diff --git "a/Low Fermi Level/paper/all_papers.json" "b/Low Fermi Level/paper/all_papers.json" new file mode 100644--- /dev/null +++ "b/Low Fermi Level/paper/all_papers.json" @@ -0,0 +1 @@ +[ { "title": "0901.1299v1.Fermi_surface_topology_and_low_lying_quasiparticle_structure_of_magnetically_ordered_Fe1_xTe.pdf", "content": "arXiv:0901.1299v1 [cond-mat.supr-con] 9 Jan 2009Fermi surface topology and low-lying quasiparticle struct ure of magnetically ordered\nFe1+xTe\nY. Xia,1D. Qian,1,2L. Wray,1,3D. Hsieh,1G.F. Chen,4J.L. Luo,4N.L. Wang,4and M.Z. Hasan1,5,6\n1Joseph Henry Laboratories of Physics, Department of Physic s, Princeton University, Princeton, NJ 08544\n2Department of Physics, Shanghai Jiao Tong University, Shan ghai 200030, China\n3Lawrence Berkeley National Laboratory, University of Cali fornia, Berkeley, CA 94305\n4Beijing National Laboratory for Condensed Matter Physics,\nInstitute of Physics, Chinese Academy of Sciences, Beijing , China\n5Princeton Center for Complex Materials, Princeton Univers ity, Princeton, NJ 08544\n6Princeton Institute for the Science and Technology of Mater ials, Princeton University, Princeton, NJ 08544∗\n(Dated: November 21, 2018)\nWe report the first photoemission study of Fe 1+xTe - the host compound of the newly discovered\niron-chalcogenidesuperconductors(maximumT c∼27K).Ourresultsrevealapairofnearlyelectron-\nhole compensated Fermi pockets, strong Fermi velocity reno rmalization and an absence of a spin-\ndensity-wave gap. A shadow hole pocket is observed at the ”X” -point of the Brillouin zone which is\nconsistent with a long-range ordered magneto-structural g roundstate. No signature of Fermi surface\nnesting instability associated with Q=( π/2,π/2) is observed. Our results collectively reveal that\nthe Fe 1+xTe series is dramatically different from the undoped phases o f the high T cpnictides and\nlikely harbor unusual mechanism for superconductivity and magnetic order.\nPACS numbers:\nThe discovery of superconductivity in the pnictides\n(FeAs-based compounds) has generated interest in un-\nderstandingthegeneralinterplayofquantummagnetism,\nelectronic structure and superconductivity in iron-based\nlayered compounds [1, 2]. The recent observation of un-\nusual superconductivity and magnetic order in the struc-\nturally simpler compounds such as FeSe and Fe 1+xTe are\nthe highlights of current research [3, 4, 5, 6, 7]. The ex-\npectation is that these compounds may provide a way\nto isolate the key ingredients for superconductivity and\nthe nature of the parent magnetically order state which\nmaypotentiallydifferentiatebetweenvastlydifferentthe-\noretical models [8, 9, 10]. The crystal structure of these\nsuperconductors comprises of a direct stacking of tetra-\nhedral FeTe layers along the c-axis and the layers are\nbonded by weak van der Waals coupling. Superconduc-\ntivity with transition temperature up to 15K is achieved\nin the Fe 1+x(Se,Te) series [3, 4] and T cincreases up to\n27K under a modest application of pressure [5]. Den-\nsity functional theories (DFT) predict that the electronic\nstructure is very similar to the iron-pnictides and mag-\nnetic order in FeTe originates from very strong Fermi\nsurface (FS) nesting leading to the largest SDW gap in\nthe series. Consequently, the doped FeTe compounds\nare expected to exhibit T cmuch higher than that ob-\nserved in the iron-pnictides if superconductivity would\nindeed be originating from the so called ( π,0) spin fluc-\ntuations [10]. These predictions critically base their ori-\ngin on the Fermi surfacetopologyand the band-structure\ndetails, however, no experimental results on the Fermi-\nology and band-structure exist on this sample class to\n∗Electronic address: mzhasan@Princeton.edu0 100 200 3000.0030.0060.009\nH//c\nH//a-bχ (emu/mol)\nT (K)SrFe As2 2 FeTe\n(a) (b) (c)Fe Te1+x\nab\nc\nFIG. 1: Magneto-structural transition and long-range orde r\nin Fe1+xTe: (a)Temperature dependence of the magnetic sus-\nceptibility [6]. (b) The spins in SrFe 2As2are collinear and\nQAFpoints along the ( π,0) direction. (c) The ordering vec-\ntor,QAFin FeTe, is rotated by 45◦and points along the\n(π\n2,π\n2) direction [15].\nthis date. Here we report the first angle-resolved photoe-\nmission (ARPES) study of the Fe 1+xTe - the host com-\npound of the superconductor series. Our results reveal a\npair of nearly electron-hole compensated Fermi pockets,\nstrong band renormalization and a remarkable absence\nof the predicted large SDW gap. Although the observed\nFermi surface topology is broadly consistent with the\nDFT calculations, no Fermi surface nesting instability\nassociated with the magnetic order vector was observed.\nAn additional hole pocket is observed in our data which\ncan be interpreted to be associated with a local-moment\nmagneto-structuralgroundstateinclearcontrasttointer-\nband nesting. Our measurements reported here collec-\ntively suggest that the FeTe compound series is dramati-\ncally different from the parent compound of the pnictide\nsuperconductorsand mayharbornovelformsofmagnetic\nand superconducting instabilities not present in the high\nTcpnictides.2\n0□meV -20□meV -40□meV -60□meV\n(a) (b) (c) (d)\nΓM\nXLDA Schematic\n(f)Γ\nMX( ,□0)π\n(□□□, )π\n2π\n2Γ M\n(e)ΓM\nEF\n(g)Band\nSchematic\nFIG. 2: Fermi surface topology : (a) Fermi surface topol-\nogy of FeTe. (a-d) Connection to the underlying band-\nstructure is revealed by considering the evolution of the de n-\nsity of states, n(k), as the chemical potential is rigidly lo w-\nered by 20, 40 and 60 meV. The Fermi surface consists of hole\npockets centered at Γ, and electron pockets centered at M. An\nadditional hole-like feature is observed at M in (b-d), whic h\nis attributed to a band lying below E F(see schematic in (e)).\n(e) presents a schematic of the band dispersions in the Γ-M\ndirection forming the observed electron and hole pockets. A\nqualitative agreement is observed between the (f) the exper -\nimentally measured FS topology and (g) the LDA calculated\n[10] FS although measured bands are strongly renormalized.\nSingle crystals of Fe 1+xTe were grown from a mixture\nofgroundedFe and Te powderusing the Bridgemantech-\nnique. The powder was heated to 920◦C in an evacuated\ntube thenslowlycooled, formingsinglecrystals. Theiron\nconcentrationwasmeasured by ICP (inductively-coupled\nplasma) technique and a value of x was determined to be\nless than 0.05. High-resolution ARPES measurements\nwere then performed using linearly-polarized 40eV pho-\ntons on beamline 10.0.1 of the Advanced Light Source\nat the Lawrence Berkeley National Laboratory. The en-\nergy and momentum resolution was 15meV and 2% of\nthe surface Brillouin Zone (BZ) using a Scienta analyzer.\nThe in-plane crystal orientation was determined by Laue\nx-raydiffractionpriortoinsertingintotheultra-highvac-\nuum measurement chamber. The magnetic order below\n65K was confirmed by DC susceptibility measurements\n(Fig-1). The samples were cleaved in situat 10K under\npressures of less than 5 ×10−11torr, resulting in shiny\nflat surfaces. Cleavage properties were characterized by\nSTM topography and by examining the optical reflection\nproperties.\nFigure 2 presents the momentum dependence of the\nphotoemission intensity n(k) from the sample at 10K at\nseveral different binding energies (0, 20, 40 and 60 meV)\nintegratedovera finite energywindow ( ±5 meV) at each\nbinding energy. The non-zero spectral intensity at the\nFermi level ( µ) confirmed that the low temperature state\nof Fe1+xTe is metallic-like. At the Fermi level, electrons\nare mostly distributed in one broad hole-like pocket at Γ\nand another similar size electron-like pocket around M.\nTo reveal the band shapes we present a gradual binding-energy evolution of the band features and present the\ndata in a way to simulate the effect of rigidly lowering\nof the chemical potential down to 60meV. The FS (at\n0 meV) is seen to consist of circular hole pockets cen-\nteredatΓ, andellipticalelectronpocketsattheMpoints.\nAn elliptical-like feature at the M point expands as the\nbinding energy is increased, suggesting that the associ-\nated band is hole-like. We will subsequently show that\nthis band lies below µand the M-point Fermi pockets\nare only electron-like ( 2(e)). By considering the domi-\nnant intensity patterns in panel-(a-d), apart from a weak\nfeature at ”X”-point, the FS topology is similar to that\nexpected from the DFT calculations ( 2(g)) [10]. In ad-\ndition, the area of the hole-like FS pocket at Γ and the\nelectron-like FS pocket at M are approximately equal in\nsize suggesting nearly equal number of electron and hole\ncarrier densities in this material. Therefore, if the ex-\ncess Fe atoms are contributing to the carrier density it\nis likely to be small and beyond our k-resolution of the\nexperiment.\nIn order to systematically study the low lying energy\nband structure, ARPES spectra are taken along different\nk-spacecut directionsin the 2DBrillouinzone(BZ). Two\ndifferent electron-photon scattering geometries are used\nto ensure that all bands are imaged. In a σscattering\ngeometry where the polarization vector is parallel to the\nkyaxis, photoelectron signal is predominantly from the\ndxy,dxz, anddz2energybands due to the dipole emission\nmatrix elements [11, 12]. Similarly, when the polariza-\ntion vector is parallel to kx, theπ-geometry, the dyzand\ndx2−y2states are predominantly excited. Figure 3 (a)\nand (b) present scans along the Γ −Mdirection in the\nσandπgeometries. In both sets of spectra, one finds\na broad hole-like band centered at Γ. However, near M\nthe scanstakenat twodifferent geometriesaredrastically\ndifferent. In cut 1, two hole-like bands ( α2,α3as marked\nin panel (a)) are observed to be approaching µ. Under\ntheπgeometry the band emission pattern near M is dra-\nmatically different (see the β1band in cut 2), while the\nα2andα3band signals are significantly weaker. The po-\nlarization dependence suggests that the β1band should\nhavedyzordx2−y2orbital character.\nIn order to fully resolve the broad band feature cen-\ntered at Γ, high resolution scans are performed along\ndifferent k-cut directions through the zone center. Fig-\nure 3(c) presents one cut in the Γ −Mdirection inside\nthe first zone, together with the corresponding energy\ndistribution curve (EDC). Two hole-like bands are re-\nsolved, labeled α2andα3. Since there are traces of mul-\ntiple bands near M, one might wonder whether the α3\nband crosses µnear the Mpoint, thus forming a hole-like\nFermi pocket. To systematically investigate this, a series\nof spectra are taken along the Γ −Mdirection with π-\ngeometry ( 3(d)-(f)) through multiple k-cuts intersecting\ntheM-point Fermi pocket. At M, cut-4 shows a strong\nelectron-like band forming the FS pocket. This band can\nbe attributed to the β1band (also observed in cut-2). As\none moves away from M, the band intensity becomes in-3\nFIG. 3: Low-lying band topology : ARPES spectra along the Γ −Mdirection in the (a) σand (b)π-scattering geometries.\nvFof the quasiparticle band forming the Γ hole pocket is calcul ated from the dashed line. (c) The EDC of a high momentum\nresolution scan through Γ shows two hole-like bands crossin g the E F. (d)-(f) Cuts along the M-Γ direction through the electron\nFS pocket show that the hole band near M remains at least 10meV below E F. For each scan direction the corresponding\nEDC is also presented, within the momentum range marked by gr een dashed lines. Comparison with (g) a band dispersion\nschematic of the LDA result [10] shows a good agreement betwe en experiment and theory. (h) The directions of the six scans\nare summarized and identified in the first BZ. The direction of the electric field is shown.\ncreasinglymore hole-like, indicating the emergenceofthe\nα2band. Nevertheless, the hole band lies completely and\nconsistently below the Fermi level, with some weak elec-\ntron quasiparticle intensity above the band maximum.\nTheβ1intensity becomes the weakest near the edge of\nthe M pocket (cut-6). But even at that location, the hole\npocket lies at least 10meV below the chemical potential.\nThe result shows that there are no hole pocket features\nin the FS near Mwhich supports the interpretation that\nthis material is nearly electron-hole compensated. Ad-\nditionally, one can map the observed bands to the band\nstructure estimated by DFT calculations (schematic in\n3(g) [10]). The calculation finds three bands ( α1-α3)\ncrossing Fermi level near Γ and two near ( β1,β2) M,\nforming the electron and hole pockets. While the DFT\ncalculated bands agree fairly well with our data, within\nour resolution or the sample surface roughness, we have\nnot succeeded in fully resolving the α1band near Γ. The\nβ2band, which forms the second M electron pocket, is\nobserved in the Fermi surface topology (Fig. 2). The\noverall band narrowing is about factor of 2 compared to\nthe DFT calculations highlighting the importance of cor-\nrelationeffects. This isalsoconsistentwith asmallFermi\nvelocity (v Fermi∼0.7 eV·˚A) observed (Fig.3(a)).\nWe now revisit the details of the FS map (Fig.-2) and\ndiscuss a weak feature observed at X=(π\n2,π\n2) between\ntwo Γ points (Fig. 4(e)). Such a feature is not expected\nfrom the DFT calculations. A similar feature, whose ori-\ngin is debated, has been observed in AFe 2As2(A=Ba,\nSr, Ca), which is attributed to a 2 ×1 surface reconstruc-\ntion of the Aatom layer [13, 14], in addition to a weak\nbulk structural distortion. However, in FeTe there are\nno additional atoms (such as Sr or Ba) between the Felayers and the crystal cleaves at a weak van der Waals\nbond bewteen two adjacent layers no strong 2 ×1 long-\nrange ordered A-type reconstruction is expected except\nfor a weak bulk-like structural orthorhombicity tied to\nthe magnetic order (magneto-structural effect). Recent\nneutron[15, 16] andx-raydiffraction[17, 18]studieshave\nshown that FeTe undergoes a bulk structural distortion\nfrom the tetragonal to weakly-monoclinic or orthorhom-\nbic phasenear65K,accompaniedbylong-rangemagnetic\norder Q AF=(π\n2,π\n2).\nIn the parent compound of the pnictide superconduc-\ntors such as the SrFe 2As2or BaFe 2As2, the SDW vector\nQSDW=(π, 0) coincides with a Fermi surface nesting of\nvector connecting the hole-pocket at Γ and the electron\npocket at M. Currently, it is believed that this nesting\nis responsible for opening a gap in the low temperature\nphysical properties [6, 11, 19]. In the case of FeTe, while\na nesting vector can indeed be drawn along QSDW=(π,\n0) between a pair of electron and hole pockets, all avail-\nableneutronscatteringmeasurementsreportthat thean-\ntiferromagnetic ordering vector is 45◦away from that in\nSrFe2As2, namely, in Fe 1+xTe,QAF=(π\n2,π\n2). The order-\ning shows a commensurate to incommensurate cross-over\nif the concentration of excess iron, x, is increased. An-\nother remarkable difference is that the magnetic suscep-\ntibility is Curie-Weiss like in FeTe suggesting that the\nmagnetism is of local-moment origin. Within a local\nmoment-like AFM long-range ordered state which also\ncouples to a weak structural distortion one should ex-\npect relatively intense shadow Fermi surfaces along the\nNeel vector QAF. The X-point Fermi surface we ob-\nserve thus can be related to the vector observed in neu-\ntron scattering X= Γ+QAF. The weak shadow-like X4\n-0.05 0.00\n0.0 0.5 1.0 1.5-0.3-0.2-0.10.0\nE (eV)BΓ1 X\nk(Å )-1SrFe As2 2 FeTe\nΓ1\nXMSrFe As2 2 FeTeFe Sr(a) (b)\n(c)\n2 1□Sr□Order/c180 No□Sr□□OrderM\nXΓ\nQSDW\nMXΓ\nQSDW?\n(e)QAFQAFE (eV)BGold\nM\nΓ\n(d)\n(f)Norm.□Intensity\nFIG. 4:Electronic structure and Magnetism : While (a)\nthe electron and hole pockets in SrFe 2As2can be kinemat-\nically nested by QAF=(π,0), there exists (b) no FS pocket\nwhich can be nested by QAF= (π\n2,π\n2) in the FeTe BZ. A\nFS pocket observed at X in SrFe 2As2has been attributed to\nthe (c) 2 ×1 surface order of the Sr atomic layer (red), which\noccurs in addition to a weak bulk distortion (blue) across th e\nSDW transition. However, there are no additional Sr atoms\nbetween the Fe layers in FeTe, so no Sr order is possible, al-\nthough a weak bulk distortion is not excluded. (d) The EDCs\nmeasured at near the Γ and M pockets exhibit no evidence of\nenergygaps. Nevertheless, (e)aweakFSpocketisobserveda t\nX,(f) correspondingtotwohole-like bandsdispersingtowa rds\nthe chemical potential which is consistent with a long-rang e\nlocal moment magneto-structural order.\npocket FS might therefore arise from a band folding due\nto long-range magnetic order. However, unlike SrFe 2Se2,\nwhere the Γ electron pocket nests with the M hole pock-\nets viaQSDW= (π,0) [20], an analogousnesting channel\nis unavailable at (π\n2,π\n2) in the FeTe ( 4(b)) clearly ruling\nout FS nesting as the origin of magnetic order. In the\nabsence of nesting FS gapping is not expected in FeTe\nwhich is consistent with our results in Fig-2 and 3. A\nlarge low temperature specific heat value [6] is thus con-sistent with our observation of a Fermi surface in the\ncorrelated magneto-structurally ordered state. To exam-\nine the band dispersion character of the X pocket, figure\n4 (f) presents a spectra along the Γ −Xdirection. Re-\nsults show two bands dispersing towards µ, forming the\nhole pocket in the ”X”-FS whose shapes are indeed very\nsimilar to the bands that form the central Γ-pocket FS.\nFurther evidence against nesting comes from the absence\nof a large gap at the overall low temperature electron\ndistributions presented in Fig-2 and 3 ( 4(d)). Our re-\nsults seem to suggest that the groundstate is a nearly\nelectron-hole compensated semimetal. Absence of a gap\nis consistent with recent bulk optical conductivity, spe-\ncific heat and Hall measurements in FeTe [6] (while most\nmeasurements do report a gap in SrFe 2As2[19]).\nA recent density functional calculation [21] suggests\nthat the excess Fe is in a valence state near Fe+and\ntherefore donates electron to the system. Due to the in-\nteraction of the magnetic moment of excess Fe with the\nitinerant electrons of FeTe layer a complex magnetic or-\ndering pattern is realized. In this scenario, excess Fe\nwould lead to an enlargement of the electron pocket FS,\nhowever, within our resolution electron and hole Fermi\nsurface pockets are measured to be of very similar in size\nsuggestingalackofsubstantialelectrondopingduetoex-\ncess Fe. Finally, we note that a gaplessyet long-rangeor-\nderedlocal-momentmagneto-structuralgroundstatecon-\nsistent with ARPES and neutron data taken together is\ncaptured in both first-principles electronic structure [8]\nandmany-bodyspinmodel[9]calculations. However,the\nbroad agreement of DFT calculations with experimental\nband-structure data except for about a factor 2 renor-\nmalization is also remarkable. A complete understand-\ning of electron correlation, moment localization and the\ntrue nature of the gapless antiferromagnetic state would\nrequire further systematic ARPES study [22].\nIn conclusion, we have presented the first ARPES\nstudy of Fermi surface topology and band structure of\nFe1+xTe. Our results reveal a pair of nearlyelectron-\nhole compensated Fermi pockets, strong renormalization\nand the remarkable absence of a spin-density-wave gap.\nThe observed shadow hole pocket is consistent with a\nlong-rangeorderedlocal moment-likemagneto-structural\ngroundstate whereas, most remarkably, no Fermi surface\nnesting instability associated with the antiferromagnetic\norderwasobserved. Contraryto band theorysuggestions\n[23], results collectively suggest that the Fe 1+xTe series\nis dramatically different from the undoped phases of the\nhigh T cpnictides and likely harbor a novel mechanism\nfor superconductivity and quantum magnetism.\n[1] Y. Kamihara, T. Watanabe, M. Hirano, and H. Hosono,\nJ. Am. Chem. Soc. 130, 3296 (2008); Z.A. Ren et al.,\nChin. Phys. Lett. 25, 2215 (2008).\n[2] G.F. Chen et al., Phys. Rev. Lett. 100, 247002 (2008);X. H. Chen et al., Nature 453, 761 (2008).\n[3] F.C. Hsu et al., Proc. Natl. Acad. Sci. 105, 14262 (2008).\n[4] M.H. Fang et al., Phys. Rev. B 78, 224503 (2008).\n[5] Y. Mizuguchi et al., Appl. Phys. Lett. 93, 152505 (2008).5\n[6] G.F. Chen et al., arXiv:0811.1489 (2008).\n[7] C. Xu and S. Sachdev, Nature Phys. 4, 898 (2008).\n[8] F.J. Ma et al., arXiv:0809.4732 (2008).\n[9] C. Fang et al., arXiv:0811.1294 (2008).\n[10] A. Subedi et al., Phys. Rev. B 78, 134514 (2008).\n[11] D. Hsieh et al., arXiv:0812.2289 (2008).\n[12] D. Qian et al., Phys. Rev. Lett. 97, 186405 (2006).\n[13] M.C. Boyer et al., arXiv:0806.4400 (2008).\n[14] Y. Yin et al., arXiv:0810.1048 (2008).\n[15] S. Li et al., arXiv:0811.0195 (2008).[16] W. Bao et al., arXiv:0809.2058 (2008).\n[17] K. Yeh et al., arXiv:0808.0474 (2008).\n[18] Y. Mizuguchi et al., arXiv:0810.5191 (2008).\n[19] W.Z. Hu et al., Phys. Rev. Lett. 101, 257005 (2008).\n[20] J. Zhao et al., Phys. Rev. B 78, 140504 (2008).\n[21] L. Zhang et al., arXiv:0810.3274 (2008).\n[22] Y. Xia, M.Z. Hasan et al.,.\n[23] M. Johannes, Physics 1, 28 (2008)" }, { "title": "1903.04107v1.Spontaneous_charge_current_in_a_doped_Weyl_semimetal.pdf", "content": "arXiv:1903.04107v1 [cond-mat.mes-hall] 11 Mar 2019Journal of the Physical Society of Japan\nSpontaneous Charge Current in a Doped Weyl Semimetal\nYositake Takane\nDepartment of Quantum Matter, Graduate School of Advanced S ciences of Matter,\nHiroshima University, Higashihiroshima, Hiroshima 739-8 530, Japan\n(Received )\nA Weyl semimetal hosts low-energy chiral surface states, wh ich appear to connect a pair of Weyl nodes\nin reciprocal space. As these chiral surface states propaga te in a given direction, a spontaneous circulating\ncurrent is expected to appear near the surface of a singly con nected Weyl semimetal. This possibility is\nexamined by using a simple model with particle-hole symmetr y. It is shown that no spontaneous charge\ncurrent appears when the Fermi level is located at the band ce nter. However, once the Fermi level deviates\nfrom the band center, a spontaneous charge current appears t o circulate around the surface of the system\nand its direction of flow is opposite for the cases of electron doping and hole doping. These features are\nqualitatively unchanged even in the absence of particle-ho le symmetry. The circulating charge current is\nshown to be robust against weak disorder.\n1. Introduction\nA Weyl semimetal possesses a pair of, or pairs of, non-\ndegenerate Dirac cones with opposite chirality.1–10)The\npair of Dirac cones can be nondegenerate if time-reversal\nsymmetry or inversion symmetry is broken. The elec-\ntronic property of a Weyl semimetal is significantly influ-\nenced by the position of a pair of Weyl nodes in recipro-\ncal and energy spaces, where a Weyl node represents the\nband-touchingpointofeachDiraccone.Inthe absenceof\ntime-reversalsymmetry,apairofWeylnodesisseparated\nin reciprocal space. In this case, low-energy states with\nchirality appear on the surface of a Weyl semimetal3)if\nthe Weyl nodes are projected onto two different points in\nthe corresponding surface Brillouin zone. A notable fea-\nture of such chiral surface states is that they propagate\nonly in a given direction, which depends on the position\nof the Weyl nodes. This gives rise to an anomalous Hall\neffect.5)If inversion symmetry is also broken in addition\nto time-reversal symmetry, a pair of Weyl nodes is also\nseparated in energy space. In this case, the propagating\ndirection of chiral surface states is tilted accordingto the\ndeviation of the Weyl nodes. If only inversion symmetry\nis broken,no chiralsurface state appearssince the pairof\nWeyl nodes coincides in reciprocal space. To date, some\nmaterials have been experimentally identified as Weyl\nsemimetals.11–18)\nLet us focus on the case in which a Weyl semimetal\nwith a pair of Weyl nodes at k±= (0,0,±k0) is in the\nshape of a long prism parallel to the zaxis. In this case,\nchiral surface states appear on the side of the system. In\nthe presence of inversion symmetry, they typically prop-\nagate in a direction perpendicular to the zaxis; thus,\nwe expect that a spontaneous charge current appears to\ncirculate in the system near the side surface. If inver-\nsion symmetry is additionally broken, the propagating\ndirection is tilted to the zdirection; thus, an electron in\nthe chiral surface state shows spiral motion around the\nsystem.19)Thus, we expect that a spontaneous charge\ncurrent has a nonzero component in the zdirection. This\nlongitudinalcomponentmust becanceledoutbythe con-tribution from bulk states if they are integrated over a\ncross section parallel to the xyplane.\nIn this paper, we theoretically examine whether a\nspontaneous charge current appears in the ground state\nof a Weyl semimetal. Our attention is focused on the\ncase where time-reversal symmetry is broken. We calcu-\nlatethespontaneouschargecurrentinducedneartheside\nof the system by using a simple model with particle-hole\nsymmetry. We find that no spontaneous charge current\nappears when the Fermi level, EF, is located at the band\ncenter, which is set equal to 0 hereafter, implying that\nthe contribution from chiral surface states is completely\ncanceled out by that from bulk states. However, once EF\ndeviates from the band center, the spontaneous charge\ncurrent appears to circulate around the side surface of\nthe system and its direction of flow is opposite for the\ncases of electron doping (i.e., EF>0) and hole doping\n(i.e.,EF<0). The circulating chargecurrent is shown to\nbe robust against weak disorder. In the absence of inver-\nsion symmetry, we show that chiral surface states induce\nthe longitudinal component of a spontaneous charge cur-\nrent near the side surface, which is compensated by the\ncontribution from bulk states appearingbeneath the side\nsurface.Thislongitudinalcomponentisshowntobe frag-\nile against disorder.\nIn the next section, we present a tight-binding model\nfor Weyl semimetals and show the absence of a sponta-\nneous charge current when the Fermi level is located at\nthe band center. In Sect. 3, we derive a tractable con-\ntinuum model from the tight-binding model and analyt-\nically determine the magnitude of a spontaneous charge\ncurrent induced by the deviation of the Fermi level from\nthe band center. In Sect. 4, we numerically study the\nbehaviors of a spontaneous charge current by using the\ntight-binding model. We also examine the effect of disor-\nder on the spontaneous charge current. The last section\nis devoted to a summary and discussion. We set /planckover2pi1= 1\nthroughout this paper.\n1J. Phys. Soc. Jpn.\n2. Model\nLet us introduce a tight-binding model for Weyl\nsemimetals on a cubic lattice with lattice constant a. Its\nHamiltonian is given by H=H0+Hx+Hy+Hzwith4,5)\nH0=/summationdisplay\nl,m,n|l,m,n/an}bracketri}hth0/an}bracketle{tl,m,n|, (1)\nHx=/summationdisplay\nl,m,n{|l+1,m,n/an}bracketri}hthx/an}bracketle{tl,m,n|+h.c.},(2)\nHy=/summationdisplay\nl,m,n{|l,m+1,n/an}bracketri}hthy/an}bracketle{tl,m,n|+h.c.},(3)\nHz=/summationdisplay\nl,m,n{|l,m,n+1/an}bracketri}hthz/an}bracketle{tl,m,n|+h.c.},(4)\nwhere the indices l,m, andnare respectively used to\nspecify lattice sites in the x,y, andzdirections and\n|l,m,n/an}bracketri}ht ≡[|l,m,n/an}bracketri}ht↑,|l,m,n/an}bracketri}ht↓] (5)\nrepresents the two-component state vector with ↑,↓cor-\nresponding to the spin degree of freedom. The 2 ×2 ma-\ntrices are\nh0=/bracketleftbigg2tcos(k0a)+4B 0\n0 −2tcos(k0a)−4B/bracketrightbigg\n,(6)\nhx=/bracketleftbigg−Bi\n2A\ni\n2A B/bracketrightbigg\n, (7)\nhy=/bracketleftbigg−B1\n2A\n−1\n2A B/bracketrightbigg\n, (8)\nhz=/bracketleftbigg−t+iγ0\n0t+iγ/bracketrightbigg\n, (9)\nwhere 00. Here,qlabels the eigenstates\nin descending order (i.e., 0 ≥ǫv\n1≥ǫv\n2≥ǫv\n3≥...) in the\nvalence band and in ascending order (i.e., 0 ≤ǫc\n1≤ǫc\n2≤\nǫc\n3≤...) in the conduction band. Equation (14) allows\nus to setǫv\nq=−ǫc\nqwith\n|q/an}bracketri}htc= Γph|q/an}bracketri}htv. (15)\nWe introduce the charge current operator jαdefined\non an arbitrary site, where α=x,y,zspecifies the direc-\ntion of flow. For example, the current operator jxon the\n(l,m,n)th site is given by\njx=−e(−i)/bracketleftbig\n|l+1,m,n/an}bracketri}hthx/an}bracketle{tl,m,n|−h.c./bracketrightbig\n.(16)\nIt may be more appropriate to state that this is defined\non the link connecting the ( l,m,n)th and (l+1,m,n)th\nsites. We can show that any jαis invariant under the\ntransformation of Γ phas\nΓ−1\nphjαΓph=jα. (17)\nIn the ground state with EF= 0, the expectation value\nof anyjαis expressed as\n/an}bracketle{tjα/an}bracketri}ht=/summationdisplay\nq≥1v/an}bracketle{tq|jα|q/an}bracketri}htv. (18)\nBy using the relations given above and the completeness\nof the set of eigenstates consisting of {|q/an}bracketri}htv}and{|q/an}bracketri}htc},\nwe can show that\n/an}bracketle{tjα/an}bracketri}ht=1\n2/summationdisplay\nq≥1[v/an}bracketle{tq|jα|q/an}bracketri}htv+c/an}bracketle{tq|jα|q/an}bracketri}htc]\n=1\n2tr{jα}= 0, (19)\nindicating that a spontaneous charge current completely\nvanishes everywhere in the system at EF= 0. That is,\nalthough chiral surface states carry a circulating charge\n2J. Phys. Soc. Jpn.\ncurrent, their contribution is completely canceled out by\nthat from bulk states.\nNote that the above argument based on particle-hole\nsymmetry is not restricted to the two-orbital model used\nin this study and is also applicable to the four-orbital\nmodel introduced in Ref. 20.\n3. Analytical Approach\nBefore performing numerical simulations in the case\nofEF/ne}ationslash= 0, we analytically study the behaviors of chi-\nral surface states in a cylindrical Weyl semimetal. To do\nso, we apply the analytical approach given in Ref. 21,\nwhich was developed to describe unusual electron states\nin a Weyl semimetal: chiral surface states22)and chiral\nmodes along a screw dislocation.23,24)It is convenient to\nmodify the tight-binding Hamiltonian by taking the con-\ntinuum limit in the xandydirections, leaving the lattice\nstructure in the zdirection so that the resulting model\nhas a layered structure. After the partial Fourier trans-\nformation in the zdirection, the Hamiltonian is reduced\nto\nH=/bracketleftbigg˜Λ+2γsin(kza)˜A(ˆkx−iˆky)\n˜A(ˆkx+iˆky)−˜Λ+2γsin(kza)/bracketrightbigg\n,(20)\nwhere˜Λ = ∆(kz) +˜B(ˆk2\nx+ˆk2\ny) withˆkx=−i∂x,ˆky=\n−i∂y,˜A=Aa, and˜B=Ba2. We adapt this model\nto a cylindrical Weyl semimetal of radius Rby using\nthe cylindrical coordinates ( r,φ) withr=/radicalbig\nx2+y2and\nφ= arctan(y/x). Let Ψ(r,φ) =t(F,G) be an eigenfunc-\ntion ofHfor a given kz. It is convenient to rewrite F\nandGasF=eiλφf(r) andG=ei(λ+1)φg(r), where\nλis the azimuthal quantum number. Then, in terms of\nψ(r,φ) =t(f,g) for givenkzandλ, the eigenvalue equa-\ntion is written as\n/bracketleftbigg\n∆(kz)−˜BDλ˜A/parenleftbig\n−i∂r−iλ+1\nr/parenrightbig\n˜A/parenleftbig\n−i∂r+iλ\nr/parenrightbig\n−∆(kz)+˜BDλ+1/bracketrightbigg\nψ=˜Eψ,(21)\nwhere˜E=E−2γsin(kza) and\nDλ=∂2\nr+1\nr∂r−λ2\nr2. (22)\nAs demonstrated in Ref. 21, if ˜Bis finite but very\nsmall, the eigenvalue equation, Eq. (21), can be decom-\nposed into two separate equations: the Weyl and sup-\nplementary equations. The Weyl equation for fandgis\ngiven by\n(Dλ−Λ−)f= 0,(Dλ+1−Λ−)g= 0,(23)\nwhile the supplementary equation is\n(Dλ−Λ+)f= 0,(Dλ+1−Λ+)g= 0,(24)\nwhere\nΛ−=−˜E2−∆2\n˜A2,Λ+=˜A2\n˜B2.