Chen, Y. Kawazoe, and P. Jena, Nano Lett. 9, 3867 (2009); J. T. Robinson, J. S. Burgess, C. E. Junkermeier, S. C. Badescu, T. L. Reinecke, F. K. Perkins, M. K. Zalalutdniov, J. W. Baldwin, J. C. Cul- bertson, P. E. Sheehan, and E. S. Snow, ibid. 10, 3001 (2010). 36G.-F. Zhang, Y. Li, and C. Wu, arXiv:cond- mat/1403.0563. 12
2206.05041v1.Quantum_heat_engine_based_on_a_spin_orbit_and_Zeeman_coupled_Bose_Einstein_condensate.pdf
Quantum heat engine based on a spin-orbit and Zeeman-coupled Bose-Einstein condensate Jing Li,1,E. Ya Sherman,2, 3, 4and Andreas Ruschhaupt1 1Department of Physics, University College Cork, T12 H6T1 Cork, Ireland 2Departamento de Qu´ ımica-F´ ısica, UPV /EHU, Apartado 644, 48080 Bilbao, Spain 3IKERBASQUE, Basque Foundation for Science, 48011 Bilbao, Spain 4EHU Quantum Center, University of the Basque Country UPV /EHU We explore the potential of a spin-orbit coupled Bose-Einstein condensate for thermodynamic cycles. For this purpose we propose a quantum heat engine based on a condensate with spin-orbit and Zeeman coupling as a working medium. The cooling and heating are simulated by contacts of the condensate with an external magnetized media and demagnetized media. We examine the condensate ground state energy and its dependence on the strength of the synthetic spin-orbit and Zeeman couplings and interatomic interaction. Then we study the eciency of the proposed engine. The cycle has a critical value of spin-orbit coupling related to the engine maximum e ciency. Introduction Quantum cycles are of much importance both for fundamental research and for applications in quantum- based technologies[1, 2]. Quantum heat engines have been demonstrated in recent on several quantum platforms, such as trapped ions [3, 4], quantum dots [5] and optomechanical os- cillators [6–9]. Well-developed techniques for experimental control make Bose-Einstein condensates (BECs) [10] a suit- able system for a quantum working medium of a thermal ma- chine [11–13]. Recently, a quantum Otto cycle was experimentally realized using a large quasi-spin system with individual cesium (Cs) atoms immersed in a quantum heat bath made of ultracold ru- bidium (Rb) atoms [14, 15]. Several spin heat engines have been theoretically and experimentally implemented using a single-spin qubit [16], ultracold atoms [17], single molecule [18], a nuclear magnetic resonance setup [19] and a single- electron spin coupled to a harmonic oscillator flywheel [20]. These examples have motivated our exploration of the spin- orbit coupled BEC considered in this paper. Spin-orbit coupling (SOC) links a particle’s spin to its motion, and artificially introduces charge-like physics into bosonic neutral atoms [21]. The experimental generation [22–25] of SOC is usually accompanied by a Zeeman field, which breaks various symmetries of the underlying system and induces interesting quantum phenomena, e.g. topological transport[26]. In addition, in the spin-orbit coupled BEC sys- tem, more studies on moving solitons [27–29], vortices [30], stripe phase [31] and dipole oscillations [32] have been re- ported. In this paper, we propose a BEC with SOC as a working medium in a quantum Stirling cycle. The classic Stirling cy- cle is made of two isothermal branches, connected by two isochore branches. The BEC is characterized by SOC, Zee- man splitting, a self-interaction, and is located in a quasi- one-dimensional vessel with a moving piston that changes the length of the vessel. The external ”cooling” and ”heating” reservoirs are modelled by the interaction of the spin-1 /2 BEC with an external magnetized and demagnetized medias. The expansion and compression works depend on the SOC and Corresponding author: jli@ucc.ieZeeman coupling. A main goal is to examine the condensate ground state energy and its dependence on the strength of the synthetic spin-orbit, Zeeman couplings, interatomic interac- tion and length of the vessel. For the semiquantitative analy- sis, perturbation theory is applied to understand the e ects of SOC and Zeeman splitting. We further analyze several impor- tant parameters and investigate how they a ect the e ciency of the cycle, e.g. the critical SOC strength for di erent self- interactions. Model of the heat engine: Working medium We consider a quasi-one dimensional BEC, extended along the xaxis and tightly confined in the orthogonal directions. The mean-field energy functional of the system is then given by E=R+1 1"dx with spin-independent self-interaction of the Manakov’s sym- metry [33]: "= yH0 +g 2(j "j2+j #j2)2; (1) where ( "; #)T(here T stands for transposition) and the wavefunctions "and #are related to the two pseudo- spin components. The parameter grepresents the strength of the atomic interaction which can be tuned by atomic swave scattering length using Feshbach resonance [34, 35] with g> 0,g<0, and g=0 giving the repulsive, attractive, and no atomic interaction, respectively. The Hamiltonian H0in Eq. (1) of the spin-1 /2 BEC, trapped in an external potential V(x), is given by H0=ˆp2 2mˆ0+ ~ˆpˆx+~ 2ˆz+V(x); (2) with ˆ p=i~@xbeing the momentum operator in the longi- tudinal direction, ˆ x;zbeing the Pauli matrices, and ˆ 0being the identity matrix. Here is the SOC constant and is the Zeeman field. We choose a convenient length unit , an energy unit~2=(m2) and a time unit m2=~and express the following equations in the corresponding dimensionless variables. The coupled Gross-Pitaevskii equations are now given by i@ @t "= 1 2@2 @x2+ 2+g n(x)+V(x)! "i @ @x #;(3) i@ @t #= 1 2@2 @x2 2+g n(x)+V(x)! #i @ @x ";(4)arXiv:2206.05041v1 [cond-mat.quant-gas] 10 Jun 20222 a a aP =0 >0A B CD magnetization source magnetization source B C demagnetization source D Ademagnetization source 1 2(a) (b) FIG. 1. (a) The schematic diagram of the quantum Stirling cycle based on the Zeeman and SOC. (b) Visualization of the demagne- tization (left) and magnetization (right) processes with the external sources; the blue dots represent the BEC atoms and the orange dots represent the external source. where the density is given by n(x)=j "j2+j #j2. We fix the norm N=R1 1n(x)dx=1. We consider a hard-wall potential V(x) of half width a: V(x)=0;(jxja); V(x)=1(jxj>a):(5) This potential is analogous to a piston in a thermodynamic cycle and it allows one to define the work of the quantum cy- cle. The ground state of the BEC then depends on the half width a, the detuning , the interactions gand the SOC , i.e. ;g(a;), and the corresponding total ground state energy of the BEC is then denoted as E ;g(a;). We define also the pres- sure P ;g(a;) as a measure of the energy E ;g(a;) stored per total length 2 a: P ;g(a;)@E ;g(a;) 2@a: (6) In the special case of  = 0 and for the spin-independent self-interaction proportional to n(x), the energy [36, 37] is given by E ;g(a;0)=E0;g(a;0) 2=2 resulting in independent pressure P ;g(a;0). Notice that at both nonzero and, the system is characterized by a magne- tostriction in the form M ;g(a;)=@P ;g(a;)=@: Model of the heat engine: Quantum Stirling cycle We con- sider a quantum Stirling cycle keeping the interaction gand the SOC fixed during the whole process. The key idea is that the external ”cooling” and ”heating” reservoirs are modelled by the interaction of the spin 1 /2 BEC with an external mag- netized media (see Fig. 1(b), right) resp. demagnetized media (see Fig. 1(b), left). This external, (de)magnetized source leads to a random magnetic field in the condensate and be- cause of the Zeeman-e ect this corresponds to a detuning ofthe condensate to with some probability density distribution p(). We assume that this external source brings the system to a stationary state with the condensate described by a density operator ˆ=Z p()j ;g(a;)ih ;g(a;)jd: (7) The probability density distribution of the demagnetizing source pdm() is centered around hidmR pdm()d =0 while the one of the magnetizing source pm() is centered around a positive value him>0. As an increase in de- creases the BEC energy [10] by an dependent amount, the demagnetization source plays the role of a “hot thermal bath” here and the magnetization source plays the role of a “cold thermal bath”. In general there could exist a stationary exter- nal magnetic field leading to an additional detuning during the cycle. We neglect this possibility in the following in order to simplify the notation. The realization of the Stirling cycle is described by a four- stroke protocol, illustrated in Fig. 1(a). We start at point Awith the BEC being in contact with the demagnetiza- tion source, leading to an e ective detuning centered around hidm=0. The potential is of half width a1. The BEC state is given by Eq. 7 with p()=pA()pdm(). Quantum “isothermal” expansion stroke A !B:Dur- ing this stroke, the working medium stays in contact with the external demagnetization source while the potential ex- pands adiabatically from a1toa2without excitation in the BEC. The probability density distribution p() stays con- stant during this ”isothermal” stroke, i.e. we have pA()= pB()=pdm() (e ective detuning centered around hidm= 0). The average work done during this “isothermal” ex- pansion stroke can be then calculated as [38] hWei=R pdm() E ;g(a1;)E ;g(a2;) d. Quantum isochore cooling stroke B !C:The contact with the demagnetization source is switched o and the work- ing medium is brought into contact with the magnetization source while keeping a2constant. The probability distribu- tion p() is changed to pC()pm(), this corresponds to a ”cooling” (as the total energy of the BEC is lowered). The average heat exchange in this stroke can be calculated as hQci=R (pm()pdm())E ;g(;a2)d. Quantum “isothermal” compression stroke C ! D: During this stroke, the working medium stays in contact with the external magnetization source while the BEC com- presses adiabatically from potential half width a2toa1with- out excitation in the BEC. The probability density distribu- tion p() remains constant during this ”isothermal” stroke, i.e. we have pD()=pC()=pm() leading to an ef- fective detuning centered around him>0. The average work done during this “isothermal” compression is hWci=R pm() E ;g(a2;)E ;g(a1;) d. Quantum isochore heating stroke D !A:The contact with the magntetization source is switched o and the work- ing medium is brought again into contact with the demagne- tization source while keeping a1constant. The probability distribution p() is changed back to pA()=pdm(), this corresponds to a ”heating” (as the total energy of the BEC is3 increased). The average heat exchange in this stroke can be calculated ashQhi=R (pdm()pm())E(;a1)d. To study this quantum cycle, it is important to examine and understand the dependence of the BEC ground-state energy on the di erent parameters. This will be done in the following. Perturbation theory for the ground state energy The com- plex BEC system used in the thermodynamic cycle does not have an exact analytical solution. However, we can obtain analytical insight by considering perturbation theory of the ground state energy E ;0(a;) of the non-selfinteracting BEC (i.e. g=0) at small (and nonzero ), as well as at small  (and nonzero ). In the case of small then 1=a, the Hamiltonian in Eq. (1) can be written as H0=H0;0+H0 0whereH0=ˆp2=2+ ˆz=2+V(x) and the perturbation term being H0 0= ˆpˆx. The eigenstate basis of H0;0is given by (0) n;#(x)=0; n(x)T, (0) n;"(x)= n(x);0T, where n(x) are the eigenstates of the potential in Eq.(5). The first-order correction to the energy vanishes and the second-order correction becomes: (0) 2=X n>1jh (0) n;"(x)jH0 0j (0) 0;#(x)ij2 (n21)2=(8a2)+ : (8) Thus, the total ground state energy E ;0(a;) of the system up to second order in is given by E ;0(a;)2 8a2 22 2 4a2+2 2 8a42(a;) cot(a;) 2;(9) where(a;)p 28a2:We can simplify Eq. (9) by approximating the expression up to first order in : E ;0(a;)2 8a2 2 2 2+26 32 a `sr!2 : (10) The first three terms on the right-hand side of Eq. (10) corre- spond to kinetic energy, Zeeman energy (at =0) and SOC energy (at  =0). Here we introduced the spin rotation length `sr1= with a=`sr1. Alternatively, in the case of large then >1=aand small detuning , the Hamiltonian can be written as H0=H0;1+H0 1 whereH0=ˆp2=2+ ˆpˆx+V(x);and the perturbation term H0 1=  ˆz=2:The unperturbedH0;1has pairs of degenerate eigenstates (0) aand (0) bwith the energy E ;0(a;0): (0) a(x)= n(x)ei x 1 1! ; (0) b(x)= n(x)ei x 1 1! :(11) Based on the perturbation theory for degenerate states and tak- ing into account that the diagonal matrix elements of the per- turbation, h (0) ijzj (0) ii=2=0, we obtain at a=`sr1 the ground state energy in the form: E ;0(a;)2 8a2 2 22 4`sr=a (2a=`sr)22 sin 2a `sr! :(12) When we look at the corresponding pressure following from Eq. (12), we can calculate approximately the pressure di er- encePbetween the points BandCin the cycle (at a2, see Fig. 1.56 1.58 1.60 1.62 1.64-0.04-0.020.000.02 1.0 1.2 1.4 1.6 1.8 2.00.00.20.40.60.81.01.2FIG. 2. Pressure P ;0(a2;) versus potential half width afor the cases of = 0; =1:6 (solid black, essentially, -independent),  = 1; =1 (dashed blue) and  = 1; =1:6 (dot-dashed red). (Inset) The pressure di erence between points CandB,P=P ;0(a2;1) P ;0(a2;0). 0.5 1 1.5 21.41.61.82.2.2 FIG. 3. Critical c(g;) versus detuning for di erent nonlineari- ties: attractive g=1 (black solid line), non-interaction g=0 (blue dashed line), and repulsive g=1 (dot-dashed red line). 1). The di erencePjumps from negative to positive at cer- tain widths where 2 a2=`sr(n+1)or (n+1)=(2a2) with n=1;2;3;4;:::. In addition, there is always an between two consecutive “jump points” where Pbecomes zero. We will denote the first corresponding value of ;where the change of Pfor negative to positive occurs, as the critical c(g;). Energy and pressure We examine now the exact numerical values of energy and pressure where we fix a1=1 and a2=2. The corresponding pressure is illustrated in Fig. 2 for a non- interacting BEC ( g=0). The shown pressure P ;0(a2;0) for  =0 does not depend on the strength of SOC as discussed above. We can also see that the pressure P ;0(a2;1) is approx- imately equal to the pressure P ;0(a2;0) at a2;providing crossing of the red and dotted lines; this corresponds then to a critical c(0;1)1:6. The corresponding di erence in pres- surePis shown in detail in the inset; it can be seen that P changes from negative to positive at c(0;1) as one expects it from the perturbation theory above. In Fig. 3, the relations between the critical c(g;) and de- tuning for di erent nonlinearities gare plotted. From the perturbation theory for g=0 and for small , one expects a value of c(g;)=a21:57. The figure shows that the ex-4 act cis increasing with increasing for all cases of g. There is a competition between SOC and Zeeman field, therefore, a larger detuning requires automatically a larger (and there- fore a larger c) to have an e ect. We also see that cis larger (smaller) for attractive g=1 (repulsive g=1) for all . The heuristic reason is that there is a (kind of) compression (ex- pansion) of the wavefunction for g<0 (g>0) and, therefore, a weaker (stronger) e ect of SOC. This requires heuristically a larger (smaller) (and therefore c) to show an e ect. Work, heat and e ciency of the engine Here we are mainly interested in the properties of the cycle originating from the BEC and not in the details of the (de)magnetization source. Therefore, we assume that the probability density distribu- tions pdmresp. pmare strongly peaked around hidm=0 resp.him=  0>0 such that we approximate pdm()=() and pm()=(0) (whereis the Dirac distribution). In this case, the black-solid line and the blue-dashed line in Fig. 2 present an example of the expansion and com- pression strokes of the cycle shown in the schematic Fig. 1. The work done during the “isothermal” expansion pro- cess in Fig. 1,hWei, is then given by the energy di er- ences:hWei=E ;g(a1;0)E ;g(a2;0). The cooling heat exchange from BtoChQcithrough contact with the mag- netization source, becomes hQci=E ;g(a2;0)E ;g(a2;0): The workhWcidone during the compression stroke is then hWci=E ;g(a2;0)E ;g(a1;0). The heat in the last stroke can be calculated by hQhi=E ;g(a1;0)E ;g(a1;0). The total work then becomes: A=hWci+hWei=I ABCDP ;g(a;0)da: (13) For small 0, A=0Z2a2 2a1M ;g(a;0!0)da: (14) As defined above, at = c(g;0), the pressures at a2for  = 0 and 0>0 approximately coincide. If > c(g;0); the pressure-dependencies on afor = 0 and 0>0 cross at a certain half width eawith a1<ea<a2. In that case, the work done at the interval ( ea;a2) provides a negative contri- bution while the contribution of the interval ( a1;ea) can still increase. In the following, we restrict our analysis to the case  c(g;0) while we expect a maximum of the total work close to c(g;0). The e ciency of each quantum cycle is now defined as =A hQhi: (15) At small 1=(2a2) we may approximate the e ciency of the quantum cycle in terms of 0as 2666642 22 00BBBB@1 a2 11 a2 21CCCCA+2 43 0377775 2; (16) where the coe cientis =(a1;0) a4 1cot(a1;0) 2(a2;0) a4 2cot(a2;0) 2:(17) 0.0 0.5 1.0 1.50.00.20.40.60.81.0 0.0 0.5 1.0 1.5 2.00.00.20.40.60.81.0(a) (b) FIG. 4. E ciencyversus with 0=0:5 (solid black), 0= 1:0 (solid blue) and 0=2:0 (solid red); the dotted vertical lines denote the critical SOC strength c(g;0). (a) g=0; results based on perturbation theory in Eq. (16) (blue, red and black dashed lines); the dashed pink line is given by Eq. (18) for 0!0. (b) g=1. In the limit of 0!0, the e ciencysimplifies to =2 26 32 a2 2a2 1 2: (18) It is worth noticing that Eq. (18) has two limits with respect to the value of a2:Let us define cas the e ciency at the critical c. First, Eq. (18) is applicable only at a2<; thus, limiting the criticalcto the values of the order of 0.1. Secondly, for g<0 the value of a2is limited to 2 =jgj[39], thuscis limited correspondingly. (Note that Eq. (18) is not directly applicable tog,0 BEC). Figure 4 shows that the e ciencygrows as increases. The approximate e ciency in Eq. (16) is a quadratic function of , and this is in good agreement with the numerical results in Fig. 4(a) for the case g=0. In the limit of 0!0, the eciency 2, see Eq. (18). This limit case is also shown by the dashed pink line in Fig. 4(a). As one expects a maximum of the total work close to c(g;0), one expects also that the e ciency reaches the maximum at close to c(g;0). The e ciencycat a critical cwith respect to 0 is shown in Fig. 5. The e ciency decreases with increasing . This corresponds to Eq. (16) when = c=a2for all three cases of g(see Fig. (3)). Discussion and conclusions Here we return to the physical units and discuss the possibility of experimental realization of the present Stirling cycle. In the one-dimensional realization5 0.5 1 1.5 20.50.60.70.80.9 FIG. 5. E ciencycat c(g;0) versus 0. Nonlinearities: attrac- tiveg=1 (black solid), non-interaction g=0 (blue dashed), and repulsive g=1 (dot-dashed red). Values of c(g;0) are the same as those in Fig. 3. considered above, with the physical unit of length , the re- sulting dimensionless coupling constant gcan be estimated as 2Naat=sp;where spis the condensate cross-section, physi- cally corresponding to the piston cross-section. Here aatis the interatomic scattering length (typically of the order of 10 aB, where aBis the Bohr radius) dependent on the Feshbach res- onance realization, and N 103is the total number of atoms in the condensate. A reasonable for optical setups is of the order of 10 m. Thus, the choice of a1;a2of the order of 10 m allows one to achieve dimensionless and0of the order of unity [25], and thus explore the operational regimes of the Stirling cycle up to the critical values. In summary, we have explored the potential of a spin-orbit coupled Bose-Einstein condensate in a thermodynamic Stirling-like cycle. It takes advantage of both the non- commuting synthetic spin-orbit and Zeeman-like contribu- tions. The ”cooling” and ”heating” is assumed to originate by interaction with external magnetization and demagnetiza- tion media. We have examined the ground-state energy of the condensate and how the corresponding pressure depends on the di erent parameters of the system. We have studied the eciency of the corresponding engine in the dependence on the strength of these spin-related couplings. The cycle is characterized by a critical spin-orbit coupling, corresponding, essentially, to the maximum e ciency. The dependence of the eciency on the spin-dependent coupling and nonlinear self-interaction paves the way to applications of these cycles. While we have concentrated here on e ects originating from the BEC, it will be interesting to study the details of the e ects of the external magnetization and demagnetization sources in the future. ACKNOWLEDGMENTS We are grateful to C. Whitty and D. Rea for commenting on the manuscript. J.L. and A.R. acknowledge support from the Science Foundation Ireland Frontiers for the Future Re- search Grant “Shortcut-Enhanced Quantum Thermodynam- ics” No.19 /FFP/6951. The work of E.S. is financially sup- ported through the Grant PGC2018-101355-B-I00 funded by MCIN /AEI/10.13039 /501100011033 and by ERDF “A way of making Europe”, and by the Basque Government through Grant No. IT986-16. [1] J. Goold, M. Huber, A. Riera, L. del Rio, and P. Skrzypczyk, Journal of Physics A: Mathematical and Theoretical 49, 143001 (2016). [2] S. De ner and S. Campbell, Quantum Thermodynamics , 2053- 2571 (Morgan & Claypool Publishers, 2019). [3] O. Abah, J. Roßnagel, G. Jacob, S. De ner, F. Schmidt-Kaler, K. Singer, and E. Lutz, Phys. Rev. Lett. 109, 203006 (2012). [4] J. Roßnagel, S. T. Dawkins, K. N. Tolazzi, O. Abah, E. Lutz, F. Schmidt-Kaler, and K. Singer, Science 352, 325 (2016). [5] B. Sothmann, R. S ´anchez, and A. N. Jordan, Nanotechnology 26, 032001 (2014). [6] K. Zhang, F. Bariani, and P. Meystre, Phys. Rev. Lett. 112, 150602 (2014). [7] C. Bergenfeldt, P. Samuelsson, B. Sothmann, C. Flindt, and M. B ¨uttiker, Phys. Rev. Lett. 112, 076803 (2014). [8] C. Elouard, M. Richard, and A. Au `eves, New Journal of Physics 17, 055018 (2015). [9] T. Hugel, N. B. Holland, A. Cattani, L. Moroder, M. Seitz, and H. E. Gaub, Science 296, 1103 (2002). [10] C. J. Pethick and H. Smith, Bose–Einstein Condensation in Dilute Gases , 2nd ed. (Cambridge University Press, 2008) p. 316–347. [11] C. Charalambous, M. A. Garcia-March, M. Mehboudi, and M. Lewenstein, New Journal of Physics 21, 083037 (2019). [12] N. M. Myers, F. J. Pe ˜na, O. Negrete, P. Vargas, G. D. Chiara,and S. De ner, New Journal of Physics 24, 025001 (2022). [13] J. Li, T. Fogarty, S. Campbell, X. Chen, and T. Busch, New Journal of Physics 20, 015005 (2018). [14] F. Schmidt, D. Mayer, Q. Bouton, D. Adam, T. Lausch, J. Net- tersheim, E. Tiemann, and A. Widera, Phys. Rev. Lett. 122, 013401 (2019). [15] Q. Bouton, S. Nettersheim, Jensand Burgardt, D. Adam, E. Lutz, and A. Widera, Nature Communications 12, 2063 (2021). [16] K. Ono, S. N. Shevchenko, T. Mori, S. Moriyama, and F. Nori, Phys. Rev. Lett. 125, 166802 (2020). [17] J.-P. Brantut, C. Grenier, J. Meineke, D. Stadler, S. Krinner, C. Kollath, T. Esslinger, and A. Georges, Science 342, 713 (2013). [18] W. H ¨ubner, G. Lefkidis, C. D. Dong, D. Chaudhuri, L. Chotorl- ishvili, and J. Berakdar, Phys. Rev. B 90, 024401 (2014). [19] J. P. S. Peterson, T. B. Batalh ˜ao, M. Herrera, A. M. Souza, R. S. Sarthour, I. S. Oliveira, and R. M. Serra, Phys. Rev. Lett. 123, 240601 (2019). [20] D. von Lindenfels, O. Gr ¨ab, C. T. Schmiegelow, V . Kaushal, J. Schulz, M. T. Mitchison, J. Goold, F. Schmidt-Kaler, and U. G. Poschinger, Phys. Rev. Lett. 123, 080602 (2019). [21] Y . Zhang, M. E. Mossman, T. Busch, P. Engels, and C. Zhang, Frontiers of Physics 11, 118103 (2016). [22] Y .-J. Lin, K. Jim ´enez-Garc ´ıa, and I. B. Spielman, Nature 471,6 83 (2011). [23] C. Wang, C. Gao, C.-M. Jian, and H. Zhai, Phys. Rev. Lett. 105, 160403 (2010). [24] T.-L. Ho and S. Zhang, Phys. Rev. Lett. 107, 150403 (2011). [25] D. L. Campbell, G. Juzeli ¯unas, and I. B. Spielman, Phys. Rev. A84, 025602 (2011). [26] D. Xiao, M.-C. Chang, and Q. Niu, Rev. Mod. Phys. 82, 1959 (2010). [27] V . Achilleos, D. J. Frantzeskakis, P. G. Kevrekidis, and D. E. Pelinovsky, Phys. Rev. Lett. 110, 264101 (2013). [28] Y . Xu, Y . Zhang, and B. Wu, Phys. Rev. A 87, 013614 (2013). [29] Y . V . Kartashov and V . V . Konotop, Phys. Rev. Lett. 118, 190401 (2017). [30] J. Radi ´c, T. A. Sedrakyan, I. B. Spielman, and V . Galitski, Phys. Rev. A 84, 063604 (2011). [31] S. Sinha, R. Nath, and L. Santos, Phys. Rev. Lett. 107, 270401(2011). [32] J.-Y . Zhang, S.-C. Ji, Z. Chen, L. Zhang, Z.-D. Du, B. Yan, G.- S. Pan, B. Zhao, Y .-J. Deng, H. Zhai, S. Chen, and J.-W. Pan, Phys. Rev. Lett. 109, 115301 (2012). [33] S. V . Manakov, Sov. Phys. JETP 65, 248 (1974). [34] E. Tiesinga, A. J. Moerdijk, B. J. Verhaar, and H. T. C. Stoof, Phys. Rev. A 46, R1167 (1992). [35] P. Courteille, R. S. Freeland, D. J. Heinzen, F. A. van Abeelen, and B. J. Verhaar, Phys. Rev. Lett. 81, 69 (1998). [36] D. S ´anchez and L. Serra, Phys. Rev. B 74, 153313 (2006). [37] I. V . Tokatly and E. Y . Sherman, Annals of Physics 325, 1104 (2010). [38] P. Talkner, E. Lutz, and P. H ¨anggi, Phys. Rev. E 75, 050102 (2007). [39] C. Sulem and P.-L. Sulem, in The Nonlinear Schr¨ odinger Equa- tion: Self-Focusing and Wave Collapse (Applied Mathematical Sciences Book 139) (Springer, 1999).
0810.5405v1.Berry_Phase_in_a_Single_Quantum_Dot_with_Spin_Orbit_Interaction.pdf
arXiv:0810.5405v1 [cond-mat.mes-hall] 30 Oct 2008APS/123-QED Berry Phase in a Single Quantum Dot with Spin-Orbit Interact ion Huan Wang∗, Ka-Di Zhu† Department of Physics, Shanghai Jiao Tong University, Shan ghai 200240, People’s Republic of China (Dated: November 3, 2018) Berry phase in a single quantum dot with Rashba spin-orbit co upling is investigated theoretically. Berry phases as functions of magnetic field strength, dot siz e, spin-orbit coupling and photon-spin coupling constants are evaluated. It is shown that the Berry phase will alter dramatically from 0 to 2πas the magnetic field strength increases. The threshold of ma gnetic field depends on the dot size and the spin-orbit coupling constant. PACS numbers: 71.70.Ej, 03.65.Vf I. INTRODUCTION Due to its important role in encoding information, the phase of wavefunction attracts a lot of interest in infor- mation science. Those properties can also be used in fu- ture quantum information and quantum computer. Thus selectingone kindofphaseswhich canbe manipulated by quantumeffectisveryimportant. Berryphaseisbelieved to be a promising candidate. As a quantum mechanical system evolvescyclically in time such that it return to its initial physical state, its wavefunction can acquire a ge- ometric phase factor in addition to the familiar dynamic phase[1, 2]. If the cyclic change of the system is adia- batic, this additional factor is known as Berry’s phase[3], and is, in contrast to dynamic phase, independent of en- ergy and time. Fuentes-Guridi et al.[4] calculated the Berry phase of a particle in a magnetic field in consid- eration of the quantum nature of the light field. Yi et al.[5] studied the Berry phase in a composite system and showed how the Berry phases depend on the coupling be- tween the two subsystems. In a recent paper, San-Jose et al.[6] have described the effect of geometric phases in- duced by either classical or quantum electric fields acting on single electron spins in quantum dots. Wang and Zhu [7] have investigated the voltage-controlled Berry phases in two vertically coupled InGaAs/GaAs quantum dots. Most recently, observations of Berry phases in solid state materials are reported[8, 9, 10]. Leek et al.[10] demon- strated the controlled Berry phase in a superconducting qubit which manipulates the qubit geometrically using microwave radiation and observes the phase in an inter- ference experiment. Spin-relatedeffectshavepotentialapplicationsinsemi- conductor devices and in quantum computation. Rashba et al.[11] have described the orbital mechanisms of electron-spin manipulation by an electric field. Sonin[12] has demonstrated that an equilibrium spin current in a 2D electron gas with Rashba spin-orbit interaction can result in a mechanical torque on a substrate near an edge of the medium. Serebrennikov[13] considered that the ∗Email: wanghuan2626@sjtu.edu.cn †Email: zhukadi@sjtu.edu.cncoherent transport properties of a charge carrier. The transportation will cause a spin precession in zero mag- netic fields and can be described in purely geometric terms as a consequence of the corresponding holonomy. The spin-orbit interaction in semiconductor heter- structures is increasingly coming to be seen as a tool whichcan manipulate electronicspin states[14, 15]. Two basic mechanisms of the spin-orbit coupling of 2D elec- trons are directly related to the symmetry properties of QDs. They stem from the structure inversion asymme- try mechanism described by the Rashba term[11, 16] and the bulk inversion asymmetry mechanism described by the Dresselhaus term[17]. Recently, Debald and Emary [18] have investigated a spin-orbit driven Rabi oscilla- tion in a single quantum dot with Rashba spin-orbit cou- pling. However, the influence of spin-orbit interaction on Berry phase in a single quantum dot is still lacking. In the present paper we will give a detail study on the Berry phase evolution of a single quantum dot with spin- orbit interaction in a time-dependent quantized electro- magnetic environment. We will borrow quantum optics method to investigate the impact of the spin-orbit inter- action and spin-photon interaction on Berry phase. The paper is organized as follows. In Sec.II, we give the model Hamiltonian including both spin-orbit inter- action and spin-photon interaction and calculated Berry phases as functions of magnetic field strength, dot size, spin-orbit coupling and photon-spin coupling constants. In Sec.III, we draw the figures of the Berry phase as a function of magnetic field strength and some discussions are given. The final conclusion is presented in Sec.IV. II. THEORY We consider a simple two-dimensional quantum dot with parabolic lateral confinement potential in a perpen- dicular magnetic field Bwhich points along zdirection. Then the electron system can be described by the Hamil- tonian [18], Hs=(p+e cA)2 2m∗+m∗ 2ω2 0(x2+y2)+1 2gµBBσz,(1) wherepis the linear momentum operatorof the electron, A(r) =B 2(−y,x,0)isthevectorpotentialinthesymmet-2 ric gauge,ω0is the characteristic confinement frequency, andσ= (σx,σy,σz) is the vector Pauli matrices. m∗ is the effective mass of the electron and gits gyromag- netic factor. µBis the Bohr magneton. In the second quantized notation, Eq.(1) becomes Hs= (a+ xax+a+ yay+1)/planckover2pi1/tildewideω+/planckover2pi1ωc 2i(a+ xay−axa+ y) +1 2gµBBσz,(2) whereωc=eB m∗cand/tildewideω2=ω2 0+ω2 c 4. If we set a+=1√ 2(ax−iay),a−=1√ 2(ax+iay),(3) Then, the Hamiltonian (2) can be written as Hs=n+/planckover2pi1ω++n−/planckover2pi1ω−+1 2gµBBσz, (4) whereω±=/tildewideω±ωc/2,n+=a+ +a+andn−=a+ −a−. In what follows we include the spin-orbit interaction which is described as Rashba Hamiltonian in this system [11] Hso=−α /planckover2pi1[(p+e cA)×σ]z, (5) whereαis the spin-orbit coupling constant which can be controlledbygatevoltageinexperiment. Onsubstituting Eq.(3) into Eq.(5) and then Hso=α /tildewidel[γ+(σ+a++σ−a+ +)−γ−(σ−a−+σ+a+ −)],(6) whereγ±= 1±1 2(/tildewidel/lB)2,/tildewidel= (/planckover2pi1/m∗/tildewideω)1 2andlB= (/planckover2pi1/m∗ωc)1 2. The Hamiltonians of photons and the coupling to the electron spin can be written as follows: Hp=/planckover2pi1ωpb+b, (7) Hp−s=gc(σ++σ−)(b†+b), (8) whereb+(b) andωpare the creation (annihilation) op- erator and energy of the photons, respectively. gcis the spin-photon coupling constant. Hence we obtain the to- tal Hamiltonian of the electron and photons: H=Hs+Hso+Hp+Hp−s =/planckover2pi1ω+a+ +a++/planckover2pi1ω−a+ −a−+1 2gµBBσz +α /tildewidel[γ+(σ+a++σ−a+ +)−γ−(σ−a−+σ+a+ −)] +/planckover2pi1ωpb+b+gc(σ++σ−)(b†+b).(9) Performing a unitary rotation of the spin such that σz→ −σzandσ±→ −σ∓, we arrive at the Hamiltonian H=/planckover2pi1ω+a+ +a++/planckover2pi1ω−a+ −a−−1 2gµBBσz +α /tildewidel[γ−(σ+a−+σ−a+ −)−γ+(σ−a++σ+a+ +)] +/planckover2pi1ωpb+b−gc(σ++σ−)(b†+b).(10)We now derive an approximation form of this Hamilto- nian by borrowing the observation from quantum optics that the terms preceded by γ+in Eq.(10) are counterro- tating, and thus negligible under the rotating-wave ap- proximation when the spin-orbit coupling is small com- pared to the confinement [18]. The last term in Eq.(10) treats in the conventional rotaing-waveapproximation of quantum optics. H=/planckover2pi1ω+a+ +a++/planckover2pi1ω−a+ −a−+1 2|g|µBBσz +λ(σ+a−+σ−a+ −)+/planckover2pi1ωpb+b−gc(σ+b+σ−b+),(11) whereλ=αγ−//tildewidel. Sincegis negative in InGaAs, we choose the absolute value |g|ofg. It is obvious that the ω+mode is decoupled from the rest of the system, giving H=/planckover2pi1ω+n++HJCwhere HJC=/planckover2pi1ω−a+ −a−+1 2|g|µBBσz+/planckover2pi1ωpb+b +λ(σ+a−+σ−a+ −)−gc(σ+b+σ−b+).(12) This is the well-known two mode Jaynes-Cummings model of quantum optics. In general this Hamiltonian can not be solved exactly except ωp=ω−. In what fol- lows, for the sake of analytical simplicity, we consider ωp=ω−which we can use a frequency-controllable laser and a special circuit to satisfy this condition in real ex- periments. In order to solve the above Hamiltonian, we define the normal-mode operators: A=e1a−+e2b, (13) K=e2a−−e1b, (14) where e1=λ/radicalbig λ2+g2c,e2=−gc/radicalbig λ2+g2c,(15) withe2 1+e2 2= 1. The new operators satisfy the commu- tation relations[19] [A,A†] = 1,[NA,A] =−A,[NA,A†] =A†, [K,K†] = 1,[NK,K] =−K,[NK,K†] =K†, [A,K] = 0,[A,K†] = 0,[NA,NK] = 0,(16) whereNA=A†A(NK=K†K) is the number opera- tor related to the normal-mode operator A(K). Intro- ducing the number-sum operator S=NA+NKand the number-difference D=NA−NK, we can verify that the Hamiltonian (12) transforms into the follow- ing Hamiltonian:(i) S=na+nbis a conserved quan- tity (na=a† −a−andnb=b†b); (ii) the operator D can be written in terms of the generators {Q+,Q−,Q}of the SU(2) Lie algebra, D= 2(e2 1−e2 2)Q0+2e1e2(Q++Q−),(17)3 whereQ−=a−b†,Q+=a† −b, andQ0=1 2(a† −a−−b†b), with [Q−,Q+] =−2Q0and [Q0,Q±] =±Q±;(iii) the commutation relation between the operators S and D is null, i.e., [S,D] = 0; and consequently, (iv) the Hamilto- nianHJCsimplifies to HJC=H0+V, where H0=/planckover2pi1ωp(S+1 2σz), V=1 2δσz+λA(σ−A++σ+A),(18) with [H0,V] = 0.λA=/radicalbig λ2+g2cis an effective coupling constantand δ=ωp−|g|µBB//planckover2pi1. The aboveHamiltonian can be solved exactly. The eigenstates of this Hamilto- nian are given by |Ψ(n,±)/an}bracketri}ht=cosθ(n,±)|n,↑/an}bracketri}ht+sinθ(n,±)|n+1,↓/an}bracketri}ht,(19) tanθ(n,±)= (δ±∆n)/2λA/radicalbig (n+1),(20) where ∆ n=/radicalbig δ2+4λ2 A(n+1) and | ↑>(| ↓>) is the spin-up (down) state. According to Ref.[4], since only the quasi-mode Ais coupled with the spin of the electron, so the phase shift operatorU(ϕ) =e−iϕA†Ais introduced. Applied adi- abatically to the Hamiltonian (18), the phase shift op- erator alters the state of the field and gives rise to the following eigenstates: |ψ(n,±)>=e−inϕcosθ(n,±)|n,↑>+ e−i(n+1)ϕsinθ(n,±)|n+1,↓>.(21) Changingϕslowly from 0 to 2 π, the Berry phase is cal- culated as Γ l=i/integraltext2π 0l/an}bracketle{tψ|∂ ∂ϕ|ψ/an}bracketri}htldϕwhich is given by Γl= 2π[sinθ(n,l)]2. (22) This Berry phase is composed of two parts. One is in- duced by spin-orbit interaction, the other is induced by quantized light. Therefore if we can measure the total Berry phase and either part of two Berry phase, we will measure the other part of Berry phase. III. NUMERICAL RESULTS For the illustration of the numerical results, we choose the typical parametersof the InGaAs: g=−4,m∗/me= 0.05 (meis the mass of free electron). The dot size is defined by l0=/radicalbig /planckover2pi1/m∗ω0. Figure 1 depicts the Berry phases Γ +as a function of the magnetic field strength for three spin-orbit couplings. In Figure 1, we can find that all the Berry phases change almost from 0 to 2 πas the magnetic field strength varies from 20 mTto 50mT. When other parametersare fixed, the spin-orbit coupling constantchangesas α= 0.4×10−12eVm, 0.8×10−12eVm and 1.2×10−12eVm, the Berry phases Γ +will have a FIG. 1: The Berry phase Γ +as a function of magnetic field strength Bwith three spin-orbit coupling constants ( α= 0.4×10−12eVm, 0.8×10−12eVmand 1.2×10−12eVm). The other parameters used are g=−4,m∗/me= 0.05,gc= 0.01meV,l0= 80nm, andn= 0. FIG. 2: The Berry phases of Γ +and Γ −as a function of magnetic field strength B. The parameters used are α= 0.4×10−12eVm,g=−4,m∗/me= 0.05,gc= 0.01meV, l0= 80nm, andn= 0. slight movement in the figure. When B <20mTand B >50mT, the Berry phase changes gradually, while when 20mT <B < 50mT, the Berry phase changes dra- matically. As the coupling constant increases, the Berry phase changes from sharply to slowly. The Sh¨ ordinger equationhastwodifferenteigenenergieswhen n= 0. The two eigenenergies will give two different Berry phases. Figure 2 illustrates these two Berry phases. In Figure 2, when the others parameter are fixed, one of the Berry phasechangesfrom0to2 π, whiletheotherchangesfrom 2πto 0 as the magnetic field strength varies from 20 mT to 50mT. Two Berry phases have an intersecting point at approximatively B= 33mT, which is corresponding to the resonant point.4 FIG. 3: The Berry phases Γ +as a function of Bwith three three different dot sizes ( l0= 70nm,80nm,90nm). The pa- rameters used are α= 0.4×10−12eVm,g=−4,m/me= 0.05,gc= 0.01meV, andn= 0. FIG. 4: The Berry phases Γ +as a function of B with three different light coupling constants ( gc= 0.01meV,0.02meV,0.03meV). The parameters used are α= 0.4×10−12eVm,g=−4,m/me= 0.05,l0= 80nm, and n= 0. Figure 3 shows the effect of the dot size on the Berry phase. When dot size becomes large from 70nm to 90nm, although all three Berry phases change from 0 to 2 π, the threshold points of the magnetic field have a large move- ment. When the dot size is 70nm, the Berry phase will change dramatically at approximately 40mT, while the dot sizes are 80nm and 90nm, the turning points are ap- proximately at 30mT and 20mT, respectively. This im- plies that the bigger the dot, the smaller the threshold of the magnetic field strength. Figure 4 illustrates the in- fluence of spin-photon coupling constant on Berry phase. As the coupling constant becomes large, the Berry phase becomes less drastic as shown in Figure 4. In a recent paper, Giuliano et al.[20] have designed an experimental arrangement, which is capacitively coupledthe dot to one arm of a double-path electron interferome- ter. The phase carried by the transported electrons may be influenced by the dot. The dot’s phase gives raise to an interference term in the total conductance across the ring. More recently, Leek et al.[10] have measured Berry phase in a Ramsey fringe interference experiment. Our experimentalsetupproposedhereisanalogouswith these two arrangements as shown in Figure 5. A beam light is split into two beams, one of the beams passes through the dot, and interferes with the other one. Accurate con- trol of the light field for dot is achieved through phase and amplitude modulation of laser radiation coupled to the dot. We choose a special designed electric circuit to ensure the magnetic and laser vary synchronistically. Through detecting the interfered light, we can measure the Berry phase. FIG. 5: A sketch of a possible experimental setup to detect the Berry phase. a, b and c are three mirror, d is a beam splitter. IV. CONCLUSIONS In conclusion, we have theoretically investigated the Berry phase in a single quantum dot in the presence of Rashba spin-orbit interaction. Berry phases as functions of magnetic field strength, dot size, spin-orbit coupling and photon-spin coupling constants are evaluated. It is shown that for a given quantum dot, the spin-orbit coupling constant and photon-spin coupling constant the Berry phase will alter dramatically from 0 to 2 πas the magnetic field strength increases. The threshold of mag- neticfieldisdependent onthe Rashbaspin-orbitcoupling constant, spin-photon coupling constantand the dot size. Wealsoproposeapracticablemethod todetectthe Berry phase in such a quantum dot system. Finally, we hope that our predictions in the present work can be testified by experiments in the near future.5 Acknowledgments This work has been supported in part by National Natural Science Foundation of China (No.10774101) andthe National MinistryofEducationProgramforTraining PhD. [1] A.Shapere and F.Wilczek, Geometric Phases in Physics, (World Scientific, Singapore, 1989) [2] J.Anandan, Nature (London)360 , 307(1992). [3] M.V.Berry, Proc.R.Soc.London Ser. A 392 , 45(1984). [4] I.Fuentes-Guridi, A.Carollo, S.Bose and V.Vedral, Phys.Rev.Lett.89 , 220404(2002). [5] X.X.Yi, L.C.Wang, and T.Y.Zheng, Phys.Rev.Lett.92 , 150406(2004). [6] P.San-Jose, B.Scharfenberger, G.Shon, A.Shnirman, an d G.Zarand, arXiv:cond-mat/0710.3931(2007). [7] H.Wang and K.D.Zhu, Euro.Phys.Lett82 , 60006(2008). [8] Y.Zhang, Y.W.Tan, H.L.Stormer and P.Kim, Nature (London)438 ,201(2005). [9] M.M¨ ott¨ onen, J.J.Vartiainen, and J.P.Pekola, Phys. Rev. Lett.100, 177201(2008). [10] P.J.Leek, J.M.Fink, A.Blais, Science318 ,1889(2007).[11] E.I.Rashba, Sov.Phys.Solid State2 , 1109(1960) [12] E.B.Sonin, Phys. Rev. Lett. 99 ,266602(2007). [13] Y.A.Serebrennikov, Phys.Rev.B73 ,195317(2006). [14] J.C.Egues, G.Burkard, and D.Loss, Phys.Rev.Lett.89 , 176401(2002). [15] E.I.Rashba, and Al.L.Efros, Phys. Rev. Lett.91 , 126405 (2003). [16] Y.A.Bychkov and E.I.Rashba, JEPT Lett.39 , 78(1984). [17] G.Dresselhaus Phys.Rev.B 100 , 580(1955). [18] S.Debald and C.Emary Phys. Rev. Lett.94 ,226803(2005) [19] M.A.Marchiolli, R.J.Missori and J.A.Roversi arXiv: quant-ph/0404.008v1 (2004). [20] D.Giuliano, P.Sodano, and A.Tagliacozzo Phys.Rev.B67 , 155317(2003).