(25)\nHere,fandgarerelated by the originaleigenvalue equa-\ntion with a finite but very small ˜B. Note that the restric-\ntion on ˜B(i.e.,˜Bis very small) does not significantly\naffect the behaviors of chiral surface states. Indeed, the\nenergy of chiral surface states does not depend on ˜Bas\nseen in Eq. (28).We hereafter focus on chiral surface states, which ap-\npear only in the case of |∆(kz)|>|˜E|. The solutions\nof both the Weyl and supplementary equations are ex-\npressed by modified Bessel functions in this case. By su-\nperposing two solutions that asymptotically increase in\nan exponential manner, we can describe spatially local-\nized states near the side. With η≡/radicalbig\n∆2−˜E2/˜Aand\nκ≡˜A/˜B, the general solution for given λandkzis writ-\nten as\nψ=a/bracketleftBiggI|λ|(ηr)\n−i∆−˜E√\n∆2−˜E2I|λ+1|(ηr)/bracketrightBigg\n+b/bracketleftbiggI|λ|(κr)\niI|λ+1|(κr)/bracketrightbigg\n,\n(26)\nwhere the first and second terms respectively arise from\nthe Weyl and supplementary equations. The boundary\ncondition of ψ(R) =t(0,0) requires\n∆−˜E/radicalbig\n∆2−˜E2=−I|λ+1|(κR)\nI|λ|(κR)I|λ|(ηR)\nI|λ+1|(ηR),(27)\nindicating that a relevant solution is obtained only in the\ncase of ∆(kz)<0, which holds when kz∈(−k0,k0). The\neigenvalue of energy is approximately determined as\nE=˜A\nR/parenleftbigg\nλ+1\n2/parenrightbigg\n+2γsin(kza). (28)\nIn the case of γ= 0, the dispersion is flat (i.e., inde-\npendent of kz), representing a characteristic feature of\nthe chiral surface state. An electron in the chiral surface\nstate propagates in the anticlockwise direction viewed\nfrom above. The dispersion becomes dependent on kzif\nγ/ne}ationslash= 0, indicating that the groupvelocity is tilted to the z\ndirection. Consequently, an electron in the chiral surface\nstate circulates around the side surface in a spiral man-\nner.19)This implies that a spontaneous charge current\nin thezdirection can be induced near the side surface if\nγ/ne}ationslash= 0.\nNow, we roughly determine the magnitude of a spon-\ntaneous circulating current in the system consisting of\nNlayers. The circulating charge current carried by each\nchiral surface state is\nJ0\nφ=−e˜A\n2πR, (29)\nwhich flows in the clockwise direction. Note that the\nchiral surface state with E(λ) appears only when kz∈\n(−k0,k0). Ifkzdeviates from this interval, the state is\ncontinuously transformed to a bulk state, which is spa-\ntially extended over the entire system. Let us determine\nthe total chargecurrent Jφ. SinceJφvanishes at EF= 0,\nweneedtocollectthecontributionsto Jφarisingfromthe\nchiral surface states with E(λ) satisfying 0 0. Here, it is assumed that the contribution from\nbulk states in the same interval of energy is not impor-\ntant since their wavelength is relatively long and hence\nthey cannot induce a short-wavelength response local-\nized near the surface. If EF<0, the total charge current\nis obtained by collecting the contributions to Jφarising\nfrom the states with E(λ) satisfying EF< E <0 and\nthen reversingits sign. In addition to the condition for λ,\nit is important to note that the chiral surface states are\n3J. Phys. Soc. Jpn.\n-1 0 1\n 0(a)-πE/A\nkza\n-1 0 1\n 0(b)-πE/A\nkza\nFig. 2. Energy dispersion as a function of kzin the cases of (a)\nγ/A= 0 and (b) 0 .1. For the purely bulk states, only a quarter of\nthe corresponding branches are shown for clarity.\nstabilized only when |kz|0 near\nl= 1 andjy<0 nearl= 50 in the case of EF>0,\nwhereasjy<0 nearl= 1 andjy>0 nearl= 50 in\nthe case of EF<0. This indicates that the spontaneous\ncharge current circulates around the system in the clock-\nwise direction viewed from above when EF>0, whereas\nit circulates in the anticlockwise direction when EF<0\n(see Fig. 1). That is, its direction of flow is opposite for\nthe cases of EF>0 andEF<0. We also observe that jy\nincreaseswithincreasing EFinaccordancewithEq.(30).\nEquation (30) predicts |Jy|/N≈0.024×eAin the case\nofEF/A=±0.2, whereJyrepresents the total charge\ncurrent induced near each side surface of height N. This\nresult is consistent with those shown in Figs. 3(a) and\n3(d).\nSecondly, let us examine the behaviors of a sponta-\nneous charge current in the longitudinal direction. We\ncalculate the distribution of the charge current in the\nzdirection through the cross section parallel to the\nxyplane at the center of the system (broken line in\nFig. 1). Precisely speaking, jzon each link connecting\nthe (l,m,15)th and ( l,m,16)th sites is calculated for\n1≤l≤50 and 1 ≤m≤50. Figure 4 shows the results\nforjznormalized by eAforγ/A= 0.1 andEF/A= 0.2,\n0.1,−0.1, and−0.2. The results for γ/A= 0 are not\nshown since jzvanishes everywhere in this case. Again,\nFig. 4 indicates that jzincreases with increasing EFand\nthat its sign is opposite for the cases of EF>0 and\nEF<0. Note that a relatively large current appears\nnear the side surface, particularly near the corners, while\na small current flowing in the opposite direction is dis-\ntributed beneath the side surface. The former is induced\nby chiral surface states, while the latter originates from\nbulk states. These two contributions cancel each other\nout if they are integrated over the cross section; thus,\nthe total charge current in the zdirection completely\nvanishes.\nFinally, we examine the effect of disorder25–30)on the\nspontaneous charge current by adding the impurity po-\ntential term\nHimp=/summationdisplay\nl,m,n|l,m,n/an}bracketri}ht/bracketleftBigg\nV(l,m,n)\n1 0\n0V(l,m,n)\n2/bracketrightBigg\n/an}bracketle{tl,m,n|\n(31)\nto the Hamiltonian H, whereV1andV2are as-\nsumed to be uniformly distributed within the interval\nof [−W/2,+W/2]. Previous studies have shown that a\nWeyl semimetal phase is robust against weak disorder\nup to a critical disorder strength, Wc,25,26)and that chi-\nral surface states also persist as long as W < W c.29)In\nthe case of B/A= 0.5,t/A= 0.5, andk0a= 3π/4,\nthe critical disorder strength is Wc/A∼4. The ensemble 0 10 20 30 40 50 0 10 20 30 40 50-0.0040.0000.004\n(a)\nlmjz-0.0040.0000.004\n 0 10 20 30 40 50 0 10 20 30 40 50-0.0040.0000.004\n(b)\nlmjz\n 0 10 20 30 40 50 0 10 20 30 40 50-0.0040.0000.004\n(c)\nlmjz\n 0 10 20 30 40 50 0 10 20 30 40 50-0.0040.0000.004\n(d)\nlmjz\nFig. 4. (Color online) Spatial distribution of jznormalized by\neAin the cross section parallel to the xyplane; (a) EF/A= 0.2,\n(b) 0.1, (c)−0.1, and (d) −0.2.\naverages, /an}bracketle{tjy/an}bracketri}htand/an}bracketle{tjz/an}bracketri}ht, are calculated over 500 samples\nwith different impurity configurations at EF/A= 0.1\nfor a given value of W/A. In calculating jyandjzfor a\ngiven impurity configuration,we take accountofonly the\ncontribution from electron states with an energy Esat-\nisfying 0 -0.010.000.01\n 0 10 20 30 40 50 0 10 20 30-0.010.000.01\n(b)\nln< jy >\n 0 10 20 30 40 50 0 10 20 30-0.010.000.01\n(c)\nln< jy >\nFig. 5. (Color online) Spatial distribution of /angbracketleftjy/angbracketrightnormalized by\neAin the cross section parallel to the xzplane at EF/A= 0.1; (a)\nW/A = 3, (b) 4, and (c) 4 .5.\n 0 10 20 30 40 50 0 10 20 30 40 50-0.0020.0000.002\n(a)\nlm< jz >-0.0020.0000.002\n 0 10 20 30 40 50 0 10 20 30 40 50-0.0020.0000.002\n(b)\nlm< jz >\nFig. 6. (Color online) Spatial distribution of /angbracketleftjz/angbracketrightnormalized by\neAin the cross section parallel to the xyplane at EF/A= 0.1; (a)\nW/A = 2, and (b) 3.bust against disorder up to W/A∼4 but is suppressed\nwhenW/Aexceeds 4. This behavior is consistent with\nan observation reported previously.29)Figure 6 shows\nthe results for /an}bracketle{tjz/an}bracketri}htnormalized by eAforγ/A= 0.1 and\nEF/A= 0.1withW/A= 2and3.We observethat /an}bracketle{tjz/an}bracketri}htis\nsignificantlysuppressedinthecaseof W/A= 3,although\nthe circulatingchargecurrentis almost unaffected in this\ncase. This indicates that the charge current in the zdi-\nrectionis morefragilethan the circulatingchargecurrent\nagainstthe mixingofchiralsurfacestatesandbulk states\ndue to disorder.\n5. Summary and Discussion\nWe theoretically studied a spontaneous charge current\ndue to chiral surface states in the ground state of a Weyl\nsemimetal. We analytically and numerically determined\nthemagnitudeofthechargecurrentinducedneartheside\nsurface of the system. It is shown that no spontaneous\ncharge current appears when the Fermi level, EF, is lo-\ncated at the band center. It is also shown that, once EF\ndeviates from the band center, the spontaneous charge\ncurrent appears to circulate around the side surface of\nthe system and its direction of flow is opposite for the\ncases of electron doping and hole doping. The circulating\ncurrent is shown to be robust against weak disorder.\nLet us focus on the two features revealed in this paper:\nthe appearance of a spontaneous charge current except\nat the band center and the reversal of its direction of\nflow as a function of EF. As they are derived by using a\nmodel possessingparticle-hole symmetry, a natural ques-\ntion arises: do these features manifest themselves even in\nthe absence of particle-hole symmetry? The answer is\nyes. The disappearance of the spontaneous charge cur-\nrent reflects the fact that the contribution from chiral\nsurface states is completely canceled out by that from\nbulk states. As the spontaneous charge current due to\nchiral surface states is localized near the side surface,\nthis cancellation should be mainly caused by bulk states\nwith a short wavelength, occupying the bottom region\nof the energy band far from the band center. Hence, if\nEFis varied near the band center, the contribution from\nbulk states is almost unaffected but that from chiral sur-\nface states is significantly changed, depending on EFin\na roughly linear manner. This behaviorshould take place\nregardlessofthepresenceorabsenceofparticle-holesym-\nmetry. Thus, we expect that the features of the sponta-\nneouschargecurrentstillmanifest themselveseveninthe\nabsence of particle-hole symmetry, although the point of\nthe disappearance shifts away from the band center.\nAcknowledgment\nThis work was supported by JSPS KAKENHI Grant\nNumbers JP15K05130 and JP18K03460.\n1) R. Shindou and N. Nagaosa, Phys. Rev. Lett. 87, 116801\n(2001).\n2) S. Murakami, New J. Phys. 9, 356 (2007).\n3) X. Wan, A. M. Turner, A. Vishwanath, and S. Y. Savrasov,\nPhys. Rev. B 83, 205101 (2011).\n4) K.-Y. Yang, Y.-M. Lu, and Y. Ran, Phys. Rev. B 84, 075129\n6J. Phys. Soc. Jpn.\n(2011).\n5) A. A. Burkov and L. Balents, Phys. Rev. Lett. 107, 127205\n(2011).\n6) A. A. Burkov, M. D. Hook, and L. Balents, Phys. Rev. B 84,\n235126 (2011).\n7) W. Witczak-Krempa and Y. B. Kim, Phys. Rev. B 85, 045124\n(2012).\n8) P.Delplace, J.Li, and D.Carpentier, Europhys.Lett. 97, 67004\n(2012).\n9) G. Hal´ asz and L. Balents, Phys. Rev. B 85, 035103 (2012).\n10) A. Sekine and K. Nomura, J. Phys. Soc. Jpn. 82, 033702 (2013).\n11) H. Weng, C. Fang, Z. Fang, B. A. Bernevig, and X. Dai, Phys.\nRev. X5, 011029 (2015).\n12) S.-M. Huang, S.-Y. Xu, I. Belopolski, C.-C. Lee, G. Chang , B.\nWang, N. Alidoust, G. Bian, M. Neupane, C. Zhang, S. Jia,\nA. Bansil, H. Lin, and M. Z. Hasan, Nat. Commun. 6, 7373\n(2015).\n13) S.-Y. Xu, I. Belopolski, N. Alidoust, M. Neupane, G. Bian , C.\nZhang, R. Sankar, G. Chang, Z. Yuan, C.-C. Lee, S.-M. Huang,\nH. Zheng, J. Ma, D. S. Sanchez, B. Wang, A. Bansil, F. Chou,\nP. P. Shibayev, H. Lin, S. Jia, and M. Z. Hasan, Science 349,\n613 (2015).\n14) B.-Q. Lv, H.-M. Weng, B.-B. Fu, X.-P. Wang, H. Miao, J. Ma,\nP. Richard, X.-C. Huang, L.-X. Zhao, G.-F. Chen, Z. Fang, X.\nDai, T. Qian, and H. Ding, Phys. Rev. X 5, 031013 (2015).\n15) B. Q. Lv, N. Xu, H. M. Weng, J. Z. Ma, P. Richard, X. C.\nHuang, L. X. Zhao, G. F. Chen, C. E. Matt, F. Bisti, V. N.\nStrocov, J. Mesot, Z. Fang, X. Dai, T. Qian, M. Shi, and H.\nDing, Nat. Phys. 11, 724 (2015).\n16) S.-Y. Xu, N. Alidoust, I. Belopolski, Z. Yuan, G. Bian, T. -R.\nChang, H. Zheng, V. N. Strocov, D. S. Sanchez, G. Chang, C.Zhang, D. Mou, Y. Wu, L. Huang, C.-C. Lee, S.-M. Huang,\nB. Wang, A. Bansil, H.-T. Jeng, T. Neupert, A. Kaminski, H.\nLin, S. Jia, and M. Z. Hasan, Nat. Phys. 11, 748 (2015).\n17) S. Souma, Z. Wang, H. Kotaka, T. Sato, K. Nakayama, Y.\nTanaka, H. Kimizuka, T. Takahashi, K. Yamauchi, T. Oguchi,\nK. Segawa, and Y. Ando, Phys. Rev. B 93, 161112 (2016).\n18) K. Kuroda, T. Tomita, M.-T. Suzuki, C. Bareille, A. A. Nu-\ngroho, P. Goswami, M. Ochi, M. Ikhlas, M. Nakayama, S.\nAkebi, R. Noguchi, R. Ishii, N. Inami, K. Ono, H. Kumigashira ,\nA. Varykhalov, T. Muro, T. Koretsune, R. Arita, S. Shin, T.\nKondo, and S. Nakatsuji, Nat. Mater. 16, 1090 (2017).\n19) P. Baireuther, J. A. Hutasoit, J. Tworzyd/suppress lo, and C. W. J.\nBeenakker, New J. Phys. 18, 045009 (2016).\n20) M. M. Vazifeh and M. Franz, Phys. Rev. Lett. 111, 027201\n(2013).\n21) Y. Takane, J. Phys. Soc. Jpn. 86, 123708 (2017).\n22) R.Okugawa and S.Murakami, Phys.Rev.B 89, 235315 (2014).\n23) K.-I. Imura and Y. Takane, Phys. Rev. B 84, 245415 (2011).\n24) H. Sumiyoshi and S. Fujimoto, Phys. Rev. Lett. 116, 166601\n(2016).\n25) C.-Z. Chen, J. Song, H. Jiang, Q.-F. Sun, Z. Wang, and X. C.\nXie, Phys. Rev. Lett. 115, 246603 (2015).\n26) H. Shapourian and T. L. Hughes, Phys. Rev. B 93, 075108\n(2016).\n27) S. Liu, T. Ohtsuki, and R. Shindou, Phys. Rev. Lett. 116,\n066401 (2016).\n28) E. V. Gorbar, V. A. Miransky, I. A. Shovkovy, and P. O. Sukh a-\nchov, Phys. Rev. B 93, 235127 (2016).\n29) Y. Takane, J. Phys. Soc. Jpn. 85, 124711 (2016).\n30) Y. Yoshimura, W. Onishi, K. Kobayashi, T. Ohtsuki, and K. -I.\nImura, Phys. Rev. B 94, 235414 (2016).\n7" }, { "title": "1006.5179v1.Competition_between_paramagnetism_and_diamagnetism_in_charged_Fermi_gases.pdf", "content": "arXiv:1006.5179v1 [cond-mat.stat-mech] 27 Jun 2010Competition between paramagnetism and diamagnetism in cha rged Fermi gases\nXiaoling Jian, Jihong Qin, and Qiang Gu∗\nDepartment of Physics, University of Science and Technolog y Beijing, Beijing 100083, China\n(Dated: November 10, 2018)\nThe charged Fermi gas with a small Lande-factor gis expected to be diamagnetic, while that with\na larger gcould be paramagnetic. We calculate the critical value of th eg-factor which separates\nthe dia- and para-magnetic regions. In the weak-field limit, gchas the same value both at high and\nlow temperatures, gc= 1/√\n12. Nevertheless, gcincreases with the temperature reducing in finite\nmagnetic fields. We also compare the gcvalue of Fermi gases with those of Boltzmann and Bose\ngases, supposing the particle has three Zeeman levels σ=±1,0, and find that gcof Bose and Fermi\ngases is larger and smaller than that of Boltzmann gases, res pectively.\nPACS numbers: 05.30.Fk, 51.60.+a, 75.10.Lp, 75.20.-g\nI. INTRODUCTION\nMagnetism of electron gases has been considerably\nstudied in condensed matter physics. In magnetic field,\nthe magnetization of a free electron gas consists of two\nindependent parts. The spin magnetic moment of elec-\ntrons results in the paramagnetic part (the Pauli para-\nmagnetism), while the orbital motion due to charge de-\ngreeoffreedom in magneticfield induces the diamagnetic\npart (the Landau diamagnetism)[1]. The Pauli paramag-\nnetism and the Landau diamagnetism compete with each\nother. For electrons whose Lande-factor g= 2, the zero-\nfield paramagnetic susceptibility is two times stronger\nthan the diamagnetic susceptibility, so the free electron\ngas exhibits paramagnetism altogether.\nMagnetic properties of relativistic Fermi gases have\nalso been under extensive investigation. Daicic et al.\ndeveloped statistical mechanics for the magnetized pair-\nfermion gases and found that the intrinsic spin causes\nimportant effects upon the relativistic para- and dia-\nmagnetism[2].\nThe study of ultracold atoms has stimulated renewed\nresearch interest in the magnetism of quantum gases[3–\n6]. When atomic gases are confined in optical traps[3, 4],\ntheir spin degree of freedom becomes active, leading\nto the manifestation of magnetism. Theoretically, the\nparamagnetism[7] and ferromagnetism[8–10] in a neutral\nspin-1 Bose gas have ever been studied. In experiments,\nmagnetic domains have been directly observed in87Rb\ncondensate, a typical ferromagneticspinor condensate[5].\nVery recently, the exploration of magnetism in quantum\ngases has been extended to the Fermi gas. It is already\nobserved that an ultracold two-component Fermi gas of\nneutral6Li atoms exhibits ferromagnetism caused by re-\npulsive interactions between atoms[6].\nFurthermore, it is possible to realize charged quantum\ngases from neutral ultracold atoms. So far, cold plasma\nhas been created by photoionization of cold atoms[11].\nThe temperatures of electrons and ions are as low as 100\n∗Corresponding author: qgu@ustb.edu.cnmK and 10 µK, respectively. The ions can be regarded\nas charged Bose or Fermi gases. Once the quantum gas\nhas both the spin and charge degrees of freedom, there\narises the competition between paramagnetism and dia-\nmagnetism, as in electrons. Different from electrons, the\ng-factor for different magnetic ions is diverse, ranging\nfrom 0 to 2.\nTheg-factor is a characteristic parameter to measure\nthe intensity of paramagnetic effect. It is expected that\nthequantumgasdisplaysdiamagnetisminsmall gregion,\nbut paramagnetism in large gregion. The main purpose\nof this paper is to calculate the critical value of g. We\nalso present a comparison with results of charged spin-1\nBose gases which have been obtained previously[12].\nII. THE MODEL\nWe consider a charged Fermi gas of Nparticles, with\nthe effective Hamiltonian written as\n¯H−µN=DL/summationdisplay\nj,kz,σ(ǫjkz+ǫσ−µ)njkzσ,(1)\nwhereµis the chemical potential and ǫjkzis the quan-\ntized Landau energy in the magnetic field B,\nǫjkz= (1\n2+j)/planckover2pi1ω+/planckover2pi12k2\nz\n2m∗(2)\nwithj= 0,1,2,...labeling different Landau levels and\nω=qB/(m∗c) being the gyromagnetic frequency of a\nfermionwithcharge qandmassm∗.DL=qBSxy/(2π/planckover2pi1c)\nmarks the degeneracy of each Landau level with Sxythe\ntotal section area in x, y directions of the system. ǫσ\ndenotes the Zeeman energy,\nǫσ=−g/planckover2pi1q\nm∗cσB=−gσ/planckover2pi1ω, (3)\nwheregis the Lande-factor and σrefers to the spin- z\nindex of Zeeman state |F,σ/angbracketright.\nThe grand thermodynamic potential of the Fermi gas\nis expressed as\nΩT/negationslash=0=−1\nβDL/summationdisplay\nj,kz,σln[1+e−β(ǫjkz+ǫσ−µ)],(4)2\nwhereβ= (kBT)−1. Performing Taylor expansions and\nthen integrating out kz, Eq. (4) becomes\nΩT/negationslash=0=−ωV\n/planckover2pi12/parenleftbiggm∗\n2πβ/parenrightbigg3/2\n×��/summationdisplay\nl=1/summationdisplay\nσ(−1)l+1l−3\n2e−lβ(/planckover2pi1ω\n2−gσ/planckover2pi1ω−µ)\n1−e−lβ/planckover2pi1ω,(5)\nwhereVis the volume of the system. For simplicity, the\nfollowing notation is introduced,\nFσ\nτ[α,δ] =∞/summationdisplay\nl=1(−1)l+1lα/2e−lx(ησ+δ)\n(1−e−lx)τ,(6)\nwherex=β/planckover2pi1ωandησ= (/planckover2pi1ω/2−µ+ǫσ)/(/planckover2pi1ω). Then\nEq. (5) is rewritten as\nΩT/negationslash=0=−ωV\n/planckover2pi12/parenleftbiggm∗\n2πβ/parenrightbigg3/2/summationdisplay\nσFσ\n1[−3,0].(7)\nFor a system with the given particle density n=N/V,\nthe chemical potential is obtained according to the fol-\nlowing equation\nn=−1\nV/parenleftbigg∂Ω\n∂µ/parenrightbigg\nT,V=x/parenleftbiggm∗\n2πβ/planckover2pi12/parenrightbigg3/2/summationdisplay\nσFσ\n1[−1,0].(8)\nA similar treatment has been employed to study diamag-\nnetism of the charged spinless Bose gas[13] and extended\nto the study of competition of diamagnetism and param-\nagnetism in charged spin-1 Bose gases[12]. This method\nis more applicable at relatively high temperatures.\nTo determine gc, we need calculate the magnetization\ndensityMas a function of the magnetic field Band tem-\nperatureT. The system is paramagnetic when M >0\nwhile diamagnetic when M <0.Mcan be derived from\nthe thermodynamic potential by the standard procedure\nMT/negationslash=0=−1\nV/parenleftbig∂Ω\n∂B/parenrightbig\nT,V, which yields\nMT/negationslash=0=/planckover2pi1q\nm∗c/parenleftbiggm∗\n2πβ/planckover2pi12/parenrightbigg3/2/summationdisplay\nσ/braceleftbigg\nFσ\n1[−3,0]\n+x(gσ−1\n2)Fσ\n1[−1,0]−xFσ\n2[−1,1]/bracerightbigg\n.(9)\nFor carrying out numerical calculations, it is useful\nto make the parameters dimensionless, such as M=\nm∗cM/(n/planckover2pi1q),ω=/planckover2pi1ω/(kBT∗), andt=T/T∗. HereT∗\nis the characteristic temperature of the system, which is\ngiven bykBT∗= 2π/planckover2pi12n2\n3/m∗. Then we have x=ω/t,\nησ= 1/2−µ′/ω−gσandµ′=µ/(kBT∗) in the notation\nFσ\nτ[α,δ]. Equations (8) and (9) can be re-expressed as\n1 =ωt1/2/summationdisplay\nσFσ\n1[−1,0] (10)/s45/s48/s46/s53/s48/s48/s46/s48/s48/s48/s46/s53/s48\n/s48/s46/s48/s48/s48/s46/s52/s48/s48/s46/s56/s48\n/s48/s46/s48 /s48/s46/s52 /s48/s46/s56/s45/s48/s46/s52/s56/s45/s48/s46/s52/s52/s40/s99/s41\n/s32/s32/s77/s40/s97/s41\n/s40/s98/s41\n/s32\n/s32/s32/s77\n/s112\n/s32/s32/s77\n/s100\n/s103\nFIG. 1: (a) The total magnetization density ( M), (b) the\nparamagnetization density ( Mp), and (c) the diamagnetiza-\ntion density ( Md) as a function of gatt= 0.2. The dotted\nand dashed lines correspond to ω= 5 and 10, respectively.\nand\nMT/negationslash=0=t3/2/summationdisplay\nσ/braceleftbigg\nFσ\n1[−3,0]+x(gσ−1\n2)\n×Fσ\n1[−1,0]−xFσ\n2[−1,1]/bracerightbigg\n.(11)\nIn following discussions, ωis used to indicate the magni-\ntude of magnetic field since it is proportional to B.\nIII. RESULTS AND DISCUSSIONS\nFirst, we look at a spin-1\n2Fermi gas, setting σ=±1\nto present the two Zeeman levels. The dimensionless\nmagnetization density Mas a function of gis shown in\nFig. 1(a). As expected, Mis negative in the small g\nregion, which means that the diamagnetism dominates.\nFor each given value of ω,Mgrows with gand changes\nits sign from negative to positive in the larger gregion,\nindicating that the paramagnetic effect is enhanced due\nto increase of g. This phenomenon is also observed in the\nBose system[12]. gcis just the value of gwhereM= 0.3\n/s48 /s53 /s49/s48 /s49/s53 /s50/s48/s48/s46/s50/s56/s48/s46/s51/s53/s48/s46/s52/s50/s48/s46/s52/s57\n/s32 /s32/s61/s32/s50/s48\n/s32 /s32/s61/s32/s49/s48\n/s32 /s32/s61/s32/s53\n/s32/s32/s103\n/s99\n/s49/s47/s116\nFIG. 2: Plots of the critical value of g-factor as a function of\n1/t. The magnetic field is chosen as ω= 20 (solid line), 10\n(dashed line) and 5 (dotted line), respectively.\nThe total magnetization Mshown in Fig. 1(a) con-\nsists of both the paramagnetic and diamagnetic contri-\nbutions. Figure 1(b) depicts the pure paramagnetic con-\ntribution to M, which is calculated by Mp=gmwhere\nm=n1−n−1. Figure 1(c) plots the diamagnetic con-\ntribution to M,Md=M−Mp. For each fixed value\nofω, the diamagnetization Mdis slightly weakened with\nincreasingg. Interestingly, for a charged spin-1 Bose gas\nin a constant magnetic field, the diamagnetism is not\nsuppressed but enhanced as gbecomes larger[12]. Com-\nparing Figs 1(a), 1(b) and 1(c), it can be seen clearly\nthat the increase of Mwithgis mainly owing to the\nparamagnetization Mp.\ngcis an important parameter to describe the competi-\ntionbetweenthediamagnetismandparamagnetism. Fig-\nure 2 plotsgcfor the chargedspin-1\n2Fermi gas. The tem-\nperature is described by 1 /tand the magnetic fields are\nchosen to be relatively larger, since the method adopted\nhere is more applicable at higher temperatures or in\nstronger fields. In the high temperature limit, gctends\nto a universal value, gc|t→∞= 0.28868, regardless of the\nstrength of the magnetic field. In a fixed magnetic field,\ngcincreasesmonotonicallyasthe temperaturefalls down.\nButthetrendofincreasingissloweddownifthemagnetic\nfield is weakened.\nPresent method can not produce valid results in the\nlowtemperatureregionifasmallmagneticfieldischosen.\nIn this case that gω≪t≪µ′, the grand thermodynamic\npotential in Eq. (4) can be calculated using the Euler-\nMaclaurin formula\n∞/summationdisplay\nj=0ψ(j+1\n2)≈/integraldisplay∞\n0ψ(x)dx+1\n24ψ′(0),(12)\nand then Eq. (4) is transformed into\nΩT/negationslash=0=−V\nβλ3/summationdisplay\nσf5\n2(z)+Vβ\n24λ3(/planckover2pi1ω)2/summationdisplay\nσf1\n2(z),(13)/s48 /s53 /s49/s48 /s49/s53 /s50/s48/s48/s46/s51/s53/s48/s46/s52/s48/s48/s46/s52/s53/s48/s46/s53/s48\n/s32/s61/s32/s53\n/s32/s32/s103\n/s99\n/s49/s47/s116/s32/s61/s32/s49/s48\nFIG. 3: The gc−1/tcurves for charged spin-1 gases obey-\ning the Bose-Einstein (BE, dotted line), Maxwell-Boltzman n\n(MB, solid line) and Fermi-Dirac (FD, dashed line) statisti cs,\nrespectively. The magnetic field is chosen as ω= 10 and 5.\nwhereλ=hβ1\n2/(2πm∗)1\n2,z=eβ(µ−ǫσ)and the Fermi-\nDirac integral fn(z) is normally defined as\nfn(z)≡1\nΓ(n)/integraldisplay∞\n0xn−1\nz−1ex+1dx, (14)\nwhere Γ(n) is a usual gamma function and x=βǫ. Then\nthe magnetization density, M, is obtained from grand\nthermodynamic potential in Eq. (13),\nMT/negationslash=0=gt3/2/summationdisplay\nσσf3\n2(z′)−ωt1/2\n12/summationdisplay\nσf1\n2(z′)\n−gω2t−1/2\n24/summationdisplay\nσσf−1\n2(z′), (15)\nwherez′=e(µ′+gσω)/t. After some algebra, we get from\nEq. (15) that gc|ω→0= 1/√\n12≈0.28868. So in the weak\nfield limit, gchas the same value both at the high and\nlow temperature limit. This implies that the gc−1/t\ncurve is likely to flatten out in the weak-field limit.\nTo proceed, we make a comparison between Fermi\nand Bose gases. Considering that gcfor charged spin-\n1 Bose gas has already been studied[12], we discuss a\nFermi gas with three sublevels, σ=±1,0 (a pseudo-\nspin-1 Fermi gas) to ensure the comparability. Results\nfor the spin-1 Boltzmann gas are also obtained. Fig-\nure 3 shows the gc−1/tcurves for the three kinds of\ncharged spin-1 gases. In a given value of ω,gcof all the\nthree gases displays similar temperature-dependence: gc\nincreases monotonously as treduces. In the high tem-\nperature limit, gcgoes to the same value in all magnetic\nfields,gc|t→∞= 1/√\n8, reflecting that the Bose-Einstein\n(BE) and Fermi-Dirac (FD) statistics coincide with the\nMaxwell-Boltzmann (MB) statistics in this case.\nFigure 3 also demonstrates the difference among the\nthree kinds of statistics. For each given fixed value of ω,4\nthegc−1/tcurves of Bose and Fermi gases always locate\nat the two sides of that of the Boltzmann gas. Given the\nsame temperature and magnetic field, gcof Fermi gas\nis the smallest. According to our previous research[12],\ngc|t→0= 1/2 for the Boltzmann gas regardless of the\nmagnetic field. This means that gcof the Fermi gas does\nnever exceed 1 /2 in the low temperature, no matter how\nstrong the field is.\nIV. SUMMARY\nIn summary, we study the interplay between paramag-\nnetism and diamagnetismofchargedFermi gasessuppos-\ning that the Lande-factor gis a variable. The gas under-\ngoes a shift from diamagnetism to paramagnetism at thecritical value of gandgcincreases monotonically as the\ntemperature tdecreases in a fixed magnetic field ω, and\nthe rise ingcis lowered as ωis reduced. We conjecture\nthatgcholds a constant at all temperatures in the weak\nfield limit. For a spin-1\n2Fermi gas,gc|ω→0= 1/√\n12. We\nalso briefly compare gcof charged spin-1 gases obeying\nthe Fermi-Dirac, Bose-Einstein and Maxwell-Boltzmann\nstatistics. The gc−1/tcurves of Boltzmann gases are al-\nways between those of Bose and Fermi gases in the same\nmagnetic field. In the high temperature limit, gcof all\nthe three gases tends to the same value.\nThisworkwassupportedbytheFokYingTungEduca-\ntion Foundation of China (No. 101008), the Key Project\nof the Chinese Ministry of Education (No. 109011), and\nthe Fundamental Research Funds for the Central Univer-\nsities of China.\n[1] L.D. Landau, E.M. Lifshitz, Statistical Physics. Part 1 ,\nButterworth-Heinemann, Oxford, 1980.\n[2] J. Daicic , N.E. Frankel, R.M. Gailis, V. Kowalenko,\nPhys. Rep. 63 (1994) 237 and references therein.\n[3] D.M. Stamper-Kurn, M.R. Andrews, A.P. Chikkatur, S.\nInouye, H.-J. Miesner, J. Stenger, W. Ketterle, Phys.\nRev. Lett. 80 (1998) 2027.\n[4] J. Stenger, S. Inouye, D.M. Stamper-Kurn, H.-J. Mies-\nner, A.P. Chikkatur, W. Ketterle, Nature 396 (1998) 345.\n[5] L.E. Sadler, J.M. Higbie, S.R. Leslie, M. Vengalattore,\nD.M. Stamper-Kurn, Nature 443 (2006) 312.\n[6] G.B. Jo, Y.R. Lee, J.H. Choi, C.A. Christensen, T.H.\nKim, J.H. Thywissen, D.E. Pritchard, W. Ketterle, Sci-\nence 325 (2009) 1521.\n[7] K. Yamada, Prog. Theor. Phys. 76 (1982) 443; M.V.Simkin, E.G.D. Cohen, Phys. Rev. A 59 (1999) 1528.\n[8] T.-L. Ho, Phys. Rev. Lett. 81 (1998) 742; T. Ohmi, K.\nMachida, J. Phys. Soc. Jpn. 67 (1998) 1822.\n[9] Q. Gu, R.A. Klemm, Phys. Rev. A 68 (2003) 031604(R);\nC. Tao, P. Wang, J. Qin, Q. Gu, Phys. Rev. B 78 (2008)\n134403.\n[10] K. Kis-Szabo, P. Szepfalusy, G. Szirmai, Phys. Rev. A\n72 (2005) 023617; S. Ashhab, J. Low Tempt. Phys. 140\n(2005) 51.\n[11] T.C. Killian, S. Kulin, S.D. Bergeson, L.A. Orozco, C.\nOrzel, S.L. Rolston, Phys. Rev. Lett. 83 (1999) 4776.\n[12] X. Jian, J. Qin, Q. Gu, unpublished.\n[13] G.B. Standen, D.J. Toms, Phys. Rev. E 60 (1999) 5275." }, { "title": "2303.14534v1.Spin_momentum_locking_and_ultrafast_spin_charge_conversion_in_ultrathin_epitaxial_Bi___1_x__Sb__x__topological_insulator.pdf", "content": "Spin-momentum locking and ultrafast spin-charge conversion in\nultrathin epitaxial Bi 1\u0000xSbxtopological insulator.\nE. Rongione,1, 2L. Baringthon,1, 3, 4D. She,1, 3, 4G. Patriarche,4R. Lebrun,1\nA. Lemaître,4M. Morassi,4N. Reyren,1M. Mičica,2J. Mangeney,2J. Tignon,2\nF. Bertran,3S. Dhillon,2, 5P. Le Fèvre,3H. Jaffrès,1, 5and J.-M. George1, 5\n1Unité Mixte de Physique, CNRS, Thales,\nUniversité Paris-Saclay, F-91767 Palaiseau, France\n2Laboratoire de Physique de l’Ecole Normale Supérieure,\nENS, Université PSL, CNRS, Sorbonne Université,\nUniversité Paris Cité, F-75005 Paris, France\n3Synchrotron SOLEIL, L’Orme des Merisiers,\nDépartementale 128, F-91190 Saint-Aubin, France\n4Université Paris-Saclay, CNRS, Centre de Nanosciences\net de Nanotechnologies, F-91120 Palaiseau, France\n5Corresponding authors: sukhdeep.dhillon@phys.ens.fr,\nhenri.jaffres@cnrs-thales.fr, jeanmarie.george@cnrs-thales.fr\n(Dated: March 28, 2023)\n1arXiv:2303.14534v1 [cond-mat.mtrl-sci] 25 Mar 2023Abstract\nThe helicity of 3D topological insulator surface states has drawn significant attention\ninspintronicsowingtospin-momentumlockingwherethecarriers’spinisorientedper-\npendicular to their momentum. This property can provide an efficient method to con-\nvert charge currents into spin currents, and vice-versa, through the Rashba-Edelstein\neffect. However, experimental signatures of these surface states to the spin-charge\nconversion are extremely difficult to disentangle from bulk state contributions. Here,\nwe combine spin- and angle-resolved photo-emission spectroscopy, and time-resolved\nTHz emission spectroscopy to categorically demonstrate that spin-charge conversion\narises mainly from the surface state in Bi 1\u0000xSbxultrathin films, down to few nanome-\nters where confinement effects emerge. We correlate this large conversion efficiency,\ntypically at the level of the bulk spin Hall effect from heavy metals, to the complex\nFermi surface obtained from theoretical calculations of the inverse Rashba-Edelstein\nresponse. Both surface state robustness and sizeable conversion efficiency in epitaxial\nBi1\u0000xSbxthinfilmsbringnewperspectivesforultra-lowpowermagneticrandom-access\nmemories and broadband THz generation.\nKeywords: topological insulator, ARPES, spin-resolved ARPES, THz-TDS, spin-charge conversion, surface\nstates\n2INTRODUCTION\nThe discovery of metallic quantum states at the surface of 3D topological insulators\n(TIs) [1–3] has opened exciting new functionalities in spintronics owing to their topological\nprotection and spin-momentum locking (SML) properties [4, 5]. Indeed, the combination\nof band inversion and time reversal symmetry (TRS) results in a peculiar spin texture in\nmomentumspace. Injectingacurrentinthesestatesresultsthereforeinanoutofequilibrium\nspin density (also called spin-accumulation) along the transverse direction [4–6]. This is the\nRashba-Edelstein effect (REE) [7] which can be used to exert a spin-orbit torque (SOT)\nonto the magnetization of an adjacent ferromagnet (FM) [8]. The reciprocal phenomenon,\nby which a spin density produces an in-plane transverse charge current, is called the Inverse\nRashba-Edelstein effect (IREE) [6, 9].\nImportantly, the resulting spin-charge conversion (SCC) efficiencies in topological surface\nstates (TSS) combining strong spin-orbit coupling (SOC) and SML is expected to be at least\noneorderofmagnitudelargercomparedtothespinHalleffect(SHE)of5 dheavymetals[10–\n12]. SCC has been demonstrated in a range of Bi-based TI coumpounds, including bismuth\nselenide Bi 2Se3, bismuth telluride Bi 2Te3, Bi 2(Se,Te) 3[13] or Bi 1\u0000xSbx(BiSb) [14]. To\nbenefit fully from IREE, the charge currents should be confined in the surface states and\nany current flowing through the bulk states should be avoided. The prerequisites are hence\ni) a sizeable bandgap typically larger than 0.2 to 0.3 eV, and ii) a perfect control of the\nFermi level position, usually achieved by stoichiometry and/or strain engineering. In this\nrespect, Bi 1\u0000xSbxalloys, although displaying clear topological surface states [15–17], have\nbeen mostly neglected for spintronic applications as a result of their modest bulk bandgap\n(about 40 meV for x= 0:07) and relatively complex band structure. However, quantization\neffects in ultrathin films have shown to lead to much larger gap [18, 19] while retaining their\nband inversion near the M point in the x= 0:07\u00000:3composition range [20], unlike pure\nBi[21–23]. BiSbthereforehasconsiderablepotentialascandidateforspintronicsapplications\nas well as for recently engineered efficient spintronic THz emitters [24–31].\nIn this letter, we report on our detailed investigation of the surface state SML properties\nof ultrathin (111)-oriented Bi 1\u0000xSbxepitaxial films. They exhibit a topological phase as\nrecently confirmed by our angular-resolved photo-emission spectroscopy (ARPES) measure-\nments [18]. Here, we focus on spin-resolved ARPES (SARPES) performed on ultrathin BiSb\n3films and extract the in-plane spin texture for the different electron and hole pockets charac-\nterizing the complex BiSb Fermi surface. Moreover, the SCC mediated by the BiSb surface\nstates is probed at the sub-picosecond timescale using an adjacent metallic Co layer acting\nas a spin injector. Unprecedentedly large SCC is measured with efficiencies beyond the level\nof carefully optimized Co/Pt systems. Our results also indicate that surface state related\nIREE is the mechanism responsible for SCC mechanism. Tight-binding (TB) calculation\nand linear response theory account for our findings.\nI. SPIN-RESOLVED ARPES\nSARPES is the method of choice to probe the spin-textured Fermi contour of TI surfaces.\nUltrathin epitaxial Bi 1\u0000xSbxfilms withx= 0:07, 0.1, 0.15, 0.21, 0.3, 0.4 and thicknesses\ndown to 2.5 nm were grown by molecular beam epitaxy (MBE) on clean 7\u00027reconstructed\nSi(111) surfaces. They all exhibit a non-trivial topological phase. Details of the growth [32]\nare quickly recalled in the Methods section where scanning transmission electron microscopy\n(STEM) and energy-dispersive X-ray spectroscopy (EDX) characterizations are also dis-\ncussed. We focus first on the 5 nm thick Bi 0:85Sb0:15sample. Fig. 1a shows its experimental\nFermi surface within the 2D surface Brillouin zone, as measured by ARPES. It is composed\nof three pockets labelled P1(the hexagonal electron pocket surrounding \u0000), and two elon-\ngatedP2hole andP3electron pockets along each of the six equivalent \u0000M directions (one\nhas been chosen as the kxaxis). Fig. 1b displays the energy dispersion along the \u0000M di-\nrection. The signature of two surface S1andS2states [17, 18] are visible in the bandgap.\nThe valence band state energy dispersion is also visible, from a series of confined states with\nenergy splitting increasing with reducing thickness [18]. The corresponding experimental\nSARPES polarization map for the \u001byspin-polarization component is given in Fig. 1c. For\nthis experiment, the incident photon energy is 20eV (5 meV resolution).\nOwing to the electron analyzer movable entrance optics, the \u001bxand\u001byspin polarizations\nwere measured on a large part of the surface first Brillouin zone (see Methods). On Fig. 1e,\nthe spin polarization is represented as a vector field. For clarity, the measured polarization is\nonly displayed for positions where the DOS is the largest. The norm is roughly proportional\ntothespin-resolvedDOS(s-DOS).Thevectorfieldissuperimposedoveracolormapyielding\nthe sum of all the signals measured by the spin detector, which is proportional to the\n4DOS at the Fermi level. For a clearer representation, the vector fields were averaged over\n0:02\u00020:02Å\u00002areas, corresponding to twenty data points. The experimental data reveals\nthe helical spin texture of the inner P1Fermi contour, the opposite spin polarization of the\nP2hole pocket, and finally, the same spin chirality for weaker P3pocket. We recover from\nexperiments the symmetry property imposing the orthogonality between the spin direction\nandtheverticalsymmetryplanes \u001b\u0000M\nVcontainingthe \u0000Mlines. Suchpropertyremainspartly\ntrue concerning the in-plane spin components for the \u0000K directions even if the corresponding\nvertical planes \u001b\u0000K\nVdo not represent perfect symmetry operators. The lack of symmetry for\n\u001b\u0000K\nVleads to the warping term responsible for the appearance of a \u001bzcomponent [2], not\npresently discussed.\nWe now compare the S1andS2surface state spin-texture with a TB model. Calculations\nare implemented by considering relaxed bulk lattice parameters following the Vegard’s law\n(refer to Ref. [18] and Methods). The electronic band structure and energy band dispersion\nalongM\u0000M are plotted in the Suppl. Mat. ??showing a good agreement with ARPES\ndata. The \u001bys-DOS at the Fermi surface, originating from S1andS2, and projected on\nthe top surface (first BL, n= 1) is displayed in Fig. 1d. The color code represents the \u001by-\nDOSN\u001by\nDOSprojected onto the first BL. Around \u0000, positive (negative) values are observed\nfor positive (negative) kx. The sign of this Rashba field is opposite around the M point\n(close to 0.8 Å\u00001). Two electron ( P1,P3) and one hole ( P2) pockets emerge, in very good\nagreementagainwith(S)ARPESdata. ByintroducingtheRashbasurfacepotentialimposed\nby the surface symmetry breaking, we are able to reproduce the spin-resolved map over the\nprojected 2D-Brillouin onto the first BiSb BL (see Methods).\nII. ULTRAFAST SPIN-CHARGE CONVERSION IN Bi 1\u0000xSbx/Co PROBED BY\nTHz-TDS EMISSION SPECTROSCOPY.\nWith the clear demonstration of the spin resolved surface states, we now discuss the\ndynamical spin-charge conversion in Bi 1\u0000xSbxcapped by a thin Co layer as a spin-injector.\nTHz emission spectroscopy in the time domain (THz-TDS) has recently emerged as a pow-\nerful spectroscopic technique to investigate ultrafast SCC in materials with strong SOC [24]\nand, in particular, in TI/FM structures [25–27, 31]. The thin FM layer is excited by a\nfemtosecond laser pump leading to ultrafast demagnetization. This generates both spin\n5-0.10.00.10.2\n-\n0.20.00.20.4\n-\n0.2-0.10.00.10.20.30.4-0.10.00.1\n-0.20.00.20.4-0.10.00.1\n0\n.00.20.40.60.8-0.4-0.3-0.2-0.10.0ky (A-1)(a)M\n(d)P\n1P2P3\nky (A-1)k\nx (A-1)(e)0°\n°P\n1P2P3ky (A-1)(c)°\nk\nx (A-1)Binding energy (eV)(b)0 °\n° S1S2ΓFigure 1.Spin-resolved surface states of a 5-nm thick Bi 0:85Sb0:15film grown on Si(111).\n(a) High-resolution ARPES map at the Fermi energy (integrated over 20 meV). (b) ARPES energy\ndispersion along the \u0000M direction ( kxdirection). As a guide to the eye, red and blue lines underline\ntheS1andS2surface states. All ARPES measurements were performed at 20 K. (c) \u001bypolarization\nDOS measured at the Fermi level, and (d) corresponding TB modelling of the s-DOS projected on\nthe first BL. The experiment is performed at room temperature, while the calculations are at T= 0.\nThe color bar indicates the spin polarization \u001bybetween -1 and +1 in (d). The color scale in (c) is\nproportional to the polarization with scale extrema of P\u0001S=\u00060:1. (e) Color map representing the\nmeasured ARPES intensity (arb. units, proportional to the DOS at the Fermi level) close to Fermi\nenergy integrated on 25 meV with arrows representing the measured spin polarization direction and\namplitude at room temperature. The different electron and hole pockets labeled P1,P2andP3in\n(d) are easily identified in the experimental measurements.\ndensity ^\u0016sand spin-currentJsdiffusing toward the BiSb/Co interface owing to the two spin\npopulations and mobilities introduced by the sp-band spin-splitting in Co [33]. The spin is\nafterwards converted into a transverse charge flow Jcon sub-picosecond timescales leading\nto a THz transient emission that is directly probed in the time-domain (see Refs. [24, 33]\nand Methods). Two SCC mechanisms can contribute: the IREE from the surface states and\nthe ISHE from the bulk. The THz electric field can be expressed as:\n6Figure 2. SCC and THz emission from Bi 1\u0000xSbx/Co bilayers. (a) THz time-\ntrace from Bi 0:85Sb0:15(5)/Co(2)/AlO x(3) for\u0006Bcompared to Co(2)/Pt(4) (grown on high\nresistivity Si substrates). THz phase reversal is a signature of SCC-mediated THz emis-\nsion. (b) Spectral components of the THz emission from Bi 0:85Sb0:15(5)/Co(2)/AlO x(3) and\nCo(2)/Pt(4) for +B. Inset) Normalized THz efficiency \u0011THzdependence on the azimuthal\nangle\u001efor Bi 0:85Sb0:15(5)/Co(2)/AlO x(3). (c) THz efficiency \u0011THzas a function of the\nBi0:79Sb0:21layer thickness (2.5, 5 and 15 nm) compared to Co(2)/Pt(4) (high resistivity Si sub-\nstrates). (d) THz signals from Bi 0:79Sb0:21(15)/Al(5)/Co(4)/AlO x(3), Bi 0:79Sb0:21(15)/AlO x(3),\nBi0:79Sb0:21(15)/Co(4)/AlO x(3)onBaF 2(Suppl.Mat. ??)andCo(2)/Pt(4). Timetracesareshifted\nin time for clarity. Inset) Fluorescence map obtained from a TEM cross-section for Bi 0:79Sb0:21\n(15)/Co(4)/AlOx(3)(brownframe)andBi 0:79Sb0:21(15)/Al(5)/Co(4)/AlOx(3)(greenframe)grown\non BaF 2for the elements and the color code given below the maps.\n78\n>>><\n>>>:EISHE\nTHz(\u0012)/@JISHE\nc\n@t/i!\u0012SHE\u0012\nJs\u0002M(\u0012)\njMj\u0013\nEIREE\nTHz(\u0012)/@JIREE\nc\n@t/i!\u0003IREE\u0012\n^\u0016s\u0002M(\u0012)\njMj\u0013 (1)\nwhere M(\u0012)is the Co layer magnetization vector, ^\u0016sis the spin-accumulation vector relaxing\nonS1andS2and!isthefrequency. M(\u0012)canbecontrolledbyanexternalin-planemagnetic\nfieldBat an angle \u0012from theyaxis. In the above equation, \u0012SHE=Jc=Jsis the spin Hall\nangle scaling the bulk ISHE-mediated SCC, whereas \u0003IREEis the Rashba-Edelstein length\nscaling the IREE from the surface states (refer to Methods and Suppl. Mat. ??).\nOn Fig. 2a, we report the THz signal acquired in reflection geometry from\nBi0:85Sb0:15(5)/Co(2)sample(numbersinparenthesisarethicknessesinnm)atroomtemper-\nature with a saturating in-plane magnetic field B'\u0006100mT insuring that the magnetiza-\ntion follows the external field within much less than a degree. It is compared to the emission\nfrom our optimized Co(2)/Pt(4) metallic ISHE-type sample. In both cases, we observe a\nshort picosecond THz pulse with some minor oscillations within a 3 ps wide envelope. The\nTHz signal phase changes sign when reversing the magnetic field, in full agreement with\nSCC-mediated THz emission (Suppl. Mat. ??). We note several differences: first, the am-\nplitude from the BiSb/Co layer is 1.5 times larger than the Co/Pt revealing the large SCC\nefficiency in BiSb. For the same magnetic field orientation, the BiSb/Co THz signal phase\nis opposite to the one found in the Co/Pt sample (Fig. 2a): this sign inversion is related to\nthe inverted layer stacking of the FM and non magnetic layer, giving an opposite phase and\nindicating that Pt and BiSb share the same conversion sign.\nWe also report in Fig. 2b the THz spectra (Fourier transform) obtained from Co(2)/Pt(4)\nand Bi 1\u0000xSbx(5)/Co(2), demonstrating a relative power enhancement by a factor '2:3\nfor the Bi 1\u0000xSbx/Co sample (field enhancement by a factor \u00021.5). The THz amplitude\nof Bi 1\u0000xSbx/Co is also shown to scale linearly with the pump fluence (Suppl. Mat. ??),\nmeasured up to a few tens of µJ.cm\u00002. The azimuthal angular dependence of the THz\nemission obtained by rotating the sample in the plane by an angle \u001e, while keeping fixed the\nmagnetic field, is shown in the inset of Fig. 2b. The emission is almost isotropic revealing a\npure SCC phenomenon with no evidence of non-linear optical effects such as shift or surge\ncurrent contributions ( e.g.photon drag, photogalvanic effects, etc.). This is in contrast\nwith the recent report given by Park et al.[30] where additional but small non-magnetic\n8contributions were observed. At normal incidence of the optical pump pulse, a pure isotropic\nTHz response vs.\u001eis expected from the linear response theory for both ISHE and IREE\nscenario (Suppl. Mat. ??) as discussed here. The role of the capping layer (metallic Au or\noxidizedAl,AlO x)hasbeencarefullyexcludedbycontrolexperiments(Suppl.Mat. ??). Two\nadditional control samples were grown: i) a Bi 0:79Sb0:21(15nm) sample free of ferromagnetic\nCo and only capped with naturally oxidized AlO x(3nm) almost emitting no THz radiation\n(Fig. 2d) and ii) a sample with a 5 nm thick Al metallic spacer inserted between BiSb and\nCo. The Al insertion strongly reduces the THz emission. It may be explained either by a\nlarger near-infrared (NIR) and THz absorption in Al, or by interfacial spin-loss or weakening\nof the surface SCC induced by the strong degradation of the interface quality. Indeed, TEM\npictures displayed in the inset of Fig. 2d reveal a strong intermixing of the Al interlayer that\nmay induce a loss of an efficient spin-injection and/or the alteration of the surface states\n(more details in Suppl. Mat. ??).\nTo get a better insight into the SCC mechanism, the thickness dependence of the THz\nefficiency,\u0011THz, is displayed in Fig. 2c for the Bi 0:79Sb0:21series (2.5, 5 and 15 nm thick\nlayers).\u0011THzrepresents the spin-injection and conversion efficiencies and is obtained by\nwithdrawing the NIR and THz absorptions in the heterostructure following the procedure\nproposed in Refs. [34–36] (Suppl. Mat. ??). Strikingly, \u0011THzfor BiSb remains constant for\nthe whole thickness series where the BiSb bandgap widens dramatically by several hundreds\nof meV as the thickness is reduced, as calculated previously [18]. Moreover, in agreement\nwith our ARPES results [18], this demonstrates the absence of coupling between the top and\nbottom surface states [37] down to 2.5 nm, unlike previously argued in the case of sputtered\nBiSb materials [28]. The characteristic evanescence length is indeed ultrashort, typically 2\nBL (0.8 nm) near \u0000as confirmed by density functional theory [38] for pure Bi and confirmed\nby our TB calculation (Suppl. Mat. ??). Such behavior is therefore more in favor of an\ninterfacial origin of the SCC and thus strongly hints towards IREE from the surface states\nrather than ISHE from bulk.\nFurthermore, when considering the possible contribution of bulkISHE, BiSb thickness\nhas to be compared with its spin diffusion length. It has been evaluated in bulk BiSb at\n\u0015sf'8 nm by Sharma et al.[28], which is much larger than 2.5 nm. If bulk states were\nto contribute viaISHE, the spin current would therefore flow across the whole layer depth,\nand upper and lower interfaces would contribute similarly to the SCC but with an opposite\n9Figure 3. Sb content dependence of the THz efficiency \u0011THzand calculated IREE re-\nsponse for \u001cs= 10fs.(a)\u0011THzfrom Bi 1\u0000xSbx(5)/Co(2-3)/AlO x(3) withx=0.07, 0.15 and 0.3.\n(b)\u0011THzfrom Bi 1\u0000xSbx(15)/Co(4)/Au(4-6) with x=0.1, 0.21, 0.3 and 0.4. (c-d) Values of the IREE\nlength \u0003IREE\nxyillustrating the conversion efficiency vs.energy\"for (c) Bi 0:79Sb0:21layers with thick-\nnesses2.5, 5and15nmand(d)Bi 1\u0000xSbx(5nm)forx=0.07, 0.15and0.3. Theinsetpresentsthefull\nIREE response outside the bandgap in the bulk valence and conduction bands for Bi 0:79Sb0:21(5nm).\nThe calculations have been performed on for a spin-relaxation time \u001cs=\u001c= 10fs after integration\nof the bands from to the top BL to the middle of the layer.\nsign. The net charge current would drop to zero, as tanh\u0010\ntTI\n\u0015sf\u0011\ntanh\u0010\ntTI\n2\u0015sf\u0011\n/t2\nTI\n2\u00152\nsf[39] for\nsmall TI thicknesses tTIwhen considering the multiple spin current reflections at the TI\ninterfaces [36]. This is in contrast with the thickness-independent SCC observed here, in the\nultrathin limit. We thus anticipate that the bulk states are hardly involved, and that the\n10surface states at the FM/TI interface are mainly responsible for a net charge current through\nIREE. This conclusion is also supported by i) the large increase of the surface state DOS\nof BiSb at the interface compared to the DOS bulk states as the BiSb thickness increases\n(see Suppl. Mat. ??) and by ii) the assumption of the preservation of BiSb surface states in\nexchange contact with Co as it seems to be the case.\nWe now focus on the THz efficiency \u0011THz vs.the Sb content x, plotted for\nBi1\u0000xSbx(5)/Co(2)/AlO x(3) in Fig. 3a and for Bi 1\u0000xSbx(15)/Co(4)/Au in Fig. 3b\n(Suppl. Mat. ??). In each series, a large \u0011THzis measured, comparable to that of Co/Pt\nwith a maximum for x= 0:3for the 15 nm series, i.e.at the limit of the topological phase\ndiagram [20]. Importantly, \u0011THzremains similar for all the samples for the tTI= 5nm series\n(Fig. 3a), highlighting again the prominent role of the surface states in the SCC. It also\nemphasizes the robustness of the IREE over a wide range of Sb content. The persistence of\na large signal outside the bulk topological window ( x= 0:4for 15 nm thickness, in Fig. 3b)\nindicates that bulk states may start to contribute by shunting a part of the spin current.\nIII. SCC AND IREE TENSOR FROM LINEAR RESPONSE THEORY\nWe now compare the enhanced THz emission observed on ultrathin BiSb films to that\nobtained from the linear response theory described by the IREE response, namely \u0003IREE\nxy\nmatching with the so-called inverse Edelstein length (see Suppl. Mat. ??and Methods).\nWe consider an extended formalism to the one recently developed to address the direct\nREE response as given in Refs. [40–42]. We evaluate here the intraband response to an\nongoing spin-current relaxing onto the Fermi surface generating, viaan out-of-equilibrium\nspin-density, a charge current according to: Jx\nc= \u0003IREE\nxyJy\ns;zwith \u0003IREE\nxy =P\nn;k\u0010\n\u001by\nnkvx\nnk\u001cs@fnk\n@\"\u0011\nP\nn;k\u0010@fnk\n@\"\u0011\nwhereyis the direction of the spin injected with a flow along zwhereasxis the direction of\nthein-planechargecurrent. Intheaboveexpression, nisthebandindex, vx\nnkthecorrespond-\ning band velocity along xand\u001csis the typical (spin) relaxation time assumed to be constant\nonto the Fermi surface. fnkis the occupation number for the band nand wavevector k\nwhereas\u0000@fnk\n@\"=NDOS(\";k;n)represents the local DOS in the kspace.\nCalculations are performed on bare BiSb bilayers (BLs) free of any Co overlayer as in-\nvestigated by SARPES. We show, on Fig. 3c, the energy dependence of \u0003IREE\nxyobtained\nfor Bi 0:79Sb0:21of different film thicknesses: 2.5, 5 and 15 nm (see Methods). The Fermi\n11level position corresponds to \"= 0. One observes that \u0003IREE\nxyis largely enhanced in the\nbandgap region in the (-0.2 - 0.2) eV window where the S2surface state and S1TSS are lo-\ncated, however without being able to differentiate their individual contributions (see Fig. ??\nfor the corresponding DOS in the Suppl. Mat. ??).\u0003IREE\nxyremains constant vs.BiSb film\nthickness in this window range, as experimentally observed in THz data of Fig. 2c. Last,\n\u0003IREE\nxyis maximum at the Fermi level, where it reaches values in the range of the equiva-\nlent SCC efficiency of Pt, product of the spin-Hall angle by the spin-diffusion length \u0015sfas\n\u0012SHE\u0002\u0015sf\u00190:2\u00000:3nm for the same spin or momentum relaxation time ( 10fs). Increas-\ning the spin-relaxation onto the surface states to 30 fs would yield \u0003IREE\nxy\u00191nm. \u0003IREE\nxy\nis displayed in Fig. 3c as a function of the Sb content xfor the 5 nm series. Although one\nobserves a slight increase of the SCC response from x=0.1 to 0.3 in the gap window (still in\nline with the presence of the surface states, and possibly indicating a volume contribution),\none may conclude that the SCC remains roughly constant at the vicinity of the Fermi en-\nergy (\"=\"F= 0). This is observed experimentally from THz-TDS measurements, and thus\nsuggests an interfacial IREE nature of the SCC, at least for the thinner films. One cannot\ntotally rule out a certain ISHE contribution arising from the bulk propagating bands asso-\nciated to a very short spin-diffusion length. Nonetheless, the strong impedance mismatch\nand subsequent spin-backflow at Co/BiSb interface would be strongly in disfavor of such\nscenario.\nCONCLUSIONS\nAlthough less investigated compared to other Bi-based families because of its modest\nbandgap, BiSb ultrathin films still exhibit very robust surface states, as revealed by our\nspin-resolved ARPES measurements. This is in part related to the confinement effects with-\nout being detrimental to the surface states. In particular, the presence of a topological\nsurface state S1in a wide Sb-composition and thickness ranges is clearly observed, dis-\nplaying helical spin texture at the Fermi surface with a specific opposite chirality between\nthe electron pocket near the \u0016\u0000point and the six hole pockets away from \u0016\u0000.Viaultrafast\nTHz emission spectroscopy, we demonstrate that the complex spin-texture gives rise to a\nvery efficient spin-charge conversion, resulting from the spin-injection from a Co overlayer\nexcited by an ultrashort laser pulse, mainly occurring viainverse Rashba-Edelstein effect\n12(IREE) owing to the strong localization of the surface states. Our results address the role\nof spin-textured hybridized Rashba-like surface states offering unprecedented SCC efficiency\ndespite the breaking of the TRS symmetry due to the local exchange interactions imposed\nby the magnetic contact. These results hold promise for efficient and integrated structures\nbased on BiSb. Future investigations will concern the dynamics of the spin relaxation onto\nthe BiSb surface states excited by ultrashort pulses.\nACKNOWLEDGMENTS\nWe thank T. Kampfrath and G. Bierhance (Freie Universität Berlin) for very fruitful\ndiscussions. This work was supported by a grant overseen by the French National Research\nAgency (ANR) as part of the “Generic Project Call - 2021” Programme (ANR-21-CE24-\n0011 TRAPIST), public grant overseen by the French National Research Agency (ANR)\nas part of the “Investissements d’Avenir” program (Labex NanoSaclay, reference: ANR-10-\nLABX-0035 SPICY) and the program ESR/EquipEx+ (Grant No. ANR-21-ESRE-0025).\nWe acknowledge financial support from the Horizon 2020 Framework Programme of the\nEuropean Commission under FET-Open Grant No. 863155 (s-Nebula) and Grant No. 64735\n(Extreme-IR).\nAUTHOR CONTRIBUTIONS\nJ.-M.G., H.J., P.LF., S.D. and A.L. conceived and designed the experiment. J.-M.G.\nsupervised the project. L.B., D.S., Ma.Mo., A.L., N.R., F.B., P.LF. and J.-M.G. grew\nthe samples and performed ARPES and SARPES experiments at Synchrotron Soleil. G.P,\nL.B. and A.L. performed TEM cross-section and SR-TEM microscopy. E.R, S.D., Ma.Mi.,\nJu.M., J.T. performed THz-TDS experiments at room temperature. H.J. performed TB\ncalculations. E.R., P.LF., S.D., A.L., R.L., N. R., H.J, J.-M.G. analysed the data. E.R.,\nP.LF., S.D., A.L., N. R., R.L., H.J., J.-M.G. wrote the paper.\n13METHODS\nMBE growth of Bi 1\u0000xSbx.The Bi 1\u0000xSbxsamples were grown by Molecular Beam\nEpitaxy (MBE). A Si(111) substrate was annealed in ultra-high vacuum (UHV) at 1370 K\nin order to obtain a 7 \u00027 surface reconstruction, observed by reflection high energy electron\ndiffraction (RHEED). The Bi 1\u0000xSbxalloy is grown by co-deposition from two Knudsen cells.\nEach cell flux is calibrated prior to the deposition using a quartz microbalance and the\nrelative deposition rates are directly used to estimate the Sb-concentration x. RHEED mea-\nsurements are performed throughout the deposition. Up to 5 nm, we observe a 2D-growth;\nfor thicker films, we stop the evaporation at 7 nm for a 10 min intermediate annealing at\n500 K before completing the film growth up to the targeted thickness. This procedure is\nfurther described in Ref. [18]. BiSb films with thicknesses from 2.5 to 15 nm and with var-\nious compositions (0.03< x<0.3) were grown using this method and their crystallographic\nquality was demonstrated by RHEED, X-ray-diffraction or Scanning Transmission Electron\nMicroscopy (STEM) [18].\nAll our samples were grown in the MBE chamber of the CASSIOPEE beamline installed\non the SOLEIL synchrotron. After their elaboration, they can be transferred in UHV to\nan ARPES (Angle-Resolved PhotoEmission Spectroscopy) or a SARPES (Spin-Resolved\nARPES) experiments where we can characterize their electronic structure. After photoe-\nmission measurements, some of the sample were transferred again into the MBE chamber\nfor Co electron beam deposition. These bilayer systems were used for SCC measurements\nusing THz emission spectroscopy.\nSEM - FIB milling and EDX. Lamellae for STEM observation were prepared from\nthe sample using Focused Ion Beam (FIB) ion milling and thinning. Prior to FIB ion\nmilling, the sample surface was coated with 50 nm of carbon to protect the surface from the\nplatinum mask deposited used for the ion milling process. Ion milling and thinning were\ncarried out in a FEI SCIOS dual-beam FIB-SEM. Initial etching was performed at 30 keV,\nand final polishing was performed at 5 keV. The lamellae were prepared following the two\ndifferent zone axis ( h110iandh112i) of the BaF 2substrate. All samples were observed in an\naberration-corrected FEI TITAN 200 TEM-STEM operating at 200 keV. The convergence\nhalf-angle of the probe was 17.6 mrad and the detection inner and outer half-angles for\nHAADF-STEM were 69 mrad and 200 mrad, respectively. All micrographs where 2048 by\n142048 pixels. The dwell time was 8 µs and the total acquisition time 41 s. EDX measurements\nwere performed in the Titan microscope featuring the Chemistem system, that uses a Bruker\nwindowless Super-X four-quadrant detector and has a collection angle of 0.8 sr.\n(Spin-)Angular-resolved photoemission spectroscopy. The photoemission experi-\nments were performed on the CASSIOPEE beamline installed on the SOLEIL storage ring\n(Saint-Aubin, France). The beamline hosts two endstations. A high-resolution ARPES end-\nstation, which was used in this work for the measurement of the Fermi surface and of the\nband dispersion, using 20 eV incident photons with a linear horizontal polarization. It is\nequipped with a Scienta R4000 electron analyzer. The photon spot size on the sample is\nof the order of 50 \u000250µm2and the overall kinetic energy resolution (taking into account\nboth the photon energy and the electron kinetic energy resolutions) was of the order of 10\nmeV. The second endstation is a spin-resolved ARPES experiment, where the beam size\nis around 300\u0002300µm2. It is equipped with a MBS A1-analyzer with a 2D detector for\nARPES measurements. Close to this 2D detector, a 1 \u00021 mm2hole collects photoelectron\nwith well-defined kinetic energy and momentum. They are sent into a spin manipulator\nable to orient any spin component along the magnetization axis of a FERRUM VLEED\nspin-detector, made of a Fe(100)-p(1 \u00021)O surface [43, 44] deposited on a W-substrate.\nThe spin polarization along the selected direction is proportional to the difference of\nthe two signals collected for opposite magnetizations of the Fe-oxide target. To reduce as\nmuch as possible the measurement asymmetries stemming from the instrument ( i.e., not\ndue to the spin polarization), four measurements per polarization direction are acquired,\nreversing both the ferrum magnetization direction, and the electron spin direction. This four\nmeasurements are combined into a geometrical average. The polarization is then determined\nbyP=S\u00001(I\u001b\n+\u0000I\u001b\n\u0000)=(I\u001b\n++I\u001b\n\u0000).