2307.12177v2.Spin_orbit_coupling_induced_phase_separation_in_trapped_Bose_gases.pdf
Spin-orbit-coupling-induced phase separation in trapped Bose gases Zhiqian Gui,1Zhenming Zhang,2Jin Su,3Hao Lyu,4and Yongping Zhang1,∗ 1Department of Physics, Shanghai University, Shanghai 200444, China 2CAS Key Laboratory of Quantum Information, University of Science and Technology of China, Hefei 230026, China 3Department of Basic Medicine, Changzhi Medical College, Changzhi 046000, China 4Quantum Systems Unit, Okinawa Institute of Science and Technology Graduate University, Onna, Okinawa 904-0495, Japan In a trapped spin-1 /2 Bose-Einstein condensate with miscible interactions, a two-dimensional spin- orbit coupling can introduce an unconventional spatial separation between the two components. We reveal the physical mechanism of such a spin-orbit-coupling-induced phase separation. Detailed features of the phase separation are identified in a trapped Bose-Einstein condensate. We further analyze differences of phase separation in Rashba and anisotropic spin-orbit-coupled Bose gases. An adiabatic splitting dynamics is proposed as an application of the phase separation. I. INTRODUCTION Phase separation is a generic phenomenon from clas- sical physics to quantum physics, for example, the oil-water separation and spin Hall effect [1]. Two- component atomic Bose-Einstein condensates (BECs) provide a tunable platform for the investigations of phase separation [2–8]. The two components can be real- ized by using different atomic species or same species with different electric hyperfine states. Such a sys- tem features intra- and inter-component interactions. When the inter-component interactions dominate over the intra-component interactions, two components pre- fer to be phase-separated in order to minimize the inter- component interactions [9, 10]. The interactions for phase separation are called immiscible. The immiscibility of two-component BECs are completely tunable in exper- iments. Phase separation effect induces rich physics in quantum gases, such as the formation of vector solitons and vortex-soliton structures, coherent spin dynamics, and pattern formations [11–19]. In a two-component BEC, an artificial spin-orbit cou- pling can be synthesized between different hyperfine states via Raman lasers [20–23]. Such a Raman-induced spin-orbit coupling is one dimensional. Rashba spin- orbit coupling, which is two dimensional, has also been experimentally realized in BECs [24, 25]. The imple- mentation of spin-orbit coupling in BECs gives rise to exotic quantum phases and rich superfluid properties, which opens an avenue for simulating topological mat- ters and exploring superfluid dynamics [25–32]. In a Raman-type spin-orbit-coupled BEC, a stripe phase [33] can exist in miscible interactions [26, 28]. In contrast, for a Rashba spin-orbit-coupled BEC, the stripe phase may appear in the immiscible regime [34]. Very inter- estingly, Refs. [35, 36] have numerically found a spatially phase-separated ground state in a Rashba-coupled and harmonically trapped BEC with miscible interactions. Such a ground state in the miscible regime is unexpected ∗yongping11@t.shu.edu.cnfor a usual two-component BEC without spin-orbit cou- pling. Reference [35] identifies that the exotic phase sep- aration satisfies a combined symmetry of parity and a spin flip. The existence of this state is attributed by Refs. [36, 37] to a spin-dependent force. The force is in- trinsic in the presence of Rashba spin-orbit coupling and drives the two components moving in opposite directions. The force concept provides an intuitive picture for the un- expected phase separation. However, its weakness is ob- vious. The force is proportional to the square of Rashba spin-orbit coupling strength. Therefore, a large strength is expected to generate a larger spatial separation. In contrast, numerical results show that the separation de- creases with an increasing strength [36, 37]. So far, the physical origin of the unconventional phase separation in the miscible regime is yet to be addressed. It calls for an unambiguous interpretation, since the phase separation has already found a broad application in other excited states. In Refs. [38, 39], a spin-orbit-coupled bright soli- ton is found to be spatially separated in center-of-mass between the two components. Dynamics of the separa- tion in bright solitons is analyzed by varying spin-orbit coupling strength [40]. A spin-orbit-coupled single-vortex state, in which each component carries a singly quan- tized vortex, shows spatial separation between two com- ponents, and the separation is inversely proportional to spin-orbit coupling strength [41]. Recently, dynamics of the separation is triggered by a sudden quench of spin- orbit coupling strength in a trapped BEC [42]. All men- tioned separations occur in the miscible regime, causing a counterintuitive expectation. In this paper, we provide the physical mechanism for the unconventionally spin-orbit-coupling-induced phase separation. Eigenstates of a two-dimensional spin-orbit coupling have a momentum-dependent relative phase φ(⃗k) between the two components. Closely around a fixed momentum ⃗k0, the relative phase may present a linear dependence φ(⃗k)∝(⃗k−⃗k0)·⃗ r0with a constant ⃗ r0. The linear dependence is a momentum kick to move two components relatively. The superposition of these eigenstates distributing around ⃗k0constitutes a spatially separated wave packet. The separation, whose amplitude can be calculated, is a completely single-particle effect ofarXiv:2307.12177v2 [cond-mat.quant-gas] 12 Oct 20232 spin-orbit coupling. A weakly trapped BEC with two- dimensional spin-orbit coupling is a perfect platform to simulate the phase separation. The miscible interactions force atoms to condense at a certain spin-orbit-coupled momentum state with a momentum-dependent relative phase. Meanwhile, weak traps broaden the condensed momentum so that the condensation occupies momen- tum states in a narrow regime, which give rise to the linear dependence of the relative phase. We numerically identify detailed features of spin-orbit-coupling-induced phase separation in a trapped BEC with miscible inter- actions by analyzing its ground states. The phase sepa- ration matches with the single-particle prediction when spin-orbit coupling strength dominates. We also com- pare the separation differences between Rashba and an anisotropic spin-orbit coupling. Finally, as an applica- tion of the phase separation, we propose an adiabatic splitting dynamics. The paper is organized as follows. In Sec. II, the phys- ical mechanism of spin-orbit-coupling-induced phase sep- aration is unveiled. The separation amplitudes are pre- dicted. From the mechanism, we know that the sepa- ration is a single-particle effect. In Sec. III, we identify separated features of ground states in a trapped BEC with Rashba spin-orbit coupling by the imaginary-time evolution method and the variational method. In Sec. IV, we reveal the effect of the anisotropy of spin-orbit cou- pling on the phase separation. In Sec. V, we propose an adiabatic dynamics to dynamically split two components basing on the phase separation. For the completeness of our discussion, immiscible-interaction-induced phase separation is shown in Sec. VI. The conclusion follows in Sec. VII. II. SPIN-ORBIT-COUPLING-INDUCED PHASE SEPARATION Rashba spin-orbit-coupling-induced phase separation is a completely single-particle effect. We reveal the phys- ical origin of such phase separation. The Rashba spin- orbit-coupled Hamiltonian is HSOC=p2 x+p2 y 2+λ(pxσy−pyσx), (1) where pxandpyare the momenta along the xandydirec- tions respectively, λis the spin-orbit coupling strength, andσx,yare spin-1/2 Pauli matrices. The eigenenergy of the Hamiltonian has two bands. The lower band is E=k2 x+k2 y 2−λq k2x+k2y, (2) with associated eigenstates being Φ =1√ 2eikxx+ikyy eiφ 2 e−iφ 2 . (3) Since the Hamiltonian possesses continuously transla- tional symmetry, the eigenstates are plane waves withkx,ybeing the quasimomenta along the xand ydi- rections, respectively. The outstanding feature is that Rashba spin-orbit coupling generates a relative phase φ between the two components, which satisfies tan(φ) =kx ky. (4) It is noted that ( kx, ky) = (0 ,0) is a singularity, closely around which the relative phase cannot be defined. Therefore, the eigenstate in Eq. (3) works beyond the regime around the singularity. We construct a wave packet by superposing these eigenstates, Ψ =Z∞ −∞dkxdkyG(k−¯k)Φ, (5) with the superposition coefficient Gbeing a momentum- dependent localized function centering around ¯k. For a straightforward illustration, we take a Gaussian distribu- tion as an example, G(k−¯k) =1 2πp ∆x∆ye−(kx−¯kx)2 2∆x−(ky−¯ky)2 2∆y.(6) The Gaussian distributed superposition coefficient is cen- tered at ¯k= (¯kx,¯ky) with the packet widthsp ∆x,yalong xandydirections. If the widths are narrow, the superpo- sition mainly happens around ¯k. Therefore, we analyze the eigenstates around ¯k, and the relative phase becomes φ(k)≈φ(¯k) + (kx−¯kx)∂φ ∂kx ¯k+ (ky−¯ky)∂φ ∂ky ¯k,(7) which is linearly dependent on the momenta kx,y. This is true since expanding any continuous function around a certain parameter point leads to dominant linear- dependence. Such linear dependence in Eq. (7) induced a momentum kick, generating the relative motion between the two components. After substituting the Gaussian distribution in Eq. (6) and φin Eq. (7) into Eq. (5) and performing integration, we get the wave packet, Ψ =1√ 2ei¯kxx+i¯kyy × e−∆x 2h x+1 2∂φ(¯k) ∂kxi2−∆y 2h y+1 2∂φ(¯k) ∂kyi2+iφ(¯k) 2 e−∆x 2h x−1 2∂φ(¯k) ∂kxi2−∆y 2h y−1 2∂φ(¯k) ∂kyi2−iφ(¯k) 2 .(8) The outstanding feature of the resultant wave packet is that the two components have a relative position dis- placement. The displacements along the xandydirec- tions are ∂φ(k) ∂kx ¯k=¯ky ¯k2x+¯k2y,∂φ(k) ∂ky ¯k=−¯kx ¯k2x+¯k2y. (9) The nonzero displacements give rise to a phase separation between two components. From the construction of the3 phase-separated wave packets, we can see that the origin of the phase separation is the existence of the momentum- dependent relative phase in eigenstates and the occupa- tion of these eigenstates confined in a narrow momentum regime. Rashba spin-orbit coupled BEC is an ideal platform to generate such a phase-separated state. The lower band in Eq. (2) has infinite energy minima which locate at the quasimomenta satisfying k2 x+k2 y=λ2; therefore, kx=λcos(θ) and ky=λsin(θ) with θbeing an angle. The interacting atoms spontaneously choose one of en- ergy minima to condense and form a BEC [33, 35, 36]. This means that θis spontaneously chosen to be a value ¯θ. In real atomic BEC experiments, traps are inevitable. A weak harmonic trap naturally broadens the BEC mo- mentum giving rise to a Gaussian distribution centered at (¯kx,¯ky) =λ(cos( ¯θ),sin(¯θ)). Furthermore, the broad- ening is narrow so that Eq. (7) is satisfied. Conse- quently, the Rashba-coupled BEC presents as a phase separated state in Eq. (8) with ∂φ(k)/∂kx|¯k= sin( ¯θ)/λ and∂φ(k)|/∂ky|¯k=−cos(¯θ)/λ. The position displace- ment is inversely proportional to the spin-orbit coupling strength λ, which clearly indicates Rashba spin-orbit- coupling-induced phase separation. When the strength goes to zero ( λ≈0), the momentum ( ¯kx,¯ky)≈(0,0) becomes a singularity so that the eigenstate in Eq. (3) is not physical. Without the spin-orbit coupling, the BEC becomes the conventional one, and there is no phase separation between two components. If the strength is enhanced gradually from zero, the position displace- ments should continuously increase from zero to catch up with the predicted value (sin( ¯θ)/λ,−cos(¯θ)/λ). When the strength λis large enough, the position displace- ments decrease towards zero again since they are in- versely proportional to λ. In this case, the plane-wave phase ( ¯kxx+¯kyy) dominates, while the relative phase in the eigenstates [Eq. (3)] is independent of λ, i.e., tan(φ) = ¯kx/¯ky= cot( ¯θ). Consequently, the effect of the relative phase is obliterated by the plane-wave phase, and phase separation disappears. According to the above mechanism of the phase sepa- ration, if there is no weak trap to broaden the condensed momentum, the spin-orbit-coupled BEC can not present the position displacement. This is why a spatially ho- mogeneous BEC with spin-orbit coupling does not show phase separation as studied in most literature. Never- theless, in order to broaden the condensed momentum without traps, we may consider spatially localized exci- tation states, such as bright solitons and vortices. These spatially self-trapped states naturally broadens the con- densed momentum. Therefore, the resultant phase sep- aration between two components in Rashba spin-orbit- coupled bright solitons and quantum vortices, which have been numerically revealed in Refs. [38, 39, 41], can be understood by a generalization of our mechanism. Inter- estingly, the position displacement of quantum vortex is inversely proportional to the spin-orbit coupling strength as uncovered numerically in [41], can be explained unam-biguously. We emphasize that the spin-orbit-coupling-induced phase separation only works for a two-dimensional spin- orbit coupling. For an one-dimensional spin-orbit cou- pling, i.e., the Raman-induced one, the single-particle Hamiltonian is H′=p2 x/2+λpxσz+Ωσxwith Ω being the Rabi frequency due to Raman lasers [20, 23]. The lower energy band of this system is E=k2 x/2−p λ2k2x+ Ω2 with eigenstates being Φ = eikxx(−sin(Θ) ,cos(Θ))T. Here, tan(Θ) = Ω /(λkx), and Tis the transpose oper- ator. It is noted that there is no momentum-dependent relative phase in the eigenstates. Therefore, the Raman- induced spin-orbit coupling, in principle, can not gener- ate the phase separation. In above, we have revealed the physical mecha- nism of Rashba spin-orbit-coupling-induced phase sep- aration. We demonstrate that a weakly trapped spin- orbit-coupled BEC satisfies requirements of the mech- anism. Quantum phase in trapped spin-orbit-coupled BECs may be phase separated states. The spin-orbit- coupling-induced phase separation is a single-particle ef- fect. The role of nonlinearity in the BEC is to sponta- neously choose one energy minimum for condensation. In the following, we study ground states of a trapped spin- orbit-coupled BEC and identify features of spin-orbit- coupling-induced phase separation. III. RASHBA SPIN-ORBIT-COUPLING-INDUCED PHASE SEPARATION IN TRAPPED BECS We consider a quasi-two-dimensional spin-1/2 BEC with Rashba spin-orbit coupling. The trap frequency ωz along the zdirection is assumed to be very large so that the dynamics is completely frozen into the ground state of thez-directional harmonic trap. Such the strong trap can be implemented by an optical lattice in the zdirection in experiments. After integrating the atomic state along zdirection, we are left with a quasi-two-dimensional sys- tem. Rashba spin-orbit coupling can be artificially im- plemented by an optical Raman lattice [24], generating the Hamiltonian HSOC shown in Eq. (1). The spin- orbit-coupled BEC is described by the following Gross- Pitaevskii (GP) equation, i∂Ψ ∂t= (HSOC+V+Hint) Ψ. (10) with Ψ = (Ψ 1,Ψ2)Tbeing the two-component wave function. The harmonic trap in the x−yplane is V=1 2ω2(x2+y2) with ωthe dimensionless trap fre- quency. Hintdenotes nonlinear interactions, Hint= g|Ψ1|2+g12|Ψ2|20 0 g12|Ψ1|2+g|Ψ2|2 ,(11) The GP equation is dimensionless, and the units of length, time, momentum and energy are chosen as lz=4 - 808y / lz( a ) - 202lzky/s47 /s1115( c ) - 8 0 8- 808 x / lz( b ) - 2 0 2- 202 lzkx/s47 /s1115( d ) FIG. 1. Ground state of a trapped Rashba spin-orbit-coupled BEC with miscible interactions. (a) and (b) Density distri- butions |Ψ1|2and|Ψ2|2in the coordinate space. (c) and (d) Density distributions |Ψ1|2and|Ψ2|2in the momentum space. The parameters are ω/ωz= 0.1,lzλ/ℏ= 0.2,g= 12 and g12= 8. p ℏ/(mωz), 1/ωz,ℏ/lzandℏωz, respectively. With the units, the inter-and intra-component interaction coeffi- cients become g=Na√ 8π/lzandg12=Na12√ 8π/lz. Here, Nis the atom number, and aand a12are corresponding s-wave scattering lengths, respectively. The wave functions satisfy the normalization condition,R dxdy(|Ψ1|2+|Ψ2|2) = 1. In numerical calculations, ex- perimentally accessible parameters are used. The typical trap frequency is ωz= 2π×200 Hz, leading to the unites of length and time lz= 0.76µmand 1 /ωz= 0.8msre- spectively. a∼100a0with a0being the Bohr radius, andN∼300, lead to g∼10. The spin-orbit coupling strength can be changed by tuning parameters of Raman lasers in experiments [24]. When g > g 12, the interactions are miscible. We first study ground states of the system in this regime by performing the imaginary-time evolution of the GP equation. The evolution is numerically implemented by the split-step Fourier method. The window of two- dimensional space is chosen as ( x, y)∈[−6π,6π] and is discretized into a 256 ×256 grid. A typical result is shown in Fig. 1. As expected from the prediction in the previous section, the ground state is phase-separated. The two components are spatially separated along the xdirection, as shown by Figs. 1(a) and 1(b). The ground state spontaneously chooses ¯θ= −π/2 so that the atoms condense at ( ¯kx,¯ky) = (0 ,−λ), which can be seen from the momentum-space density dis- tributions in Figs. 1(c) and 1(d). In this case, according to Eq. (9), the position displacement occurs along the x direction, and the first component shifts by 1 /(2λ) on the right side and the second component shifts oppositely by 1/(2λ) on the left side. In the presence of interactions, it is impossible to con- struct analytical wave function of ground state from the procedure demonstrated in the previous section. Never- theless, the single-particle wave functions in Eq. (8) and -101- 2-100 1 2 0.0580.0950 1 2 0.0560.063x/lz(a) lz k'y//s295(b)Δ'x/lzl z /s108/ /s295(c)Component 1C omponent 2Δ 'y/lzl z /s108/ /s295(d)FIG. 2. Rashba spin-orbit-coupling-induced phase sepa- ration in a trapped BEC. The parameters are ω/ωz= 0.1, g= 12 and g12= 8. (a) The center of mass for the two com- ponents along the xdirection as a function of the spin-orbit coupling strength λ. The solid lines are from the variational method, and the circles are obtained by the imaginary-time evolution of the GP equation. The dashed lines are ±1/(2λ) predicted from the single-particle model. The red (blue) color represents the first (second) component. (b) The condensed momentum ¯k′ yas a function of λ. The red solid line and blue circles are obtained by the imaginary-time evolution and the variational method, respectively. The variational parameters ∆′ x,yare shown in (c) and (d). phase-separated results shown in Fig. 1 stimulate us to use a trial wave function to study the phase separation by the variational method [37]. The trial wave function is assumed to be Ψ(x, y) =(∆′ x∆′ y)1 4 √ 2πei¯k′ yy e−∆′ x 2(x−δx)2−∆′ y 2y2 e−∆′ x 2(x+δx)2−∆′ y 2y2 .(12) Here, we have assumed that the atoms spontaneously condenses at (0 ,¯k′ y) in momentum space and therefore the phase separation only happens along the xdirection with the relative position displacement 2 δx. 1/p∆′x,y characterize the widths of the wave packet along the x, y directions. The unknown parameters ¯k′ y, δx,∆′ x,yare to be determined by minimizing the energy functional, E=Z dxdyΨ∗(HSOC+V)Ψ +Z dxdyhg 2(|Ψ1|4+|Ψ2|4) +g12|Ψ1|2|Ψ2|2i , (13) Substituting the trial wave function into the energy func-5 4 6 8 1 0 - 1 . 0 - 0 . 8 0 . 8 1 . 0 0 . 1 0 . 2 0 . 3 - 0 . 7 0 . 0 0 . 7 x / l z g 1 2 ( a ) x / l z /s119 / /s119 z ( b ) FIG. 3. The center of mass of two components for a non- dominant lzλ/ℏ= 0.5. The solid lines are obtained by the variational method and the circles are from the imaginary- time evolution of the GP equation. The red and blue colors represent the first and second components, respectively. (a) The center-of-mass as a function of the inter-component in- teraction coefficient g12.ω/ωz= 0.1 and g= 12. (b) The center-of-mass as a function of the trap frequency ω.g= 12 andg12= 8. tional Eleads to E=¯k′2 y 2+∆′ x+ ∆′ y 4 1 +ω2 ∆′x∆′y +1 2ω2δ2 x +λ(∆′ xδx−¯k′ y)e−∆′ xδ2 x+p∆′x∆′y 8π g+g12e−2∆′ xδ2 x . (14) By minimizing the energy functional with respect to the unknown parameters, ∂E/∂X = 0 ( X=¯k′ y, δx,∆′ x,y), we obtain all information of the trial wave function. The phase separation can be characterized by the center of mass of each component, ¯r1,2=Z r|Ψ1,2(r)|2dr, (15) withr= (x, y). In Fig. 2(a), the solid lines show ¯x1,2=±δxcalculated from the variational method, while the results obtained by the imaginary-time evolution of the GP equation are demonstrated by the circles. We find that the results from the two calculation methods agree very well. Without spin-orbit coupling ( λ= 0), the conventional BEC has ¯ x1,2= 0 and condensates at ¯k′ y= 0, as shown in Fig. 2(b). With the growth of λ, ¯k′ yalways increases linearly [see Fig. 2(b)]. The displace- ment ¯ xfirst increases drastically to a maximum value and then declines to the predicted ±1/(2λ) obtained by the single-particle model [see the dashed lines in Fig. 2(a)]. The dependence of the displacement on λexactly follows the expectation in the previous section. In the dramatic increase regime for ¯ x, the variational parameters ∆′ x,y also change dramatically [see Figs. 2(c) and 2(d)]. Rashba spin-orbit coupling introduces an intrinsic force, F=dp dt=− [r, HSOC], HSOC = 2λ2(p×ez)σz, (16)withezbeing the unit vector along the zdirection and pthe atom momentum. The force originates from spin- orbit-coupling-induced anomalous velocity [43–45]. Con- sidering the ground states shown in Fig. 2, the force op- erator in momentum space becomes Fx= 2λ2¯k′ yσzand Fy= 0. The two components feel opposite force Fxalong thexdirection. Ground states must compensate the in- trinsic force to reach equilibrium. It can be implemented by displacing two component opposite to the force. Since ¯k′ y<0 in the case shown in Fig. 2, the first component is displaced towards to the right side and the second to the left side. The force concept has been used in Refs. [36, 37] to explain the phase separation. Since the force is pro- portional to λ2, it seems that a large displacement would be induced for a large λ. However, as shown in Fig. 2(a), the dependence of the displacement on λdoes not fol- low the force. We can see that the intrinsic force cannot explain the phase separation in the large λregime. Figure 2(a) shows that the separation follows the single-particle prediction ±1/(2λ) when λdominates. When λis weak, the displacement also depends on other parameters, such as nonlinear coefficients and the har- monic trap. In Fig. 3(a), we plot the displacement ¯ xas a function of the inter-component interaction coefficient g12for a non-dominant λ. The displacement slightly rises with an increasing g12, and it reaches the maxi- mum when g12=g. If g12> g, the interactions be- come immiscible, leading to ground states different from the trial wave function in Eq. (12). The dependence of the displacement on the trap frequency is shown in Fig. 3(b). We find that the displacement decreases as the trap frequency increases. This is because the dis- placement requires more kinetic energy in a tight trap. It is notice that there is a slight mismatching between the results from the variational method (the solid line) and the imaginary-time evolution (the circles) in Fig. 3. The origin of such the mismatching is that the Gaus- sian profile in the trial wave function in Eq. (12) cannot exactly describe the imaginary-time-evolution-generated wave function as shown in Fig. 1. IV. THE ANISOTROPIC SPIN-ORBIT-COUPLING-INDUCED PHASE SEPARATION IN TRAPPED BECS Rashba-spin-orbit-coupling-induced phase separation has been analyzed in the previous section. In the two- dimensional spin-orbit-coupled BEC experiment [24], the spin-orbit coupling strengths are tunable, which leads to an anisotropic coupling. It has been revealed that the anisotropic spin-orbit coupling has a great impact on ground states of a spatially homogeneous BEC [46]. In this section, we study anisotropic-spin-orbit-coupling- induced phase separation. The single-particle Hamilto- nian of the anisotropic spin-orbit coupling is H′ SOC=p2 x+p2 y 2+λ1pxσy−λ2pyσx, (17)6 FIG. 4. The anisotropic spin-orbit-coupling-induced phase separation in a trapped BEC. The parameters are ω/ωz= 0.1, g= 12, and g12= 8. (a1) The lower band of H′ SOCin Eq. (17) with lzλ1/ℏ= 0.3 and lzλ2/ℏ= 0.6. (a2) and (a4) show corresponding ground-state density distributions of the first component |Ψ1|2in coordinate and momentum spaces, respectively. (a3) and (a5) are for the second component |Ψ2|2. (b1)-(b5) are same as (a1)-(a5) but with lzλ1/ℏ= 0.6 and lzλ2/ℏ= 0.3. In (a2), (a3), (b2), and (b3), white stars represent the center of wave packets predicted by the single-particle model. with the anisotropic strengths λ1̸=λ2. The lower band ofH′ SOCis E=k2 x+k2 y 2−q λ2 1k2x+λ2 2k2y. (18) with the associated eigenstates being the same as Eq. (3) but having the different relative phase which can be writ- ten as tan(φ) =λ1kx λ2ky. (19) According to the mechanism of the spin-orbit-coupling- induced phase separation, the anisotropic coupling can generate position displacements related to the derivatives of the relative phase. The displacements along the xand ydirections are ∂φ(k) ∂kx ¯k=λ1λ2¯ky λ2 1¯k2x+λ2 2¯k2y, ∂φ(k) ∂ky ¯k=−λ1λ2¯kx λ2 1¯k2x+λ2 2¯k2y. (20)Here, ¯k= (¯kx,¯ky) is the momentum at which the atoms condense. The lowest energy minima of the lower band depend on the anisotropy. When λ1< λ 2, the two minima locate at ( ¯kx,¯ky) = (0 ,±λ2) [see Fig. 4(a1)]. They locate at ( ¯kx,¯ky) = (±λ1,0) when λ1> λ 2[see Fig. 4(b1)]. With the miscible interactions, the BEC spontaneously chooses one of these two minima to con- dense. The ground state that spontaneously condenses at (¯kx,¯ky) = (0 ,−λ2) for λ1< λ 2is demonstrated in Figs. 4(a2)-(a5). We obtain ground states by the imaginary-time evolution of the GP equation with H′ SOC. From the single-particle prediction in Eq. (20), the phase separation of this ground state happens only along the xdirection, and the center-of-mass of the first compo- nent is λ1/(2λ2 2) and that of the second component is −λ1/(2λ2 2) [see the white stars in Figs. 4(a2) and 4(a3)]. Density distributions shown in Figs. 4(a2) and 4(a3) clearly indicate the phase separation following the pre- dictions. The ground state that spontaneously condenses at (¯kx,¯ky) = ( −λ1,0) for λ1> λ 2is demonstrated in Figs. 4(b2)-(b5). The single-particle mechanism in7 0.00 .51 .01 .52 .0-0.40.00.4l z /s1082//s295x/lz- 0.40.00.4y /lz FIG. 5. Anisotropic-spin-orbit-coupling-induced phase sepa- ration as a function of λ2with a fixed lzλ1/ℏ= 1. Circles are for the first component and crosses are for the second compo- nent. The blue (red) color represents separation along the x (y) direction. Other parameters are ω/ωz= 0.1,g= 12 and g12= 8. Eq. (20) predicts that for this ground state the separa- tion happens along the ydirection and the center-of-mass are∓λ2/(2λ2 1) for two components [see the white stars in Figs. 4(b2) and 4(b3)]. The results from the imaginary- time evolution shown in Figs. 4(b2) and 4(b3) match with the single-particle predictions. These analyses have shown that the center of mass of each component strongly depends on the ratio of the spin-orbit coupling strengths. To reveal the depen- dence of phase separation on λ2/λ1, we calculate ground states with a fixed λ1and a changeable λ2by using the imaginary-time evolution. The results are summarized in Fig. 5, where the circles (crosses) represent the center of mass for the first (second) component. For λ2< λ1= 1, the phase separation occurs along the ydirection and |¯y|increases with the increase of λ2[see red circles and crosses in Fig. 5], while ¯ xis zero [see blue circles and crosses in Fig. 5]. When λ2= 0, the spin-orbit coupling becomes one-dimensional, there is no phase separation due to the absence of the relative phase. The results change for λ2> λ 1= 1 and the phase separation along thexdirection is observed. In this case, the separation decreases with λ2increasing. For a very large λ2, the separation disappears since the spin-orbit coupling effec- tively turns to be one-dimensional. The results in Fig. 5 demonstrate that the maximum separation happens for λ1=λ2which is Rashba spin-orbit coupling. This is also expected from the single-particle prediction in Eq. (20). V. ADIABATIC SPLITTING DYNAMICS We have shown that ground states of a trapped BEC with two-dimensional spin-orbit coupling and miscible in-teractions are phase-separated. As an important appli- cation, we study adiabatic dynamics of the phase separa- tion. As pointed out by previous works, a linear coupling between two component favors miscibility regardless of interactions [47, 48]. Therefore, a miscible-to-immiscible transition may occur by decreasing the coupling. The adiabatic dynamics is stimulated by slowly switching off the linear coupling. Theoretically, the process is de- scribed by the time-dependent GP equation, i∂Ψ ∂t= [HSOC+ Ω(t)σx+V+Hint] Ψ. (21) Here, Ω( t)σxrepresents the linear coupling between the two components, and can be experimentally achieved by using a radio-frequency coupling [6]. The time-dependent Rabi frequency is Ω(t) = Ω 0(1−t/τq), (22) with Ω 0being the initial value of the linear coupling and τqis the quench duration. At t= 0, the presence of Ω 0 greatly suppresses the ground-state phase separation. We obtain ground state by the imaginary-time evolution of Eq. (21) with Ω( t) = Ω 0. A typical ground state is shown in insets (a) and (b) of Fig. 6, and the separation be- tween two components is not obvious. Using this ground state as initial state, we evolve the time-dependent GP equation. The center-of-mass ¯ xfor two components is recorded during the time evolution in Fig. 6. By de- creasing the linear coupling adiabatically, the separation between two component gradually increases. When it is completely switched off, i.e., t=τq, the separation 04 08 01 20-0.6-0.30.00.30.6x/lz/s119 z t -808-808x /lzy/lz -808-808x /lzy/lz -808-808x /lzy/lz -808-808x /lzy/lz( a ) ( b ) ( c ) ( d ) FIG. 6. Adiabatic splitting dynamics of a trapped BEC with Rashba spin-orbit coupling by slowly switching off the linear coupling. The parameters are ω/ωz= 0.1,g= 12, g12= 8, Ω0/ωz= 3, and ωzτq= 150. The red (blue) dots represent the center-of-mass of the first (second) component. Insets (a,b) [(c,d)] are density distributions of the first and second components at t= 0 [t=τq], respectively.8 FIG. 7. Immiscible-interaction-induced phase-separated ground states in a trapped BEC with Rashba spin-orbit coupling. The parameters are g= 4, g12= 8 and ω/ωz= 0.1. (a1)-(a4) When lzλ/ℏ= 0.5 the ground state is a half-quantum vortex state. (a1) [a(3)] and (a2) [(a4)] are the coordinate (momentum) space density distributions of the first and second components respectively. (b1)-(b4) When lzλ/ℏ= 1.5 the ground state is a stripe state. (b1) [b(3)] and (b2) [(b4)] are the coordinate [momentum] space density distributions of the first and second components respectively. is maximized [see corresponding density distributions in insets (c) and (d)]. The two components can realize a dynamically spatial splitting, which move along opposite directions. Such adiabatic splitting dynamics are remi- niscent of a kind of “atomic spin Hall effect” [49]. VI. IMMISCIBLE INTERACTIONS INDUCED PHASE SEPARATION In all above, the interactions are miscible ( g > g 12), which support atoms to condense at a particular momen- tum state. On the other hand, immiscible interactions (g < g 12) prefer a spatial separation between two com- ponents in order to minimize the inter-component inter- actions proportional to g12. In the presence of spin-orbit coupling, the immiscible-interaction-induced phase sepa- ration presents interesting features [35, 36, 50]. In Fig. 7, we show two different kinds of immiscible-interaction- induced phase separated ground states with different val- ues of spin-orbit coupling strength λ. For λ= 0.5, the ground state obtained by the imaginary-time evolution is a half-quantum vortex state which has been first re- vealed in Refs. [35, 50]. The first component distribution has a Gaussian shape [see Figs. 7(a1) and 7(a3)], and the second component is a vortex with a winding num- berw= 1 [see Figs. 7(a2) and 7(a4)]. The first compo- nent is filled in the density dip of the second one, form- ing a spatial separation along the radial direction. For λ= 1.5, the ground state becomes a stripe state which has been first revealed in Refs. [33, 34]. The ground state condenses simultaneously at two different momenta [see Figs. 7(b3) and 7(b4)]. Such momentum occupation gen- erates spatially periodic modulations in density distribu- tions [Figs. 7(b1) and 7(b2)]. Meanwhile, stripes of the two components are spatially separated.We emphasize that phase separations induced by spin- orbit coupling and immiscible interaction have different physical origins. The spin-orbit-coupling-induced phase separation only works for a two-dimensional spin-orbit coupling. However, phase separations have been also studied for a BEC with a one-dimensional spin-orbit cou- pling, the mechanism of which is different. In the pi- oneered spin-orbit-coupled experiment, experimentalists observed a spatial separation between two dressed states with a Raman-induced spin-orbit coupling [20]. The spin-orbit coupling generates two energy minima, whose occupations can be considered as two dressed states. In the dressed state space, atomic interactions turn to be immiscible between two dressed states in the presence of the Rabi frequency. The phase separation happens in the dressed state space due to immiscibility. In ad- dition, Ref. [47] reveals the existence of phase separa- tion in a spin-1 BEC with the Raman-induced spin-orbit coupling. The single-particle Hamiltonian of the system isH= (px+λ′Fz)2/2 + Ω′Fx+ϵF2 z. Here, Fx,y,z are the spin-1 Pauli matrices, λ′is the spin-orbit coupling strength, Ω′is the Rabi frequency, and ϵis the quadratic Zeeman shift. The spinor interactions include density- density part with the coefficient c0and spin-spin part with the coefficient c2. In particular, a very negative quadratic Zeeman shift ϵ=−λ2/2 was considered. With such a large negative ϵ, the occupation in the second component can be eliminated. The spinor only occu- pies the first and third components. Interestingly, the spinor interactions between the first and third compo- nents are immiscible for a negative spin-spin interaction (c2<0). Different phase-separated states between the first and third component are due to immiscible interac- tions [47] .9 VII. CONCLUSION In summary, we have revealed the physical mechanism of spin-orbit-coupling-induced phase separation. The mechanism, which is very different from the conventional immiscible-interaction-induced separation, is a complete single-particle effect of spin-orbit coupling. We have ana- lyzed separation features in a trapped BEC with Rashba spin-orbit coupling and miscible interactions and studied the effects of the anisotropy of spin-orbit coupling on theseparation. All features can be explained by the single- particle mechanism. As an interesting application of the phase separation, we propose an adiabatic dynamics that can dynamically split two components spatially. ACKNOWLEDGMENTS This work was supported by National Natural Sci- ence Foundation of China with Grants No.12374247 and 11974235. H.L. acknowledges support from Okinawa In- stitute of Science and Technology Graduate University. [1] J. Sinova, S. O. Valenzuela, J. Wunderlich, C. H. Back, and T. Jungwirth, Spin Hall effects, Rev. Mod. Phys. 87, 1213 (2015). [2] Y. Shin, M. W. Zwierlein, C. H. Schunck, A. Schirotzek, and W. Ketterle, Observation of Phase Separation in a Strongly Interacting Imbalanced Fermi Gas, Phys. Rev. Lett.97, 030401 (2006). [3] S. B. Papp, J. M. Pino, and C. E. Wieman, Tunable Miscibility in a Dual-Species Bose-Einstein Condensate, Phys. Rev. Lett. 101, 040402 (2008). [4] G. Thalhammer, G. Barontini, L. De Sarlo, J. Catani, F. Minardi, and M. Inguscio, Double Species Bose- Einstein Condensate with Tunable Interspecies Interac- tions, Phys. Rev. Lett. 100, 210402 (2008). [5] S. Tojo, Y. Taguchi, Y. Masuyama, T. Hayashi, H. Saito, and T. Hirano, Controlling phase separation of binary Bose-Einstein condensates via mixed-spin-channel Fesh- bach resonance, Phys. Rev. A 82, 033609 (2010). [6] E. Nicklas, H. Strobel, T. Zibold, C. Gross, B. A. Mal- omed, P. G. Kevrekidis, and M. K. Oberthaler, Rabi Flopping Induces Spatial Demixing Dynamics, Phys. Rev. Lett. 107, 193001 (2011). [7] K. Jim´ enez-Garc´ ıa, A. Invernizzi, B. Evrard, C. Frapolli, J. Dalibard, and F. Gerbier, Spontaneous formation and relaxation of spin domains in antiferromagnetic spin-1 condensates, Nat. Commun. 10, 1422 (2019). [8] L. He, P. Gao, and Z.-Q. Yu, Normal-Superfluid Phase Separation in Spin-Half Bosons at Finite Temperature, Phys. Rev. Lett. 125, 055301 (2020). [9] M. Trippenbach, K. G´ oral, K. Rzazewski, B. Malomed, and Y. B. Band, Structure of binary Bose-Einstein con- densates, J. Phys. B 33, 4017 (2000). [10] D. S. Petrov, Quantum Mechanical Stabilization of a Col- lapsing Bose-Bose Mixture, Phys. Rev. Lett. 115, 155302 (2015). [11] U. Shrestha, J. Javanainen, and J. Ruostekoski, Pulsat- ing and Persistent Vector Solitons in a Bose-Einstein Condensate in a Lattice upon Phase Separation Insta- bility, Phys. Rev. Lett. 103, 190401 (2009). [12] K. J. H. Law, P. G. Kevrekidis, and L. S. Tuckerman, Sta- ble Vortex–Bright–Soliton structures in Two-Component Bose-Einstein Condensates, Phys. Rev. Lett. 105, 160405 (2010). [13] K. L. Lee, N. B. Jørgensen, I.-K. Liu, L. Wacker, J. J. Arlt, and N. P. Proukakis, Phase separation and dynam- ics of two-component Bose-Einstein condensates, Phys.Rev. A 94, 013602 (2016). [14] A. Roy, M. Ota, A. Recati, and F. Dalfovo, Finite- temperature spin dynamics of a two-dimensional Bose- Bose atomic mixture, Phys. Rev. Res. 3, 013161 (2021). [15] Y. Eto, M. Takahashi, M. Kunimi, H. Saito, and T. Hirano, Nonequilibrium dynamics induced by misci- ble–immiscible transition in binary Bose–Einstein con- densates, New J. Phys. 18, 073029 (2016). [16] N. R. Bernier, E. G. Dalla Torre, and E. Demler, Un- stable Avoided Crossing in Coupled Spinor Condensates, Phys. Rev. Lett. 113, 065303 (2014). [17] W. E. Shirley, B. M. Anderson, C. W. Clark, and R. M. Wilson, Half-Quantum Vortex Molecules in a Binary Dipolar Bose Gas, Phys. Rev. Lett. 113, 165301 (2014). [18] S. I. Mistakidis, G. C. Katsimiga, P. G. Kevrekidis, and P. Schmelcher, Correlation effects in the quench-induced phase separation dynamics of a two species ultracold quantum gas, New J. Phys. 20, 043052 (2018). [19] R. S. Lous, I. Fritsche, M. Jag, F. Lehmann, E. Kir- ilov, B. Huang, and R. Grimm, Probing the Interface of a Phase-Separated State in a Repulsive Bose-Fermi Mix- ture, Phys. Rev. Lett. 120, 243403 (2018). [20] Y.-J. Lin, K. Jim´ enez-Garc´ ıa, and I. B. Spielman, Spin– orbit-coupled Bose–Einstein condensates, Nature (Lon- don)471, 83 (2011). [21] N. Goldman, G. Juzeli¯ unas, P. ¨Ohberg, and I. B. Spiel- man, Light-induced gauge fields for ultracold atoms, Rep. Prog. Phys. 77, 126401 (2014). [22] H. Zhai, Degenerate quantum gases with spin–orbit cou- pling: a review, Rep. Prog. Phys. 78, 026001 (2015). [23] Y. Zhang, M. E. Mossman, T. Busch, P. Engels, and C. Zhang, Properties of spin–orbit-coupled Bose–Einstein condensates, Front. Phys. 11, 118103 (2016). [24] Z. Wu, L. Zhang, W. Sun, X.-T. Xu, B.-Z. Wang, S.-C. Ji, Y. Deng, S. Chen, X.-J. Liu, and J.-W. Pan, Realization of two-dimensional spin-orbit coupling for Bose-Einstein condensates, Science 354, 83 (2016). [25] A. Vald´ es-Curiel, D. Trypogeorgos, Q.-Y. Liang, R. P. Anderson, and I. B. Spielman, Topological features with- out a lattice in Rashba spin-orbit coupled atoms, Nat. Commun. 12, 593 (2021). [26] Y. Li, L. P. Pitaevskii, and S. Stringari, Quantum Tri- criticality and Phase Transitions in Spin-Orbit Coupled Bose-Einstein Condensates, Phys. Rev. Lett. 108, 225301 (2012).10 [27] A. Manchon, H. C. Koo, J. Nitta, S. M. Frolov, and R. A. Duine, New perspectives for Rashba spin–orbit coupling, Nat. Mater. 14, 871 (2015). [28] J.-R. Li, J. Lee, W. Huang, S. Burchesky, B. Shteynas, F. C ¸. Top, A. O. Jamison, and W. Ketterle, A stripe phase with supersolid properties in spin–orbit-coupled Bose–Einstein condensates, Nature (London) 543, 91 (2017). [29] M. A. Khamehchi, K. Hossain, M. E. Mossman, Y. Zhang, T. Busch, M. M. Forbes, and P. Engels, Negative-Mass Hydrodynamics in a Spin-Orbit–Coupled Bose-Einstein Condensate, Phys. Rev. Lett. 118, 155301 (2017). [30] Y. V. Kartashov and D. A. Zezyulin, Stable Multiring and Rotating Solitons in Two-Dimensional Spin-Orbit- Coupled Bose-Einstein Condensates with a Radially Pe- riodic Potential, Phys. Rev. Lett. 122, 123201 (2019). [31] M. Hasan, C. S. Madasu, K. D. Rathod, C. C. Kwong, C. Miniatura, F. Chevy, and D. Wilkowski, Wave Packet Dynamics in Synthetic Non–Abelian Gauge Fields, Phys. Rev. Lett. 129, 130402 (2022). [32] A. Fr¨ olian, C. S. Chisholm, E. Neri, C. R. Cabrera, R. Ramos, A. Celi, and L. Tarruell, Realizing a 1D topo- logical gauge theory in an optically dressed BEC, Nature (London) 608, 293 (2022). [33] C. Wang, C. Gao, C.-M. Jian, and H. Zhai, Spin-Orbit Coupled Spinor Bose-Einstein Condensates, Phys. Rev. Lett.105, 160403 (2010). [34] T.-L. Ho and S. Zhang, Bose-Einstein Condensates with Spin-Orbit interaction, Phys. Rev. Lett. 107, 150403 (2011). [35] H. Hu, B. Ramachandhran, H. Pu, and X.-J. Liu, Spin- Orbit Coupled Weakly Interacting Bose-Einstein Con- densates in Harmonic Traps, Phys. Rev. Lett. 108, 010402 (2012). [36] Y. Zhang, L. Mao, and C. Zhang, Mean-Field Dynamics of Spin-Orbit Coupled Bose-Einstein Condensates, Phys. Rev. Lett. 108, 035302 (2012). [37] S.-W. Song, Y.-C. Zhang, L. Wen, and H. Wang, Spin–orbit coupling induced displacement and hidden spin textures in spin-1 Bose–Einstein condensates, J. Phys. B 46, 145304 (2013). [38] H. Sakaguchi, B. Li, and B. A. Malomed, Creation of two-dimensional composite solitons in spin-orbit-coupled self-attractive Bose-Einstein condensates in free space,Phys. Rev. E 89, 032920 (2014). [39] Y. Xu, Y. Zhang, and C. Zhang, Bright solitons in a two-dimensional spin-orbit-coupled dipolar Bose- Einstein condensate, Phys. Rev. A 92, 013633 (2015). [40] Y.-J. Wang, L. Wen, G.-P. Chen, S.-G. Zhang, and X.- F. Zhang, Formation, stability, and dynamics of vec- tor bright solitons in a trapless Bose–Einstein conden- sate with spin–orbit coupling, New J. Phys. 22, 033006 (2020). [41] M. Kato, X.-F. Zhang, and H. Saito, Vortex pairs in a spin-orbit-coupled Bose-Einstein condensate, Phys. Rev. A95, 043605 (2017). [42] R. Ravisankar, H. Fabrelli, A. Gammal, P. Muruganan- dam, and P. K. Mishra, Effect of Rashba spin-orbit and Rabi couplings on the excitation spectrum of binary Bose-Einstein condensates, Phys. Rev. A 104, 053315 (2021). [43] S. Mardonov, M. Modugno, and E. Y. Sherman, Dynam- ics of spin-orbit coupled bose-einstein condensates in a random potential, Phys. Rev. Lett. 115, 180402 (2015). [44] S. Mardonov, E. Y. Sherman, J. G. Muga, H.-W. Wang, Y. Ban, and X. Chen, Collapse of spin-orbit-coupled bose-einstein condensates, Phys. Rev. A 91, 043604 (2015). [45] S. Mardonov, V. V. Konotop, B. A. Malomed, M. Mod- ugno, and E. Y. Sherman, Spin-orbit-coupled soliton in a random potential, Phys. Rev. A 98, 023604 (2018). [46] T. Ozawa and G. Baym, Ground-state phases of ultra- cold bosons with Rashba-Dresselhaus spin-orbit coupling, Phys. Rev. A 85, 013612 (2012). [47] S. Gautam and S. K. Adhikari, Phase separation in a spin-orbit-coupled Bose-Einstein condensate, Phys. Rev. A90, 043619 (2014). [48] I. M. Merhasin, B. A. Malomed, and R. Driben, Transi- tion to miscibility in a binary Bose–Einstein condensate induced by linear coupling, J. Phys. B 38, 877 (2005). [49] M. C. Beeler, R. A. Williams, K. Jim´ enez-Garc´ ıa, L. J. LeBlanc, A. R. Perry, and I. B. Spielman, The spin Hall effect in a quantum gas, Nature (London) 498, 201 (2013). [50] B. Ramachandhran, B. Opanchuk, X.-J. Liu, H. Pu, P. D. Drummond, and H. Hu, Half-quantum vortex state in a spin-orbit-coupled Bose-Einstein condensate, Phys. Rev. A85, 023606 (2012).
1601.06935v2.Double_Quantum_Spin_Vortices_in_SU_3__Spin_Orbit_Coupled_Bose_Gases.pdf
arXiv:1601.06935v2 [cond-mat.quant-gas] 24 Sep 2016Double-quantum spin vortices in SU(3) spin-orbit coupled B ose gases Wei Han,1,2Xiao-Fei Zhang,1Shu-Wei Song,3Hiroki Saito,4Wei Zhang∗,5Wu-Ming Liu†,2and Shou-Gang Zhang‡1 1Key Laboratory of Time and Frequency Primary Standards, National Time Service Center, Chinese Academy of Sciences, Xi’an 710600, China 2Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China 3State Key Laboratory Breeding Base of Dielectrics Engineer ing, Harbin University of Science and Technology, Harbin 150080 , China 4Department of Engineering Science, University of Electro- Communications, Tokyo 182-8585, Japan 5Department of Physics, Renmin University of China, Beijing 100872, China We show that double-quantum spin vortices, which are charac terized by doubly quantized circu- lating spin currents and unmagnetized filled cores, can exis t in the ground states of SU(3) spin-orbit coupled Bose gases. It is found that the SU(3) spin-orbit cou pling and spin-exchange interaction play important roles in determining the ground-state phase diagram. In the case of effective fer- romagnetic spin interaction, the SU(3) spin-orbit couplin g induces a three-fold degeneracy to the magnetized ground state, while in the antiferromagnetic sp in interaction case, the SU(3) spin-orbit coupling breaks the ordinary phase rule of spinor Bose gases , and allows the spontaneous emergence of double-quantum spin vortices. This exotic topological d efect is in stark contrast to the singly quantized spin vortices observed in existing experiments, and can be readily observed by the current magnetization-sensitive phase-contrast imaging techniq ue. PACS numbers: 03.75.Lm, 03.75.Mn, 67.85.Bc, 67.85.Fg I. INTRODUCTION The recent experimental realization of synthetic spin- orbit (SO) coupling in ultracold quantum gases [1–10] is considered as an important breakthrough, as it provides new possibilities for ultracold quantum gases to be used as quantum simulation platforms, and paves a new route towards exploring novel states of matter and quantum phenomena [11–19]. It has been found that the SO cou- pling can not only stabilize various topological defects, such as half-quantum vortex, skyrmion, composite soli- ton and chiral domain wall, contributing to the design and exploration of new functional materials [20–23], but alsoleadtoentirelynew quantum phases, suchasmagne- tized phase and stripe phase [24–26], providing support for the study of novel quantum dynamical phase transi- tions [27, 28] and exotic supersolid phases [29–31]. All the intriguing features mentioned above are based on the characteristics that the SO coupling (either of the NIST[1], Rashba[24]orWeyl[32]types)makestheinter- nal states coupled to their momenta via the SU(2) Pauli matrices. However, if the (pseudo)spin degree of freedom involves more than two states, the SU(2) spin matrices cannot describe completely all the couplings among the internal states. For example, a direct transition between the states |1∝angbracketrightand|−1∝angbracketrightis missing in a three-component system [24, 33]. From this sense, an SU(3) SO coupling with the spin operator spanned by the Gell-Mann matri- ∗wzhangl@ruc.edu.cn †wliu@iphy.ac.cn ‡szhang@ntsc.ac.cnces is more effective in describing the internal couplings among three-component atoms [33, 34]. The SU(3) SO coupled system has no analogue in ordinary condensed matter systems, hence may lead to new quantum phases and topological defects. In this article, we show that a new type of topological defects, double-quantum spin vortices, can exist in the ground states of SU(3) SO coupled Bose-Einstein con- densates (BECs). It is found that the SU(3) SO coupling leads to two distinct ground-state phases, a magnetized phase or a lattice phase, depending on the spin-exchange interaction being ferromagnetic or antiferromagnetic. In the magnetized phase, the SU(3) SO coupling leads to a groundstatewiththree-folddegeneracy,in starkcontrast to the SU(2) case where the degeneracy is two, thus may offer new insights into quantum dynamical phase transi- tions [27]. In the lattice phase, the SU(3) SO coupling breaks the ordinary phase requirement 2w 0= w1+w−1 for ordinary spinor BECs, where w iis the winding num- ber of thei-th spin component [35–37], and induces three types of exotic vortices with cores filled by different mag- netizations. The interlaced arrangementof these vortices leads to the spontaneous formation of multiply quantized spin vortices with winding number 2. This new type of topological defects can be observed in experiments us- ing magnetization-sensitive phase-contrast imaging tech- nique. II. SU(3) SPIN-ORBIT COUPLING We consider the F= 1 spinor BECs with SU(3) SO coupling. Using the mean-field approximation, the Hamiltonian can be written in the Gross-Pitaevskii form2 as H=/integraldisplay dr/bracketleftbigg Ψ†/parenleftbigg −/planckover2pi12∇2 2m+Vso/parenrightbigg Ψ+c0 2n2+c2 2|F|2/bracketrightbigg ,(1) where the order parameter Ψ= [Ψ1(r),Ψ0(r),Ψ−1(r)]⊤ is normalized with the total particle num- berN=/integraltextdrΨ†Ψ. The particle density is n=/summationtext m=1,0,−1Ψ∗ m(r)Ψm(r), and the spin density vectorF= (Fx,Fy,Fz) is defined by Fν(r) =Ψ†fνΨ withf= (fx,fy,fz) being the vector of the spin-1 ma- trices given in the irreducible representation [35, 38–40]. TheSOcouplingtermischosenas Vso=κλ·p, whereκis the spin-orbit coupling strength, p= (px,py) represents 2D momentum, and λ= (λx,λy) is expressed in terms ofλx=λ(1)+λ(4)+λ(6)andλy=λ(2)−λ(5)+λ(7), withλ(i)(i= 1,...8) being the Gell-Mann matrices, i.e., the generators of the SU(3) group [41]. Note that the SU(3) SO coupling term in the Hamiltonian involves all the pairwise couplings between the three states. This is distinct from the previously discussed SU(2) SO coupling in spinor BECs, where the states Ψ 1(r) and Ψ −1(r) are coupled indirectly [24, 42, 43]. The parameters c0 andc2describe the strengths of density-density and spin-exchange interactions, respectively. The Hamiltonian with SU(3) SO coupling can be re- alized using a similar method of Raman dressing as in the SU(2) case [1, 9, 44]. As shown in Fig. 1(a), three laser beams with different polarizations and frequencies, intersecting at an angle of 2 π/3, are used for the Ra- man coupling. Each of the three Raman lasers dresses one hyperfine spin state from the F= 1 manifold ( |F= 1,mF= 1∝angbracketright,|F= 1,mF= 0∝angbracketrightand|F= 1,mF=−1∝angbracketright) to the excited state |e∝angbracketright[See Fig. 1(b)]. When the standard rotatingwaveapproximationisusedandtheexcitedstate is adiabatically eliminated due to far detuning, one can obtain the effective Hamiltonian in Eq. (1), as discussed in Appendix A. By diagonalizing the kinetic energy and SO coupling terms, we can obtain the single-particleenergyspectrum, which can provide useful information about the ground stateofBosecondensates. FortheSU(2)case,itisknown that the single-particle spectrum with the NIST type SO coupling acquires either a single or two minima, depend- ing on the strength of the Raman coupling [1], while for the case of Rashba type there exist an infinite number of minima locating on a continuous ring in momentum space [45]. For the SU(3) SO coupling discussed here, we find that there are in general three discrete minima re- siding on the vertices of an equilateral triangle [See Figs. 1(c)-1(d)]. This unique property of the energy band im- plies the possibilityofa three-folddegeneratemany-body magnetized state [27] or a topologically nontrivial lattice state, depending on the choices among the three minima made by the many-body interactions. FIG. 1: (Color online) Scheme for creating SU(3) spin-orbit coupling in spinor BECs. (a) Laser geometry. Three laser beams with different frequencies and polarizations, inters ect- ing at an angle of 2 π/3, illuminate the cloud of atoms. (b) Leveldiagram. EachofthethreeRamanlasers dresses onehy- perfine Zeeman level from |F= 1,mF= 1/angbracketright,|F= 1,mF= 0/angbracketright and|F= 1,mF=−1/angbracketrightof the87Rb 5S1/2,F= 1 ground state. δ1,δ2andδ3correspond to the detuning in the Raman tran- sitions. (c) Triple-well dispersion relation. The SU(3) sp in- orbit coupling induces three discrete minima of the single- particle energy band on the vertices of an equilateral trian gle in thekx-kyplane. (d) Projection of the first energy band on a 2D plane. Units with /planckover2pi1=m= 1 are used for simplicity. III. PHASE DIAGRAM Next, we discuss the phase diagram of the many-body ground states. For the case of SU(2) SO coupling, it is shown that two many-body ground states, magnetized state and stripe state, can be stabilized in a homoge- neous system [7, 24, 26]. Although the Rashba SO cou- pling provides infinite degenerate minima in the single- particle spectrum, a many-body ground state condensed in one or two points in momentum space is always ener- getically favorable due to the presence of spin-exchange interaction [24]. As a result, a lattice state with the con- densates occupying three or more momentum points for SU(2) SO coupling is unstable, unless a strong harmonic trap is introduced [21, 42, 46]. For the present case of SU(3) SO coupling, we first analytically calculate the possible ground states using a variational approach with a trial wave function Ψ= α1Ψ1+α2Ψ2+α3Ψ3, where3 Ψ1=1√ 3 1 1 1 e−i2κx, (2a) Ψ2=1√ 3 e−iπ 3 eiπ 3 eiπ eiκ(x−√ 3y),(2b) Ψ3=1√ 3 eiπ 3 e−iπ 3 eiπ eiκ(x+√ 3y), (2c) correspond to the many-body states with all particles condensing on one of the three minima of the single- particlespectrum, and αi=1,2,3areexpansioncoefficients. Substituting Eqs. (2a)-(2c) into the interaction energy functional E=/integraldisplay dr/parenleftigc0 2n2+c2 2|F|2/parenrightig , (3) one obtains E N=/parenleftbiggc0 2+4c2 9/parenrightbigg ¯n−7c2 9¯n/summationdisplay i/negationslash=j|αi|2|αj|2,(4) where ¯n=|α1|2+|α2|2+|α3|2is the mean particle den- sity. By minimizing the interaction energy with respect to the variation of |αi|2, one finds that the spin-exchange interaction plays an important role in determining the phase diagram. Whenc2>0, it favors |α1|2=|α2|2=|α3|2= ¯n/3, indicating that the ground state is a triangular lattice phasewithanequallyweightedsuperpositionofthe three single-particle minima. On the other hand, as c2<0, the system prefers a state with either |α1|2=¯n,|α2|2= |α3|2= 0, or |α2|2= ¯n,|α1|2=|α3|2= 0, or |α3|2= ¯n,|α1|2=|α2|2= 0, indicating that the ground state occupies one single minimum in the momentum space, and corresponds to a three-fold degenerate magnetized phase. Note that the variational wave function Eqs. (2a)-(2c) is a good starting point as the SO coupling is strong enough to dominate the chemical potential. For the case with weak SO coupling, one must rely on numerical sim- ulations to determine the many-body ground state. In such a situation, we find a stripe phase with two minima in momentum space occupied for c2≫κ2, which will be discussed latter. The many-body ground states can be numerically ob- tained by minimizing the energy functional associated with the Hamiltonian Eq. (1) via the imaginarytime evo- lution method. It is found that the numerical results are consistent with the analytical analysis discussed above for rather weak interaction with c2/lessorsimilarκ2. Figure 2 il- lustrates the two possible ground states of spinor BECs with SU(3) SO coupling. When c2>0, the three compo- nents are immiscible and arranged as an interlaced tri- angular lattice with the spatial translational symmetry FIG. 2: (Color online) Two distinct phases present in SU(3) spin-orbit coupled BECs. (a)-(d) The topologically nontri vial lattice phase for antiferromagnetic spin interaction ( c2>0) with (a) the density and phase of Ψ 1represented by heights and colors, (b) the phase within one unit cell showing the positions of vortices (white circles) and antivortices (bl ack circles), (c) the corresponding momentum distributions, a nd (d) the structural schematic drawing of the phase separatio n. (e)-(f) The three-fold degenerate magnetized phase for fer ro- magnetic spin interaction ( c2<0) with (e) the density and phase distributions of Ψ 1and (f) the corresponding momen- tum distributions. spontaneously broken [See Figs. 2(a)-2(d)]. This lattice is topologically nontrivial and embedded by vortices and antivortices as shown in Fig. 2(b). From this result, we conclude that a lattice phase can be stabilized in a uniform SU(3) SO coupled BEC, which is in clear con- trast to the SU(2) case where a strong harmonic trap is required[21,42,46]. Moredetailsonthestructureofvor- tices aswell astheir unique spin configurationswill be in- vestigated later. On the other hand, as c2<0, the three components are miscible, and the system forms a magne- tized phase with the spatial transitional symmetry pre- served but the time-reversal symmetry broken [See Figs. 2(e)-2(f)]. This magnetized phase occupies one of the three minima of the single-particle spectrum by sponta- neous symmetry breaking, hence is three-fold degenerate instead of doubly degenerate in the SU(2) case [26, 27]. For strong antiferromagnetic spin interaction with c2≫κ2, however, a stripe phase is identified with two of three minima occupied in the momentum space. We take the states with two or three minima occupied in the momentum space as trial wave functions, and per- form imaginary time evolution to find their respective optimized ground state energy. A typical set of results are summarized in Figure 3(a), showing the energy com- parison with different values of interatomic interactions. Obviously, one finds that the stripe phase will has lower energy than the lattice phase when the interatomic in- teraction exceeds a critical value. Due to the finite mo- mentum in vertical direction of the stripe [See Fig. 3(d)], both the spatial translational and time-reversal symme- tries are broken [See Figs. 3(b)-3(c)]. This is distinct4 FIG. 3: (color online) (a) Energy comparison between the lattice and stripe phases. The energy difference ∆ Ebetween the numerical simulation and the variational calculation a re shown by solid (lattice state) and dashed (stripe state) lin es. (b)-(d) The ground-state density, phase and momentum dis- tributions of the stripe phase with the parameters c2= 20κ2 andc0= 10c2. from the stripe phase induced by SU(2) SO coupling, where the time-reversal symmetry is preserved [24]. IV. PHASE REQUIREMENT The vortex configuration of spinor BECs depends on the phase relation between the three components. We next discuss the influence of SO coupling on the phase requirement of the vortex configuration. We first assume that the spinor order parameter of a vortex in the polar coordinate ( r,θ) can be described as ψj(r,θ) =φjeiwjθ+αj, (5) wherej= 0,±1 andφj≥0. A. Without spin-orbit coupling In the absence of SO coupling, the phase-dependent terms in the Hamiltonian are Hphase=Ephase kin+Ephase int =−1 2/integraldisplay Ψ∗1 r2∂2 ∂θ2Ψdr+2c2/integraldisplay ℜ(ψ∗ 1ψ∗ −1ψ2 0)dr,(6) where the first term results from the kinetic energy and the second from the spin-exchange interaction. Substi-tuting Eq. (5) into (6), one obtains Ephase kin=/summationdisplay j=1,0,−1w2 j/integraldisplayπφ2 j rdr, (7) Ephase int= 2c2/integraldisplay φ1φ−1φ2 0rdr /integraldisplay cos[(w 1−2w0+w−1)θ+(α1−2α0+α−1)]dθ.(8) It is easy to read from Eq. (7) that the system favors small winding numbers energetically. Moreover, from Eq. (8) the energy minimization requires the winding number and phase satisfy the following relations w1−2w0+w−1= 0, (9a) α1−2α0+α−1=nπ, (9b) wherenis odd forc2>0 and even for c2<0. The phase requirementofEq. (9a)indicatesthatthefollowingtypes of winding combination, such as ∝angbracketleft±1,×,0∝angbracketright,∝angbracketleft0,×,±1∝angbracketright, ∝angbracketleft±1,0,∓1∝angbracketright,∝angbracketleft±1,±1,±1∝angbracketright,∝angbracketleft±2,±1,0∝angbracketrightand∝angbracketleft0,±1,±2∝angbracketrightare allowed in a spinor BEC, where the symbol “ ×” denotes the absence of the Ψ 0component. B. With SU(2) spin-orbit coupling For the case of SU(2) SO coupling, we take the Rashba type as an example, and write the Hamiltonian as Esoc=/integraldisplay κψ† 0−i∂x−∂y0 −i∂x+∂y0−i∂x−∂y 0−i∂x+∂y0 ψdr,(10) whereψ= [ψ1,ψ0,ψ−1]⊤. Substituting Eq. (5)into(10), one can obtain Esoc=/integraldisplay drdθ/bracketleftig (φ0r∂rφ1−w1φ0φ1)ei[(w1−w0+1)θ+(α1−α0−π 2)] −(φ1r∂rφ0+w0φ1φ0)e−i[(w1−w0+1)θ+(α1−α0−π 2)] +(φ0r∂rφ−1+w−1φ0φ−1)ei[(w−1−w0−1)θ+(α−1−α0−π 2)] −(φ−1r∂rφ0−w0φ−1φ0)e−i[(w−1−w0−1)θ+(α−1−α0−π 2)]/bracketrightig . (11) In order to minimize the SO coupling energy, it is pre- ferred that w1−w0+1 = 0, (12a) w−1−w0−1 = 0, (12b) α1−α0−π 2=mπ, (12c) α−1−α0−π 2=nπ. (12d) Then the SO coupling energy is rewritten as Esoc= 2π/integraldisplay [φ0r∂rφ1−φ1r∂rφ0−(w1+w0)φ0φ1]drcosmπ +2π/integraldisplay [φ0r∂rφ−1−φ−1r∂rφ0+(w−1+w0)φ0φ−1]drcosnπ, (13)5 wheremandnare odd or even, which can be deter- mined by minimizing the energy expressed in Eq. (13). It is found that the SU(2) SO coupling does not vio- late the ordinary requirement on the winding combina- tion in Eq. (9a), but introduces further requirements in Eqs.(12a)-(12b). Asaresult, while ∝angbracketleft−1,0,1∝angbracketright,∝angbracketleft−2,−1,0∝angbracketright and∝angbracketleft0,1,2∝angbracketrightare still allowed, some winding combinations such as ∝angbracketleft±1,±1,±1∝angbracketright,∝angbracketleft±1,×,0∝angbracketright,∝angbracketleft0,×,±1∝angbracketright,∝angbracketleft1,0,−1∝angbracketright, ∝angbracketleft2,1,0∝angbracketrightand∝angbracketleft0,−1,−2∝angbracketrightare forbidden. Obviously, one can see that the SO coupling break the chiral symmetry, thus may lead to chiral spin textures. C. With SU(3) spin-orbit coupling For the case of SU(3) SO coupling, the effective Hamil- tonian can be written as Esoc=/integraldisplay κψ† 0−i∂x−∂y−i∂x+∂y −i∂x+∂y0−i∂x−∂y −i∂x−∂y−i∂x+∂y0 ψdr.(14) Substituting Eq. (5) into (14), we get Esoc=/integraldisplay drdθ/bracketleftig (φ0r∂rφ1−w1φ0φ1)ei[(w1−w0+1)θ+(α1−α0−π 2)] −(φ1r∂rφ0+w0φ1φ0)e−i[(w1−w0+1)θ+(α1−α0−π 2)] +(φ0r∂rφ−1+w−1φ0φ−1)ei[(w−1−w0−1)θ+(α−1−α0−π 2)] −(φ−1r∂rφ0−w0φ−1φ0)e−i[(w−1−w0−1)θ+(α−1−α0−π 2)] +(φ−1r∂rφ1+w1φ−1φ1)ei[(w1−w−1−1)θ+(α1−α−1−π 2)] −(φ1r∂rφ−1−w−1φ1φ−1)e−i[(w1−w−1−1)θ+(α1−α−1−π 2)]/bracketrightig . (15) By minimizing the SO coupling energy, one obtains the following relations w1−w0+1 = 0, (16a) w−1−w0−1 = 0, (16b) w1−w−1−1 = 0, (16c) α1−α0−π 2=mπ, (16d) α−1−α0−π 2=nπ, (16e) α1−α−1−π 2=lπ. (16f) Then the SO coupling energy can be rewritten as Esoc=2π/integraldisplay [φ0r∂rφ1−φ1r∂rφ0−(w1+w0)φ0φ1]drcosmπ +2π/integraldisplay [φ0r∂rφ−1−φ−1r∂rφ0+(w−1+w0)φ0φ−1]drcosnπ +2π/integraldisplay [φ−1r∂rφ1−φ1r∂rφ−1+(w1+w−1)φ−1φ1]drcoslπ, (17)wherem,nandlare odd or even, which can be determined from Eq. (17). However, the three winding requirements Eqs. (16a)-(16c) can not be satisfied simul- taneously. Thus the SU(3) SO coupling may choose two out of the three winding requirements for the following three cases: Case I: w1−w0+1 = 0, (18a) w−1−w0−1 = 0, (18b) α1−α0−π 2=mπ, (18c) α−1−α0−π 2=nπ. (18d) Case II: w1−w0+1 = 0, (19a) w1−w−1−1 = 0, (19b) α1−α0−π 2=mπ, (19c) α1−α−1−π 2=lπ. (19d) Case III: w−1−w0−1 = 0, (20a) w1−w−1−1 = 0, (20b) α−1−α0−π 2=nπ, (20c) α1−α−1−π 2=lπ. (20d) For case I, the winding combination ∝angbracketleft−1,0,1∝angbracketrightis allowed, while∝angbracketleft1,0,−1∝angbracketrightis not allowed, indicating the chiral sym- metry is broken. For case II and case III, one can find that the SU(3) SO coupling breaks the ordinary require- ment on the winding combination in Eq. (9a), thus new winding combinations, such as ∝angbracketleft0,1,−1∝angbracketrightand∝angbracketleft1,−1,0∝angbracketright, are possible. V. VORTEX CONFIGURATIONS The vortex configurations of spinor BECs can be clas- sified according to the combination of winding numbers and the magnetization of vortex core [35–37]. For ex- ample, a Mermin-Ho vortex has winding combination ∝angbracketleft±2,±1,0∝angbracketrightwith a ferromagnetic core, where the plus and minus signs represent different chirality of the vor- tices [47], and the expression of ∝angbracketleftw1,w0,w−1∝angbracketrightindicates that the components of Ψ 1, Ψ0and Ψ −1in the wave function acquire winding numbers of w 1, w0and w −1, respectively. Using this notation, a polar-core vortex has winding combination ∝angbracketleft±1,0,∓1∝angbracketrightwith an antiferromag- netic core, and a half-quantum vortex has winding com- bination ∝angbracketleft±1,×,0∝angbracketrightwith a ferromagnetic core, where the symbol “ ×” denotes the absence of the Ψ 0component.6 FIG. 4: (Color online) Vortex configurations in antiferro- magnetic spinor BECs with SU(3) spin-orbit coupling. (a) Vortex arrangement among the three components of the con- densates. One can identify three types of vortices, includ- ing a polar-core vortex with winding combination /angbracketleft−1,0,1/angbracketright (blue line) and two ferromagnetic-core vortices with wind- ing combinations /angbracketleft1,−1,0/angbracketright(green line) and /angbracketleft0,1,−1/angbracketright(red line). (b)-(d) Spherical-harmonic representation of the t hree types of vortices. The surface plots of |Φ(θ,φ)|2for (b) the polar-core vortex /angbracketleft−1,0,1/angbracketright, (c) the ferromagnetic-core vor- tex/angbracketleft1,−1,0/angbracketright, and (d) the ferromagnetic-core vortex /angbracketleft0,1,−1/angbracketright are shown with the colors representing the phase of Φ(θ,φ). Here,Φ(θ,φ) =/summationtext1 m=−1Y1m(θ,φ)ΨmandY1mis the rank-1 spherical-harmonic function. In the lattice phase induced by the SU(3) SO cou- pling with antiferromagnetic spin interaction, there ex- ists three types of vortices: one is a polar-core vortex with winding combination ∝angbracketleft−1,0,1∝angbracketright, and the other two are ferromagnetic-core vortices with winding combina- tions∝angbracketleft1,−1,0∝angbracketrightand∝angbracketleft0,1,−1∝angbracketright[See Fig. 4(a)]. However, the vortex configurations with opposite chirality of each type, such as ∝angbracketleft1,0,−1∝angbracketright,∝angbracketleft−1,1,0∝angbracketrightand∝angbracketleft0,−1,1∝angbracketright, are not allowed, because the chiral symmetry is intrincically bro- kenin SU(3) SO coupledsystems, asdiscussed inSec. IV. Surprisingly, one finds that the two types of ferromagnetic-core vortices ∝angbracketleft1,−1,0∝angbracketrightand∝angbracketleft0,1,−1∝angbracketrightvio- late the conventional phase requirement 2w 0= w1+w−1 for ordinary spinor BECs [35–37]. This can be under- stood by noting that the relative phase among different wave function components are no longer uniquely deter- mined by the spin-exchange interaction but also affected by the SU(3) SO coupling, as qualitatively explained in Sec. IV. Thus, the interlaced arrangement of the three types of vortices forms a new class of vortexlattice which has no analogue in systems without SO coupling. The configurations of the three types of vortices in- duced by the SU(3) SO coupling with antiferromagnetic interaction are essentially different from those usually observed in ferromagnetic spinor BECs, as can be il- lustrated by the spherical-harmonic representation [35]. From Figs. 4(b)-4(d) one can find that for the polar-corevortex,theantiferromagneticorderparametervariescon- tinuously everywhere, while for the ferromagnetic-core vortex,themagneticorderparameteracquiresasingular- ity at the vortex core. In contrast, in the ordinary ferro- magnetic spinor BECs, the ferromagnetic order parame- ter varies continuously everywhere for the ferromagnetic- corevortex,buthasasingularityatthecoreforthepolar- core vortex [35]. VI. DOUBLE-QUANTUM SPIN VORTICES Spin vortex is a complex topological defect resulting from symmetry breaking, and is characterized by zero net mass current and quantized spin current around an unmagnetized core [35, 38, 48–51]. It is not only different from the magnetic vortex found in magnetic thin films [52–54], but alsofromthe 2Dskyrmion[55,56] dueto the existence of singularity in the spin textures [57]. Single- quantum spin vortex with the spin current showing one quantum ofcirculation has been experimentally observed in ferromagnetic spinor BECs [58]. Multi-quantum spin vortices with l(l≥2) quanta circulating spin current, however, are considered to be topologically unstable and have not been discovered yet [35]. A particularly important finding of our present work is that the polar-core vortex in the lattice phase has a spin current with two quanta of circulation around the unmagnetized core, hence can be identified as a double- quantum spin vortex. Figure 5 presents the transverse magnetization F+=Fx+iFy, longitudinal magnetiza- tionFz, and amplitude of the total magnetization |F| in the lattice phase, which are experimentally observ- able by magnetization-sensitive phase-contrast imaging technique [59]. From these results, one can find two dis- tinct types of topological defects, double-quantum spin vortex (DSV) and half skyrmion (HS) [60, 61], which correspond to the polar-core vortex with winding com- binations ∝angbracketleft−1,0,1∝angbracketrightand the ferromagnetic-core vortex with winding combinations ∝angbracketleft1,−1,0∝angbracketrightor∝angbracketleft0,1,−1∝angbracketright, re- spectively. In particular, for the double-quantum spin vortex, the core is unmagnetized and the orientation of the magnetization along a closed path surrounding the core acquires a rotation of 4 π. This finding indicates that a regular lattice of multi-quantum spin vortices can emerge spontaneously in antiferromagnetic spinor BECs with SU(3) SO coupling. By exploring the effect of a small but finite temperature, we confirm that the double- quantum spin vortices are robust against thermal fluctu- ations and hence are observable in experiments, as dis- cussed in Appendix B. The emergence of spin current with two quanta of cir- culation can be analytically understood by expanding the wave function obtained by the variational methods around the center of a double-quantum spin vortex. We suppose that the wave function of the lattice phase is7 FIG. 5: (Color online) Double-quantum spin vortex in an- tiferromagnetic spinor BECs with SU(3) spin-orbit couplin g. (a) Spatial maps of the transverse magnetization with color s indicating the magnetization orientation. (b) Longitudin al magnetization. (c) Amplitude of the total magnetization |F|. Two kinds of topological defects, double-quantum spin vor- tex (DSV) and half skyrmion (HS) are marked by big and small circles, respectively. The transverse magnetizatio n ori- entation arg F+along a closed path (indicated by big circles) surrounding the unmagnetized core shows a net winding of 4π, revealing the presence of a double-quantum spin vortex. written as ψ=1 3 1 1 1 e−i2κx+1 3 e−iπ 3 eiπ 3 eiπ eiκ(x−√ 3y)+1 3 eiπ 3 e−iπ 3 eiπ eiκ(x+√ 3y). (21) Then one can expand ψaround the center of a vortex with winding number ∝angbracketleft−1,0,1∝angbracketright, e.g., at the location of (x,y) = (0,π/(3√ 3κ)). Substituting x=ǫcosθandy= π/(3√ 3κ)+ǫsinθintoψand expanding with respect to the infinitesimal ǫ, we obtain ψ= −iκe−iθǫ−1 2κ2ei2θǫ2 1−κ2ǫ2 −iκeiθǫ−1 2κ2e−i2θǫ2 +O/parenleftbig ǫ3/parenrightbig .(22) Notice that the second-order terms with e±i2θhave no essential influence on the phases, thus the wind- ing number for each component can still be represented as∝angbracketleft−1,0,1∝angbracketright[See Figs. 6(a)-6(c)]. However, since the first-order terms are canceled out when calculating the transverse magnetization F+=√ 2[ψ∗ 1ψ0+ψ∗ 0ψ−1], the second-order terms play a dominant role, leading to the emergence of spin current with two quanta of circulation around an unmagnetized core F+∝ǫ2e−i2θ, (23) as illustrated in Fig. 6(d). FIG. 6: (color online) (a)-(c) Phases of the polar-core vort ex described by the wave function in Eq. (22), displaying the winding combination /angbracketleft−1,0,1/angbracketright. (d) Direction of the trans- verse magnetization, indicating the emergence of spin curr ent with two quanta of circulation. VII. CONCLUSION To summarize, we have mapped out the ground-state phase diagram of SU(3) spin-orbit coupled Bose-Einstein condensates. Several novel phases are discovered includ- ing a three-fold degenerate magnetized phase, a vortex lattice phase, as well as a stripe phase with time-reversal symmetry broken. We also investigate the influence of SU(3) spin-orbit coupling on the phase requirement of thevortexconfiguration,anddemonstratethattheSU(3) spin-orbit coupling breaks the ordinary phase rule of spinor Bose-Einstein condensates, and allows the sponta- neous emergence of stable double-quantum spin vortices. As a new member in the family of topological defects, double quantum spin vortex has never been discovered in any other systems. Our work deepen the understand- ing of spin-orbit phenomena, and will attract extensive interest of scientists in the cold atom community. ACKNOWLEDGMENTS This work was supported by NKRDP under grants Nos. 2016YFA0301500, 2012CB821305; NSFC under grants Nos. 61227902, 61378017, 11274009, 11434011, 11434015, 11447178, 11522436, 11547126; NKBRSFC under grant No. 2012CB821305; SKLQOQOD un- der grant No. KF201403; SPRPCAS under grant Nos. XDB01020300, XDB21030300; JSPS KAKENHI8 grant No. 26400414 and MEXT KAKENHI grant No. 25103007. W. H. and X.-F. Z. contributed equally to this work. APPENDIX A: DERIVING THE EFFECTIVE HAMILTONIAN We consider spinor Bose-Einstein condensates (BECs) illuminated by three Raman laser beams, which couple two of the three hyperfine spin components respectively, as illustrated in Figs. 1(a)-1(b) of the main text. The internal dynamics of a single particle under this scheme can be described by the Hamiltonian H=3/summationdisplay j=1/parenleftbigg/planckover2pi12k2 2m+εj/parenrightbigg |j∝angbracketright∝angbracketleftj|+n/summationdisplay l=1El|l∝angbracketright∝angbracketleftl| +3/summationdisplay j=1n/summationdisplay l=1/bracketleftig Ωjei(Kj·r+ωjt)Mlj|l∝angbracketright∝angbracketleftj∝angbracketright+h.c./bracketrightig ,(A1) where/planckover2pi1kis the momentum of the particles, and εjand Elare the energies of the ground and excited states, re- spectively. In the atom-light coupling term, Kjandωj are the wave vectors and frequencies of the three Raman lasers with Ω jthe corresponding Rabi frequencies, and Mljis the matrix element of the dipole transition. One can see that this Hamiltonian is similar to that used in the scheme for creating 2D spin-orbit (SO) coupling in ultracold Fermi gases [9], thus can be readily realized in Bose gases. Taking the standard rotating wave approxi- mation to get rid of the time dependence of the Hamilto- nian, and adiabatically eliminating the excited states for far detuning, the Hamiltonian can be rewritten as H= /planckover2pi12(k+K1)2 2m+δ1Ω12 Ω13 Ω21/planckover2pi12(k+K2)2 2m+δ2Ω23 Ω31 Ω32/planckover2pi12(k+K3)2 2m+δ3 ,(A2) whereδ1,δ2andδ3arethetwo-photondetunings, andthe realparametersΩ jj′= Ωj′jdescribe the Raman coupling strength between hyperfine ground states |j∝angbracketrightand|j′∝angbracketright, which can be expressed as [9, 62] Ωjj′=−/radicalbigIjIj′ /planckover2pi12cǫ0/summationdisplay m′∝angbracketleftj′|erq|m′∝angbracketright∝angbracketleftm′|erq|j∝angbracketright ∆.(A3) Here,Ijis the intensity of each Raman laser, and ∆ denotes the one-photon detuning. Other parameters c, ǫ0andein Eq. (A3) are the speed of light, permittiv- ity of vacuum and elementary charge, respectively. In Eq. (A3),q=x,y,zis an index labeling the components ofrin the spherical basis, and |m′∝angbracketrightdescribes the middle excited hyperfine spin state in the Raman process. For simplicity, we assume Ω = Ω 12= Ω13= Ω23, which can always be satisfied by adjusting the system parameters, such as the laser intensity. Introducing a unitary transformation U=1√ 3 1 1 1 −e−iπ 3−eiπ 31 −eiπ 3−e−iπ 31 (A4)and a time-dependent unitary transformation U(t) = ei/parenleftbigg /planckover2pi12K2 0 2m+δ2−Ω/parenrightbigg t, the effective Hamiltonian becomes H= /planckover2pi12k2 2m+δ1−δ20 0 0/planckover2pi12k2 2m0 0 0/planckover2pi12k2 2m+δ3−δ2+3Ω +Vso,(A5) where the laser vectors K1=−K0ˆ ey,K2=√ 3K0 2ˆ ex+ K0 2ˆ eyandK3=−√ 3K0 2ˆ ex+K0 2ˆ eyare defined with K0= 2mκ//planckover2pi1. The spin-dependent uniform potential induced by the Raman detuning δiand Raman coupling strength Ω can be eliminated by applying a Zeeman field, leading to H= /planckover2pi12k2 2m+ǫ10 0 0/planckover2pi12k2 2m0 0 0/planckover2pi12k2 2m+ǫ2 +Vso,(A6) whereǫ1=δ1−δ2+∆1+∆2andǫ2=δ3−δ2−∆1+∆2+ 3Ω with ∆ 1and ∆ 2denoting the linear and quadratic Zeeman energy respectively. By tuning the detuning, the Zeeman energy and the Raman coupling strength, one can reach the regime ∆ 1=δ3−δ1+3Ω 2and ∆ 2=δ2− δ1+δ3+3Ω 2which satisfying ǫ1=ǫ2= 0. Then we have H=/planckover2pi12k2 2m+Vso, (A7) which is the single-particle Hamiltonian with SU(3) SO coupling considered in the main text. APPENDIX B: STABILITY OF THE DOUBLE-QUANTUM SPIN VORTEX STATES In order to verify the stability of the phases discovered in this manuscript, we have explored the effects of a small but finite temperature, and concluded that the double-quantum spin vortex states are robust against the thermal fluctuations. In particular, we considered a random fluctuation ∆ φin the real-time evolution of the Gross-Pitaevskii equation, which causes an energy fluctuation about ∆ E= 0.03EgwithEgthe ground-state energy. An estimation shows that this level of fluctuation corresponds to the energy scale kBTwithT∼300 nK, which is higher enough for a usual system of Bose-Einstein condensates in realistic experiments. According to numerical simulations, we find that the structure of the double-quantum spin vortex state is stable under this level of fluctuation in tens of millisenconds [See Fig. 7], suggesting that this phase is indeed observable in experiments.9 FIG. 7: (color online) Stable double-quantum spin vortex under a random fluctuation. The images are taken at t= 20msin the real-time evolution, with thermal fluctuation in the en- ergy scale of kBTwithT∼300 nK. (a) Spa- tial maps of the transverse magnetization with colors indicating the magnetization orientation. (b) Longitudinal magnetization. (c) Amplitude of the total magnetization |F|. It is shown that the double-quantum spin vortices are topologi- cally stable under external fluctuations with a fairly long lifetime of tens of ms. [1] Y. J. Lin, K. Jim´ enez-Garc´ ıa, and I. B. Spielman, Natur e 471, 83 (2011). [2] P. Wang, Z. Q. Yu, Z. Fu, J. Miao, L. Huang, S. Chai, H. Zhai, and J. Zhang, Phys. Rev. Lett. 109, 095301 (2012). [3] L. W. Cheuk, A. T. Sommer, Z. Hadzibabic, T. Yefsah, W. S. Bakr, and M. W. Zwierlein, Phys. Rev. Lett. 109, 095302 (2012). [4] J. Y. Zhang, S. C. Ji, Z. Chen, L. Zhang, Z. D. Du, B. Yan, G. S. Pan, B. Zhao, Y. J. Deng, H. Zhai, S. Chen, and J. W. Pan, Phys. Rev. Lett. 109, 115301 (2012). [5] C. Qu, C. Hamner, M. Gong, C. Zhang, and P. Engels, Phys. Rev. A 88, 021604(R) (2013). [6] Z. Fu, L. Huang, Z. Meng, P. Wang, L. Zhang, S. Zhang, H. Zhai, P. Zhang, and J. Zhang, Nat. Phys. 10, 110 (2014). [7] S. C. Ji, J. Y. Zhang, L. Zhang, Z. D. Du, W. Zheng, Y. J. Deng, H. Zhai, S. Chen, and J. W. Pan, Nat. Phys. 10, 314 (2014). [8] A. J. Olson, S.-J. Wang, R. J. Niffenegger, C.-H. Li, C. H. Greene, and Y. P. Chen, Phys. Rev. A 90, 013616 (2014). [9] L. Huang, Z. Meng, P. Wang, P. Peng, S.-L. Zhang, L. Chen, D. Li, Q. Zhou, and J. Zhang, Nat. Phys. 12, 540 (2016). [10] Z. Wu, L. Zhang, W. Sun, X.-T. Xu, B.-Z. Wang, S.- C. Ji, Y. Deng, S. Chen, X.-J. Liu, and J.-W. Pan, arXiv:1511.08170 (2015). [11] A. M. Dudarev, R. B. Diener, I. Carusotto, and Q. Niu, Phys. Rev. Lett. 92, 153005 (2004). [12] V. Galitski and I. B. Spielman, Nature 494, 49 (2013). [13] J. Dalibard, F. Gerbier, G. Juzeli¯ unas, and P. ¨Ohberg, Rev. Mod. Phys. 83, 1523 (2011). [14] N. Goldman, G. Juzeli¯ unas, P. ¨Ohberg, and I. B. Spiel- man, Rep. Prog. Phys. 77, 126401 (2014). [15] H. Zhai, Int. J. Mod. Phys. B 26, 1230001 (2012). [16] H. Zhai, Rep. Prog. Phys. 78, 026001 (2015). [17] W. Yi, W. Zhang, and X. Cui, Sci. China Phys. Mech. Astron. 58,014201 (2015). [18] J. Zhang, H. Hu, X. J. Liu, and H. Pu, Annu. Rev. Cold At. Mol. 2, 81 (2014).[19] C. Wu, Mod. Phys. Lett. B 23, 1 (2009). [20] C. Wu, I. Mondragon-Shem, and X. F. Zhou, Chin. Phys. Lett. 28, 097102 (2011). [21] H. Hu, B. Ramachandhran, H. Pu, and X. J. Liu, Phys. Rev. Lett. 108, 010402 (2012). [22] H. Sakaguchi, B. Li, and B. A. Malomed, Phys. Rev. E 89, 032920 (2014). [23] W. Han, G. Juzeli¯ unas, W. Zhang, and W. M. Liu, Phys. Rev. A91, 013607 (2015). [24] C. Wang, C. Gao, C. M. Jian, and H. Zhai, Phys. Rev. Lett.105, 160403 (2010). [25] T. L. Ho and S. Zhang, Phys. Rev. Lett. 107, 150403 (2011). [26] Y. Li, L. P. Pitaevskii, and S. Stringari, Phys. Rev. Let t. 108, 225301 (2012). [27] T. F. J. Poon and X. J. Liu, Phys. Rev. A 93, 063420 (2016). [28] A. Polkovnikov, K. Sengupta, A. Silva, and M. Vengalat- tore, Rev. Mod. Phys. 83, 863 (2011). [29] Y. Li, G. I. Martone, L. P. Pitaevskii, and S. Stringari, Phys. Rev. Lett. 110, 235302 (2013). [30] G. I. Martone, Y. Li, L. P. Pitaevskii, and S. Stringari, Phys. Rev. A 86, 063621 (2012). [31] K. Sun, C. Qu, Y. Xu, Y. Zhang, and C. Zhang, Phys. Rev. A93, 023615 (2016). [32] B. M. Anderson, G. Juzeli¯ unas, V. M. Galitski,and I. B. Spielman, Phys. Rev. Lett. 108, 235301 (2012). [33] T. Graß, R. W. Chhajlany, C. A. Muschik, and M. Lewenstein, Phys. Rev. B 90, 195127 (2014). [34] R. Barnett, G. R. Boyd, andV.Galitski, Phys.Rev.Lett. 109, 235308 (2012). [35] Y. Kawaguchi and M. Ueda, Phys. Rep. 520, 253 (2012). [36] T. Isoshima, K. Machida, and T. Ohmi, J. Phys. Soc. Jpn.70, 1604 (2001). [37] T. Mizushima, N. Kobayashi, and K. Machida, Phys. Rev. A70, 043613 (2004). [38] D. M. Stamper-Kurn and M. Ueda, Rev. Mod. Phys. 85, 1191 (2013). [39] D. S. Wang, Y. R. Shi, K. W. Chow, Z. X. Yu, and X. G. Li, Eur. Phys. J. D 67, 242 (2013).10 [40] D. S. Wang, Y. Q. Ma, and X. G. Li, Commun. Nonlinear Sci. Numer. Simulat. 19, 3556 (2014). [41] G. B. Arfken, H. J. Weber, and F. E. Harris, Mathemat- ical Methods for Physicists (7th ed.) (2000). [42] Z. F. Xu, Y. Kawaguchi, L. You, and M. Ueda, Phys. Rev. A86, 033628 (2012). [43] Z. Lan and P. Ohberg, Phys. Rev. A 89, 023630 (2014). [44] Y. Zhang, L. Mao, and C. Zhang, Phys. Rev. Lett. 108, 035302 (2012). [45] T. D. Stanescu, B. Anderson, andV. Galitski, Phys. Rev. A78, 023616 (2008). [46] S. Sinha, R. Nath, and L. Santos, Phys. Rev. Lett. 107, 270401 (2011). [47] H. Saito, Y. Kawaguchi, and M. Ueda, Phys. Rev. Lett. 96, 065302 (2006). [48] H. Saito, Y. Kawaguchi, and M. Ueda, Phys. Rev. A 75, 013621 (2007). [49] Y. Kawaguchi, H. Saito, K. Kudo, and M. Ueda, Phys. Rev. A82, 043627 (2010). [50] J. Lovegrove, M. O. Borgh, and J. Ruostekoski, Phys. Rev. A86, 013613 (2012). [51] S.-W. Su, I.-K. Liu, Y.-C. Tsai, W. M. Liu, and S.-C. Gou, Phys. Rev. A 86, 023601 (2012). [52] A. Hubert and R. Schafer, Magnetic Domains (Springer,Berlin, 1998). [53] T. Shinjo, T. Okuno, R. Hassdorf, K. Shigeto, and T. Ono, Science 289, 930 (2000). [54] A. Wachowiak, J. Wiebe, M. Bode, O. Pietzsch, M. Mor- genstern, and R. Wiesendanger, Science 298, 577 (2002). [55] L. S. Leslie, A. Hansen, K. C. Wright, B. M. Deutsch, and N. P. Bigelow, Phys. Rev. Lett. 103, 250401 (2009). [56] Jae-yoon Choi, Woo Jin Kwon, and Yong-il Shin, Phys. Rev. Lett. 108, 035301 (2012). [57] S. Yi and H. Pu, Phys. Rev. Lett. 97, 020401 (2006). [58] L. E. Sadler, J. M. Higbie, S. R. Leslie, M. Vengalattore , and D. M. Stamper-Kurn, Nature 443, 312 (2006). [59] Other experimental techniques, such as selective abso rp- tion and time-of-flight measurement, can also be used to observe the double-quantum spin vortex states. We sug- gest the magnetization-sensitive phase-contrast imaging method just because it has advantage in observing the phase information of the spin-vortex structure. [60] G. E. Brown and M. Rho, The Multifaceted Skyrmion (World Scientific, Singapore, 2010). [61] M. DeMarco and H. Pu, Phys. Rev. A 91, 033630 (2015). [62] T. A.Savard, S. R.Granade, K. M. O’Hara, M. E. Gehm, and J. E. Thomas, Phys. Rev. A 60, 4788 (1999).
1208.2923v2.Chaos_driven_dynamics_in_spin_orbit_coupled_atomic_gases.pdf
Chaos-driven dynamics in spin-orbit coupled atomic gases Jonas Larson,1, 2,Brandon M. Anderson,3and Alexander Altland2 1Department of Physics, Stockholm University, Se-106 91 Stockholm, Sweden 2Institut f ur Theoretische Physik, Universit at zu K oln, K oln, De-50937, Germany 3Joint Quantum Institute, national Institute of Standards and technology and the University of Maryland, Gaithersburg, Maryland 20899-8410, USA (Dated: May 24, 2022) The dynamics, appearing after a quantum quench, of a trapped, spin-orbit coupled, dilute atomic gas is studied. The characteristics of the evolution is greatly in uenced by the symmetries of the system, and we especially compare evolution for an isotropic Rashba coupling and for an anisotropic spin-orbit coupling. As we make the spin-orbit coupling anisotropic, we break the rotational sym- metry and the underlying classical model becomes chaotic; the quantum dynamics is a ected ac- cordingly. Within experimentally relevant time-scales and parameters, the system thermalizes in a quantum sense. The corresponding equilibration time is found to agree with the Ehrenfest time, i.e. we numerically verify a log(h1) scaling. Upon thermalization, we nd the equilibrated distributions show examples of quantum scars distinguished by accumulation of atomic density for certain energies. At shorter time-scales we discuss non-adiabatic e ects deriving from the spin-orbit coupled induced Dirac point. In the vicinity of the Dirac point, spin uctuations are large and, even at short times, a semi-classical analysis fails. PACS numbers: 03.75.Kk, 03.75.Mn I. INTRODUCTION The physics of ultracold atomic gases has greatly ad- vanced in recent years [1]. The high control of system pa- rameters, together with the isolation of the system from its environment, have made it possible to use such se- tups to simulate various theoretical models of condensed matter physics [1, 2]. Of signi cance in many condensed matter models is the response to external magnetic elds. Since atoms are neutral, there is no direct way to imple- ment a Lorentz force in these systems. Early experiments created a synthetic magnetic eld via rotation [3]. While simple theoretically, these methods are impractical for certain setups, and they are limited to weak, uniform elds. The rst experimental demonstration of laser- induced synthetic magnetic elds for neutral atoms [4], on the other hand, paves the way for an avenue of new situations to be studied in a versatile manner [5{7]. Ow- ing to numerous fundamental applications in the con- densed matter community [8, 9], maybe the most im- portant direction appears when the laser elds induce a synthetic spin-orbit (SO) coupling. Indeed, a certain kind of SO-coupling for neutral atoms has already been demonstrated [10], and it is expected that more general SO-couplings will be attainable within the very near fu- ture [11, 12]. While SO-couplings can in principle bear identical forms in condensed matter and cold atom models, there is an inevitable di erence, often overlooked, between these two systems. The presence of a con ning potential for the atomic gas can qualitatively change the physics [1, 3], jolarson@fysik.su.seand has only recently been addressed [13{17]. Fur- thermore, most of these studies are concerned with ground/stationary state properties of the system [13{ 15], while few works discuss dynamics or non-equilibrium physics. Notwithstanding, the experimental isolation of these systems suggests that they are well suited for stud- ies of closed quantum dynamics [18]. Historically, some of the nest experiments regard- ing dynamics of closed quantum systems have been per- formed in quantum optics [19, 20]. An early example proved quantization of the electromagnetic eld by mak- ing explicit use of quantum revivals [21]. Such quantum recurrences, in general connected to integrability or small system sizes, are now well understood. The situation be- comes more complex for non-integrable systems [18] or systems with a large number of degrees-of-freedom [22]. One particularly interesting question is whether any ini- tial state relaxes to an asymptotic state, and if so, what are then the properties of this \equilibrated" state and the mechanism behind the equilibration. Both these questions have inspired numerous publications during the last decade, both theoretical [23, 24] as well as ex- perimental [25{27]. A rule of thumb is that in order for a closed quantum system to thermalize , i.e. all ex- pectation values can be obtained from a microcanonical state, its underlying classical Hamiltonian should be non- integrable [18]. While true in most cases studied so far, exceptions to this hypothesis has been found [28]. More- over, the behavior near the transition from regular to chaotic dynamics, classically explained by Kolmogorov- Arnold-Moser theory [29], is not well understood for a quantum system [30]. It is therefore desirable to study a system where these two regimes can be explored by tuning an external parameter, and for which the exper- imental methods in terms of preparation and detectionarXiv:1208.2923v2 [quant-ph] 28 Jan 20132 are already well developed. Motivated by the above arguments, in this paper we consider dynamics of a trapped SO-coupled cold dilute atomic gas. The SO-coupling is assumed tunable from isotropic (Rashba-like) to anisotropic, and hence the sys- tem can be tuned between regular and chaotic. Note that even though this crossover is generated by a change in the form of the SO-coupling, the con ning trap causes the system to become non-integrable. We distinguish between short and long time evolution, where by \long time" we mean times similar to the Ehrenfest time. In fact, the corresponding time-scale for the thermalization is found to agree with the Ehrenfest time, and thereby scale as log( h1)=whereis the maximum Lyaponov exponent. This scaling for the thermalization has been conjectured in Ref. [31], but was not numerically ver- i ed in these works. At shorter times when the wave packet remains localized, we especially study the rapid changes in the spin as the wave packet evolves in the vicinity of the Dirac point (DP). For energies below the DP (E < 0), we utilize an adiabatic model derived in the Born-Oppenheimer approximation (BOA) [32]. Aside from some special initial states, we encounter thermal- ization in all cases. These exceptions correspond to states evolving within a regular \island" in the otherwise chaotic sea. Among the thermalized states, the equi- librated distributions are found to show quantum scars originating from periodic orbits of the underlying classi- cal model. The experimental relevance of all our theoret- ical predictions are discussed and put in a state-of-the-art experimental perspective. The paper is outlined as follows. The following sec- tion introduces the system Hamiltonian and discusses its symmetries. Section II B derives the adiabatic model by imposing the BOA. A semi-classical analysis, demon- strating classical chaos for anisotropic SO-couplings, is presented in Sec. III. The following section considers the full quantum model at short times, Sec. IV A, and long times, Sec. IV B. Section IV C contains a discussion re- garding experimental relevance of our results. Finally, Sec. V gives some concluding remarks. II. SPIN-ORBIT COUPLED COLD ATOMS A. Model spin-orbit Hamiltonian Several proposals exist for implementing spin-orbit couplings in cold atoms [33{35]. In general, these syn- thetic spin-orbit elds are generated through the appli- cation of optical and Zeeman elds to produce a set of dressed states that are well separated energetically from the remaining dressed states [5]. We denote these states as pseudo-spin, but emphasize that there is no connection to real space rotations. Spatial variation of the dressed states will couple the pseudo-spin to the orbital motion of the atom. An atom prepared in a pseudo-spin state will therefore see an e ective Hamiltonian, provided theatom is suciently cold. For a speci c con guration of optical elds, one can induce the e ective Hamiltonian [35] ^HSO=^p2 2m+1 2m!2r2+vx^px^x+vy^py^y;(1) where ^p= (^px;^py) is the momentum operator, ^r= (^x;^y) is the position operator, mis the mass of the atom, and !the frequency of a harmonic trap. The operator ^ i is thei-th Pauli matrix in pseudo-spin space, and the velocitiesvicouple pseudo-spin to an e ective momen- tum dependent Zeeman eld, B(p) = (vxpx;vypy). This momentum-dependent Zeeman eld can simulate any combination of the Rashba [38] and Dresselhaus [39] SO- couplings experienced in semiconductor quantum wells and systems alike. In the absence of a trap, != 0, the spectrum of (1) is E(px;py) =1 2m p2 x+p2 y +q (vxpx)2+ (vypy)2(2) with the corresponding eigenfunctions j ;pi=eim(vxx+vyy)j'i (3) where j'i=1p 2 ei'=2j"iei'=2j#i ; (4) is a spinor with helicity =1 and'= arctan(vypy=vxpx). These states have well de ned mo- mentum, but have no velocity since h_ri=hrpHi= 0, provided the optical elds are maintained. Note further that the eigenstates are parametrically dependent on px andpy. We remark that for an isotropic SO-coupling, vx=vy, the Hamiltonian (1) is equivalent to the dual E"Jahn- Teller model, frequently appearing in chemical/molecular physics and condensed matter theories [37]. With a simple unitary rotation of the Pauli matrices, the SO- coupling attains the more familiar Rashba form [38] (or equivalently Dresselhaus form [39]). For vx6=vy, i.e. when the SO-coupling is anisotropic, the model becomes the dualE( x+ y) Jahn-Teller model [37]. In par- ticular, the ^ z-projection of total angular momentum, ^Jz=^Lz+^z 2, is a constant-of-motion for the isotropic but not for the anisotropic model. More precisely, break- ing of the SO isotropy implies a reduction in symmetry fromU(1) toZ2. Throughout we will use dimensionless parameters where the oscillator energy Eo= h!sets the energy- scale,l=p h=m! the length-scale, and the characteristic time is=!1. We note that for typical experimental setups,!10100 Hz and m(v2 x+v2 y)=h110 kHz. Moreover, in what follows we will refer to pseudo-spin simply as spin. When necessary, we introduce a param- eterhserving as a dimensionless Planck's constant, i.e. hh. In this way, hcontrols the strength of Planck's con- stant and by varying it we can explore how the dynamics depends on  h.3 B. Adiabatic model The large ratio of the SO energy to trapping energy, typicallymv2=h!101000, suggests that a BOA [32] will be valid for experimental implementations. The sep- aration of timescales of the spin and orbital degrees of freedom implies that in some regimes we can factorize the wavefunction as the product of spin and orbital wave- functions. A spin initially aligned with the adiabatic momentum-dependent magnetic eld B(p) will remain locked to that eld at future times, provided the center of mass motion avoids the DP. We then solve for the spin wavefunction at an instantaneous orbital con guration and use this answer to nd an adiabatic potential for the orbital motion. This is in analogy with the traditional BOA, where the electronic and nuclear wavefunctions are approximated as a product, and the electron degrees of freedom instantaneously adjust to the adiabatic potential given by the nuclear degrees of freedom. In our BOA, we have chosen the adiabatic states [32] for the orbital motion to be the spin-helicity states, given by (4). If we project the Hamiltonian into the basis j'i, we arrive at the adiabatic potential ^H() ad=^x2 2+^y2 2+^p2 x 2+^p2 y 2+q v2x^p2x+v2y^p2y:(5) The trap thus takes the role of kinetic energy and (5) can be pictured as a particle in a (dual) adiabatic potential V(^px;^py) =^p2 x 2+^p2 y 2+q v2x^p2x+v2y^p2y: (6) shown in Fig. 1 for both the isotropic (a) and anisotropic (b) cases. We have neglected non-adiabatic corrections arising from the vector potential and the Born-Huang term [40]. For example, an additional scalar potential Vnad(px;py)(vxvy)2(p2 x+p2 y) v2xp2x+v2yp2y2: (7) will emerge from the action of the SO-coupling on the spinorj'i. This term is order Vnadh'jr2 pj'i1=p2. There will also be an additional vector potential term A1=p. The non-adiabatic corrections diverge near the DP, but then fall o rapidly at nite p. The adiabatic approximation, i.e. BOA, will be valid if the particle avoidsp= 0. We will show later that this condition is met if the particle is in the lower band, =1, and has energyE < 0. Imposing the BOA, any state propagating on the lower adiabatic potential will be denoted ( px;py;t), and it is understood that (px;py;t) =(px;py;t)j'i: (8) The real space wave function ( x;y;t ) is given as usual from the Fourier transform of (px;py;t). The time-evolution follows from (px;py;t) =exp i^H() adt (px;py;0). It is also clear that the state (px;py;t) determines the spin orientation which is inherent in the ket-vector j'ii. More explicitly, the time-evolved Bloch vector R(t) = (Rx(t);Ry(t);Rz(t))(h^xi;h^yi;h^zi) (9) takes the form Rx(t) =Z dpxdpyj(px;py;t)j2cos('); Ry(t) =Z dpxdpyj(px;py;t)j2sin('); (10) Rz(t) = 0 in the BOA, and it is remembered that the parameter ' depends on pxandpy. Note that the Bloch vector pre- cesses in the equatorial spin xy-plane. If the wave packet (px;py;t) is sharply localized, a crude approximation for the Bloch vector is given by Rx(t) =vxpx(t)q (vxpx(t))2+ (vypy(t))2; (11) Ry(t) =vypy(t)q (vxpx(t))2+ (vypy(t))2; (12) Rz(t) = 0; (13) where p (t) =R dpxdpyj(px;py;t)j2p with =x; y. III. CLASSICAL DYNAMICS Quantum chaos is often de ned by having an under- lying chaotic classical model. For the full model (1), the spin degrees-of-freedom cannot be eliminated in a straightforward manner in the vicinity of the Dirac point and as a consequence it is not a priori clear what the underlying classical model would be in this regime. On the other hand, in the BOA, the adiabatic Hamiltonian ^H() adcan serve as our classical model Hamiltonian. Still, it should be noted that we assume h^H() adi0, such that the spectrum contains a suciently large number of en- ergies below E= 0. Furthermore, we point out that jus- ti cation of the BOA does not necessarily imply approval of a semi-classical approximation which depends on the system energy and the actual shape of the dual poten- tialV(px;py). Nevertheless, as we will demonstrate in the following, for the chosen parameters, the agreement is indeed very good. The classical equations-of-motion of the Hamiltonian4 V ( )p+V ( )p-V ( )p+ V ( )p-py pypxpx(a) (b) FIG. 1. Adiabatic potentials of the isotropic (a) and anisotropic (b) SO-coupled models. In both gures, the E= 0 plane is the one including the DP at px=py= 0. A nec- essary, but not sucient, condition for the validity of the BOA is that E < 0. In (a), the lower adiabatic potential V(px;py) has the characteristic sombrero shape. By con- sidering an anisotropic SO-coupling, the rotational symme- try is broken and V(px;py) possesses two global minima at (px;py) = (0;vy). −20 0 20−20−1001020 x−20 0 20−50050 xpx(a) (b) FIG. 2. Two examples of classical trajectories (( x(t);Px(t)) for regular (a) and chaotic (b) dynamics. In (a), typical for regular motion the trajectories evolve upon a tori. Contrary, in (b) the trajectory is much more irregular which is char- acteristic for the chaotic evolution. The regular motion is calculated for the SO-coupling strengths vx=vy= 30, and the chaotic motion with vx= 20 andvy= 30. In both cases, the energy is E=192. FIG. 3. Poincar e sections of the Rashba SO-coupled adiabatic model (5) for the intersections y= 0 (a) and py= 0 (b). The initial energy is E=192, the SO-coupling strengths vx=vy= 30, and the number of simulated semi-classical trajectories 18. ^H() adare _x=pxv2 xpxq v2xp2x+v2yp2y; (14) _px=x; (15) _y=pyv2 ypyq v2xp2x+v2yp2y; (16) _py=y: (17) For the Rashba SO-coupling, vx=vy=v, there is one unstable x point ( px;py) = (0;0) and a seam of sta- ble x points p2 x+p2 y=v2, see Fig. 1 (a). For the anisotropic case, vy> vx, there are three unstable x points, (px;py) = (0;0) and (px;py) = (vx;0), while there are two stable x points ( px;py) = (0;vy), see Fig. 1 (b). The classical energy E(x;px;y;py) =p2 x=2 +p2 y=2 + x2=2 +y2=2q v2xp2x+v2yp2ydetermines a hypersurface in phase space for any given energy E(x;px;y;py) =E0. The semi-classical trajectories ( x(t);px(t);y(t);py(t)) live on this surface. For the integrable case, vx=vy, these surfaces form di erent tori characteristic for quasi- periodic motion. As the rotational symmetry is slightly broken,vx6=vy, the tori deforms and the motion loses its quasi-periodic structure [29]. This is the generic crossover from regular to chaotic classical dynamics. As an example of this generic behavior, we show in Fig. 2 two randomly sampled trajectories in the xpx-plane for regular (a) and chaotic (b) evolution. For all results of this section, we solve the set of coupled di erential equa- tions (14) using the Runge-Kutta (4,5) algorithm mod- i ed by Gear's method , suitable for sti equations. We have also numerically veri ed our results employing dif- ferent algorithms [41]. As will be discussed further below, even in the chaotic regime, periodic orbits may persist and will greatly a ect the dynamics, both at a classical and a quantum level [42]. Such orbits are not, however, visible from Fig. 2. The semi-classical behavior of classical dynamical sys- tems is favorable visualized using Poincar e sections [43].5 Corresponding sections for the system (14)-(17) are de- picted in Figs. 3 and 4. In the rst gure we display the Poincar e sections in the xpxplane for the intersections determined by y= 0 (a) or py= 0 (b) of the isotropic model with the SO-coupling amplitudes vx=vy= 30. The initial energy is taken as E=192, well below the DP, consistent with the BOA. In (b), the section de ned bypy= 0, the evolution results in ellipses in the Poincar e section, characteristic of quasi periodic motion. The structure of the Poincar e section for y= 0 (a) is somewhat more complex. This can be understood from the sombrero shape of the adiabatic potential V(px;py); for givenx=x0,px=p0 x,y= 0, and energy E0, there are four possible values of py, and this multiplicity of possiblepy's allow the \curves" in Fig. 3 (a) to cross. It should be noted that any single curve does not cross itself. Furthermore, by adding the pyvalues to Fig. 3 we have veri ed that neither of the corresponding three dimensional curves cross. Figure 4 presents two examples for anisotropic SO- couplings, both with vx= 20 and vy= 30. The quasi-periodic evolution is lost and the dynamics become mixed, with regions of both chaos and regular dynamics. The same conclusions were found in Ref. [44] where a re- lated Jahn-Teller model was studied. The two lower plots consider the same energies as in Fig. 3, i.e. E=192, while for (a) and (b) E=88. Expectedly, the higher energy increases the accessible volume of phase space. For both energies we nd islands free from chaotic tra-jectories. As will be demonstrated in the next section, within these islands the evolution is regular and the sys- tem does not thermalize. The plots also demonstrate clear structures also appearing in the chaotic regimes of the Poincar e sections in which the density of solutions changes. IV. QUANTUM DYNAMICS The idea of this section is to analyze how the corre- sponding quantum evolution is a ected by whether the classical dynamics is regular or chaotic. Of particular im- portance is the long time evolution in which the system state may or may not equilibrate. However, we study also the short time dynamics arising for a localized wave packet traversing the Dirac point. In this regime, clearly the classical results of the previous section does not hold. To study the system beyond the classical approxima- tion, we solve the time-dependent Schr odinger equation, represented by the Hamiltonians (1) or (5), to obtain the corresponding wave function ( x;y;t ) at time t. Note that for the full model (1), the wave function con- tains the spin degree-of-freedom ( x;y;t ) = "(x;y;t )j" i+ #(x;y;t )j#i. The non-equilibrium initial state ap- pears after a quench in the center of the trap. We prepare the system in a quasi-ground state for a shifted trap, and att= 0 suddenly move the trap center to xs=ys= 0, V(x;y) =(xxs)2 2+(yys)2 2; xs6= 0 and=orys6= 0;t<0; xs=ys= 0; t0:(18) By \quasi-ground state" in an anisotropic SO-coupled system, we consider an initial state predominantly pop- ulated in one of the two minima of the adiabatic po- tentialV(px;py). This seems experimentally reasonable where small uctuations will favor one of the two min- ima. For the isotropic case, the phase of ( px;py;t= 0) is taken randomly in agreement with symmetry breaking. Given the evolved states ( x;y;t ), we are interested in the Bloch vector (10) or its components, and the distri- butionsj(px;py;t)j2andj (x;y;t )j2. The numerical calculation is performed employing the split-operator method [45] which relies on factorizing, for short times t, the time-evolution operator into a spatial and a momentum part. For small SO-couplings vxand vy, the method is relatively fast, while as vxand/orvyare increased the time-steps tmust be considerably reduced and the necessary computational power rises rapidly. In addition, for large vxandvy, the grid sizes of position and momentum space must be increased, which also in- creases the computation time. Thus, we will limit the analysis to SO-couplings vx; vy30. Furthermore, we have found by convergence tests that the full model (1)requires much smaller time-steps tthan the adiabatic one (5), and most of our simulations will therefore be re- stricted to energies E < 0 for which the BOA is justi ed. The full quantum simulations are complemented by the semi-classical truncated Wigner approximation (TWA), which has turned out very ecient in order to re- produce quantum dynamics [46]. The TWA considers a set ofNdi erent initial values ( xi;yi;pxi;pyi) ran- domly drawn from the distributions j (x;y;0)j2and j(px;py;0)j2. These are then propagated according to the classical equations-of-motion (14). The propagated set (xi(t);yi(t);pxi(t);pyi(t)) gives the semi-classical dis- tributions, from which expectation values can be evalu- ated. A. Short time dynamics Before investigating the prospects of thermalization, we rst consider short time dynamics, by which we mean time-scales where the wave packet remains localized. In6 FIG. 4. Poincar e sections of the anisotropic SO-coupled adia- batic model (5) for y= 0 (a) and (c), and for py= 0 (b) and (d). The initial energies are E=88 (a) and (b), andE=192 (c) and (d), and the SO-coupling strengths vx= 20 andvy= 30 for both cases. The corresponding maxi- mum Lyaponov exponents have been derived to 0:12 and = 0:090 respectively. The number of semi-classical trajec- tories is the same as for Fig. 3, namely 18. this respect, it is tempting to think of the dynamics as semi-classical. However, in the vicinity of the the DP any classical description would fail. Equivalently, the spin degrees-of-freedom will show large uctuations which are dicult to capture classically. The short time dynamics is consequently most interesting for situations with ener- giesE > 0 where both the semi-classical approximation and the BOA break down, implying that the simulation 0 5 10 15 20 25 30−101 tRα−101Rα (a) (b)FIG. 5. Bloch vector components Rx(dashed lines) and Ry (solid lines). For the upper plot (a), the trap has been dis- placed in th y-direction,xs= 0 andys= 28, while in the lower plot (b) the displace direction is the perpendicular, xs= 28 andys= 0. In both gures, vx= 10 andvy= 15, and the average energy E280. is performed using the full model Hamiltonian (1). For these energies, the wave packet can traverse the DP and population transfer between the two adiabatic potentials V(^px;^py) typically occurs. It is known that such non- adiabatic transitions can play important roles for the dy- namics, and that the actual transition probabilities be- tween the two potentials may be extremely sensitive to small uctuations in the state [31, 48]. In this subsection we especially address such non-adiabatic e ects. There are indeed several relevant time-scales in the dy- namics: (i) The spin precession time Tspgives the typical time for spin evolution and is proportional to the e ective magnetic eldjB(p)j, (ii) the classical oscillation period Tcl= 2, and (iii) the thermalization time Tth, which estimates the time it takes for the system to thermalize, i.e. when expectation values become approximately time independent. Normally, the magnitudes of these times follow the list above (in growing order), except in the vicinity of the DP where TspTclor evenTspTcl very close to the DP. While the rst two are well de ned, de ning the last one is non-trivial. We can say that ( i) and (ii) characterizes short time-time scales, and ( iii) long time-scales. As will be numerically demonstrated, the thermalization time turns out to scale as log( h1)=, wherehis the e ective dimensionless Planck's constant andthe maximum Lyaponov exponent. This suggests that the thermalization time agrees with the Ehrenfest time TE= log(V=h)=; (19) withVthe e ective occupied phase space volume. TE is also the typical time-scale where semi-classical (TWA) expectation values no longer agree with quantum expec- tation values, which can be seen as a breakdown of Ehren- fest's theorem [49]. From the form of the non-adiabatic coupling (7), it fol- lows that transitions between the adiabatic states (4) are restricted to the vicinity of the DP. These non-adiabatic transitions are manifested as rapid changes in the Bloch7 vector (10). In Fig. 6 we present two examples of the Bloch vector evolution (in both examples Rz(t)0). In Fig. 6 (a), the trap has been shifted in the y-direction. For short times, the shift of the trap induces a build-up of momentum in the opposite y-direction as a consequence of the Ehrenfest theorem. This adds with the non-zero y-component of momentum before the quench. The av- erage momentum in the x-direction remains zero and as a consequence Rx(t)0, see Eq. (11). These dynamics change qualitatively if the trap is shifted in the x-direction instead of the y-direction. For suciently large shifts of xs, the wave packet will set o along the adiabatic potentials and encircle the DP. The spin dynamics should therefore not display the same type of \jumps" that appear when the wave packet traverses the DP. Moreover, since the average momentum in the x-direction is in general non-zero, Rx(t) will also be non- zero. The results are demonstrated in Fig. 6 (b). Com- pared to the rst example in (a), the wave packet does not spend much time near the DP so the wave packet delocalization occurs more slowly. To a large extent the evolution is driven by harmonicity, in contrast to the ex- ample of Fig. 6 (a) where the anharmonicity of the Born- Huang term, and the non-adiabatic transitions near the DP, push the system away from semi-classical evolution. The gure demonstrates how the dynamics can depend on the initial conditions, in both (a) and (b), E280 but the wave packet broadening starts earlier in (a) than in (b). This type of state-dependence has been discussed in Ref. [47]; generically there is a period tswhere the width of the wave packet stays nearly constant, followed by a rapid broadening. The time-scale tsdepends strongly on the initial conditions, while the proceeding evolution aftertsseems pretty generic for chaotic systems. B. Long time dynamics; thermalization Whenever we consider an anisotropic SO-coupling, vx6=vy, from the Figs. 3 and 4 it is clear how the adiabatic classical model becomes chaotic. Beyond the adiabatic model, it has been shown [50] that the full anisotropic model, i.e., E( x+ y) Jahn-Teller model, is chaotic in the sense of level repulsion [51] of eigenen- ergies. For the isotropic E"Jahn-Teller model, on the other hand, the level repulsion e ect is not as evi- dent, however a weak repulsion also in this model signals emergence of quantum chaos [52]. The goal of this subsection is to study the long time dynamics of the system; speci cally if equilibration oc- curs, and if so, does the equilibrated state mimic a ther- mal state. A distinguishing property of thermal states is, for example ergodicity, i.e., the distributions j (x;y;t )j2 andj(px;py;t)j2spread out over their accessible energy shells. Moreover, for a thermally equilibrated state, the distributions show seemingly irregular interference struc- tures on scales of the order of the Planck cells, which normally become even ner in the Wigner quasi distri- FIG. 6. (Color online) Distributions j (x;y;t f)j2((a) and (c)) andj(px;py;tf)j2((b) and (d)) at tf= 400 for the Rashba SO-coupled model. At time t= 0, the trap is sud- denly displaced from x0=y0= 16 tox0=y0= 0. The initial ground state is then quenched into a localized excited state. The upper two plots (a) and (b) display the results from full quantum simulations of the adiabatic model (5), while the lower plots (c) and (d) show the corresponding semi-classical TWA distributions. The average semi-classical energy E192 with a standard deviation E22. The dimensionless SO-coupling strengths vx=vy= 30. bution [53{55]. Non-thermalized states, on the contrary, typically leave much more regular traces of quantum in- terference in their distributions. While such often sym- metrical structures are absent for thermalized states, we will demonstrate that thermalized distributions may still show clear density uctuations on scales larger than the Planck cells. These are examples of quantum scars and they are remnants of classical periodic orbits [42]. We begin by considering the adiabatic isotropic model withvx=vy= 30, and trap shifts xs=ys= 16. Af- ter a quench of the trap position, the initial energy is E=h^H() adi 192. This energy corresponds to the energy of the Poincar e section presented in Fig. 3. The resulting distributions are shown in Fig. 7 (a) and (b) after a propagation time tf= 400 . The nal time tfap- proximates 60 classical oscillations. Both the real space densityj (x;y;t )j2and momentum density j(px;py;t)j2 reveal clear interference patterns as anticipated. The DP at the origin ( px;py) = (0;0) repels the wave function forming a \hole." The lack of zero momentum states in- duces a mass ow in real space and a similar \hole" in its distribution. The classically energetically accessible8 FIG. 7. (Color online) Same as Fig. 6 but for the anisotropic SO-coupled model with vx= 20 andvy= 30. The largely populated regions are so called quantum scars and derive from properties of the underlying classical model, i.e. they are not outcomes of some coherent quantum mechanism. regions are given by x2+y22Emax+v2 y; p2 x+p2 y2q v2xp2x+v2yp2y2Emax;(20) whereEmaxis the maximum energy component notice- ably populated by the state. The quantum results are compared with the TWA dis- tributions displayed in the lower plots (c) and (d) of the same Fig. 7. The same kind of ring-shape is obtained, and the concentration in density appears at the same lo- cations for both the quantum and classical simulations. Expectedly, the quantum interference taking place within the wave packet is not captured by the TWA. This fol- lows since single semi-classical trajectories are treated independently, i.e. added incoherently, while a quantum wave packet must be considered as one entity. For a TWA approach of the full isotropic E"Jahn-Teller model (1) we refer to Ref. [56]. The situation is drastically changed when we break the rotational U(1) symmetry by assuming vx6=vy. The result for low initial energy is depicted in Fig. 7 (a) and (b). The energy is comparable to the potential barrier separating the two minima in the adiabatic potentials, and as a consequence, the wave packet is predominantly localized in the left minima. The density modulations seems now much more irregular in comparison to Fig. 6. In the seemingly random density distribution, some clear density maxima emerge, both in momentum as well as in real space. These density accumulations derive from periodic orbitals of the underlying classical model and FIG. 8. (Color online) Same as Fig. 7 but for an initial energy E > 0. The dimensionless SO-couplings vx= 14 andvy= 21, while the shifts xs=ys= 16 giving an average energy E= h^HSOi36:5. 0100200300∆x(t) 0 10 20 30 40 50 60 70 800100200300 t∆x(t+δ)(a) (b)h FIG. 9. Examples of the phase space area  x(t) for di erent h-values (h= 1;2;3; :::; 10). The upper plot (a) gives  x(t) without shifting the time, while for the lower one (b) time has been shifted by = log(h)=. The arrow indicates increasing h-values. It is clear how the spread in  x(t) between di erent hvalues is suppressed when we shift the time. The trap shifts xs=ys= 19 resulting in an energy E88. The maximum Lyaponov exponent = 0:18. are termed quantum scars [42, 57, 58]. The appearance of scars is an example of the classically chaotic model leaving a trace in its quantum counterpart. The scars are also captured in the semi-classical TWA, shown in Fig. 7 (c) and (d), supporting their classical origin. When we shift the trap for larger values on xsand ys, the energy is increased and at some point the BOA breaks down. An example, obtained from integrating the full model (1), is presented in Fig. 8. For these higher energies there are no signs of quantum scars. As for the situation of Fig. 7, the spread of the wave packet and the irregular interference patterns indicates thermalization. This far we have demonstrated thermalization for the anisotropic SO coupled model, but not discussed corre- sponding time-scales. One related question is how the evolution of various expectation values scale with h(di- mensionless Planck constant). It has been argued that the Ehrenfest time, Eq. (19), can be a measure of the thermalization time [31]. We will now explore how the phase space area  (t) =  p ( =x; y), where9 −7 −6.5 −6 −5.5 −5 −4.5 −4−0.02−0.0100.010.020.030.04 x|ψ(x,0)| FIG. 10. (Color online) Sections of j (x;y= 0)jfor di erent values on the dimensionless Planck's constant h:h= 1 (black solid line), h= 2 (blue dotted line), and h= 3 (red dashed line). The nal time tf= 80,xs=ys= 16, and vx= 14 andvy= 20. As a comparison between classical and quan- tum results, we also include the TWA results as a green solid line, calculated for h= 1. The green line has been shifted downward with 0.02 for clarity.  and p are the variances of ^ and ^p respectively, evolves for di erent values of h. Since  x(t) and y(t) behave similarly we focus only on  x(t). For thermaliza- tion, x(t)y(t) is an e ective measure of the covered phase space volume, and for large times tit should more or less approach the accessible phase space volume as the distribution spreads over the whole energy shell. We have chosen to study  x(t) since it uctuates relatively little before reaching its asymptotic value. In Fig. 9 (a) we dis- play x(t) for 10 di erent values on hranging from h= 1 toh= 10. The arrow in the plot shows the direction of increasingh's. As is seen, by increasing hthe wave packet broadening starts earlier and the state equilibrates faster. If the Ehrenfest time TEsets the typical time scale in the process, by shifting the time with = log(h)=we should recover a \clustering" of the curves. This is indeed ver- i ed in Fig. 9 (b) where the curves have been shifted in time by. The corresponding Lyaponov exponent has been optimized in order to minimize the spread in the curves. The obtained value = 0:18 is somewhat larger than the numerically calculated one = 0:12 but still of the same order. The picture also makes clear that the wave packet broadening kicks in after some time tsas anticipated above. The route to thermalization can typically be di- vided into; ( i) a classical drift, and ( ii) quantum dif- fusion [31]. The role of the quantum di usion for ther- malization was analyzed in Ref. [31], where it was found to \smoothen" the phase space distributions preventing sub-Planck structures. For the classical drift there is no lower bound on the neness of density structures that can form, and characteristic for classical chaotic dynamics is that ever ner formations build-up as a result of the typ- ical \stretching-and-folding" mechanism. However, in aquantum chaotic system, when the structures reach the Planck cell regime, the quantum pressure becomes too strong and the quantum di usion then prevents any fur- ther structures to form. Thus, Planck's constant sets a lower bound for the uctuations in the distributions. This quantum smoothening is demonstrated in Fig. 10, where we plot a section of j (x;y= 0)jfor di erent values on the scaled dimensionless Planck's constant h(= 1;2;3 for black, blue, and red lines respectively). The e ect is clearly seen in the gure. A similar pattern is found (not shown) also for the momentum distributions. For the classical system, corresponding to h= 0, there is no lower limit on how ne the structures can be. We indicate this by also plotting the TWA results in the same gure as a green line (note that the green line has been shifted downward in order to separate it from the quantum re- sults). The number of trajectories used for the gure is 250 000, and if we would like to produce ner structures (by propagating the system for longer times) we would need many more trajectories and the simulation would rapidly become very time consuming. Related to the above discussion a note on quantum phase space distributions is in order. It is well known that sub-Planck structures are common in the Wigner distribution [53]. This is not contradicting any quan- tum uncertainty relation. After all, the Wigner distribu- tion is not a proper probability distribution, despite the fact that its marginal distributions reproduce the cor- rect real and momentum space probability distributions. The Husimi Q-function, while not possessing the proper marginal distributions, is positive de nite and lacking singularities, and it is indeed found that the Q-function does not support sub-Planck structures [60]. We nish this subsection by analyzing the dynamics in the islands of the Poincar e sections of Fig. 4 where the classical theory predicts regular evolution. From Fig. 4 (c) we have that for px20 andxy0 the evolu- tion should be regular. We can achieve such a situation by using the quench-shifts xs= 20 andys= 0. As for the examples above, we propagate the state for a time tf= 400, and the resulting distributions are given in Fig. 11. The striking di erence with Figs. 7 and 8 is evi- dent; no irregular structure is apparent, but clear regular interference patterns are. We have veri ed that the in- terference structure prevails also after doubling the time, tf= 800. C. Proposed experimental realization Much of the above dynamics can be observed in a sys- tem of cold atoms with synthetic SO-coupling, for exam- ple, a system of87Rb with a synthetic eld induced by the 4-level scheme [33]. In this system, the recoil energy Er=mv2h50 kHz. The synthetic eld limits the lifetime of the experiment to tl1s[4, 10]. To push the experiment into the long time regime, we will use a trapping frequency of !=2= 30 Hz. These parame-10 FIG. 11. (Color online) Same as Fig. 7 but for the shifts xs= 20 andys= 0. For the given dimensionless parameters, the initial state is such that its dynamics should be regular according to the corresponding Poincar e section, Fig. 4. The energy E250. ters will give a dimensionless value of vy=q Er h!11, withvxtunable between 0 and 11. The large trapping frequency will provide a sucient number of oscillations for thermalization to occur. We could consider values of vy30 by decreasing the trapping frequency to 10Hz, but then the lifetime of the system may be at the boarder for thermalization. The condensate can be adiabatically loaded to one of the two states at the bottom of the momentum-space potential, de ned by p=mvy^y. The quench can then be preformed by shifting the minimum of the real-space trapping potential. We then let the system evolve until we reach either the thermalization time, or the lifetime of the experiment. The momentum distribution can be measured with a destructive time-of- ight (TOF) mea- surement [4, 10], which should reveal thermalization as well as signatures of quantum scars. Repeated experi- mental measurements allow for time-resolved calculation of expectation values. Similarly, the quantum spin jumps near the DP, as discussed in Sec. IV A, can be observed using a spin-resolved TOF measurement. As a nal remark, for a weakly interacting gas we work near a Feschbach resonance [61]. However, for re- alistic parameters [62], we estimate a scattering length as3109m,N5105atoms, and a transverse harmonic trapping frequency !z100 Hz. For these parameters, the characteristic scale of the non-linearity ish1kHz, which is smaller than the recoil energy above, suggesting the non-linear term will play only a minor role. We have numerically veri ed that the results do not change qualitatively by solving the corresponding non-linear Gross-Pitaevskii equation. Indeed, we nd the deviations with a non-linearity are not large enough to be seen by eye. V. CONCLUSIONS In this paper we studied dynamics, deriving from a quantum quench, in anisotropic SO-coupled cold gases, focusing primary on aspects arising from the fact thatthe underlying classical model is chaotic. The evolution of the initially localized wave packet on its way to equili- bration has been analyzed, and we have shown how a clas- sical period of limited spreading is followed by a collapse regime dominated by rapid spreading. After the collapse period, the wave packet is maximally delocalized, but still possesses quantum interference structures. At the Ehrenfest time, the state has approximately equilibrated as is seen in the decay of expectation values, as well as seemingly irregular density uctuations both in real and momentum space. We showed that the ne structure of these uctuations are limited by the quantum di usion, and thereby the size of the Planck's constant h. For the isotropic model, after the collapse no thermalization is found, as is expected from the integrability of the under- lying classical model. For smaller energies, when the wave packet predom- inantly populates one of the dual potential wells, ther- malization is again seen. Here, however, an additional phenomenon appears in terms of quantum scars. These density enhancements emerge along classically periodic orbits. They are classical in nature and long lived. Quan- tum scars have also been studied in di erent cold atom settings; atoms in an optical lattice and con ned in an anisotropic harmonic trap [58]. The results on thermal- ization presented in this work is most likely also applica- ble to the set-up of Ref. [58]. We also demonstrated that for certain ne tuned initial states, the dynamics stays regular even in the anisotropic model. In the classical picture, these solutions correspond to the ones belong- ing to regular islands in the otherwise chaotic Poincar e sections. We argue that the present system is ideal for studies of quantum chaos and quantum thermalization for nu- merous reasons. The system parameters can be tuned externally by adjusting the wavelength of the lasers in- ducing the SO-coupling, and as we discussed in Sec. IV C the SO dominated regime is reachable in current exper- iments. Moreover, both state preparation and detection are relatively easily performed in these setups. Equally important, the system is well isolated from any environ- ment and coherent dynamics can be established up to hundreds of oscillations which is well beyond the themral- ization time. The energy of the state is simply controlled by the trap displacement, and it should for example be possible to give the system small energies such that the atoms reside mainly in one potential well where quantum scars develop. We nish by pointing out that the present model is also di erent from most earlier studies on quantum thermal- ization [18, 24] in the sense that the dynamics is essen- tially \single-particle" and not arising from many-body physics. Related to this, we have numerically veri ed that adding a non-linear term gj (x;y;t )j2to the Hamil- tonian does not change our results qualitatively for mod- erate realistic interaction strengths g. In order to enter into the regime where interaction starts to a ect the re- sults, one would need a condensate with a large number11 of atoms (millions of atoms) or alternatively externally tune the scattering length via the method of Feshback resonances. ACKNOWLEDGMENTS The authors thank Ian Spielman for helpful comments. SFB/TR 12 is acknowledged for nancial support. JLacknowledges Vetenskapsr adet (VR), DAAD (Deutscher Akademischer Austausch Dienst), and the Royal Re- search Council Sweden (KVA) for nancial help. BA acknowledges the sponsorship of the US Department of Commerce, National Institute of Standards and Technol- ogy, and was supported by the National Science Founda- tion under Physics Frontiers Center Grant PHY-0822671 and by the ARO under the DARPA OLE program. [1] M. Lewenstein, A. Sanpera, V. Ahu nger, B. Damski, A. Sen(De), and U. Sen, Adv. Phys. 56, 243 (2007); I. Bloch, J. Dalibard, and W. Zwerger, Rev. Mod. Phys. 80, 885 (2008). [2] M. Greiner, O. Mandel,T. Esslinger, T. W. H ansch, and I. Bloch, Nature 415, 39 (2002). [3] C. J. Pethick and H. Smith, Bose-Einstein Condensa- tion in Dilute Gases , (Cambridge University Press, Cam- bridge, 2008). [4] Y. J. Lin, R. L. Compton, K. Jimenez-Garcia, J. V. Porto, and I. B. Spielman, Nature 462, 628 (2009); Y. J. Lin, R. L. Compton, A. R. Perry, W. D. Phillips, J. V. Porto, and I. B. Spielman, Phys. Rev. Lett. 102, 130401 (2009). [5] J. Dalibard, F. Gerber, G. Juzelunas, and P. Ohberg, Rev. Med. Phys. 83, 1523 (2011). [6] G. Juzeli unas, J. Ruseckas, P. Ohberg, and M. Fleis- chhauer, Phys. Rev. A 73, 025602 (2006). [7] N. R. Cooper and J. Dalibard, Euro. Phys. Lett. 95, 66004 (2011). [8] M. Sigrist and K. Ueda, Rev. Mod. Phys. 63, 239 (1991); I. Zutic, J. Fabian, S. Das Sarma, Rev. Mod. Phys. 76, 323 (2004) [9] M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010). [10] Y. J. Lin, K. Kimenez-Garcia, and I. Spielman, Nature 471, 83 (2011). [11] G. Juzeli unas, J. Ruseckas, and J. Dalibard, Phys. Rev. A81, 053403 (2010). [12] B. M. Anderson, G. Juzeli unas, I. B. Spielman, and V. M. Galitski, Phys. Rev. Lett. 108, 235301 (2012). [13] H. Hu, B. Ramachandhran, H. Pu, and X.-J. Liu, Phys. Rev. Lett. 108, 010402 (2012). [14] S. Sinha, R. Nath, and L. Santos, Phys. Rev. Lett. 107, 270401 (2011). [15] J. Radic, T. A. Sedrakyan, I. B. Spielman, and V. Galit- ski, Phys. Rev. A 84, 063604 (2011); H. Hu, H. Pu, and X.-J. Liu, Phys. Rev. Lett. 108, 010402 (2012); H. Hu and X.-J. Liu, Phys. Rev. A 85, 013619 (2012). [16] J. Larson and E. Sj oqvist, Phys. Rev. A 79, 043627 (2009); Y. Zhang, L. Mao, and C. Zhang, Phys. Rev. Lett. 108, 035302 (2012). [17] S. K. Ghosh, J. P. Vyasanakere and V. B. Shenoy, Phys. Rev. A 84, 053629 (2011) [18] A. Polkovnikov, K. Sengupta, A. Silva, and M. Vengalat- tore, Rev. Mod. Phys. 83, 863, (2011); V. I. Yukalov, Laser Phys. Lett. 8, 485 (2011). [19] D. Leibfried, R. Blatt, C. Monroe, and D. Wineland, Rev. Mod. Phys. 75, 281 (2003); S. Haroche and J.-M. Rai-mond, Exploring the Quantum , (Oxford University Press, Oxford, 2006); R. Islam, E. E. Edwards, K. Kim, S. Ko- renblit, C. Noh, H. J. Carmichael, G. D. Lin, L. M. Duan, C. C. J. Wang, J. K. Freericks, and C. Monroe, Nature Commun. 2, 377 (2011). [20] J. M. Fink, M. G oppl, M. Baur, R. Bianchetti, P. J. Leek, A. Blais, and A. Wallra , Nature 454, 315 (2008). [21] H. Walther, B. T. H. Varcoe, T. H. Benjamin B. G. En- glert, and T. Becker, Rep. Prog. Phys. 69, 1325 (2006). [22] Q. Y. He, M. D. Reid, B. Opanchuk, R. Polk- inghorne, L. E. C. Rosales-Zarate, and P. D. Drummond, arXiv:1112.0380. [23] J. M. Deutsch, Phys. Rev. A 43, 2046 (1991); M. Sred- nicki, Phys. Rev. E 50, 888 (1994). [24] M. Rigol, V. Dunjko, and M. Olshanii, Nature 452, 854 (2008). [25] T. Kinoshita, T. Wenger, and S. D. Weiss, Nature 440, 900 (2006). [26] V. Milner, J. L. Hanssen, W. C. Campbell, and M. G. Raizen, Phys. Rev. Lett. 86, 1514 (2001); M. F. Ander- sen, A. Kaplan, T. Gr unzweig, and N. Davidson, Phys. Rev. Lett. 97, 104102 (2006); M. Gring, M. Kuhnert, T. Langen, T. Kitagawa, B. Rauer, M. Schreitl, I. Mazets, D. A. Smith, E. Demler, and J. Schmiedmayer, Science 337, 1318 (2012). [27] S. Chaudhury, A. Smith, B. E. Anderson, S. Ghose, and P. S. Jessen, Nature 461, 768 (2009). [28] C. Gogolin, M. P. M uller, and J. Eisert, Phys. Rev. Lett. 106, 230502 (2011). [29] R. C. Hilborn, Chaos and Nonlinear Dynamics , 2nd ed. (Oxford University Press, Oxford, 1980). [30] V. A. Yurovsky and M. Oslhanii, Phys. Rev. Lett. 106, 025303 (2011); M. Olshanii, K. Jacobs, M. Rigol, V. Dun- jko, H. Kennard, and V. A. Yurovsky, Nature Commun. 3, 641 (2012). [31] A. Altland and F. Haake, Phys. Rev. Lett. 108, 073601 (2012); A. Altland and F. Haake, New J. Phys. 14, 073011 (2012). [32] A. Bohm, A. Mostafazadeh, H. Koizumi, Q. Niu, and J. Zwanziger, The Geomtric Phases in Quantum Systems , (Springer Verlag, Berlin, 2003); M. Baer, Beyond Born- Oppenheimer , (John Wiley & Sons, New Jersey, 2006). [33] D. L. Campbell, G. Juzeli unas, and I. B. Spielman, Phys. Rev. A 84025602 (2011). [34] Z. F. Xu, L. You, Phys. Rev. A 85, 043605 (2012) [35] J. Ruseckas, G. Juzeli unas, P. Ohberg, and M. Fleis- chhauer, Phys. Rev. Lett. 95, 010404 (2005); T. D. Stanescu, C. Zhang, and V. Galitski, Phys. Rev. Lett. 99, 110403 (2007); T. D. Stanescu, B. Anderson, and12 V. Galitski, Phys. Rev. A 78, 023616 (2008). [36] H. C. Longuet-Higgens, U. Opik, M. H. L. Pryce, and R. A. Sack, Proc. R. Soc. London Ser. A 244, 1 (1958); J. Larson, Phys. Rev. A 78, 033833 (2008). [37] R. Englman, The Jahn-Teller E ect in Molecules and Crystals , (Wiley, New York, 1972); G. Grosso and G. P. Parravicini, Solid State Physics , (Academic Press, 2003). [38] Y. A. Bychkov and E. I. Rashba, J. Phys. C 17, 6039 (1984). [39] G. Dresselhaus, Phys. Rev. 100, 580 (1955). [40] M. V. Berry and R. Lim, J. Phys. A: Math. Gen. 23, L655 (1990). [41] Regular Runge-Kutta (4,5), and the Adams-Bashforth- Moulton PECE solver. [42] E. J. Heller, Phys. Rev. Lett. 53, 1515 (1984); M. V. Berry, Proc. Roy. Soc. London Ser. A 423, 219 (1989); L. Kaplan and E. J. Heller, Annals of Phys. 264, 171 (1998); L. Kaplan, Nonlinearity 12, 1 (1998). [43] S. H. Strogatz, Nonlinear Dynamics and Chaos , (West- view Press, Cambridge, 2000). [44] R. S. Markiewicz, Phys. Rev. E 64, 026216 (2001). [45] M. D. Fleit, J. A. Fleck, and A. Steiger, J. Comput. Phys. 47, 412 (1982). [46] A. Polkovnikov, Annals of Phys. 325, 1790 (2010). [47] W.-G. Wang and B. Li, Phys. Rev. E 66, 056208 (2002). [48] D. Wang, T. Hansson, A. Larson, H. O. Karlsson, and J. Larson, Phys. Rev. A 77, 053808 (2008). [49] S. Habib, K. Shizume, and W. H. Zurek, Phys. Rev. Lett. 80, 4361 (1998). [50] E. Majernikova and S. Shpyrko, Phys. Rev. E 73, 057202 (2006). [51] F. Haake, Quantum Signatures of Chaos , (Springer- Verlag, Berlin, 2010). [52] E. Majernikova and S. Shpyrko, Phys. Rev. E 73, 066215(2006). [53] W. H. Zurek, Nature 412, 712 (2001). [54] Ph. Jacquod, I. Adagideli, and C. W. Beenakker, Phys. Rev. Lett. 89, 154103 (2002). [55] Normally, the sub-Planck cell structures are studied in the phase space distributions. However, there exist sev- eral di erent phase psace distributions having di erent properties. For example, the sub-Planck cell structures typically appear in the Wigner distributions, while in the HusimiQ-function, being strictly positive, small uctua- tions have been \averaged out" and it is normally much smoother than the Wigner distribution. [56] J. Larson, E. N. Ghassemi, and A. Larson, arXiv:1111.4647. [57] P. B. Wilkinson, T. M. Fromhold, L. Eaves, F. W. Sheard, N. Miura, and T. Takamasu, Nature 380, 608 (1996); A. M. Burke, R. Akis, T. E. Day, G. Speyer, D. K. Ferry, and B. R. Bennett, Phys. Rev. Lett. 104, 176801 (2010). [58] T. M. Fromhold, C. R. Tench, S. Bujkiewicz, P. B. Wilkinson, and F. W. Sheard, J. Opt. B: Quant. Semi- class. Opt. 2, 628 (2999); R. G. Scott, S. Bujkiewicz, T. M. Fromhold, P. B. Wilkinson, and F. W. Sheard, Phys. Rev. A 66, 023407 (2002). [59] J. C. Sprott, Chaos and Time-Series Analysis , (Oxford University Press, Oxford, 2003). [60] D. J. O'Dell and J. Larson, To be submitted. [61] G. Roati, M. Zaccanti, C. D'Errico, J. Catani, M. Mod- ugno, A. Simoni, M. Inguscio, and G. Modugno, Phys. Rev. Lett. 99, 010403 (2007). [62] R. A. Williams, L. J. LeBlanc, K. Jimenez-Garcia, M. C. Beeler, A. R. Perry, W. D. Phillips, and I. B. Spielman, Science 335, 314 (2012).