\nThe 1\u00021 mm2hole introduces an integration on both the kinetic energy and the wave\nvector. For the kinetic energy, it corresponds to 0.23% of the used pass energy (10 eV in our\ncase), so around 23 meV. Convoluted with the energy resolution of the analyzer (10 meV for\nthis pass energy and an entrance slit of 400 µm), it gives an overall kinetic energy resolution\nof 25 meV. For the wave vector, the 1 mm aperture corresponds to an integration on 4%\nof the total (30\u000e) angular range, which gives 1.2\u000e. At 20 eV photon energy, for electrons at\nthe Fermi level, this gives a k-resolution of around 0.048 Å\u00001. This explains the relatively\n15broad features of the SARPES in Fig. 1c,e compared to Fig. 1a.\nThe analyzer optics is movable and can collect electrons in a large 2D (30\u000e\u000230\u000e) angular\nrange. To map the spin texture at the Fermi level, the analyzer is set to the appropriate\nkinetic energy while the optics is moved by 0.2\u000estep along two X and Y perpendicular\ndirections. The two in-plane spin components are measured at each step.\nTHz emission spectroscopy. Ultrafast near-infrared (NIR) pulses ( '100 fs) centered\nat\u0015NIR=810 nm are derived from a Ti:Sapphire oscillator to photo-excite the spin carriers\ndirectly from the front surface (Co side). Average powers of up to 600 mW were used with a\nrepetition rate of 80 MHz (the energy per pulse is around 3 nJ). The typical laser spot size\non the sample was about 200 µm\u0002200µm. The optical pump is initially linearly polarized\nand irradiate the TI/FM heterostructure under normal incidence. The generated THz pulses\nwere also collected from the front surface of the samples ( i.e., reflection geometry) using a\nset of parabolic mirrors of 150 and 75 mm focal length to focus on the detection crystal. The\nsamples were placed on a mount with a small magnetic field (around 100 mT) in the plane of\nthe thin films. Both the sample orientation (angle \u001e) and the in-plane magnetic field (angle\n\u0012) can be independently rotated in the sample plane. Standard electro-optic sampling was\nused to detect the electric field of the THz pulses, using a 500 µm-thickh1 1 0iZnTe crystal.\nA chopper was placed at the focal point between the second and third parabolic mirror to\nmodulate the THz beam at 6 kHz for heterodyne lock-in detection. A mechanical delay line\nwas used to sample the THz ultrafast pulse as a function of time. The THz propagation\npath was enclosed in a dry-atmosphere purged chamber (typically <2% humidity) to reduce\nwater absorption of the THz radiation.\nTight-binding calculations of Bi 1\u0000xSbxmultilayers. We have developed a tight\nbinding (TB) model in order to describe the Bi 1\u0000xSbxelectronic band structure as well as\ntheir surface topological properties [45]. This approach is indeed well suited for TI and gives\na fair description of the surface state spin texture in close agreement with the one derived\nfrom Density Functional Theory (DFT) developed for pure Bi surfaces [21, 22, 46]. The\nrhombohedral A7 structure is described by two atoms per unit cell, forming then a bilayer\n(BL) of thickness of about 0.4 nm. Bi 1\u0000xSbxslabs are obtained by stacking the BL along\nthe (111) direction ( zaxis) with two different plane-to-plane distances. We constructed our\nHamiltonian on the basis of the work of Ref. [45] using the generalization of the sp3TB-\n16model Hamiltonian proposed for bulk Bi and Sb crystals [47], adapted to Bi 1\u0000xSbxalloys [15]\nand complemented by the introduction of additional surface potential terms when dealing\nwith thin layers (treatment in slabs) [45, 48, 49]. In particular, the hopping parameters for\nthe BiSb alloys are obtained by using the virtual crystal approximation (VCA) according\nto [15]:\nVBiSb\nC=x VSb\nC+ (1\u0000x2)VBi\nC (2)\nwherexis the antimony content and VSb\nCandVBi\nCare the respective hopping parameters of\nSb and Bi taken from Ref. [47]. One notes ^\u001b\u000bthe spin index on each atom where \u000bstands\nfor the directional index. The hopping terms among the atomic orbitals are decomposed\ninto inter- and intra-BL hopping terms. The inter-BL off-diagonal hopping term between\natoms (plane) 1 and atoms (plane) 2 consists of the nearest-neighbor coupling in the bulk\nBiSb Hamiltonian, whereas the intra-BL hopping term consists of two parts which represents\nrespectively the third and second nearest neighbor contributions. We considered the overall\nTB Hamiltonian according to:\n^H=^H0+^HSO+^H\r (3)\nwhere ^H0= \u0006j\u0017\ni\u0016ji\u0016iVj\u0017\ni\u0016hj\u0017jrepresents the hopping Hamiltonian ( i,jare the atomic posi-\ntions,\u0016,\u0017are the orbitals), ^HSO=~\n4m2c2\u0010\u0000 !rV(r)\u0002^p\u0011\n\u0001^\u001bthe SOC term and ^H\rthe Rashba\nsurface potential induced by the deformation of the surface orbitals due to the local electric\nfield. Indeed, due to the symmetry breaking at the surface, a Rashba SOC term must be\ntaken into account in the Hamiltonian at the two surface planes. We model such effect for\nthesp3basis by using the approach of Ast and Gierz [49] for ^H\rconsidering two additional\nsurface hopping terms \rspand\rppacting respectively between the s\u0000pzandpx\u0000pz(or\npy\u0000pz) surface orbitals. We thus add the ^H\rHamiltonian term of the form:\n^H\r=8\n>>>>><\n>>>>>:\rsp (i;i)\u0011(s;pz)\n\u0006\rsp1 (i;j)\u0011(s;pz)\n\u0006\rppcos(\u0012) (i;j)\u0011(px;pz)\n\u0006\rppsin(\u0012) (i;j)\u0011(py;pz)(4)\n17where the + (-) sign corresponds to the uppermost (lowermost) atomic plane and \u0012is the\nangle between the direction joining the two atoms considered and the x-direction. We then\nrestrained ourselves to the in-plane surface hopping as for a pure 2D system. The best\nagreement with ARPES results is found by adding, as proposed in Ref. [49], additional\non-sites\u0000pzcoupling\rsp=\u00000:2eV, and surface hopping terms \rsp1=0.3 eV and \rpp=-\n0.6 eV forx=0.15 slightly departing from the values given for pure Bi, i.e.\rsp=0.45 eV and\n\rpp=-0.27 eV [45], with opposite sign for the top and bottom surfaces due to the opposite\ndirection of the potential gradient. We emphasize that this surface terms are required to\ncorrectly reproduce the surface state dispersion as observed by ARPES experiments. The\nsize of the Hamiltonian ^H(kx;ky)to diagonalize is 16N\u000216NwhereNis the number of\nbilayers (BLs). Once the Green function of the multilayer system is defined as:\n^G(\";kx;ky) =h\n\"+i\u000e\u0000^H(kx;ky)i\u00001\n(5)\nthe partial density of state (DOS) NDOS(\")vs.the energy \"equalsN(n;\") =\n\u0000(1=\u0019)ImTr[^G(\";n;kx;ky)]whereas the spin density of states (spin-DOS) with spin along\nthe\u000bdirection is s\u000b(\") = (1=\u0019)ImTr[^\u001b\u000b^G(\";kx;ky)].\u000eis the typical energy broadening ( '\n10 meV) and the trace ( Tr) is applied over the considered sp3orbitals on a given BL index\n(n2[1;N]). The energy zero ( \"=0) refers to the Fermi level position.\nThe modelling of IREE were performed by TB method. We have summed the contribu-\ntions from the different Fermi surface pockets within the 2D-BZ after i) having introduced\nthe Rashba potentials at the BiSb surfaces, required to match both the TSS electronic dis-\npersion and SML measured in (S)ARPES experiments [18]; and after ii) having considered\na same (spin) relaxation time of \u001cs=\u001c0= 10fs involved in the intraband transitions. The\ninverse Edelstein length has been then evaluated on the whole Fermi surface according to the\nfollowing expression \u0003xy(\") =P\nnR\nd2kh nkj^vx\u001c0j nkih nkj^\u001byj nkiNDOS(\";k;n)\u001cP\nnR\nd2kNDOS(\";k;n)(see Suppl. Mat. ??).\nThe typical value of \u001cs= 10fs corresponds to an energy broadening \u0000 =~=(2\u001cs)'\n50meV) and a Fermi velocity of about 5\u0002105m.s\u00001like extracted from ARPES data. The\ntrace on\u0014xyis performed by summing the BL contributions from the top surface down to\nthe middle of the BL, typically matching the typical finite TSS evanescence or extension\n18length.\n[1] M.Z.HasanandC.L.Kane, “ Colloquium : topologicalinsulators,” Reviews ofModernPhysics,\nvol. 82, pp. 3045–3067, Nov. 2010.\n[2] Y. Ando, “Topological Insulator Materials,” JournalofthePhysical SocietyofJapan, vol. 82,\np. 102001, Oct. 2013.\n[3] X.-L. Qi and S.-C. Zhang, “Topological insulators and superconductors,” Reviews ofModern\nPhysics, vol. 83, pp. 1057–1110, Oct. 2011.\n[4] Y. Shiomi, K. Nomura, Y. Kajiwara, K. Eto, M. Novak, K. Segawa, Y. Ando, and E. Saitoh,\n“Spin-Electricity Conversion Induced by Spin Injection into Topological Insulators,” Physical\nReviewLetters, vol. 113, p. 196601, Nov. 2014.\n[5] Y. Ando and M. Shiraishi, “Spin to Charge Interconversion Phenomena in the Interface and\nSurface States,” JournalofthePhysical SocietyofJapan, vol. 86, p. 011001, Jan. 2017.\n[6] W. Han, Y. Otani, and S. Maekawa, “Quantum materials for spin and charge conversion,” npj\nQuantum Materials, vol. 3, p. 27, Dec. 2018.\n[7] V. Edelstein, “Spin polarization of conduction electrons induced by electric current in two-\ndimensional asymmetric electron systems,” SolidStateCommunications, vol. 73, pp. 233–235,\nJan. 1990.\n[8] A. R. Mellnik, J. S. Lee, A. Richardella, J. L. Grab, P. J. Mintun, M. H. Fischer, A. Vaezi,\nA. Manchon, E.-A. Kim, N. Samarth, and D. C. Ralph, “Spin-transfer torque generated by a\ntopological insulator,” Nature, vol. 511, pp. 449–451, Jul 2014.\n[9] S. Zhang and A. Fert, “Conversion between spin and charge currents with topological insula-\ntors,”Physical ReviewB, vol. 94, p. 184423, Nov. 2016.\n[10] I. M. Miron, K. Garello, G. Gaudin, P.-J. Zermatten, M. V. Costache, S. Auffret, S. Bandiera,\nB. Rodmacq, A. Schuhl, and P. Gambardella, “Perpendicular switching of a single ferromag-\nnetic layer induced by in-plane current injection,” Nature, vol. 476, pp. 189–193, Aug. 2011.\n[11] L. Liu, O. J. Lee, T. J. Gudmundsen, D. C. Ralph, and R. A. Buhrman, “Current-Induced\nSwitching of Perpendicularly Magnetized Magnetic Layers Using Spin Torque from the Spin\nHall Effect,” Physical ReviewLetters, vol. 109, p. 096602, Aug. 2012.\n[12] L. Liu, C.-F. Pai, Y. Li, H. W. Tseng, D. C. Ralph, and R. A. Buhrman, “Spin-Torque\n19Switching with the Giant Spin Hall Effect of Tantalum,” Science, vol. 336, pp. 555–558, May\n2012.\n[13] K. Kondou, R. Yoshimi, A. Tsukazaki, Y. Fukuma, J. Matsuno, K. S. Takahashi, M. Kawasaki,\nY. Tokura, and Y. Otani, “Fermi-level-dependent charge-to-spin current conversion by Dirac\nsurface states of topological insulators,” NaturePhysics, vol. 12, pp. 1027–1031, Nov 2016.\n[14] N. H. D. Khang, Y. Ueda, and P. N. Hai, “A conductive topological insulator with large\nspin Hall effect for ultralow power spin–orbit torque switching,” NatureMaterials, vol. 17,\npp. 808–813, Sept. 2018.\n[15] J. C. Y. Teo, L. Fu, and C. L. Kane, “Surface states and topological invariants in three-\ndimensional topological insulators: Application to Bi 1\u0000xSbx,”Phys.Rev.B, vol. 78, p. 045426,\nJul 2008.\n[16] D. Hsieh, D. Qian, L. Wray, Y. Xia, Y. S. Hor, R. J. Cava, and M. Z. Hasan, “A topological\nDirac insulator in a quantum spin Hall phase,” Nature, vol. 452, pp. 970–974, Apr. 2008.\n[17] H. M. Benia, C. Straßer, K. Kern, and C. R. Ast, “Surface band structure of Bi 1\u0000xSbx(111),”\nPhysical ReviewB, vol. 91, p. 161406, Apr. 2015.\n[18] L. Baringthon, T. H. Dang, H. Jaffrès, N. Reyren, J.-M. George, M. Morassi, G. Patriarche,\nA. Lemaitre, F. Bertran, and P. Le Fèvre, “Topological surface states in ultrathin Bi 1\u0000xSbx\nlayers,”Physical ReviewMaterials, vol. 6, p. 074204, July 2022.\n[19] S. Ito, M. Arita, J. Haruyama, B. Feng, W.-C. Chen, H. Namatame, M. Taniguchi, C.-M.\nCheng, G. Bian, S.-J. Tang, T.-C. Chiang, O. Sugino, F. Komori, and I. Matsuda, “Surface-\nstate Coulomb repulsion accelerates a metal-insulator transition in topological semimetal\nnanofilms,” ScienceAdvances, vol. 6, no. 12, p. eaaz5015, 2020.\n[20] B. Lenoir, H. Scherrer, and T. Caillat, “An overview of recent developments for BiSb Alloys,”\ninSemiconductors andSemimetals, vol. 69, pp. 101–137, Elsevier, 2001.\n[21] T. Hirahara, T. Nagao, I. Matsuda, G. Bihlmayer, E. V. Chulkov, Y. M. Koroteev, P. M.\nEchenique, M. Saito, and S. Hasegawa, “Role of Spin-Orbit Coupling and Hybridization Effects\nin the Electronic Structure of Ultrathin Bi Films,” Phys.Rev.Lett., vol. 97, p. 146803, Oct\n2006.\n[22] Y. M. Koroteev, G. Bihlmayer, E. V. Chulkov, and S. Blügel, “First-principles investigation\nof structural and electronic properties of ultrathin Bi films,” Phys.Rev.B, vol. 77, p. 045428,\nJan 2008.\n20[23] I. Aguilera, H.-J. Kim, C. Friedrich, G. Bihlmayer, and S. Blügel, “ Z2topology of bismuth,”\nPhysical ReviewMaterials, vol. 5, p. L091201, Sept. 2021.\n[24] T. Seifert, S. Jaiswal, U. Martens, J. Hannegan, L. Braun, P. Maldonado, F. Freimuth,\nA. Kronenberg, J. Henrizi, I. Radu, E. Beaurepaire, Y. Mokrousov, P. M. Oppeneer, M. Jour-\ndan, G. Jakob, D. Turchinovich, L. M. Hayden, M. Wolf, M. Münzenberg, M. Kläui, and\nT. Kampfrath, “Efficient metallic spintronic emitters of ultrabroadband terahertz radiation,”\nNaturePhotonics, vol. 10, pp. 483–488, Jul 2016.\n[25] X. Wang, L. Cheng, D. Zhu, Y. Wu, M. Chen, Y. Wang, D. Zhao, C. B. Boothroyd, Y. M. Lam,\nJ.-X. Zhu, M. Battiato, J. C. W. Song, H. Yang, and E. E. M. Chia, “Ultrafast Spin-to-Charge\nConversion at the Surface of Topological Insulator Thin Films,” Advanced Materials, vol. 30,\nno. 52, p. 1802356, 2018.\n[26] M. Tong, Y. Hu, Z. Wang, T. Zhou, X. Xie, X. Cheng, and T. Jiang, “Enhanced Terahertz\nRadiation by Efficient Spin-to-Charge Conversion in Rashba-Mediated Dirac Surface States,”\nNanoLetters, vol. 21, no. 1, pp. 60–67, 2021. PMID: 33331788.\n[27] X.Chen,H.Wang,C.Wang,C.Ouyang,G.Wei,T.Nie,W.Zhao,J.Miao,Y.Li,L.Wang,and\nX. Wu, “Efficient Generation and Arbitrary Manipulation of Chiral Terahertz Waves Emitted\nfrom Bi 2Te3–Fe Heterostructures,” Advanced Photonics Research, vol. 2, no. 4, p. 2000099,\n2021.\n[28] V. Sharma, W. Wu, P. Bajracharya, D. Q. To, A. Johnson, A. Janotti, G. W. Bryant, L. Gund-\nlach, M. B. Jungfleisch, and R. C. Budhani, “Light and microwave driven spin pumping across\nFeGaB–BiSb interface,” Phys.Rev.Materials, vol. 5, p. 124410, Dec 2021.\n[29] H. Park, K. Jeong, I. Maeng, K. I. Sim, S. Pathak, J. Kim, S.-B. Hong, T. S. Jung, C. Kang,\nJ. H. Kim, J. Hong, and M.-H. Cho, “Enhanced Spin-to-Charge Conversion Efficiency in\nUltrathin Bi 2Se3Observed by Spintronic Terahertz Spectroscopy,” ACSApplied Materials &\nInterfaces, vol. 13, no. 19, pp. 23153–23160, 2021. PMID: 33945256.\n[30] H. Park, S. Rho, J. Kim, H. Kim, D. Kim, C. Kang, and M. Cho, “Topological Surface-\nDominated Spintronic THz Emission in Topologically Nontrivial Bi 1\u0000xSbxFilms,”Advanced\nScience, p. 2200948, May 2022.\n[31] E. Rongione, S. Fragkos, L. Baringthon, J. Hawecker, E. Xenogiannopoulou, P. Tsipas,\nC. Song, M. Mičica, J. Mangeney, J. Tignon, T. Boulier, N. Reyren, R. Lebrun, J.-M. George,\nP. Le Fèvre, S. Dhillon, A. Dimoulas, and H. Jaffrès, “Ultrafast Spin-Charge Conversion at\n21SnBi 2Te4/Co Topological Insulator Interfaces Probed by Terahertz Emission Spectroscopy,”\nAdvanced OpticalMaterials, vol. 10, no. 7, p. 2102061, 2022.\n[32] L.Baringthon, Élaboration et caractérisation de phases topologiques de l’alliage Bi 1\u0000xSbxpour la conversion spin charge.\nPhD thesis, Université Paris-Saclay, 2022.\n[33] T.H.Dang, J.Hawecker, E.Rongione, G.BaezFlores, D.Q.To, J.C.Rojas-Sanchez, H.Nong,\nJ. Mangeney, J. Tignon, F. Godel, S. Collin, P. Seneor, M. Bibes, A. Fert, M. Anane, J.-M.\nGeorge, L. Vila, M. Cosset-Cheneau, D. Dolfi, R. Lebrun, P. Bortolotti, K. Belashchenko,\nS. Dhillon, and H. Jaffrès, “Ultrafast spin-currents and charge conversion at 3 d-5dinter-\nfaces probed by time-domain terahertz spectroscopy,” Applied PhysicsReviews, vol. 7, no. 4,\np. 041409, 2020.\n[34] O.Gueckstock, L.Nádvorník, M.Gradhand, T.S.Seifert, G.Bierhance, R.Rouzegar, M.Wolf,\nM. Vafaee, J. Cramer, M. A. Syskaki, G. Woltersdorf, I. Mertig, G. Jakob, M. Kläui, and\nT.Kampfrath,“TerahertzSpin-to-ChargeConversionbyInterfacialSkewScatteringinMetallic\nBilayers,” Advanced Materials, vol. 33, no. 9, p. 2006281, 2021.\n[35] M. Meinert, B. Gliniors, O. Gueckstock, T. S. Seifert, L. Liensberger, M. Weiler, S. Wim-\nmer, H. Ebert, and T. Kampfrath, “High-Throughput Techniques for Measuring the Spin Hall\nEffect,”Phys.Rev.Applied, vol. 14, p. 064011, Dec 2020.\n[36] J. Hawecker, E. Rongione, A. Markou, S. Krishnia, F. Godel, S. Collin, R. Lebrun, J. Tignon,\nJ. Mangeney, T. Boulier, J.-M. George, C. Felser, H. Jaffrès, and S. Dhillon, “Spintronic THz\nemitters based on transition metals and semi-metals/Pt multilayers,” AppliedPhysicsLetters,\nvol. 120, no. 12, p. 122406, 2022.\n[37] Y. Zhang, K. He, C.-Z. Chang, C.-L. Song, L.-L. Wang, X. Chen, J.-F. Jia, Z. Fang, X. Dai,\nW.-Y. Shan, S.-Q. Shen, Q. Niu, X.-L. Qi, S.-C. Zhang, X.-C. Ma, and Q.-K. Xue, “Crossover\nof the three-dimensional topological insulator Bi 2Se3to the two-dimensional limit,” Nature\nPhysics, vol. 6, pp. 584–588, Aug. 2010.\n[38] H. Ishida, “Decay length of surface-state wave functions on Bi(111),” Journal ofPhysics:\nCondensed Matter, vol. 29, p. 015002, nov 2016.\n[39] J. Sinova, S. O. Valenzuela, J. Wunderlich, C. H. Back, and T. Jungwirth, “Spin Hall effects,”\nRev.Mod.Phys., vol. 87, pp. 1213–1260, Oct 2015.\n[40] J.-C. Rojas-Sánchez and A. Fert, “Compared Efficiencies of Conversions between Charge and\nSpin Current by Spin-Orbit Interactions in Two- and Three-Dimensional Systems,” Phys.Rev.\n22Appl., vol. 11, p. 054049, May 2019.\n[41] A. Johansson, B. Göbel, J. Henk, M. Bibes, and I. Mertig, “Spin and orbital Edelstein effects\nin a two-dimensional electron gas: Theory and application to SrTiO 3interfaces,” Phys.Rev.\nRes., vol. 3, p. 013275, Mar 2021.\n[42] A. Roy, F. T. Cerasoli, A. Jayaraj, K. Tenzin, M. B. Nardelli, and J. Sławińska, “Long-range\ncurrent-induced spin accumulation in chiral crystals,” npjComputational Materials, vol. 8,\np. 243, Nov 2022.\n[43] R. Bertacco and F. Ciccacci, “Oxygen-induced enhancement of the spin-dependent effects in\nelectron spectroscopies of Fe(001),” Phys.Rev.B, vol. 59, pp. 4207–4210, Feb 1999.\n[44] M. Escher, N. B. Weber, M. Merkel, L. Plucinski, and C. M. Schneider, “FERRUM: A New\nHighly Efficient Spin Detector for Electron Spectroscopy,” e-Journal ofSurfaceScienceand\nNanotechnology, vol. 9, pp. 340–343, 2011.\n[45] K. Saito, H. Sawahata, T. Komine, and T. Aono, “Tight-binding theory of surface spin states\non bismuth thin films,” Phys.Rev.B, vol. 93, p. 041301, Jan 2016.\n[46] Y. M. Koroteev, G. Bihlmayer, J. E. Gayone, E. V. Chulkov, S. Blügel, P. M. Echenique, and\nP. Hofmann, “Strong Spin-Orbit Splitting on Bi Surfaces,” Phys.Rev.Lett., vol. 93, p. 046403,\nJul 2004.\n[47] Y. Liu and R. E. Allen, “Electronic structure of the semimetals Bi and Sb,” Phys.Rev.B,\nvol. 52, pp. 1566–1577, Jul 1995.\n[48] L.PetersenandP.Hedegård,“Asimpletight-bindingmodelofspin–orbitsplittingof sp-derived\nsurface states,” SurfaceScience, vol. 459, no. 1, pp. 49–56, 2000.\n[49] C. R. Ast and I. Gierz, “ sp-band tight-binding model for the Bychkov-Rashba effect in a two-\ndimensional electron system including nearest-neighbor contributions from an electric field,”\nPhys.Rev.B, vol. 86, p. 085105, Aug 2012.\n23" }, { "title": "0903.4406v2.p_type_Bi2Se3_for_topological_insulator_and_low_temperature_thermoelectric_applications.pdf", "content": " \n1 \n p-type Bi 2Se3 for topological insula tor and low temperature \nthermoelectric applications \n \nY.S. Hor1, A. Richardella2, P. Roushan2, Y. Xia2, J.G. Checkelsky2, \nA. Yazdani2, M.Z. Hasan2, N.P. Ong2, and R.J. Cava1 \n \n1Department of Chemistry, Princeton University \n2Department of Physics, Princeton University \n \nAbstract \n \nThe growth and elementary properties of p-type Bi 2Se3 single crystals are reported. Based \non a hypothesis about the defect chemistry of Bi 2Se3, the p-type behavior has been induced \nthrough low level substitutions (1% or less) of Ca for Bi. Scanning tunneling microscopy is \nemployed to image the defects and establish their charge. Tunneling and angle resolved \nphotoemission spectra show that the Fermi level has been lowered into the valence band by \nabout 400 meV in Bi 1.98Ca0.02Se3 relative to the n-type material. p-type single crystals with ab \nplane Seebeck coefficients of + 180 µV/K at room temperature ar e reported. These crystals show \na giant anomalous peak in the Seebeck coeffi cient at low temperatures, reaching +120 µVK-1 at 7 \nK, giving them a high thermoelectric power fact or at low temperatures. In addition to its \ninteresting thermoelectric properties, p-type Bi 2Se3 is of substantial interest for studies of \ntechnologies and phenomena proposed for topological insulators. \n \n2 \n Introduction \nBi2Se3 is one of the binary e nd-members of the (Bi,Sb) 2(Te,Se) 3 family of thermoelectric \nmaterials. Decades of work in the chemistry, physics, and processing of these materials has led to complex formulations of compounds and microstr uctures optimized for use as thermoelectrics \nunder various conditions (see e.g. refs. 1 and 2). Due to the superiority of Bi\n2Te3-based materials \nfor these applications, Bi 2Se3, while well studied, has not been subject to the same degree of \nintensive research as has its heavier mass analog. One of the major issues for Bi 2Se3-based \nthermoelectrics has been the di fficulty in making the material p-type. Unlike Bi 2Te3, which can \nsimply be made n- or p-type through variation of the Bi:T e ratio, the defect chemistry in Bi 2Se3 \nis dominated by charged selenium vacancies, wh ich act as electron donors [3], resulting in n-type \nbehavior for virtually all of the reported transport st udies (see e.g. refs. 3-11). p-type behavior for \npure Bi 2Se3 was reported in an early study [9], but neve r since. Modern studies have shown that \np-type behavior is possible when beginning with an n-type host material in the (Bi 2-xSbx)Se 3 \nsolid solution with x = 0.4, and then doping that composition with small amounts of Pb to create \na quaternary p-type material [10]. \nThe present study is motivated by the desire to find a chemically less complex p-type \nBi2Se3-based material to address the recent emergence of Bi 2Se3 as one of the prime candidates \nfor the study of topological su rface states (see e.g. refs. 12-16), and for thermoelectric \napplications. The character and stab ility of the surface states in Bi 2Se3 at room temperature has \nmotivated the suggestion that they may be usef ul for quantum computi ng applications if the \nFermi level can be lowered to the Dirac point through p-type doping of the normally n-type \ncompound [15,17,18]. Therefore high quality p-type crystals are important for both fundamental \nand applied research on Bi 2Se3. \n3 \n Unlike the case for Bi 2Te3, where the chemical similarity of Bi and Te leads to antisite \ndefects as the primary source of carrier doping in binary compounds , there is little tendency for \nBi/Se mixing in Bi 2Se3, and the primary structural defect giving rise to electron doping is doubly \ncharged selenium vacancies V Se•• [3]: \nSeSe → VSe•• + Se(g) + 2e ′ \nOrdinarily, one compensates for the presence of donors in this chemical family through doping \nwith Pb on the Bi site, as Pb has one fewer elect ron than Bi. This substitution does not, however, \nlead to the formation of p-type material for Bi 2Se3. We hypothesize that this is due to the fact \nthat Pb in Bi 2Se3 is ambipolar, much as is the case for Cu [11]. This indicates that a more ionic \nsubstitution on the Bi site may be required for hole-doping of Bi 2Se3, suggesting the use of Ca \nsubstitution, with the defect reaction: 2Ca → 2Ca\nBi′ + 2h• \n Bi 2Se3 \nwith Ca substitution for Bi creating a negatively charged defect (Ca Bi′) that in turn generates \nholes (h•) to compensate the electrons created by the Se vacancies. Our results, described below, \nsupport this scenario as a good representation of the defect chemistry in this compound. \nExperimental \nThe single crystals of Bi 2-xCaxSe3 were grown via a process of two-step melting, starting \nwith mixtures of high purity elements (B i, 99.999 %, Se, 99.999 %, Ca, 99.8 %). First, \nstoichiometric mixtures of Bi and Se were me lted in evacuated quartz ampoules at 800 ºC for 16 \nhours. The melts were then stirred before bein g allowed to solidify by air-quenching to room \ntemperature. Second, the stoichiometric amount of Ca was added in the form of pieces, using \ncare to avoid direct contact of th e added Ca with the quartz. The materials were then heated in \nevacuated quartz ampoules at 400 ºC for 16 hours followed by 800 ºC for a day. The crystal \n4 \n growth process involved cooling from 800 to 550 ºC over a period of 24 hours and then \nannealing at 550 ºC for 3 days. The crystals were then furnace cooled to room temperature. They \nwere cleaved very easily along the basal plane, and were cut into approximately 1.0 × 1.0 × 6.0 \nmm3 rectangular bar samples for the thermal and el ectronic transport measurements. Resistivity, \nSeebeck coefficient, and ther mal conductivity measurements were performed in a Quantum \nDesign PPMS, using the standard four-probe t echnique, with silver paste cured at room \ntemperature used for the contacts. Hall Effect m easurements to determine carrier concentrations \nwere performed at 1.5 K in a home-built apparatus. In all cases, the elect ric- and thermal-currents \nwere applied in the basal plane ( ab plane in the hexagonal setting of the rhombohedral cell) of \nthe crystals. In order to probe the electronic states of nati ve and Ca related defects, Bi 2Se3 and \nBi1.98Ca0.02Se3 samples were studied in a cryogenic scan ning tunneling microscope (STM) at 4.2 \nK. The samples were cleaved in situ in ultrahigh vacuum to expose a pristine surface. The \nsurfaces of the p-type crystals are fully stable in th e STM experiments. High-resolution angle \nresolved photoemission spectrosc opy (ARPES) measurements we re performed using 22-40 eV \nphotons at Beamline-12 of the A dvanced Light Source (Berkeley Laboratory) and on beamline 5 \nat the Stanford Synchrotron Radiation Laborat ory. The energy and momentum resolutions were \n15 meV and 2 % of the surface Brillouin Zone, re spectively, obtained using a Scienta analyzer. \nThe samples were cleaved at 10 K under pressures of less than 5 × 10-11 torr, resulting in shiny \nflat surfaces. The spectra employed are those clos est to the time of cleavage (about 10 minutes), \nwhich are the most repr esentative of the bulk. \nResults \n The rhombohedral crystal structure of Bi 2Se3 consists of hexagonal planes of Bi and Se \nstacked on top of each other along the [001] crysta llographic direction (hexagonal setting), with \n5 \n the atomic order: Se(1)-Bi-Se(2)-Bi-Se(1), wh ere (1) and (2) are refer to different lattice \npositions [19]. The unit cell consists of three of these units stacked on top of each other with \nweak van-der-Waals bonds between Se(1)-Se(1) layers, making the (001) plane the natural \ncleavage plane. Scanning tunneli ng spectroscopy (dI/dV) was perf ormed to measure the density \nof states. The spectra of Bi 2Se3 were n-type, as expected, due to th e presence of Se vacancies \nwith the Fermi energy near the conduction band, c onsistent with previous STM results [20]. In \ncontrast, a clear shift in the position of the Ferm i level towards the valance band occurs in the \nBi1.98Ca0.02Se3 sample, as shown in Fig. 1(a) . The finite density of states inside the gap is due to \nsurface states, for which novel topological properties have been predicted [12-18]. \nWe were able to identify vari ous defects and the si gn of their charge state from the STM \ntopographies of the filled states and unoccupied states. The STM topogr aphies of the native \nBi2Se3 (001) surface are dominated by one type of defect, which a ppears as a bright triangular \nprotrusion in the topographies of the unoccupied stat es (Fig. 1(b)). On aver age, these defects are \nabout 40 Å across, but vary in size between def ects indicating that they are located in various \nlayers beneath the surface. Given that no othe r defects were observed, we attribute these \ntriangular defects to Se vacancies. Figs. 1(c) a nd (d) show the topography of the unoccupied and \noccupied states of the Ca-doped Bi 2Se3, respectively. Comparison between 1(b) and (c) makes it \nclear that the density of tria ngular defects has been reduced significantly in the Ca-doped \nsamples. In addition, the STM topography of the Bi 1.98Ca0.02Se3 (001) surface shows distinct \ndefects that were not present in Bi 2Se3 samples. In topographic imag es of the empty states, the \nshape of these three-fold symmetric defects resemb les a cloverleaf (Fig. 1(c)). Based on their two \ndistinct spatial extents we conclude that they are located at two di fferent crystallographic \npositions; most likely the smaller one is located in a layer nearer the surface and other one in a \n6 \n layer deeper beneath the surface. Because thes e defects occur only in Ca-doped samples, we \nidentify them as Ca-related defects. The charge state of the defects can be inferr ed by observing the bending of the host bands \ncaused by the coulomb field surrounding a charge d defect. A positively charged defect lowers \nthe electronic energy level in its neighboring re gion, leading to a depression area in the STM \ntopography of the filled states, and an enhancement in the topogr aphy of the unoccupied states. \nWe expect this to be observed in imaging the ionized donors, and the opposite effect for imaging \nnegatively charged defects, such as acceptors. The cloverleaf shape defects are surrounded by a \nregion of enhancement in the topography of the fille d states (Fig. 1(d)), implying that they are \nnegatively charged, consistent wi th a Ca acceptor. A comparison between the triangular defect \nclose to the center of the image in Fig. 1(c) and (d) shows a region of depression in the \ntopography of the filled states (c), and hence imp lies the presence of a positively charged defect. \nThis observation is expected for Se vacancies, which are known to be electron donors and in \ntheir ionized state be positively charged. Thus th e STM data support the defect chemistry model \ndescribed above for Ca-doped Bi\n2Se3, with the added observation that the concentration of \ndefects in general is lower in the Ca-doped samples than for native Bi 2Se3. \nThe ARPES data showing valence band energy dispersion curves in the vicinity of the Γ \npoint of the Brillouin Zone for n- and p- type Bi 2Se3 are presented in Figs. 2a and 2b. For the n-\ntype crystal, the valence band is clearly obse rved below E F, as is a small pocket of the \nconduction band near the Γ point. For the p-type crystal, only