2207.06347v1.Giant_orbital_Hall_effect_and_orbital_to_spin_conversion_in_3d__5d__and_4f_metallic_heterostructures.pdf
PHYSICAL REVIEW RESEARCH 4, 033037 (2022) Giant orbital Hall effect and orbital-to-spin conversion in 3 d,5d,a n d4 fmetallic heterostructures Giacomo Sala*and Pietro Gambardella† Department of Materials, ETH Zurich, 8093 Zurich, Switzerland (Received 4 April 2022; revised 21 June 2022; accepted 23 June 2022; published 13 July 2022) The orbital Hall effect provides an alternative means to the spin Hall effect to convert a charge current into a flow of angular momentum. Recently, compelling signatures of orbital Hall effects have been identified in 3 d transition metals. Here, we report a systematic study of the generation, transmission, and conversion of orbitalcurrents in heterostructures comprising 3 d,5d,a n d4 fmetals. We show that the orbital Hall conductivity of Cr reaches giant values of the order of 5 ×10 5[¯h 2e]/Omega1−1m−1and that Pt presents a strong orbital Hall effect in addition to the spin Hall effect. Measurements performed as a function of thickness of nonmagnetic Cr, Mn, andPt layers and ferromagnetic Co and Ni layers reveal how the orbital and spin currents compete or assist eachother in determining the spin-orbit torques acting on the magnetic layer. We further show how this interplaycan be drastically modulated by introducing 4 fspacers between the nonmagnetic and magnetic layers. Gd and Tb act as very efficient orbital-to-spin current converters, boosting the spin-orbit torques generated by Cr by afactor of 4 and reversing the sign of the torques generated by Pt. To interpret our results, we present a generalizeddrift-diffusion model that includes both spin and orbital Hall effects and describes their interconversion mediatedby spin-orbit coupling. DOI: 10.1103/PhysRevResearch.4.033037 I. INTRODUCTION The interconversion of charge and spin currents underpins a variety of phenomena and applications in spintronics, in-cluding spin-orbit torques, spin pumping, the excitation ofmagnons, and the tuning of magnetic damping [ 1,2]. The spin Hall effect (SHE) mediates this interconversion through thecombination of intrinsic and extrinsic scattering processes,all of which require sizable spin-orbit coupling [ 3]. Recent theoretical work has shown that the intrinsic SHE is accompa-nied by a complementary process involving the orbital angularmomentum, the so-called orbital Hall effect (OHE), whichconsists in the flow of orbital momentum perpendicular tothe charge current [ 4–10]. According to theoretical calcula- tions, the OHE is more common and fundamental than theSHE because it does not require spin-orbit coupling and canthus occur in a wider range of materials. The intrinsic SHEthen emerges as a by-product of the OHE resulting from theorbital-to-spin conversion in materials with nonzero spin-orbitcoupling. In this case, the spin Hall conductivity has the samesign as the product between the orbital conductivity and theexpectation value of spin-orbit coupling: σ S∼σL/angbracketleftL·S/angbracketright.T h e OHE was first predicted in 4 dand 5 dtransition elements [11,12] and recently in light metals [ 4] and their interfaces *giacomo.sala@mat.ethz.ch †pietro.gambardella@mat.ethz.ch Published by the American Physical Society under the terms of the Creative Commons Attribution 4.0 International license. Further distribution of this work must maintain attribution to the author(s) and the published article’s title, journal citation, and DOI.[13] as well as in two-dimensional (2D) materials [ 14,15]. The theoretical orbital Hall conductivity of light elements iscomparable to or even larger than the spin Hall conductivityof Ta, W, and Pt, which provide a strong SHE [ 4]. The OHE is thus intrinsically more efficient than the SHE, and orbitalcurrents are expected to contribute to magnetotransport effectssuch as the anisotropic, spin Hall, and unidirectional mag-netoresistance as well as spin-orbit torques [ 6,16–20]. The ubiquity and strength of the OHE, besides making it funda-mentally interesting, broaden the material palette available forspintronic applications and provide an additional handle tooptimize the efficiency of spin-orbit torques. Yet, differentlyfrom spins, nonequilibrium orbital currents do not coupledirectly to the magnetization of magnetic materials and cantorque magnetic moments only indirectly through spin-orbitcoupling [ 19–21]. Optimizing the orbital-to-spin conversion is thus a prerequisite for taking advantage of large orbitalcurrents. The prediction of the OHE in light elements has triggered intense research on current-induced orbital effects. Recentexperiments have identified signatures of the OHE in ma-terials with low [ 18,19] and high [ 20] spin-orbit coupling and revealed its contribution to spin-orbit torques [ 16,18], whose strength can be tuned by improving the orbital-to-spinconversion ratio [ 19,22]. However, experimental values of the orbital Hall conductivity are smaller than theoretical estimates[16,18,19,23], and a systematic investigation of orbital effects as a function of the type and thickness of nonmagnetic, ferro-magnetic, and spacer layers is still missing. Here, we present a comprehensive study of the interplay of the OHE and SHE in structures combining different light andheavy nonmagnetic metals (NM =Cr, Mn, Pt), ferromagnets (FM=Co, Ni), and rare-earth spacers ( X=G d ,T b ) .W e 2643-1564/2022/4(3)/033037(14) 033037-1 Published by the American Physical SocietyGIACOMO SALA AND PIETRO GAMBARDELLA PHYSICAL REVIEW RESEARCH 4, 033037 (2022) (a) (b) (a)LS MTSTLSJCxz L SE MTSTLSLSJC E NM FML S FIG. 1. (a) The spin Hall effect and orbital Hall effect induced by an electric field Ein a nonmagnet (NM) produce spin ( TS)a n d orbital ( TLS) torques on the magnetization Mof an adjacent ferro- magnet (FM). The strength of the torques depends on the intensityof the spin and orbital currents and on the spin-orbit coupling of the ferromagnet. The schematic shows the direction of the induced spin (S, blue dots and crosses) and orbital ( L, red circling arrows) angular momenta when the spin and orbital Hall conductivities σ S,L>0. (b) The insertion of a spacer layer may increase the orbital torque relative to the spin torque by converting the orbital current (red) intoa spin current (blue) prior to their injection into the ferromagnet. provide evidence of the OHE in Pt and Mn and report giant values of the orbital Hall conductivity in Cr, which extrapolateto the theoretical limit of 10 6[¯h 2e](/Omega1m)−1in Cr films thicker than the orbital diffusion length [ 4], which we estimate to be/greaterorsimilar20 nm. Because of the simultaneous presence of strong OHE and SHE in Pt and Cr, we argue that experimental resultsare best described by a combined spin-orbital conductivityrather than by separating the two effects. We show that theinterplay between orbital and spin currents can be tailoredby varying the thickness of the ferromagnetic layer as wellas by inserting a Gd or Tb conversion layer between thenonmagnet and the ferromagnet. Rare-earth spacers do notgenerate significant spin-orbit torques by themselves, but theyenhance the torque efficiency up to four times when Cr isthe source of spin and orbital currents and reverse the signof the torques generated by Pt. The latter effect is attributedto the OHE overcoming the SHE in Pt. Finally, we present aphenomenological extension of the spin drift-diffusion modelthat includes orbital effects and the conversion between spinand orbital moments, which accounts for both the thicknessdependence and sign change of the spin-orbit torques gen-erated by the interplay of OHE and SHE in NM/FM andNM/ X/FM heterostructures. II. BACKGROUND According to the theory of the OHE, an electric field ap- plied along the xdirection in a material with orbital texture inkspace induces interband mixing that results in electron states with finite orbital angular momentum [ 5–7]. Electrons occupying these nonequilibrium states carry the angular mo-mentum as they travel in real space. Therefore, although thetotal orbital momentum vanishes, a nonzero orbital current isproduced along the z(y) direction with orbital polarization parallel to ±y(±z), similar to the SHE [see Fig. 1(a)]. Thelatter occurs concomitantly with the OHE when the nonmag- net has nonzero spin-orbit coupling /angbracketleftL·S/angbracketright NM. The primary spin current injected into the adjacent ferromagnet exerts adirect torque on the local magnetization (spin torque). Or-bitals, instead, act indirectly through the spin-orbit couplingof the ferromagnet that converts the orbital current into asecondary spin current. We refer to the torque generated bythis secondary spin current as orbital torque. The (indepen-dence) dependence of the (spin) orbital torque on /angbracketleftL·S/angbracketright FM is the key difference between SHE and OHE. In the SHE scenario, the angular momentum is entirely generated in thenonmagnet, and the ferromagnet behaves almost as a passivelayer since it only contributes to the properties of the NM/FMinterface. In contrast, the OHE in a NM/FM bilayer dependson both the interfacial and bulk properties of the ferromagnet,which is directly involved in the torque generation. Sincethe orbital conductivity is typically large [ ≈10 5(/Omega1m)−1][4] but the spin-orbit coupling of 3 dferromagnets is relatively weak [ 24], the orbital torque efficiency in NM/FM bilayers is finite but small. Alternatively, the orbital torque may beenhanced by realizing most of the orbital-to-spin conversionin a spacer layer sandwiched between the nonmagnet and theferromagnet [Fig. 1(b)]. The effectiveness of this approach depends on the conversion efficiency of the spacer, its spinand orbital diffusion lengths, and the quality of the additionalinterfaces, as discussed later. Here, we summarize fundamental theoretical predictions and experimental confirmations of the OHE. We list ap-proaches to distinguish orbital and spin effects by meansof torque measurements in heterostructures with differentelements, thickness, and stacking order. Furthermore, we es-tablish a parallel between known spin-transport effects andpossible orbital counterparts that have not been observed yetbut could contribute to answering open questions about orbitaltransport. (i) Large orbital Hall conductivities have been predicted in several 3 d,4d, and 5 dtransition elements [ 4,11,12] and 2D materials [ 9,10]. Experimental evidence is so far limited to Cr [19,25], Cu [ 20,21,26], Zr [ 18], and Ta [ 20]. Recent experi- ments on V [ 23,27] can also be reinterpreted in light of the OHE. The coexistence of the OHE and the SHE, especially inheavy metals, makes it difficult to distinguish the two effects. (ii) The spin and orbital torques are expected to add constructively (destructively) when /angbracketleftL·S/angbracketright NM·/angbracketleftL·S/angbracketrightFM>0 (<0). This competition can be tailored by properly choosing the ferromagnet, as recently observed in Refs. [ 19,20]. (iii) In a NM/FM bilayer, the orbital Hall efficiency should depend on the thickness of both the nonmagnet ( tNM) and the ferromagnet ( tFM). In contrast, the spin Hall efficiency is nominally independent of the latter and results in an inversedependence of the spin torque on t FM[1]. The dependence of the orbital Hall efficiency on tNMhas been addressed in Ref. [ 19], but the role of tFMis still unknown. (iv) The spin diffusion in transition metals with strong SHE is typically limited to a few nanometers [ 28]. Al- though recent measurements suggest longer orbital dif-fusion lengths [ 19,29], the length scale of the orbital diffusion and its conversion into spins remain to be es-tablished. These quantities and the nature of the mech-anisms underlying the orbital scattering may be ad- 033037-2GIANT ORBITAL HALL EFFECT AND ORBITAL-TO-SPIN … PHYSICAL REVIEW RESEARCH 4, 033037 (2022) dressed by torque measurements in thick nonmagnetic films and by nonlocal transport measurements, which couldalso verify the existence of the inverse OHE. (v) Spacer layers between the nonmagnet and the ferro- magnet can alter spin torques in several ways, namely, byintroducing an additional interface with different spin scat-tering properties, by suppressing the spin backflow, and bymodifying the spin memory loss [ 1,30–33]. Such effects are expected to influence also the orbital torque. In addition,spacers can either increase or decrease the orbital torquedepending on the sign of their orbital and spin Hall conduc-tivities, and spin-orbit coupling, which converts orbitals intospins and vice versa. Pt spacers have been shown to increasethe orbital torques in light metal systems [ 19,21]; however, Pt is also a well-known SHE material. A systematic investigationof the enhancement or suppression of spin and orbital currentsin materials with different combinations of orbital and spinconductivities is required. (vi) The spin diffusion in multilayer structures is usu- ally modeled by semiclassical drift-diffusion equations thataccount for, e.g., spin backflow at interfaces, the spin-orbittorque dependence on the thickness of the nonmagnet, and thespin Hall and unidirectional magnetoresistance [ 34–39]. The model has not been extended yet to the OHE, which requiresthe inclusion of the spin-orbital interconversion mediated byspin-orbit coupling. (vii) The orbital transmission at the NM/FM interface is more sensitive to the interface quality than spins and, hence,to growth conditions and stacking order [ 7,18]. It is an open question whether the transmission can be described by a singleparameter equivalent to the spin-mixing conductance, whichwe dub orbital mixing conductance. (viii) The SHE generates dampinglike and fieldlike spin- orbit torques of comparable strength [ 1,40]. So far, no theoretical or experimental work has determined with cer-tainty the relative magnitude of the two components ofthe orbital torque. Assessing their strength may help us tounderstand the mechanism of accumulation, transfer, and con-version of orbitals. (ix) The OHE has been attributed to an intrinsic scattering mechanism in elements with orbital texture. The analogy withthe SHE [ 41,42] suggests that also extrinsic processes may contribute to the generation of orbital currents. Measuring theorbital Hall efficiency as a function of the element resistivitymay reveal extrinsic orbital effects. (x) The transmission and absorption of spins and orbitals at the interface with an insulating ferromagnet [ 43], e.g., yttrium iron garnet (YIG), may be fundamentally differentsince the latter do not interact with the magnetization. Earlyexperiments reported spin pumping effects in YIG/light metalbilayers, but they were interpreted in terms of the inverse SHE[44]. (xi) The generation and accumulation of orbitals at the NM/FM interface can modulate the longitudinal resistance bythe combination of direct and inverse OHE, as recently foundin Ref. [ 22]. Compared with the spin Hall magnetoresistance [35], such orbital Hall magnetoresistance may have a different dependence on the type of ferromagnet, its thickness, and thethickness of the nonmagnet.(xii) Orbital accumulation might also give rise to a unidi- rectional magnetoresistance, in analogy to the unidirectionalspin Hall magnetoresistance [ 45]. The underlying mechanism, however, would be intrinsically different since orbitals wouldnot directly alter the magnon population, whereas orbital-dependent scattering might contribute to the conductivityin addition to spin-dependent scattering [ 46]. On the other hand, the injection into the ferromagnet of electrons withfinite orbital momentum may induce an additional source oflongitudinal magnetoresistance analogous to the anisotropicmagnetoresistance [ 17]. Orbital effects are thus rich and intertwined with spin transport, allowing for additional means to tune the spin-orbittorque efficiency as well as to understand the transport ofangular momentum in thin-film heterostructures. In the fol-lowing, we address points (i)–(vii) listed above. We providecomprehensive evidence for the occurrence of giant OHEsin 3dand 5 dtransition metals, reveal the interplay of the OHE and SHE in ferromagnets of variable thickness withand without spacer layers, and establish a phenomenologicalframework to analyze and efficiently exploit the interplay ofspin and orbital currents in metallic heterostructures. III. EXPERIMENTS We studied NM/FM and NM/ X/FM multilayers where NM =Cr, Mn, or Pt, FM =Co or Ni, and X=Gd or Tb. The samples were grown by magnetron sputtering on a SiNsubstrate, capped with either Ti(2) or Ru(3.5) (thicknessesin nanometers), and patterned in Hall-bar devices by opticallithography and lift-off. All samples have in-plane magneti-zation. Current-induced spin-orbit torques were quantified bythe harmonic Hall voltage method [ 40] using angle-scan mea- surements [ 51]. We detected the first- and second-harmonic Hall voltage while applying an alternate current with 10 Hzfrequency and rotating a constant magnetic field in the easyplane of the magnetization [ xyplane; see Fig. 2(a)]. The harmonic signals were measured as a function of currentamplitude and field strength [Figs. 2(b) and2(c)]. The second- harmonic resistance depends on the field angle φasR 2ω xy= /Theta1Tcosφ+/Phi1T(2 cos3φ−cosφ). Here, /Theta1Tis the sum of the dampinglike spin-orbit field BDLand contribution from the thermal gradient along z, and /Phi1Tdepends on the fieldlike spin-orbit field BFLand the Oersted field BOe. Thus the anal- ysis of R2ω xymeasured at different magnetic fields allows for the separation of torques and thermal effects, yielding themagnitude of the spin-orbit fields for a given electric field[51]. These spin-orbit fields exert spin and orbital torques on the magnetization T DL=MsBDLm×(p×m) and TFL= MsBFLm×p, where pis the net spin polarization direction, mis the magnetization vector, and Msis the saturation mag- netization. In the following, we consider uniquely BDLsince, apart from Ni/Cr and Co/Pt samples, BFLwas too small to distinguish from the Oersted field. The difficult detection ofB FLin our samples originates from the very small planar Hall coefficient (of the order of 1 m /Omega1) to which /Phi1Tis propor- tional. To compare samples with different elements, thickness,and stacking order, we converted B DLinto a spin-orbital 033037-3GIACOMO SALA AND PIETRO GAMBARDELLA PHYSICAL REVIEW RESEARCH 4, 033037 (2022) (b) (c)JCzmϕ Vxy(a)B xy 0.0 0.5 1.0 1.5 2.0-2.2-2.02.02.22.4 Co/Cr(12) Ni/Cr(12)ΘT/RAHE-3 (10 ) 1/Beff (1/T)0 60 120 180 240 300 3600.00.10.2 180 mT 1350 mTRxy2ω (mΩ) ϕ (° ) FIG. 2. (a) Schematic of the harmonic Hall effect measurements. An alternating current Jcflows along the Hall bar and generates transverse first- and second-harmonic Hall signals that depend on the angle φrelative to xof a magnetic field Bof constant ampli- tude. (b) Representative second-harmonic Hall resistance measuredin Co(2) /Cr(12) during the rotation of Bin the xyplane. The solid lines are fits to the function R 2ω xy=/Theta1Tcosφ+/Phi1T(2 cos3φ−cosφ) (see text). (c) Dependence of /Theta1T(dampinglike field BDL+ther- mal signal) normalized to the anomalous Hall resistance ( RAHE) on the effective field given by the sum of the applied magnetic field and demagnetizing field. Data are shown for Co(2) /Cr(12) and Ni(4)/Cr(12). The slope of the linear fit (solid lines) is proportional toBDL, while its intercept with the vertical axis corresponds to the thermal contribution, which is field independent and can be easilydistinguished. conductivity according to the formula ξLS=2e ¯hMstFMBDL E, (1) where eis the electron charge, ¯ his Planck’s constant, tFMis the thickness of the ferromagnet, and E=ρJcis the applied electric field ( ρis the longitudinal resistivity, and Jcis the current density) [ 1,52]. The normalization to the applied elec- tric field avoids the ambiguities intrinsic to the calculation ofthe current density in a heterostructure. Since in our samplesthe ferromagnet lies below the nonmagnet, we invert the signof the measured B DLto follow the convention that Pt has positive spin Hall conductivity. In the literature, Eq. ( 1)i s usually referred to as the spin Hall conductivity orspin-orbit torque efficiency , which is related to the effective spin Hall angle of the NM layer by θLS=ρξLS(Appendix A). Here, we point out that, when the SHE and OHE are consideredtogether, both spin and orbital currents influence ξ LSand their individual quantitative contributions cannot be disentangledbecause the spin-orbit torques depend on the total nonequi-librium spin angular momentum in the ferromagnet (primaryspins+converted spins) but not on the orbital component. This reasoning implies the impossibility of determining sep-arately the spin and orbital Hall conductivities of a materialby measuring nonequilibrium effects on an adjacent ferro-magnet, even for transparent interfaces. Thus we call ξ LSthe spin-orbital conductivity and θLSthe spin-orbital Hall angle. However, spin and orbital effects can still be distinguished(b)σ > 0S σ > 0LL S σ < 0S σ > 0LL S0.10.2 Ni/Cr Ni/Mn 02468 1 0 1 2 1 4 1 6-2-10 Co/Cr Co/MnξLS (105 Ω-1m-1) t (nm)NMNM02468 1 0 1 201234 Co/Pt Ni/PtξLS (105 Ω-1m-1) t (nm)(a) FIG. 3. (a) Spin-orbital conductivity as a function of the thick- ness of the Pt layer in Co(2) /Pt(tNM) and in Ni(4)/Pt(5). The solid line is a fit to the drift-diffusion equation [Eq. ( 2)]. The sign of the spin and orbital Hall conductivities in the nonmagnet is indi- cated and color-coded in the schematic representing the generation,transmission, and conversion of orbital (red) and positive (blue) or negative (white) spin currents. (b) The same as (a) in FM /Cr(t NM) and FM /Mn(tNM), where FM =Co(2) or Ni(4). at a qualitative level, as discussed in the following. We also note that a finite OHE could explain, at least in part, the largevariability of the spin-orbit torque efficiency found in sampleswith different ferromagnets, thicknesses, stacking order, andpreparation conditions [ 1]. IV . OHE in Cr, Mn, and Pt A. Dependence of ξLSon the thickness of the NM layer Figure 3compares ξLSmeasured in FM/NM bilayers, where FM is an in-plane magnetized Co(2) or Ni(4) layer andNM is a Cr, Mn, or Pt layer of variable thickness t NM.W e find that the two 3 dlight metals generate sizable spin-orbit torques, similar to previous measurements in materials withweak spin-orbit coupling such as V , Cr, and Zr [ 16,23,44,53]. The torques are remarkably strong in Cr-based samples, forwhich ξ LSreaches values similar to those for Co/Pt. To the best 033037-4GIANT ORBITAL HALL EFFECT AND ORBITAL-TO-SPIN … PHYSICAL REVIEW RESEARCH 4, 033037 (2022) TABLE I. Sign of the spin-orbit coupling /angbracketleftL·S/angbracketrightand orbital and spin Hall conductivity σL,Sin selected transition metals (see Refs. [ 4,11,12,47–49]). A positive orbital (spin) Hall conductivity means that a charge current along +xinduces an orbital current (spin current) along +zwith orbital (spin) angular momentum along −y [50]. Cr Mn Co Ni Gd Tb Pt /angbracketleftL·S/angbracketright –– ++ –– + σL ++ + + σS –– ++ –– + of our knowledge, this is the highest torque efficiency reported in the literature for a FM/NM bilayer made of light elements.However, the dependence of ξ LSon the type of ferromagnet and on tNMis very different in Cr and Mn with respect to Pt. Co/Pt(tNM) and Ni/Pt(5) have torque efficiencies of com- parable magnitude and identical sign. In contrast, when Cror Mn is used, ξ LSchanges sign when Co is replaced with Ni [see Fig. 3(b)]. A comparison between Fig. 3and Table I indicates that Co/Cr and Co/Mn behave as expected withinthe framework of the SHE, namely, the sign of the torques isopposite to Co/Pt because the spin Hall conductivity σ Shas opposite sign in Cr and Mn relative to Pt. The same argu-ment, however, cannot explain the positive sign of ξ LSin the Ni-based samples since the direction of the spin polarizationinduced by the SHE is fixed and determined by /angbracketleftL·S/angbracketright NM. The sign change can be accounted for only by consideringthe OHE and the opposite sign of the spin and orbital Hallconductivities of Cr and Mn. In this case, the negative ξ LS measured in Co/Cr and Co/Mn indicates that in these samples the spin torque overwhelms the orbital torque. The positivespin-orbital conductivity found with Ni shows instead that theorbital-to-spin conversion in this ferromagnet is so efficientas to make the orbital torque stronger than the spin torque[19]./angbracketleftL·S/angbracketright FMis indeed predicted to be larger in Ni than in Co and positive [ 7,20,54]; thus a larger amount of the orbital current can be converted into a spin current of opposite signto the primary spin current generated by Cr or Mn. Thereforethe torques exerted on Co are mostly generated outside theferromagnet thanks to the orbital-to-spin conversion occurringin the nonmagnet. In contrast, the torques on Ni result fromthe orbital-to-spin conversion inside the ferromagnet. The variation of ξ LSwith the thickness tNMof Cr and Mn is also different from the thickness dependence of thetorque efficiency in heavy elements [ 1,52]. In Co/Pt( t NM),ξLS saturates at about 9 nm [Fig. 3(a)]. The fit to the drift-diffusion equation ξLS(t)=σLS/bracketleftbigg 1−sech/parenleftbiggt λ/parenrightbigg/bracketrightbigg (2) yields a diffusion length λ=2.2 nm and an intrinsic spin- orbital Hall conductivity σLS=3.5×105[¯h 2e](/Omega1m)−1.T h i s value, which agrees with previous works [ 1,3,41,52,55], as- sumes a transparent interface and is thus an underestimationof the intrinsic spin-orbital Hall conductivity of Pt. In Cr andMn,ξ LSincreases with tNMand does not saturate, even at tNM=15 nm. The intrinsic spin-orbital Hall conductivity ofCr is thus significantly larger than ξLSreported in Fig. 3(b). Indeed, fitting ξLSin Co/Cr(tNM) with λfixed in the range 15– 25 nm yields 5 ×105<|σLS|<12×105(/Omega1m)−1, in good agreement with the predicted giant orbital Hall conductivityof Cr [ 4]. The trend of ξ LS(tNM) hints at two alternatives. The first possibility is that the spin ( λS) and orbital ( λL) diffusion lengths of Cr and Mn are larger than the typical spin diffusionlength of heavy elements. For example, λ Sis found to be about 13 and 11 nm in Cr and Mn, respectively, in Ref. [ 44], whereas λS=1.8 nm and λL=6.1 nm in Cr according to Ref. [ 19]. Alternatively, we argue that it suffices to have a large orbital diffusion length and a nonzero /angbracketleftL·S/angbracketrightNMfor spins to accumulate over long distances, even if the spin diffusionlength in the nonmagnet is short (see Sec. VI). Spin torque measurements cannot distinguish between the two possibili-ties. Nonetheless, the trends in Fig. 3suggest the possibility to increase the spin-orbital conductivity in FM/Cr sampleswith large t NMup to and beyond the maximal efficiency of Co/Pt. This possibility has gone unnoticed so far because thinnonmagnetic films ( t NM≈5 nm) are typically considered in torque measurements. A very long orbital diffusion length in Mn may also ex- plain why ξLSis smaller in Mn than in Cr at any thickness and independently of the ferromagnet. This result contrastswith theoretical calculations that predict large and comparableorbital conductivities in Cr and Mn [ 4] but agrees with the spin pumping measurements of Ref. [ 44]. We notice that Ref. [ 4] considered the bcc structure to calculate the orbital conduc-tivity of Mn, but different crystalline phases can compete andcoexist in Mn thin films [ 56]. This difference may account for the small experimental value of ξ LS. Alternatively, the small spin-orbital conductivity may be determined by a differentquality of the FM/Cr and FM/Mn interfaces, to which theorbital current is very sensitive [ 7,18], possibly because of Co and Mn intermixing [ 57]. Owing to the larger resistivity of Mn compared with Cr, however, we note that the effectivespin-orbital Hall angle of Co(2)/Mn(9) is θ LS=−0.03, which is comparable to θLS=−0.05 of Co(2)/Cr(9) (Appendix A). We also notice that interfacial effects (interfacial torques, spin memory loss, and spin transparency) can influence thestrength of the torques and hence the spin-orbital conductivity,as shown by the different ξ LSmeasured in Co/Pt and Ni/Pt samples [ 58]. However, interfacial effects cannot explain our results, namely, the sign change of ξLSwith the ferromagnet and its monotonic increase with tNM, because they should be independent of the thickness of the nonmagnet and becomenegligible in thick films. Overall, these measurements provide strong evidence of the OHE and orbital torques in Cr and Mn, in agreement withtheoretical predictions and previous studies of Cr-based sam-ples [ 4,19,25]. Additionally, they show that the spin-orbital diffusion length is much longer in light elements than in Pt,a difference that could be exploited to boost the effectivespin-orbital conductivity beyond the limit of FM/Pt samples. B. Dependence of ξLSon the thickness of the FM layer Theoretical calculations of the spin and orbital transfer at the FM/NM interface predict a different dependence of the 033037-5GIACOMO SALA AND PIETRO GAMBARDELLA PHYSICAL REVIEW RESEARCH 4, 033037 (2022) spin and orbital torque on the thickness tFMof the ferromagnet [59]. The former is dominant when tFMis small, whereas the orbital torque can be comparable to or larger than the spintorque in thick ferromagnets. As a consequence, the totaltorque may change sign when t FMincreases if /angbracketleftL·S/angbracketrightNM· /angbracketleftL·S/angbracketrightFM<0, as, for instance, in the case of Co/Cr. To test this possibility, we measured the torque on the magnetizationof Co( t FM)/Cr(9) and Co( tFM)/Pt(5) as a function of tFM. Figures 4(a) and4(b) show the torque per unit electric field calculated as T=MsBDL E. The sign of the torque is opposite in the two sets of samples and does not change in the exploredthickness range. This result might suggest that in Co( t FM)/Cr the orbital torque is always negligible compared with thespin torque. However, a careful analysis indicates a differentscenario. After taking into account the dead magnetic layer(0.5 and 0.3 nm in the samples with Cr and Pt, respectively;see Appendix B), we tentatively fit the dependence of Ton the ferromagnet thickness to ∼1/t FM. This scaling should reflect the inverse proportionality of the torque amplitude to the mag-netic volume when the current-induced angular momentumis generated outside the ferromagnet. In this case, the spinHall conductivity is constant and solely determined by thecharge-to-spin conversion efficiency of the nonmagnet [ 1], and Eq. ( 1) yields T=¯h 2eξLS tFM. (3) Equation ( 3) captures well the variation of the torque only fortFM<1–2 nm, in both Co( tFM)/Cr and Co( tFM)/Pt. The discrepancy at large thicknesses suggests the presence of atorque mechanism additional to the spin current injection fromthe nonmagnetic layer. This possibility is corroborated by thethickness dependence of the spin-orbital conductivity, whichis different in the two series of samples. In Co( t FM)/Pt,|ξLS| is approximately constant up to 3 nm and increases at larger tFMby about 20% [see Fig. 4(c)]. In Co( tFM)/Cr, instead, |ξLS|initially increases as the ferromagnet becomes thicker, possibly due to the formation of a continuous Co/Cr interface;then it decreases starting from t FM=1 nm and drops by more than 50% at tFM=3 nm relative to the maximum. Beyond this thickness, it remains approximately unchanged. The distinctthickness dependence in Co( t FM)/Cr and Co( tFM)/Pt cannot be ascribed to strain [ 60] since Co is grown on an amorphous substrate. In addition, strain-induced effects should be similarin the two sample series. Moreover, it cannot be attributed to avariation of the interface quality. Since the latter is expected toimprove as Co becomes thicker, the spin-orbital conductivityshould increase or remain approximately constant for t FM> 1 nm. Furthermore, we exclude that the measured trend de-pends on uncertainties in the saturation magnetization due toproximity effects since ξ LSdepends on the areal magnetization [see Eq. ( 1)], which is free from ambiguity (see Appendix B). Finally, we rule out self-torques due to the SHE inside the Colayer [ 49] since control measurements in Co(7)/Ti(3) do not give evidence of torques within the experimental resolution. Alternatively, we propose that the decrease of the spin- orbital conductivity with t FMin Co/Cr results from the competition between spin and orbital torques in the ferromag-net. As sketched in Figs. 4(d) and4(e), the spin and orbital currents J SandJLdecay inside the ferromagnet on a lengthzJL JSJLS zJL JSJLS tFM tFM|ξ |LS(a) (b) (c) (d) (e) |ξ |LS02468 1 0 1 2-1.6-0.4-0.200.2 Co/Cr Co/PtT (ΩTm-2) tFM (nm) 0123456789 1 0 1 1 1 201234|ξLS| (105 Ω-1m-1) t (nm)FM02468 1 0 1 2-0.04-0.0200.020.04T (ΩTm-2) tFM (nm) FIG. 4. (a) Dependence of the spin-orbit torque normalized to the applied electric field on tFMin Co( tFM)/Cr(9) and Co( tFM)/Pt(5). The solid lines are fits to1 tFM. (b) Enlarged view of (a). (c) Dependence of |ξLS|ontFMin the two sample series. (d) and (e) Schematics showing qualitatively the interplay of the spin JSand orbital JLcurrents, which are injected into the ferromagnet from the interface with the nonmag- netic metal and decay with the distance z. Part of the orbital moments is converted into spin moments and generates a spin current JLSwith the same (opposite) polarization as the primary spin current in Pt/Co (Cr/Co). JSyields the spin torque, and JLSyields the orbital torque. The spin-orbital conductivity ξLSis constant when the orbital-to-spin conversion is negligible (dashed line). It increases with tFMwhen JS andJLSadd up and decreases when JSandJLScompete (solid line). scale determined by the respective dephasing lengths. In the absence of orbital-to-spin conversion, the spin-orbital conduc-tivity, which depends on the absorption of the injected spincurrent ξ LS∼JS(0)−JS(tFM), increases rapidly with tFMand 033037-6GIANT ORBITAL HALL EFFECT AND ORBITAL-TO-SPIN … PHYSICAL REVIEW RESEARCH 4, 033037 (2022) remains constant afterwards because spin dephasing occurs within a few atomic layers from the interface [ 61,62]. On the other hand, if we assume that the orbital current is transmittedover a distance longer than its spin counterpart [ 59] and that part of it is also converted into a spin current J LS, then ξLS∼ JS(0)−JS(tFM)±JLS(tFM) can increase or decrease with tFM depending on the relative sign of JSandJLS, i.e., on the prod- uct/angbracketleftL·S/angbracketrightNM·/angbracketleftL·S/angbracketrightFM. Since the latter is positive (negative) in Co( tFM)/Pt [Co( tFM)/Cr], our qualitative model envisages an increase (decrease) of the spin-orbital conductivity with tFM, in agreement with our measurements and the thickness dependence predicted in Ref. [ 59]. The dependence of both T andξLSontFMshows that Co, rather than being a passive layer subject to an externallygenerated spin current, participates in the overall generationof spin-orbit torques. The active role of the ferromagnet inval-idates the assumption on which Eq. ( 3) rests and explains the deviation of the torque measured at large thicknesses from the1/t FMdependence. Interestingly, these measurements point to a non-negligible OHE in Pt, in accordance with the measure-ments discussed next. V . ORBITAL-TO-SPIN CONVERSION IN A SPACER LAYER The results presented in Sec. IVshow that the spin-orbital conductivity of a light metal can be maximized by a properchoice of the ferromagnet and its thickness. There is, however,a limitation from both a practical and theoretical point of view.According to Hund’s third rule, light metals have oppositespin-orbit coupling relative to ferromagnetic Fe, Co, and Ni;thusξ LScannot be maximized in such bilayers. As proposed in Sec. II, this optimization may be possible, instead, if the orbital current is converted into the spin current prior to the in-jection into the ferromagnetic layer [Fig. 1(b)]. This approach requires materials with high spin-orbit coupling between thelight metal and the ferromagnet [ 21]. Although the additional layer can itself be a source of spin current, we show in thefollowing how thickness-dependent measurements reveal theunderlying orbital-to-spin conversion and indicate the optimalconversion conditions. Figure 5shows the spin-orbital conductivity measured in Co(2)/X(t X)/Cr(9) and Co(2) /X(tX)/Pt(5) as a function of the rare-earth thickness, where Xis either Gd or Tb. We find a drastic change of the magnitude and sign of the torquesupon increasing t X. As the rare-earth layer becomes thicker in Co/X(tX)/Cr,|ξLS|first increases, reaching its maximum magnitude at about t=3 nm, and then decreases [notice the negative sign of ξLSin Fig. 5(a)]. At this thickness, |ξLS| of Co/ X(3)/Cr is three to four times larger than in Co/Cr and is thus comparable to or larger than the highest spin-orbital conductivity of Co/Pt [cf. Figs. 3(a) and 5(a)]. In Co/X(t X)/Pt, instead, ξLSdecreases rapidly with tX, changes sign at 2 nm, and saturates thereafter. This variation, which issimilar in samples containing Gd and Tb, is in direct contrastwith the widespread assumption that the positive spin Hallconductivity of Pt determines the sign and magnitude of thedampinglike spin-orbit torque in Pt heterostructures. Indeed, our findings cannot be attributed to the sole SHE in the nonmagnetic layer, nor can they be attributed to thespin-orbit torques generated by the rare-earth layer, which,(b)σ < 0S σ > 0LL S L S σ > 0S σ > 0L(a) 0123456-5-4-3-2-10 Co/Gd/Cr Co/Tb/CrξLS (105 Ω-1m-1) t (nm)X 02468 1 0-3-2-10123 Co/Gd/Pt Co/Tb/PtξLS (105 Ω-1m-1) tX (nm) FIG. 5. (a) Dependence of the spin-orbital conductivity on the thickness of the rare-earth spacer in Co(2) /Gd(tX)/Cr(9) and Co(2)/Tb(tX)/Cr(9). The schematic depicts the conversion of the orbital current into a spin current. Since the spin and orbital Hallconductivities of Cr are opposite and the spin-orbit coupling of Gd and Tb is negative, the primary and converted spin currents have the same sign. (b) The same as (a) in Co(2) /Gd(t X)/Pt(5) and Co(2)/Tb(tX)/Pt(5). In this case, the primary (blue) and converted (white) spin currents have opposite sign because Pt has positive spin and orbital Hall conductivities and Gd and Tb have negativespin-orbit coupling. although present, are too small to explain the sizable change ofξLSin the trilayers with respect to the Co/Cr and Co/Pt bilayers (see control measurements of Co/Tb and Co/Gd in theConclusions). Moreover, samples with inverted position of Gdand Tb with respect to the Co layer present spin-orbital Hallconductivities similar to the samples without the spacer, whichindicates that the rare-earth layer is not the dominant source ofspin-orbit torques (see the Conclusions). Instead, the results inFig. 5can be rationalized by considering the combination of OHE, SHE, and orbital-to-spin conversion in the spacer. Thenet spin current transferred from Cr or Pt to Co depends onthe transmission at the interface, the spin and orbital diffusionin the rare-earth layer, and its orbital-to-spin conversion effi-ciency. Whereas the first two effects always diminish the spincurrent reaching the ferromagnet, the orbital-to-spin conver-sion enhances it when /angbracketleftL·S/angbracketright NM·/angbracketleftL·S/angbracketrightX>0 and weakens it when /angbracketleftL·S/angbracketrightNM·/angbracketleftL·S/angbracketrightX<0. This is the case for samples containing Cr and Pt, respectively (see Table I). The length scale over which the effect takes place is determined by thecombination of the spin and orbital diffusion lengths of Gdand Tb. When the spacer is thin relative to these two lengths,the orbital-to-spin conversion supplies the spin current withmore spins than those lost by scattering. On the other hand, 033037-7GIACOMO SALA AND PIETRO GAMBARDELLA PHYSICAL REVIEW RESEARCH 4, 033037 (2022) spin-flip events become dominant at large thicknesses and decrease the transmitted spin current. The spin-orbital con-ductivity saturates then to a finite value determined by theSHE of the rare-earth layer, as indicated by the similar ξ LS measured in samples with either Cr or Pt and thick spacers (tX/greaterorequalslant6n m ) . These findings highlight the importance of achieving ef- ficient orbital-to-spin conversion. This can be pursued bysandwiching a rare-earth spacer of optimal thickness betweenthe ferromagnet and the nonmagnet because rare-earth met-als are effective enhancers of the conversion but not strongsources of spin-orbit toques [ 63]. Remarkably, our results also provide evidence of a strong OHE in Pt. VI. GENERALIZED DRIFT-DIFFUSION MODEL OF ORBITAL AND SPIN CURRENTS To shed light on the interplay between spin and orbital currents, we developed a 1D model that takes into accountthe generation and diffusion of both spin and orbital angu-lar momenta as well as their interconversion mediated byspin-orbit coupling. We consider first a single nonmagneticlayer where an electric field Eapplied along xinduces the SHE and OHE. Let μ=μ S,Lbe the spin or orbital chemical potential and Jμ=JS,Lbe the corresponding spin or orbital current along zwith spin and orbital polarization along y. The generation, drift, and diffusion of spins and orbitals aregoverned by [ 34–39] d 2μ dz2=μ λ2μ, (4) Jμ=−σ 2edμ dz+σHE, (5) where λμis the diffusion length, σis the longitudinal electri- cal conductivity, and σHis the spin or orbital conductivity, i.e., the off-diagonal element of the conductivity tensor. Solvingthese equations yields μ=Ae z/λμ+Be−z/λμ, with the coeffi- cients AandBobtained by imposing the boundary condition thatJμvanishes at the edges of the nonmagnet. In this form, however, the equations of the spin and orbital components areindependent and cannot account for the orbital-to-spin andspin-to-orbital conversion mediated by spin-orbit coupling.To capture this process, we add a phenomenological termto Eq. ( 4) for the spin (orbital) chemical potential that is proportional to its orbital (spin) counterpart, i.e., d 2μS dz2=μS λ2 S±μL λ2 LS, (6) d2μL dz2=μL λ2 L±μS λ2 LS, (7) JS=−σ 2edμS dz+σSE, (8) JL=−σ 2edμL dz+σLE, (9) where the +(−) sign corresponds to negative (positive) spin- orbit coupling. Physically, this additional term represents theconversion between spins and orbitals at a rate proportional tothe respective chemical potential. Thus, even when the SHE is negligible, a finite spin imbalance is produced in responseto the orbital accumulation. The parameter controlling thisprocess is the coupling length λ LS, which is a measure of both the efficiency and length scale over which the conversion takesplace. We remark that Eqs. ( 6)–(9) are phenomenological and based on the hypothesis that spin and orbital transport canbe described on an equal footing. They assume implicitly thepossibility of defining spin and orbital potentials and currentseven if the spin and orbital angular momenta are not conservedin the presence of spin-orbit coupling and the crystal field[64]. In this regard, we notice that the spin diffusion model has found widespread use in the quantitative analysis of spin-orbit torques [ 1,52,65], spin Hall magnetoresistance [ 35], and surface spin accumulation [ 55] despite the nonconservation of spin angular momentum. Moreover, there is a fundamentaldifference between spin and orbital transport that makes theapproximations underlying the orbital drift-diffusion modelless critical. Contrary to intuition, the crystal field does notquench the nonequilibrium orbital moment as efficiently as it suppresses the equilibrium orbital moment. This is because the orbital moment is carried by a relatively narrow subsetof conduction electron states, namely, its transport is medi-ated by “hot spots” in kspace. Since the orbital degeneracy of the hot spots is in general protected against the crystalfield splitting, the orbital momentum can be transported overlonger distances than its spin counterpart [ 5,59]. This orbital transport mechanism has no spin equivalent and is supportedby the experimental evidence that orbital diffusion lengths innonmagnets and dephasing lengths in ferromagnets are signif-icantly longer than the corresponding spin lengths, as shownin this paper and in Refs. [ 19,29]. Further theoretical work is required to ascertain the limits of our spin-orbital modeland determine how to capture analytically the spin-orbitalinterconversion. However, our model is consistent with theBoltzmann approach proposed in Ref. [ 66] and also repro- duces the experimental results, as explained in the following. To solve the coupled equations ( 6) and ( 7), we substitute the former into the latter and obtain d 4μS dz4−/parenleftbigg1 λ2 S+1 λ2 L/parenrightbiggd2μS dz2+/parenleftbigg1 λ2 Sλ2L−1 λ4 LS/parenrightbigg μS=0.(10) The solution to Eq. ( 10) reads μS(z)=Aez/λ1+Be−z/λ1+Cez/λ2+De−z/λ2, (11) where 1 λ2 1,2=1 2⎡ ⎣1 λ2 S+1 λ2 L±/radicalBigg/parenleftbigg1 λ2 S−1 λ2 L/parenrightbigg2 +4 λ4 LS⎤ ⎦ (12) are the combined spin-orbital diffusion lengths that result from the coupling of the spin and orbital degrees of freedomintroduced by λ LS. Equation ( 11) is the generalization of the standard diffusion of spins valid in the absence of spin-orbitalinterconversion. Two additional exponentials appear becauseof the coupling between LandS. For the same reason, the spin-orbital diffusion lengths λ 1,2are a combination of the spin, orbital, and coupling lengths. The same formal solutionas Eqs. ( 11) and ( 12) holds for the orbital chemical potential 033037-8GIANT ORBITAL HALL EFFECT AND ORBITAL-TO-SPIN … PHYSICAL REVIEW RESEARCH 4, 033037 (2022) because our model treats μSandμLon an equal footing. However, the eight unknown coefficients in Eq. ( 11) (four forμSand four for μL) are in general different between μS andμL. They are found by imposing that the spin and orbital currents vanish at the edges of the nonmagnet and that the pairof solutions for μ SandμL[Eq. ( 11)] satisfies Eqs. ( 6) and ( 7) at any z. Then, we find that the spin chemical potential at the surface of the nonmagnet increases with the thickness tNMas μS(tNM)=2eλ1/parenleftBiggσS∓σL λ2 LSγ2 1−γ2 γ1/parenrightBigg E σtanh/parenleftbiggtNM 2λ1/parenrightbigg +2eλ2/parenleftBiggσS∓σL λ2 LSγ1 1−γ1 γ2/parenrightBigg E σtanh/parenleftbiggtNM 2λ2/parenrightbigg , (13) where γi=1 λ2 i−1 λ2S. Equation ( 13) captures the interplay be- tween the SHE and OHE, which reinforce or weaken each other depending on the sign of the spin-orbit coupling and onλ LS. In comparison, in the absence of coupling between Sand L,E q .( 13) would read μS(tNM)=2eλSσS σEtanh/parenleftbiggtNM 2λS/parenrightbigg , (14) consistent with the standard spin drift-diffusion model. We note that Eqs. ( 11) and ( 13) are valid under the condition λLS>√λSλLbecause for smaller values of λLSthe solution to Eq. ( 10) is a linear combination of complex exponential functions, i.e., μSandμLhave an oscillatory dependence on z. Similar oscillations have been predicted in Ref. [ 9]. However, we argue that complex solutions to Eq. ( 12) are incompati- ble with experimental results since an oscillatory dependenceof spin-orbit torques or spin Hall magnetoresistance on thethickness of the nonmagnetic layer has never been observed.The condition λ LS>√λSλLalso implies that the conversion between spins and orbitals cannot occur on a length scaleshorter than the shortest distance over which either spins ororbitals diffuse. At the same time, it shows that the conver-sion is always less efficient than the intrinsic spin and orbitalrelaxation. We apply our model to study the interplay of nonequi- librium spins and orbitals induced by the SHE and OHEin two exemplary situations. First, we take a single non-magnetic layer with negative spin-orbit coupling, e.g., Cr.Figure 6shows the spin and orbital chemical potentials in three different conditions. In Fig. 6(a), the OHE is turned off (σ L=0), and the SHE is active [ σS=−105(/Omega1m)−1]. In Figs. 6(b)and6(c), the situation is opposite, namely, σL=105 (/Omega1m)−1andσS=0. In all cases, orbitals (spins) accumulate at the interfaces even if the OHE (SHE) is set to zero. Since/angbracketleftL·S/angbracketright NM<0, the two chemical potentials are of opposite sign. When λLSdecreases, the spin accumulation resulting from the orbital conversion increases approximately as λ−2 LS [Fig. 6(b) and Eq. ( 13)]. Interestingly, we find that even if λSis small, spins accumulate on a long distance because the spin-orbital diffusion lengths λ1,2are dominated by λL[cf. Figs. 6(b) and6(c)]. AsλLSdecreases and the spin-orbit conversion becomes more efficient, both the spin accumulation and the orbitalaccumulation increase at the sample edges. This effect might(a) (b) σ= 0L σ= 0S (c) -30030 0 5 10 15 20-0.80.00.8L (μeV ) S (μeV) z (nm) λ = 5 n mL λ = λ = 2 nmLSσ= 0S λ = λ = 2 nmLS λ = 2 nmS-0.20.00.2 05 1 0 1 5 2 0-10010L (μeV) S (μeV) z (nm)-10010 05 1 0 1 5 2 0-202 λLS= 10 nm λLS= 3 nmL (μeV) λLS= 10 nm λLS= 3 nmS (μeV) z (nm) FIG. 6. (a) Orbital and spin chemical potentials in a single non- magnetic layer with tNM=20 nm, σS=−105(/Omega1m)−1,σL=0, λS=λL=2n m ,a n d λLS=10 nm. (b) The same as (a) with σS=0, σL=105(/Omega1m)−1,λS=λL=2n m ,a n d λLS=3 or 10 nm. (c) The same as (a) with σS=0,σL=105(/Omega1m)−1,λS=2n m , λL=5n m , andλLS=10 nm. In all cases, the resistivity of the NM layer was set to ρ=56×10−8/Omega1m as measured for Cr, and an electric field E=5×104V/m was considered. seem counterintuitive, because spin-orbit coupling usually induces dissipation of angular momentum. In our model, how-ever, the dissipation of SandLis included in the parameters λ SandλL, respectively, whereas λLSdescribes the nondissipa- tive exchange of angular momentum between the orbital andspin reservoirs. Thus λ LSeffectively increases the spatial ex- tent of orbital and spin accumulation. Formally, this happensbecause one of the two spin-orbital diffusion lengths λ 1,2in- creases while the other changes weakly when λLSis reduced. As a consequence, more spins and orbitals can accumulate atthe sample edges. This result is similar to the model withoutspin-orbit coupling, which predicts an increase in μ Swith the spin diffusion length: μS(tNM/greatermuchλS)∼λS[see Eq. ( 14)]. Next, we consider a trilayer structure representative of the samples Co(2) /X(tX)/Cr(9) and Co(2) /X(tX)/Pt(5). We model the spatial variations of μSandμLby four equations of the same type as Eq. ( 11), two for the rare-earth layer and two for Cr (or Pt). We assume that μS,μL,JS, and JLare continuous at the X/NM interface ( z=tX) and impose the constraint that JSandJLrelate to μSandμL, respectively, 033037-9GIACOMO SALA AND PIETRO GAMBARDELLA PHYSICAL REVIEW RESEARCH 4, 033037 (2022) (a) (b) -1.5-1.0-0.50.0 01234560510JS (109 A/m2) JL (109 A/m2) tX (nm)012 02468 1 0048JS (109 A/m2) JL (109 A/m2) tX (nm)σ > 0S σ > 0L Pt X CotXCr X CotXσ < 0S σ > 0L FIG. 7. (a) Calculated spin and orbital currents at the FM/ X interface as a function of tXin Co(2) /X(tX)/Pt(5). (b) The same as (a) for Co(2) /X(tX)/Cr(9). The orbital and spin Hall conductivities of Pt and Cr are indicated above the graphs. The parameters usedto calculate the orbital and spin currents can be found in Table II (Appendix C). through the mixing conductance GS,L: JX S(tX)=GX S eμS(tX), (15) JX L(tX)=GX L eμL(tX). (16) In doing so, we introduce the orbital equivalent of the spin mixing conductance, which is expected to depend on thespin-orbit coupling of the ferromagnet and to influence thestrength of the orbital torque. Thus, in our model, G Ltakes into account the additional orbital-to-spin conversion occur-ring in the ferromagnet or at the interface. Furthermore, weonly consider the real part of G S,Lsince the fieldlike torque is small in our samples. Finally, we assume a finite SHE inboth the nonmagnetic and rare-earth layers, but smaller inthe latter, whereas the OHE is present only in the nonmagnet(see Appendix Cfor a list of the parameters). We set σ S>0 in Pt, σS<0 in Cr and in the spacer, and σL>0 in both Cr and Pt. The spin-orbit coupling is assumed positive in Ptand negative in Cr and in the rare-earth layer. With thesereasonable assumptions, we can reproduce qualitatively theresults of Fig. 5, namely, the enhancement of the spin-orbital conductivity upon insertion of a rare-earth spacer betweenCo and Cr and the sign change of the torques when thesame layer is sandwiched between Co and Pt. Figure 7shows the calculated spin and orbital currents, to which spin-orbittorques are proportional, that reach the FM/ Xinterface as a function of the rare-earth thickness t X. In both the case of Cr and the case of Pt the orbital current decreases monoton-ically as t Xincreases because of the orbital diffusion away from the X/NM interface. In contrast, the spin current variesdifferently with tXdepending on whether Cr or Pt is chosen because the primary spin current and the current obtainedupon orbital-to-spin conversion in the rare-earth element havethe same sign with Cr and have opposite sign with Pt. Thuscalculations based on a generalized drift-diffusion model con-firm the interpretation of the data in Fig. 5, which cannot be explained without the inclusion of the OHE. We believethat a better quantitative agreement with the measurementscould be obtained by including additional effects that we havedisregarded, namely, the interfacial resistance ( μ SandμLnot continuous), the interfacial spin and orbital scattering ( JSand JLnot continuous), and the thickness dependence of the spacer resistivity and, possibly, of the diffusion lengths. The modelcould be extended to account for the orbital conversion in theferromagnet, which is hidden here behind the orbital mixingconductance. Finally, it may be employed to investigate othertransport effects such as the spin Hall magnetoresistance andits orbital counterpart. VII. CONCLUSIONS Our measurements of spin-orbit torques in FM/NM and FM/X/NM multilayers with light and heavy metals provide comprehensive evidence for strong OHE effects in 3 dand 5 d metals and establish a systematic framework to analyze andefficiently exploit the interplay of spin and orbital currents.Owing to the entanglement of the orbital and spin degreesof freedom in materials with finite spin-orbit coupling, thisinterplay is best described by combined spin-orbital conduc-tivity ( ξ LS) and diffusion length ( λLS) parameters rather than by considering the OHE and SHE as two separate effects.The experimental values of ξ LSfor the different systems and control samples are summarized in Fig. 8. Corresponding values of the spin-orbital Hall angle θLSare reported in Ap- pendix A. We found strong spin-orbit torques produced by the light elements Cr and Mn, whose sign depends on the adja-cent ferromagnet, in contrast with torques generated by theSHE. The spin-orbital conductivity increases with the thick-ness of the light metal layer without indications of saturation.This trend is compatible with spin-orbital diffusion lengthsλ LS/greaterorsimilar20 nm in these elements and extrapolates to a giant intrinsic spin-orbital conductivity as predicted by theory [ 4]. Because of the competition between spin and orbital torques,the spin-orbital conductivity varies with the thickness of theferromagnet in a monotonic or nonmonotonic way dependingon the relative sign of /angbracketleftL·S/angbracketright NMand/angbracketleftL·S/angbracketrightFM. Furthermore, we show that the interplay between spin and orbital torquescan be drastically enhanced by inserting a 4 fspacer layer between the nonmagnet and the ferromagnet. As summarizedin Fig. 8, the inclusion of a Tb (Gd) spacer results in a fourfold (threefold) increase of the torques generated by Cr and Mnthat cannot be attributed to spin currents generated by therare-earth element. Instead, the enhancement results from theconversion of the orbital current into a secondary spin currentof the same sign as the primary spin current. The orbital-to-spin conversion has a striking effect in Pt, when the primaryspin current generated by the SHE and the secondary spincurrents generated by the OHE interfere destructively. Thiseffect results in the reversal of the spin-orbit torque generatedby Pt when the orbital-to-spin conversion rate is stronger than 033037-10GIANT ORBITAL HALL EFFECT AND ORBITAL-TO-SPIN … PHYSICAL REVIEW RESEARCH 4, 033037 (2022) FM/Cr(9) Co(2)/Gd(3)/Cr(9) Co(2)/Tb(3)/Cr(9) Co(2)/Gd(4) Co(2)/Tb(4) Gd(3)/Co(2)/Cr(9)-4-3-2-101(a) (b) (c)23ξLS (105 Ω-1m-1) CoNi Cr FM/Mn(9) Co(2)/Gd(3)/Mn(9) Gd(3)/Co(2)/Mn(9)CoNi Mn FM/Pt(5) Co(2)/Gd(3)/Pt(5) Co(2)/Tb(3)/Pt(5) Gd(4)/Co(2)/Pt(5) Tb(4)/Co(2)/Pt(5)NiCo Pt FIG. 8. Comparison of the effective spin-orbital conductivity ξLSmeasured in NM/FM and NM/ X/FM layers where (a) NM =Cr, (b) NM =Mn, and (c) NM =Pt; FM =Co,Ni,X=Gd,Tb. The thickness of each layer is indicated in nanometers in parentheses. The results of control experiments on X/FM and NM/FM/ Xlayers are also shown. the primary spin current. These findings indicate the presence of a strong OHE and SHE in Cr, Mn, and Pt and highlightthe importance of orbital-to-spin conversion phenomena indifferent types of heterostructures. The largest ξ LS=−4.3× 105(/Omega1m)−1andθLS≈0.25 are found in Co(2)/Gd(3) /Cr(9) and Co(2)/Tb(3) /Cr(9) layers. Both of these parameters are larger compared with Co/Pt and previous measurements, in-dicating that optimization of the thickness of 3 dmetal layers and the insertion of 4 fspacers lead to giant spin-orbital Hall effects and ensuing spin-orbit torques. The fits of ξ LSas a function of thickness indicate that the spin-orbital conductiv-ity of Cr saturates at values of the order of 10 6(/Omega1m)−1,i n agreement with theoretical estimates [ 4]. Finally, we propose an extended drift-diffusion model that treats the orbital andspin moment on an equal footing and includes the orbital-to- FM/Cr(9) Co(2)/Gd(3)/Cr(9) Co(2)/Tb(3)/Cr(9)Co(2)/Gd(4) Co(2)/Tb(4) Gd(3)/Co(2)/Cr(9)-0.3-0.2-0.10.00.10.2θH CoNi Cr FM/Mn(9) Co(2)/Gd(3)/Mn(9) Gd(3)/Co(2)/Mn(9)MnCoNi FM/Pt(5) Co(2)/Gd(3)/Pt(5) Co(2)/Tb(3)/Pt(5) Gd(4)/Co(2)/Pt(5) Tb(4)/Co(2)/Pt(5)Co Ni Pt075150225300ρ (μΩ cm) Co Ni FIG. 9. Resistivity (top) and effective spin-orbital Hall angle (bottom) of the samples in Fig. 8in the main text.spin conversion mediated by spin-orbit coupling. The model explains both the monotonic and nonmonotonic behavior ofξ LSobserved in the FM/NM and FM/ X/NM multilayers as a function of thickness and spin-orbit coupling of the con-stituent layers. It also shows how the spatial profiles of theorbital and spin accumulation are determined by the combinedspin-orbital diffusion lengths and spin and orbital mixing con-ductances. Overall, our results provide a useful framework tomaximize the orbital-to-spin conversion efficiency, interpretexperimental results, and address open fundamental questionsabout orbital transport. (a) (b) 01234560.00.51.01.52.02.5 Co/Gd( tX)/Cr Co/Tb( tX)/CrMs*t (mA ) tX (nm)02468 1 00.00.40.81.21.62.02.4 Co/Gd( tX)/Pt Co/Tb( tX)/PtMs*t (mA) tX (nm) 02468 1 0 1 20481216 Co(tFM)/Cr Co(tFM)/PtMs*t (mA) tFM (nm)(c) FIG. 10. (a) Dependence of the areal magnetization on the thickness of the rare-earth layer in Co(2) /X(tX)/Cr(9) samples. (b) The same as (a) in Co(2) /X(tX)/Pt(5) samples. (c) Dependence of the areal magnetization on the thickness of the ferromagnet in Co(tFM)/Cr(9) and Co( tFM)/Pt(5) samples. 033037-11GIACOMO SALA AND PIETRO GAMBARDELLA PHYSICAL REVIEW RESEARCH 4, 033037 (2022) TABLE II. Parameters used in the drift-diffusion model to calculate the spin and orbital currents in a FM/ X/NM trilayer, where NM is either Cr or Pt. λS,Lis the spin or orbital diffusion length, λLSis the spin-orbital conversion length, σS,Lis the spin or orbital Hall conductivity, α=±1 is the sign of the spin-orbit coupling, GS,Lis the spin or orbital mixing conductance, and ρis the electrical resistivity. The thickness of the Cr (Pt) layer was 9 (5) nm. An electric field E=5×104V/m was considered in both cases. λNM LλXLλNMSλXSλNMLSλXLS σNM L σX L σNM S σX S GL GS ρNM ρX (nm) (nm) (nm) (nm) (nm) (nm) [( /Omega1m)−1][ (/Omega1m)−1][ (/Omega1m)−1][ (/Omega1m)−1]αNMαX[(/Omega1m2)−1][ (/Omega1m2)−1](/Omega1m) ( /Omega1m) C r 8262 2 0 2 . 5 8 .2×1050 −0.7×105−0.15×105−1−13×1014101456×10−8115×10−8 P t 12222 2 . 5 8 .8×10503 .5×105−0.15×105+1−13×1014101433×10−8115×10−8 ACKNOWLEDGMENT We acknowledge the support of the Swiss National Science Foundation (Grant No. 200020_200465). APPENDIX A: EFFECTIVE SPIN-ORBITAL HALL ANGLE Figure 9shows the effective spin-orbital Hall angle of the samples presented in Fig. 8in the main text. The Hall angle was calculated according to θLS=ξLSρ, where ρis the resistivity of the entire stack. However, we refrain from esti-mating the resistivity of the individual layers and comparingquantitatively θ LSof the NM layers alone in this way, because the resistivity of the heterostructures depends strongly on in-terfaces and the thickness of all layers. Similarly to ξ LS,w e interpret θLSas a parameter that describes the simultaneous occurrence of the OHE, the SHE, and orbital-to-spin conver-sion. The values reported in Figs. 8and9are measured in samples with the thickness specified in the axis labels. APPENDIX B: SATURATION MAGNETIZATION Figure 10shows the surface saturation magnetization of samples belonging to the series Co(2) /X(tX)/Cr(9), Co(2)/X(tX)/Pt(5), Co( tFM)/Cr(9), and Co( tFM)/Pt(5) as a function of the corresponding thickness. The magnetiza-tion was measured by superconducting quantum interferencedevice (SQUID) magnetometry on blanket films grown si-multaneously to the measured devices. The measurementyields the magnetic moment of the sample, which, after nor-malization to the sample area, defines the areal saturationmagnetization M stFM. This parameter is to be preferred over the volume saturation magnetization, since the latter dependson the thickness of the ferromagnetically active material. Thisis in turn difficult to define with certainty in the studied sam-ples because of interdiffusion at interfaces, proximity effects,and possible ferrimagnetic coupling. Such a complexity, how-ever, does not impinge on the calculation of the spin-orbitalconductivity because the quantity appearing in Eq. ( 1)i st h e areal saturation magnetization M stFM, not the volume magne- tization. Figures 10(a) and 10(b) show that the areal magnetiza- tion decreases upon increasing the thickness of either Gdor Tb in both Cr- and Pt-based samples. We attribute thisreduction to the antiferromagnetic interaction between Co and the rare-earth layer. We note that the ferrimagnetic couplingcannot explain our results, namely, the trends presented inFig. 5. First, the torque efficiency rises by a factor of 3–4 when the thickness of the rare earth t Xincreases from 0 to 3 nm, while the areal magnetization decreases only by 20%.Second, the areal magnetization decreases monotonically with t X, while the trends in Fig. 5are not monotonic with respect to the thickness. For example, the spin-orbital conductivity ofCo(2)/X(t X)/Pt(5) saturates in the limit of large tX, whereas the magnetization does not. The areal magnetization of Co( tX)/Cr(9) and Co( tX)/Pt(5) samples increases linearly with tX, as expected. The linear fits yield dead layers of about 0.5 and 0.3 nm in the two series, re-spectively. The dead layer is likely located at the substrate/Cointerface and is probably thinner in the Co( t FM)/Pt(5) series because of proximity effects with Pt. These values have beentaken into account in the torque calculation in Fig. 4. Finally, the saturation magnetization of Co(2) /NM( t NM) and Ni(4) /NM( tNM) was found to be independent of tNM, except for Co(2) /Pt(tNM), where the areal magnetization in- creases by 7% from tPt=1n mt o tPt=12 nm (not shown). APPENDIX C: PARAMETERS OF THE DRIFT-DIFFUSION MODEL Table IIlists all the parameters used for the calculation of the orbital and spin currents in the Co(2) /X(tX)/Cr(9) and Co(2)/X(tX)/Pt(5) samples (Fig. 7). Some of them have been measured (spin diffusion length of Pt; spin Hall conductivityof Cr and Pt from Co/Cr and Co/Pt samples, respectively;and resistivity). Others have been chosen in accordance withthe literature (spin mixing conductance, sign of the spin-orbitcoupling, orbital conductivity). The remaining parameters,mostly involving orbitals and the spin-orbital interconver-sion, are not available in the literature and have been chosensuch that the calculations agree qualitatively with the mea-surements. From this perspective, the extended drift-diffusionmodel can be used to estimate the order of magnitude ofthe unknown parameters. For instance, the spin and orbitaldiffusion length and the spin-orbital coupling length of therare-earth layer must be of the order of a few nanometers atmost for the model to reproduce the measurements in Fig. 5. [1] A. Manchon, J. Železný, I. M. Miron, T. Jungwirth, J. Sinova, A. Thiaville, K. Garello, and P. Gambardella, Current-inducedspin-orbit torques in ferromagnetic and antiferromagnetic sys- tems, Rev. Mod. Phys. 91, 035004 (2019) . 033037-12GIANT ORBITAL HALL EFFECT AND ORBITAL-TO-SPIN … PHYSICAL REVIEW RESEARCH 4, 033037 (2022) [2] V . E. Demidov, S. Urazhdin, G. de Loubens, O. Klein, V . Cros, A. Anane, and S. O. Demokritov, Magnetization oscillationsand waves driven by pure spin currents, Phys. Rep. 673,1 (2017) . [3] J. Sinova, S. O. Valenzuela, J. Wunderlich, C. H. Back, and T. Jungwirth, Spin Hall effects, Rev. Mod. Phys. 87, 1213 (2015) . [4] D. Jo, D. Go, and H. W. Lee, Gigantic intrinsic orbital Hall effects in weakly spin-orbit coupled metals, P h y s .R e v .B 98, 214405 (2018) . [5] D. Go, D. Jo, C. Kim, and H. W. Lee, Intrinsic Spin and Orbital Hall Effects from Orbital Texture, Phys. Rev. Lett. 121, 086602 (2018) . [6] D. Go and H.-W. Lee, Orbital torque: Torque generation by orbital current injection, Phys. Rev. Research 2, 013177 (2020) . [7] D. Go, F. Freimuth, J.-P. Hanke, F. Xue, O. Gomonay, K.-J. Lee, S. Blügel, P. M. Haney, H.-W. Lee, and Y . Mokrousov, Theoryof current-induced angular momentum transfer dynamics inspin-orbit coupled systems, P h y s .R e v .R e s . 2, 033401 (2020) . [8] D. Go, D. Jo, H.-W. Lee, M. Kläui, and Y . Mokrousov, Orbi- tronics: Orbital currents in solids, Europhys. Lett. 135, 37001 (2021) . [9] L. Salemi, M. Berritta, and P. M. Oppeneer, Quantitative com- parison of electrically induced spin and orbital polarizations inheavy-metal/3 d-metal bilayers, P h y s .R e v .M a t e r i a l s 5, 074407 (2021) . [10] P. Sahu, S. Bhowal, and S. Satpathy, Effect of the inversion symmetry breaking on the orbital Hall effect: A model study,Phys. Rev. B 103, 085113 (2021) . [11] T. Tanaka, H. Kontani, M. Naito, T. Naito, D. S. Hirashima, K. Yamada, and J. Inoue, Intrinsic spin Hall effect and orbital Halleffect in 4 dand 5 dtransition metals, Phys. Rev. B 77, 165117 (2008) . [12] H. Kontani, T. Tanaka, D. S. Hirashima, K. Yamada, and J. Inoue, Giant Orbital Hall Effect in Transition Metals: Origin ofLarge Spin and Anomalous Hall Effects, Phys. Rev. Lett. 102, 016601 (2009) . [13] D. Go, D. Jo, T. Gao, K. Ando, S. Blügel, H.-w. Lee, and Y . Mokrousov, Orbital Rashba effect in a surface-oxidized Cufilm, P h y s .R e v .B 103, L121113 (2021) . [14] S. Bhowal and S. Satpathy, Intrinsic orbital moment and predic- tion of a large orbital Hall effect in two-dimensional transitionmetal dichalcogenides, Phys. Rev. B 101, 121112(R) (2020) . [15] L. M. Canonico, T. P. Cysne, A. Molina-Sanchez, R. B. Muniz, and T. G. Rappoport, Orbital Hall insulating phase in transitionmetal dichalcogenide monolayers, Phys. Rev. B 101, 161409(R) (2020) . [16] Z. C. Zheng, Q. X. Guo, D. Jo, D. Go, L. H. Wang, H. C. Chen, W. Yin, X. M. Wang, G. H. Yu, W. He, H.-W. Lee, J. Teng,and T. Zhu, Magnetization switching driven by current-inducedtorque from weakly spin-orbit coupled Zr, Phys. Rev. Res. 2 , 013127 (2020) . [17] H. W. Ko, H. J. Park, G. Go, J. H. Oh, K. W. Kim, and K. J. Lee, Role of orbital hybridization in anisotropic magnetoresistance,Phys. Rev. B 101, 184413 (2020) . [18] J. Kim, D. Go, H. Tsai, D. Jo, K. Kondou, H. W. Lee, and Y . C Otani, Nontrivial torque generation by orbital angular momen-tum injection in ferromagnetic-metal/Cu/Al 2O3trilayers, Phys. Rev. B 103, L020407 (2021) . [19] S. Lee, M.-G. Kang, D. Go, D. Kim, J.-H. Kang, T. Lee, G.-H. Lee, J. Kang, N. J. Lee, Y . Mokrousov, S. Kim, K.-J. Kim, K.-J.Lee, and B.-G. Park, Efficient conversion of orbital Hall current to spin current for spin-orbit torque switching, Commun. Phys. 4, 234 (2021) . [20] D. Lee, D. Go, H.-J. Park, W. Jeong, H.-W. Ko, D. Yun, D. Jo, S. Lee, G. Go, J. H. Oh, K.-J. Kim, B.-G. Park,B.-C. Min, H. C. Koo, H.-W. Lee, O. Lee, and K.-J. Lee,Orbital torque in magnetic bilayers, Nat. Commun. 12, 6710 (2021) . [21] S. Ding, A. Ross, D. Go, L. Baldrati, Z. Ren, F. Freimuth, S. Becker, F. Kammerbauer, J. Yang, G. Jakob, Y . Mokrousov, andM. Kläui, Harnessing Orbital-to-Spin Conversion of InterfacialOrbital Currents for Efficient Spin-Orbit Torques, Phys. Rev. Lett.125, 177201 (2020) . [22] S. Ding, Z. Liang, D. Go, C. Yun, M. Xue, Z. Liu, S. Becker, W. Yang, H. Du, C. Wang, Y . Yang, G. Jakob, M. Kläui, Y .Mokrousov, and J. Yang, Observation of the Orbital Rashba-Edelstein Magnetoresistance, Phys. Rev. Lett. 128, 067201 (2022) . [23] Q. Guo, Z. Ren, H. Bai, X. Wang, G. Yu, W. He, J. Teng, and T. Zhu, Current-induced magnetization switching in perpendicu-larly magnetized V/CoFeB/MgO multilayers, Phys. Rev. B 104, 224429 (2021) . [24] K. V . Shanavas, Z. S. Popovi ´c, and S. Satpathy, Theoretical model for Rashba spin-orbit interaction in delectrons, Phys. Rev. B 90, 165108 (2014) . [25] C.-Y . Hu, Y .-F. Chiu, C.-C. Tsai, C.-C. Huang, K.-H. Chen, C.- W. Peng, C.-M. Lee, M.-Y . Song, Y .-L. Huang, S.-J. Lin, andC.-F. Pai, Toward 100% spin-orbit torque efficiency with highspin-orbital Hall conductivity Pt-Cr alloys, ACS Appl. Electron. Mater. 4, 1099 (2022) . [26] H. An, Y . Kageyama, Y . Kanno, N. Enishi, and K. Ando, Spin- torque generator engineered by natural oxidation of Cu, Nat. Commun. 7, 13069 (2016) . [27] T. Wang, W. Wang, Y . Xie, M. A. Warsi, J. Wu, Y . Chen, V . O. Lorenz, X. Fan, and J. Q. Xiao, Large spin Hall anglein vanadium film, Sci. Rep. 7, 1306 (2017) . [28] J. Bass and W. P. Pratt Jr, Spin-diffusion lengths in metals and alloys, and spin-flipping at metal/metal interfaces: An experi-mentalist’s critical review, J. Phys.: Condens. Matter 19, 183201 (2007) . [29] Y .-G. Choi, D. Jo, K.-h. Ko, D. Go, and H.-w. Lee, Observation of the orbital Hall effect in a light metal Ti, arXiv:2109.14847 . [30] J.-C. Rojas-Sánchez, N. Reyren, P. Laczkowski, W. Savero, J.-P. Attané, C. Deranlot, M. Jamet, J.-M. George, L. Vila, and H.Jaffrès, Spin Pumping and Inverse Spin Hall Effect in Platinum:The Essential Role of Spin-Memory Loss at Metallic Interfaces,Phys. Rev. Lett. 112, 106602 (2014) . [31] X. Tao, Q. Liu, B. Miao, R. Yu, Z. Feng, L. Sun, B. You, J. Du, K. Chen, S. Zhang, L. Zhang, Z. Yuan, D. Wu, and H.Ding, Self-consistent determination of spin Hall angle and spindiffusion length in Pt and Pd: The role of the interface spin loss,Sci. Adv. 4, eaat1670 (2018) . [32] L. Zhu, D. C. Ralph, and R. A. Buhrman, Spin-Orbit Torques in Heavy-Metal–Ferromagnet Bilayers with Varying Strengths ofInterfacial Spin-Orbit Coupling, Phys. Rev. Lett. 122, 077201 (2019) . [33] C. O. Avci, G. S. D. Beach, and P. Gambardella, Effects of transition metal spacers on spin-orbit torques, spin Hall magne-toresistance, and magnetic anisotropy of Pt/Co bilayers, Phys. Rev. B 100, 235454 (2019) . 033037-13GIACOMO SALA AND PIETRO GAMBARDELLA PHYSICAL REVIEW RESEARCH 4, 033037 (2022) [34] P. C. van Son, H. van Kempen, and P. Wyder, Boundary Resis- tance of the Ferromagnetic-Nonferromagnetic Metal Interface,Phys. Rev. Lett. 58, 2271 (1987) . [35] Y .-T. Chen, S. Takahashi, H. Nakayama, M. Althammer, S. T. B. Goennenwein, E. Saitoh, and G. E. W. Bauer, Theory of spinHall magnetoresistance, P h y s .R e v .B 87, 144411 (2013) . [36] P. M. Haney, H. W. Lee, K. J. Lee, A. Manchon, and M. D. Stiles, Current induced torques and interfacial spin-orbitcoupling: Semiclassical modeling, P h y s .R e v .B 87, 174411 (2013) . [37] V . P. Amin and M. D. Stiles, Spin transport at interfaces with spin-orbit coupling: Formalism, Phys. Rev. B 94, 104419 (2016) . [38] V . P. Amin and M. D. Stiles, Spin transport at interfaces with spin-orbit coupling: Phenomenology, P h y s .R e v .B 94, 104420 (2016) . [39] K.-W. Kim, Spin transparency for the interface of an ultrathin magnet within the spin dephasing length, Phys. Rev. B 99, 224415 (2019) . [40] K. Garello, I. M. Miron, C. O. Avci, F. Freimuth, Y . Mokrousov, S. Blügel, S. Auffret, O. Boulle, G. Gaudin, and P. Gambardella,Symmetry and magnitude of spin-orbit torques in ferromagneticheterostructures, Nat. Nanotechnol. 8, 587 (2013) . [41] E. Sagasta, Y . Omori, M. Isasa, M. Gradhand, L. E. Hueso, Y . Niimi, Y . Otani, and F. Casanova, Tuning the spin Hall effectof Pt from the moderately dirty to the superclean regime, Phys. Rev. B 94, 060412(R) (2016) . [42] H. Moriya, A. Musha, S. Haku, and K. Ando, Observation of the crossover between metallic and insulating regimes of thespin Hall effect, Commun. Phys. 5, 12 (2022) . [43] H. Nakayama, M. Althammer, Y .-T. Chen, K. Uchida, Y . Kajiwara, D. Kikuchi, T. Ohtani, S. Geprägs, M. Opel, S.Takahashi, R. Gross, G. E. W. Bauer, S. T. B. Goennenwein,and E. Saitoh, Spin Hall Magnetoresistance Induced by aNonequilibrium Proximity Effect, Phys. Rev. Lett. 110, 206601 (2013) . [44] C. Du, H. Wang, F. Yang, and P. C. Hammel, Systematic variation of spin-orbit coupling with d-orbital filling: Large inverse spin Hall effect in 3 dtransition metals, Phys. Rev. B 90, 140407(R) (2014) . [45] C. O. Avci, K. Garello, A. Ghosh, M. Gabureac, S. F. Alvarado, and P. Gambardella, Unidirectional spin Hall magnetoresis-tance in ferromagnet-normal metal bilayers, Nat. Phys. 11, 570 (2015) . [46] C. O. Avci, J. Mendil, G. S. D. Beach, and P. Gambardella, Origins of the Unidirectional Spin Hall Magnetoresistance inMetallic Bilayers, P h y s .R e v .L e t t . 121, 087207 (2018) . [47] K. Ueda, C.-f. Pai, A. J. Tan, M. Mann, and G. S. D. Beach, Effect of rare earth metal on the spin-orbit torque in magneticheterostructures, Appl. Phys. Lett. 108, 232405 (2016) . [48] Q. Y . Wong, C. Murapaka, W. C. Law, W. L. Gan, G. J. Lim, and W. S. Lew, Enhanced Spin-Orbit Torques in Rare-EarthPt/[Co/Ni] 2/Co/Tb Systems, Phys. Rev. Applied 11, 024057 (2019) . [49] W. Wang, T. Wang, V . P. Amin, Y . Wang, A. Radhakrishnan, A. Davidson, S. R. Allen, T. J. Silva, H. Ohldag, D. Balzar, B. L.Zink, P. M. Haney, J. Q. Xiao, D. G. Cahill, V . O. Lorenz, andX. Fan, Anomalous spin-orbit torques in magnetic single-layerfilms, Nat. Nanotechnol. 14, 819 (2019) .[50] M. Schreier, G. E. W. Bauer, V . I. Vasyuchka, J. Flipse, K.- i. Uchida, J. Lotze, V . Lauer, A. V . Chumak, A. A. Serga, S.Daimon, T. Kikkawa, E. Saitoh, B. J. van Wees, B. Hillebrands,R. Gross, and S. T. B. Goennenwein, Sign of inverse spin Hallvoltages generated by ferromagnetic resonance and temperaturegradients in yttrium iron garnet platinum bilayers, J. Phys. D: Appl. Phys. 48, 025001 (2015) . [51] C. O. Avci, K. Garello, M. Gabureac, A. Ghosh, A. Fuhrer, S. F. Alvarado, and P. Gambardella, Interplay of spin-orbittorque and thermoelectric effects in ferromagnet-normal-metalbilayers, Phys. Rev. B 90, 224427 (2014) . [52] M. H. Nguyen, D. C. Ralph, and R. A. Buhrman, Spin Torque Study of the Spin Hall Conductivity and Spin Diffusion Lengthin Platinum Thin Films with Varying Resistivity, Phys. Rev. Lett.116, 126601 (2016) . [53] T. C. Chuang, C. F. Pai, and S. Y . Huang, Cr-induced Perpen- dicular Magnetic Anisotropy and Field-Free Spin-Orbit-TorqueSwitching, Phys. Rev. Applied 11, 061005 (2019) . [54] V . P. Amin, J. Li, M. D. Stiles, and P. M. Haney, Intrinsic spin currents in ferromagnets, P h y s .R e v .B 99, 220405(R) (2019) . [55] C. Stamm, C. Murer, M. Berritta, J. Feng, M. Gabureac, P. M. Oppeneer, and P. Gambardella, Magneto-Optical Detection ofthe Spin Hall Effect in Pt and W Thin Films, P h y s .R e v .L e t t . 119, 087203 (2017) . [56] K. Ounadjela, P. Vennegues, Y . Henry, A. Michel, V . Pierron- Bohnes, and J. Arabski, Structural changes in metastableepitaxial Co/Mn superlattices, Phys. Rev. B 49, 8561 (1994) . [57] G. M. Luo, H. W. Jiang, C. X. Liu, Z. H. Mai, W. Y . Lai, J. Wang, and Y . F. Ding, Chemical intermixing at FeMn/Cointerfaces, J. Appl. Phys. (Melville, NY) 91, 150 (2002) . [58] W. Zhang, W. Han, X. Jiang, S.-H. Yang, and S. S. P. Parkin, Role of transparency of platinum-ferromagnet interfaces in de-termining the intrinsic magnitude of the spin Hall effect, Nat. Phys. 11, 496 (2015) . [59] D. Go, D. Jo, K.-W. Kim, S. Lee, M.-G. Kang, B.-G. Park, S. Blügel, H.-W. Lee, and Y . Mokrousov, Long-range orbitaltransport in ferromagnets, arXiv:2106.07928 . [60] P. Chowdhury, P. D. Kulkarni, M. Krishnan, H. C. Barshilia, A. Sagdeo, S. K. Rai, G. S. Lodha, and D. V . Sridhara Rao, Ef-fect of coherent to incoherent structural transition on magneticanisotropy in Co/Pt multilayers, J. Appl. Phys. (Melville, NY) 112, 023912 (2012) . [61] M. D. Stiles and A. Zangwill, Anatomy of spin-transfer torque, Phys. Rev. B 66, 014407 (2002) . [62] A. Ghosh, S. Auffret, U. Ebels, and W. E. Bailey, Penetration Depth of Transverse Spin Current in Ultrathin Ferromagnets,Phys. Rev. Lett. 109, 127202 (2012) . [63] N. Reynolds, P. Jadaun, J. T. Heron, C. L. Jermain, J. Gibbons, R. Collette, R. A. Buhrman, D. G. Schlom, and D. C. Ralph,Spin Hall torques generated by rare-earth thin films, Phys. Rev. B95, 064412 (2017) . [64] J. Shi, P. Zhang, D. Xiao, and Q. Niu, Proper Definition of Spin Current in Spin-Orbit Coupled Systems, P h y s .R e v .L e t t . 96, 076604 (2006) . [65] L. Liu, T. Moriyama, D. C. Ralph, and R. A. Buhrman, Spin- Torque Ferromagnetic Resonance Induced by the Spin HallEffect, Phys. Rev. Lett. 106, 036601 (2011) . [66] S. Han, H.-W. Lee, and K.-W. Kim, Orbital Dynamics in Cen- trosymmetric Systems, P h y s .R e v .L e t t . 128, 176601 (2022) . 033037-14
1211.2055v1.Metal_insulator_transition_in_three_band_Hubbard_model_with_strong_spin_orbit_interaction.pdf
arXiv:1211.2055v1 [cond-mat.str-el] 9 Nov 2012Metal-insulator transition in three-band Hubbard model wi th strong spin-orbit interaction Liang Du and Xi Dai Beijing National Laboratory for Condensed Matter Physics, and Institute of Physics, Chinese Academy of Sciences, Beij ing 100190, China Li Huang Beijing National Laboratory for Condensed Matter Physics, and Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China and Science and Technology on Surface Physics and Chemistry Lab oratory, P.O. Box 718-35, Mianyang 621907, Sichuan, China (Dated: June 19, 2018) Recent investigations suggest that both spin-orbit coupli ng and electron correlation play very crucial roles in the 5 dtransition metal oxides. By using the generalized Gutzwill er variational methodanddynamicalmean-fieldtheorywith thehybridizati onexpansioncontinuoustime quantum Monte Carlo as impurity solver, the three-band Hubbard mode l with full Hund’s rule coupling and spin-orbit interaction terms, which contains the essentia l physics of partially filled t2gsub-shell of 5dmaterials, is studied systematically. The calculated phas e diagram of this model exhibits three distinct phase regions, including metal, band insulator an d Mott insulator respectively. We find that the spin-orbit coupling term intends to greatly enhance the tendency of the Mott insulator phase. Furthermore, the influence of the electron-electron intera ction on the effective strength of spin-orbit coupling in the metallic phase is studied in detail. We concl ude that the electron correlation effect on the effective spin-orbit coupling is far beyond the mean-fi eld treatment even in the intermediate coupling region. I. INTRODUCTION The Mott metal-insulator transition (MIT) induced by electron-electron correlation has attracted intensive re- searchactivities in the past several decades1–4. Although the main features of Mott transition have already been captured by single-band Hubbard model, most of Mott transition in realistic materials have multi-orbital nature and should be described by multi-band Hubbard model. Unlike the situation in single-band case, where the Mott transition is completely driven by the local Coulomb in- teraction U, the Mott transition in multi-band case is affected by not only Coulomb interaction but also crys- tal field splitting and Hund’s rule coupling among dif- ferent orbitals5–7. The interplay between Hund’s rule coupling and crystal field splitting generates lots of inter- esting phenomena in the multi-band Hubbard model, for examples, orbital selective Mott transition, high-spin to low-spin transition and orbital ordering. Therefore, most ofthe intriguingphysicsin 3 dor4dtransitionmetalcom- pounds can be well described bythe multi-band Hubbard model with both Hund’s rule coupling and crystal field splitting. In the present paper, we would like to concentrate our attention on the Mott physics in another group of inter- esting compounds, the 5 dtransition metal compounds, where spin-orbit coupling (SOC), the new physical in- gredient in Mott physics, plays an important role. Com- pared to 3 dorbitals, the 5 dorbitals are much more ex- tended and the correlation effects are not expected to be important here. While as firstly indicated in reference8, the correlation effects can be greatly enhanced by SOC,which is commonly strong in 5 dmaterials. The first well studied 5 dMott insulator with strong SOC is Sr 2IrO4, wheretheSOCsplitsthe t2gbandsinto(upper) jeff= 1/2 doublet and (lower) jeff= 3/2 quartet bands and greatly suppresses their bandwidths8–12. Since there are totally five electrons in its 5 dorbitals, the jeff= 1/2 bands are half filled and the jeff= 3/2 bands are fully occupied, which makes the system being effectively a jeff= 1/2 single-band Hubbard model with reduced bandwidth. Therefore the checkerboard anti-ferromagnetic ground state of Sr 2IrO4can be well described by the single-band Hubbard model with half filling. Here, we will focus on the 5 dmaterials with four elec- trons in the t2gsub-shell. These materials include the newly discovered BaOsO 3, CaOsO 3and NaIrO 3etc13. All these materials share one important common feature: in low temperature, these materials are insulators with- out magnetic long-range order. The origin of the insu- lator behavior can be due to two possible reasons, the strong enough Coulomb interaction and SOC. We will have Mott insulator in the former and band insulator in the latter case respectively. Therefore it is interesting to study the features of metal-insulator transition in a generict2gsystem occupied by four electrons with both Coulomb interaction and SOC. In the present paper, we study the t2gHubbard model withSOCandfourelectronsfillingbyusingrotationalin- variantGutzwillerapproximation(RIGA)anddynamical mean-field theory combined with the hybridization ex- pansion continuous time quantum Monte Carlo (DMFT + CTQMC) respectively. The paramagnetic U−ζphase diagram is derived carefully. Further, the interplay be-2 tween SOC ζand Coulomb interaction Uis analyzed in detail. We will mainly focus on the following two key is- sues: (i) How does the SOC affect the boundary of Mott transitions in this three-band model? (ii) How does the Coulomb interaction modify the effective SOC strength? Thispaperisorganizedasfollows. InSec. II,thethree- band model is specified, and the generalized multi-band Gutzwiller variational wave function is introduced. In Sec. IIIA, the calculated results, including U−ζphase diagram, quasi-particle weight and charge distribution, for the three-band model are presented. The effect of Coulomb interaction on SOC is analyzed in Sec. IIIB. Finally we make conclusions in section IV. II. MODEL AND METHOD The three-band Hubbard model with full Hund’s rule coupling and SOC terms is defined by the Hamiltonian: H=−/summationdisplay ij,aσtijd† i,aσdj,aσ+/summationdisplay iHi loc=Hkin+Hloc,(1) whereσdenotes electronic spin, and arepresents the threet2gorbitals with a= 1,2,3 corresponding to dyz,dzx,dxyorbitals respectively. The first term de- scribes the hopping process of electrons between spin- orbital state “ aσ” on different lattice sites iandj. Local Hamiltonian terms Hi loc=Hi u+Hi soccontain Coulomb interaction Hi uand SOC Hi soc(In the following, the site index is suppressed for sake of simplicity). Hu=U/summationdisplay ana↑na↓+U′/summationdisplay a<b,σσ′naσnbσ′−Jz/summationdisplay a<b,σnaσnbσ −Jxy/summationdisplay a<b/parenleftig d† a↑da↓d† b↓db↑+d† a↑d† a↓db↑db↓+h.c./parenrightig ,(2) Hsoc=/summationdisplay aσ/summationdisplay bσ′ζ/angbracketleftaσ|lxsx+lysy+lzsz|bσ′/angbracketrightd† aσdbσ′,(3) whereU(U′) denotes the intra-orbital (inter-orbital) Coulomb interaction, Jzterm describes the longitudinal part of the Hund’s coupling. While the other two Jxy terms describe the spin-flip and pair-hopping process re- spectively. ζis SOC strength, l(s) is orbital (spin) an- gular momentum operator. Here we assume the studied system experiences approximately cubic symmetry ( Oh symmetry), in which two parameters U′andJxyfollow the constraints U′=U−2JandJxy=Jz=J. Here we chooseJ/U= 0.25 for the systems studied in this paper unless otherwise noted. This lattice model is solved in the framework of RIGA14–17and DMFT(CTQMC)18,19 methods respectively, which are both exact in the limit of infinite spacial dimensions20,21. In this work, a semi- elliptic bare density of states ρ(ǫ) = (2/πD)/radicalbig 1−(ǫ/D)2 is adopted, which corresponds to Bethe lattice with infi- nite connectivity. In the present study, the energy unitis set to be half bandwidth D= 1 and all orbitals are assumed to have equal bandwidth. Next, we will briefly introduce the recently developed RIGA method16. The generalized Gutzwiller trial wave function |ΨG/angbracketrightcan be constructed by acting a many- particle projection operator Pon the uncorrelated wave function |Ψ0/angbracketright22–24, |ΨG/angbracketright=P|Ψ0/angbracketright, (4) with P=/productdisplay iPi=/productdisplay i/summationdisplay ΓΓ′λΓΓ′|Γ/angbracketrightii/angbracketleftΓ′|. (5) |Ψ0/angbracketrightis normalized uncorrelated wave function in which Wick’s theorem holds. |Γ/angbracketrightiare atomic eigenstates on siteiandλΓΓ′are Gutzwiller variational parameters. In our work, |Γ/angbracketrightiare eigenstates of atomic Hamiltonian Hu, each|Γ/angbracketrightiis labeled by good quantum number N,J,Jz, whereNis total number of electrons, Jis total angular momentum, Jzis projection of total angular momen- tum along zdirection. The non-diagonal elements of the previously defined variational parameter matrix λΓΓ′are assumed to be finite only for state |Γ/angbracketright,|Γ′/angbracketrightbelonging to the sameatomic multiplet, i.e, with the same three quan- tum labels15. In the following, we assume the local Fock terms are absent in |Ψ0/angbracketright, /angbracketleftΨ0|c† iαciβ|Ψ0/angbracketright=δαβ/angbracketleftΨ0|c† iαciα|Ψ0/angbracketright=δαβn0 iα.(6) For general case, a local unitary transformation matrix Ais needed to transform the original diα-basis into the so-called natural cim-basis16, i.e,dα=/summationtext mAαmcm. In the original single particle basis ( dyz↑,dyz↓,dzx↑,dzx↓, dxy↑,dxy↓), SOC term is expressed as : Hsoc=−ζ 2 0 0−i0 0 1 0 0 0 i−1 0 i0 0 0 0 −i 0−i0 0−i0 0−1 0i0 0 1 0i0 0 0 .(7) Then the transformation matrix Ais as follows: A=1√ 6 −√ 3 0 1 0 0 −√ 2 0−1 0√ 3−√ 2 0 −i√ 3 0−i0 0 i√ 2 0−i0−i√ 3−i√ 2 0 0 2 0 0 −√ 2 0 0 0 2 0 0√ 2 ,(8) where the t2gorbitals have been treated as a system with leff= 1. In the natural single particle basis, SOC matrix is transformed into: Hsoc= −ζ/2 0 0 0 0 0 0−ζ/2 0 0 0 0 0 0 −ζ/2 0 0 0 0 0 0 −ζ/2 0 0 0 0 0 0 ζ0 0 0 0 0 0 ζ .(9)3 Meanwhile the Coulomb interaction term is transformed as: /summationdisplay αβδγUαβδγd† αd† βdδdγ=/summationdisplay mnkl˜Umnklc† mc† nckcl,(10) with ˜Umnkl=/summationdisplay αβδγUαβδγA† mαA† nβAδkAγl.(11) Fixn0in each orbital initial guess of renormalization factor R construct Gutzwiller effective one-particle Hamiltonian, ˆHeff 0=/summationtext i/negationslash=j/summationtext αβ/summationtext γδtαβ ijR† αγˆc† iγˆcjδRδβ+/summationtext iαηαˆc† iαˆciα and solve the Hamiltonian to get ∂E/∂R. ∂E ∂λΓΓ′=/summationdisplay δβ/parenleftigg ∂Ekin ∂Rδβ∂Rδβ ∂λΓΓ′+∂Ekin ∂R† βδ∂R† βδ ∂λΓΓ′/parenrightigg +∂Eloc ∂λΓΓ′+/summationdisplay αηα∂n′ α ∂λΓΓ′= 0 solve the equations to get renormalization factor R check ifRis self-consistent. calculate quantities: total energy, occupation ...YesNo FIG. 1. Flowchart of the RIGA self-consistent loop to mini- mize total energy E(n0) with respect to |Ψ0/angbracketrightandλΓΓ′. In this paper, we define expectation value with uncor- related wave function: O0=/angbracketleftΨ0|ˆO|Ψ0/angbracketright, (12) while expectation value with Gutzwiller wave function is defined as: O=OG=/angbracketleftΨG|ˆO|ΨG/angbracketright. (13) During the minimization process, two following con- straints are forced, /angbracketleftΨ0|P†P|Ψ0/angbracketright= 1, (14) and /angbracketleftΨ0|P†Pniα|Ψ0/angbracketright=/angbracketleftΨ0|niα|Ψ0/angbracketright. (15) In the present paper, the second constraint is satisfied in the following way. We first calculate the total energy of the trial wave function with both the left-hand andright-hand side of the above equation equaling to some desired occupation number n0 α. Then we minimize the energy respect to n0 αat the last step. The remaining task is to minimize the variational ground energy E=Ekin+Elocwith respect to λΓΓ′ and|Ψ0/angbracketright, and fulfill the previous two constraints. Here, Ekin=/angbracketleftΨG|Hkin|ΨG/angbracketright=/summationdisplay ij/summationdisplay γδ˜tγδ ij/angbracketleftΨ0|c† iγcjδ|Ψ0/angbracketright, (16) and Eloc=/angbracketleftΨG|Hloc|ΨG/angbracketright= Tr(φ†Hlocφ),(17) with˜t,Randφdefined as: ˜tγδ ij=/summationdisplay αβtαβ ijR† αγRδβ (18) R† αγ=Tr/parenleftbig φ†c† αφcγ/parenrightbig /radicalig n0γ(1−n0γ), (19) φII′=/angbracketleftI|P|I′/angbracketright/radicalbig /angbracketleftΨ0|I′/angbracketright/angbracketleftI′|Ψ0/angbracketright, (20) where|I/angbracketright(|I′/angbracketright) stands for a many-body Fock state and n0 γ=/angbracketleftΨ0|nγ|Ψ0/angbracketright. The flowchart of RIGA method is shown in Fig.1. For fixedn0 αin each orbital, minimizing Ewith respect to |Ψ0/angbracketrightandλΓΓ′can be divided into two steps in each iter- ative process. Firstly, fix Gutzwiller variational parame- tersλΓΓ′and find optimal uncorrelated wave function by solving effective single particle Hamiltonian, Heff 0=/summationdisplay i/negationslash=j/summationdisplay γδ˜tγδ ijc† iγcjδ+/summationdisplay iαηαc† iαciα,(21) where Lagrange parameters ηαare used to minimize the variational energy fulfilling Gutzwiller constraints. Sec- ondly, we fix the uncorrelated wave function, and opti- mize the variational energy with respect to Gutzwiller variational parameters λΓΓ′, ∂E ∂λΓΓ′=/summationdisplay δβ/parenleftigg ∂Ekin ∂Rδβ∂Rδβ ∂λΓΓ′+∂Ekin ∂R† βδ∂R† βδ ∂λΓΓ′/parenrightigg +∂Eloc ∂λΓΓ′+/summationdisplay αηα∂n′ α ∂λΓΓ′= 0, (22) wheren′ α=/angbracketleftΨ0|P†Pnα|Ψ0/angbracketright. In this way, λΓΓ′and|Ψ0/angbracketright are self-consistently determined. For the fix n0 αalgorithm, we need to scan the n0 αto get the global ground state of the studied system. In this paper, because SOC will split the t2gorbitals into two fold jeff= 1/2 and four fold jeff= 3/2 states, we can introduce an alternative variable δn0to determine n0 αfor4 each orbital. The occupation polarization δn0is defined as: δn0=n0 3/2−n0 1/2, (23) in which n0 3/2andn0 1/2stand for the average occupa- tion number of lower ( jeff= 3/2) and upper ( jeff= 1/2) orbitals respectively. Since total electron number of the system is fixed to be 4 n0 3/2+ 2n0 1/2= 4, we have 0≤δn0≤1.δn0(n0) corresponding to ground state is denoted by δn0 g(n0 g). In the present paper, we also use DMFT+CTQMC method to crosscheck our results derived by RIGA. For DMFT+CTQMC method, the system temperature is set to beT= 0.025 (corresponding to inverse temperature β= 40). In each DMFT iteration, typically 4 ×108 QMC samplings have been performed to reach sufficient numerical accuracy25. III. RESULTS AND DISCUSSION A.U-ζphase diagram In this subsection, we mainly focus on phase diagram for the three-band model proposed in Eq.(1). The ob- tainedU−ζphase diagrams with J/U= 0.25 are shown in Fig.2. The upper panel shows the phase diagram cal- culated by zero temperature RIGA method, while the calculated results by DMFT+CTQMC method at finite temperature is shown in the lower panel. The results ob- tained by two different methods are consistent with each other quite well except that DMFT+CTQMC can not distinguish between band insulator and Mott insulator, which will be explained later. Apparently, there exists three different phasesin U−ζplane: metallic state in the lower left corner, band insulator in the lower right region and Mott insulator in the upper right region. The gen- eral shape of the phase diagram can be easily understood by considering two limiting cases: (i) For ζ= 0, one has a degenerate three-band Hubbard model populated by 4 electrons per site. The model will undergo an interaction driven Mott transition at critical Uc/D∼11.0 with each band filled by 4 /3 electrons. (ii) For non-interacting case (U= 0.0), the model is exactly soluble. The three bands are degenerate and filled by 4 /3 electrons at ζ=0.0. Fi- niteζwill split the three degenerate bands into a (lower) jeff= 3/2quadrupletand(upper) jeff= 1/2doubletwith energy separation being 1 .5ζ. Increasing ζwill pump electrons from upper to lower orbitals until the upper bands are completely empty and the system undergoes a metal to band insulator transition, which is expected at ζ/D= 1.33. In order to clarify the way we determine the metal, band insulator, and Mott insulator phases by RIGA method, in Fig.3 we plot the total energy and quasipar- ticle weight as a function of δn0defined in the previous section , where the SOC strength is fixed at ζ/D= 0.7, U/D ζ/D 0 2 4 6 8 10 12 0 0.2 0.4 0.6 0.8 1 1.2 1.4MetalBand InsulatorMott InsulatorU/D ζ/D 0 2 4 6 8 10 12 0 0.2 0.4 0.6 0.8 1 1.2 1.4MetalInsulator FIG. 2. (Color online) Phase diagram of three-band Hub- bard model with full Hund’s coupling terms in the plane of Coulomb interaction U(J/U= 0.25) and spin-orbit coupling ζ. Upper panel: The phase diagram is calculated by RIGA at zero temperature. Lower panel: The phase diagram is calcu- lated by DMFT+CTQMC with finite temperature T= 0.025. and from top to bottom the Coulomb interaction is U/D= 1.0,3.0,6.0. The ground state of the system is the state with the lowest energy respect to δn0. The typical solution for the metal phase is shown in Fig.3a, where the energy minimum occurs at 0 < δn0 g<1.0 cor- responding to the case that all orbitals being partially occupied. While for a band insulator, as shown in Fig.3b as a typical situation, the energy minimum happens at δn0 g= 1.0 corresponding to the case that the jeff= 3/2 orbitals are fully occupied and jeff= 1/2 orbitals are empty, and more over the quasiparticle weight Zkeeps finite when δn0approching unit. And finally the situ- ation of a Mott insulator is illustrated in Fig.3c, where the quasiparticle weight Zvanishes at some critical δn0, above which the system is in Mott insulator phase and can no longer be described by the Gutzwiller variational method. While in the DMFT+CTQMC calculations, the phase boundary between metal and insulator is identified by measuring the imaginary-time Green function at τ= β/226,27. SinceG(β/2) can be viewed as a representa-5 1.601.802.00 (a) 0.91.0 (d) 7.407.607.80E (b) 0.50.60.7 Z (e) 16.116.216.3 0.0 0.2 0.4 0.6 0.8 δn0(c) 0.0 0.2 0.4 0.6 0.8 1.00.00.10.20.3 δn0(f)U=1 U=1,j=3/2 U=1,j=1/2 U=3 U=3,j=3/2 U=3,j=1/2 U=6 U=6,j=3/2 U=6,j=1/2 FIG. 3. (Color online) Total energy E(δn0) and quasipar- ticle weight Z(δn0) as a function of occupation polarization δn0=n0 3/2−n0 1/2for different values of interaction strength U/D= 1,3,6 (J/U= 0.25) at fixed ζ/D= 0.7 and zero temperature, where n0 3/2is average occupation number of jeff= 3/2 quadruplet, n0 1/2is average occupation number of jeff= 1/2 doublet. tion of the integrated spectral weight within a few kBT ofEF, so it can be used as an important criterion to judge whether metal-insulator transition occurs. The corresponding results for SOC strength ζ/D= 0.5 are shown in Fig.4. Clearly seen in this figure, the critical Uc is about 3.5, and both the jeff= 3/2 andjeff= 1/2 orbitals undergo metal-insulator transitions simultane- ously. Since there is chemical potential ambiguity in the insulator phase, it is difficult to further distinguish Mott insulator from band insulator by DMFT+CTQMC method. Therefore we only calculatethe phase boundary between the metal and insulator phase, which is in good agreement with the results obtained by RIGA. Now, we come back to discussed the phase diagram obtained by RIGA and DMFT. When both the Coulomb interaction Uand SOC ζare finite, the phase diagram looks a bit complicated. By considering different values ofSOC, we divide the phase diagramverticallyinto three regions. Firstly, for 0 .00< ζ/D < 0.24, with increasing Coulomb interaction U, our calculation by RIGA pre- dicts a transition from metal to Mott insulator. The transition is characterizedby the vanishing of quasiparti- cle weight as discussed previously. The critical Coulomb interaction Ucdecreasesdrasticallywith the increment of SOC, the effect of SOC tends to enhance the Mott MIT greatly. The DMFT results show very similar behavior in this region as shown in lower panel of Fig.2, except that the Ucobtained by DMFT has weaker dependence on the strength of SOC compared to that of RIGA. From the view point of DMFT, the suppression of the metallic phase by SOC can be explained quite clearly. For the effective impurity model in DMFT, the metallic phase corresponds to a solution when the local moments on the0.20.40.60.81.0 1 2 3 4 5 6 7 8 9 10|G(β/2)| U/Dζ=0.5j=3/2 j=1/2 FIG. 4. (Color online) The imaginary-time Green function atτ=β/2 as a function of Coulomb interaction strength U. The SOC strength ζis chosen to be 0.5 as a illustration. The calculation is done by DMFT+CTQMC method at β= 40. In this figure the normalized quantities by G(β/2) atU/D= 1.0 are shown and the arrows correspond to phase transition points. impurity site are fully screened by the electrons in heat- bath through the Kondo like effect. With the SOC, there is an additional channel to screen the local spin moment otherthanthe Kondoeffect, whichleadstothe formation of spin-orbital singlets. This additional screening chan- nel, which is completely local, will thus compete with the Kondo effect and suppress the metallic solution. There is no net local moment left in this type of Mott phase, and the ground state is simply a product state of local spin-orbital singlets on each site. For 0.24< ζ/D < 1.33, two successive phase transi- tions are observed with the increment of U. The transi- tion from metallic to band insulator phases occurs firstly, and followed by another transition to the Mott insula- tor phase. In the intermediate Uregion, the effective band width of system is reduced by the correlation ef- fects, whichdrivesthe systemintoabandinsulatorphase with relative small band splitting induced by SOC. Fur- ther increasing interaction strength Uwill push the sys- tem to the Mott limit. Although this process is believed to be a crossover rather than a sharp phase transition, our RIGA calculation provides a mean field description for these two different insulators, where in the band in- sulator phase the interaction effects only renormalize the effective band structure and do not suppress the coher- ent motion of the electrons entirely. Similar behavior can also be obtained by DMFT method, where the quasipar- ticle weight determined by DMFT selfenergy keeps finite for the band insulator phase and vanishes for the Mott insulator phase. At last, for ζ/D >1.33 region, the orbitals are fully polarized with electrons fully occupied jeff= 3/2 bands and fully empty jeff= 1/2 bands at U= 0.0, indicating that the system is in the band insulator state already6 in the non-interacting case. Similar band insulator to Mott insulator transition will be induced with further increment of Uin the RIGA description, as discussed before. 0.00.20.40.60.8Zζ=0.1, j=3/2 ζ=0.5, j=3/2 ζ=1.5, j=3/2 0.00.20.40.60.8 0 1 2 3 4 5 6 7Z U/Dζ=0.1, j=1/2 ζ=0.5, j=1/2 ζ=1.5, j=1/2 FIG. 5. (Color online) Quasiparticle renormalization fact ors Zof the lower orbitals ( jeff= 3/2 quadruplet) and upper orbitals ( jeff= 1/2 doublet) as function of Coulomb inter- actionU(J/U= 0.25) for different values of SOC ( ζ/D= 0.1,0.5,1.5). The dashed lines label the critical Ufor transi- tion to Mott state. The results are obtained by zero temper- ature RIGA method. 0.650.700.750.800.850.900.951.00 0 1 2 3 4n(jeff= 3/2) U/Dζ=0.1, RIGA ζ=0.1, DMFT ζ=0.5, RIGA ζ=0.5, DMFT ζ=1.5, RIGA ζ=1.5, DMFT FIG. 6. (Color online) Occupation number of the lower orbitals ( jeff= 3/2 quadruplet) with increasing Coulomb U(J/U= 0.25) for selected SOC ( ζ/D= 0.1,0.5,1.5). Both thecalculated results byRIGAandDMFT(CTQMC) are pre- sented. For several typical SOC parameters ( ζ/D= 0.1,0.5,1.5) in the three regions defined above, we study the evolutions of quasiparticle weight and band specific occupancy with Coulomb interaction. The quasiparti- cle weight for selected SOC with increasing Uis plotted in Fig.5. The upper (lower) panel shows the quasiparti- cle weight for jeff= 3/2 (1/2) orbitals. Note the quasi- particle weight in RIGA is defined as the eigenvalues of the Hermite matrix R†R. For both ζ/D= 0.1 and 1.5,1.71.92.1/angbracketleftL2/angbracketright(a) 1.01.41.8/angbracketleftS2/angbracketright(b) 0.00.51.01.5 0 0 .5 1 1 .5 2 2 .5 3 3 .5 4/angbracketleftJ2/angbracketright U/D(c)DMFT RIGA DMFT RIGA DMFT RIGA FIG. 7. (Color online) Expectation value of orbital angular momentum /angbracketleftL2/angbracketright, spin angular momentum /angbracketleftS2/angbracketright, and total an- gular momentum /angbracketleftJ2/angbracketrightas function of Coulomb interaction U with fixed spin-orbit coupling strength ζ/D= 0.7. It is de- rived by RIGA at zero temperature and DMFT+CTQMC at β= 40 respectively. the quasiparticle weights decrease from 1 to 0 monoton- ically with the increasing interaction strength UandJ until the transition to Mott insulator phase. While for ζ/D= 0.5, there exists a kink at U/D= 2.7 in the lower panel (jeff= 1/2), which corresponds to the transition from metal to band insulating state. For transition to Mott insulating state, quasiparticle weights for all the orbitals reach zero simultaneously, with Uc/D= 6.7 for ζ/D= 0.1,Uc/D= 4.5 for ζ/D= 0.5, and Uc/D= 4.0 forζ/D= 1.5. The occupation number of the (lower) jeff= 3/2 or- bitals as a function of on-site Coulomb interaction U is ploted in Fig.6 for three typical SOC strength. For ζ/D= 0.1, to some extent, the occupation behavior is similar to ζ= 0 case, in which the occupation number is only slightly changed by the interaction. The situa- tion is quite different for ζ/D= 0.5, where the occupa- tion of the jeff= 3/2 orbital increases with interaction at the beginning and decreases slightly after the tran- sition to the band insulator phase. The non-monotonic behavior here is mainly due to the competition between the repulsive interaction Uand Hund’s rule coupling J. The effect of Uwill always enhance the splitting of the local orbitals to reduce the repulsive interaction among these orbitals. While the Hund’s rule coupling intents to distribute the electrons more evenly among different orbitals. For ζ/D= 1.5 case, occupation number in the two subsets is fully polarized at U= 0 and the effect of Hund’s coupling term will reduce the occupation of the jeff= 3/2 orbital monotonically. At last, the expectation value of the total orbital angu- lar momentum /angbracketleftL2/angbracketright, spin angular momentum /angbracketleftS2/angbracketrightand total angular momentum /angbracketleftJ2/angbracketrightas a function of Coulomb interaction Uare plotted in Fig.7, where ζ/Dis fixed to 0.70. In the non-interacting case, all the three ex-7 pectation values can be calculated exactly and they will approach the atomic limit with the increment of inter- actionUandJ. In the atomic limit the SOC strength is much weaker than the Hund’s coupling J, the ground state can be well described by the LScoupling scheme, where the four electrons will first form a state with total orbital angular momentum L= 1 and total spin momen- tumS= 1,andthenformaspin-orbitalsingletstatewith total angularmomentum J= 0. From Fig.7, we can find that the system approaches the spin-orbital singlet quite rapidly after the transition to the band insulator phase. 0.81.01.21.41.61.82.02.22.42.62.8 0 0.5 1 1.5 2 2.5ζeff/ζ U/Dζ=0.1, RIGA ζ=0.1, HFA ζ=0.5, RIGA ζ=0.5, HFA FIG. 8. (Color online) Evolution of effective SOC strength with increasing Coulomb U(J/U= 0.25) for selected SOC (ζ/D= 0.1,0.5), A comparison of results derived by RIGA and HFA are presented. B. Effective spin-orbit coupling In the multi-orbital system, the interaction effects will mainly cause two consequences for the metallic phases: (1) It will introduce renormalizationfactor for the energy bands; (2) It will modify the local energy level for each orbital which splits the bands. For the present model, the second effect will modify the effective SOC, which is another very important problem for the spin orbital coupled correlation system. Within the Gutzwiller vari- ational scheme used in the present paper, the effective SOC can be defined as: ζeff=−1 2∂Eint(δn0) ∂δn0−1 2∂Esoc(δn0) ∂δn0,(24) whereEintandEsocare the ground state expectation values of interaction and SOC terms in the Hamilto- nian respectively. Note the second term is different from the bare SOC ζ0unlessnαis a good quantum num- ber. If the interaction energy is treated by Hartree Fock mean field approximation (HFA), the above equation givesζeff=−∂EHF int(δn0)/(2∂δn0) +ζ0, which will al- ways greatly enhance the spin-orbital splitting with the increasing Uas found by some works based on LDA+Umethod28,29. In this section, we compare the results ob- tained by RIGA and HFA. As shown in Fig.8, the effec- tive SOC obtained by HFA increases quite rapidly with the interaction UandJ. While the results obtained by RIGA show very different behavior. For weak SOC strength, i.e. ζ/D= 0.1, the effective SOC obtained by RIGA increases first then decrease. This interesting non-monotonicbehaviorreflectsthecompetitionbetween the repulsive interaction U, which intends to increase the occupation difference for jeff= 1/2 andjeff= 3/2 or- bitals, and the Hund’s rule coupling J, which intents to decrease the occupation difference. While for relatively strong SOC strength, the effective SOC increases with interaction U(andJ) monotonically all the way to the phase boundary indicating the repulsive interaction U plays a dominate role here. But compared to HFA, the enhancement of effective SOC induced by the interaction is much weaker even for the latter case. This is mainly due to the reduction of the high energy local configura- tions in the Gutzwiller variational wave function com- pared to the Hatree Fock wave function, which greatly reduces the interaction energy and its derivative to the orbital occupation. IV. CONCLUDING REMARKS The Mott MIT in three-band Hubbard model with full Hund’s rule coupling and SOC is studied in detail using RIGA and DMFT+CTQMC methods. First, we propose the phase diagram with the strength of electron-electron interaction and SOC. Three different phases have been found in the U−ζplane, which are metal, band insula- tor and Mott insulator phases. For 0 .00< ζ/D < 0.24, increasing Coulomb interaction will induce a MIT tran- sition from metal to Mott insulator. For 0 .24< ζ/D < 1.33,effect of Uwill causetwosuccessivetransitions, first frommetaltobandinsulator, thentoMottinsulator. For ζ/D >1.33, a transition from band insulator to Mott in- sulatorisobserved. Fromthephasediagram,wefindthat the critical interaction strength Ucis strongly reduced by the presence of SOC, which leads to the conclusion that the SOC will greatly enhance the strong correlation ef- fects in these systems. Secondly, we have studied the ef- fect of electron-electron interaction on the effective SOC. Our conclusion is that the enhancement of effective SOC found in HFA is strongly suppressed once we go beyond the mean field approximationand include the fluctuation effects by RIGA or DMFT methods. ACKNOWLEDGMENT We acknowledge valuable discussions with profes- sor Y.B. Kim and professor K. Yamaura, and finan- cial support from the National Science Foundation of China and that from the 973 program under Contract No.2007CB925000 and No.2011CBA00108. The DMFT8 + CTQMC calculations have been performed on the SHENTENG7000 at Supercomputing Center of Chinese Academy of Sciences (SCCAS). 1M. Imada, A. Fujimori, and Y. Tokura, Rev. Mod. Phys. 70, 1039 (1998). 2A. Georges, G. Kotliar, W. Krauth, and M. J. Rozenberg, Rev. Mod. Phys. 68, 13 (1996). 3G. Kotliar, S. Y. Savrasov, K. Haule, V. S. Oudovenko, O. Parcollet, and C. A. Marianetti, Rev. Mod. Phys. 78, 865 (2006). 4L. de’ Medici, J. Mravlje, and A. Georges, Phys. Rev. Lett. 107, 256401 (2011). 5L. de’ Medici, S. R. Hassan, M. Capone, and X. Dai, Phys. Rev. Lett. 102, 126401 (2009). 6P. Werner and A. J. Millis, Phys. Rev. Lett. 99, 126405 (2007). 7T. Kita, T. Ohashi, and N. Kawakami, Phys. Rev. B 84, 195130 (2011). 8B. J. Kim, H. Jin, S. J. Moon, J. Y. Kim, B. G. Park, C. S. Leem, J. Yu, T. W. Noh, C. Kim, S. J. Oh, J. H. Park, V. Durairaj, G. Cao, and E. Rotenberg, Phys. Rev. Lett. 101, 076402 (2008). 9D. Pesin and L. Balents, Nature Physics 6, 376 (2010). 10B. J. Kim, H. Ohsumi, T. Komesu, S. Sakai, T. Morita, H. Takagi, and T. Arima, Science 323, 1329 (2009). 11H. Watanabe, T. Shirakawa, and S. Yunoki, Phys. Rev. Lett. 105, 216410 (2010). 12G. Jackeli and G. Khaliullin, Phys. Rev. Lett. 102, 017205 (2009). 13M. Bremholm, S. Dutton, P. Stephens, and R. Cava, Journal of Solid State Chemistry 184, 601 (2011).14J. B¨ unemann, W. Weber, and F. Gebhard, Phys. Rev. B 57, 6896 (1998). 15J. B¨ unemann, F. Gebhard, T. Ohm, S. Weiser, and W. Weber, Phys. Rev. Lett. 101, 236404 (2008). 16N. Lanat` a, H. U. R. Strand, X. Dai, and B. Hellsing, Phys. Rev. B 85, 035133 (2012). 17X. Deng, L. Wang, X. Dai, and Z. Fang, Phys. Rev. B 79, 075114 (2009). 18P. Werner and A. Millis, Phys. Rev. B 74, 1 (2006). 19E. Gull, A. J. Millis, A. I. Lichtenstein, A. N. Rubtsov, M. Troyer, and P. Werner, Rev. Mod. Phys. 83, 349 (2011). 20W. Metzner and D. Vollhardt, Phys. Rev. Lett. 59, 121 (1987). 21W. Metzner and D. Vollhardt, Phys. Rev. B 37, 7382 (1988). 22M. C. Gutzwiller, Phys. Rev. Lett. 10, 159 (1963). 23M. C. Gutzwiller, Phys. Rev. 134, A923 (1964). 24M. C. Gutzwiller, Phys. Rev. 137, A1726 (1965). 25L. Huang, L. Du, and X. Dai, Phys. Rev. B 86, 035150 (2012). 26A. Liebsch, Phys. Rev. Lett. 91, 226401 (2003). 27L. Huang, Y. Wang, and X. Dai, Phys. Rev. B 85, 245110 (2012). 28G.-Q. Liu, V. N. Antonov, O. Jepsen, and O. K. Ander- sen., Phys. Rev. Lett. 101, 026408 (2008). 29A. Subedi, Phys. Rev. B 85, 020408 (2012).