the vale nce band is observed. This \nchemical potential shift indicates that the Ca doped crystal is hole doped relative to the n-type \nBi2Se3 crystal. In the p-type doped samples, the surface state bands are not observed, indicating \nthat the Fermi level is below the energy of th e Dirac point. Judging from the position of the \n7 \n strongest valence bands, the ener gy dispersion curves through the Γ point show that the chemical \npotential in p-type Bi 1.98Ca0.02Se3 is shifted by approximately 400 meV relative to the n-type \ncrystal (Fig. 2c). Lower hol e doping levels, which would be obtained by tuning the Ca \nconcentration, are expected to lead to materials whose chemical potential is very close to the \nDirac point, and therefore of interest in fundamental and a pplied studies of the topological \nsurface states [15]. \nThe 2–300 K resistivities in the ab plane for undoped Bi 2Se3 and Bi 2-xCaxSe3 crystals for \nx = 0.005, 0.02 and 0.05 are shown in Fig. 3. All show the weakly meta llic resistivities \ncommonly seen in high carrier concentration sma ll band gap semiconductors, with resistivities in \nthe 0.3 to 1.5 m Ω-cm range at temperatures near 10 K. The lightly doped x = 0.005 material is p-\ntype with a carrier concentrati on at room temperature determin ed by Hall effect measurements \n(lower inset, Fig. 3) to be approximately 1 × 1019 cm-3. This crystal has a resistivity ratio, \nρ300/ρ4.2 of about 3, as does the undoped n-type crystal. The upper inset shows the low \ntemperature resistivity region in more detail for n- and p-type crystals. The decreasing resistivity \nseen on cooling is arrested at these low te mperatures and rises slightly below 20-30 K, \nsuggesting that carriers have been fro zen out in this temperature regime. \n The corresponding ab plane Seebeck coefficients are shown in Fig. 4. The room \ntemperature Seebeck coefficient for the n-type undoped Bi 2Se3 crystal, -190 µVK-1, is large \ncompared to those usually reported for this mate rial, which are typically in the -50 to -100 µVK-1 \nrange (3-11). The carrier concentration for this n-type Bi 2Se3 is 8 × 1017 cm-3 (lower inset, Fig. \n2). The magnitude of the Seebeck coefficient decreases smoothl y with temperature until around \n15 K (inset Fig. 4), where it beco mes distinctly more negative in a small peak with an onset \ntemperature roughly corresponding to that of the upturn in the resistivity. A similar peak in \n8 \n Seebeck coefficient has been observed previously in this temperature range in n-type Bi 2Se3 [3]. \nThe p-type materials show similar but more unusua l behavior. The room temperature Seebeck \ncoefficient for the lightly doped p-type material is again very high, reaching approximately +180 \nµVK-1. This value is larger than what is observed in the quaternary p-type (Bi,Sb,Pb) 2Se3 \nmaterials [10] and remains large and p-type over the full temperature range of measurement. \nCrystals with Ca-doping levels as low as Bi 1.9975Ca0.0025Se3 are p-type. A dramatic low \ntemperature peak in the Seebeck coeffi cient (inset Fig. 3) is seen for all p-type compositions \nstudied. The low temperature Seebec k coefficient peaks of -60 µVK-1 and +120 µVK-1 for pure \nand doped Bi 2Se3 at 7 K represent very high values when compared to other small band gap \nsemiconductors at this low temperature. Inte restingly, similarly anomalous low temperature \npeaks have been observed in Bi 1-xSbx alloys, another material in th e vicinity of a Dirac point [12-\n14], in the critical “zero gap” comp osition region near x = 0.07 [21]. \n The total thermal conductivities ( κ) in the ab plane are shown in Fig. 5. They display the \nbehavior typical of high quality crystals, in creasing with decreasing temperature as the phonon \ncontribution to κ grows, until low temperatures, where the phonon mean free path becomes \ncomparable to intra-defect distan ces. The low temperature regime for n- and p-type crystals is \nshown in the inset. The thermal conductivity for the n-type crystal reaches the relatively high \nvalue of about 100 WK-1m-1 at 10 K, and all the Ca-doped materials show a somewhat lower \nmaximum thermal conductivity, in the 40-50 WK-1m-1 range. \n The efficiency of a material for thermo electric cooling or power generation at \ntemperature T is described by the thermoelect ric figure of merit ZT [22]. The materials \nparameter is Z = (S2/ρ)(1/κ), where the power factor, define d as the square of the Seebeck \ncoefficient S divided by the electrical resistiv ity, has been separated from the total thermal \n9 \n conductivity κ. Due to the very high thermal conductiv ities of the single crystals, the Bi 2Se3-\nbased n- and p-type materials are of limited use as practical thermoelectric materials in single \ncrystal form, especially at low temperatures. The resulting low values of the thermoelectric \nfigure of merit are shown in the inset to Fig. 6. The Seebeck coefficients and electrical \nresistivities do, however, suggest that thes e materials may have potential for use as \nthermoelectrics in the very low temperature regime . To illustrate this fact we have plotted the \ntemperature dependences of the power factors in the main pane l of Fig. 6. The anomalous low \ntemperature peaks in all the ma terials, especially for the p-type material, where the power factor \nat 7 K is higher than it is at 300 K, and comparable to that of the n-type material at 300 K, are \nparticularly intriguing for possible low temperatur e thermoelectric applications. The underlying \norigin of these peaks, and also the more subtle anomalies seen in the transport properties in the \n30-40 K range in all material s, is not presently known. \nConclusions \n Through consideration of th e defect chemistry of Bi 2Se3, we have identified Ca as a \ndopant that when present in sub-percent quantities results in the formation of a p-type material. \nThe use of p-type Bi 2Se3 for the fundamental study of topologi cal surface states and for devices \nbased on these states for quantum computing will be of considerable future interest. The dramatic low temperature peaks found in the Seebeck coefficients in Bi\n1-xSbx and Bi 2Se3, \nmaterials that have strong spin orbit coupling and nearby Dirac points [12-16,19], suggests that \nthese characteristics may be related. In addi tion, processing of th e microstructure of n- and p-\ntype Bi 2Se3 to reduce the thermal conductivity to the < 1 WK-1m-1 range typically observed for \npolycrystalline materials in this family, if it can be performed while maintaining the power \nfactors observed in the current work, promises the possibility of low temperature thermoelectric \n10 \n applications. Finally, the development of more s ophisticated crystal grow th methods for lightly-\ndoped p-type Bi 2-xCaxSe3 crystals will be of interest for th e study of topological surface states. \n \nAcknowledgements \n This work was supported by the AFOSR, grant FAA9550-06-1-0530, and also by the \nNSF MRSEC program, grant DMR-0819860. \n11 \n References \n[1] H. Scherrer and S. Scherrer, in Handbook of Thermoelectrics , D.M. Rowe, ed. CRC press, \nNY (1994) pp 211-237. [2] M. Stordeur, Ibid, pp. 239-255. \n[3] J. Navrátil, J. Horák, T. Plechacek, S. Kamba, P. Losta´ k, J. S. Dyck, W. Chen, and C. Uher, \nJ. Sol. Stat. Chem. 177 1704 (2004). \n[4] A. Von Middendorf, H.F. Ko ehler, and G. Landwehr, Phys. Status Solidi B 57 203 (1973). \n[5] L.P. Caywood, Jr. and G.R. Miller, Phys. Rev. B 2 3209 (1970). \n[6] V.F. Boechko and V. I. Isarev, Inorg. Mater. 11 1288 (1975). \n[7] M. Stordeur, K.K. Ketavong, A. Prei muth, H. Sobotta, and V. Reide, Phys. Status Solidi B \n169 505 (1992). \n[8] A. Al Bayaz, A. Giani, A. Fou caran, F. Pascal-Delannoy, A. Boyer, Thin Solid Films 441 1 \n(2003). \n[9] H. Köhler , and A. Fabricius, Phys. Status Solidi B 71 487 (1975). \n[10] J. Kašparová, È. Drašar, A. Krejèová, L. Beneš P. Lošt’ák, Wei Chen, Zhenhua Zhou, and \nC. Uher, J. Appl. Phys. 97 103720 (2005). \n[11] A. Vaško , L. Tichy, J. Horák and J. Weissenstein, Appl. Phys. 5 217 (1974). \n[12] L. Fu and C.L. Kane, Phys. Rev. B 76 045302 (2007). \n[13] S. Murakami, New J. Phys. 9 356 (2007). \n[14] D. Hsieh, D. Qian, L. Wray, Y. Xia, Y. S. Hor, R. J. Cava and M. Z. Hasan, Nature 452 970 \n(2008). [15] Y. Xia, L. Wray, D. Qian, D. Hsieh, A. Pal, H. Lin, A. Bansil, D. Grauer, Y. S. Hor, R. J. \nCava, and M.Z. Hasan, arXiv:0812.2078 (2008). \n12 \n [16] Haijun Zhang, Chao-Xing Liu, Xiao-Lia ng Qi, Xi Dai, Zhong Fang, Shou-Cheng Zhang, \narXiv: 0812.1622 (2008). [17] J. Zaanen, Science 323 888 (2009). \n[18] L. Fu and C. L. Kane, Phys. Rev. Lett. 100 096407 (2008). \n[19] S. K. Mishray, S. Satpathy, and O. Jepsen, J. Phys.: Condens. Mattter 9 461 (1997). \n[20] S. Urazhdin, D. Bilc, S. H. Tessmer, S. D. Mahanti, T. Kyratsi, and M. G. Kanatzidis, Phys. \nRev. B 66 161306(R) (2002). \n[21] B. Lenoir, A. Dauscher, X. Devaux, R. Ma rtin-Lopez, Yu.I. Ravich, H. Scherrer and S. \nScherrer, Proceedings of the 15th International Conference on Thermoelectrics, IEEE (1996) pp. \n1-13. [22] G.J. Snyder and E.S. Toberer, Nature Materials 7 105 (2008). \n \n13 \n Figures \nFig. 1. (Color on line) (a) Spatially averaged density of st ates (dI/dV) measurements showing \nthe shift of E F between n-type Bi 2Se3 and p-type Bi 2Ca0.02Se3. (b) STM topography of the empty \nstates of Bi 2Se3 (V B = +1 V, and I = 10 pA) showing tria ngular shaped defects observed at \nvarious intensities, indicating they are located in different layers beneath the surface. (c) \nTopography of the empty states of Bi 1.98Ca0.02Se3 (VB = +2.0 V, and I = 10 pA) showing clover \nleaf looking defects and a substantial reduction in the density of the triangular defects, which \ndominated the undoped samples. (d) Topography of the filled states over the same area as in (c) \n(VB = -1.0 V, and I = 10 pA). The area around th e triangular defect near center shows a \ndepression around it in the filled states, demonstrat ing it is positively charged. In contrast, the \nCa-related defects exhibit an area of enhancem ent around them, indicati ng they are negatively \ncharged. All STM topographies are 500 Å by 500 Å. Fig. 2. (color on line) (a) The valence band structure of n-type Bi\n2Se3 measured by ARPES near \nthe Fermi energy and the Γ point of the Brillouin Zone. The ch emical potential (the Fermi level) \nis about 0.4 eV above the top of the valence band, about 0.3 eV above the surface Dirac point \n(15). (b) The band stru cture of a Ca-doped Bi 2Se3 crystal (Bi 1.98Ca0.02Se3) measured under \nsimilar conditions. The chemical potential is with in 30 meV of the valence band. For (a) and (b) \nthe Γ point is 0.0 on the horizontal axes and the band dispersions are shown in the ±M directions. \n(c) The momentum-selective densit y of states (DOS) near the Γ point, compared for the n- and p-\ntype crystals, covering a large bi nding energy range. A large shift of the chemical potential with \nCa doping can be traced by consid ering the shift of DOS peaks in the valence band. The drop of \nthe chemical potential by about 380 meV conf irms the hole doped nature of the Ca doped \nsamples. \n14 \n Fig. 3. (color on line) Temperatur e dependent resistivity be tween 2 and 300 K in the ab plane of \nsingle crystals of Bi 2Se3 and Ca-doped variants. The stoichiometric material is n-type and the \nCa-doped materials are all p-type. The upper inset shows the resistivities in the low temperature \nregime. The lower inset shows the Hall effect data employed to de termine the carrier \nconcentrations for undoped Bi 2Se3 and one Ca-doped crystal. \nFig. 4. (color on line) Temperature dependent Seeb eck coefficients between 2 and 300 K in the \nab plane of single crystals of Bi 2Se3 and Ca-doped variants. The inset shows the low temperature \nregion for both pure and lightly Ca-doped Bi 2Se3. The undoped Bi 2Se3 is n-type, with a carrier \nconcentration of 8 × 1017 cm-3, and the Ca-doped materi al with x = 0.005 is p-type with a carrier \nconcentration of 1 × 1019 cm-3. \nFig. 5. (color on line) Temperature dependent to tal thermal conductivities between 2 and 300 K \nin the ab plane of single crystals of Bi 2Se3 and Ca-doped variants. The inset shows the low \ntemperature region for both pure and lightly Ca-doped Bi 2Se3. \nFig. 6. (color on line) Temperature dependent ther moelectric power factor between 2 and 300 K \nin the ab plane of single crystals of Bi 2Se3 and Ca-doped variants. All materials show a peak at \nlow temperatures but the peak in the p-type variant with 1 % Ca doping (Bi 2-xCaxSe3 with x = \n0.02) is particularly dramatic. The inset shows the thermoelectric figure of merit in the same \ntemperature range. \n15 \n \n \n \n \nFigure 1 \n \n \n16 \n \n \n \n \nFigure 2 \n17 \n \n \nFigure 3 \n \n18 \n \n \nFigure 4 \n \n19 \n \n \nFigure 5 \n \n20 \n \n \nFigure 6 " }, { "title": "2404.06532v1.Solvable_models_of_two_level_systems_coupled_to_itinerant_electrons__Robust_non_Fermi_liquid_and_quantum_critical_pairing.pdf", "content": "Solvable models of two-level systems coupled to itinerant electrons:\nRobust non-Fermi liquid and quantum critical pairing\nEvyatar Tulipman,1,∗Noga Bashan,1,∗J¨ org Schmalian,2, 3and Erez Berg1\n1Department of Condensed Matter Physics, Weizmann Institute of Science, Rehovot 76100, Israel\n2Karlsruher Institut f¨ ur Technologie, Institut f¨ ur Theorie der Kondensierten Materie, 76049, Karlsruhe, Germany\n3Karlsruher Institut f¨ ur Technologie, Institut f¨ ur Quantenmaterialien und Technologien, 76021, Karlsruhe, Germany\n(Dated: April 11, 2024)\nStrange metal behavior is traditionally associated with an underlying putative quantum critical\npoint at zero temperature. However, in many correlated metals, e.g., high- Tccuprate supercon-\nductors, strange metallicity persists at low temperatures over an extended range of microscopic\nparameters, suggesting the existence of an underlying quantum critical phase , whose possible phys-\nical origins remain poorly understood. Systematic investigations of physical scenarios giving rise to\nsuch a critical, non-Fermi liquid (NFL) phase are therefore crucial to better understand this puz-\nzling behavior. In a previous work [1], we considered a solvable large- Nmodel consisting of itinerant\nelectrons coupled to local two-level systems (TLSs) via spatially random interactions, inspired by\nthe possibility of emergent metallic glassiness due to frustrated competing orders, and found that\nthe system hosts an NFL phase with tunable exponents at intermediate couplings. In this work,\nwe expand our investigation to the following: (i) We study the extent to which this NFL phase is\ngeneric by considering various deformations of our theory, including coupling of electrons to multiple\noperators of the TLSs and arbitrarily directed TLS-fields. We find that the physical picture obtained\nin [1] qualitatively persist in a wide region of parameter space, showcasing the robustness of the\nNFL phase; (ii) We analyze the superconducting instability due to coupling of TLSs to electrons,\nand find a rich structure, including quantum critical pairing associated with the NFL phase and\nconventional BCS pairing in the weak and strong coupling limits; (iii) We elaborate on the analysis\nof Ref. [1], including single-particle, transport and thermodynamic properties.\nI. INTRODUCTION\nOne of the central problems in condensed matter\nphysics concerns the low-temperature anomalous normal-\nstate transport properties of correlated metals such as\nhigh- Tccuprate superconductors and others [2–6]. A\nhallmark of the anomalous behavior is the linear-in-\ntemperature scaling of the dc resistivity, known as\n‘strange metal’ behavior. Such behavior stands at odds\nwith conventional Fermi-liquid theory, where a T2-scaling\nis predicted, and is believed to indicate that quantum\nfluctuations are so pronounced as to completely invali-\ndate the Landau Fermi-liquid quasiparticle picture [7–\n10].\nThe traditional theoretical approach to describe\nstrange metals and other non-Fermi liquids (NFLs) in-\nvolves coupling a Fermi surface to bosonic collective fluc-\ntuations of an order parameter, sometimes leading to\nNFL behavior when tuning the system to a quantum crit-\nical point (QCP) [11–13]. In this case, the NFL behavior\nmanifests in a critical fan, emanating from a single (crit-\nical) point at T→0. Interestingly, while consistent with\nsome materials, e.g. heavy fermion systems [14], there\nare numerous examples, e.g. cuprates [15–18] as well as\ntwisted bilayer graphene, organic superconductors, and\nother systems [6, 19–26], where NFL behavior persists\nover an extended region of non-thermal parameters at\n∗These authors contributed equally to this work.T→0 and thus cannot be ascribed to a single QCP.\nRather, the extended NFL behavior raises the possibility\nof a quantum critical phase . Since no general prescription\nexists for this scenario, capturing such behavior within a\ncontrolled, physically motivated theory is considerably\nchallenging.\nOne potential route to realize an extended critical NFL\nphase requires an efficient source of scattering for itin-\nerant electrons over an extended region of non-thermal\nparameters, e.g., by identifying a physical setting where\n“critical” low-energy excitations (i.e., with support at\nthe lowest energies) exist. To this end, in a recent work\n[1], we demonstrated that an NFL phase can arise when\nitinerant electrons are interacting with fluctuations of\na metallic glass (e.g., charge or stripe glass), described\nas a collection of two-level systems (TLSs) that corre-\nspond to quasi-local collective excitations, analogously\nto the excitations in structural glasses1. Physically, our\ntheory is motivated by the complex phase diagrams of\nsuch strongly correlated materials that often host multi-\nple frustrated, competing orders which can give rise to\nglassiness, even in the absence of impurity disorder [28].\nThe presence of disorder could further stabilize such ex-\ntended NFL behavior, as was observed in Ref. [29].\nIn fact, it has been long recognized that inelastic scat-\ntering of electrons off of local TLSs can result in a T-\nlinear resistivity in the weak coupling limit [30]. Nonethe-\n1For a recent detailed analysis of the density of tunneling defects\nin metallic glasses, see Ref. [27].arXiv:2404.06532v1 [cond-mat.str-el] 9 Apr 20242\nless, going beyond weak coupling, the interaction with\nelectrons may dramatically alter the properties of the\nTLSs, e.g., by renormalization of the bare TLS param-\neters or inducing inter-TLS correlations analogously to\nthe Ruderman–Kittel–Kasuya–Yosida (RKKY) mecha-\nnism [31–33]. Clearly these effects can further affect the\nelectrons themselves.\nTo study the evolution of the physical picture beyond\nweak coupling, in Ref. [1], we considered a large- Nthe-\nory consisting of itinerant electrons interacting with local\nTLSs via spatially random couplings. We found that the\nnon-Gaussian saddle point of the theory hosts a robust\nNFL phase with tunable exponents, which is not asso-\nciated with quantum criticality; see Sec. V for a sum-\nmary of our findings. Roughly speaking, the electrons\nconstitute an ohmic bath for the TLSs which results in\na renormalization of the TLS energy splitting towards\nlower energies. At sufficiently strong couplings, a signifi-\ncant portion of the TLSs are renormalized to low energies,\nwhich in turn provides an efficient source of scattering for\nthe electrons, resulting in an NFL behavior over a finite\nrange of coupling strengths.\nIn Ref. [1], we have considered the special case where\nelectrons are coupled to a single operator of TLSs (i.e.,\nσxof spin-1 /2 Pauli operators), which allowed us to fo-\ncus on the effects of inelastic electron-TLS scattering in a\nsimple setting. Specifically, we considered the case where\nelectrons are coupled to σxof the TLSs, which we shall\nrefer to as the ‘ x-model’ henceforth. Aiming for a broader\nunderstanding, it is natural to ask whether the behavior\nof the x-model persist to the generic case, where elec-\ntrons may couple to all three operators of the TLSs. An-\nother important aspect of the physical picture concerns\nTLS-induced pairing, which is expected to take over at\nsufficiently low temperatures.\nIn this work, we investigate a class of large- Nmod-\nels, generalizing our study of the x-model of Ref. [1] to\nmore generic settings. We begin by considering the ef-\nfect of coupling of spinless electrons to all operators of\nthe TLSs. Importantly, the low-energy behavior is qual-\nitatively identical to that of the x-model within a wide\nrange of parameters, showcasing the robustness of the\nNFL phase found in Ref. [1]. Towards a more realistic\nscenario, we generalize our analysis to spinful electrons\nand find a transition to a superconducting ground state\ndue to TLS-induced pairing. Remarkably, the rich phe-\nnomenology of the normal state in our model is also man-\nifested in the form of the critical temperature Tc, exhibit-\ning various crossovers, where in particular the transition\nfrom the NFL phase assumes a quantum critical form.\nIn addition to these key finding, we further study various\naspects of the model, including transport and thermody-\nnamic properties and 1 /Ncorrections.\nThe structure of the paper is as follows. In Sec. II\nwe present the model and discuss physically motivated\nchoices of parameters. In Sec. III we provide a brief sum-\nmary of our main results. In Sec. IV we provide a map-\nping between our model to a set of decoupled spin-boson(SB) models. In Sec. V we provide an extensive review\nof the x-model, previously studied in Ref. [1], which cor-\nresponds to the special case where the electrons interact\nonly with the ˆ xcomponent of the TLSs. In Sec. VI we\nanalyze the general case where the electrons interact with\nall operators of the TLSs (dubbed the xyz−model), and\nin Sec.VII, to the most general case where the field acting\non the TLS is allowed to point in arbitrary directions. In\nSec. VIII, Sec. IX and Sec. X we use our results to study\ntransport, thermodynamic and superconducting proper-\nties of the model, respectively. In Sec. XI we consider\nseveral finite- Ncorrections. Sec. XII contains a sum-\nmary of our work and discuss possible implications and\nfurther directions.\nII. MODEL\nWe consider the following Hamiltonian, defined on a\nd-dimensional hypercubic lattice:\nH=NX\nk;i=1(εk−µ)c†\nikcik−MX\nr;l=1hl,r·σl,r+Hint,(1)\nwhere\nHint=NX\nr;i,j=1\"\n1\nN1/2Vij,r+1\nNMX\nl=1gijl,r·σl,r#\nc†\nircjr.\n(2)\nEach site contains Nelectronic “orbitals” i= 1, ..., N ,\nandMspecies of TLSs l= 1, ..., M where σl,ris a vec-\ntor of spin-1 /2 Pauli operators at position r.εkand\nµdenote the electronic dispersion and chemical poten-\ntial, respectively, and are assumed to be diagonal and\ni−independent in orbital space. For simplicity, we con-\nsider spinless fermions. We will reintroduce the spin in-\ndex later when we discuss superconductivity. Thus, elec-\ntrons create a dynamic local field that acts on the TLSs\nwhile, at the same time, they scatter off those local de-\ngrees of freedom.\nEach TLS is subjected to a field hl,r= (hx\nl,0, hz\nl),\nwith hzbeing the asymmetry and hxthe tunneling rate\nbetween the two states associated with each TLS, both\ntaken to be independent random variables drawn from a\nprobability distribution Pβa(ha) with a=x, z. Below, we\nrefer to this as the physical basis for h. Note that hy≡0\nin order to respect time-reversal symmetry. We focus on\npower-law distributions, Pβa(ha)∝(ha)βa, supported on\nthe interval 0 < ha< ha\nc. It suffices to consider positive\nfields as the sign of hacan be absorbed into the definition\nofσz.ha\ncdenote the TLS bare bandwidth and βa>\n−1 is a tunable parameter. The generalization to other\ndistributions is straightforward. In some cases, it will\nbe convenient to rotate to the eigenbasis where h=hˆz,\nwith h=p\n(hx)2+ (hz)2being the energy splitting of\nthe TLS. We call this the diagonal basis of h.\nWhile we present throughout this work general results\nfor arbitrary βxandβz, there are two physically moti-3\nvated choices for the splitting distributions. Working in\nthe “physical” basis (i.e. the eigenbasis of σz), we as-\nsume that the distribution of the level asymmetry, hz,\nhas finite weight around hz= 0 [34], which corresponds\nto a uniform distribution with βz= 0. It is then natural\nto consider the cases where the width of the distribu-\ntion of the tunneling rate, hx, is either comparable or\nnegligible compared to the asymmetry, corresponding to\nβx= 0, hx\nc=hz\nc≡hc, or setting hx\nc≡0, respectively. In\nthe latter the diagonal basis coincides with the physical\nbasis, while in the former, changing to the diagonal basis\nresults in a linear distribution of the eigenvalues h, i.e.\nβz= 1.\nThe couplings gijl,r=\u0010\ngx\nijl,r, gy\nijl,r, gz\nijl,r\u0011\nare taken\nto be uncorrelated Gaussian random variables with zero\nmean and variance g2\na. For a=x, z, we consider real-\nvalued couplings with\nga\nijl,rga\ni′j′l′,r′=g2\naδr,r′δll′(δii′δjj′+δij′δji′), (3)\nand to ensure that His time reversal symmetric, gymust\nbe purely imaginary with\ngy\nijl,rgy\ni′j′l′,r′=−g2\nyδr,r′δll′(δii′δjj′−δij′δji′).(4)\nThe gaare all real valued. The different components\nare uncorrelated, i.e. gaga′= 0 if a̸=a′.(·) denotes\naveraging over realizations of the coupling constants.\nSimilarly, the onsite potential disorder, Vij,r, is nor-\nmally distributed with zero mean and variance V2. Note\nthat setting the couplings gxandgzto be uncorrelated in\nthe physical basis or in the eigenbasis is not equivalent in\nthe cases where both hx, hz>0. Here we first consider\nsimple variants of the model where gxandgzare uncor-\nrelated in the diagonal basis (i.e. setting hx≡0) and\nlater show that the qualitative physical picture persists\nwhen they are uncorrelated in the physical basis (with\nhx̸= 0).\nThe Fermi energy, EF, sets the largest energy scale in\ntheory, and also corresponds to the cutoff energy of the\nelectronic bath, traditionally denoted by ωc(=EF) in the\nspin-boson literature [35–37].\nThe TLS bandwidth satisfies hc≪EF, and we re-\nstrict the on-site disorder strength, V2, such that Γ =\n2πρFV2≪EF(Γ being the elastic scattering rate),\ntherefore considering ‘good metals.’ We do not restrict\nthe interaction strengths ga, namely, our study covers\nthe range from weak to strong coupling. We focus on the\nlow-energy limit of the model, defined by ω, T≪hc,R,\nwhere hc,Ris the renormalized cutoff of the TLSs, to be\ndefined below.\nThe dimensionless coupling parameters that will be\nused in the following are related to the interaction\nstrengths by ( a=x, y, z ):\nαa=ρ2\nFg2\na\nπ2, (5)\nλa=M\nNρFg2\na\nhc,R. (6)The parameters αa(defined in accordance to the spin-\nboson literature conventions) represent the strength of\nthe dissipation acting on the TLSs, while λaquantify\nthe strength of the scattering of electrons by TLSs at\nlow energies.\nThroughout this work we consider the limit N, M → ∞\nwith a fixed ratio M/N . We will see below that the\nlimit M→ ∞ enables us to (i) reduce the electron’s self-\nenergy to a summation over rainbow diagrams containing\nonly two-point correlation functions of the TLSs, which\nis not clear a priori as Wick’s theorem does not hold\nfor the TLSs; and (ii) invoke self-averaging of the TLSs,\nsuch that sums over the TLS flavors can be replaced by\naverages over the splitting distribution:PM\nl=1f(hl)→\nM´\nf(h)Pβ(h)dh. Importantly, since the splitting dis-\ntribution is independent of position, the self-averaging\nassumption translates to statistical translation invariance\nof the model. Note further that the N→ ∞ limit is es-\nsential for the mapping of our model to the SB model,\nwhere the bosonic bath coupled to the TLSs is composed\nof particle-hole pairs, see Sec. IV.\nIII. BRIEF SUMMARY OF RESULTS\nIn the following sections we expound on the properties\nof different variants of the model. However, for the ben-\nefit of the reader, we first briefly outline the key conclu-\nsions of our work. We first describe the physical picture\nof the x-model and then show that this picture qualita-\ntively persists to generic variants of the model.\nNormal state. Consider the normal state properties\nat low T, corresponding to regions (I) and (II) in Fig. 1.\nIn region (I), as the dimensionless coupling α=αxis\nincreased, the system crosses over from a FL, MFL, and\nNFL, up to a critical value α≈1 where the TLS freeze\natT= 0. These regimes are defined by the exponent of\nthe single-particle scattering rate: Σ′′(ω)−Σ′′(0)∝ |ω|γ,\nγ(α, β) = (1 + β)(1−α) (shown for β= 1 in Fig. 1).\nThis is also manifested in the dc resistivity: ρ−ρ0∝Tγ.\nIn region (II), the TLSs are frozen at T= 0, such that\nscattering off of TLS is mainly elastic. At finite Tfor\nα≳1, however, residual quantum fluctuations of the\nTLSs provides a source for inelastic scattering, leading to\nan additional sequence of NFL-MFL-FL crossovers with\nan inelastic scattering exponent 2( α−1). This residual\ncontribution corresponds to a weak T-variation of the dc\nresistivity as shown in Fig. 1. The behavior in the critical\nregion α≈1 and T→0, separating regions (I) and (II),\nis more involved and show logarithmic T-dependence of\nthe single-particle scattering rate (and the dc resistivity)\nthat smoothly interpolates between the two regions.\nSuperconductivity. Considering the model with\nspinful electrons, a superconducting transition occurs be-\nlow a critical temperature Tcdue to TLS-induced pairing;\nsee Fig. 1. Interestingly, Tcis a non-monotonic function\nof the coupling αx, with remarkably rich pairing phe-\nnomenology. Specifically, at intermediate couplings (cor-4\nFigure 1. (a,b) Illustration of the lattice model Eqs. (1,2). (a) a unit cell containing a large number of electronic states\nand local two-level systems interacting via random couplings; (b) the electrons hop between unit cells. (c)Phase diagram\nof the x−model in α−Tplane ( α=αxis the dimensionless coupling strength) for a linear splitting distribution ( β= 1).\nRegion (I), defined by T≪T⋆∼hc,R,hc,Rbeing the renormalized cutoff of the TLS-splitting, is characterized by the leading\ninelastic scattering exponent 2(1 −α) as manifested in the dc resistivity, see (d); the system crosses over from a Fermi-liquid\n(FL) for α <1/2, to a marginal Fermi-liquid (MFL) for α= 1/2, and a non Fermi-liquid (NFL) for 1 /2< α < 1. At α≈1,\nthe TLSs undergo a freezing transition at T= 0. In Region (II), for α≳1 at finite T, scattering off of the TLSs is mainly\nelastic with small inelastic corrections that scale ∼T2(α−1), namely, another set of crossovers from NFL (1 < α < 3/2) to MFL\n(α= 3/2) to FL ( α >2); see (e). Region (II) is defined up to T∼T⋆⋆, where T⋆⋆is the scale at which standard FL behavior\nbecomes dominant. The gray line denotes the transition temperature to the superconducting state, Tc, in the spinful version\nof the x-model. (d,e) Resistivity as a function of Tfor regions (I) and (II), respsectively, where ρ∞is the resistivity due to\nsaturated classical TLSs. Note that, in region (II), the α-dependent coefficient ηα∝(hc/EF)2≪1, corresponding to the weak\nT-variation of the dc resistivity.