1110.6798v1.Spin_Orbit_Coupled_Quantum_Gases.pdf
November 1, 2011 0:15 WSPC/INSTRUCTION FILE Review_SOC Spin-Orbit Coupled Quantum Gases Hui Zhai Institute for Advanced Study, Tsinghua University, Beijing, 100084, China In this review we will discuss the experimental and theoretical progresses in studying spin-orbit coupled degenerate atomic gases during the last two years. We shall rst review a series of pioneering experiments in generating synthetic gauge potentials and spin-orbit coupling in atomic gases by engineering atom-light interaction. Realization of spin-orbit coupled quantum gases opens a new avenue in cold atom physics, and also brings out a lot of new physical problems. In particular, the interplay between spin-orbit coupling and inter-atomic interaction leads to many intriguing phenomena. Here, by reviewing recent theoretical studies of both interacting bosons and fermions with isotropic Rashba spin- orbit coupling, the key message delivered here is that spin-orbit coupling can enhance the interaction e ects, and make the interaction e ects much more dramatic even in the weakly interacting regime. Keywords : Cold Atoms, Synthetic Gauge Potential, Spin-Orbit Coupling, Super uidity, Feshbach Resonance Many interesting phenomena in condensed matter physics occur when electrons are placed in an electric or magnetic eld, or possess strong spin-orbit (SO) cou- pling. However, in the cold atom systems, neutral atoms neither possess gauge coupling to electromagnetic elds nor have SO coupling. Recently, by controlling atom-light interaction, one can generate an arti cial external abelian or non-abelian gauge potential coupled to neutral atoms. The basic principle is based on the Berry phase e ect1and its non-abelian generalization2. An important application of this scheme is creating an e ective SO coupling in degenerate atomic gases. Since 2009, Spielman's group in NIST has successfully implemented this principle and gener- ated both synthetic uniform gauge eld6, magnetic eld7, electric eld8and SO coupling9. We shall discuss the experimental progresses along this line in Sec. 1. The e ects of SO coupling in electronic systems have been extensively discussed in condensed matter physics before and are also important topics nowadays. One of the most famous example is recently discovered topological insulators3;4;5. In this review, we try to convey the message that SO coupling in degenerate atomic gases will bring out new physics which have not been considered before, mainly due to the interplay between SO coupling and the unique properties of atomic gases. For bosonic atoms, SO coupled interacting bosons is a system never explored in physics before. For fermionic atoms, since a lot of intriguing physics have been revealed during the last ten years by utilizing Feshbach resonance technique to achieve in- teraction as strong as Fermi energy, the interplay between resonance physics and SO coupling is de nitely a subject of great interests. 1arXiv:1110.6798v1 [cond-mat.quant-gas] 31 Oct 2011November 1, 2011 0:15 WSPC/INSTRUCTION FILE Review_SOC 2Hui Zhai A key point we want to emphasize in this review is that a nearly isotropic SO coupling will dramatically enhance the e ects of inter-particle interactions, so that the interaction e ects are not weak even in the regime where the interaction strength itself is small . This is because an isotropic Rashba SO coupling or a nearly isotropic SO coupling signi cantly changes the low-energy states of single particle Hamiltonian, as we shall discuss in Sec 2. In Sec. 3, we discuss many-body system of bosons. Because the single particle ground state has large degeneracy, it is the inter-particle interaction that selects out a unique many-body ground state and determines its low-energy uctuations. In Sec. 4, we discuss many-body system of fermions. Because the low-energy density-of-state (DOS) is largely enhanced, the interaction e ects become much more profound, in particular for weak attractions. A brief summary and future perspective are given in Sec. 5. 1. Synthetic Gauge Potentials and Spin-Orbit Coupling In 2009, Spielman's group in NIST rst realized a uniform vector potential in Bose- Einstein condensate (BEC) of87Rb6. In this experiment, two counter propagating Raman laser beams couple jF;m Fi=j1;1ilevel of87Rb toj1;0ilevel, and couple j1;0ilevel toj1;1ilevel, as shown in Fig. 1(a) and (b), which can be described by the Hamiltonian H=0 B@k2 x 2m+1 2ei2k0x0 2ei2k0xk2 x 2m 2ei2k0x 0 2ei2k0xk2 x 2m21 CA (1) wherek0= 2=,is the wave length of two lasers. 2 k0is therefore the momentum transfer during the two-photon processes. 1=  1+!+2, and2=  1+!2, where  1denotes the linear Zeeman energy, !denotes the frequency di erence of two Raman lasers, and  2is the quadratic Zeeman energy. Applying a unitary transformation to wave function  = U , where U=0 @ei2k0x0 0 0 1 0 0 0ei2k0x1 A (2) and the e ective Hamiltonian becomes He =UHUy=0 B@(kx+2k0)2 2m+1 20 2k2 x 2m 2 0 2(kx2k0)2 2m21 CA: (3) When both 1and2are large, the single particle energy dispersion of He is shown as Fig. 1(c), which displays a single energy minimum at nite kx. In this regime the low energy physics is dominated by a single dressed state described by1 2m(kxAx)2, whereAxis a constant. This leads to a uniform vector gauge eld. In a follow up experiment, Spielman's group applied a Zeeman eld gradient along ^ydirection to this system7. In this case, Axbecomes a function of yinsteadNovember 1, 2011 0:15 WSPC/INSTRUCTION FILE Review_SOC Spin-Orbit Coupled Quantum Gases 3 /Minus3/Minus2/Minus1123k/Slash1k00.51.01.5E-101(b)BEC (a)(c)(d)/Minus3/Minus2/Minus1123k/Slash1k0/Minus224E Fig. 1. (a) A schematic of NIST experiment, in which two counter propagating Raman beams are applied to87Rb BEC. (b) A schematic of how three F= 1 levels are coupled by Raman beams. (c) Dispersion in the regime of uniform vector potential. (d) Dispersion in the regime of non-abelian gauge eld. of a constant. It gives rise to a non-zero synthetic magnetic eld Bz=@yAx6= 0. They observed a critical magnetic eld above which many vortices are generated in the BEC7. In another experiment8, they made Axtime dependent which gives rise to a non-zero electric led Ex=@tAx6= 0. They observed collective oscillation of BEC after a pulse of electric eld8. By tuning the Zeeman energy and the laser frequency, one can also reach the regime where  1+!2, and thus20, while122is still large. In 2011, Spielman's group rst reached this regime and showed that a SO coupling can be generated9. As shown in Fig. 1(d), in this regime the low-energy physics contains two energy minima which are dominated by j1;1iandj1;0istates, respectively. Therefore we can deduce the low-energy e ective Hamiltonian by keeping both j1;1iandj1;0i, and rewrite the Hamiltonian as H= k2 x 2m+h 2 2ei2k0x 2ei2k0xk2 x 2mh 2! (4) whereh=2. Similarly, by applying a unitary transformation to the wave function with U=eik0x0 0eik0x (5) one reaches an e ective Hamiltonian that describes SO coupling HSO=UHUy= (kx+k0)2 2m+h 2 2 2(kxk0)2 2mh 2! =1 2m(kx+k0z)2+ 2x+h 2z: (6)November 1, 2011 0:15 WSPC/INSTRUCTION FILE Review_SOC 4Hui Zhai In fact, upon a pseudo-spin rotation x!zandz!x, the above Hamiltonian is equivalent to HSO=1 2m(kx+k0x)2 2z+h 2x; (7) where the rst term can be viewed as an equal weight of Rashba ( kxx+kyy) and Dresselhaus ( kxxkyy) SO couplinga. The Hamiltonian of Eq. 7 can also be viewed as a Hamiltonian with synthetic non-abelian gauge eld, since the vector potentialAx=k0zdoes not commute with the scale potential  = 2x+h 2z. Later on, Campbell et al. discussed how to generalize the NIST scheme to create SO coupling in both ^ xand ^ydirection, and nearly isotropic Rashba SO coupling10. Xu and You recently introduce a dynamics generalization of the NIST scheme11. Sau et al. discussed an explicit method to create Rashba SO coupling in fermionic 40K system at nite magnetic eld, where a magnetic Feshbach resonance is avail- able12. Beside the NIST scheme, there are also other theoretical proposals that use -type or tripod system to generate synthetic magnetic eld13;14;15;16, non-abelian gauge eld with a monopole17and SO coupling18;19;20;21. However, for those pro- posals using dark states13;14;15;16;20;21, collisional stability is a concern since there is always one eigen-state whose energy is below the dark state manifold, and multi- particle collision can lead to decay out of the dark state manifold. A recent review paper by Dalibard et al. has discussed di erent proposals in detail22. 2. Single Particle Properties with Rashba Spin-Orbit Coupling In the rest part of this review we will consider SO coupling in both ^ xand ^ydirections, whose Hamiltonian is given by H0=1 2m (kxxx)2+ (kyyy)2+k2 z (8) and in particular, we will consider the most symmetric Rashba case x=y= >0, where the physics is most interesting. In this case, the Hamiltonian can be rewritten as H0=1 2m k2 ?2k?+2+k2 z (9) Obviously, spin is no longer a good quantum number. However \helicity" is a good quantum number. \Helicity" means that the spin direction is either parallel or anti-parallel to the in-plane momentum direction. For these two helicity branches, their dispersion are given by k=1 2m(k2 ?2k?+2+k2 z) (10) aIn many literature, Rashba SO coupling denotes kxykyx, while Dresselhaus SO coupling denoteskxy+kyx. They are equivalent to the notations used in this paper by a pseudo-spin rotationx!yandy!x.November 1, 2011 0:15 WSPC/INSTRUCTION FILE Review_SOC Spin-Orbit Coupled Quantum Gases 5 (a)(b)/Minus2/Minus10120.00.51.01.52.0 k/UpTee/Slash1ΚE/Slash1ER012340.000.050.100.150.20E/Slash1ERDOS/Slash1/LParen1Κm/RParen1 Fig. 2. Energy dispersion with kz= 0 (a) and density-of-state (b) for Rashba spin-orbit coupled single particle Hamiltonian. In (a), \helicity" is \ + " for red solid line and is \ " for blue dashed line.ER=2=(2m) is introduced as energy unit. where k?= (kx;ky) andk?=q k2x+k2y. This Hamiltonian displays a symmetry of simultaneously rotation of spin and momentum along ^ zdirection. As shown in Fig. 2(a), helicity plus branch has lower energy. The single par- ticle energy minimum has nite in-plane momentum k?=, and all the single particle states with same k?=andkz= 0 but di erent azimuthal angle are degenerate ground states. The single particle DOS also has non-trival feature. In a three-dimensional system without SO coupling, DOS vanishes aspat low ener- gies. However, in this system, as shown in Fig. 2(b), the low-energy DOS becomes a constant when <E R, similar to a conventional two-dimensional system. Without SO coupling, the single particle Hamiltonian has a unique ground state atk= 0. At zero-temperature, bosons are all condensed at k= 0 state. If the inter- action is weak, interaction e ect is perturbative which only creates nite quantum depletion to the zero-momentum condensate. Also, because the vanishing DOS at low energies, two-body bound state appears only when the attractive interaction is above a threshold. Far below the threshold, when the attractive interaction is weak, the strength of fermion pairing and the fermion super uid transition temperature are both exponentially small. That is to say, without SO coupling, not surprisingly, the interaction e ect is weak in the regime where the interaction strength is small. With SO coupling, the e ects of interaction are signi cantly enhanced even when its strength is still weak, because SO coupling signi cantly changes the single par- ticle behaviors as discussed above. First, because the single particle ground state is not unique now, the ground state of a boson condensate is also not unique if there is no interactions. That is to say, it is the interaction that selects out a unique ground state among many possibilities. In this sense, the e ect of interaction is non-perturbative even for very weak interactions. Secondly, because the DOS is a constant at low energies, a two-body bound state appears for any weak attractions. Therefore, for weak attractive interactions, both pairing gap and super uid transi- tion temperature are largely enhanced by SO coupling. In particular, the super uid transition temperature can reach the same order as the Fermi temperature even forNovember 1, 2011 0:15 WSPC/INSTRUCTION FILE Review_SOC 6Hui Zhai weak attractions. These points will be discussed in the next two sections in more detail. On the other hand, we shall always keep in mind that in real experiment, one can never achieve a perfect isotropic Rashba SO coupling. Therefore, in our following discussions, we always rst obtain interesting results in the isotropic Rashba limit, and then discuss how robust the results are if there is a little anisotropy, namely, a slight di erence between xandy. 3. Spin-Orbit Coupled Bose Gases We will rst consider (pseudo-)spin-1 =2 bosons. We shall rst consider the mean- eld ground stateband then discuss the e ects of quantum and thermal uctuations on top of mean- eld saddle points. Let us rst consider the most simpli ed form of interactions ^Hint=Z d3r gn2 1(r) +gn2 2(r) + 2g12n1(r)n2(r) (11) By minimizing Gross-Pitaevskii energy functional, Wang et al. found that the mean- eld ground state has two di erent phases depending on the sign of g12g23. When g12<g, the system is in the \plane wave phase", where all bosons are condensed into a single plane wave state, and the direction of plane wave is spontaneously chosen in thexyplane. For instance, if the plane wave momentum is along ^ xdirection, the condensate wave function is given by =r 2eix1 1 : (12) The density of each component is uniform, but their phases modulate from zero to 2periodically, as found from numerical solution of Gross-Pitaevskii equation and shown in Fig. 3(a). This state spontaneously breaks time-reversal, rotational symmetry and the U(1) symmetry of super uid phase. When g12>g, all bosons are condensed into a superposition of two plane wave states with opposite momentums, whose condensate wave function is given by =p 2 eix1 1 +eix1 1 =pcos(x) isin(x) (13) In this phase, the spin density n1n2=cos(2x) which has a periodic modulation in space, as shown in Fig. 3(b), and therefore is named as \stripe super uid". The direction of the stripe is also spontaneously chosen in the xyplane. Here, without loss of generality, we choose it along ^ xdirection. In this state, the high density regime of one component coincides with the low-density regime of the other component, so that the inter-component repulsive interaction is maximumly avoided. That is the bThere are also proposals of non-mean- eld fragmented state, like N00N state, as ground state of SO coupled bosons24, however, the conventional wisdom is that such a state is very fragile when external perturbations are present.November 1, 2011 0:15 WSPC/INSTRUCTION FILE Review_SOC Spin-Orbit Coupled Quantum Gases 7 (b) (a) Fig. 3. (a) Phase of condensate wave function in the \plane wave phase"; (b) Spin density in the \stripe super uid" phase. reason why the spin stripe state is favored when g12is larger than g. In addition to theU(1) super uid phase, this state also breaks the rotational symmetry (but keeps the re ection symmetry), and the translational symmetry along the stripe direction. Hence, symmetry wise, it can also been called \smectic super uid". In contrast to the \plane wave phase", this state does not break time-reversal symmetry. In principle, the condensate wave function can be any superposition of single particle ground states as =X 'c'ei(cos'x+sin'y)1 ei' (14) One may wonder why only a single state or a superposition of a pair of states are favored by interactions. In fact, if there are more than a pair of single particle states in the superposition, the condensate wave function will exhibit interesting structure, such as various types of skyrmion lattices. And if all the degenerate states enter the condensate wave function with equal weight and speci ed relative phases, condensate will exhibit interesting structure of half vortices, as rst proposed by Stanescu, Anderson, Galitski24and Wu, Mondragon-Shem25. However, such a state is not energetically favorable in spin-1 =2 case for a uniform (or nearly uniform) system. This is because that the interaction part can be rewritten as ^Hint=Z d3rg+g12 2n2(r) +gg12 2S2 z(r) ; (15) wheren(r) =n1(r) +n2(r) andSz=n1(r)n2(r). TheS2 z-term can be satis ed by either choosing the \plane wave phase" or the \stripe phase", while the n2- term with positive coecient always favors a uniform density. One can easily show that, if there are more than a pair of states in the superposition, the condensate density will always have non-uniform modulation, which causes energy of n2-term. Similar situation has also been found for spin-1 Hamiltonian26;23. However, there are several situations where skyrmion lattices or half vortices are found as ground state. When a strong harmonic con nement potential V(r) =m!r2 ?=2 is applied to the system, condensate density is no longer uniform because of the trapping potential and the requirement from n2-term becomes less restrictive. In addition,November 1, 2011 0:15 WSPC/INSTRUCTION FILE Review_SOC 8Hui Zhai one notes in the limit of zero interaction, the ground state in a harmonic trap can be solved exactly and it is a half-vortex state31;32;54. Recently, three groups have found that if one largely increases the con nement potential so that a=p ~=m! is comparable to 1 =, or reduces the interaction energy to be comparable to ~!, the ground state will evolve continuously to skyrmion lattice phases, and nally to half vortex phases31;32;54. Besides, Xu et al. and Kawakami et al. found in spin-2 case, because of an addition interaction term (which favors cyclic phase in absence of SO coupling29;30), there exists a parameter regime where various types of skyrmion lattice phases are ground state even for a uniform system. Back to the discussion of a nearly uniform spin-1 =2 systemc, the \plane wave phase" and the \stripe phase" are in fact two very robust mean- eld states. In real situation, the interactions between two pseudo-spin states have much more complicated form than the simpli ed form of Eq. (11). Both Yip and Zhang et al. considered a speci c type of complicated interaction form using a concrete real- ization of Rashba SO coupling, and they found the ground state is still either the \plane wave phase" or the \stripe phase"34;80. And also, the SO coupling is always not perfectly isotropic, say, x>y, then the Hamiltonian itself does not have rota- tional symmetry anymore. The single particle energy has two minima at ( ;0;0) instead of a continuous circle of degenerate states. At mean- eld level, the e ect of anisotropic SO coupling is to pin the direction of plane wave or stripe into certain direction (^xdirection for this case). The NIST experimental situation discussed in Sec. 1 corresponds to the case y= 0 and also with a Zeeman eld. As shown by Ho and Zhang36, and also in the experimental paper9, the phase diagram of this system also only contains such two phases. To go beyond the mean- eld description, three di erent approaches have been tried so far. The rst is the e ective eld theory approach37, which treats Gaussian uctuations of low-energy modes. The second is Bogoliubov approach41;42;43, which more focuses on the gapless phonon excitations. And the third is the renormalization approach38;39;40, which discusses how scattering vertices are renormalized by high order processes. These di erent approaches address beyond-mean- eld e ects from di erent perspectives, and the results are consistent with each other where they overlap. Taking e ective eld theory approach as an example, for the \stripe" phase, the super uid phase and the relative phase ubetween two momentum components are two low-lying modes: 'ST=p 2ei ei(x+u)1 1 +ei(x+u)1 1 : (16) In fact, the relative phase udescribes the phonon mode of stripe order. The dy- cFor typical experimental parameters, the BEC is in the nearly uniform regime rather than strong harmonic con nement regime.November 1, 2011 0:15 WSPC/INSTRUCTION FILE Review_SOC Spin-Orbit Coupled Quantum Gases 9 namics ofandu elds are governed by an e ective energy function37 HST e = 2m" (@x)2+(@y)2 2+ (@xu)2+ @2 yu2 42# ; (17) where >1 is a constant. For the \plane wave" phase, the superfuid phase is the only low energy mode 'PW=pei(x+)1 1 (18) and its e ective energy is derived as37 HPW e = 2m (@x)2+1 42(@2 y)2 : (19) One notes that in the \stripe" phase, the quadratic term ( @yu)2is absent in Eq. (17), and in the \plane wave" phase, the quadratic term ( @y)2is absent in Eq. (19). This is in fact a manifestation of rotational symmetry in this system. Similar e ective theory has also been found for \FFLO" state in fermion super uid44;46. Such an energy function is also the classical energy of smectic liquid crystal. In two dimensions, the nite temperature phase transition is driven by prolif- eration of topological defects. For usual XYmodel, both the energy of topological vortex and its entropy logarithmically depend on system size. Thus, only above a critical temperature, entropy wins energy and the topological defects proliferate which drives system into normal phase. However, in the \stripe" phase, because the absence of ( @yu)2term, the energy for a topological defect of u, i.e. a dislocation in the stripe, no longer logarithmically depends on system size. Hence it will lose to en- tropy at any nite temperature, and the proliferation of these dislocations will melt the stripe order. This restores translational symmetry and drives the system into a boson paired super uid phase37. Such a boson pairing state can also be predicted by looking at pairing instability of normal state with renormalized interactions38. By considering the renormalization e ects of scattering vertex from high energy states, Gopalakrishnan et al.38and Ozawa, Baym39;40found that the scattering amplitudes between two states with opposite momentum become smaller and even vanishing, which makes the \stripe phase" more favorable at the low-density limit, and meanwhile leads to more signi cant the uctuation of stripe order38. For same reason, in the \plane wave" phase, the super uid phase will immediately disorder at nite temperature and the system becomes normal. In addition, the e ective theory Eq.(17) and Eq. (19) also imply that there is a Goldstone mode which has linear dispersion along ^ xdirection (direction of stripe or plane wave momentum), and quadratic dispersion along ^ ydirection (direction perpendicular to the direction of stripe or plane wave momentum). Same results have been reached by Bogoliubov calculation41;42;43. Similar behaviors of Goldstone modes also exist in \FFLO" phase of fermion super uid44;45;46.November 1, 2011 0:15 WSPC/INSTRUCTION FILE Review_SOC 10Hui Zhai When the SO coupling is anisotropic ( x6=y), there is no rotational symmetry in the Hamiltonian. Therefore, one will have ( @yu)2in the stripe phase, and ( @y)2 in the plane wave phase. However, the coecients of those terms are propositional to 1(y=x)2. In the regime y=xis very close to unity, the energy of a disloca- tion is much smaller than the energy of a vortex or a half vortex. Hence, the system will undergo two Kosterlitz-Thouless phase transitions. At a lower critical tempera- ture, dislocations proliferate and the system becomes paired super uid. Then, at a higher critical temperature, vortices proliferate and the system becomes normal. A completed phase diagram in term of interaction parameter, SO coupling anisotropy and temperature is given by Jian and Zhai37. For SO coupled bosons, many questions remain open. Recently several works begin to address the questions about the super uid critical velocity42, vortices in presence of external rotation47;48;49, the e ects of dipolar interactions50, the collec- tive modes51, super uid to Mott insulator transition in a lattice52, the interplay between magnetic eld and SO coupling53, and the dynamical e ects nearby Dirac point due to SO coupling54;55;56. The research along this direction will de nitely reveal more interesting physics and stimulate more interesting experiments. 4. Spin-Orbit Coupled Fermi Gases across a Feshbach Resonance Even for a non-interacting system, the thermodynamic behavior of a Fermi gas is dramatically changed by a strong SO coupling because of the change of the low- energy DOS57. In this section, we focus on Fermi gas with attractive interaction, and in particular, across a Feshbach resonance. The interaction part is modeled as Hint=gZ d3r y "(r) y #(r) #(r) "(r) (20) where 1 g=m 4~2asX k1 ~2k2=m: (21) Here we relate the bare interaction gtos-wave scattering length via Eq. (21), which is the same as the scheme widely used in free space without SO coupling. This is based on the assumption that the interaction Hamiltonian is not changed by SO coupling, and the ashere should be understood as scattering length in free space.d Using this interaction Hamiltonian, Vyasanakere and Shenoy rst studied the two-body problem across a Feshbach resonance59. Because the low-energy DOS is now a constant, an arbitrary weak attractive interaction will give rise to a bound state. Similar situations are two-body problem in two-dimension and Cooper prob- lem in three-dimension in absence of SO coupling, where DOS are also constants. dThis is equivalent to assume that the s-wave pseudo-potential is still a valid approximation for a short-range realistic potential in presence of SO coupling. In fact, the validity of such an approximation is not quite obvious, and it has only been examined recently by Cui58.November 1, 2011 0:15 WSPC/INSTRUCTION FILE Review_SOC Spin-Orbit Coupled Quantum Gases 11 Fig. 4. (a) Gap size  as a function of =kFfor di erent values of 1 =(kFas). (b) Size of Cooper pair inxyplanel?and along ^zdirectionlzas a function of =kFat resonance with as=1; (c) Super uid transition temperature Tc=TFas a function of 1 =(kFas) for di erent values of =kF. Reprinted from arXiv: 1105.2250 (Phys. Rev. Lett. to be published). The binding energy can be easily calculated by looking at poles of T-matrix59;61 or by reducing the two-body Schr odinger equation to a self-consistent equation62. The two-body properties at the BCS side and at resonance regime are signi cantly changed by SO coupling. At weakly interacting BCS side with small negative as, the binding energy behaves as Eb~22 2m4 e2e2 jasj: (22) where a large binding energy can always been reached by increasing the strength of SO coupling. At resonance with as=1, 1=is the only length scale in the two-body problem and therefore one has a universal result Eb=0:88~22 2m: (23) While for the BEC side with small positive as, at leading order Ebis still given by ~2=(2ma2 s) and is not a ected by SO coupling. Moreover, because of SO coupling, the two-body wave function has both singlet and triplet components. For a two- body bound state with zero center-of-mass momentum, the wave function behaves as59 = s(r)j"##"i +  a(r)j""i + a(r)j##i (24) where s(r) and a(r) are symmetric and anti-symmetric functions, respectively. Furthermore, by looking at binding energy with nite center-of-mass momentum, one can determine the e ective mass of molecules (two-body bound state). Hu et al. 61and Yu and Zhai62found that at the BCS limit, the e ective mass of molecule nally saturates to 4 m, and at resonance, the e ective mass is a universal number of 2:40m. In the BEC limit, the e ective mass saturates to 2 mas conventional case. Generalizing conventional BEC-BCS crossover mean- eld theory to the case with SO coupling and equal population hn"i=hn#i, one can show that the system remains gapped for all kFas, although there are triplet p-wave components60;61;62;63,November 1, 2011 0:15 WSPC/INSTRUCTION FILE Review_SOC 12Hui Zhai and the pair wave function obtained from the mean- eld theory nally approaches the wave function of two-body bound state (i.e. molecule wave function)60;62. Hence, it is still a crossover as 1 =(kFas) changes from negative to positive. However, on the other hand, the change of low-energy DOS and the presence of two-body bound state at the BCS side and resonance regime will signi cantly change the properties of crossover60;61;62;64. For example, as shown in Fig. 4(a), the pairing gap at the BCS side is dramatically enhanced when =kFis comparable or larger than unitye. For such a strong SO coupling, the DOS at Fermi energy becomes a constant, and is much larger than the DOS in absence of SO coupling. That is the reason why the pairing e ects become much dramatic even for same interaction strength. For another example, in absence of SO coupling, Fermi energy is the only energy scale at resonance, and therefore the size of Cooper pair is a universal constant times 1 =kF. SO coupling introduces another scale at resonance, which is 1 =. For large , as the pairing gap approaches two-body binding energy, the size of Cooper pairs also approaches 1 =. Fig. 4(b) shows that the behavior of lundergoes a crossover from 1=kFto 1=as=kFincreases. This plot also shows that the size of Cooper pair in thexyplane is di erent from the size along ^ zdirection, namely, the Cooper pairs are anisotropic.61;62. In addition, one can also show that the super uid transition temperature at the BCS side can also be enhanced a lot by SO coupling. For large enough SO coupling, it eventually approaches the BEC temperature of molecules with mass 4 m62;64, which is a sizable fraction of Fermi temperature. The critical temperature across resonance is rst estimated by Yu and Zhai62as shown in Fig. 4(c). If SO coupling is slightly anisotropic, DOS at very low-energy will nally vanish. However, there is still a large energy window .~22=(2m) where the DOS is greatly enhanced by SO coupling. Hence, pairing gap will still be enhanced as long as the density of fermions is not extremely low. Moreover, although it is no longer true that arbitrary small attraction can cause a bound state, the critical value for the appearance of a bound state will move from unitary point as=1to the BCS side with negative as59. Once the bound state is present, it will in uence the universal behavior of pair size at resonance and super uid critical temperature as discussed above. For the imbalanced case, the phase diagram becomes more richer. Several groups have studied the phase diagram in presence of a Zeeman eld65;66;67;68;69and in various other circumstance70;71;72;73;74;75;76;77;78;79;80 f. They have shown that, eIn cold atom system, SO coupling is generated by atom-light coupling, and therefore is on the order of the inverse of the laser wave length. And since the laser wave length and the inter-particle distance are comparable (between 0:1mand1m) in atomic gases, the strength of SO coupling in cold atom systems can naturally reach the regime =kF1. fIskin and Subasi studied SO coupled Fermi gas with mass imbalance72. They use mixture of di erent species as motivation of this study. We caution that the current way of generating SO coupling is based on light coupling of di erent internal state of atoms, which can not generate SO coupling if di erent internal states are di erent atomic speciesNovember 1, 2011 0:15 WSPC/INSTRUCTION FILE Review_SOC Spin-Orbit Coupled Quantum Gases 13 instead of a crossover, there are phase transitions between topological and non- topological phases65;66;67;68. They have also discussed how SO coupling in uences the competition between a uniform super uid and a phase separation66;67. 5. Summary and Future Developments SO coupled quantum gases with interactions are new systems in cold atom physics. Moreover, SO coupled bosonic system has never been thought in physics before. Currently our understanding of this system is still very limited, and many questions remain open. However, even from our limited experience with this new system, one can already get a feeling that this system has many unusual behaviors. This gives a lot of opportunities for theorists and experimentalists. Acknowledgements I would like to thank Chao-Ming Jian and Zeng-Qing Yu for collaboration on this subject, and thank Xiaoling Cui for helpful discussions. This work is supported by Tsinghua University Initiative Scienti c Research Program, NSFC under Grant No. 11004118 and No. 11174176, NKBRSFC under Grant No. 2011CB921500. References 1. M. V. Berry, Proc. Roy. Soc. London A 392, 45 (1984) 2. F. Wilczek and A. Zee, Phys. Rev. Lett. 52, 2111 (1984) 3. X. L. Qi and S. C. Zhang, Physics Today, 63, 33 (2010); 4. M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010) 5. X. L. Qi, S. C. Zhang, arXiv: 1008.2026, to be published in Rev. Mod. Phys 6. Y. J. Lin, R. L. Compton, A. R. Perry, W. D. Phillips, J. V. Porto, and I. B. Spielman, Phys. Rev. Lett. 102, 130401 (2009) 7. Y. J. Lin, R. L. Compton, K. Jimnez-Garca,, J. V. Porto, I. B. Spielman, Nature 462, 628 (2009) 8. Y. J Lin, R. L. Compton, K. Jimnez-Garca, W. D. Phillips, J. V. Porto, I. B. Spielman, Nature Physics, 7, 531, (2011) 9. Y.-J. Lin, K. Jimnez-Garca, I. B. Spielman, Nature, 471, 83 (2011) 10. D. L. Campbell, G. Juzeli unas, I. B. Spielman, arXiv: 1102.3945 11. Z. F. Xu and L. You, arXiv:1110.5705 12. J. D. Sau, R. Sensarma, S. Powell, I. B. Spielman, S. Das Sarma, Phys. Rev. B 83, 140510(R) (2011) 13. R. Dum and M. Olshanii, Phys. Rev. Lett. 76, 1788 (1996). 14. G. Juzeli unas and P. Ohberg, Phys. Rev. Lett. 93, 033602 (2004); 15. G. Juzeli unas, J. Ruseckas, P. Ohberg, and M. Fleischhauer, Phys. Rev. A 73, 025602 (2006). 16. K. J. G unter, M. Cheneau, T. Yefsah, S. P. Rath, and J. Dalibard, Phys. Rev. A 79, 011604(R) (2009) 17. J. Ruseckas, G. Juzeli unas, P. Ohberg, and M. Fleischhauer, Phys. Rev. Lett. 95, 010404 (2005); 18. S. L. Zhu, H. Fu, C. J. Wu, S. C. Zhang, and L. M. Duan, Phys. Rev. Lett. 97, 240401 (2006);November 1, 2011 0:15 WSPC/INSTRUCTION FILE Review_SOC 14Hui Zhai 19. X. J. Liu, X. Liu, L. C. Kwek, and C. H. Oh, Phys. Rev. Lett. 98, 026602 (2007); 20. T. D. Stanescu, C. Zhang, and V. Galitski, Phys. Rev. Lett. 99, 110403 (2007). 21. G. Juzeliunas, J. Ruseckas, J. Dalibard, Phys. Rev. A 81, 053403 (2010) 22. J. Dalibard, F. Gerbier, G. Juzeli unas, P. Ohberg, arXiv: 1008.5378, to be published in Rev. Mod. Phys. 23. C. Wang, C. Gao, C. M. Jian, and H. Zhai, Phys. Rev. Lett. 105, 160403 (2010) 24. T. D. Stanescu, B. Anderson, and V. Galitski, Phys. Rev. A 78, 023616 (2008) 25. C. J. Wu, I. Mondragon-Shem, arXiv:0809.3532v1 26. T. L. Ho, Phys. Rev. Lett. 81, 742 (1998) 27. Z. F. Xu, R. L u, and L. You, Phys. Rev. A 83, 053602 (2011) 28. T. Kawakami, T. Mizushima, and K. Machida, Phys. Rev. A 84, 011607 (2011) 29. C. V. Ciobanu, S. K. Yip, and T. L. Ho, Phys. Rev. A 61, 033607 (2000) 30. M. Koashi and M. Ueda, Phys. Rev. Lett. 84, 1066 (2000) 31. C. J. Wu, I. Mondragon-Shem, X. F. Zhou, Chin. Phys. Lett. 28, 097102 (2011) 32. H. Hu, H. Pu, X. J. Liu, arXiv:1108.4233 33. S. Sinha, R. Nath, L. Santos, arXiv: 1109.2045 34. S. K. Yip, Phys. Rev. A 83, 043616 (2011) 35. Y. Zhang, L. Mao, C. Zhang, arXiv: 1102.4045 36. T. L. Ho, S. Zhang, Phys. Rev. Lett. 107, 150403 (2011) 37. C. M. Jian and H. Zhai, Phys. Rev. B 84, 060508 (2011) 38. S. Gopalakrishnan, A. Lamacraft, P. M. Goldbart, arXiv: 1106.2552 39. T. Ozawa, G. Baym, arXiv:1107.3162 40. T. Ozawa, G. Baym, arXiv:1109.4954 41. Q. Zhou and X. L. Cui, to be published 42. Q. Zhu, C. Zhang and B. Wu, arXiv: 1109.5811 43. R. Barnett, S. Powell, T. Gra , M. Lewenstein and S. Das Sarma, arXiv:1109.4945 44. L. Radzihovsky, A. Vishwanath, Phys. Rev. Lett. 103, 010404 (2009), 45. L. Radzihovsky and S. Choi, Phys. Rev. Lett. 103, 095302 (2009) 46. Radzihovsky, arXiv: 1102.4903 47. X. Q. Xu, J. H. Han, arXiv: 1107.4845 48. X. F. Zhou, J. Zhou, C. J. Wu, arXiv: 1108.1238 49. J. Radic, T. Sedrakyan, I. Spielman, V. Galitski, arXiv: 1108.4212 50. Y. Deng, J. Cheng, H. Jing, C.-P. Sun, S. Yi, arXiv: 1110.0558 51. E. van der Bijl, R.A. Duine, arXiv: 1107.0578 52. T. Grass, K. Saha, K. Sengupta, M. Lewenstein, arXiv: 1108.2672 53. M. Burrello and A. Trombettoni, Phys. Rev. Lett. 105, 125304 (2010) 54. M. Merkl, A. Jacob, F. E. Zimmer, P. Ohberg, and L. Santos, Phys. Rev. Lett. 104, 073603 (2010) 55. M. J. Edmonds, J. Otterbach, R. G. Unanyan, M. Fleischhauer, M. Titov, P. Ohberg, arXiv: 1106.5925 56. D. W. Zhang, Z. Y. Xue, H. Yan, Z. D. Wang, S. L. Zhu, arXiv: 1104.0444 57. L. He and Z. Q. Yu, Phys. Rev. A 84, 025601 (2011) 58. X. L. Cui, to be published 59. J. P. Vyasanakere, V. B. Shenoy, Phys. Rev. B 83094515 (2011) 60. J. P. Vyasanakere, S. Zhang, V. B. Shenoy, Phys. Rev. B 84, 014512 (2011) 61. H. Hu, L. Jiang, X. J. Liu, H. Pu, arXiv: 1105.2488, to be published in Phys. Rev. Lett. 62. Z. Q. Yu and H. Zhai, arXiv: 1105.2250, to be published in Phys. Rev. Lett. 63. L. Han, C. A. R. S a de Melo, arXiv: 1106.3613; 64. J. P. Vyasanakere, V. B. Shenoy, arXiv: 1108.4872November 1, 2011 0:15 WSPC/INSTRUCTION FILE Review_SOC Spin-Orbit Coupled Quantum Gases 15 65. M. Gong, S. Tewari, C. Zhang, arXiv: 1105.1796, to be published in Phys. Rev. Lett. 66. M. Iskin and A. L. Subas, Phys. Rev. Lett. 107, 050402 (2011); 67. W. Yi, G. C. Guo, arXiv: 1106.5667; 68. K. Seo, L. Han, C. A. R. S a de Melo, arXiv:1108.4068; 69. R. Liao, Y. X. Yu and W. M. Liu, arXiv: 1110.5818 70. S. L. Zhu, L. B. Shao, Z. D. Wang, L. M. Duan, Phys. Rev. Lett. 106, 100404 (2011) 71. L. Jiang, T. Kitagawa, J. Alicea, A. R. Akhmerov, D. Pekker, G. Refael, J. Ignacio Cirac, E. Demler, M. D. Lukin, P. Zoller, Phys. Rev. Lett. 106, 220402 (2011) 72. M. Iskin, A. L. Subasi, arXiv:1107.2376 73. M. Iskin and A. L. Subas, arXiv: 1108.4263 74. G. Chen, M. Gong, C. Zhang, arXiv: 1107.2627; 75. L. Dell'Anna, G. Mazzarella and L. Salasnich, arXiv: 1108.1132 76. L. He and X. G. Huang, arXiv: 1109.5777 77. S. K. Ghosh, J. P. Vyasanakere and V. B. Shenoy, arXiv: 1109.5279 78. J. Zhou, W. Zhang and W. Yi, arXiv:1110.2285 79. J. N. Zhang. Y. H. Chan and L. M. Duan, arXiv: 1110.2241 80. K. Zhou and Z, Zhang, arXiv: 1110.3565
2010.01970v1.Detection_of_the_Orbital_Hall_Effect_by_the_Orbital_Spin_Conversion.pdf
Detection of the Orbital Hall E ect by the Orbital-Spin Conversion Jiewen Xiao,1Yizhou Liu,1and Binghai Yan1, 1Department of Condensed Matter Physics, Weizmann Institute of Science, Rehovot 7610001, Israel (Dated: October 6, 2020) The intrinsic orbital Hall e ect (OHE), the orbital counterpart of the spin Hall e ect, was pre- dicted and studied theoretically for more than one decade, yet to be observed in experiments. Here we propose a strategy to convert the orbital current in OHE to the spin current via the spin-orbit coupling from the contact. Furthermore, we nd that OHE can induce large nonreciprocal magne- toresistance when employing the magnetic contact. Both the generated spin current and the orbital Hall magnetoresistance can be applied to probe the OHE in experiments and design orbitronic devices. I. INTRODUCTION The intrinsic orbital Hall e ect (OHE), where an elec- tric eld induces a transverse orbital current, was pro- posed by the Zhang group [1] soon after the prediction of the intrinsic spin Hall e ect (SHE) [2, 3]. The SHE was soon observed [4, 5], later applied for the spintronic devices[6, and references therein], and also led to the sem- inal discovery of the quantum SHE, i.e., the 2D topolog- ical insulator [7, 8]. Di erent from the SHE, the OHE does not rely on the spin-orbit coupling (SOC), and thus, it was predicted to exist in many materials [1, 9{18] with either weak or strong SOC, for example, in metals Al, Cu, Au, and Pt. In an OHE device, the transverse orbital current leads to the orbital accumulation at transverse edges, similar to the spin accumulation in a SHE device. Zhang et al [1] proposed to measure the edge orbital accumulation by the Kerr e ect. Recently, Ref. 19 predicted the orbital torque generated by the orbital current. However, the OHE is yet to be detected in experiments until today. The detection of the orbital is rather challenging, because the orbital is highly non-conserved compared to the spin, especially at the device boundary. A very recent work by us proposed [20] that the lon- gitudinal current through DNA-type chiral materials is orbital-polarized, and contacting DNA to a large-SOC material can transform the orbital current into the spin current. Thus, we are inspired to conceive a similar way to detect the transverse OHE by converting the orbital to the spin by the SOC proximity. In this article, we propose two ways to probe the OHE, where the strong SOC from the contact transforms the orbitronic problem to the spintronic measurement. One way is to generate spin current or spin polarization from the transverse orbital current by connecting the edge to a third lead with the strong interfacial SOC. Then the edge spin polarization and spin current is promising to be mea- sured by the Kerr e ect [4] and the inverse SHE [21{23], respectively. The other way is to introduce a third mag- netic lead and measure the magnetoresistance. We call binghai.yan@weizmann.ac.il FIG. 1. Illustration of the orbital-spin conversion and the orbital Hall magnetoresistance (OHME). (a) The orbital Hall e ect and the spin polarization/current genera- tion. Opposite orbitals (red and blue circular arrows) from the left lead de ect into opposite boundaries. The red and blue backgrounds represent the orbital accumulation at two sides. Because of the SOC region (yellow) at one side, the orbital current is converted into the spin current (indicated by black arrows). (b) The two-terminal (2T) OHMR. The third lead is magnetized but open. (c) The three-terminal (3T) OHMR. The third lead is magnetized and conducts cur- rent. The 2T/3T conductance between di erent leads relies on the magnetization sensitively. The thickness of grey curves represent the relative magnitude of the conductance. it the orbital Hall magnetoresistance (OHMR), similar to the spin Hall magnetoresistance [24, 25]. In our pro- posal, the OHE refers to orbitals that resemble atomic- like orbitals, which naturally couple to the spin via the atomic SOC. We rst demonstrate detection principles in a lattice model by transport calculations. Then we incor- porate these principles into the metal copper, which has negligible SOC and avoids the co-existence of the SHE, as a typical example of realistic materials. In the copper based device, we demonstrate the resultant spin polariza- tion/current and very large OHMR (0 :31:3 %), which are measurable by present experiment techniques. II. RESULTS AND DISCUSSIONS A. Methods and General Scenario To detect the OHE, we introduce an extra contact with the strong SOC on the boundary of the OHE material,arXiv:2010.01970v1 [cond-mat.mtrl-sci] 5 Oct 20202 as shown in Figure 1. This device can act for both two- terminal (2T) and three-terminal (3T) measurements (or more terminals). In theoretical calculations, we com- pletely exclude SOC from all leads so that we can well de ne the spin current. We also remove SOC in the OHE material, the device regime in the center, to avoid the ex- istence of SHE. Only nite atomic SOC is placed in the interfacial region (highlighted by yellow in Figure 1) be- tween the OHE and the third lead. We rst prove the principle by a simple square-lattice model that hosts OHE. As shown in the inset of Figure 2(a), a tight-binding spinless model is constructed, withthree orbitals s,pxandpyassigned to each site. Under the above basis, the atomic orbital angular momentum operator ^Lzis written as ^Lz= h2 40 0 0 0 0i 0i03 5 (1) And three eigenstates p(pxipy)=p 2;scorrespond to eigenvalues Lz=1;0, respectively. After consider- ing the nearest neighboring hopping, the Hamiltonian is written as H(kx;ky) =0 @Es+ 2tscoskxa+ 2tscoskya2itspsinkxa 2itspsinkya 2itspsinkxa E px+ 2tpcoskxa+ 2tpcoskya 0 2itspsinkya 0 Epy+ 2tpcoskxa+ 2tpcoskya1 A (2) whereEs,EpxandEpyare onsite energies of s,pxandpy orbitals.ts,tp,tp,tspare electron hopping integrals betweensorbitals,type oriented porbitals,type orientedporbitals, and sandporbitals, respectively. In the following calculations, their values are speci ed as Es= 1:3,Epx=Epy=1:9,ts=0:3,tp= 0:6, tp= 0:3, andtsp= 0:5, in the unit of eV. To realize the OHE, it requires the inter-orbital hopping to induce the transverse Lzcurrent. Since pxandpyorbital are orthogonal under the square lattice geometry, the inter- orbital hopping tspbecomes the critical parameter that controls the existence of the OHE. Then we introduce the atomic SOC on the boundary to demonstrate the OHE detection by soc^Sz^Lz, where ^Szis the spin operator. We estimate the OHE conductivity ( OH) with the orbital Berry curvature in the Kubo formula [26, 27], OH=e hX nZd3k (2)3fnk Lz n(k) (3) Lz n(k) = 2h2X m6=nIm[hunkjjLzyjumkihumkj^vxjunki (EnkEmk)2] (4) where Lzn(k) is the \orbital" Berry curvature for the nthband with Bloch state junkiand energy eigenvalue Enk.fnkis the Fermi-Dirac distribution function. vx is thexcomponent of the band velocity operator while jLzyis the orbital current operator in the ydirection, de- ned asjLzy= (^Lz^vy+ ^vy^Lz)=2. Therefore, the above formula indicates that the interband perturbation in- duces the orbital Berry curvature, further reiterating the importance of inter-orbital hopping. We also note that, the orbital Berry curvature is even under the time- reversal symmetry or the spatial inversion symmetry, Lzn(k) = Lzn(k). For the device schematically presented in Figure 1, wecalculated the conductance by the Landauer-B uttiker for- mula [28] with the scattering matrix from lead ito lead j, Gi!j=e2 hX n2j;m2ijSnmj2; (5) whereSmnis the scattering matrix element from the m- th eigenstate in lead ito thentheigenstate in lead j. In all three leads ( i;j= 1;2;3), spin (Sz="#) is a con- served quantity because of the lack of SOC. We turn o the inter-orbital hopping in leads so that Lzis also conserved, i.e., Lzcommutes with the Hamiltonian (See Supplementary Materials). Therefore, with the spin and orbital conserved leads, we can specify the conductance in eachSzandLzchannel, and de ne the orbital- and spin-polarized conductance as: Gij Sz=Gi!j"Gj!j#(6) Gij Lz=Gi!j+Gi!j; (7) whereGij Sz(Lz)is the conductance from lead ito the Sz(Lz) channel of lead j.Gi!j0is omitted here since Lz= 0 contributes no polarization. We performed the conductance calculations with the quantum transport package Kwant [29]. As illustrated in Figure 1(a), electrons with the op- posite orbital angular momentum de ect into transverse directions in the OHE region, resulting in the transverse orbital current. Therefore, orbital accumulates at two sides, and the orbital polarization emerges. To detect the orbital polarization, atomic SOC is added at one side, as highlighted by yellow in Figure 1(a). After electron de- ecting into the SOC region, the right-handed orbital (red circular arrows) is converted to the up spin polar- ization. If a third lead is further attached, the SOC re- gion converts the orbital current into the spin current.3 If the third lead exhibits magnetization along z(Mz) (Figure 1(b) and 1(c)), inversely, the OHE induces the OHMR, relying on whether Mzis parallel or anti-parallel to the generated spin polarization. In the 2T measure- ment (Figure 1(b)), the conductance from lead 1 to lead 2 (G1!2) changes when the Mzdirection is reversed. And the changing direction of G1!2depends sensitively on the size of the device, due to the complex orbital accu- mulation and re ection with an open lead. While for its spin counterpart, the SHE-induced magnetoresistance is commonly measured in a 2T setup [24, 25]. In the 3T device (Figure 1(c)), the situation is simpler since the transverse orbital current can ow into the third lead. IfMzand spin polarization is parallel (anti-parallel), the transverse orbital current matches (mismatches) the lead magnetization, resulting in the high (low) G1!3and low (high)G1!2accordingly. We point out that the 3T mea- surement is usually more favorable than 2T, since the 3T device avoids the 2T reciprocity constrain [30] and the conductance change [ G=G(Mz)G(Mz)] is also relatively larger in the third lead, as discussed in the fol- lowing. B. Spin Polarization and Spin Current Generated by the OHE The band structure weighted by the orbital Berry cur- vature for the square lattice is plotted in Figure 2(a). The highest band corresponds to the sorbital dispersion, while two lower bands are dominated by porbitals. The orbital Berry Curvature concentrates near the point, Mpoint and - Mline in the Brillouin zone, where band hybridization is strong. After integrating Lzin the Bril- louin zone, the orbital Hall conductivity is derived and presented in Figure 2(a). It shows that, due to the inter- orbital hopping tsp, states below ( porbitals) and above ( s orbital) Fermi level both exhibits signi cant OH. How- ever, iftspis turned o so that Lzis conserved, both Lz andOHvanishes. Based on the square lattice with nite tsp, the 2T de- vice is constructed, as shown in Figure 2(b). Without SOC at two sides, the orbital density distribution is plot- ted in Figure 2(c), which shows that opposite orbitals accumulate and polarize at two boundaries. With SOC turned on, spin density appears and largely concentrates on the local SOC atoms, which is promising to be de- tected by the Kerr e ect [4]. Since SOC couples the p+ (p) orbital to the"(#) spin and forms the jjm=3 2i (jjm=3 2i) state, the spin density near the SOC region largely follows the orbital density pattern: positive at the upper side and negative at the lower side. To ver- ify that the spin polarization is directly induced by the OHE rather than SOC, we turned o the OHE by setting tsp= 0 eV and preserve the SOC at the interface. The supplementary Figure S2 shows that both the orbital and spin polarization disappear. On the basis of 2T device, a third lead is attached to FIG. 2. Orbital-spin conversion in the two terminal (2T) and three terminal (3T) device. (a) Band struc- ture of the square lattice with tsp= 0:5 eV (left) and the orbital Hall conductivity with tsp= 0:5 eV andtsp= 0:0 eV (right). In the inset, the tight binding model of the square lattice is presented. (b) 2T and 3T detection devices, where larger spheres at two sides represent SOC regions. The yel- low spheres at left, right and upper sides represent leads. (c) Orbital and spin density distribution in the 2T setup, at the energy level of 0.2 eV. (d) Total, orbital and spin conductance from lead 1 to lead 3 with (left) and without (right) SOC. the SOC side to form a 3T device, as shown in Figure 2(b). Therefore, rather than the orbital accumulation, the orbital current will ow into the third lead and gen- erate the spin current. Figure 2(d) shows that the orbital current from lead 1 to lead 3 ( G13 Lz) exists with and with- out SOC at the interface. For instance, for states above Fermi level, Lz= +1 states are more easily transported into lead 3 than Lz=1 states, and thus polarizes the lead, being consistent with the positive orbital polariza- tion at the upper side in Figure 2(c). On the other hand, for the spin conductance, it only appears when turning on SOC, and the energy dependence of G13 Szlargely follows the orbital conductance, further demonstrating the spin generation process from the orbital. If we increase the SOC strength, G13 Szincreases accordingly, because of the higher orbital-spin conversion eciency (see Figure S3). We also test orbital non-conserved leads with nonzero tsp, whose spin conductance remains the similar feature (see Figure S4). C. Orbital Hall Magnetoresistance As discussed above, the current injected into lead 3 is spin-polarized. When lead 3 is magnetized along the z axis, we expect the existence of magnetization-dependent conductance, i.e. G13(Mz)6=G13(Mz). From the cur-4 FIG. 3. Orbital Magnetoresistance with magnetic leads. (a) 2T and 3T detection devices with the exchange eld Mz in the open and conducting lead 3. Larger spheres represent the interfacial SOC region. (b) Total conductance from lead 1 to lead 2 in Mz eld in the 2T device, with the dephasing term set to 0.001. In the inset, the dephasing dependent  G12for the peak inside the circle is presented. (c) Total conductance from lead 1 to lead 2 in Mz eld in the 3T device. (d) Total, (e) Spin and (f) Orbital conductance from lead 1 to lead 3 in Mz eld in the 3T device. In all these calculations, SOC and Mzis set to 0.2 eV and 0.4 eV, respectively. rent conservation [30], we deduce the relation, G13=G12(8) where GijGij(Mz)Gij(Mz). To demonstrate this,Mzis introduced to lead 3 as an exchange eld to the spin, as shown in the 3T setup in Figure 3(a). Results in Figure 3(c) and 3(d) indicate that  G12and G13 can reach several percentage of the total conductance at some energies. We also con rm that  G12and G13are proportional to the exchange eld strength (see Figure S5). To understand the orbital induced magnetoresistance, the spin and orbital conductance from lead 1 to lead 3 are calculated. As shown in Figure 3(e), G13 Szalmost changes its sign when ipping Mzin lead 3, as expected. And G13 Sz now inversely a ects G13 Lzbecause of the interfacial SOC. When further comparing Figure 3(d) and 3(f), we found that the change of the magnitude of G13 Lzis proportional to the change of total conductance  G13. Therefore, it veri es the scenario in Figure 1(c): when the orbital matches the spin in magnetic leads, G13 Lzand thusG13is higher while G12is accordingly lower. Thus, it indicates the essential role of the orbital in connecting charge and spin in the transport.However, the 2T results exhibit qualitatively di erent features from the 3T results. According to the reciprocity relation [30], the 2T conductance obeys G12(Mz) = G12(Mz). Only when the current conservation is bro- ken, one may obtain the 2T magentoresistance. There- fore, we introduce a dephasing term ito leak electrons into virtual leads [31] to release the above constrain. As shown in the inset of Figure 3(b), the  G12is zero at = 0, rst increases quickly and soon decreases as fur- ther increasing . In the large limit, the system is to- tally out of coherence and thus, the conductance cannot remember the spin and orbital information. We note that the dephasing exists ubiquitously in experiments due to the dissipative scattering for example by electron-phonon interaction and impurities. For the same Mz, the 2T G12(Figure (3b)) roughly exhibits the opposite sign compared to the 3T  G12(Fig- ure (3c)) in the energy window investigated. Unlike that G13follows the change of G13 Lz(Figure (3f)), the change direction of G12depends on the geometry of the 2T de- vice (see Figure S6). The magnitude of the 2T  G12 also depends sensitively on the value of . Its peak value (Figure (3b)), with around 0.001, is comparable with the 3T value in the same parameter regime. However,5 the 3T conductance avoids the strict constrain of the 2T reciprocity, and the existence of 3T OHMR does not rely on the dephasing. Furthermore, in the 3T setup, the magnetoresistance ratio  G13=G13in the third lead is also larger than  G12=G12, because of the lower total conductance of G13. Therefore, we propose that the 3T setup may be more advantageous to detect the OHE. D. Realistic Material Cu Based on the simple square lattice model, we demon- strate two main phenomena, the OHE-induced spin po- larization / spin current current assisted by the atomic SOC on the boundary and the existence of OHMR. We further examine them in a realistic material Cu. This light noble metal is predicted to exhibit the strong OHE. As shown in Figure 4(a), the 2T (without lead 3) and 3T devices are composed of Cu (without SOC) in both the scattering region and leads, and the heavy metal Au (with SOC) at two boundaries. We adopted the tight- binding method to describe the Cu and leads, where 9 atomic orbitals ( s,px,py,pz,dz2,dx2y2,dxy,dyz,dzx) are assigned to each site. The nearest-neighboring and the second-nearest-neighboring hoppings are considered with the Slater-Koster type parameters from Ref. 32. For the heavy metal Au at two sides, the SOC strength is set to 0.37 eV as suggested by Ref. 32. With the tight- binding approach, the rst-principles band structure of Cu is reproduced (see Figure S7). As shown in Figure 4(b), the orbital Berry curvature concentrates on the dorbital region (4 eV2 eV) due to the orbital hybridization, consistent with previ- ous works [12, 15]. After integrating Lz, Figure 4(b) shows that the orbital Hall conductivity is around 6000 (h=e)( cm)1in thedorbital region, even larger than the spin Hall conductivity of Pt. Near the Fermi level, the orbital Hall conductivity is determined by the s-orbital derived bands and reduces to around 1000 ( h=e)( cm)1. For the 2T device, the orbital and spin density at Fermi level are plotted in Figure 4(c). The orbital polarization exists at two sides as a consequence of the OHE. With the heavy metal Au attached, spin polarization is generated, which concentrates on Au atoms and follows the orbital density pattern. To con rm that the spin polarization is induced by the OHE, we arti cially turn o the inter- orbital hopping in Cu to eliminate the OHE, but still keep the SOC in the Au region. Result show that both the orbital and spin polarization disappear (see Figure S8), in accordance with our prediction. For the 3T device, we add a third Cu lead to one SOC side and calculate the spin conductance from lead 1 to lead 3 (G13 Sz). As shown in Figure 4(d), the generatedspin conductance displays an energy-dependence similar to the bulk OH. Near the Fermi level, the spin polar- ization rate can reach 4 %, and it is even around 20% in thedorbital region. Therefore, a sizable spin current can also be generated from the OHE by adding an inter- facial SOC layer. Similarly, when arti cially switching o the OHE of Cu but keeping the Au part, the spin cur- rent disappears, eliminating the contribution of the SHE brought by the thin Au layer (see Figure S9). We also studied the OHMR by applying an exchange eldMzin the lead 3. We choose Mz= 0:95 eV ac- cording to the approximate spin splitting in the tran- sition metal Co (see Figure S10). As shown in Figure 4(e) and 4(f), the 3T OHMR is rather large, where we nd G12=G120:3% and G13=G131:3% at the Fermi level. In experiment, the SHE magentoresistance is around 0:050:5% (see Ref. 33 for example). Therefore, the sizable OHMR in copper can be fairly measurable by present experimental techniques. We should point out that similar e ects can be generalized to other OHE ma- terials like Li and Al [15]. III. SUMMARY In summary, we have proposed the OHE detec- tion strategies by converting the orbital to spin by the interfacial SOC, and inducing the strong spin cur- rent/polarization. Inversely, the OHE can also generate the large nonreciprocal magnetoresistance when employ- ing the magnetic contact. We point out that, compared to the two-terminal one, the three-terminal OHMR does not require the dephasing term , and may be more ad- vantageous to detect the OHE. Using the device setup based on the metal Cu, we demonstrate that the gener- ated spin polarization and OHMR are strong enough to be measured in the present experimental condition. Our work will pave a way to realize the OHE in experiment, and further design orbitronic or even orbitothermal de- vices for future applications. IV. ACKNOWLEDGEMENT We honor the memory of Prof. Shoucheng Zhang. This article follows his earlier works on the intrinsic orbital Hall e ect and spin Hall e ect. B.Y. acknowledges the nancial support by the Willner Family Leadership Insti- tute for the Weizmann Institute of Science, the Benoziyo Endowment Fund for the Advancement of Science, Ruth and Herman Albert Scholars Program for New Scientists, and the European Research Council (ERC) under the European Union's Horizon 2020 research and innovation programme (Grant No. 815869, NonlinearTopo). [1] B. A. Bernevig, T. L. Hughes, and S.-C. Zhang, Physical Review Letters 95, 066601 (2005).[2] S. Murakami, N. Nagaosa, and S.-C. Zhang, Science 301, 1348 (2003).6 FIG. 4. The orbital-spin conversion and orbital magnetoresistance in Cu. (a) 2T/3T device based on real materials, where the scattering region and leads are treated as Cu and the SOC region is treated as Au. (b) The band structure of Cu, weighted by the orbital Berry curvature Lz, and the corresponding orbital Hall conductivity. (c) Orbital and spin density distribution at the Fermi level in the 2T setup. (d) Spin conductance from lead 1 to lead 3 in nonmagnetic leads in the 3T setup. (e) Total conductance from lead 1 to lead 2 in Mzexchange eld in the 3T setup. (f) Magnetoresistance ratio  G12=G12and G13=G13in the 3T setup. [3] J. Sinova, D. Culcer, Q. Niu, N. Sinitsyn, T. Jungwirth, and A. H. MacDonald, Physical review letters 92, 126603 (2004). [4] Y. K. Kato, R. C. Myers, A. C. Gossard, and D. D. Awschalom, science 306, 1910 (2004). [5] J. Wunderlich, B. Kaestner, J. Sinova, and T. Jungwirth, Physical review letters 94, 047204 (2005). [6] T. Jungwirth, J. Wunderlich, and K. Olejn k, Nature materials 11, 382 (2012). [7] C. L. Kane and E. J. Mele, Physical review letters 95, 226801 (2005). [8] B. A. Bernevig, T. L. Hughes, and S.-C. Zhang, science 314, 1757 (2006). [9] G. Y. Guo, Y. Yao, and Q. Niu, Phys. Rev. Lett. 94, 226601 (2005). [10] H. Kontani, T. Tanaka, D. S. Hirashima, K. Yamada, and J. Inoue, Physical Review Letters 100, 096601 (2008), cond-mat/0702447. [11] H. Kontani, T. Tanaka, D. S. Hirashima, K. Yamada, and J. Inoue, Physical Review Letters 102, 016601 (2008), 0806.0210. [12] T. Tanaka, H. Kontani, M. Naito, T. Naito, D. S. Hi- rashima, K. Yamada, and J. Inoue, Physical Review B 77(2008), 10.1103/physrevb.77.165117. [13] I. V. Tokatly, Physical Review B 82, 161404 (2010), 1004.0624. [14] D. Go, D. Jo, C. Kim, and H.-W. Lee, Physical Review Letters 121, 086602 (2018).[15] D. Jo, D. Go, and H.-W. Lee, Phys. Rev. B 98, 214405 (2018). [16] V. T. Phong, Z. Addison, S. Ahn, H. Min, R. Agarwal, and E. J. Mele, Phys. Rev. Lett. 123, 236403 (2019). [17] L. M. Canonico, T. P. Cysne, A. Molina-Sanchez, R. Muniz, and T. G. Rappoport, arXiv preprint arXiv:2001.03592 (2020). [18] S. Bhowal and S. Satpathy, Phys. Rev. B 101, 121112 (2020). [19] D. Go and H.-W. Lee, Physical Review Research 2, 013177 (2020). [20] Y. Liu, J. Xiao, J. Koo, and B. Yan, arXiv:2008.08881 (2020). [21] E. Saitoh, M. Ueda, H. Miyajima, and G. Tatara, Ap- plied physics letters 88, 182509 (2006). [22] S. O. Valenzuela and M. Tinkham, Nature 442, 176 (2006). [23] H. Zhao, E. J. Loren, H. M. van Driel, and A. L. Smirl, Phys. Rev. Lett. 96, 246601 (2006). [24] S. Y. Huang, X. Fan, D. Qu, Y. P. Chen, W. G. Wang, J. Wu, T. Y. Chen, J. Q. Xiao, and C. L. Chien, Phys. Rev. Lett. 109, 107204 (2012). [25] M. Weiler, M. Althammer, F. D. Czeschka, H. Huebl, M. S. Wagner, M. Opel, I.-M. Imort, G. Reiss, A. Thomas, R. Gross, and S. T. B. Goennenwein, Phys. Rev. Lett. 108, 106602 (2012). [26] D. Xiao, M.-C. Chang, and Q. Niu, Rev. Mod. Phys. 82, 1959 (2010).7 [27] N. Nagaosa, J. Sinova, S. Onoda, A. H. MacDonald, and N. P. Ong, Reviews of Modern Physics 82, 1539 (2010). [28] M. B uttiker, Phys. Rev. Lett. 57, 1761 (1986). [29] C. W. Groth, M. Wimmer, A. R. Akhmerov, and X. Waintal, New Journal of Physics 16, 063065 (2014).[30] M. B uttiker, IBM Journal of Research and Development 32, 317 (1988). [31] M. B uttiker, Physical Review B 33, 3020 (1986). [32] D. A. Papaconstantopoulos et al. ,Handbook of the band structure of elemental solids (Springer, 1986). [33] C. O. Avci, K. Garello, A. Ghosh, M. Gabureac, S. F. Alvarado, and P. Gambardella, Nature Physics 11, 570 (2015), 1502.06898.