\nresponding, e.g., to the NFL phase of region (I) in Fig. 1)\nTcassumes an algebraic, quantum critical scaling form,\nand, as the coupling is further increased beyond a certain\nthreshold (but still at intermediate values), crosses over\nto an Allen-Dynes-like, strong coupling form [38]. In ad-\ndition, Tcassumes the standard BCS form at weak cou-\npling, but also at very strong coupling (e.g., for α >3/2\nin Fig. 1), which corresponds to pairing due to the resid-\nual quantum fluctuations of the nearly frozen spins.\nRobustness. To test the extent to which the physical\npicture of the x-model is generic, we allow for interactions\nwith other operators of the TLSs. In Fig. 2, we show how\nthe normal state T= 0 phase diagram of the x-model\nchanges upon introducing coupling to the y(top row) and\nz(bottom row) operators of the TLSs for constant and\nlinear TLS-splitting distributions (left and right columns,\nrespectively). We shall refer to these variants as the xy-\nandxz-models.\nIn all cases, the qualitative behavior of the x-model\npersists provided that the largest coupling is orthogonal\nto the direction of the TLS-field (i.e. to hzσz), namely,\nthe characterizing exponent γ(α) varies from 1 + βto\n0 as the dominant coupling is increased, up to a crit-ical value at which the TLSs freeze. This sequence of\ncrossovers corresponds to region (I) in Fig. 1, while the\nresidual crossovers of region (II) are expected to qualita-\ntively change for sufficiently strong perturbations due to\nnon-universal corrections. Note that when the dominant\ncoupling is parallel to the TLS-field (i.e., when αz> αx),\nthe TLSs are essentially static and the system shows\nFermi-liquid behavior with TLS-induced elastic scatter-\ning along with weak FL-like corrections.\nIV. MAPPING TO SPIN-BOSON MODEL\nIn this section, we use an effective action approach\nto map our theory to the SB model. We set V2= 0\nfor simplicity. An alternative diagrammatic derivation of\nthe mapping is given in App. A.\nWe begin by considering the spin coherent-state path\nintegral representation for the TLSs. The partition func-\ntion is given by\nZ[h,g] =ˆ\nD[σ, c,c]e−S, (7)5\nFigure 2. T= 0 normal-state phase diagram of the xy-model\n(top) and xz-model (bottom), for constant ( β= 0) and linear\n(β= 1) TLS splitting ditributions. The color represents the\nexponent in the low- Tbehavior of the resistivity: ρ−ρ0∝Tγ\n(analogous to region (I) in Fig. 1). The dashed line denotes\nthe BKT transition over which the TLSs freeze, and beyond\nwhich there is a non-universal version of the NFL phase (sim-\nilarly to region (II) of Fig. 1 for αx>1).\nwith the action, S=S0+Sint:\nS0=X\nrMX\nl=1SBerry[σl,r]−X\nrMX\nl=1ˆ\nτhl,r·σl,r+NX\ni=1X\nkˆ\nτcik(∂τ+εk−µ)cik, (8)\nSint=1\nNX\nrNX\ni,j=1MX\nl=1ˆ\nτgijl,r·σl,rcircjr. (9)\nHere we kept the same symbols σl,rfor the unit vectors\nthat result from the coherent state representation of the\nPauli operators. SBerry denotes the Berry’s phase of the\nTLSs, see e.g. [39].\nTo proceed, we average over the random couplings us-\ning the replica method and introduce the bilocal fields\nGr,r′(τ, τ′) =1\nNX\nicir(τ)cir′(τ′), (10)\nχa,r(τ, τ′) =1\nMX\nlσa\nl,r(τ)σa\nl,r(τ′). (11)\nThe constraints, (10) and (11), are enforced via conju-\ngated fields, Σ and Π, respectively. Notice that, for now,\nwe are considering spinless fermions. In this case, there is\nno pairing instability to leading order in 1 /N. Later on,\nwe shall consider a model of spinful fermions, where the\nanomalous part of the Green’s function must be consid-\nered, and an instability towards a superconducting state\nwith an intra-flavor, on-site order parameter occurs [40].\nTo proceed, we integrate over the fermions and\nsubstitute a replica-diagonal Ansatz , which allows\nus to express the partition function as Z[h] =´\nD[G,χ,Σ,Π,σ]e−Seff, where the effective action is\ngiven by\nSeff=−NTr ln\u0000\nG−1\n0−Σ\u0001\n−Nˆ\nτ,τ′X\nr,r′X\nσGr,r′(τ, τ′) Σr,r′(τ, τ′) +M\n2ˆ\nτ,τ′X\nr,aχa,r(τ, τ′) Πa\nr(τ′, τ)\n+M\n2ˆ\nτ,τ′X\nrX\nag2\naGr(τ, τ′)Gr(τ′, τ)χa,r(τ, τ′) +X\nrMX\nl=1SBerry [σl,r]−ˆ\nτX\nrMX\nl=1hl,r·σl,r\n−1\n2ˆ\nτ,τ′X\nr,aΠa\nr(τ′, τ)MX\nl=1σa\nl,r(τ)σa\nl,r(τ′). (12)\nIn the limit of large MandN, with fixed ratio M/N ,\nwe can analyze the problem in the saddle point limit.\nPerforming the variation with respect to Gand Σ gives\nΣr,r′(τ) =δr,r′M\nNX\nag2\naGr,r(τ)χa,r(τ) (13)\nas well as\nGr,r′(iω) =\u0000\nG−1\n0(iω)−Σ (iω)\u0001−1\f\f\f\nr,r′. (14)\nHere, we have used thermal equilibrium to write the\nsaddle-point equations with time-translation-invariantcorrelation functions and their Fourier transforms. In\naddition, the stationary point that follows from the vari-\nation with respect to χis\nΠa,r(τ) =−g2\naGr,r(τ)Gr,r(−τ). (15)\nThe Berry phase term SBerry, that reflects the fact that\nno Wick theorem exists for Pauli operators, implies that\nthe TLSs cannot be simply integrated over as a Gaussian\nintegral. However, it allows us to recast the TLS problem\nto that of Mdecoupled TLSs per site r,PM\nr,l=1Sr,l[σr,l],\ncoupled to a bosonic bath of particle-hole excitations.6\nEach TLS is governed by the spatially local effective ac- tion\nSr,l[σ] =SBerry [σ]−ˆ\nτhl,r·σ(τ)−ˆ\nτ,τ′Πr(τ′−τ)σa(τ)σa(τ′). (16)\nThis is indeed the action of the spin-boson model after\nthe bosonic bath degrees of freedom have been integrated\nout [37]. The latter give rise to the non-local in time\ncoupling Πa\nr(τ′, τ) that is, in general, different for each\nsite. Of course, in our problem the origin of the bath\nfunction are not bosons but the conduction electrons. For\nthe solution of this local problem this makes, however, no\ndifference. Sr,lstill depends on the random configuration\nhl,rof the fields.\nFor a given realization of the fields hl,rthe problem\nis not translation invariant and correlation functions likeD\nσa\nl,r(τ)σa\nl,r(0)E\nfluctuate in space. However, to deter-\nmine the self energy in Eq. (13) we only need to know\nthe average χa,r(��) of this correlation function over the\nMflavors. To proceed we assume that the model is self-\naveraging in the M→ ∞ limit, such that sums over\nthe TLS flavors can be replaced with averaging over the\nTLS splitting distribution (PM\nl=1→M´\nP(hr)dhr).\nSince the splitting distribution is independent of position,\nthe self-averaging assumption translates to a statistical\ntranslation invariance of the model, at least for the av-\nerage of interest. Hence, χa,r(τ) =χa(τ) is independent\nonr. The same must then hold for the bath function\nΠa,r(τ) = Π a(τ). From the saddle point equations (15)\nit follows that the local fermionic Green’s function and\nthrough Eq. (14) the self energy are both space indepen-\ndent. Hence we can go to momentum space and find that\nthe theory is goverened by a momentum-independent\nself-energy and the Dyson equation for the electrons read\nΣ (τ) =M\nNX\nag2\naχa(τ)G(τ), (17)\nGk(iω) =1\niω−εk−Σ (iω), (18)\nwhere G(τ) =´\nkGk(τ) is the local Green’s function.\nFor a momentum independent fermionic self energy we\nobtain in the limit of large electron bandwidth\nG(iω) =ˆ\nkGk(iω)≈ −iπρFsgn (ω). (19)\nThe particle-hole correlation function can now be evalu-\nated. We find\nΠa(ω) =ρ2\nFg2\na\n2π|ω|, (20)\nirrespective of the electronic self-energy. We thus con-\nclude that each TLS is coupled to an Ohmic bath ofparticle-hole excitations that is independent of the back\nreaction of the TLS on the electronic degrees of freedom.\nThis is a consequence of the fact that the Π aare inde-\npendent of Σ. Thus, we have shown that the (spatially\nlocal) TLS-correlator\nχa(τ−τ′) =1\nMX\nl⟨σa\nl(τ)σa\nl(τ′)⟩, (21)\nis determined by the behavior of Mdecoupled SB models.\nThe strategy of the solution of our model in the large-\nNlimit is therefore: (i) solve the spin-boson problem\nwith ohmic bath for a given realization of the field h, (ii)\naverage over the TLS distribution function of the fields,\nand (iii) use the resulting propagator χa(ω) of the TLSs\nto determine the fermionic self energy from Eq.(18). The\nnon-linear character of the problem is rooted in the rich\nphysics of the spin-boson problem, along with the averag-\ning over the distribution functions of the fields h. Given\nthe importance of the spin-boson model for our analysis\nwe will give a summary of this model in the next section.\nV. THE x−MODEL\nWe begin with a review of the solution of the model for\nthe simplest case where the electrons interact only with\nσxof the TLSs’ pseudospins, i.e., we are working in the\ndiagonal basis and setting gz, gy= 0, as in Ref. [1]. This\nspecial case allows for a more transparent discussion of\nthe key steps of our analysis. In addition, we will see\nthat the more general problem reduces in many cases to\nthis model in the limit of sufficiently low energies. In\nterms of the mapping provided in Sec. IV, the model is\nmapped into the SB model with one bath. Throughout\nthis section, since αy, αz= 0 we will use the notation\nα=αxfor simplicity.\nA. A simple view of the physical picture for α <1\nConsider first the weak coupling limit, α→0, where\nthe effect of interactions can be studied perturbatively.\nTo leading order in g2the decay rate of an electron with\nenergy ωis proportional to the amount of TLSs at acces-\nsible energies, namely,\nΣ′′(ω)∝ˆω\n0Pβ(h)dh∝ |ω|1+β, (22)7\nwhere Σ′′denotes the imaginary part of the electronic\nretarded self energy ( F′′will denote the imaginary part\nof the retarded function Fthroughout the paper). For\nβ= 0 this weak coupling analysis yields marginal Fermi\nliquid behavior. On the other hand, for any β >0, i.e.\nfor distribution functions that vanish for h→0 one only\nfinds Fermi liquid behavior. Increasing the strength of\nthe interaction modifies the behavior of the TLSs in two\nmain aspects: it renormalizes the energy splitting of each\nTLS, such that hℓ→hR(hℓ); and broadens the TLS spec-\ntral function. The broadening has negligible effect on the\nfrequency dependence of Σ′′(at sufficiently low energies).\nHowever, as we show in detail below, the renormalization\nof the energy splitting leads to a renormalization of the\nbare TLS-splitting distribution, decreasing its exponent\nfrom βtoβ−(1 + β)α. Thus, increasing the interac-\ntion strength transfers the spectral weight of the TLSs\ntowards lower energies (at the limit α→1, the spectral\nweight is pushed to zero energy, signaling a BKT tran-\nsition of the TLSs to a localized phase). Consequently,\nthe naive perturbative argument will hold for the renor-\nmalized splitting distribution, resulting in a tunable ex-\nponent as a function of α.\nB. Summary of the single-bath spin-boson model\nThe single-bath spin-boson model (1bSB), or the\nCaldeira-Leggett model [37, 41], is given by the Hamil-\ntonian (note that the commonly used convention in the\nSB literature swaps σx↔σzrelative to our convention)\nH1bSB =−hσz+gσxϕ+Hϕ. (23)\nThe bosonic field ϕ=P\ni(ai+a†\ni) is to be inter-\npreted in terms of a bath of oscillators whose spectral\nfunction, dictated by the Hamiltonian Hϕ, is assumed to\nbe of power law form below some high energy cutoff ωc:\nΠ(ω≪ωc)∝ |ω|s. The cases where s <1, s= 1, s > 1\nare respectively called the subohmic, ohmic, and super-\nohmic baths. Throughout our work we will exclusively\nbe interested in the ohmic case Π( ω) =π\n2αω, using this\nas the definition of the dimensionless coupling constant\nα. For extensive reviews, see, e.g., Refs. [35, 37].\nHistorically, this model was proposed as a toy model\nfor the study of quantum dissipation and decoherence\n[42]. For an ohmic bath, it was found that the spin grad-\nually loses its coherence as αis increased and becomes\noverdamped (in terms of the one-point function ⟨σx(t)⟩)\nbeyond α= 1/2 [35, 36]. At α= 1, the spin undergoes\na Berezinskii-Kosterlitz-Thouless (BKT) phase transition\nafter which it becomes localized in one of the two ˆ xstates\n[35, 37].\nFor our purposes, the most important corollary is that\nin the delocalized regime ( α <1), the TLS-splitting, h, is\nrenormalized due to the high-frequency modes of the bath\nwhich must adjust to different positions whenever the hσz\nterm attemps to flip the TLS between the two ˆ xstates(similarly to the Frank-Condon effect of electron-phonon\ncoupling). This renormalization process, along with the\nBKT transition, are governed by the beta functions of α\nand the rescaled splitting ˜h≡h/ωc, which, to order ˜h2,\nare given by\ndα\ndℓ=−α˜h2, (24)\nd˜h\ndℓ= (1−α)˜h, (25)\nwhere eℓis the renormalization group rescaling factor.\nThe flow dictated by Eqs. (24) and (25) on the α−˜h\nplane is shown in Fig. 3: For αnear 1 there exists a\nconstant of flow, x≡(1−α)2−˜h2, such that the BKT\nseperatrix corresponds to the rightmost of the two x= 0\nlines, with the localized (strong coupling) phase to the\nright of it. The effective energy scales, i.e. the renormal-\nized splittings, hR, in the different regions of the phase\ndiagram are given by\nhR=cαωc×\n\n\u0010\nh\nωc\u00111\n1−αIh/ωc≪(1−α)\ne−bωc\nh IIh/ωc≈(1−α) or\nh/ωc≫ |1−α|\ne−π√\n|x|III h/ωc→α−1\n0 IV else(26)\nwhere cαis a numerical prefactor which cannot be de-\ntermined solely from the RG equations (but can be ex-\ntracted using exact techniques such as bosonization or\nBethe Ansatz [43–46]), and b=b(α, x) is a slowly vary-\ning function whose value is of order 1 away from the BKT\ntransition (both are given explicitly in App. B). The lo-\ncalization of the TLSs in region IV is due to the fact that\nthe effective tunneling hRbetween the the two ˆ xstates\nflows to zero in that region.\nUsing this information on the effective low-energy the-\nory, we now turn to the study of correlation functions. As\nwe apply the RG process, the operator σxremains sup-\nported at low energies, such that correlation functions of\nσxdepend solely on parameters of the low-energy the-\nory:αandhR, and not explicitly on horωc. That is, at\nzero temperature, the x-susceptibility can be expressed\nin terms of a one-parameter scaling form [47],\nχ′′\nx(ω)≡1\nωfα\u0012ω\nhR\u0013\n, (27)\nwith hRgiven by (26), and fα(x) being an α-dependent\nscaling function. This relation between the frequency and\nthehRdependence of the susceptibility will be crucial for\nobtaining exact results in the rest of this section. Unlike\nσx, the correlation functions of the operators σz, σycon-\ntain substantial spectral weight at high energies (of order\nEF), and hence cannot be reduced to a single-parameter\nscaling form as in Eq. (27), and will depend on ωcexplic-\nitly2. This fact will be important when discussing the\n2The difference between σxandσy, σzis due to the fact that after8\nFigure 3. Schematic flow diagram of the 1bSB model.\nWe present the RG flow of Eqs. (24) and (25). The solid\nred line is the BKT speratrix between the dynamic phase\n(I−III) and the localized phase (IV), and the dashed black\nline represents the set of points which map into the isotropic\nKondo model. For our purposes, we separate the dynamic\nphase into subregions according to the functional dependence\nof the renormalized scale hRon the bare ˜h. In region I, the\nflow is nearly vertical (the beta function of αis small), and the\nrenormalization of hRis power law like. In region II, the flow\nbehaves similarly to that of the isotropic Kondo model, and so\nthe renormalized scale is exponential, with a weakly varying\nprefactor in the exponent. Finally, in region III the flow slows\ndown significantly due to the vicinity to the BKT transition,\nand the renormalized scale is exponential and depends on the\ndistance from the transition, vanishing exactly on the critical\nline.\nsubleading contributions to the electronic self-energy in\nSec. VI.\nAnother approach to the solution of the 1bSB model is\nvia a mapping to the anisotropic Kondo model (AKM) or\nthe resonant level model (RLM) [35, 49–51]. Remarkably,\natα= 1/2, known as the “Toulouse Point,” the model\nmaps to the non-interacting point of the RLM, and is\nthus exactly solvable. At this point the effective energy\nscale is hR=πh2/4ωc, and the temperature dependent\nsusceptibility is given by\nχ′′\nx(ω, T) =4\nπ2hR\nω2+ 4h2\nR\u0014\nImΦ ( ω, T)−2hR\nωReΦ ( ω, T)\u0015\n,\n(28)\nwhere\nΦ (ω, T) =ψ\u00121\n2+hR\n2πT\u0013\n−ψ\u00121\n2+hR\n2πT−iω\n2πT\u0013\n(29)\nintegrating out the fast modes of the bath, the operators σy, σz\nin the emergent low energy theory are highly dressed polaronic\nversions of the bare ones which move the high-energy bath modes\nalong with the TLS. However, since the bath couples to the TLS\nin the x−basis, integrating it out does not dress σx, and thus\nthe “high energy” and “low energy” σxoperators coincide. See\n[48] for more details.(ψ(z) being the Digamma function). At zero temperature\nthe scaling form takes the exact form\nf1/2(x) =4\nπ21\nx2+ 4\u0002\nxtan−1(x) + ln\u0000\n1 +x2\u0001\u0003\n.(30)\nHaving summarized the essential properties of the SB\nmodel, we shall now turn to evaluate the electronic prop-\nerties of the x−model using the fact that each TLS is\nequivalent to a spin with a randomly distributed split-\ntingh. Note also that the scaling behavior at T= 0 can\nbe extended to sufficiently small T(to be defined below),\nas can be verified explicitly in the Toulouse point and at\nweak coupling, which enables us to estimate the finite- T\nproperties of the model in the following.\nC. Averaged TLS-susceptibility and electronic self\nenergy\nOnce we have determined the TLS susceptibility\nχi(ω), we can determine the fermionic self energy from\nEq.(18). We compute the imaginary part of the retarded\nself energy using\nΣ′′(ω) =M\nNX\na=x,y,zg2\naˆ∞\n−∞dν\n2πχ′′\na(ν)G′′\nR(ω+ν)\n×\u0012\ncoth\u0010ν\n2T\u0011\n−tanh\u0012ν+ω\n2T\u0013\u0013\n. (31)\nIf we use Eq. (19) for the fermion propagator we obtain\nthat at T= 0 (here gy=gz= 0 and gx=g)\nΣ′′(ω) =−M\nNρFg2ˆω\n0dν \n1\nMMX\ni=1χ′′\nx,i(ν)!\n.(32)\nSince each of the MTLSs contributing to this sum has a\nrandomly distributed splitting hi, the self energy of each\nelectron will be a randomly distributed variable. How-\never, since M≫1 the central limit theorem guarantees\nthat this random variable will be normally distributed,\nwith width of order M−1/2. We can thus replace this\nrandom variable by its expectation value (we will revisit\nthis assumption, and consider when it breaks down for\nlarge but finite M, at the end of this section):\n1\nMMX\ni=1χ′′\nx,i(ω)→χ′′x(ω) =ˆ\nPβ(h)χ′′\nx(ω, h)dh. (33)\nOur task thus reduces to the calculation of the TLS\nsusceptibility averaged over the distribution of splittings\nPβ(h). We note however that the distribution which is\nof interest to us is not that of the bare splittings, but\nof the renormalized splittings hR. We can instead define\nthe renormalized distribution\nPr(hR) =\u0012dhR\ndh\u0013−1\nPβ(h), (34)9\nwhich is nonzero up to the renormalized bandwidth\nhc,R≡hR(hc). Combined with Eq. (27), we may write\nthe averaged susceptibility as\nχ′′x(ω) = sgn( ω)ˆ∞\n|ω|/hc,RPr\u0012|ω|\nx\u0013fα(x)\nx2dx. (35)\nWe will now evaluate this integral for the different func-\ntional forms of hR(h) corresponding to the regions I-IV\nin Fig. 3 at zero temperature.\n1.α <1−hc/EF\nIn region I of Fig. 3, where the flow of αis weak, hR\nis a power of h. Thus the effect of the renormalization\nwould be to alter the exponent in the distribution:\nPr(hR) =γhγ−1\nR\nhγ\nc,R, γ≡(1 +β)(1−α).(36)\nFor energy scales far below the renormalized cutoff ω≪\nhc,R, the result would not be sensitive to the exact form of\nthe scaling function fα, and we retrieve results similar to\nthose presented in the perturbative argument in Sec. V A,\nalbeit with a modified exponent:\nχ′′x(ω) = sgn( ω)Aαγ|ω|γ−1\nhγ\nc,R, (37)\nwith Aα≡´∞\nω/hc,Rfα(x)\nxγ+1dx. Since the scaling func-\ntion fα(x→0)∝x2(corresponding to the universal\n1/t2decay of the real-time correlation function at late\ntimes [35, 37, 48]), we may continue the lower limit of\nthe integral to 0 (provided γ < 2), such that Aα=´∞\n0fα(x)\nxγ+1dx+O\u0010\u0010\n|ω|\nhc,R\u0011γ\u0011\n3.\nRemarkably, observe that at low energies, the leading\nfrequency dependence of the response of the whole col-\nlection of TLSs (which is the effective degree of freedom\ncoupled to the electrons) is independent of broadening ef-\nfects of the individual TLSs. Rather, it is governed solely\nby the renormalized distribution Pr, while the functional\nform of the susceptibility of each individual TLS fα(x)\nis absorbed into the prefactor Aα4. Hence, for ω≪hc,R,\nwe find that\nΣ′′(ω) =−AαM\nNρFg2\f\f\f\fω\nhc,R\f\f\f\fγ\n(38)\n3This result follows from dimensional analysis: If the upper in-\ntegration limit can be continued to ∞, the only place where an\nenergy scale appears is in the normalization of the distribution,\n1/hγ\nc,R, which sets the frequency dependence since the suscepti-\nbility is dimensionless.\n4Note that the low-frequency dependence of χ′′\nxserves as a “cut-\noff” for χ′′x: The exponent in the averaged susceptibility does not\nexceed that of the individual susceptibilities, namely, χ′′x∝ |ω|\nfor all γ >2. This is equivalent to the cases where the integral\ninAαdiverges at the lower limit, resulting in a modification of\nthe frequency dependence.=−λAαhc,R\f\f\f\fω\nhc,R\f\f\f\fγ\n. (39)\nWe see that the self-energy depends on the parameters\nαandβonly via γ. In particular, for any initial β≥\n0, the self energy realizes a MFL form upon tuning the\ncoupling to α=β\n1+β, and realizes any NFL exponent γ <\n1 by increasing αtowards 1. Note, however, that as we\nincrease the coupling α, the effective TLS bandwidth hc,R\ndecreases such that smaller NFL exponents are restricted\nto narrower low-energy intervals; see Fig. 1.\nThe temperature dependence of Σ′′at low Tand zero\nfrequency follows from similar considerations. The con-\ntribution of an individual TLS to the self energy can be\nwritten as a scaling function:\nΣ1(ω= 0, T, h) =−ρFg2M\nNfΣ(T/h R). (40)\nWe can thus perform the averaging over hRat this stage,\nand analogously to Eq. (35) we find that\nΣ′′(ω= 0, T) =A′\nαM\nNρFg2\u0012T\nhc,R\u0013γ\n, (41)\nwith A′\nα=γ´∞\n0fΣ(x)\nx1+γdx. Since fΣ(x≪1)∝x2and\nfΣ(x→ ∞ )→1, this is well defined for 0 < γ < 2. This\nsuggests an ω/T scaling of the form Σ′′∝max (|ω|, T)γ.\n2.α >1 +hc/EF\nForα > 1 +hc/EFall the TLSs have undergone the\nBKT transition and are in the localized phase, where the\ndominant TLS contribution at T= 0 is an elastic scat-\ntering term. However, residual quantum fluctuations of\nthe TLSs at finite frequencies provide a weak inelastic\nscattering mechanism which becomes the leading contri-\nbution to the temperature dependence of the dc resistiv-\nity. The finite frequency behavior follows from scaling\nconsiderations (App. D 4, [48]) and is given by\nχ′′x(ω) = (1 −ηα)δ(ω) + 2ηα(α−1)E2−2α\nF\n|ω|3−2α(42)\nηα=2α(1 +β)\n(α−1)(3 + β)\u0012hc\nEF\u00132\n≪1 (43)\nwith ηα∝(hc/EF)2≪1. We thus find in this regime\nthat the leading inelastic contribution to the self energy\nis of the form\nΣ′′(ω) =−ρFg2M\nN \n1−ηα+ηα\u0012|ω|\nEF\u00132α−2!\n.(44)\nThe low energy excitations of the system are thus those\nof a NFL for 1 < α < 3/2, a MFL at α= 3/2 and FL for\nα >3/2. However, note that unlike in the regime α <1,\nhere the elastic contribution is much larger than the in-\nelastic. Note that while this behavior persists up to a10\nlarge energy scale (a fraction of EF), it is expected to be\nthe dominant contribution to the self energy only below\nan energy scale of order 1 /α4−2α(hc/EF)α−1\n2−α, defined as\nthe scale at which conventional FL-like corrections to re-\nsistivity become comparable (i.e. assuming ρcontains an\nadditional T2/EFcontribution).\n3.α≈1\nIn the “critical” region |1−α|< h c/EFthe above\ndescriptions are not valid, since the flow of hslows down\nand becomes comparable to the flow of α. This slowdown\nleads to a logarithmic behavior of the self energy, which\ninterpolates between the regimes described by Eq. (39)\nand Eq. (42). As an example of the behavior in region\nII of Fig. 3, we consider specifically the case α= 1. The\nrenormalized scale is given by hR=c1ωcexp\u0000\n−πωc\n2h\u0001\n,\nand as a result the renormalized distribution becomes\nlogarithmic:\nPr(hR) =(1 +β) log1+β\u0010\nωc\nhc,R\u0011\nhRlog2+β\u0010\nωc\nhR\u0011, (45)\nwhere we ignore subleading corrections (in log( ωc/hc,R)),\nrelated to the prefactor c1; see App. D 2. Inserting\nEq. (45) into Eq. (35), we obtain that\nχ′′x(ω) =1\nω(1+β) log1+β\u0012ωc\nhc,R\u0013ˆ∞\n|ω|\nhc,Rf1(x)\nxlog2+β\u0010\nωc\n|ω|x\u0011dx.\n(46)\nSince the function f1(x) must decay faster than 1 /xfor\nx≫1 (due to the sum rule Eq. (C4)), and also since\n|ω|/ωc≪ |ω|/hc,R, we may ignore the xinside the log\nin the denominator and then, as in the previous case,\ncontinue the lower limit of integration to 0. The resulting\nself energy is given by\nΣ′′(ω) =−M\nNρFg2\n2\nlog\u0010\nωc\nhc,R\u0011\nlog\u0010\nωc\n|ω|\u0011\n1+β\n. (47)\nNote that when αis not exactly 1, the only difference\nwould be that the factor of π/2 in the exponent of hR\nwill vary slightly. Repeating the above calculations, this\nwill only alter the value of hc,R, but not the functional\nform of the self energy.\nFor 1 < α < 1 +hc/EF, we must distinguish the TLSs\ninto those which are dynamical ( h/E F> α−1), and those\nwhich are localized/frozen ( h/E F≤α−1). The contri-\nbution of the dynamical ones will be similar to that of the\nα= 1 case (shown explicitly in App. D 3, the exact result\nis somewhat more involved, but maintains the logarith-\nmic form of (47)), while the frozen ones will contribute an\nelastic scattering term to leading order (i.e. a delta func-\ntion peak around ω= 0), plus higher order Fermi-liquidterms which we ignore. Defining mα≡α−1\nhc/ωcas the frac-\ntion of frozen TLSs, the self energy will include both a\nconstant elastic contribution along with the NFL contri-\nbution attained earlier. For simplicity, setting β= 1 we\nfind that the self energy will be:\nΣ′′(ω) =−M\nNρFg2\nmα+ (1−mα)Bα\nlog\u0010\nωc\nhc,R\u0011\nlog\u0010\nωc\n|ω|\u0011\n2\n.\n(48)\nAsαapproaches 1 + hc/ωcfrom below, both the relative\nweight, 1 −mα, of the inelastic contribution, as well as\nthe energy scale, hc,R, vanishes.\nVI. THE xyz−MODEL\nWe consider a generalized variant of the model, where\nwe allow electron-TLS coupling in arbitrary directions,\ni.e.,g2\na>0 for a=x, y, z (keeping the field hparallel to\nthezdirection), which we dub ‘ xyz−model’. Remark-\nably, we will show that throughout much of the parame-\nter space (of gx−gy−gz), the behavior is qualitatively\nsimilar to that of the x−model, namely, increasing the\ncouplings will generically drive the model towards a BKT\ntransition, leading to a tunable exponent in the electronic\nself-energy which depends on the distance from the tran-\nsition.\nTo proceed we recall that the electronic self energy in\nthe multi-channel model is given by\nΣ′′(ω) =−M\nNX\na=x,y,zρFg2\naˆω\n0χ′′a(ν)dν\n2π.(49)\nAs before, the TLSs are decoupled such that the dynam-\nics of each TLS are determined by solving an independent\nSB model coupled to three ohmic baths. The correspond-\ning multibath SB model (mbSB) for a single TLS is\nHmbSB =−hσz+X\na=x,y,zgaσa\nxϕa+X\na=x,y,zHϕa,(50)\nwhere the bosonic field of (23) is generalized to three\nfields, ϕx, ϕy, ϕz, corresponding to three independent\nbaths which couple to the three spin directions. In our\ncase, all three baths are ohmic and have the same cutoff\n(because ωc=EF), such that the interaction strengths\nare measured via the three dimensionless couplings, αa≡\nρ2\nFg2\na/π2,a=x, y, z . Let us point out that the model\n(50) has two high-symmetry points: the U(1) symmetric\npoint, corresponding to αx=αy; and the SU(2) sym-\nmetric point corresponding to h= 0, αx=αy=αz(for\nrelevant works see, e.g., Refs. [52–54]). These points will\nnot be of particular interest to us, as they are both un-\nstable fixed points (see below) and require fine tuning.\nSimilarly to the 1bSB model, the low-energy proper-\nties of the model can be obtained from an analysis of the11\nRG flow. Here, the RG equations can be derived pertur-\nbatively in two out the of three couplings; while one of\nthe couplings, αa, is allowed to be arbitrarily large, the\nRG equations are valid to linear order in αb̸=a[52–54].\nLoosely speaking, the RG analysis is valid near the axes\nin the ( αx, αy, αz)-coordinate system. The beta functions\nfor the mbSB model are given by\nd˜h\ndℓ= (1−αx−αy)˜h, (51)\ndαa\ndℓ=−\n2X\nb̸=aαb+ (1−δaz)˜h2\nαa.(52)\nIn order to simplify the analysis, we first discuss the cases\nwhere only two of the baths are active (by setting αz= 0\norαy= 0), treating one bath as dominant and the other\nas a perturbation. We then generalize the discussion to\nthe case where all three baths are active, where, as we will\nshow, the physical pictures can be essentially reduced to\nthe simplified cases (of the two active baths). Further, as\nthe model does not contain any other stable fixed points,\nwe expect that the RG analysis will qualitatively capture\nthe physics for all couplings.\nIt is useful to think about the RG of the multi-bath\ncases as a two-step process: the first step describes the\n‘fast’ flow of the couplings, where the dominant bath as-\nsumes a weakly renormalized coupling while the other ir-\nrelevant baths flow to weak coupling; in the second step,\nthe baths renormalize the TLS-fields according to the\nrenormalized couplings. We now proceed to analyze the\ndifferent cases, where the usefulness 2-step perspective\nbecomes apparent.\nA. The xy−model\nWe start by setting αz= 0, i.e., the case where there\nare two active baths acting in the direction perpendicular\nto the field. The beta functions are given by\nd˜h\ndℓ= (1−αx−αy)˜h, (53)\ndαx\ndℓ=−2αyαx−˜h2αx, (54)\ndαy\ndℓ=−2αxαy−˜h2αy. (55)\nIt is insightful to define the bath anisotropy parameter,\nθ≡\u0012αx−αy\nαx+αy\u00132\n, (56)\nwhose beta function is\ndθ\ndℓ= (αx+αy)θ(1−θ). (57)\nThe anisotropy is thus relevant whenever the couplings\nare not finely tuned to the U(1) symmetric point θ= 0,\nFigure 4. Initial flow of the couplings in the xy-model.\nθ=\u0010\nαx−αy\nαx+αy\u00112\nis the bath-anisotropy parameter.