2206.00784v2.Substrate_Effects_on_Spin_Relaxation_in_Two_Dimensional_Dirac_Materials_with_Strong_Spin_Orbit_Coupling.pdf
Substrate E ects on Spin Relaxation in Two-Dimensional Dirac Materials with Strong Spin-Orbit Coupling Junqing Xu1,and Yuan Ping1,y 1Department of Chemistry and Biochemistry, University of California, Santa Cruz, CA 95064, USA (Dated: December 6, 2022) Understanding substrate e ects on spin dynamics and relaxation in two-dimensional (2D) mate- rials is of key importance for spintronics and quantum information applications. However, the key factors that determine the substrate e ect on spin relaxation, in particular for materials with strong spin-orbit coupling, have not been well understood. Here we performed rst-principles real-time density-matrix dynamics simulations with spin-orbit coupling (SOC) and quantum descriptions of electron-phonon and electron-impurity scattering for the spin lifetimes of supported/free-standing germanene, a prototypical strong SOC 2D Dirac material. We show that the e ects of di er- ent substrates on spin lifetime ( s) can surprisingly di er by two orders of magnitude. We nd that substrate e ects on sare closely related to substrate-induced modi cations of the SOC- eld anisotropy, which changes the spin- ip scattering matrix elements. We propose a new electronic quantity, named spin- ip angle "#, to characterize spin relaxation caused by intervalley spin- ip scattering. We nd that the spin relaxation rate is approximately proportional to the averaged value of sin2 "#=2 , which can be used as a guiding parameter of controlling spin relaxation. INTRODUCTION Since the long spin di usion length ( ls) in large-area graphene was rst reported by Tombros et al.[1], sig- ni cant advances have been made in the eld of spin- tronics, which has the potential to realize low-power electronics by utilizing spin as the information car- rier. Various 2D materials have shown promising spin- tronic properties[2], e.g., long lsat room temperatures in graphene[3] and ultrathin black phosphorus[4], spin- valley locking (SVL) and ultralong spin lifetime sat low temperatures in transition metal dichalcogenides (TMDs)[5] and germanene[6], and persistent spin helix in 2D hybrid perovskites[7]. Understanding spin relaxation and transport mecha- nism in materials is of key importance for spintronics and spin-based quantum information technologies. One critical metric for ideal materials in such applications is spin lifetime ( s), often required to be suciently long for stable detection and manipulation of spin. In 2D- material-based spintronic devices, the materials are usu- ally supported on a substrate. Therefore, for the design of those devices, it is crucial to understand substrate ef- fects on spin relaxation. In past work, the substrate ef- fects were mostly studied for weak SOC Dirac materials like graphene[8{12]. How substrates a ect strong SOC Dirac materials like germanene is unknown. In partic- ular, the spin relaxation mechanism between weak and strong SOC Dirac materials was shown to be drastically di erent. [6] Therefore, careful investigations are required to unveil the distinct substrate e ects on these two types of materials. Here we focus on the dangling-bond-free insulating jxu153@ucsc.edu yyuanping@ucsc.edusubstrates, which interact weakly with the material thus preserve its main physical properties. Insulating sub- strates can a ect spin dynamics and relaxation in sev- eral aspects: (i) They may induce strong SOC elds, so called internal magnetic elds Binby breaking inversion symmetry[9] or through proximity e ects[10]. For ex- ample, the hexagonal boron nitride substrate can induce Rashba-like elds on graphene and dramatically accel- erate its spin relaxation and enhance the anisotropy of sbetween in-plane and out-of-plane directions[8]. (ii) Substrates may introduce additional impurities [11, 12] or reduce impurities/defects in material layers, e.g., by encapsulation[13]. In consequence, substrates may change the electron-impurity (e-i) scattering strength, which a ects spin relaxation through SOC. (iii) Ther- mal vibrations of substrate atoms can introduce addi- tional spin-phonon scattering by interacting with spins of materials[9]. Previously most theoretical studies of substrate e ects on spin relaxation were done based on model Hamil- tonian and simpli ed spin relaxation models[9, 11, 12]. While those models provide rich mechanistic insights, they are lack of predictive power and quantitative ac- curacy, compared to rst-principles theory. On the other hand, most rst-principles studies only simulated the band structures and spin polarizations/textures of the heterostructures[14{16], which are not adequate for un- derstanding spin relaxation. Recently, with our newly- developed rst-principles density-matrix (FPDM) dy- namics approach, we studied the hBN substrate e ect on spin relaxation of graphene, a weak SOC Dirac material. We found a dominant D'yakonov-Perel' (DP) mechanism and nontrivial modi cation of SOC elds and electron- phonon coupling by substrates[8]. However, strong SOC Dirac materials can have a di erent spin relaxation mech- anism - Elliott-Yafet (EY) mechanism[17], with only spin- ip transition and no spin precession, unlike the DParXiv:2206.00784v2 [cond-mat.mes-hall] 4 Dec 20222 mechanism. How substrates a ect spin relaxation of ma- terials dominated by EY mechanism is the key question here. Furthermore, how such e ects vary among di er- ent substrates is another outstanding question for guiding experimental design of interfaces. In our recent study, we have predicted that mono- layer germanene (ML-Ge) is a promising material for spin-valleytronic applications, due to its excellent prop- erties including spin-valley locking, long sandls, and highly tunable spin properties by varying gates and ex- ternal elds[6]. As discussed in Ref. 6, ML-Ge has strong intrinsic SOC unlike graphene and silicene. Under an out-of-plane electric eld (in consequence broken inver- sion symmetry), a strong out-of-plane internal magnetic eld forms, which may lead to mostly EY spin relax- ation [6]. Therefore, predicting sof supported ML-Ge is important for future applications and our understand- ing of substrate e ects on strong SOC materials. Here, we examine the substrate e ects on spin relaxation in ML-Ge through FPDM simulations, with self-consistent SOC and quantum descriptions of e-ph and e-i scatter- ing processes[6, 8, 18{20]. We study free-standing ML- Ge and ML-Ge supported by four di erent insulating substrates - germanane (GeH), silicane (SiH), GaTe and InSe. The choice of substrates is based on similar lat- tice constants to ML-Ge, preservation of Dirac Cones, and experimental synthesis accessibility[21, 22]. We will rst show how electronic structures and sof ML-Ge are changed by di erent substrates - while sof ML- Ge on GeH and SiH are similar to free-standing ML- Ge, the GaTe and InSe substrates strongly reduce sof ML-Ge due to stronger interlayer interactions. We then discuss what quantities are responsible for the disparate substrate e ects on spin relaxation, which eventually an- swered the outstanding questions we raised earlier. RESULTS AND DISCUSSIONS Substrate e ects on electronic structure and spin texture We begin with comparing band structures and spin textures of free-standing and supported ML-Ge in Fig. 1, which are essential for understanding spin relaxation mechanisms. Since one of the most important e ects of a substrate is to induce an out-of-plane electric eld Ez on the material layer, we also study ML-Ge under a con- stantEzas a reference. The choice of the Ezis based on reproducing a similar band splitting to the one in ML-Ge with substrates. The band structure of ML-Ge is similar to graphene with two Dirac cones at KandK0K, but a larger band gap of 23 meV. At Ez= 0, due to time-reversal and inversion symmetries of ML-Ge, every two bands form a Kramers degenerate pair[17]. A nite Ezor a substrate breaks the inversion symmetry and in- duces a strong out-of-plane internal B eld Bin(Eq. 21), which splits the Kramers pairs into spin-up and spin-down bands[6]. Interestingly, we nd that band struc- tures of ML-Ge-SiH (Fig. 1c) and ML-Ge-GeH (Fig. S4) are quite similar to free-standing ML-Ge under Ez=-7 V/nm (ML-Ge@-7V/nm, Fig. 1b), which indicates that the impact of the SiH/GeH substrate on band structure andBinmay be similar to a nite Ez(see Fig. S4). This similarity is frequently assumed in model Hamiltonian studies[9, 11]. On the other hand, the band structures of ML-Ge-InSe (Fig. 1d) and ML-Ge-GaTe (Fig. S4) have more di erences from the free-standing one under Ez, with larger band gaps, smaller band curvatures at Dirac Cones, and larger electron-hole asymmetry of band split- tings. This implies that the impact of the InSe/GaTe substrates can not be approximated by applying an Ez to the free-standing ML-Ge, unlike SiH/GeH substrates. We further examine the spin expectation value vectors Sexpof substrate-supported ML-Ge. Sexpis parallel to Binby de nition (Eq. 21). Sexp Sexp x;Sexp y;Sexp z withSexp ibeing spin expectation value along direction i and is the diagonal element of spin matrix siin Bloch basis. Importantly, from Fig. 1e and 1f, although Sexpof ML-Ge on substrates are highly polarized along z(out-of- plane) direction, the in-plane components of Sexpof ML- Ge-InSe (and ML-Ge-GaTe) are much more pronounced than ML-Ge-SiH (and ML-Ge-GeH). Such di erences are crucial to the out-of-plane spin relaxation as discussed in a later subsection. Spin lifetimes of germanene on substrates and spin relaxation mechanism We then perform our rst-principles density-matrix calculation [6, 18{20] at proposed interfaces, and examine the role of electron-phonon coupling in spin relaxation of ML-Ge at di erent substrates. Throughout this paper, we focus on out-of-plane sof ML-Ge systems, since their in-planesis too short and less interesting. We com- pare out-of-plane sdue to e-ph scattering between the free-standing ML-Ge (with/without an electric eld) and ML-Ge on di erent substrates in Fig. 2a. Here we show electronsfor most ML-Ge/substrate systems as intrin- sic semiconductors, except hole sfor the ML-Ge-InSe interface. This choice is because electron sare mostly longer than hole sat lowTexcept for the one at the ML-Ge-InSe interface; longer lifetime is often more ad- vantageous for spintronics applications. From Fig. 2, we nd thatsof ML-Ge under Ez= 0 and -7 V/nm are at the same order of magnitude for a wide range of tem- peratures. The di erences are only considerable at low T, e.g, by 3-4 times at 20 K. On the other hand, sof supported ML-Ge are very sensitive to the speci c sub- strates. While sof ML-Ge-GeH and ML-Ge-SiH have the same order of magnitude as the free-standing ML- Ge, in particular very close between ML-Ge-GeH and ML-Ge@-7 V/nm, sof ML-Ge-GaTe and ML-Ge-InSe are shorter by at least 1-2 orders of magnitude in the whole temperature range. This separates the substrates3 FIG. 1. Band structures and spin textures around the Dirac cones of ML-Ge systems with and without substrates. (a)-(d) show band structures of ML-Ge under Ez= 0 and under -7 V/nm and ML-Ge on silicane (SiH) and on InSe substrates respectively. (e) and (f) show spin textures in the kx-kyplane and 3D plots of the spin vectors Sexp k1on the circlej !kj= 0:005 bohr1of the band at the band edge around Kof ML-Ge on SiH and InSe substrates respectively. Sexp Sexp x;Sexp y;Sexp z withSexp i being spin expectation value along direction iand is the diagonal element of spin matrix siin Bloch basis. The red and blue bands correspond to spin-up and spin-down states. Due to time-reversal symmetry, band structures around another Dirac cone atK0=Kare the same except that the spin-up and spin-down bands are reversed. The grey, white, blue, pink and green balls correspond to Ge, H, Si, In and Se atoms, respectively. Band structures of ML-Ge on germanane (GeH) and GaTe are shown in Fig. S4 in the Supporting Information, and are similar to those of ML-Ge on SiH and InSe substrates, respectively. In subplots (e) and (f), the color scales Sexp zand the arrow length scales the vector length of in-plane spin expectation value. into two categories, i.e. with a weak e ect (ML-Ge-GeH and ML-Ge-SiH) and a strong e ect (ML-Ge-GaTe and ML-Ge-InSe). We further investigate the role of electron-impurity (e- i) scattering in spin relaxation under di erent substrates, by introducing defects in the material layer. We consider a common type of impurity - single neutral Ge vacancy, whose formation energy was found relatively low in previ- ous theoretical studies[23, 24]. From Fig. 2b, we can see thatsof all ve systems decrease with impurity density ni. Since carrier scattering rates 1 p(carrier lifetime p) increases (decrease) with ni, we then obtain sdecreases withp's decrease, an evidence of EY spin relaxation mechanism. Moreover, we nd that sis sensitive to the type of the substrate with all values of ni, and for each of four substrates, sis reduced by a similar amount with di erentni, from low density limit (109cm2, where e- ph scattering dominates) to relatively high density (1012 cm2, where e-i scattering becomes more important). Since the bands near the Fermi energy are composed of the Dirac cone electrons around KandK0valleys in ML-Ge, spin relaxation process arises from intervalleyand intravalley e-ph scatterings. We then examine rel- ative intervalley spin relaxation contribution (see its de nition in the Fig. 2 caption) in Fig. 2c. being close to 1 or 0 corresponds to intervalley or intravalley scatter- ing being dominant in spin relaxation. becomes close to 1 below 70 K for electrons of ML-Ge-SiH, and below 120 K for holes of ML-Ge-InSe. This indicates that at lowTonly intervalley scattering processes are relevant to spin relaxation in ML-Ge on substrates. This is a re- sult of spin-valley locking (SVL), i.e. large SOC-induced band splittings lock up or down spin with a particular K or K' valley [6]. According to Fig. 1 and 2c, the SVL transition temperature ( TSVL; below which the propor- tion of intervalley spin relaxation rate is close to 1) seems approximately proportional to SOC splitting en- ergy SOC, e.g. for electrons (CBM) of ML-Ge-GaTe and ML-Ge-SiH, and for holes (VBM) of ML-Ge-InSe, SOC are15,24 and 40 meV respectively, while TSVLare 50, 70 and 120 K respectively. As SOCcan be tuned by Ezand the substrate, TSVLcan be tuned simultaneously. Under SVL condition, spin or valley lifetime tends to be exceptionally long, which is ideal for spin-/valley-tronic4 FIG. 2. The out-of-plane spin lifetime sof intrinsic free-standing and substrate-supported ML-Ge. (a) sof ML-Ge under Ez= 0, -7 V/nm and substrate-supported ML-Ge as a function of temperature without impurities. Here we show electron s for intrinsic ML-Ge systems except that hole sis shown for ML-Ge-InSe, since electron sare longer than hole sat lowT except ML-Ge-InSe. (b) sas a function of impurity density niat 50 K. The impurities are neutral ML-Ge vacancy with 50% at higher positions and 50% at lower ones of a Ge layer. The dashed vertical line corresponds to the impurity density where e-ph and e-i scatterings contribute equally to spin relaxation ( ni;s). And e-ph (e-i) scattering is more dominant if ni<(>)ni;s. (c) The proportion of intervalley spin relaxation contribution of (electrons of) ML-Ge-SiH and (holes of) ML-Ge-InSe without impurities. is de ned as =(inter s;z)1 (inters;z)1+w(intras;z)1, whereinter s;z andintra s;z are intervalley and intravalley spin lifetimes, corresponding to scattering processes between KandK0valleys and within a single KorK0valley, respectively. being close to 1 or 0 corresponds to dominant intervalley or intravalley spin relaxation, respectively. wis a weight factor related to what percentage of total Szcan be relaxed out by intravalley scattering itself. wbeing close to 0 and 1 correspond to the cases that intravalley scattering can only relax a small part (0) and most of excess spin (1) respectively. In Supporting Information Sec. SII, we give more details about de nition of w. (d) Electron and hole sat 20 K of ML-Ge without impurities on hydrogen- terminated multilayer Si, labeled as Si nH withnbeing number of Si layers. Si nH is silicane if n= 1, and hydrogen-terminated Silicon (111) surface if n=1. applications. Additionally, the studied substrates here are mono- layer, while practically multilayers or bulk are more com- mon, thus it is necessary to understand how schanges with the number of substrate layers. In Fig. 2d, we showsat 20 K of ML-Ge on hydrogen-terminated mul-tilayer Si, ML-Ge-Si nH, withnbeing number of Si layer. SinH becomes hydrogen-terminated Silicon (111) surface ifn=1. We nd that sare changed by only 30%-40% by increasing nfrom 1 to 3 and kept unchanged after n3. For generality of our conclusion, we also test the layer dependence of a di erent substrate. We found5 thesof ML-Ge on bilayer InSe ( n= 2) is changed by 8% compared to monolayer InSe at 20 K, even smaller change than the one at Si nH substrates. Given the dis- parate properties of these two substrates, we conclude using a monolayer is a reasonable choice for simulating the substrate e ects on sin this work. The correlation of electronic structure and phonon properties to spin relaxation at di erent substrates We next analyze in detail the relevant physical quan- tities, and determine the key factors responsible for sub- strate e ects on spin relaxation. We focus on results under lowTas spin relaxation properties are superior at lowerT(the realization of SVL and longer s). First, to have a qualitative understanding of the material-substrate interaction strength, we show charge density distribution at the cross-section of interfaces in Fig. 3a-d. It seems that four substrates can be catego- rized into two groups: group A contains GeH and SiH with lower charge density distribution in the bonding re- gions (pointed by the arrows); group B contains GaTe and InSe with higher charge density distribution in the bonding regions. In Fig. S5, we investigate the charge density change  e(de ned by the charge density dif- ference between interfaces and individual components). Consistent with Fig. 3, we nd that  efor GaTe and InSe substrates overall has larger magnitude than the one for GeH and SiH substrates. Therefore the material- substrate interactions of group B seem stronger than those of group A. Intuitively, we may expect that the stronger the interaction, the stronger the substrate e ect is. The FPDM simulations in Fig. 2a-b indeed show that the substrate e ects of group B being stronger than those of group A on s, consistent with the above intuition. Next we examine electronic quantities closely related to spin- ip scattering responsible to EY spin relaxation. Qualitatively, for a state k1, its spin- ip scattering rate 1 s(k1) is proportional to the number of its pair states k2allowing spin- ip transitions between them. The num- ber of pair states is approximately proportional to den- sity of states (DOS) around the energy of k1. Moreover, for EY mechanism, it is commonly assumed that spin relaxation rate is proportional to the degree of mixture of spin-up and spin-down states (along the zdirection here), so called \spin-mixing" parameter[17] b2 z(see its de nition in Sec. SII), i.e., 1 s/ b2 z , where b2 z is the statistically averaged spin mixing parameter as de ned in Ref. 6. Therefore, we show DOS, energy-resolved spin- mixingb2 z(") and b2 z as a function of temperature in Fig. 3e-g. We nd that in Fig. 3e DOS of ML-Ge-GeH and ML- Ge-SiH are quite close to that of ML-Ge@-7V/nm, while DOS of ML-Ge-GaTe and ML-Ge-InSe are 50%-100% higher around the band edge. Such DOS di erences are qualitatively explained by the staggered potentials of ML-Ge-GaTe and ML-Ge-InSe being greater than thoseof ML-Ge-GeH and ML-Ge-SiH according to the model Hamiltonian proposed in Ref. 25. In Fig. 3f-g, b2 zof ML- Ge-GeH and ML-Ge-SiH are found similar to ML-Ge@- 7 V/nm, and not sensitive to energy and temperature. On the contrast, for ML-Ge-GaTe and ML-Ge-InSe, their b2 z(") and b2 z increase rapidly with energy and temper- ature. Speci cally, we can see at 300 K, b2 z of ML-Ge- GaTe and ML-Ge-InSe are about 4-20 times of the one of ML-Ge-GeH and ML-Ge-SiH in Fig. 3g. Thus the one order of magnitude di erence of sbetween group A (ML- Ge-GeH and ML-Ge-SiH) and group B (ML-Ge-GaTe and ML-Ge-InSe) substrates at 300 K can be largely ex- plained by the substrate-induced changes of DOS and b2 z . On the other hand, at low T, e.g., at 50 K, b2 z of ML-Ge-GaTe and ML-Ge-InSe are only about 1.5 and 2.5 times of the ones of ML-Ge-GeH and ML-Ge-SiH, and DOS are only tens of percent higher. However, there is still 1-2 order of magnitude di erence of sbetween di erent substrates. Therefore, the substrate e ects on scan not be fully explained by the changes of b2 z and DOS, in particular at relatively low temperature. We then examine if substrate-induced modi cations of phonon can explain the changes of spin relaxation at dif- ferent substrates, especially at low T. We emphasize that at lowT, since spin relaxation is fully determined by intervalley processes (Fig. 2c), the related phonons are mostly close to wavevector K. From Fig. 4, we nd that the most important phonon mode for spin relaxation at lowThas several similar features: (i) It contributes to more than 60% of spin relaxation (see Fig 4a). (ii) Its energy is around 7 meV in the table of Fig. 4a. (iii) Its vibration is exural-like, i.e., atoms mostly vibrate along the out-of-plane direction as shown in Fig. 4b- d. Moreover, for this mode, the substrate atoms have negligible thermal vibration amplitude compared to the one of the materials atoms. This is also con rmed in the layer-projected phonon dispersion of ML-Ge-InSe in Fig. 4e. The purple box highlights the critical phonon mode around K, with most contribution from the mate- rial layer. (iv) The critical phonon mode does not couple with the substrate strongly, since its vibration frequency does not change much when substrate atoms are xed (by comparing Fig. 4e with f). We thus conclude that the substrate-induced modi cations of phonons and ther- mal vibrations of substrate atoms seem not important for spin relaxation at low T(e.g. below 20 K). Therefore, neither the simple electronic quantities b2 and DOS nor the phonon properties can explain the sub- strate e ects on spin relaxation at low T. The determining factors of spin relaxation derived from spin- ip matrix elements On the other hand, with a simpli ed picture of spin- ip transition by the Fermi's Golden Rule, the scattering rate is proportional to the modulus square of the scat- tering matrix elements. For a further mechanistic un-6 FIG. 3. Charge density, density of states (DOS), and spin mixing parameters of free-standing and substrate-supported ML-Ge. Cross-section views of charge density at interfaces of ML-Ge on (a) GeH, (b) SiH, (c) GaTe, and (d) InSe. The Ge layers are above the substrate layers. The unit of charge density is e=bohr3. Charge densities in the regions pointed out by black arrows show signi cant di erences among di erent systems. (e) DOS and (f) energy-solved spin-mixing parameter along zaxis b2 z(") of ML-Ge under Ez=-7 V/nm and on di erent substrates. "edgeis the band edge energy at the valence band maximum or conduction band minimum. The step or sudden jump in the DOS curve corresponds to the edge energy of the second conduction/valence band or the SOC-induced splitting energy at K. (g) The temperature-dependent e ective spin-mixing parameter b2 z of various ML-Ge systems. derstanding, we turn to examine the modulus square of the spin- ip matrix elements, and compare their qual- itative trend with our FPDM simulations. Note that most matrix elements are irrelevant to spin relaxation and we need to pick the \more relevant" ones, by de ning a statistically-averaged function. Therefore, we propose an e ective band-edge-averaged spin- ip matrix element jeg"#j2(Eq. 8). Here the spin- ip matrix element can be for general scattering processes; in the following we focus on e-ph process for simplicity. We also propose a so-called scattering density of states DSin Eq. 9, which measures the density of spin- ip transitions and can be roughly re- garded as a weighted-averaged value of the usual DOS. Based on the generalized Fermi's golden rule, we approx- imately have 1 s/jeg"#j2DSfor EY spin relaxation (see the discussions above Eq. 11 in \Methods" section). As shown in Fig. 5a, 1 sis almost linearly propor- tional tojeg"#j2DSat 20 K. As the variation of DSamong ML-Ge on di erent substrates is at most three times (see Fig. 3e and Fig. S6), which is much weaker than the large variation of 1 s, this indicates that the substrate- induced change of sis mostly due to the substrate- induced change of spin- ip matrix elements. Although jeg"#j2was often considered approximately proportionalto b2 , resulting in 1 s/ b2 , our results in Fig. 3 in the earlier section indicate that such simple approx- imation is not applicable here, especially inadequate of explaining substrate dependence of sat lowT. To nd out the reason why jeg"#j2for di erent sub- strates are so di erent, we rst examine the averaged spin- ip wavefunction overlap jo"#j2(with the reciprocal lattice vector G= 0), closely related to jeg"#j2(Eq. 18 and Eq. 17). From Fig. 5b, 1 sandjo"#j2have the same trend, which implies jeg"#j2andjo"#j2may have the same trend. However, in general, the G6=0elements ofjo"#j2 may be important as well, which can not be unambigu- ously evaluated here. (See detailed discussions in the subsection \Spin- ip e-ph and overlap matrix element" in the \Methods" section). To have deeper intuitive understanding, we then pro- pose an important electronic quantity for intervalley spin- ip scattering - the spin- ip angle "#between two electronic states. For two states ( k1;n1) and (k2;n2) with opposite spin directions, "#is the angle between Sexp k1n1 andSexp k2n2or equivalently the angle between Bin k1and Bin k2. The motivation of examining "#is that: Suppose two wavevectors k1andk2=k1are in two opposite valleys7 (a) Substrate !K(meV) Contribution Ge@-7V/nm 7.7 78% Ge-GeH 6.9 70% Ge-SiH 7.1 64% Ge-GaTe 6.4 90% Ge-InSe 7.2 99% FIG. 4. (a) The phonon energy at wavevector Kof the mode that contributes the most to spin relaxation, and the per- centage of its contribution for various systems at 20 K. We consider momentum transfer K, as spin relaxation is fully de- termined by intervalley processes between KandK0valleys. (b), (c) and (d) Typical vibrations of atoms in 3 3 supercells of (b) ML-Ge@-7 V/nm, (c) ML-Ge-SiH, and (d) ML-Ge- InSe of the most important phonon mode at Karound 7 meV (shown in (a)). The red arrows represent displacement. The atomic displacements smaller than 10% of the strongest are not shown. (e) The layer-projected phonon dispersion of ML- Ge-InSe within 12 meV. The red and blue colors correspond to the phonon displacements mostly contributed from the ma- terial (red) and substrate layer (blue) respectively. The green color means the contribution to the phonon displacements from the material and substrate layers are similar. The pur- ple boxes highlight the two most important phonon modes aroundKfor spin relaxation.(f) Phonon dispersion of ML- Ge-InSe within 12 meV with substrate atoms (InSe) being xed at equilibrium structure and only Ge atoms are allowed to vibrate. Qand -Qrespectively and there is a pair of bands, which are originally Kramers degenerate but splitted by Bin. Due to time-reversal symmetry, we have Bin k1=Bin k2, which means the two states at the same band natk1 andk2have opposite spins and "#between them is zero. Therefore, the matrix element of operator bAbe- tween states ( k1;n) and (k2;n) -Ak1n;k2nis a spin- ipone and we name it as A"# k1k2. According to Ref. 26, with time-reversal symmetry, A"# k1k2is exactly zero. In general, for another wavevector k3within valley - Qbut notk1,A"# k1k3is usually non-zero. One critical quan- tity that determines the intervalley spin- ip matrix ele- mentA"# k1k3for a band within the pair introduced above is"# k1k3. Based on time-independent perturbation theory, we can prove that A"# between two states is approxi- mately proportional to sin "#=2 . The derivation is given in subsection \Spin- ip angle "#for intervalley spin relaxation" in \Methods" section. As shown in Fig. 5c, 1 sof ML-Ge on di erent substrates at 20 K is almost linearly proportional to sin2("#=2)DS, where sin2("#=2) is the statistically- averaged modulus square of sin "#=2 . This indicates that the relation jeg"#j2/sin2("#=2) is nearly perfectly satis ed at low T, where intervalley processes dominate spin relaxation. We additionally show the relations be- tween1 sandjeg"#j2DS,jo"#j2DSandsin2("#=2)DSat 300 K in Fig. S7. Here the trend of 1 sis still approxi- mately captured by the trends of jeg"#j2DS,jo"#j2DSand sin2("#=2)DS, although not perfectly linear as at low T. Since"#is de ned by Sexpat di erent states, sis highly correlated with Sexpand more speci cally with the anisotropy of Sexp(equivalent to the anisotropy of Bin). Qualitatively, the larger anisotropy of Sexpleads to smaller"#and longersalong the high-spin-polarization direction. This nding may be applicable to spin re- laxation in other materials whenever intervalley spin- ip scattering dominates or spin-valley locking exists, e.g., in TMDs[5], Stanene[27], 2D hybrid perovskites with persis- tent spin helix[7], etc. At the end, we brie y discuss the substrate e ects on in-plane spin relaxation ( s;x), whereas only out-of- plane spin relaxation was discussed earlier. From Table SI, we nd that s;xof ML-Ge@-7V/nm and supported ML-Ge are signi cantly (e.g., two orders of magnitude) shorter than free-standing ML-Ge, but the di erences betweens;xof ML-Ge on di erent substrates are rela- tively small (within 50%). This is because: With a non- zeroEzor a substrate, the inversion symmetry broken induces strong out-of-plane internal magnetic eld Bin z (>100 Tesla), so that the excited in-plane spins will pre- cess rapidly about Bin z. The spin precession signi cantly a ects spin decay and the main spin decay mechanism becomes DP or free induction decay mechanism[28] in- stead of EY mechanism. For both DP and free induc- tion decay mechanisms[20, 28], s;xdecreases with the uctuation amplitude (among di erent k-points) of the Bincomponents perpendicular to the xdirection. As the uctuation amplitude of Bin zof ML-Ge@-7V/nm and supported ML-Ge is large (Table SI; much greater than the one ofBin y), theirs;xcan be much shorter than the value of ML-Ge at zero electric eld when EY mechanism dominates. Moreover, since the uctuation amplitude of Bin zof ML-Ge on di erent substrates has the same or-8 FIG. 5. The relation between 1 sand the averaged modulus square of spin- ip e-ph matrix elements jeg"#j2, of spin- ip overlap matrix elementsjo"#j2andsin2("#=2) multiplied by the scattering density of states DSat 20 K. See the de nition of jeg"#j2, jo"#j2andDSin Eq. 8, 19 and 9 respectively. "#is the spin- ip angle between two electronic states. For two states ( k;n) and (k0;n0) with opposite spin directions, "#is the angle between Sexp knandSexp k0n0.sin2("#=2) is de ned in Eq. 24. The variation ofDSamong di erent substrates is at most three times, much weaker than the variations of 1 sand other quantities shown here. der of magnitude (Table SI), s;xof ML-Ge on di erent substrates are similar. CONCLUSIONS In this paper, we systematically investigate how spin relaxation of strong SOC Dirac materials is a ected by di erent insulating substrates, using germanene as a pro- totypical example. Through FPDM simulations of sof free-standing and substrate supported ML-Ge, we show that substrate e ects on scan di er orders of magni- tude among di erent substrates. Speci cally, sof ML- Ge-GeH and ML-Ge-SiH have the same order of mag- nitude as free-standing ML-Ge, but sof ML-Ge-GaTe and ML-Ge-InSe are signi cantly shortened by 1-2 orders with temperature increasing from 20 K to 300 K. Although simple electronic quantities including charge densities, DOS and spin mixing b2 z qualitatively ex- plain the much shorter lifetime of ML-Ge-GaTe/InSe compared to ML-Ge-GeH/SiH in the relatively high T range, we nd they cannot explain the large variations ofsamong substrates at low T(i.e. tens of K). We point out that spin relaxation in ML-Ge and its inter- faces at low Tis dominated by intervalley scattering pro- cesses. However, the substrate-induced modi cations of phonons and thermal vibrations of substrates seem to be not important. Instead, the substrate-induced changes of the anisotropy of Sexpor the spin- ip angles "#which changes the spin- ip matrix elements, are much more cru- cial."#is at the rst time proposed in this article to the best of our knowledge, and is found to be a useful elec- tronic quantity for predicting trends of spin relaxation when intervalley spin- ip scattering dominates.Our theoretical study showcases the systematic inves- tigations of the critical factors determining the spin re- laxation in 2D Dirac materials. More importantly we pointed out the sharp distinction of substrate e ects on strong SOC materials to the e ects on weak SOC ones, providing valuable insights and guidelines for optimizing spin relaxation in materials synthesis and control. METHODS First-Principles Density-Matrix Dynamics for Spin Relaxation We solve the quantum master equation of density ma- trix(t) as the following:[19] d12(t) dt= [He;(t)]12+ 0 BBB@1 2P 3458 < :[I(t)]13P32;4545(t) [I(t)]45P 45;1332(t)9 = ; +H:C:1 CCCA; (1) Eq. 1 is expressed in the Schr odinger picture, where the rst and second terms on the right side of the equa- tion relate to the coherent dynamics, which can lead to Larmor precession, and scattering processes respec- tively. The rst term is unimportant for out-of-plane spin relaxation in ML-Ge systems, since Larmor preces- sion is highly suppressed for the excited spins along the out-of-plane or zdirection due to high spin polarization alongzdirection. The scattering processes induce spin9 relaxation via the SOC. Heis the electronic Hamiltonian. [H;] =HH. H.C. is Hermitian conjugate. The subindex, e.g., \1" is the combined index of k-point and band.P=Peph+Peiis the generalized scattering- rate matrix considering e-ph and e-i scattering processes. For the e-ph scattering[19], Peph 1234 =X qAq 13Aq; 24; (2) Aq 13=r 2 ~gq 12q G(12!q)q n q;(3) whereqandare phonon wavevector and mode, gq is the e-ph matrix element, resulting from the absorp- tion () or emission (+) of a phonon, computed with self-consistent SOC from rst-principles,[29] n q=nq+ 0:50:5 in terms of phonon Bose factors nq, andG  represents an energy conserving -function broadened to a Gaussian of width . For electron-impurity scattering[19], Pei 1234=Ai 13Ai; 24; (4) Ai 13=r 2 ~gi 13q G(13)p niVcell; (5) whereniandVcellare impurity density and unit cell vol- ume, respectively. giis the e-i matrix element computed by the supercell method and is discussed in the next sub- section. Starting from an initial density matrix (t0) prepared with a net spin, we evolve (t) through Eq. 1 for a long enough time, typically from hundreds of ps to a few s. We then obtain spin observable S(t) from(t) (Eq. S1) and extract spin lifetime sfromS(t) using Eq. S2. Computational details The ground-state electronic structure, phonons, as well as electron-phonon and electron-impurity (e-i) matrix elements are rstly calculated using density functional theory (DFT) with relatively coarse kandqmeshes in the DFT plane-wave code JDFTx[30]. Since all sub- strates have hexagonal structures and their lattice con- stants are close to germanene's, the heterostructures are built simply from unit cells of two systems. The lattice mismatch values are within 1% for GeH, GaTe and InSe substrates but about 3.5% for the SiH sub- strate. All heterostructures use the lattice constant 4.025 A of free-standing ML-Ge relaxed with Perdew-Burke- Ernzerhof exchange-correlation functional[31]. The in- ternal geometries are fully relaxed using the DFT+D3 method for van der Waals dispersion corrections[32]. We use Optimized Norm-Conserving Vanderbilt (ONCV) pseudopotentials[33] with self-consistent spin-orbit cou- pling throughout, which we nd converged at a ki- netic energy cuto of 44, 64, 64, 72 and 66 Ry forfree-standing ML-Ge, ML-Ge-GeH, ML-Ge-SiH, ML-Ge- GaTe and ML-Ge-InSe respectively. The DFT calcula- tions use 2424kmeshes. The phonon calculations em- ploy 33 supercells through nite di erence calculations. We have checked the supercell size convergence and found that using 66 supercells lead to very similar results of phonon dispersions and spin lifetimes. For all systems, the Coulomb truncation technique[34] is employed to ac- celerate convergence with vacuum sizes. The vacuum sizes are 20 bohr (additional to the thickness of the het- erostructures) for all heterostructures and are found large enough to converge the nal results of spin lifetimes. The electric eld along the non-periodic direction is applied as a ramp potential. For the e-i scattering, we assume impurity density is suciently low and the average distance between neigh- boring impurities is suciently long so that the interac- tions between impurities are negligible, i.e. at the dilute limit. The e-i matrix gibetween state ( k;n) and (k0;n0) isgi kn;k0n0=hknjViV0jk0n0i, whereViis the poten- tial of the impurity system and V0is the potential of the pristine system. Viis computed with SOC using a large supercell including a neutral impurity that simulates the dilute limit where impurity and its periodic replica do not interact. To speed up the supercell convergence, we used the potential alignment method developed in Ref. 35. We use 55 supercells, which have shown reasonable convergence (a few percent error of the spin lifetime). We then transform all quantities from plane wave basis to maximally localized Wannier function basis[36], and interpolate them[29, 37{41] to substantially ner k and q meshes. The ne kandqmeshes are 384384 and 576576 for simulations at 300 K and 100 K respectively and are ner at lower temperature, e.g., 1440 1440 and 24002400 for simulations at 50 K and 20 K respectively. The real-time dynamics simulations are done with our own developed DMD code interfaced to JDFTx. The energy-conservation smearing parameter is chosen to be comparable or smaller than kBTfor each calculation, e.g., 10 meV, 5 meV, 3.3 meV and 1.3 meV at 300 K, 100 K, 50 K and 20 K respectively. Analysis of Elliot-Yafet spin lifetime In order to analyze the results from real-time rst- principles density-matrix dynamics (FPDM), we com- pare them with simpli ed mechanistic models as dis- cussed below. According to Ref. [18], if a solid-state system is close to equilibrium (but not at equilibrium) and its spin relaxation is dominated by EY mechanism, its spin lifetime sdue to the e-ph scattering satis es (for simplicity the band indices are dropped)10 1 s/N2 k X kq8 < :jg"#;q k;kqj2nqfkq(1fk) (kkq!q)9 = ;;(6) =N1 kX kfk(1fk); (7) wherefis Fermi-Dirac function. !qandnqare phonon energy and occupation of phonon mode at wavevector q.g"#is the spin- ip e-ph matrix element between two electronic states of opposite spins. We will further discuss g"#in the next subsection. According to Eq. 6 and 7, 1 sis proportional to jg"# qj2 and also the density of the spin- ip transitions. Therefore we propose a temperature ( T) and chemical potential (F;c) dependent e ective modulus square of the spin- ip e-ph matrix element jeg"#j2and a scattering density of statesDSas jeg"#j2=P kqwk;kqP jg"#;q k;kqj2nqP kqwk;kq; (8) DS=N2 kP kqwk;kq N1 kP kfk(1fk); (9) wk;kq=fkq(1fk)(kkq!c); (10) where!cis the characteristic phonon energy speci ed below, and w k;kqis the weight function. The matrix element modulus square is weighted by nqaccording to Eq. 6 and 7. This rules out high-frequency phonons at lowTwhich are not excited. !cis chosen as 7 meV at 20 K based on our analysis of phonon-mode-resolved contribution to spin relaxation. w k;kqselects transitions between states separated by !cand around the band edge orF;c, which are \more relevant" transitions to spin relaxation. DScan be regarded as an e ective density of spin- ip e-ph transitions satisfying energy conservation be- tween one state and its pairs. When !c= 0, we haveDS=R d df d D2()=R d df d D() with D() density of electronic states (DOS). So DScan be roughly regarded as a weighted-averaged DOS with weight df d D(). Withjeg"#j2andDS, we have the approximate relation for spin relaxation rate, 1 s/jeg"#j2DS: (11) Spin- ip e-ph and overlap matrix element In the mechanistic model of Eq. 6 in the last section, the spin- ip e-ph matrix element between two electronic states of opposite spins at wavevectors kandkqof phonon mode reads[29]g"#;q kkq=D u"(#) k qvKS u#(") kqE ; (12) qvKS=s ~ 2!qX  e ;q@ qvKS pm; (13) @ qvKS=X leiqRl@VKS @ jrRl; (14) VKS=V+~ 4m2c2rrVp; (15) whereu"(#) kis the periodic part of the Bloch wavefunc- tion of a spin-up (spin-down) state at wavevector k.is the index of ion in the unit cell. is the index of a di- rection. Rlis a lattice vector. Vis the spin-independent part of the potential. pis the momentum operator. is the Pauli operator. From Eqs. 12-15, g"#can be separated into two parts, g"#=gE+gY; (16) wheregEandgYcorrespond to the spin-independent and spin-dependent parts of VKSrespectively, called El- liot and Yafet terms of the spin- ip scattering matrix elements respectively.[28] Generally speaking, both the Elliot and Yafet terms are important; for the current systems swith and with- out Yafet term have the same order of magnitude. For example,sof ML-Ge-GeH and ML-Ge-SiH without the Yafet term are about 100% and 70% of swith the Yafet term at 20 K. Therefore, for qualitative discussion of s of ML-Ge on di erent substrates (the quantitative calcu- lations ofsare performed by FPDM introduced earlier), it is reasonable to focus on the Elliot term gEand avoid the more complicated Yafet term gY. De neVE qas the spin-independent part of  qvKS, so thatgE=D u"(#) k VE q u#(") kqE . Expanding VE qas P GeVE q(G)eiGr, we have gE=X GeVE q(G)o"# kkq(G); (17) o"# kkq(G) =D u"(#) k eiGr u#(") kqE ; (18) whereo"# kkq(G) isG-dependent spin- ip overlap func- tion. Without loss of generality, we suppose the rst Brillouin zone is centered at . Therefore,gEis not only determined by the long-range component of o"# kkq(G), i.e.,o"# kkq(G= 0) but also the G6= 0 components. But nevertheless, it is helpful to investigate o"# kkq(G= 0) and similar to Eq. 8, we pro- pose an e ective modulus square of the spin- ip overlap matrix elementjo"#j2,11 jo"#j2=P kqwk;kqP jo"# k;kq(G= 0)j2 P kqwk;kq: (19) Internal magnetic eld Suppose originally a system has time-reversal and in- version symmetries, so that every two bands form a Kramers degenerate pair. Suppose the k-dependent spin matrix vectors in Bloch basis of the Kramers degenerate pairs are s0 kwiths(sx;sy;sz). The inversion symme- try broken, possibly due to applying an electric eld or a substrate, induces k-dependent Hamiltonian terms HISB k=BgeBin ks0 k; (20) whereBgeis the electron spin gyromagnetic ratio. Bin kis the SOC eld and called internal magnetic elds. Binsplits the degenerate pair and polarizes the spin along its direction. The de nition of Bin kis Bin k2SOC kSexp k=(Bge); (21) where Sexp Sexp x;Sexp y;Sexp z withSexp ibeing spin expectation value along direction iand is the diagonal element ofsi. SOCis the band splitting energy by SOC. Spin- ip angle "#for intervalley spin relaxation Suppose (i) the inversion symmetry broken induces Bin k (Eq. 21) for a Kramers degenerate pair; (i) there are two valleys centered at wavevectors QandQand (iii) there are two wavevectors k1andk2nearQandQrespec- tively. Due to time-reversal symmetry, the directions of Bin k1andBin k2are almost opposite. De ne the spin- ip angle "# k1k2as the angle between Bin k1andBin k2, which is also the angle between Sexp k1 andSexp k2. We will prove that for a general operator bA, A"# k1k2 2 sin2 "# k1k2=2 A## k1k2 2 ; (22) whereA"# k1k2andA## k1k2are the spin- ip and spin- conserving matrix elements between k1andk2respec- tively. The derivation uses the rst-order perturbation theory and has three steps: Step 1: The 22 matrix of operator bAbetween k1and k2of two Kramers degenerate bands is A0 k1k2. According to Ref. 26, with time-reversal symmetry, the spin- ip matrix element of the same band between kandkis exactly zero, therefore, the spin- ip matrix elements ofA0 k1k2are zero at lowest order as k1+k20, i.e.,A0;"# k1k2 A0;#" k1k20. Step 2: The inversion symmetry broken induces Bin k and the perturbed Hamiltonian HISB k(Eq. 20). The new eigenvectors Ukare obtained based on the rst-order perturbation theory. Step 3: The new matrix is Ak1k2=Uy k1A0 k1k2Uk2. Thus the spin- ip matrix elements A"# k1k2with the inversion symmetry broken are obtained. We present the detailed derivation in SI Sec. III. From Eq. 22, for the intervalley e-ph matrix elements of ML-Ge systems, we have g"# k1k2 2 sin2 "# k1k2=2 g## k1k2 2 : (23) As g"# k1k2 2 largely determines sof ML-Ge systems, the di erences of sof ML-Ge on di erent substrates should be mainly due to the di erence of sin2 "# k1k2=2 . For the intervalley overlap matrix elements, we should have o"# k1k2 2 sin2 "# k1k2=2 o## k1k2 2 . Since o## k1k2 2 is of order 1, o"# k1k2 2 is expected proportional to sin2 "# k1k2=2 and have the same order of magnitude as sin2 "# k1k2=2 . Finally, similar to Eq. 8, we propose an e ective mod- ulus square of sin2 "# k1k2=2 , sin2("#=2) =P kqwk;kqsin2 "# k;kq=2 P kqwk;kq: (24) DATA AVAILABILITY The data that support the ndings of this study are available upon request to the corresponding author. CODE AVAILABILITY The codes that were used in this study are available upon request to the corresponding author. ACKNOWLEDGEMENTS We thank Ravishankar Sundararaman for helpful dis- cussions. This work is supported by the Air Force Of- ce of Scienti c Research under AFOSR Award No. FA9550-YR-1-XYZQ and National Science Foundation under grant No. DMR-1956015. This research used resources of the Center for Functional Nanomaterials,12 which is a US DOE Oce of Science Facility, and the Scienti c Data and Computing center, a component of the Computational Science Initiative, at Brookhaven Na- tional Laboratory under Contract No. DE-SC0012704, the lux supercomputer at UC Santa Cruz, funded by NSF MRI grant AST 1828315, the National Energy Research Scienti c Computing Center (NERSC) a U.S. Depart- ment of Energy Oce of Science User Facility operated under Contract No. DE-AC02-05CH11231, and the Ex- treme Science and Engineering Discovery Environment (XSEDE) which is supported by National Science Foun- dation Grant No. ACI-1548562 [42].AUTHOR CONTRIBUTIONS J.X. performed the rst-principles calculations. J.X. and Y.P. analyzed the results. J.X. and Y.P. designed all aspects of the study. J.X. and Y.P. wrote the manuscript. ADDITIONAL INFORMATION Supplementary Information accompanies the pa- per on the npj Computational Materials website. Competing interests: The authors declare no com- peting interests. REFERENCES [1] N. Tombros, C. Jozsa, M. Popinciuc, H. T. Jonkman, and B. J. Van Wees, Nature 448, 571 (2007). [2] A. Avsar, H. Ochoa, F. Guinea, B. Ozyilmaz, B. J. van Wees, and I. J. Vera-Marun, Rev. Mod. Phys. 92, 021003 (2020). [3] M. Drogeler, C. Franzen, F. Volmer, T. Pohlmann, L. Banszerus, M. Wolter, K. Watanabe, T. Taniguchi, C. Stampfer, and B. Beschoten, Nano Lett. 16, 3533 (2016). [4] A. Avsar, J. Y. Tan, M. Kurpas, M. Gmitra, K. Watan- abe, T. Taniguchi, J. Fabian, and B. Ozyilmaz, Nat. Phys. 13, 888 (2017). [5] P. Dey, L. Yang, C. Robert, G. Wang, B. Urbaszek, X. Marie, and S. A. Crooker, Phys. Rev. Lett. 119, 137401 (2017). [6] J. Xu, H. Takenaka, A. Habib, R. Sundararaman, and Y. Ping, Nano Lett. 21, 9594 (2021). [7] L. Zhang, J. Jiang, C. Multunas, C. Ming, Z. Chen, Y. Hu, Z. Lu, S. Pendse, R. Jia, M. Chandra, et al. , Nat. Photon. 16, 529 (2022). [8] A. Habib, J. Xu, Y. Ping, and R. Sundararaman, Phys. Rev. B 105, 115122 (2022). [9] C. Ertler, S. Konschuh, M. Gmitra, and J. Fabian, Phys. Rev. B 80, 041405 (2009). [10] A. W. Cummings, J. H. Garcia, J. Fabian, and S. Roche, Phys. Rev. Lett. 119, 206601 (2017). [11] D. Van Tuan, F. Ortmann, A. W. Cummings, D. Soriano, and S. Roche, Sci. Rep. 6, 1 (2016). [12] P. Zhang and M. Wu, New J. Phys. 14, 033015 (2012). [13] J. Li, M. Goryca, K. Yumigeta, H. Li, S. Tongay, and S. Crooker, Phys. Rev. Mater. 5, 044001 (2021). [14] Z. Ni, E. Minamitani, Y. Ando, and S. Watanabe, Phys. Chem. Chem. Phys. 17, 19039 (2015). [15] T. Amlaki, M. Bokdam, and P. J. Kelly, Phys. Rev. Lett. 116, 256805 (2016). [16] K. Zollner, A. W. Cummings, S. Roche, and J. Fabian, Phys. Rev. B 103, 075129 (2021). [17] I. Zuti c, J. Fabian, and S. D. Sarma, Rev. Mod. Phys. 76, 323 (2004). [18] J. Xu, A. Habib, S. Kumar, F. Wu, R. Sundararaman,and Y. Ping, Nat. Commun. 11, 2780 (2020). [19] J. Xu, A. Habib, R. Sundararaman, and Y. Ping, Phys. Rev. B 104, 184418 (2021). [20] J. Xu, K. Li, U. N. Huynh, J. Huang, V. Vardeny, R. Sun- dararaman, and Y. Ping, arXiv:2210.17074 (2022). [21] T. Giousis, G. Potsi, A. Kouloumpis, K. Spyrou, Y. Geor- gantas, N. Chalmpes, K. Dimos, M.-K. Antoniou, G. Pa- pavassiliou, A. B. Bourlinos, et al. , Angew. Chem. 133, 364 (2021). [22] S. Lei, L. Ge, S. Najmaei, A. George, R. Kappera, J. Lou, M. Chhowalla, H. Yamaguchi, G. Gupta, R. Vajtai, et al. , ACS Nano 8, 1263 (2014). [23] J. E. Padilha and R. B. Pontes, Solid State Commun. 225, 38 (2016). [24] M. Ali, X. Pi, Y. Liu, and D. Yang, AIP Advances 7, 045308 (2017). [25] D. Kochan, S. Irmer, and J. Fabian, Phys. Rev. B 95, 165415 (2017). [26] Y. Yafet, in Solid state physics , Vol. 14 (Elsevier, 1963) pp. 1{98. [27] L. Tao and E. Y. Tsymbal, Phys. Rev. B 100, 161110 (2019). [28] M. Wu, J. Jiang, and M. Weng, Phys. Rep. 493, 61 (2010). [29] F. Giustino, Rev. Mod. Phys. 89, 015003 (2017). [30] R. Sundararaman, K. Letchworth-Weaver, K. A. Schwarz, D. Gunceler, Y. Ozhabes, and T. A. Arias, SoftwareX 6, 278 (2017). [31] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996). [32] S. Grimme, J. Antony, S. Ehrlich, and H. Krieg, J. Chem. Phys. 132, 154104 (2010). [33] D. R. Hamann, Phys. Rev. B 88, 085117 (2013). [34] S. Ismail-Beigi, Phys. Rev. B 73, 233103 (2006). [35] R. Sundararaman and Y. Ping, J. Chem. Phys. 146, 104109 (2017). [36] N. Marzari and D. Vanderbilt, Phys. Rev. B 56, 12847 (1997). [37] A. M. Brown, R. Sundararaman, P. Narang, W. A. God- dard, and H. A. Atwater, ACS Nano 10, 957 (2016).13 [38] P. Narang, L. Zhao, S. Claybrook, and R. Sundarara- man, Adv. Opt. Mater. 5, 1600914 (2017). [39] A. M. Brown, R. Sundararaman, P. Narang, A. M. Schwartzberg, W. A. Goddard III, and H. A. Atwater, Phys. Rev. Lett. 118, 087401 (2017). [40] A. Habib, R. Florio, and R. Sundararaman, J. Opt. 20, 064001 (2018).[41] A. M. Brown, R. Sundararaman, P. Narang, W. A. God- dard III, and H. A. Atwater, Phys. Rev. B 94, 075120 (2016). [42] J. Towns, T. Cockerill, M. Dahan, I. Foster, K. Gaither, A. Grimshaw, V. Hazlewood, S. Lathrop, D. Lifka, G. D. Peterson, R. Roskies, J. R. Scott, and N. Wilkins-Diehr, Comput. Sci. Eng. 16, 62 (2014).
1908.00927v2.Two_dimensional_orbital_Hall_insulators.pdf
Two-dimensional orbital Hall insulators Luis M. Canonico,1Tarik P. Cysne,2,3Tatiana G. Rappoport,2,4and R. B. Muniz1 1Instituto de Física, Universidade Federal Fluminense, 24210-346 Niterói RJ, Brazil 2Instituto de Física, Universidade Federal do Rio de Janeiro, Caixa Postal 68528, 21941-972 Rio de Janeiro RJ, Brazil 3Departamento de Física, Universidade Federal de São Carlos, Rod. Washington Luís, km 235 - SP-310, 13565-905 São Carlos, SP, Brazil 4Department of Physics and Center of Physics, University of Minho, 4710-057, Braga, Portugal (Dated: March 4, 2020) Detailed analyses of the spin and orbital conductivities are performed for different topological phases of certain classes of two-dimensional (2D) multiorbital materials. Our calculations show the existence of orbital-Hall effect (OHE) in topological insulators, with values that exceed those obtained for the spin-Hall effect (SHE). Notably, we have found non-topological insulating phases that exhibit OHE in the absence of SHE. We demonstrate that the OHE in these systems is deeply linked to exotic momentum-space orbital textures that are triggered by an intrinsic Dresselhaus- type of interaction that arises from a combination of orbital attributes and lattice symmetry. Our resultsstronglyindicatesthatotherclassesofsystemswithnon-trivialorbitaltexturesand/ororbital magnetism may also exhibit large OHE even in their normal insulating phases. The OHE, similarly to the SHE, refers to the creation of a transverse flow of orbital angular momentum that is induced by a longitudinally applied electric field [1]. It has been explored mostly in three dimensional metallic systems, where it can be quite strong [2–5]. For systems in which the spin-orbit coupling (SOC) is sizeable, the orbital and spin angular momentum degrees of freedom are coupled, establishing an interrelationship between charge, spin, and orbital angular momentum excitations. However, the OHE does not necessarily require SOC, it can be associated to the presence of orbital textures [5] and be especially significant in various materials. Chiral orbital textures in the reciprocal space have been discussed in connection with orbital magnetism at the surface of spmetals [6], photonic graphene [7] and also in topological insulators with strong SOC. More re- cently they were observed in chiral borophene [8], single- layer transition metal dichalcogenides [9] and tin tel- luride monolayers for photocurrent generation [10]. Or- bital magnetism is enhanced in surfaces [11], indicating that orbital effects can be crucial in 2D materials, which can also be evidenced by the observation of orbital tex- tures in van der Waals materials. Still, OHE remains mostly unexplored in 2D materials [12, 13]. Here, we investigate the role of orbital textures for the OHE displayed by multi-orbital 2D materials. We pre- dict the appearance of rather large OHE in these systems both in their metallic and insulating phases. The orbital Hall currents can be considerably larger than the spin Hall ones, and be present even in the absence of SHE. Their use as information carriers widens the development possibilities of novel spin-orbitronic devices. In our analyses, we consider a minimal tight-binding (TB) model Hamiltonian that involves only two orbitals (pxandpy) per atom in a honeycomb lattice [14, 15]:H=X hijiX st ijpy ispjs+X is i+I`z z ss py ispis; (1) whereiandjdenote the honeycomb lattice sites posi- tioned at~Riand~Rj, respectively. The symbol hijiindi- cates that the sum is restricted to the nearest neighbour (n.n) sites only. The operator py iscreates an electron of spinsin the atomic orbitals p=p=1p 2(pxipy) centred at ~Ri. Here,s=";#labels the two electronic spin states, and iis the atomic energy at site i, which may symbolise a staggered on-site potential that takes valuesi=VAB, when site i belongs to the A and B sub-lattices of the honeycomb arrangement, respectively. The transfer integrals t ijbetween the porbitals centred on n.n atoms are parametrised according to the standard Slater-Koster TB formalism [16]. They depend on the direction cosines of the n.n. interatomic directions, and may be approximately expressed as linear combinations of two other integrals ( VppandVpp) involving the p andporbitals, where andrefer to the usual compo- nents of the angular momentum around these axes. Since our model does not include the orbital pz, it is restricted to a sector of the `= 1angular momentum vector space spanned only by the eigenstates of `z p associated with m`=1. Within this sector it is useful to introduce a pseudo angular momentum SU(2)-algebra where the Pauli matrices act on p . In this case, there is a one-to-one correspondence between the representa- tions of the Cartesian components of the orbital angular momentum operators in this basis and the usual Pauli matrices, and `zis not conserved (details are given Sec. I of the supplementary material - SM). The last term in Eq.1 describes the intrinsic atomic SOC. This simple model describes relatively well the low-arXiv:1908.00927v2 [cond-mat.mes-hall] 2 Mar 20202 Figure 1: (a) Band structure calculations along some symme- try lines in the 2D BZ for Vpp= 0,Vpp=1 eV, andI= 0. The blue line represents the results for VAB= 0:0, and the red line for VAB= 0:8. (b) Orbital Hall conductivities cal- culated for the same sets of parameters. The insets show the in-plane contribution to the orbital angular momentum tex- tures calculated in the neighbourhoods the (left inset) and K(right inset) symmetry points of the 2D Brillouin zone, for VAB= 0:0. The left and right inset textures are associated with the lower flat and dispersive bands, respectively. energy electronic properties of novel group V based 2D materials [15, 17, 18]. Its topological characteristics were previously investigated in the context of optical lattices, and it has been verified that it exhibits a rich topological phase diagram, which includes quantum spin-Hall insu- lator (QSHI) phases [17, 19–22]. Following Ref. 21 we shall assume, for simplicity, that Vpp= 0andVpp= 1eV. Our focus is on three dis- tinct phases that manifest themselves depending on the parameters specified in Eq. (1). In the absence of SOC and sub-lattice resolved potentials, the electronic band structure consists of four gapless bulk energy bands, two of which form Dirac cones at the KandK0symmetry points of the 2D first Brillouin zone (BZ), whereas the other two are flat. Each flat band is tangent to one of the dispersive bands at the point, as Fig. 1 (a) illus- trates. Our results for the orbital Hall conductivities ( z OH), calculated as functions of energy by means of the Kubo formula [23], with the orbital current defined as J`z y= 1 2f`z;vyg, are shown in Fig. 1 (b) for VAB= 0:0(blue line), and for VAB= 0:8(red line). Details of these calculations are given in Sec. II of SM. Here we notice a strong orbital Hall conductivity, which peaks at energies close to where the flat bands touch the dispersive bands at. ForVAB6= 0, the electronic structure develops an energy gap around E= 0that eliminates the original Dirac cones in the vicinities of KandK0. The flat bands, however, remain tangent to the dispersive bands at , as shown in Fig. 1 (a), and the large OHE in this case also occurs for energies close to where they touch each other. The insets of Fig. 1 (b) depict the in-plane contribu- tion to the orbital angular momentum textures, calcu- lated on a circle around the (left inset) and K(right inset) symmetry points of the 2D first BZ. They are both computed for VAB= 0. The colours of the arrows em- phasise their in-plane azimuthal angles. At the point,the texture displays a dipole-field like structure, whereas in the vicinity of the Kpoints it is identical to the spin- texture produced by the Dresselhaus SOC in zinc blende lattice systems [24]. Here, the texture is not caused by SOC, but results only from the orbital features and crys- talline symmetry, as we shall subsequently show. In the presence of SOC, three energy gaps open: one originating from the K(K0)points, and the other two at , while the flat bands acquire a slight energy disper- sion - see Fig. SI of the SM. When the relative values ofIandVABvary, this model exhibits a rich topolog- ical phase diagram [21]. We shall focus on three phases that display distinct topological gap features. They are classified by sets of spin Chern numbers (i,j,k,l) associ- ated with the four "-spin bands, namely A1 (1,-1,1,-1), B1 (1,0,0,-1), and B2 (0,1,-1,0), according to the nota- tion of Ref. 21. We remind that the spin Chern numbers for the#-spin sector have opposite signs. This codifica- tion clearly indicates that when the system is in the A1 phase the two lateral energy gaps are topological, but the central one is not. The reverse occurs in the B2 phase, where only the central energy gap is topological. Last but not least, all the three energy gaps are topological in the B1 phase. This is explicitly verified in the left panels of Fig. 2, which show the spin Hall conductivi- ties (z SH) (red curves) calculated as functions of energy for three different sets of parameters that simulate sys- tems in each of these phases. In the absence of sublattice asymmetry ( VAB= 0) and forI= 0:2, the system as- sumes the B1 phase and becomes a QSHI within all the three energy gaps, as the quantised plateaux of z SHin Fig. 2 (a) show. For I= 1:1andVAB= 0:8, the system is in the B2 phase, which exhibits a quantised spin Hall conductivity plateau in the central energy gap, and two non topological side gaps within which it behaves as an ordinary insulator, displaying no QSHE, as Fig. 2 (c) il- lustrates. For I= 0:2andVAB= 0:8, the system takes on the A1 phase, where it becomes a QSHI for energies within the lateral energy gaps, but behaves as a conven- tional insulator inside the central gap, as portrayed in Fig. 2 (e). The corresponding orbital Hall conductivities ( z OH) calculated for the three phases (blue curves) are also de- picted in the left panels of Fig. 2, together with the respective densities of states (grey lines) represented in arbitrary units. We notice that within the lateral gaps, z OHexhibits plateaux with much higher intensities than those of the SHE. However, in contrast with the latter, the OHE is not quantised. Its plateaux heights depend uponIandVAB, increasinginmodulusasthegapwidth reduces, thoughlimitedbytheOHEvaluefor I!0(see Sec. IV of SM). A remarkable result illustrated in Fig. 2 (c) is the existence of finite OHE within the two (non- topological) side energy gaps of phase B2, where the sys- tem becomes an ordinary insulator with no QSHE. This is particularly interesting because there are no electronic3 edge states crossing these energy gaps (see Sec. V of the SM), and raises the question on how the orbital Hall cur- rent propagates through the system in this case. It is also noticeable that the OHE is an odd function of the Fermi energy ( EF)and vanishes in the central energy gaps for all three phases. This is due to symmetry lim- itations of this simplified model, which we shall address subsequently. Figure 2: Spin Hall conductivity (red), and orbital Hall con- ductivity (blue), together with the density of states (grey), calculated as functions of energy for: (a) I= 0:2and VAB= 0- B1 phase; (c) I= 1:1, andVAB= 0:8- B2 phase, and (e) I= 0:2, andVAB= 0:8- A1 phase. The densities of states are depicted in arbitrary units. Panels (b), (d) and (f) show the associated orbital textures, calculated for the lower"-spin band, with the same sets of parameters, respectively. The density plots illustrate their corresponding h`zipolarisations. It is also important to examine how disorder affects the transport properties of these systems. To simulate it we consider on-site potentials iwith values randomly distributed within [-W/2, W/2], where W is the disorder strength. We calculate the spin- and orbital-Hall conduc- tivities for different values of W using Chebyshev poly- nomial expansions and the Kubo-Bastin formula, which are efficiently implemented in the open-source software KITE. Similarly to what we have previously found forthe SHE [22], the orbital Hall plateaux are robust to rel- atively strong Anderson disorder. Details of these calcu- lations are described in sections VI, VII and VIII of the SM. We know that OHE is linked to orbital textures in re- ciprocal space [5], and to establish this relationship we have calculated these textures for the "-spin lowest en- ergybandintheentire2DfirstBZ.Theresultsareshown in panels (b), (d), and (f) for systems in the B1, B2, and A1 phases, respectively. The orbital characters for all "- spin eigenstates are depicted in Fig. SV of the SM. It is worth noticing that when either IorVABare different from zero, the orbital textures display finite out-of-plane components for each spin direction. However, due to time reversal symmetry the `zorbital polarisations for inverted spin directions are opposite, and consequently the total`zpolarisation vanishes. The structure of the in-plane texture, nevertheless, remains the same for both spin components, which means that the in-plane orbital texture survives. It is also noteworthy that both the low- est two energy bands as well as the upper ones display opposite in-plane orbital textures for this simple model. Consequently, the OHE vanishes at the onset of the cen- tral energy gap, where the accumulated in-plane orbital texture of the occupied states becomes zero. The absence of electronic states within an energy gap leads to a con- stant value for z OH[2, 25, 26] in its range, which justifies the lack of OHE in the central energy gap found for the three phases. Contour curves are also shown for certain values of EFranging from the bottom of the energy band to the beginningofthelowestenergygap. Inallphases, wenote that close to the point, where the lowest energy band value is minimum, there is virtually no in-plane orbital angular momentum texture, and the OHE is very small. AsEFincreases the in-plane orbital texture builds up, assuming a dipole-field like configuration. Eventually, whenEFapproaches the onset of the first energy gap, it develops a Dresselhaus-like arrangement near the K and K’ points, with opposite signs in each valley. In order to uncover the raison d’etre of these exotic or- bital textures that promote OHE in this systems we de- rive an effective theory near the Dirac points KandK0. Around them, the orbital angular momentum texture is perfectly captured by a linear approximation in the crys- talline momentum, whereas it requires a fourth-order ex- pansion near the point. Our effective Hamiltonian Heff can be expressed in terms of SU(2) SU(2)orbital and sub-lattice algebras, and written as: Heff=H0+HAB+ HSOC+H`. HereH0=~vF(kxx+kyy)istheusual Dirac Hamiltonian, with Fermi velocity vF=ap 3 2~,ade- notes the lattice constant, and =1for theKandK0 valleys, respectively. HSOC=sI`zrepresents the SOC, wheres=1for"and#spin electrons, respectively. HAB=VABzis the sub-lattice resolved potential. H`4 breaksthedegeneracybetween `zeigenstatesandisgiven by: H`=~vF 4(k+`++k`)p 3~vF 2a(`xx+`yy); (2) where=x+iy,=,` ( =x;y) are the orbital angular momentum matrices in the corresponding Hilbert space, k=kxiky, and`=`xi`y. As shown in the section X of the SM, in the absence ofH`each valley presents two degenerated Dirac cones. The first term in the right hand side of Eq. (2) alters the Fermi velocity of the Dirac cones and leads to an in- plane orbital texture profile similar to the one portrayed around the point. The second term, however, pro- duces a Dresselhaus-like splitting in the Dirac cones and is primarily responsible for the orbital angular momen- tum texture found in our TB calculations. Our effective theory confirms that the exotic in-plane texture exhib- ited by these 2D systems is an intrinsic property that arises solely from the px-pyorbital characteristics and crystalline symmetries. Figure 3: (a) Energy band spectrum calculated for the simple model with second n.n. hopping integrals V(2) pp=0:2. Here we keepVpp= 0,Vpp= 1,I= 0:2andVAB= 0:8. (b) SH (red line) and OH (blue line) conductivities calculated as functions of energy. The grey line depicts the density of states (DOS) in arbitrary units. The inset shows a closeup of the central energygaphighlightingthenon-zero valueoftheOHE within this energy range. We shall now address the absence of OHE in the cen- tral energy gap as results from our calculations. This limitation actually comes from a combination of electron- hole and parity symmetries, which lead to energy levels that are symmetric with respect to the zero energy for this simple model[21]. One way of breaking it is by in- troducing second n.n. hopping integrals, as Fig. 3 (a) illustrates. Here, just as a proof of concept, we kept Vpp= 0, and choose the second n.n. hopping integrals Vpp2=0:2. In this case only the central energy gap survives, and within it the system assumes an ordinary insulating phase. Fig. 3 (b) clearly shows that the SHE vanishes in this energy range, whereas the OHE is fi- nite. Here, the in-plane orbital texture associated with Figure 4: SH (red line) and OH (blue line) conductivities calculated as functions of energies for flat bismuthene: (a) without sublattice asymmetry ( VAB= 0) and (b) with VAB= 0:87. The insets highlight the non-zero values of the OHE within the corresponding central energy-gap ranges.The grey line depicts the DOS in arbitrary units. the second lowest energy band no longer cancels the con- tribution from the first band. Thus, the OHE does not vanish at the onset of the central energy gap and keeps its non-zero value constant within it. This result, al- though relatively small in this particular case, unequiv- ocally shows that it is possible to obtain a finite OHE for a non-topological insulating phase, as we previously found for the lateral energy gaps of the B2 phase. Hav- ing shown that this effect happens for our simple-model system, it is instructive to inquire into the possibility of observing it in a real system. A candidate is the recently synthesised flat bismuthene grown on SiC, whose low en- ergy electronic properties are reasonably well described by an effective TB model Hamiltonian that includes only two orbitals ( pxandpy) per atom [14, 15, 17, 18]. It is a real solid state system, typical of a promising class of 2D materials based on the group group VA elements that exhibit relatively large energy gaps. In fact, a very good TB fit of both the valence and conduction bands of flat bismuthene can be obtained with the inclusion of second n.n. hopping integrals, as Fig. SX of the SM illustrates. Results for the associated SHE and OHE cal- culated as functions of EFfor planar bismuthene em- ploying a Chebyshev polynomial expansion method are shown in Fig.4. We clearly see in this case that the spec- traarenotsymmetricwithrespecttothezeroenergyand the right-hand side gap disappear. Results for VAB= 0, depicted in Fig.4 (a), show that the remaining gaps are topological, displaying a quantised SHE, and significant OHE. For sufficiently large sublattice asymmetry, how- ever, the central gap ceases to be topological, exhibiting noSHE,asFig.4(b)illustrates. Notwithstanding, theor- bital Hall conductivity is appreciable within this energy range. This validates our original prediction that pure orbital angular momentum currents can be triggered by a longitudinally applied electric field in some normal in- sulators. In summary, we have performed detailed analyses of the spin and orbital Hall conductivities for a class of 2D5 systems, relating the corresponding OHE, SHE and or- bital textures. Our calculations show the existence of OHE in topological insulators, with values that exceed those obtained for the SHE. Remarkably, we also obtain OHE for normal insulating phases where the SHE is ab- sent and no edge states cross their energy gaps. We show that the OHE in these systems is associated with exotic momentum-space orbital textures that are caused by an intrinsic Dresselhaus-type of interaction. This is rather generalandshowthatcertain2Dinsulatingmaterialscan generateorbitalangularmomentumcurrentsthatmaybe useful for developing novel spin-orbitronic devices. We acknowledge CNPq/Brazil, FAPERJ/Brazil and INCT Nanocarbono for financial support, and NACAD/UFRJ for providing high-performance com- puting facilities. TGR acknowledges COMPETE2020, PORTUGAL2020, FEDER and the Portuguese Foun- dation for Science and Technology (FCT) through project POCI-01- 0145-FEDER-028114. TPC acknowl- edges São Paulo Research Foundation (FAPESP) grant 2019/17345-7. [1] B. A. Bernevig, T. L. Hughes, and S.-C. Zhang, Phys. Rev. Lett. 95, 066601 (2005), URL https://link.aps. org/doi/10.1103/PhysRevLett.95.066601 . [2] T. Tanaka, H. Kontani, M. Naito, T. Naito, D. S. Hirashima, K. Yamada, and J. Inoue, Phys. Rev. B 77, 165117 (2008), URL https://link.aps.org/doi/ 10.1103/PhysRevB.77.165117 . [3] H. Kontani, T. Tanaka, D. S. Hirashima, K. Yamada, and J. Inoue, Phys. Rev. Lett. 100, 096601 (2008), URL https://link.aps.org/doi/10.1103/PhysRevLett. 100.096601 . [4] H. Kontani, T. Tanaka, D. S. Hirashima, K. Yamada, and J. Inoue, Phys. Rev. Lett. 102, 016601 (2009), URL https://link.aps.org/doi/10.1103/PhysRevLett. 102.016601 . [5] D. Go, D. Jo, C. Kim, and H.-W. Lee, Phys. Rev. Lett. 121, 086602 (2018), URL https://link.aps.org/doi/ 10.1103/PhysRevLett.121.086602 . [6] D. Go, J.-P. Hanke, P. M. Buhl, F. Freimuth, G. Bihlmayer, H.-W. Lee, Y. Mokrousov, and S. Blügel, Scientific Reports 7(2017), URL https://doi.org/10. 1038/srep46742 . [7] A. Nalitov, G. Malpuech, H. Terças, and D. Solnyshkov, Physical Review Letters 114(2015), URL https://doi. org/10.1103/physrevlett.114.026803 . [8] F. C. de Lima, G. J. Ferreira, and R. H. Miwa, Nano Letters (2019), URL https://doi.org/10.1021/acs. nanolett.9b02802 . [9] Y. Chen, W. Ruan, M. Wu, S. Tang, H. Ryu, H.-Z. Tsai, R. Lee, S. Kahn, F. Liou, C. Jia, et al., Visualizing exotic orbital texture in the single-layer mott insulator 1t-tase2 (2019), arXiv:1904.11010. [10] J. Kim, K.-W. Kim, D. Shin, S.-H. Lee, J. Sinova, N. Park, and H. Jin, Nature Communications 10(2019), URL https://doi.org/10.1038/s41467-019-11964-6 .[11] M. Tischer, O. Hjortstam, D. Arvanitis, J. Hunter Dunn, F. May, K. Baberschke, J. Trygg, J. M. Wills, B. Johansson, and O. Eriksson, Phys. Rev. Lett. 75, 1602 (1995), URL https://link.aps.org/doi/10. 1103/PhysRevLett.75.1602 . [12] I. V. Tokatly, Phys. Rev. B 82, 161404 (2010), URL https://link.aps.org/doi/10.1103/PhysRevB. 82.161404 . [13] V. o. T. Phong, Z. Addison, S. Ahn, H. Min, R. Agarwal, and E. J. Mele, Phys. Rev. Lett. 123, 236403 (2019), URL https://link.aps.org/doi/10. 1103/PhysRevLett.123.236403 . [14] F. Reis, G. Li, L. Dudy, M. Bauernfeind, S. Glass, W. Hanke, R. Thomale, J. Schäfer, and R. Claessen, Science 357, 287 (2017), ISSN 0036-8075, URL http: //science.sciencemag.org/content/357/6348/287 . [15] Y. Shao, Z.-L. Liu, C. Cheng, X. Wu, H. Liu, C. Liu, J.-O. Wang, S.-Y. Zhu, Y.-Q. Wang, D.-X. Shi, et al., Nano Letters 18, 2133 (2018), pMID: 29457727, https://doi.org/10.1021/acs.nanolett.8b00429, URL https://doi.org/10.1021/acs.nanolett.8b00429 . [16] J. C. Slater and G. F. Koster, Phys. Rev. 94, 1498 (1954), URL https://link.aps.org/doi/10. 1103/PhysRev.94.1498 . [17] G. Li, W. Hanke, E. M. Hankiewicz, F. Reis, J. Schäfer, R. Claessen, C. Wu, and R. Thomale, Phys. Rev. B 98, 165146 (2018), URL https://link.aps.org/doi/ 10.1103/PhysRevB.98.165146 . [18] T. Zhou, J. Zhang, H. Jiang, I. Žutić, and Z. Yang, npj Quantum Materials 3, 39 (2018). [19] C. Wu, D. Bergman, L. Balents, and S. Das Sarma, Phys. Rev. Lett. 99, 070401 (2007), URL https://link.aps. org/doi/10.1103/PhysRevLett.99.070401 . [20] C. Wu and S. Das Sarma, Phys. Rev. B 77, 235107 (2008), URL https://link.aps.org/doi/10. 1103/PhysRevB.77.235107 . [21] G.-F. Zhang, Y. Li, and C. Wu, Phys. Rev. B 90, 075114 (2014), URL https://link.aps.org/doi/10. 1103/PhysRevB.90.075114 . [22] L. M. Canonico, T. G. Rappoport, and R. B. Muniz, Phys. Rev. Lett. 122, 196601 (2019), URL https:// link.aps.org/doi/10.1103/PhysRevLett.122.196601 . [23] G. D. Mahan, Many-particle physics (Springer Science & Business Media, 2013). [24] G.Dresselhaus, Phys.Rev. 100, 580(1955), URL https: //link.aps.org/doi/10.1103/PhysRev.100.580 . [25] P. Streda, Journal of Physics C: Solid State Physics 15, L1299 (1982), URL https://doi.org/10.1088% 2F0022-3719%2F15%2F36%2F006 . [26] M. Milletarì, M. Offidani, A. Ferreira, and R. Raimondi, Phys. Rev. Lett. 119, 246801 (2017), URL https:// link.aps.org/doi/10.1103/PhysRevLett.119.246801 . [27] See Supplemental Material, which includes Refs. [28–41] for additional analyses, and details on the numerical simulations. [28] H. Röder, R. N. Silver, D. A. Drabold, and J. J. Dong, Phys.Rev.B 55, 15382(1997), URL https://link.aps. org/doi/10.1103/PhysRevB.55.15382 . [29] A. Weiße, G. Wellein, A. Alvermann, and H. Fehske, Rev. Mod. Phys. 78, 275 (2006), URL https://link. aps.org/doi/10.1103/RevModPhys.78.275 . [30] A. Ferreira, J. Viana-Gomes, J. Nilsson, E. R. Mucci- olo, N. M. R. Peres, and A. H. Castro Neto, Phys. Rev.6 B83, 165402 (2011), URL https://link.aps.org/doi/ 10.1103/PhysRevB.83.165402 . [31] Z. Fan, A. Uppstu, and A. Harju, Phys. Rev. B 89, 245422 (2014), URL https://link.aps.org/doi/10. 1103/PhysRevB.89.245422 . [32] N. Leconte, A. Ferreira, and J. Jung, in 2D Materials , edited by F. Iacopi, J. J. Boeckl, and C. Jagadish (El- sevier, 2016), vol. 95 of Semiconductors and Semimet- als, pp. 35 – 99, URL http://www.sciencedirect.com/ science/article/pii/S0080878416300047 . [33] J. H. Garcia, A. W. Cummings, and S. Roche, Nano Letters 17, 5078 (2017), pMID: 28715194, https://doi.org/10.1021/acs.nanolett.7b02364, URL https://doi.org/10.1021/acs.nanolett.7b02364 . [34] L. M. Canonico, J. H. García, T. G. Rappoport, A. Ferreira, and R. B. Muniz, Phys. Rev. B 98, 085409 (2018), URL https://link.aps.org/doi/10. 1103/PhysRevB.98.085409 . [35] M.Anđelković, S.Milovanović, L.Covaci, andF.Peeters, arXiv preprint arXiv:1910.00345 (2019). [36] J.H.García, L.Covaci, andT.G.Rappoport, Phys.Rev. Lett.114, 116602 (2015), URL https://link.aps.org/ doi/10.1103/PhysRevLett.114.116602 . [37] J. H. Garcia and T. G. Rappoport, 2D Materi- als3, 024007 (2016), URL http://stacks.iop.org/ 2053-1583/3/i=2/a=024007 . [38] A. Bastin, C. Lewiner, O. Betbeder-matibet, and P. Nozieres, Journal of Physics and Chemistry of Solids 32, 1811 (1971), ISSN 0022-3697, URL http://www.sciencedirect.com/science/article/ pii/S0022369771801476 . [39] M. Anđelković, S. M. João, L. Covaci, T. G. Rappoport, J. M. V. P. Lopes, and A. Ferreira (In Preparation), URL http://quantum-kite.com . [40] J. Sinova, D. Culcer, Q. Niu, N. A. Sinitsyn, T. Jung- wirth, and A. H. MacDonald, Phys. Rev. Lett. 92, 126603 (2004), URL https://link.aps.org/doi/10. 1103/PhysRevLett.92.126603 . [41] F. Dominguez, B. Scharf, G. Li, J. Schäfer, R. Claessen, W. Hanke, R. Thomale, and E. M. Hankiewicz, Phys. Rev. B 98, 161407 (2018), URL https://link.aps.org/ doi/10.1103/PhysRevB.98.161407 .1 Supplementary material for “Spin and Charge Transport of Multi-Orbital Quantum Spin Hall Insulators” HAMILTONIAN IN RECIPROCAL SPACE AND ORBITAL ANGULAR MOMENTUM OPERATORS DEFINITION The Hamiltonian given by Eq. (1) of the main text can be rewritten in the reciprocal space. Using the basis f p+;A ; p;A ; p+;B ; p;B g, the hopping term reads, ~H0(~k) =0 BBB@0 0 ~A(~k)~D(~k) 0 0 ~C(~k)~B(~k) ~A(~k)~C(~k) 0 0 ~D(~k)~B(~k) 0 01 CCCA; (S1) where 2~A(~k) =txx(~k) +tyy(~k) +i txy(~k)tyx(~k) ; (S2) 2~B(~k) =txx(~k) +tyy(~k) +i tyx(~k)txy(~k) ; (S3) 2~C(~k) =txx(~k)tyy(~k) +i txy(~k) +tyx(~k) ; (S4) 2~D(~k) =txx(~k)tyy(~k)i txy(~k) +tyx(~k) ; (S5) The transfer integrals t (~k)are Fourier transforms of Slater-Koster coefficients in the honeycomb lattice, t ; (~k) =2X m=0 p ;A ^V p ;B (m)ei~k~ am; (S6) where,mruns over the three nearest-neighbours of a site in sublattice A, that are located in sublattice B, and ~ amis the vector connecting the atom at sublattice A with its mth neighbor at sublattice B. The Slater-Koster integrals are given by t (~k) =n2 (m)Vpp+ (1n2 (m))Vpp; (S7) t =n (m)n (m) VppVpp ; (S8) withn ; (m)being the direction cosine connecting the site of sublattice Awithm-th first neighbor on sublat- ticeB. In all results presented in this work, we set Vpp= 0, unless it is mentioned. This condition can be relaxed without changing any of the main conclusions of our work. In this basis, the SOC term is diagonal, Hs soc=sIdiag(1;1;1;1)and the sublattice potential is given by HAB=VABdiag(1;1;1;1). As mentioned on the main text, due the absence of the pzorbital, the electronic states are restricted to the subspace associated with m`=1only, hence the angular momentum operators can be redefined in terms of a SU(2)algebra as: lz= p+ p+ p p ; lx= p+ p + p p+ ; ly=i p p+ p+ p  (S9) KUBO FORMULA FOR LINEAR RESPONSE CONDUCTIVITY Inthemaintext,wecomputethespinandorbitalHallconductivityfordifferenttopologicalphasesotheHamiltonian of Eq. (1) in the main text. For the cases of the pristine system, we used Kubo formalism to compute both OHE and2 SHE. In this formalism, the spin Hall (SH) and orbital Hall (OH) -polarized response, in ^ydirection, to an electric field applied in the ^xdirection is given by,  OH(SH)=e ~X n6=mX s=";#Z B:Z:d2k (2)2(fm~kfn~k) X n;m;~k;s; (S10) X n;m;~k;s=~2Im" s n;~k jXy(~k) s m;~k s m;~k vx(~k) s n;~k (Es n;~kEs m;~k+i0+)2# (S11) Were OH(SH)istheorbitalHall(SpinHall)DCconductivitywithpolarizationin -direction, X n;m;~k;sistherelated gauge-invariant Berry curvature. In Eq. S11, Es n(m);~kandj s n(m);~k are eigenvalues and eigenvectors of Hamiltonian of Eq. (S1), for n(m)Bloch band, with n;m = 1;::;4(in crescent order of energy), and s=";#spin-sector. Velocity operators are defined by, vx(y)(~k) =@H(~k)=@kx(y), whereH(~k)is the tight-binding Hamiltonian in reciprocal space. The current operator in ^ydirection is defined by jXy(~k) = Xvy(~k) +vy(~k)X =2, whereX=^`(^s)for OH (SH) conductivities polarized in direction. As it was discussed in the main text, this model presents a non-vanishing z OH, even in absence of SOC, in contrast to the spin Hall ( z SH) response which depends on the presence of SOC or exchange interaction. Added to this, in the presence of an exchange term, it was shown that the model presents a non-vanishing x OHassociated with in-plane polarized orbital Hall effect. It is important to mention that equations S10 and S11 are valid only in the clean limit and do not take into account the effect of disorder. However, as was briefly pointed in the main text, the effect of disorder should not affect our results for insulating phases (Fermi energy inside an electronic gap) due to the absence of the Fermi surface, responsible to generates the leading-order contribution in the computation of vertex corrections [26]. We confirm the robustness of our results against Anderson disorder using a real-space computation method which is discussed in the next sections of the SM. ANALYSIS OF THE BAND STRUCTURE We have examined the orbital Hall conductivity properties of three distinct topological phases displayed by the Hamiltonian Hdefined by Eq. (1) in the main text. They are labelled as B1, A1, and B2 phases, according to the classifications used in Ref. 21. Figure SI shows the "-spin electron energy bands for the system in these three phases. The spin-#bands can be deduced by applying a time-reversal symmetry operation on H. Panel (a) illustrates the band structure of the B1 phase, calculated for I= 0:2Vpp, andVAB= 0. We notice that the SOC causes three energy gaps to open, one originating from the K(K0) points, and the other two at , while the flat bands acquire a slight energy dispersion. Panel (b) shows the energy bands for the system in the A1 phase, calculated with I= 0:2Vpp, and VAB= 0:8Vpp. The sub-lattice potential affects each valley differently, as expected, because it breaks the degeneracy between eigenvalues at the KandK0symmetry points. By examining the opposite spin polarisation one finds that this phase exhibits a strong spin-valley locking, as discussed in Refs. 18, 21, 22. Panel (c) displays the energy bands for the system in the B2 phase, calculated with I= 1:1VppandVAB= 0:8Vpp. In this case, Iis comparable but slightly larger than VAB, and we note that they lead to effects that are similar to those exhibited panel (b), including a strong spin-valley locking with valley polarisation stronger than in the previous case due to the relatively large values ofIandVAB. EVOLUTION OF THE ORBITAL HALL EFFECT PLATEAUX In the main text, it was mentioned that the height of the orbital Hall plateaux within the lateral gaps depends upon the SOC coupling constant and the sub-lattice resolved potential. To demonstrate this, we show in Figure SII results for the spin and orbital Hall conductivities calculated for different sets of parameters for the B1, A1, and B2 phases. The results depicted in each panel of Figure SII are obtained for a fixed value of VABand two different values of I that are represented in the left and right columns, respectively. In panel (a) we show the conductivities calculated forVAB=0;I= 0:2VppandI= 1:0Vpp, which correspond to situations in which the system is in the B1 phase. It is clear that the height of the OHE plateau decreases as the SOC increases. In fact, the height of the plateau scales3 Figure SI:"-spin electron energy bands calculated as functions of wave vectors along some symmetry directions in the 2D Brillouin zone for three distinct topological phases: (a) B1 with I= 0:2VppandVAB= 0. (b) A1 with I= 0:2Vppand VAB= 0:8Vpp(c) B2 with I= 1:1VppandVAB= 0:8Vpp. with the size of the lateral gap, being close to the maximum value of the metallic limit for very small gaps. The same trend is observed in the other two phases, in contrast with the heights of the spin Hall plateaux that remain the same in all cases . ZIGZAG NANO-RIBBONS SPECTRA Thorough the main text we analysed the spin and orbital Hall effects for the three distinct topological phases B1;A1andB2. Tofurthersubstantiateourfindingsofthenon-zeroorbitalHallconductivityinthetriviallyinsulating phases, we analysed the energy spectrum of a zigzag nano-ribbon in our system for the three distinct phases. Figure SIII shows the spectra for each of the phases studied in the main text. As expected, the number of pairs of edge states corresponds with the index Z2of each of these phases. Panel (a) shows the energy bands corresponding to the phaseB1, here the most interesting features are the pairs of edge channels that cross the gap and the fully symmetric spectrum for both spin polarizations. Panel (b) displays the spectrum of the A1phase, here we can see the strong spin-valley locking that results from the inversion symmetry breaking produced by staggered sub-lattice potential. Interestingly here we can observe the absence of edge states traversing the central gap. Finally panel (c) shows the band structure of a ribbon in the phase B2. Here we can see how due the strong spin-orbit coupling and staggered sub-lattice potential the edge states in both the lateral gaps do not cross the gap while the edge estates of the central gap are crossing again. The results are fully consistent with the spin Cher number characterization of these phases done in Ref [21]. The results of panel (c) are the most striking ones, because they indicate that differently from the spin Hall conductivity, the orbital Hall effect plateau does not require electronic conducting channels to have a constant non-quantized value. latex onecolumn undefined CHEBYSHEV POLYNOMIAL EXPANSION To study the transport and spectral properties of the honeycomb lattice with pxpyorbitals we used the Chebyshev polynomial expansion. In this numerical method, the Green and spectral functions are accurately expanded in terms of Chebyshev polynomial of first kind of the Hamiltonian matrices[28, 29]. This set of polynomials are commonly chosen due their unique convergence properties, their relation with the Fourier transform and their convenient recurrence relations that allows the iterative construction of higher order polynomials[28, 29]. In recent years this method has gained much attention in the study of the transport properties of 2D systems[30–34]. Because of its high scalability, It was used to study of topological phase transitions induced by disorder[22], and more recently, in the analysis of the electronic properties of graphene encapsulated between two twisted hBN structures[35]. The method requires a rescaling of the Hamiltonian and it spectrum to make them fit into the interval (1;1)where4 Figure SII: Spin Hall conductivity z SH(red) and orbital Hall conductivity z OH(blue)calculated for: (a) VAB= 0and I= 0:2Vpp(left) andI= 1:0Vpp(right). (b) VAB= 0:8VppandI= 0:2Vpp(left) andI= 0:5Vpp(right). (c) VAB= 0:8Vpp,I= 1:1Vpp(left) andI= 1:5Vpp(right) the Chebyshev polynomials are defined and consequently the convergence of the method is assured. This scaling is achieved by means of the transformations ~H= (Hb)=aand ~E= (Eb)=awherea(ETEB)=(2)and b(ET+EB)=2. In the later ETandEBrepresents the top and bottom limits of the spectrum, respectively, and  is a small cut-off parameter introduced to avoid numerical instabilities. With this later conditions fulfilled, the Chebyshev polynomial expansion of the density operator considering N polynomials can be written as:5 Figure SIII: Zigzag Nano-ribbons spectra for phases (a) B1withI= 0:2VppandVAB= 0, (b) A1 with I= 0:2Vppand VAB= 0:8Vpp, and (c) B2 with I= 1:1VppandVAB= 0:8Vpp.  ~E =1 p 1~E2N1X m=0gmmTm ~E ; (S12) wheregmis a kernel introduced to control the Gibbs oscillations produced by the sudden truncation of the series expansion[28, 29]. The coefficients are calculated with m=hTrTm ~H i, in whichh:::irepresents the average over different disorder configurations. The calculation of the density operator of a given system is reduced to the computation of the trace of a matrix. To further decrease the computational cost of the calculation of quantities such as the density operator, instead of calculating the full trace of the polynomial matrices[29], we simply approximate the expansion coefficient mas m1 RhRX r=1hrjTm~~H jrii (S13) In the laterjrirepresent a set of random vectors which are defined as jri=D1=2PD i=1eiijii. Herefjiigi=1;:::;D denotes the original basis set, in which orbitals and spins on the lattice sites are treated equivalently, Drepresents the dimension of the Hamiltonian matrix, and iis the phase of each of the state vectors that comprise each of the random vectors. Ris the number of random vectors used in the trace estimation and the convergence of the later goes as 1=p DR. CHEBYSHEV POLYNOMIAL EXPANSION OF KUBO FORMULA To compute the spin and orbital conductivities of disordered systems, we employed the efficient algorithm developed by J. García et. al.[36, 37], which is based in the Chebyshev expansion of the Kubo-Bastin formula[38]:  (;T) =i~ Z+1 1dEf(E;;T) Trhj (EH)j dG+ dEj dG dEj (EH)i; (S14)6 in which represents the area of the 2Dsample,f(E;;T)is the Fermi-Dirac distribution for the energy E, chemical potential and temperature T.G+(G)symbolise the advanced(retarded) one electron Green function. As it can be seen from (S14) the Kubo-Bastin formula is expressed as a current-current correlation function. Then, to adapt this formula to calculate the spin hall conductivity z SH, we define j as the current-density operator like j jx=ie ~[x;H]andj as the spin current-density as j js y=1 2fz;vygwherezis the usual Pauli’s matrix and vyis they-Component of the velocity operator. For the computation of the orbital Hall conductivity, again z OHwe define the current operator j asj jx=ie ~[x;H]and we write j as the orbital current density operator, which is defined like j js y=1 2f`z;vygwhere`zis thez. It is noteworthy to mention that for the spin hall conductivity calculations we used the open-source code from the KITE project[39]. NUMERICAL SIMULATION OF THE DISORDERED CASE It is instructive to investigate how disorder affects the OHE in these two-dimensional systems and more specifically, how it modifies the plateaux in the orbital Hall conductivity that, as discussed before, is not dominated by conducting edge states. For this purpose we include in our Hamiltonian an on-site Anderson disorder term iwhose values are randomly picked from an uniform distribution that goes from W 2;W 2 , in whichWrepresents the Anderson disorder strength and then proceeded with the aforementioned Chebyshev polynomial expansions to compute the density of states (DOS), and the transverse components of the spin and orbital conductivity tensors. In these calculations we have considered systems of 8256256orbitals, Chebyshev polynomials up to the order M= 1280and we averaged overR= 150random vectors. It is noteworthy to mention that due the large number lattice sites that we are considering, we restricted ourselves to only one disorder realization, this is based on the assumption that almost every possible configuration is contained on our system due it large size.7 Figure SIV: Spin (red) and orbital (blue) Hall conductivities calculated as functions of energy for: (a) I= 0:2VppVAB= 0, (b)I= 0:2VppVAB= 0:8Vpp, and (c)I= 1:1VppVAB= 0:8Vppin the presence of disorder. The left, central and right panels show the results obtained in the relatively weak ( W= 0:05Vpp), intermediate ( W= 0:2Vpp) and strong ( W= 0:4Vpp) disorder regimes, respectively. The grey lines represent the density of states calculated for the same set of parameters. Figure SIV shows the spin and orbital Hall conductivities calculated for both weak ( W= 0:05Vpp), intermediate (W= 0:2Vpp) and strong ( W= 0:4Vpp) disorder. Similarly to what was previously observed for the SHE [34], the orbital Hall plateau remains present, even for a relatively strong disorder that closes the lateral gaps. Our preliminary results indicate that the orbital Hall effect in two-dimensional insulators is robust against Anderson disorder. ORBITAL TEXTURE ANALYSIS In contrast with the SHE, our calculations show that the OHE is not quantised, and occurs even in the absence of metallic edge states. In order to explore the origin of the OHE in this model system, we investigated the characteristics ofitsorbitalangularmomentumin reciprocalspace withinthe2DfirstBZ.Tothisend, wecomputetheorbitaltexture in reciprocal space defined as, ~Ls n;~k= `x s n;~k^x+ `y s n;~k^y+ `z s n;~k^z; (S15)8 Where, `x;y;z s n;~k= s n;~k `x;y;z s n;~k is the expected value of angular-momentum operator in reciprocal space for states of Bloch band nand spin sector s. To study the orbital texture and how it affects the OHE, we separate the in-plane textures ( `x;y s n;~k), which are represented by arrows, see Fig. SV and panels of Fig.2 of main text, and out-of-plane textures ( `z s n;~k), which we represent as a color plot (dark blue color for `z 1and dark red color for `z 1). Following the semi-classical argument of Ref. [40], it is possible to show that z OHis a consequence of the existence of non-trivial in-plane orbital texture ( `x;y s n;~k). Some features of the function z OH(Ef)can be understood from these textures, as we briefly mentioned in the main text, and now we detailed here. Figure SV: Orbital character of the "-spin eigenstates of H[Eq. S1 or Eq. (1) of the main text] calculated for: (a) I= 0:2Vpp, andVAB= 0; (b)I= 0:2Vpp, andVAB= 0:8Vpp; (c)I= 1:1Vpp, andVAB= 0:8Vpp. Figure 2 of the main text displays both the in-plane and the out-of-plane orbital polarisations of the lowest "-spin energy band for the B1, A1 and B2 phases. Results for the #-spin bands can be easily obtained by time-reversal symmetry operation. In Figure SV we complement our analysis by showing the orbital textures of the four "-spin energy bands for each one of the three phases. The orbital projections depicted in panel (a) were calculated for I= 0:2VppandVAB= 0, and correspond to the case in which the system assumes the B1 phase. Clearly, the in-plane orbital textures of the first and second energy bands are opposite to each other, and the same happens to the third and fourth bands, which leads z OH(Ef)to be an odd function of Fermi energy, and consequently, the absence of OHE in the central gap. As it was shown in Fig. 3 of the main text, if we include second neighbors hopping in the tight-binding Hamiltonian, the particle-hole symmetry around the central gap is broken, and the cancelation of in-plane orbital texture is lost, leading to the appearance of a central plateau in the orbital Hall conductivity. It is also noteworthy that h`zi" n;~kfor the second and third bands are opposite, as well as around the K(K0) and symmetry points. Conversely, the first and fourth bands respectively exhibit h`zi" n;~k1in the vicinities of the 9 point, but virtually vanishing values around KandK0. Panel (b) displays the orbital projections of the eigenstates corresponding to the A1 phase, calculated for I= 0:2VppandVAB= 0:8Vpp. One of the main eye-catching characteristics of this phase is the opposed out-of-plane orbital polarisations around the K0andKpoints, which is a manifestation of the orbital-valley locking produced by VAB. Similarly to phase B1, the out-of-plane polarisations of the first and second "-spin energy bands are opposed to the fourth and third ones, respectively. In addition, the in-plane orbital angular momentum polarisations for this phase exhibit the same configuration as those obtained for the B1 phase,which means that, also in this phase, sigma is an odd function of Fermi energy, with no central plateau. However, due to the orbital-valley locking, the corresponding absolute values are smaller, which explains the different curve derivative of the OHE in the phase A1 when compared with the OHE of the phase B1. Finally, panel (c) shows the orbital character of the system, calculated for I= 1:1VppandVAB= 0:8Vpp, when it is in the B2 phase. In this case we find that h`zi" n;~k1for the lowest energy band, which goes along with a substantial reduction of the in-plane texture. Similarly to the previous cases, h`zi" n;~kfor the lowest and highest energy bands are inverted. However, there a noticeable change in h`zi" n;~kin comparison with the results obtained for the A1 phase, which is accompanied by a relatively strong orbital-valley locking produced by the combined action of the large values of I andVAB. LOW-ENERGY APPROXIMATION As discussed in the main text, our effective Hamiltonian in the vicinity of the K=K0point can be expressed in terms ofSU(2) SU(2)orbital and sub-lattice algebras. Expanding the matrix of Eq. S1 near valleys K= 4=3a andK0=4=3a, we obtain, up to first order in electronic momentum, the effective theory Heff=~vF(kxx+kyy) +sI`z+VABz+H`: (S16) Here,vF=ap 3 2~Vpprepresents the Fermi velocity, and ais the lattice constant; =1for theKandK0valleys, respectively, and s=1for"and#spin electrons, respectively. The last term H`breaks the degeneracy between `z eigenstates and can be separated in two contributions: H`=H`k+HD;whereH`k=~vF 4(k+`++k`)andHD=p 3~vF 2a(`xx+`yy):(S17) =x+iy,=,` ( =x;y) represent the orbital angular momentum matrices in the corresponding Hilbert space,k=kxiky, and`=`xi`y. Figure SVI shows a comparison between the energy band spectra obtained by our tight-binding (blue dashed lines) and effective models (red solid lines) calculations in the vicinities of KandK0. In the left column we notice for the three phases that our effective linear model describes rather well the two inner energy bands, but fails to properly do so for the two outer ones. This can be corrected with the inclusion of quadratic terms in our approximation, as illustrated in the right column of Figure SVI. It is noteworthy that the orbital texture near KandK0are very well described by our effective model. Nevertheless, to reproduce the orbital texture in the vicinity of , it is necessary to perform an even higher-order expansion up to 4th order. To provide insight on how H`affects the energy spectrum and orbital textures of this model, we examine the corresponding contributions of each term in Eq. S17. For simplicity, we consider only one spin sector. In this case, the energy spectrum of H0=~vF(kxx+kyy)consists of two degenerate Dirac cones that are associated with the two eigenstates of the angular momentum pseudo-spinor. Similarly to what occurs in graphene, the inclusion of HAB=VABzopens an energy gap in the spectrum, while HSOC=sI`zacts as an orbital exchange interaction, shifting upwards (downwards) the Dirac cone associated with the `zeigenvalue +1(-1). To understand how H` modifies the energy spectrum, we introduce a multiplicative factor that regulates its overall intensity and inspect the energy band structure of H0+H`for two different values of in the following situations: (i) H`K6= 0;HD= 0, (ii) H`K= 0;HD6= 0, and (iii)H`K6= 0;HD6= 0. The results for the energy bands calculated as functions of kxfor ky= 0are exhibited in Figure SVII. In panels (a) and (d) we note that H`Klifts the orbital degeneracy of the two Dirac cones for kx6= 0, by differently renormalising their corresponding Fermi velocities. Panels (b) and (d) show howHDaffects the energy bands. HDdoes not depend upon the wave vector ~k, and has the same functional form10 Figure SVI: Comparison between the tight-binding energy band calculations (blue dashed lines) with the eigenvalues of our effective Hamiltonians in the vicinities of KandK0(red solid lines). The eigenvalues obtained with the linear and quadratic order expansions are depicted in the left and right panels, respectively. The results are for: (a) I= 0:2VppVAB= 0, (b) I= 0:2VppVAB= 0:8Vpp, and (c)I= 1:1VppVAB= 0:8Vpp. of a Dresselhaus SOC for Dirac Fermions. It may be regarded as equivalent to a Dresselhaus SOC for orbital states. As expected, HDleads to a Dresselhaus-like band splitting, without opening a gap at E= 0. In panels (c) and (f) of Figure SVII, we clearly see the formation of a single Dirac cone and the two outer bands when both H`Kand HDare present. It is worth recalling that to reproduce the flat-bands, it is necessary to consider high-order terms in k. Similarly to what is observed in quantum anomalous Hall insulators, the gap opening at E= 0is a consequence of the interplay between the orbital equivalent of a SOC and an exchange interaction. There is, however, a rich phenomenology involving the contributions of the distinct terms in Eq. S16 that arises when is varied, but this goes beyond the scope of the present discussion. Finally, we examine the role of H`andHDin the orbital texture of this model system. Figure SVIII shows the orbital textures calculated for: (a) HD6= 0andH`K= 0; (b) forHD= 0andH`K6=, and (c) for the effective complete Hamiltonian without SOC and VAB. By comparing the three panels, it is clear that the orbital texture of our effective model is basically governed by the Dresselhaus-like coupling associated with the orbital angular momentum spinor, which reproduces rather well the in-plane texture of our tight-binding calculations near K.11 Figure SVII: Energy bands calculated as functions of kx(forky= 0) by means of our effective theory around the Ksymmetry point. Panels (a), (b) and (c) depict the results obtained for = 0:3in the cases: H`K6= 0;HD= 0;H`K= 0;HD6= 0and H`K6= 0;HD6= 0, respectively. Panels (d), (e), and (f) show the results calculated for the same cases, but with = 1:0. Figure SVIII: Comparison between of the in-plane texture profile for: (a) HD6= 0andH`K= 0; (b)HD= 0andH`K6= 0, and (c)HD6= 0andH`K6= 0. SECOND NEAREST NEIGHBOURS AND ORBITAL TEXTURE ANALYSIS As mentioned in the main text, the absence of OHE plateau in the central electronic spectrum gap of px-py-model of Eq. (1) is a consequence of the combination of the particle-hole and parity symmetries of spectrum which translates in cancellation of in-plane orbital texture at half-filling. To understand better the consequences of the breaking of these symmetries, we introduce a toy-model of the px-pyHamiltonian, where we have included second nearest-neighbours hopping. This model is described by, H=X hijiX st ijpy ispjs+X hhijiiX st ijpy ispjs+X isipy ispis+X ishz spy ispis; (S18) here as before, iandjrepresents the honeycomb lattice sites whose position is given by ~Riand~Rj, respectively. The12 symbolshijiandhhijiiindicates that the summations are restricted to the nearest and second nearest neighbour sites respectively. The operator py iscreates an electron of spin sin the atomic orbitals p=p=1p 2(pxipy) centred at ~Ri. Here,s=";#labels the two electronic spin states, and, now, iis the atomic energy at site i, which encodes the effect of the combination of a sublattice potential VAB, and the on-site energy of porbitals"p. This terms take values i="pVAB, when site i belongs to the A and B sub-lattices of the honeycomb arrangement, respectively. Figure SIX: Comparison between the orbital (spin- ") texture profiles of the px-pywith only nearest neighbours (a), and the orbital texture of the same model when second nearest neighbours are considered (b). Left: Orbital Texture profile for the deepest energy band. Center: Orbital Texture profile for the second lowest energy band. Right: addition of the orbital textures in Left and Right panels with the in-plane component scaled by a factor 5. As it was shown in figure 3 of the main text, the principal effect of the particle-hole and parity symmetries breaking, a consequence of the inclusion of the second nearest neighbours, is the appearance of an orbital Hall conductivity plateau in the central gap. In order to uncover the connection between the appearance of this plateau and the orbital textures, we analyse the texture profiles of the two deepest energy bands for two different cases of this model. In panel (a) of the figure SIX are shown the orbital textures of the two deepest energy bands(left and central panels) of the simple model that does not include second nearest neighbours [Eq. (1) in the main text] and their summation (right panel) in which the in-plane components of the texture are in a larger scale to make easier the analysis of their details. The in-plane component of orbital textures in left and central panels present the dipole configuration around the point and the anti-vortices in the KandK0points, and the out-of-plane component appears due to the inversion symmetry breaking produced by the inter-lattice potential. At the right panel of the figure (a), we show that the addition of the orbital textures of the left and central panels results in a zero net in-plane orbital texture. Once that orbital Hall conductivity ( z OH) appears as a consequence of dynamics of in-plane orbital texture, in presence of an external electric field, this explains the absence of OHE in the central gap of the simplified px-pymodel of Eq. (1) in the main text. Now, in panel (b) of the figure SIX, we consider the orbital texture of the Hamiltonian with the inclusion of second nearest-neighbours hopping (see. Eq. S18). Again, the left and the central panels of the figure show the orbital textures of two deepest bands and the right panel shows the sum of these two textures, with the in-plane component multiplied by a scaling factor to facilitate its visualization. To maintain the resemblance between13 the aforementioned case and this new case, we set the same Slater-Koster parameters that we used for the phase A1 of the simplified model with the addition of "p=0:3andVpp2=0:2. With these parameters, as it was shown in figure 3 (a) of the main text the energy bands of this modified model are not particle-hole symmetric, an effect caused exclusively by the inclusion of second nearest-neighbours. We note in Fig. SIX (b) that the overall features of in-plane orbital-texture of two deepest bands (left and central panels) are not qualitatively modified, i.e., still present a dipole-like texture near -point and anti-vortices textures at valleys. However, as can be seen in Fig. SIX (b), right panel, the exact cancellation of the in-plane texture of two deepest bands is lost, causing the existence of a net in- plane orbital texture which produces an OHE in the central gap of the spectrum shown in figure 3 (b) of the main text. Once shown by means of the simply pxpymodel that the orbital Hall effect is present in systems where the particle-hole and parity symmetries are absent, we now focus on a real material. For this purpose, we have chosen the flat bismuthene grown on SiC as a candidate for the observation of OHE in the central plateau. The observation of orbital-insulator phase in the central gap of bismuthene should be easier in the experimental point of view, once it corresponds to neutrality situation. In the past, this system has been studied by means of the aforementioned minimal px-pytight-binding Hamiltonian [14, 22, 41]. However, we have noticed that by including second nearest-neighbours in the tight-binding Hamiltonian used in Ref. 22 the electronic structure is better reproduced. In the bismuthene/SiC heterostructure, the break of inversion symmetry induces a small Rashba SOC, HR= 2iRX hi;jiX spy is[^z(~ ^eij)]sspjs+H:c: (S19) where~ symbolises the Pauli vector, ^eijdenotes the unit vector along the n.n. inter-site direction of ~Rj~Ri,Ris the Rashba SOC constant, and sdesignates the opposite spin direction specified by s. We will consider this term only in the fitting of tight-binding Hamiltonian to DFT spectrum and, we neglect it in transport calculations presented here. The reasons is that the non-conserving spin character of this coupling complicates the analysis of orbital texture and its typical small value does not alter the main conclusions of our discussion, as it was checked by us. In figure SX is shown a direct comparison of the DFT energy band structure obtained in Ref. 14, with and without the inclusion of Rashba spin-orbit coupling. From this figure, it is rapidly noticeable the agreement between the DFT energy bands and the tight-binding model in describing the top of the valence band and the bottom of the conduction bands and the indirect gap in the point. In the table SI are shown the two-centre integral parameters used in the description of this model. Once shown the agreement of energy band of the complete model, we are going to restrict ourselves to the situation in which the Rashba SOC is neglected and the system is subject to a staggered potential VAB= 0:87. The first of these constraints is to avoid complications in the analysis due to possible contributions to the orbital texture by the Rashba SOC, which does not conserve spin as a good quantum number, and the second one is to leave the system in a topological phase similar to the phase A1of thepxpymodel with only nearest neighbours. This will allow us to focus on the analysis of the Orbital texture and its connection with the orbital Hall conductivity. Table SI: Second Nearest-neighbour two-centre energy integrals, and spin orbit coupling constants (all in eV) for the Bi/SiC. Two-centre integrals Intrinsic SOC Rashba SOC On-site energy Vpp= +1:51522I=0:435R= 0:032p=0:279865 Vpp=0:575788 Vpp2=0:18 Vpp2=0:00658 In figure SXI is displayed the orbital textures of the two deepest energy bands of bismuthene grown over SiC in the phaseA1and without Rashba SOC. The principal difference that is noticeable when one looks at this figure is the change of the out-of-plane orbital texture of the spin- "sector, with respect to textures of A1-phase in previous cases, produced by the change of sign of the spin-orbit coupling. The in-plane components of orbital texture for the model with parameters of bismuthene shown in Fig. SXI left and centre panels, do not present noticeable differences from those of Fig. SIX (b) of the model with second nearest neighbours. Again, as can be seen from the right panel of Fig. SXI, there is a non-zero total in-plane orbital texture when we add up the textures of two deepest bands of left and central panels what again explain the existence of OHE in the central gap, as shown in Figure 4 of the main text. This suggests the recent synthesized flat bismuthene as a realistic platform to observe the orbital Hall insulator phase. The central plateau will persist by the inclusion of the Rashba term [Eq. (S19)] on Hamiltonian,14 Figure SX: Comparison between the DFT energy bands (blue doted line) and the tight-binding model with second nearest neighbours bands (red solid line) for: (a) R= 0:032eV and (b) R= 0. once the spectrum keeps the particle-hole asymmetry. When the Rashba term is included, we cannot separate the textures by spin sectors because it breaks the sz-symmetry. So the previous analysis of sum of orbital texture must be done summing the four lowest energy bands to obtain total texture related to the central plateau. But the main conclusions are the same and we do not present this analysis here. Figure SXI: Orbital Textures profile of spin- "sector for the two lowest energy bands of bismuthene over SiC in which the Rashba SOC is neglected and VAB= 0:87eV. Left: Orbital texture profile of the lowest energy band. Center: Orbital texture profile of the second lowest energy band. Right: Addition of the later texture profiles. The resultant in-plane orbital texture are scaled by a factor 5to facilitate the visualization