\nand flows towards the maximally anistropic case θ= 1,\ni.e. where the larger of the two couplings dominates and\nthe other one becomes irrelevant, as depicted in Fig. 4.\nAs we will now see, since the subdominant bath is ir-\nrelevant it can be integrated out easily, leading to a low\nenergy description similar to that of the x−model with\nrenormalized coupling.\nConsider the case αx≫αy. We focus on the regime\n1−αx≫˜h2(i.e., far enough from the BKT transi-\ntion), where simple analytical estimations can be made\nas the effect of ˜hon the flow of the couplings is neg-\nligible. Indeed, the equations can be solved by utiliz-\ning the fact that δα=αx−αyis an approximate con-\nstant along the flow. The resulting low-energy theory\nis described by the renormalized splitting, hR, and cou-\nplings, αx,R=δα≫αy,R= (hR/ωc)δααy. The BKT\ntransition is determined by the renormalized value of\nthe dominant coupling, αx,R, such that the system be-\ncomes localized when αx,R>1 +O(h/ωc), and below\nthis value the effective energy scale assumes the familiar\nform hR=ωc(h/ωc)1/(1−αx,R)(note that the exponent\ndepends on αx,Rand not on the bare value). More details\non the RG flow are shown in App. E.\nTo proceed, we note that the low energy theory we have\narrived at is nearly identical to that in the 1bSB, the one\ndifference being a remaining weak coupling to the ybath\n(αy,R≪1) which we may now treat perturbatively. The\noperator σxis only weakly dressed (it is renormalized\nonly in the short first section of the flow when αyis of\norder 1), and thus we can once again conclude that its\ncorrelation functions will assume a one parameter scaling\nform, as in Eq. (27):\nχ′′\nx(ω) =1\nωfαx,R\u0012ω\nhR\u0013\n+δχ′′\nx(ω), (58)\nwhere fαx,Ris a scaling function and the perturbative\ncorrection due to the coupling to the losing bath is of the\nform δχ′′\nx(ω) =αy,R1\nω˜fαx,R\u0010\nω\nhR\u0011\n, with a different scaling\nfunction ˜fαx,R. When averaging over the second term.,\nthe strong renormalization of the losing bath, αy,R=\n(hR/ωc)δααy, effectively enhances the exponent in the\nrenormalized distribution which results in a subleading\nfrequency dependence of the averaged δχ′′\nx; see App. F.\nIn contrast, the y−susceptibility does not assume a\none parameter scaling form. Rather, it is suppressed by\nadditional factors of hR/ωc, such that in the IR limit all\nspectral weight is shifted to frequencies of order ωc; see\nApp. F.12\nThe electronic self-energy can be evaluated following\nthe analysis of Sec. V C. Remarkably, the leading term in\n(58) assumes the same form as in the x−model:\nΣ′′(ω) =−Aαx,RρFg2\nx\f\f\f\fω\nhc,R\f\f\f\fγ\n, (59)\nwhere γ= (1+ β)(1−αx,R), and Aδαx,R=´∞\n0fαx,R(x)\nx1+γdx.\nThe subleading correction due to δχ′′\nxis of the form\nδΣ′′(ω) = −Bαyγ\nγ+δαρFg2\nx\f\f\fω\nhc,R\f\f\fγ\f\f\fω\nωc\f\f\fδα\nwith B=\n´∞\n0˜fδα(x)\nx1+γ+δαdx. An additional subleading contribution\nto Σ′′is related to the coupling to the y−susceptibility,\nwhich we denote by δΣ′′\ny. While an explicit evaluation of\nδΣ′′\nyis more challenging, the fact that δΣ′′\nyis also sublead-\ning follows from the additional “non-universal” factors of\nhR/ωcwhich it contains, similarly to the case of δχ′′\nx.\nAs mentioned before, this RG-based analysis is pertur-\nbative in the strength of the weaker coupling, αy. How-\never, at strong coupling (when αx≳αy) the value of αx,R\nis no longer equal to δα, and the BKT line αx,R= 1\nchanges accordingly. The problem has been solved nu-\nmerically for varying coupling strengths by [55], who have\nfound that at strong coupling the BKT line approaches\nthe line αx=αyasymptotically. This behavior is de-\npicted schematically in Fig. 2.\nNote that for U(1)-symmetric points, αx=αy=α,\nthe coupling to the two baths is frustrated and the system\nflows to weak coupling [53, 54, 56]. This case will be\ndiscussed in an upcoming work [57].\nB. The xz−model\nConsider now the case where αx, αz>0 and αy= 0,\ndubbed xz−model. The major difference in this case\ncompared to the xy−model is the fact that the z-bath\nis aligned with the “field” hz, making the two baths in-\nequivalent. The flow equations in this case are:\nd˜h\ndℓ= (1−αx)˜h, (60)\ndαx\ndℓ=−2αzαx−˜h2αx, (61)\ndαz\ndℓ=−2αxαz. (62)\nAs before, we neglect the effect of ˜hon the initial flow of\nthe couplings, assuming that it is sufficiently small.\nLet us start with the case αx≫αz. Then, as the\nx-bath dominates, the flow is essentially identical to the\nxy−model with zreplacing y. The main difference is\nthat σz, albeit being strongly dressed in the low energy\ntheory, has a non-zero equilibrium expectation value, i.e.,\n⟨σz⟩ ̸= 0. The z−susceptibility therefore contains a term\nproportional to ⟨σz⟩2δ(ω). Fortunately, ⟨σz⟩may beevaluated using the sum rule Eq. (C5), yielding\n⟨σz⟩∞=\u0012hR\nωc\u0013δα\na+b1−\u0010\nhR\nωc\u00111−2δα\n1−2δα\n,(63)\nwith aandbbeing numerical constants which depend\nonαx, αz. Interestingly, this static piece contributes an\nelastic scattering term to the electronic self energy:\nΣ′′\nel(ω)∝ −ρFg2\nz\n\n\u0010\nhc,R\nωc\u00112δα\nδα < 1/2,\nhc,R\nωclog2\u0010\nωc\nhc,R\u0011\nδα= 1/2,\n\u0010\nhc,R\nωc\u00112−2δα\n1/2< δα < 1,\n(64)\nwhere O(1) numerical coefficients have been suppressed\nfor clarity. The leading “inelastic” part of Σ′′due to\nthex−bath is identical to Eq. (59), with appropriate\nrenormalized value αx,R. It is worth recalling that the\ninteraction-induced elastic term of Eq. (64) adds to the\nelastic scattering term due to onsite potential in the gen-\neral model. We comment on this matter further in the\ndiscussion on the dc resistivity in Sec. VIII.\nWe move on to the second scenario, where αz≫αx. In\nthis case, the z−bath dominates such that the coupling\nflows to a finite value, αz,R=δα=αz−αx. Unlike the\nprevious case, hzis only marginally renormalized, such\nthat the low-energy parameters are given by\nhR=δα\nαzhz, (65)\nαx,R=αx\u0012hR\nωc\u0013δα\n. (66)\nNote that when δα→0,hRassumes the form shown\nin the xy−isotropic case, although the system is not\nU(1)−symmetric due to hz. For αx= 0 the TLS is static,\nand the scattering is solely elastic. The effect of αxis an\naddition of a weak inelastic term to χ′′\nz, as well as a cou-\npling of the electrons to χ′′\nx. This will result in an inelastic\ncontribution in the self energy such that Σ′′= Σ′′\nel+Σ′′\ninel\nwith\nΣ′′\nel≈ −ρFg2\nz\n2π(67)\nΣ′′\ninel(ω)∝ −|ω|γ+2δα\nhγ\nc,Rω2δαc(68)\nwhere γ= (1+ β). While this is the leading inelastic con-\ntribution, it does not give rise to any MFL/NFL behavior\nfor the considered splitting distributions with β≥0. For\nmore details, see App. F.\nC. The xyz−model\nUnderstanding the physical picture in the more general\ncase where the electrons are coupled to all opeartors of13\nthe TLSs (i.e. αa>0 for a=x, y, z andhx≡0), dubbed\nxyz−model, rests upon the fact that the anisotropy re-\nmains relevant such that one bath dominates over the\nothers at low energies. The qualitative behavior is thus\nreduced to one of the two previous cases. In the case\nwhere the coupling to the x−ory−baths is the largest,\nthe behavior is qualitatively similar to the x−model, al-\nbeit with a renormalized coupling, αR(α) (of the domi-\nnating bath), which depends on the initial values of the\ncouplings. Specifically, the leading inelastic contribution\nto the self-energy satisfies\nΣ′′(ω)−Σ′′(0)∼ −|ω|γ, γ = (1 + βz)(1−αR),(69)\nas we have demonstrated above. In addition, above a\ncritical value, αR(α)≥1, the TLSs undergo a BKT tran-\nsition to the localized phase, where most of the scattering\nis elastic. If, on the other hand, the z−bath dominates,\nthe TLSs act essentially as static impurities, with an ad-\nditional weak, FL-like inelastic contribution to the self\nenergy (as in the xz−model with dominant z−bath).\nVII. “BIASED” MODEL\nAll of our above analyses relied on the assumption that\nthe field, h, is parallel to the ˆ zdirection. Note that this\ncannot always be made the case by rotating a generic field\nh= (hx,0, hz) to point in this direction, since this would\ninduce correlations between the the couplings gx, gz, and\nin turn will lead to a mbSB model with correlated baths.\nWe thus keep the couplings uncorrelated and treat the\ncase were hx, hz̸= 0.\nA. Biased x−model\nWe start by considering the x−model with finite par-\nallel fields hx>0. We thus allow hto be randomly\ndistributed in the x−zplane with a joint distribution\nPβx,βz(hx, hz)∝hβxxhβzzforhx, hz< h c. This variant\nmaps into the “biased” 1bSB, where a field parallel to the\nbath coupling, hxσx, is added to the Hamiltonian [35, 37].\nThe main difference in this case is that hxis unaffected\nby the bath (it commutes with the interaction), while hz\nis renormalized as before. In addition, the presence of a\nnon-zero hximplies that ⟨σx(t→ ∞ )⟩ ̸= 0, which leads\nto an “elastic” delta function term in χ′′\nx(ω), namely, the\nx−susceptibility can be written as χ′′\nx≡χ′′\nel+χ′′\ninel, with\nχ′′\nx,el(ω) =\u00122\nπtan−1\u0012hx\nhR\u0013\u00132\nδ(ω), (70)\nχ′′\nx,inel (ω) =1\nωf\u0012ω\nhR,ω\nhx\u0013\n. (71)To leading order in hc,R/hc, we find that the low-energy\nself-energy is given by Σ′′≡Σ′′\nel+ Σ′′\ninel, with\nΣ′′\nel(ω) =−M\nNρFg2, (72)\nΣ′′\ninel(ω)∝ −M\nNρFg2\f\f\f\fω\nhc\f\f\f\f1+βx\f\f\f\fω\nhc,R\f\f\f\f(1+βz)(1−α)\n.(73)\nNote that since hc/hc,R∝(ωc/hc)α/(1−α)≫1, the\nelastic term contributes most of the spectral weight to\nχx, and correspondingly, Σ′′\nel≫Σ′′\ninelat low energies.\nWe provide an explicit calculation using the 2-parameter\nscaling-form along with exact evaluation at the TP in\nApp. D 5. Notice that for the reasonable case of βx≥0\nthis results in a FL for α < 1 and approaches a MFL\nbehaviour near α= 1.\nB. Biased xyz−model\nWe now treat the most general variant of our model,\nwhere the couplings and fields are all allowed to point\nin generic directions. The behavior of the biased\nxyz−model can be similarly understood from the RG\nanalysis, where Eqs. (52) are modified as [58]\nd˜hz\ndℓ= (1−αx−αy)˜hz, (74)\nd˜hx\ndℓ= (1−αz−αy)˜hx, (75)\ndαa\ndℓ=−\nX\nb̸=aαb+ (1−δaz)˜h2\nz+ (1−δax)˜h2\nx\nαa,\n(76)\nwhere ˜hx≡hx/ωc. Importantly, there are no cross terms\nbetween the fields hx, hzsuch that the qualitative be-\nhavior can be understood in terms of the approximately\nindependent flow of the individual fields. Furthermore,\nsince a bias along ˆ yis forbidden in order to respect time-\nreversal symmetry, the physical picture in the biased\nxyz−model is determined by whether or not the y−bath\ndominates. In the case where the y−bath dominates,\nboth the ‘field’, hz, and the ‘bias’, hx, will be renormal-\nized according to Eq. (26) with α→αy,R. Consequently,\nthe leading inelastic contribution to the self-energy will\ntake the form\nΣ′′(ω)−Σ′′(0)∼ −|ω|γ, (77)\nγ= (2 + βx+βz)(1−αy,R), (78)\nalong with elastic scattering terms as in Eq. (64). In\ncontrast, if the x−orz−baths dominate, the behavior\nwill be analogous the biased x−model with the appro-\npriate renormalized couplings of the dominant bath; see\nEq. (73).14\nVIII. TRANSPORT\nLet us begin by considering the electronic contribution\nto the electrical conductivity, and later incorporate the\neffect of the TLSs on the optical conductivity. Using the\nKubo formula, the real part of the conductivity (associ-\nated with the electrons) is given by\nσel(Ω) =ImΠR\nJx(Ω)\nΩ, (79)\nwhere ΠR\nJxis the retarded current correlator (along the\nxdirection). The current operator is given by J=P\na´\nkvkc†\nakcakandvk=∇kεk. The evaluation of ΠR\nJxis greatly simplified since all vertex corrections vanish due\nto the spatial randomness of the couplings to the local\nTLSs (similarly to Refs. [59, 60]). The electronic optical\nconductivity is thus given by\nσel(Ω) =1\nΩˆ\nωˆ\nkv2\nkAk(ω)Ak(Ω + ω) [f(ω)−f(Ω + ω)].\n(80)\nHereAk(ω)≡ −1\nπImGR\nk(ω) is the electronic spectral\nfunction and f(ω) denotes the Fermi distribution func-\ntion.\nIn the dc limit, the conductivity is given by\nσel(Ω→0) =v2\nFρF\n16Tˆdω\n2π1\n|Σ′′(ω)|sech2\u0010ω\n2T\u0011\n.(81)\nHence, the T-scaling of the dc resistivity follows the\nsingle-particle lifetime. It is instructive to first consider\nρ(T) in the x−model. For α < 1, the low- Tbehavior,\nT≪hc,R, is of the form\nρ(T) =ρ0+ATγ, (82)\nwhere here we have restored the on-site disorder by set-\nting V2>0, corresponding to the residual resistiv-\nity term, ρ0, and γ= (1 + β)(1−α). Similarly, for\nα= 1, we have that ρ(T)−ρ0∝1/|log (T)|1+β; and\nforα > 1 +hc/ωc, the TLSs are frozen at T= 0, and\nthus contribute to the residual resistivity, with FL-like\nfinite- Tcorrections ( ρ(T)−ρ0∝T2). In the intermedi-\nate regime, 1 < α < 1+hc/ωc, the resistivity interpolates\nsmoothly between these two behaviors. In the more gen-\neralxyz−model, the resistivity follows the behavior of\nthex−model whenever the transverse couplings αxor\nαydominate (as discussed extensively in the Sec. VI).\nWhereas, if the parallel coupling αzdominates, the scat-\ntering is mainly elastic with weak FL-like temperature\nscaling. Similarly, the biased xyz−model follows analo-\ngous behavior to that of the x−model provided that the\ny−bath dominates, and to the biased x−model if the x−\norz−bath dominates.\nWe proceed to consider the optical conductivity. In ad-\ndition to the contribution due to the itinerant electrons,\nwe also assume that each TLS carries a randomly dis-\ntributed dipole moment (recall the TLS are phenomeno-\nlogically related to charged collective degrees of freedom)which depends on the state of the TLS:\nHdipole =X\nr,lEr·\u0000\ndz\nr,lσz\nr,l+dx\nr,lσx\nr,l\u0001\n. (83)\nHereEris the local electric field and dx,z\nr,ldenote un-\ncorrelated Gaussian random dipole moments of the TLS\nflavors, with variances d2\nx,z. In total, the (longitudinal)\noptical conductivity takes the two-component form\nσ(Ω) = σel(Ω) + σTLS(Ω). (84)\nThe electronic contribution is standard and follows\nstraightforwardly from the form of the self energy. In\nparticular, at low energies, where −Σ′′(ω) =Γ\n2+c|ω|γ\n(i.e. we restore the elastic scattering term that does not\naffect any of the previous results), if the scattering is\nmainly inelastic (Γ ≪c|Ω|γ):\nσel(Ω)∼\n\n1\nΩγ γ <1,\n1\nΩ log2(1/Ω)γ= 1,\n1\nΩ2−γ γ >1.(85)\nwhile if the scattering is mainly elastic (Γ ≫c|Ω|γ):\nσel(Ω)∼1\nΓ−2γ+1c\n(γ+ 1)Γ2|Ω|γ. (86)\nAt higher energies, Ω ≳hc,R, the TLS contribution to Σ′′\nsaturates to a constant such that σel(Ω≳hc,R)∼1/Ω.\nThe TLS contribution is given by\nσTLS(Ω) = Ω\u0000\nd2\nxχ′′x(Ω) + d2\nzχ′′z(Ω)\u0001\n. (87)\nInterestingly, the TLS contribution follows the frequency\ndependence of the inelastic part of Σ′′(provided that the\ny−bath is not dominant). In particular, if the dipole mo-\nments are not negligibly small, σTLSmight constitute the\nleading frequency dependence, leading to a positive slope\nand non-monotonic behavior of the optical conductivity.\nDefining the energy scale Ω ∗=√ρFvF/da, we find that\nif the scattering is dominantly elastic and Γ ≫Ω∗then\nthere will be an increasing optical conductivity around\nzero frequency. If inelastic scattering dominates, Γ ≪Ω∗,\nthe optical conductivity will always be decreasing around\nzero frequency, but will begin increasing for frequencies\nof order Ω mIR∼ZΩ∗if the system is a FL (with Zthe\nquasiparticle weight), or Ω mIR∼(Ω∗/c)1/γif the system\nis a NFL (i.e. if γ≤1), leading to a so-called mid-\nIR peak around energies of order hc,R(assuming that\nΩmIR< hc,R) [1].\nThe assumption that led to Eq. (87) was that there\nare sufficiently many TLSs that carry a dipole moment\nand can therefore be optically excited. At the same\ntime one expects that there are TLSs that locally come\nwith a quadrupole moment. For example, they could lo-\ncally distort a state of four-fold rotation symmetry to a\nlower symmetry. In this case one can excite the TLS\nvia inelastic light scattering and the Raman response15\nfunction[61, 62] will measure directly the TLS suscep-\ntibilities\nRα,β(Ω) =\u0000\nqx\nα,β\u00012χ′′x(Ω) +\u0000\nqx\nα,β\u00012χ′′z(Ω). (88)\nHere qκ\nα,βis the quadrupole moment due to the κ-\ncomponent of the TLS pseuodspin. The individual tensor\nelements can be detected by an appropriate combination\nof the polarization of the incoming and scattered light.\nHence, the presence of TLS can, at least partially account\nfor the broad Raman continuum that has been observed\nin many correlated electron materials [61].\nIt is intriguing to examine the MFL/NFL transport\nproperties of our model through the viewpoint of Planck-\nian dissipation and the putative bound on transport\ntimes [8]. Since there is no unique definition for the\ntransport time, we consider two different approaches.\nFollowing Ref. [8], we can associate the transport time\nto the single-particle lifetime as the two are propor-\ntional in our model. In that case, the inverse transport\ntime is Planckian in the sense that 1 /τtr∼Twith an\nO(1) coefficient for the NFL phase while at the MFL\npoint the coefficient is O(1/ln(1/T)). In particular, our\nmodel trivially satisfies a ‘Planckian bound’ due to the\nKramers-Kroning relations between the real and imagi-\nnary parts of Σ5. Alternatively the inverse transport time\ncan be defined in terms of the energy scale for which\nthe dc and ac conductivities become comparable [59]:\nσ\u0000\nτ−1\ntr(T), T= 0\u0001\n∼σ(Ω = 0 , T). This procedure agrees\nwith the single-particle lifetime result for NFLs while for\nthe MFL case the transport time contains an additional\nlog correction: τ−1\ntr∼T/log2(T).\nLastly, relying on the analysis of the weakly disor-\ndered MFL (or NFL) model in Ref. [63], we note that\nthe Wiedemann-Franz law is obeyed as T→0, regard-\nless of the existence of well-defined Landau quasiparticles\n(in the absence of vertex correction, as we have here, the\nanalysis is essentially identical).\nIX. THERMODYNAMICS\nIn this section, we study thermodynamic properties of\nthe model. We mainly consider the x−model and further\ndiscuss the expected behavior in the xyz−model. It is\nworth noting that a direct evaluation of the free energy\nfrom the saddle point of the large- Neffective action is\nchallenging due to its non-Gaussian nature. Instead, we\nobtain the specific heat from the internal energy, and\n5Note that for MFLs or NFLs, the Kramers-Kroning relations\nrelate the low energy regimes of the real and imaginary parts of\nthe self energy, such that one is completely determined by the\nother (independent of the UV cutoff). This is in contrast to FLs\nwhere, since Σ′′(ω)∼ |ω|γ+O(|ω|γ) (with γ >1), the low and\nhigh energy sectors both contribute to the leading linear in ω\nbehavior of ReΣ ret(ω) (making Kramers Kroning relations not\nparticularly useful if only low-energy information is accessible).corroborate our results with an alternative derivation of\nthe specific heat from the entropy, where in particular\nwe confirm the absence of T= 0 residual entropy in our\nmodel.\nConsider the internal energy density\nU≡1\nNV⟨Hel+HTLS+Hint⟩, (89)\nwhere HelandHTLScorrespond to the first two terms in\n(1), respectively, and Vis the volume of the system. Let\nus henceforth suppress the factor 1 /(NV) and assume r=\nN/M = 1 for simplicity. By employing the equation of\nmotion for the retarded and advanced electronic Green’s\nfunctions (see App. G), we may write\n⟨Hel+Hint⟩=ˆ\nkˆ\nωωnF(ω)Ak(ω)≡Uel,0,(90)\nwhere nF(ω) is the Fermi function. Note that due to\nthe locality of the self energy, Uel,0corresponds to the\ninternal energy of non-interacting electrons. To see this,\nwe use the fact that´\nkA(ω,k) = ρF, hence Uel,0=\nρF´\nωωnF(ω). The specific heat related to Uel,0is given\nby\ncel,0=ρFˆ\nωω∂nF(ω)\n∂T=π2\n3ρFT, (91)\ni.e., the specific heat of non-interacting electrons. Inter-\nestingly, all interaction effects are encoded in the renor-\nmalized TLS part of the internal energy, UTLS=⟨HTLS⟩,\nwhich we will now evaluate. Using the sum rule Eq. (C5),\nwe may express the TLS specific heat as\ncTLS=1\n2ˆ\nωω∂χ′′x(ω, T)\n∂T. (92)\nAshc,Ris the only energy scale, for ω, T≪hc,Rone can\nwrite the TLS susceptibility as a 2-component scaling\nform, i.e. χ′′(ω, T) =1\nωF\u0010\nω\nhc,R,ω\nT\u0011\n. For concreteness,\nwe assume the following scaling form:\nχ′′x(ω, T) =χ′′x(ω, T= 0)× \n|ω|p\nω2+ (aT)2!φ\n(93)\nwith a∼ O(1) some numerical coefficient and scaling\nexponent φ >0,6which both affect the result only by a\nnumerical prefactor and not the Tdependence.\nUsing Eq. (92), we obtain that\ncTLS∼\n\nT\nhc,Rγ >1,\nT\nhc,Rloghc,R\nTγ= 1,\u0010\nT\nhc,R\u0011γ\nγ <1.(94)\n6Exact results for α≪1 and α= 1/2 indicate that φ= 1.16\nIn addition, Eq. (92) can be evaluated numerically at the\nToulouse point where the exact temperature dependence\nofχ′′(ω, T) is known analytically. The results, confirming\ntheT-dependence found in Eq. (94), are shown in Fig. 5\nforhdistributions corresponding at the Toulouse point\nto a FL, MFL and NFL.\nWe corroborate the above discussion with an alterna-\ntive derivation of the specific heat from the entropy. To\ndo so, we consider the addition of a single TLS per site\nto the theory with M= 0. In the language of the SB\nmodel, the excess entropy added to the system, defined\nbyδS≡S(M= 1)−S(M= 0), is known as the ‘impurity\ncontribution’ [36]. Importantly, δSis determined by the\nspectral function of the bath, the renormalized splitting\nand the temperature. Hence, since the TLSs are decou-\npled for any Mprovided that N≫1 (as gijlgijl′= 0\nforl′̸=l), and the particle-hole bath is ohmic, we may\nwrite the entropy of MTLSs by adding the impurity\ncontributions of the individual TLSs.\nConsidering the x−model, the impurity contribution\nof a single TLS for α <1 is given by [36, 48, 64, 65]\nδS(x) =(\nαπ\n3x+O\u0000\nx3\u0001\nx≪1,\nlog 2 x≫1,(95)\nwhere x=T/h R. The entropy of the full system (i.e. in\nthe large- M, N limit) can therefore be written as\nS(T) =S0(T) + ∆ S(T), (96)\nwhere S0=S(M= 0) denotes the contribution of the\nnon-interacting electrons and\n∆S(T) =MVˆhc,R\n0Pr(hR)δS\u0012T\nhR\u0013\ndhR.(97)\nTo evaluate ∆ S, we divide the integral over hRtohR< T\nandhR> T, denoted by S, respectively, and,\nsubstituting (36) in (97), we obtain\nS<(T)≈παγ\n3hγ\nc,RT\nhc,Rhγ−1\nc,R−Tγ−1\nγ−1, (98)\nS>(T)≈πα\n3aγ\u0012T\nhc,R\u0013γ\n, (99)\nwhere aγ≡´1\n0δS\u00001\nx\u0001\nxγ−1dx≈1/γ. Hence, for T≪\nhc,R,\n∆S(T)∼\n\nT\nhc,Rγ >1,\nT\nhc,Rloghc,R\nTγ= 1,\u0010\nT\nhc,R\u0011γ\nγ <1.(100)\nSince S0∼T, the total entropy, S, obeys the same scaling\nas ∆S, which also holds for the specific heat, in agree-\nment with Eq. (94).\nIt is also worth noting that there is no residual ex-\ntensive entropy at T= 0, in contrast to theories of\nFigure 5. Specific heat, as extracted from the internal energy\nat the Toulouse point, for values of γfor FL, MFL and NFL\nbehavior (corresponding to β= 2,1,0). As expected from the\nabove scaling arguments (99), at low temperatures the ratio\nc/Tapproaches a constant for the FL, has logarithmic diver-\ngence for the MFL, and polynomial divergence with exponent\n1−γfor the NFL.\nMFLs constructed from variants of the Sachdev-Ye-\nKitaev model [7].\nPhysically, the T−scaling of the impurity contribution\nin Eq. (96) stems from scattering of the low-energy modes\nof the bath by the TLS. Hence in the cases where αxor\nαydominate, the impurity contribution is expected to\nfollow Eq. (96) since the low energy theory is identical\nto that of the x−model (up to weak perturbations). It\ntherefore follows that S(T) satisfies Eq. (100). Moreover,\nwhen αzdominates, the system realizes a FL (with addi-\ntional static impurities) and weak renormalization of the\nsplittings, such that S(T)∝T.\nX. SUPERCONDUCTIVITY\nIn order to study the superconducting instability, we\nintroduce spinful electrons to the Hamiltonian Eq. (1,2),\nnamely, we let c†\ni,r→c†\ni,s,rwith s={↑,↓}. Note that\nsince the TLSs have no spin structure (assuming that\nthe underlying glass is non-magnetic), the couplings g\ndo not depend on spin index s, and the interaction term\nis diagonal in spin space. The mapping of the spinful\nvariant of the Hamiltonian onto a spin-boson problem is\nsimilar to the spinless case, but one must also consider\nthe anomalous bilocal field, analogously to Eqs. (10,11),\ndefined as\nFr,r′(τ, τ′) =1\nNX\nici,↓,r(τ)ci,↑,r′(τ′), (101)\nand enforced via the anomalous self-energy Φr,r′(τ, τ′)\n[40, 66]; see App. H.17\nWe study the critical temperature Tcas a function of\nthe couplings α. Approaching the SC state from the\nnormal state, where Φ = 0, we obtain Tcas the solu-\ntion for the linearized Eliashberg equation (in imaginary\nfrequency) for the local anomalous self-energy Φ( iω):\nΦ(iω) =TX\nω′DΦ(iω−iω′)\n|ω′+iΣ(iω′)|Φ(iω′), (102)\nwhere the bosonic propagator is\nDΦ(iω) =X\na=x,y,zλahc,Rχa(iω)×(−1)δa,y. (103)\nNotice that while the interactions via gxandgzmedi-\nate pairing, gyis pair breaking, as it couples to a cur-\nrent in fermion-flavor space (i.e., to an anti-symmetric\nfermionic bilinear). By rewriting Eq. (102) as Φ( iω) =P\nω′K(iω, iω′)Φ(iω′), we see that Tcis determined as the\nminimal temperature for which the largest eigenvalue of\nthe kernel Kis equal to 1.\nIn the following we first obtain Tcin the spinful variant\nof the x-model as a representative example and later com-\nment on the behavior of Tcin other variants. Specifically,\nwe estimate Tcanalytically by focusing on the “weak”\n(λ≡λx≪1) and “strong” ( λ≫1) coupling regimes,\nwhere Tcis much smaller or larger than the characteris-\ntic energy scale hc,R, respectively. Note that “weak” and\n“strong” coupling regimes do not necessarily correspond\nto small or large values of α. For simplicity we further\nassume that hc/EF≪ |1−α|(to avoid the complications\nresulting from the slowdown of the RG flow of haround\nα= 1) and mainly focus on the parametric dependence\nofTc, ignoring various O(1) coefficients.\n1. Weak coupling (λ≪1)\nIn the weak coupling regime, we rely on the α <1 and\nT= 0 form of the susceptibility given in Eq. (37), and\nassume an ω/T scaling with exponent φ >0 (similar to\nEq. (93))\nχ′′x(ω, T) = sgn( ω)γAα|ω|γ−1\nhγ\nc,R×min\u0012\n1,|ω|\nT\u0013φ\n.(104)\nWe perform analytical continuation to imaginary fre-\nquencies (see App. H) and obtain\nDΦ(iωn)∝γλ\nγ−1 \n1−\f\f\f\fωn\nhc,R\f\f\f\fγ−1!\n.(105)\nIn the FL phase γ >1, the leading piece of DΦat low T\nis constant, resulting in the conventional BCS form of Tc.\nHowever, in the MFL point or NFL phase, where γ≤1,\nDΦ(iω) diverges logarithmically or with exponent γ−1,\nrespectively. Consequently, Tccrosses over from a BCS\nform to an algebraic, quantum critical form. Explicitly,by solving the Eliashberg equation in this regime, we\nobtain\nTc∝hc,R\n\nexp\u0010\n−γ−1\nγλ\u0011\n, γ > 1 +O(p\nλ)\nexp\u0010\n−1√\nλ\u0011\n, γ= 1\n\u0010\nγλ\n1−γ\u00111\n1−γ, γ < 1− O(p\nλ))(106)\nThe results for γ≤1 are similar to those found by [67–69]\nfor other cases of quantum critical pairing. For consis-\ntency, we must require that Tc≪hc,R, which translates\nintoλ≪1. Due to the vanishing of hc,Rasα→1,\nthe problem will eventually cross over to the strong cou-\npling regime, where Tc≫hc,R, beyond some intermedi-\nate value α <1.\n2. Strong coupling (λ≫1)\nWe now consider the transition to superconductivity\nat temperatures T≫hc,R. In this regime, we obtain the\nfinite- TTLS-susceptibility via a combination of scaling\narguments with known results. For details, see App. H.\nWe find that\nχ′′x(ω, T)∝h2\nc\nE2α\nF(max ( |ω|, aαT))2−2α\nω(107)\nwhere aα=O(1). Similarly to the weak coupling limit,\nwe obtain the parametric form of Tcby performing the\nanalytical continuation and solving the linearized Eliash-\nberg equation; see App. H. Remarkably, in an analogous\nfashion to the “weak coupling” regime, we find that Tc\nexhibits a series of crossovers from a quantum critical,\nto a ‘marginal BCS’ [67], to a conventional BCS form as\nthe coupling is increased (rather than decreased, as in\nthe “weak coupling” case). Explicitly, we have that\nTc∝EF\n\n\u0010\nα2\n3−2αϵ\u00111\n3−2α, α < 3/2− O(√ϵ)\nexp\u0010\n−1\nα√ϵ\u0011\n, α= 3/2\nexp\u0000\n−2α−3\nα2ϵ\u0001\n, α > 3/2 +O(√ϵ)(108)\nwhere we defined the small parameter ϵ≡M\nN\u0010\nhc\nEF\u00112\n≪\n1. Interestingly, Tcdecreases up to α= 3, where it has\na local minimum. For larger values of α, it increases and\napproaches the limiting form Tc∝EFexp(−1/αϵ). We\nexpect our results to hold as long as α≪EF/hc(such\nthat αϵ≪1). For consistency of the strong coupling\nanalysis we must require that Tc≫hc,R. While this is\ntrivially fulfilled when α >1, for α <1 this results in the\nrequirement λ≫1, which is, as expected, complimentary\nto the weak coupling condition.\nIntuitively, the reduction of Tcfor larger values of the\ncoupling corresponds to the fact that at finite T, while\nthe TLSs are nearly frozen (i.e. resemble classical impu-\nrities), they preserve their quantum mechanical nature18\nand can thus mediate pairing. The accessible low-energy\nspectral weight for pairing diminishes with the coupling\nstrength and therefore suppresses superconductivity (this\ntrend is reversed beyond α= 3, where the increase in the\ncoupling strength is more significant than the shift of the\nremaining spectral weight to high frequencies).\n3. Superconductivity in other model variants\nFollowing from the discussion of the x-model, we com-\nment on the expected behavior of Tcin generic variants\nof the model. Let us first consider cases with gy= 0\n(i.e., without pair-breaking interactions). In the “weak\ncoupling” regime, in the sense defined above, Tcis deter-\nmined by the behavior of the dominant bath, such that\nit qualitatively follows that of the x-model if αxis domi-\nnant, or otherwise assumes a conventional BCS form (see\nFig. 2). In the “strong coupling” limit, however, the be-\nhavior of Tcis non-universal, namely, it is determined by\nthe susceptibility of the least irrelevant operator. Map-\nping out the quantitative form of Tc(α) necessitate the\nexact renormalized exponents of the TLS-susceptibilities\nand is beyond the scope of our work (given the leading ex-\nponents, the analysis is identical to that of the x-model).\nHowever, recalling that in both cases where αxorαz\ndominates, the strong coupling behavior approaches a\nBCS form, we expect that at intermediate couplings, Tc\nwill smoothly interpolate from a quantum critical to a\nBCS form as αzis increased for fixed αx. Lastly, in-\ntroducing pair-breaking interactions, i.e. a non-zero αy,\nsuppresses Tc[66].\nXI. 1/NCORRECTIONS\nIn this section, we discuss two perturbative corrections\nthat arise at leading order in 1 /N: the validity of the self-\naveraging assumption and the effect of electron-mediated\nTLS-TLS interactions, i.e. RKKY-like interactions.\nA. Validity of self averaging\nAn important assumption of our above analysis lies in\nthe self averaging of the model, which allows us to replace\nthe sum over many TLS-susceptibilities by its mean value\n[with respect to Pr(h)], due to the fact that its variance\nis suppressed by a factor of 1 /M. While this assumption\nis clearly valid in the limit M→ ∞ , for any finite (but\nstill large) Mthe standard deviation may dominate over\nthe mean at sufficiently low energies due to its different\nfrequency dependence. Indeed, consider the variance of\nthe average TLS auto-correlation function in imaginary\ntime:\nVar \n1\nMMX\ni=1\nσi\nx(τ)σi\nx(0)\u000b!\n=1\nMˆ\nPr(hR)S(τ)2dhR.(109)\nFor simplicity let us focus on the regime of interest α <1\nin the x−model. By dimensional considerations, at long\ntimes the dimensionless integral must be proportional to\n(hc,Rτ)−γ(assuming that there is no obstruction to tak-\ning the upper integration limit to ∞, which is the case for\nγ <4). As a result, by taking the square root and trans-\nforming to the frequency domain, we obtain the root-\nmean-square the TLS-susceptibility:\nq\n(χ′′x)2(ω)∼1√\nM|ω|γ/2−1\nhγ/2\nc,R. (110)\nComparing Eq. (110) to the mean in Eq. (37), we con-\nclude that statistical fluctuations can be neglected above\na parametrically small energy scale: ω∼M−1/γhc,R. For\nenergies below this scale, the self-averaging assumption\nis no longer valid and a more systematic treatment of the\n1/M(and 1 /N) fluctuations is needed to determine the\nbehavior of the model.\nB. RKKY interactions\nAnother effect arising when Nis taken to be large but\nfinite, is the emergence of RKKY-like interactions be-\ntween the different TLSs, mediated by the itinerant elec-\ntrons. We analyze this perturbative effect in the spirit of\nRef. [70]. We shall consider the x-model for simplicity,\nthe generalization to other variants is straightforward.\nIncluding the RKKY-like term,\nHRKKY =X\njkgijk,rgi′j′k′,r′\nN2g2\n×Πjk(r−r′, τ−τ′)σx\ni,r(τ)σx\ni′,r′(τ′),(111)\neach TLS will now feel the effect of a sub-ohmic bath aris-\ning from the RKKY coupling to other TLSs, in addition\nto the ohmic particle-hole bath. Following the analysis\nof Ref. [70], this contribution to the bath will be propor-\ntional to χx(iω), and thus the full bath will be of the\nform\nΠ(iω) =α|ω|+λ2\nNh2−γ\nc,R|ω|γ−1(112)\nwith sub-ohmic exponent 2 −γ.\nIn the limit of large yet finite M, N , the sub-ohmic con-\ntribution to the bath may be neglected above the small\nenergy scale ω∼\u0010\nλ2\nαN\u00111/(2−γ)\nhc,R. However, even for\nvery large Nthis energy scale will eventually approach\nhc,Rnearα→1 since λdiverges as hc,R→0. Below this\nscale, the self-consistent approximation of a TLS-induced\nsub-ohmic bath acting on itself breaks down, and a more\nsystematic analysis is needed to determine the behavior\nat very low energies. The low-energy behavior in similar19\ncases [71] suggests that this state remains non-trivial in\nthe sense that γis expected to remain less than 2.\nLastly, note that the subohmic nature of the TLS-\ninduced bath considered above is a result of perturbing\naround the N, M → ∞ saddle point. In a more realistic\nfinite-but-large- Msetting, we expect the subohmic be-\nhaviour to crossover to ohmic below a small energy scale,\ncorresponding to the lowest renormalized splitting of the\nnearby TLSs. In this case, a qualitative change in the\nbehaviour of the TLSs is less obvious, and the system\nmight remain stable to the weak RKKY-like interactions\neven at low energies.\nXII. DISCUSSION AND OUTLOOK\nIn this work, we have studied a class of large Nmod-\nels of itinerant electrons interacting with local two-level\nsystems via spatially random couplings. These mod-\nels, inspired by the possibility of metallic glassiness in\nstrongly correlated materials, exhibit a remarkably rich\nphenomenology at low energies. Most strikingly our the-\nory hosts a robust extended NFL phase in a considerable\npart of parameter space. At the crossover from FL to\nNFL our theory realizes a MFL that shows strange metal-\nlic behavior with T-linear resistivity and Tlog(1/T) spe-\ncific heat. Note that the MFL/NFL behavior does not ne-\ncessitate the existence of a quantum critical point. Physi-\ncally, the departure from FL behavior is rooted in the fact\nthat the characteristic energy of each TLS is algebraically\nsuppressed by the interaction, thus providing significant\nspectral weight of low-energy excitations which consti-\ntute an efficient scattering mechanism for the electronic\ndegrees of freedom. These abundant low-energy excita-\ntions further manifest in a rich phenomenology of the\ncritical transition temperature to the superconducting\nground state of the system.\nThe physical picture of the simplest variant of our the-\nory (the x-model), studied in Ref. [1], qualitatively per-\nsists upon relaxing several simplifying assumptions, such\nas allowing for interactions with different operators of\nthe TLSs and introducing arbitrary TLS-fields. Aiming\nat more realistic models, we further considered the ef-\nfects of relaxing additional simplifications, such as 1 /N\ncorrections, spatial correlations in gijl,rand the self av-\neraging assumption. While these tend to suppress the\nNFL behaviour found in this work below some energy\nscale suppressed by powers of N, there are physical rea-\nsons to think that this scale remains small in a realistic\nsetting. Specifically, recalling that TLSs in physical sys-\ntems are extended objects, the interaction would retain a\nhigh degree of connectivity (i.e. each TLS would interact\nwith many electrons and vice versa), which in turn could\npreserve the self averaging property, and frustrate effects\nof RKKY-like interactions.\nIt is interesting to ask what is the relation between the\ninteraction strengths ( αx,y,z) to actual physical knobs in\nrealistic systems. This is a complicated question as themicroscopic origin of such TLSs is not well understood.\nHowever, there have been many studies attempting to\nprovide a microscopic theoretical framework for under-\nstanding these objects [72–77]. It is possible that as the\nsystem approaches a glassy charge or spin ordering tran-\nsition, the shape, size and other properties of these TLSs\nchange, affecting the magnitude of their coupling to elec-\ntrons, or the relative sizes of the couplings to the x, y, z\noperators. Thus, tuning a physical knob of the system\ncould be parameterized as a nontrivial path in the space\nof couplings, leading to a nontrivial variation of the ex-\nponent in the electronic self energy.\nTo this end, another issue concerns the density of states\nof TLSs, which is parametrically larger than that of the\nelectrons (i.e. h−1\nc,R≫ρF). A direct consequence is the\nseemingly enhanced coupling λ≡M\nNρF\nhc,Rαthat appears\nin the electronic self energy. It appears, however, natural\nto expect that α∼λat least up to some intermediate\ncoupling strength. This is the case if M/N ∼h−1\nc,R/ρF≪\n1, i.e., if the TLSs are sparse compared to the electrons.\nPhysically, this seems plausible based on the mesoscopic\nconsiderations mentioned above.\nNon-Fermi liquid behavior is ultimately tied to an\nanomalous spectrum of gapless excitations. Such a spec-\ntrum is usually believed to emerge from collective modes\nwith soft long wavelength fluctuations. As we showed in\nthis paper, it can also be the result of quantum fluctu-\nations of modes that are localized in a region of size l,\nwhere each mode has an excitation gap Emin∼hRbut\nis governed by a singular distribution function P(hR)∝\nhγ−1\nRwith γ >0. Even if the correlation function for a\ngiven hRdecays rapidly in time, χhR(τ)∼exp (−hRτ),\nthe average χav(τ) =´\ndhRP(hR)χhR(τ) then decays\nlike a powerlaw ∝τ−γand the system becomes critical.\nFor the static susceptibility, χav(T) =´1/Tdτχav(τ), it\nfollows that χav(T)∝Tγ−1andC∝Tγfor the heat\ncapacity. Non-Fermi liquid behavior occurs for γ <1.\nSuch a singular distribution function was also obtained\nfrom quantum Griffiths behavior [78]. Let us therefore\ncompare and contrast our results with the ones that fol-\nlow from quantum Griffiths physics, where rare, large\ndroplets of size loccur with probability pl∝e−cldand\npossess an exponentially small gap hl∝e−bld[78]. This\nyields a power-law for\nP(hR) =ˆ\ndldplδ(hR−hl)∝hγ−1\nR, (113)\nwith non-universal exponent γ=b/c. Exponentially\nsmall gaps occur for the random transverse field Ising\nmodel [78]. However, as soon as one includes the coupling\nto conduction electrons, large droplets will freeze by the\nCaldeira-Leggett mechanism, and one rather finds super-\nparamagnetic behavior of classical droplets [79, 80]. On\nthe other hand, for systems with a continuous order pa-\nrameter symmetry power law quantum Griffiths behav-\nior becomes possible even in the presence of particle-hole\nexcitations [81]. This behavior was also seen in recent20\nnumerical simulations [29]. In contrast to this quantum\nGriffiths behavior, in our approach we consider the cou-\npling of TLSs of characteristic size of several lattice spac-\nings to conduction electrons. While isolated TLSs are\ngoverned by P(h)∝hβthat is, on its own, not suffi-\nciently singular ( β > 0), strong local quantum fluctua-\ntions due to the coupling to conduction-electrons renor-\nmalize the excitation gap h→hR∼h1/(1−α), which\nreduces the exponent β+ 1→βR+ 1 = ( β+ 1)(1 −α).\nWhile our theory does not aim to realistically describe\nany specific material, the existence of a tunable non-\nFermi liquid phase in a controlled microscopic theory\ncould shed new light on some aspects of strange metallic-\nity. It provides a novel viewpoint on the widely observed\nextended strange metal regime [15–17, 82, 83] that does\nnot rely on a putative quantum critical point. Further,\nwhile conventional wisdom typically interprets the resis-\ntivity in terms of a T−linear and a T2components, i.e.\nρ−ρ0=AT+BT2[5, 15, 16, 82, 84–87], our theory\noffers an alternative interpretation7where the exponent\nis a continuous parameter, ρ−ρ0=CTγ. Interestingly,\nthis interpretation (also known as power-law liquid) has\nbeen shown to be consistent with experimental data of\nstrange metals [26, 88–91].\nSeveral natural questions remain open. Aiming to\nbetter understand more realistic scenarios, a systematicstudy of our model for finite- Nis called for, either by\nanalytical or numerical methods. In addition, the behav-\nior deep inside the superconducting state might exhibit\ninteresting new physics, as the electrons constituting the\nOhmic bath in the normal state are becoming gapped,\nwhich has non-trivial effects on the TLSs and vice versa.\nMore broadly, one may consider various other physical\nsystems containing a coexistence of electrons and two-\nlevel systems, where the framework developed in this\nwork can be applied.\nAcknowledgements.— We thank G. Grissonnanche, S.\nA. Kivelson, C. Murthy, A. Pandey, B. Ramshaw, and B.\nSpivak for numerous discussions and for a collaboration\non prior unpublished work. We are grateful to Natan An-\ndrei, Girsh Blumberg, Andrey Chubukov, Rafael Fernan-\ndes, Tobias Holder, Yuval Oreg, and Alexander Shnirman\nfor helpful discussions. J.S. was supported by the Ger-\nman Research Foundation (DFG) through CRC TRR 288\n“ElastoQMat,” project B01 and a Weston Visiting Pro-\nfessorship at the Weizmann Institute of Science. E.B. was\nsupported by the European Research Council (ERC) un-\nder grant HQMAT (Grant Agreement No. 817799) and\nby the Israel-US Binational Science Foundation (BSF).\nSome of this work was performed at KITP, supported\nin part by the National Science Foundation under PHY-\n1748958.\n[1] N. Bashan, E. Tulipman, J. Schmalian, and E. Berg, Tun-\nable non-fermi liquid phase from coupling to two-level\nsystems (2023), arXiv:2310.07768.\n[2] C. M. Varma, Colloquium: Linear in temperature resis-\ntivity and associated mysteries including high tempera-\nture superconductivity, Reviews of Modern Physics 92,\n031001 (2020).\n[3] C. Proust and L. Taillefer, The Remarkable Underlying\nGround States of Cuprate Superconductors, Annual Re-\nview of Condensed Matter Physics 10, 409 (2019).\n[4] J. a. N. Bruin, H. Sakai, R. S. Perry, and A. P. Macken-\nzie, Similarity of Scattering Rates in Metals Showing T-\nLinear Resistivity, Science 339, 804 (2013).\n[5] A. Legros, S. Benhabib, W. Tabis, F. Lalibert´ e, M. Dion,\nM. Lizaire, B. Vignolle, D. Vignolles, H. Raffy, Z. Z.\nLi, P. Auban-Senzier, N. Doiron-Leyraud, P. Fournier,\nD. Colson, L. Taillefer, and C. Proust, Universal T -\nlinear resistivity and Planckian dissipation in overdoped\ncuprates, Nature Physics 15, 142 (2019).\n[6] Y. Cao, D. Chowdhury, D. Rodan-Legrain, O. Rubies-\nBigorda, K. Watanabe, T. Taniguchi, T. Senthil,\nand P. Jarillo-Herrero, Strange Metal in Magic-Angle\nGraphene with near Planckian Dissipation, Physical Re-\nview Letters 124, 076801 (2020).\n[7] D. Chowdhury, A. Georges, O. Parcollet, and S. Sachdev,\nSachdev-Ye-Kitaev models and beyond: Window into\n7Following the procedure in Ref. [82], the derivative dρ/dT can\ndistinguish the two behaviors provided that the T→0 resistivity\nis measured with sufficient accuracy.non-Fermi liquids, Reviews of Modern Physics 94, 035004\n(2022).\n[8] S. A. Hartnoll and A. P. Mackenzie, Colloquium: Planck-\nian dissipation in metals, Rev. Mod. Phys. 94, 041002\n(2022).\n[9] P. W. Phillips, N. E. Hussey, and P. Abbamonte, Stranger\nthan metals, Science 377, eabh4273 (2022).\n[10] S. Sachdev, Quantum statistical mechanics of the\nSachdev-Ye-Kitaev model and strange metals (2023).\n[11] H. v. L¨ ohneysen, A. Rosch, M. Vojta, and P. W¨ olfle,\nFermi-liquid instabilities at magnetic quantum phase\ntransitions, Reviews of Modern Physics 79, 1015 (2007).\n[12] C. M. Varma, Z. Nussinov, and W. van Saarloos, Singular\nor non-Fermi liquids, Physics Reports 361, 267 (2002).\n[13] S.-S. Lee, Recent Developments in Non-Fermi Liquid\nTheory, Annual Review of Condensed Matter Physics 9,\n227 (2018).\n[14] P. Gegenwart, Q. Si, and F. Steglich, Quantum criticality\nin heavy-fermion metals, Nature Physics 4, 186 (2008).\n[15] R. A. Cooper, Y. Wang, B. Vignolle, O. J. Lipscombe,\nS. M. Hayden, Y. Tanabe, T. Adachi, Y. Koike, M. No-\nhara, H. Takagi, C. Proust, and N. E. Hussey, Anomalous\nCriticality in the Electrical Resistivity of La 2–xSrxCuO 4,\nScience 323, 603 (2009).\n[16] N. E. Hussey, J. Buhot, and S. Licciardello, A tale of\ntwo metals: contrasting criticalities in the pnictides and\nhole-doped cuprates*, Reports on Progress in Physics 81,\n052501 (2018).\n[17] N. E. Hussey, H. Gordon-Moys, J. Kokalj, and R. H.\nMcKenzie, Generic strange-metal behaviour of overdoped\ncuprates, Journal of Physics: Conference Series 449,21\n012004 (2013).\n[18] R. L. Greene, P. R. Mandal, N. R. Poniatowski, and\nT. Sarkar, The Strange Metal State of the Electron-\nDoped Cuprates, Annual Review of Condensed Matter\nPhysics 11, 213 (2020).\n[19] L. Taillefer, Scattering and pairing in cuprate supercon-\nductors, Annu. Rev. Condens. Matter Phys. 1, 51 (2010).\n[20] A. Jaoui, I. Das, G. Di Battista, J. D´ ıez-M´ erida, X. Lu,\nK. Watanabe, T. Taniguchi, H. Ishizuka, L. Levitov, and\nD. K. Efetov, Quantum critical behaviour in magic-angle\ntwisted bilayer graphene, Nature Physics 18, 633 (2022).\n[21] A. Ghiotto, E.-M. Shih, G. S. S. G. Pereira, D. A.\nRhodes, B. Kim, J. Zang, A. J. Millis, K. Watanabe,\nT. Taniguchi, J. C. Hone, L. Wang, C. R. Dean, and\nA. N. Pasupathy, Quantum criticality in twisted transi-\ntion metal dichalcogenides, Nature 597, 345 (2021).\n[22] C. Pfleiderer, Non-Fermi liquid puzzle of MnSi at high\npressure, Physica B: Condensed Matter Proceedings of\nthe Second Hiroshima Workshop on Transport and The\nrmal Properties of Advanced Materials, 328, 100 (2003).\n[23] C. Pfleiderer, D. Reznik, L. Pintschovius, H. v. L¨ ohney-\nsen, M. Garst, and A. Rosch, Partial order in the non-\nFermi-liquid phase of MnSi, Nature 427, 227 (2004).\n[24] N. Doiron-Leyraud, I. R. Walker, L. Taillefer, M. J.\nSteiner, S. R. Julian, and G. G. Lonzarich, Fermi-liquid\nbreakdown in the paramagnetic phase of a pure metal,\nNature 425, 595 (2003).\n[25] S. Paschen and Q. Si, Quantum phases driven by strong\ncorrelations, Nature Reviews Physics 3, 9 (2021).\n[26] H. Zhao, J. Zhang, M. Lyu, S. Bachus, Y. Tokiwa,\nP. Gegenwart, S. Zhang, J. Cheng, Y.-f. Yang, G. Chen,\nY. Isikawa, Q. Si, F. Steglich, and P. Sun, Quantum-\ncritical phase from frustrated magnetism in a strongly\ncorrelated metal, Nature Physics 15, 1261 (2019).\n[27] F. C. Mocanu, L. Berthier, S. Ciarella, D. Khomenko,\nD. R. Reichman, C. Scalliet, and F. Zamponi, Micro-\nscopic observation of two-level systems in a metallic glass\nmodel, The Journal of Chemical Physics 158, 014501\n(2023).\n[28] J. Schmalian and P. G. Wolynes, Stripe Glasses: Self-\nGenerated Randomness in a Uniformly Frustrated Sys-\ntem, Physical Review Letters 85, 836 (2000).\n[29] A. A. Patel, P. Lunts, and S. Sachdev, Localization of\noverdamped bosonic modes and transport in strange met-\nals (2023), arXiv:2312.06751 [cond-mat.str-el].\n[30] J. L. Black, B. L. Gyorffy, and J. J¨ ackle, On the resistivity\ndue to two-level systems in metallic glasses, Philosophical\nMagazine B 40, 331 (1979).\n[31] M. A. Ruderman and C. Kittel, Indirect Exchange Cou-\npling of Nuclear Magnetic Moments by Conduction Elec-\ntrons, Physical Review 96, 99 (1954).\n[32] T. Kasuya, A Theory of Metallic Ferro- and Antifer-\nromagnetism on Zener’s Model, Progress of Theoretical\nPhysics 16, 45 (1956).\n[33] K. Yosida, Magnetic Properties of Cu-Mn Alloys, Physi-\ncal Review 106, 893 (1957).\n[34] W. A. Phillips, Tunneling states in amorphous solids,\nJournal of Low Temperature Physics 7, 351 (1972).\n[35] A. J. Leggett, S. Chakravarty, A. T. Dorsey, M. P. A.\nFisher, A. Garg, and W. Zwerger, Dynamics of the dissi-\npative two-state system, Reviews of Modern Physics 59,\n1 (1987).\n[36] T. A. Costi and G. Zar´ and, Thermodynamics of the dissi-\npative two-state system: A Bethe-ansatz study, PhysicalReview B 59, 12398 (1999).\n[37] U. Weiss, Quantum Dissipative Systems, 4th ed.\n(WORLD SCIENTIFIC, 2012).\n[38] P. B. Allen and R. C. Dynes, Transition temperature of\nstrong-coupled superconductors reanalyzed, Phys. Rev.\nB12, 905 (1975).\n[39] A. Auerbach, Interacting Electrons and Quantum Magnetism\n(1998).\n[40] I. Esterlis and J. Schmalian, Cooper pairing of incoherent\nelectrons: An electron-phonon version of the Sachdev-Ye-\nKitaev model, Physical Review B 100, 115132 (2019).\n[41] A. O. Caldeira and A. J. Leggett, Influence of Dissipation\non Quantum Tunneling in Macroscopic Systems, Physical\nReview Letters 46, 211 (1981).\n[42] A. O. Caldeira and A. J. Leggett, Quantum tunnelling in\na dissipative system, Annals of Physics 149, 374 (1983).\n[43] K. L. Hur, Entanglement entropy, decoherence, and\nquantum phase transitions of a dissipative two-level sys-\ntem, Annals of Physics 323, 2208 (2008).\n[44] G. Camacho, P. Schmitteckert, and S. T. Carr, Ex-\nact equilibrium results in the interacting resonant level\nmodel, Physical Review B 99, 085122 (2019).\n[45] V. M. Filyov and P. B. Wiegmann, A method for solving\nthe Kondo problem, Physics Letters A 76, 283 (1980).\n[46] V. V. Ponomarenko, Resonant tunneling and low-energy\nimpurity behavior in a resonant-level model, Physical Re-\nview B 48, 5265 (1993).\n[47] T. A. Costi and C. Kieffer, Equilibrium Dynamics of the\nDissipative Two-State System, Physical Review Letters\n76, 1683 (1996).\n[48] F. Guinea, Dynamics of simple dissipative systems, Phys-\nical Review B 32, 4486 (1985).\n[49] P. W. Anderson, G. Yuval, and D. R. Hamann, Exact\nResults in the Kondo Problem. II. Scaling Theory, Qual-\nitatively Correct Solution, and Some New Results on\nOne-Dimensional Classical Statistical Models, Physical\nReview B 1, 4464 (1970).\n[50] F. Guinea, V. Hakim, and A. Muramatsu, Bosonization\nof a two-level system with dissipation, Physical Review\nB32, 4410 (1985).\n[51] M. Sassetti and U. Weiss, Correlation functions for dis-\nsipative two-state systems: Effects of the initial prepara-\ntion, Physical Review A 41, 5383 (1990).\n[52] G. Zar´ and and E. Demler, Quantum phase transitions\nin the Bose-Fermi Kondo model, Physical Review B 66,\n024427 (2002).\n[53] A. H. Castro Neto, E. Novais, L. Borda, G. Zar´ and, and\nI. Affleck, Quantum Magnetic Impurities in Magnetically\nOrdered Systems, Physical Review Letters 91, 096401\n(2003).\n[54] E. Novais, A. H. Castro Neto, L. Borda, I. Affleck, and\nG. Zarand, Frustration of decoherence in open quantum\nsystems, Physical Review B 72, 014417 (2005).\n[55] H. Kohler, A. Hackl, and S. Kehrein, Nonequilibrium dy-\nnamics of a system with quantum frustration, Physical\nReview B 88, 205122 (2013).\n[56] R. Belyansky, S. Whitsitt, R. Lundgren, Y. Wang,\nA. Vrajitoarea, A. A. Houck, and A. V. Gorshkov,\nFrustration-induced anomalous transport and strong\nphoton decay in waveguide QED, Physical Review Re-\nsearch 3, L032058 (2021).\n[57] N. Bashan, E. Tulipman, J. Schmalian, and E. Berg, To\nappear.\n[58] D. V. Khveshchenko, Quantum impurity models of noisy22\nqubits, Physical Review B 69, 153311 (2004).\n[59] D. Chowdhury, Y. Werman, E. Berg, and T. Senthil,\nTranslationally Invariant Non-Fermi-Liquid Metals with\nCritical Fermi Surfaces: Solvable Models, Physical Re-\nview X 8, 031024 (2018).\n[60] A. A. Patel, J. McGreevy, D. P. Arovas, and S. Sachdev,\nMagnetotransport in a Model of a Disordered Strange\nMetal, Physical Review X 8, 021049 (2018).\n[61] T. P. Devereaux and R. Hackl, Inelastic light scatter-\ning from correlated electrons, Rev. Mod. Phys. 79, 175\n(2007).\n[62] U. Karahasanovic, F. Kretzschmar, T. B¨ ohm, R. Hackl,\nI. Paul, Y. Gallais, and J. Schmalian, Manifestation of\nnematic degrees of freedom in the raman response func-\ntion of iron pnictides, Phys. Rev. B 92, 075134 (2015).\n[63] E. Tulipman and E. Berg, A criterion for strange metal-\nlicity in the Lorenz ratio, npj Quantum Materials 8, 1\n(2023).\n[64] R. G¨ orlich and U. Weiss, Specific heat of the dissipative\ntwo-state system, Physical Review B 38, 5245 (1988).\n[65] T. A. Costi, Scaling and Universality in the Anisotropic\nKondo Model and the Dissipative Two-State System,\nPhysical Review Letters 80, 1038 (1998).\n[66] D. Hauck, M. J. Klug, I. Esterlis, and J. Schmalian,\nEliashberg equations for an electron–phonon version of\nthe sachdev–ye–kitaev model: Pair breaking in non-fermi\nliquid superconductors, Annals of Physics 417, 168120\n(2020), eliashberg theory at 60: Strong-coupling super-\nconductivity and beyond.\n[67] D. T. Son, Superconductivity by long-range color mag-\nnetic interaction in high-density quark matter, Phys.\nRev. D 59, 094019 (1999).\n[68] A. Abanov and A. V. Chubukov, Interplay between su-\nperconductivity and non-fermi liquid at a quantum criti-\ncal point in a metal. i. the γmodel and its phase diagram\natT= 0: The case 0 < γ < 1, Phys. Rev. B 102, 024524\n(2020).\n[69] Y.-M. Wu, A. Abanov, Y. Wang, and A. V. Chubukov,\nInterplay between superconductivity and non-fermi liq-\nuid at a quantum critical point in a metal. ii. the γmodel\nat a finite Tfor 0 < γ < 1, Phys. Rev. B 102, 024525\n(2020).\n[70] V. Dobrosavljevi´ c and E. Miranda, Absence of conven-\ntional quantum phase transitions in itinerant systems\nwith disorder, Phys. Rev. Lett. 94, 187203 (2005).\n[71] A. M. Sengupta and A. Georges, Non-fermi-liquid behav-\nior near a T= 0 spin-glass transition, Phys. Rev. B 52,\n10295 (1995).\n[72] V. Lubchenko and P. G. Wolynes, Intrinsic quantum exci-\ntations of low temperature glasses, Physical Review Let-\nters87, 195901 (2001).\n[73] J. Burnett, L. Faoro, I. Wisby, V. L. Gurtovoi,\nA. V. Chernykh, G. M. Mikhailov, V. A. Tulin,\nR. Shaikhaidarov, V. Antonov, P. J. Meeson, etal., Evi-\ndence for interacting two-level systems from the 1/f noise\nof a superconducting resonator, Nature Communications\n5, 4119 (2014).\n[74] D. Khomenko, C. Scalliet, L. Berthier, D. R. Reichman,\nand F. Zamponi, Depletion of two-level systems in ultra-\nstable computer-generated glasses, Physical review let-\nters124, 225901 (2020).\n[75] D. Khomenko, D. R. Reichman, and F. Zamponi, Re-\nlationship between two-level systems and quasilocalized\nnormal modes in glasses, Physical Review Materials 5,055602 (2021).\n[76] W. Ji, T. W. de Geus, E. Agoritsas, and M. Wyart, Mean-\nfield description for the architecture of low-energy excita-\ntions in glasses, Physical Review E 105, 044601 (2022).\n[77] F. C. Mocanu, L. Berthier, S. Ciarella, D. Khomenko,\nD. R. Reichman, C. Scalliet, and F. Zamponi, Micro-\nscopic observation of two-level systems in a metallic glass\nmodel, The Journal of Chemical Physics 158(2023).\n[78] M. Thill and D. Huse, Equilibrium behaviour of quantum\nising spin glass, Physica A: Statistical Mechanics and its\nApplications 214, 321 (1995).\n[79] A. Millis, D. Morr, and J. Schmalian, Local defect in\nmetallic quantum critical systems, Physical Review Let-\nters87, 167202 (2001).\n[80] A. J. Millis, D. K. Morr, and J. Schmalian, Quantum grif-\nfiths effects in metallic systems, Phys. Rev. B 66, 174433\n(2002).\n[81] T. Vojta and J. Schmalian, Quantum griffiths effects in\nitinerant heisenberg magnets, Phys. Rev. B 72, 045438\n(2005).\n[82] N. E. Hussey, R. A. Cooper, X. Xu, Y. Wang,\nI. Mouzopoulou, B. Vignolle, and C. Proust, Dichotomy\nin the T-linear resistivity in hole-doped cuprates, Philo-\nsophical Transactions of the Royal Society A: Mathemat-\nical, Physical and Engineering Sciences 369, 1626 (2011).\n[83] P. M. Lozano, G. D. Gu, Q. Li, and J. M. Tranquada,\nStrange-metal behavior in La 2−xSrxCuO 4beyond the\nstripe-percolation transition (2023), arXiv:2307.13740\n[cond-mat].\n[84] T. Park, V. A. Sidorov, F. Ronning, J.-X. Zhu,\nY. Tokiwa, H. Lee, E. D. Bauer, R. Movshovich, J. L.\nSarrao, and J. D. Thompson, Isotropic quantum scatter-\ning and unconventional superconductivity, Nature 456,\n366 (2008).\n[85] N. Doiron-Leyraud, P. Auban-Senzier, S. Ren´ e de Cotret,\nC. Bourbonnais, D. J´ erome, K. Bechgaard, and L. Taille-\nfer, Correlation between linear resistivity and Tcin\nthe bechgaard salts and the pnictide superconduc-\ntor Ba(Fe 1−xCox)2As2, Physical Review B 80, 214531\n(2009).\n[86] K. Jin, N. P. Butch, K. Kirshenbaum, J. Paglione, and\nR. L. Greene, Link between spin fluctuations and electron\npairing in copper oxide superconductors, Nature 476, 73\n(2011).\n[87] M. Berben, J. Ayres, C. Duffy, R. D. H. Hinlopen, Y.-T.\nHsu, M. Leroux, I. Gilmutdinov, M. Massoudzadegan,\nD. Vignolles, Y. Huang, T. Kondo, T. Takeuchi, J. R.\nCooper, S. Friedemann, A. Carrington, C. Proust, and\nN. E. Hussey, Compartmentalizing the cuprate strange\nmetal (2022).\n[88] Y. Dagan, M. M. Qazilbash, C. P. Hill, V. N. Kulkarni,\nand R. L. Greene, Evidence for a Quantum Phase Transi-\ntion in Pr 2−xCexCuO 4−δfrom Transport Measurements,\nPhysical Review Letters 92, 167001 (2004).\n[89] T. J. Reber, X. Zhou, N. C. Plumb, S. Parham,\nJ. A. Waugh, Y. Cao, Z. Sun, H. Li, Q. Wang, J. S.\nWen, Z. J. Xu, G. Gu, Y. Yoshida, H. Eisaki, G. B.\nArnold, and D. S. Dessau, Power Law Liquid - A\nUnified Form of Low-Energy Nodal Electronic Interac-\ntions in Hole Doped Cuprate Superconductors (2015),\narXiv:1509.01611 [cond-mat].\n[90] S. Smit, E. Mauri, L. Bawden, F. Heringa, F. Ger-\nritsen, E. van Heumen, Y. K. Huang, T. Kondo,\nT. Takeuchi, N. E. Hussey, T. K. Kim, C. Cacho,23\nA. Krikun, K. Schalm, H. T. C. Stoof, and M. S. Golden,\nMomentum-dependent scaling exponents of nodal self-\nenergies measured in strange metal cuprates and mod-\nelled using semi-holography (2021), arXiv:2112.06576\n[cond-mat].\n[91] K. Harada, Y. Teramoto, T. Usui, K. Itaka, T. Fu-\njii, T. Noji, H. Taniguchi, M. Matsukawa, H. Ishikawa,\nK. Kindo, D. S. Dessau, and T. Watanabe, Revised\nphase diagram of the high- Tccuprate superconductor Pb-\ndoped Bi 2Sr2CaCu 2O8+δrevealed by anisotropic trans-port measurements, Physical Review B 105, 085131\n(2022).\n[92] W. Gasser, E. Heiner, and K. Elk,\nGreensche Funktionen in Festk¨ orper- und Vielteilchenphysik\n(Wiley, 2002).\n[93] A. V. Chubukov and J. Schmalian, Superconductivity\ndue to massless boson exchange in the strong-coupling\nlimit, Phys. Rev. B 72, 174520 (2005).\n[94] A. J. Bray and M. A. Moore, Influence of dissipation on\nquantum coherence, Phys. Rev. Lett. 49, 1545 (1982).\nAppendix A: Diagrammatic approach to mapping\nWe now present an alternative approach for the mapping to the spin-boson model, where we demonstrate, by a\nperturbative expansion of the interaction, that the electrons constitute an Ohmic bath to the TLSs. Importantly, due\nto the spatial randomness of the couplings, the bath is that of non-interacting particle-hole pairs, i.e., it is independent\nof the electronic self energy. Consider the correlation function for a single TLS of flavor s:Sx\ns(τ)≡ ⟨Tτ{σx\ns(τ)σx\ns(0)}⟩.\nThe expansion in interaction vertices reads (we suppress the spatial index rsince all operators act on the same site):\nSx\ns(τ) =*\nTτ(\nσx\ns(τ)σx\ns(0)∞X\nn=0(−1)n\nn!\"Y\ni=1...n ˆ\ndτiX\nabcgabc\nNσx\na(τi)c†\nb(τi)cc(τi)!#)+\n, (A1)\nwhich decouples into a sum of terms of the form\nIn=ˆ\nτ1,τ2...τnX\na1b1c1a2b2c2...\u0010ga1b1c1\nNga2b2c2\nN...\u0011\nTτ\b\nσx\ns(τ)σx\ns(0)\u0000\nσx\na1(τi)...σx\nan(τn)\u0001\t\u000b\n×D\nTτn\nc†\nb1(τi1)cc1(τi1)c†\nb2(τi2)cc2(τi2)...oE\n(A2)\nBy integrating over the realizations of gabc, we note that (i) terms where all interaction TLS indices a1, ..., a n̸=sare\n“disconnected” and cancel with the vacuum diagrams; (ii) if only some (but not all) ai=s, the contribution is either\nsubleading in 1 /Nor corresponds to a self-energy insertion for the electrons (see Fig. 6). Thus if we treat the electrons\nself consistently as being fully dressed, the only relevant insertions of the interaction are those in which ai=s.\nFigure 6. Examples of diagrams considered in the mapping. a) If all insertions are of other TLSs, the bubble disconnects from\nthe “external vertices” and cancels with the vacuum diagrams. b) Other insertions of different TLSs become subleading in N.\nc) The contribution of insertions of a different TLS a̸=scan be absorbed into the full electron green function. d) Example of\na contributing diagram, with the wavy lines representing particle hole pairs (as defined in e)). Here electrons are denoted by\nsolid line, contractions over realizations of gabcby dashed lines and TLS operators, who do not admit a direct diagrammatic\nexpansion, by full circles.\nHence Incan be written as (all TLS indices = s)\nIn=ˆ\nτ1,τ2...X\nb1c1b2c2...\u0010gsb1c1\nNgsb2c2\nN...\u0011\n⟨Tτ{σx\ns(τ)σx\ns(0) (σx\ns(τi)...σx\ns(τn))}⟩24\n×D\nTτn\nc†\nb1(τi1)cc1(τi1)c†\nb2(τi2)cc2(τi2)...oE\n(A3)\nConsidering the electronic part (including the couplings gsbici), we see that the leading order contribution in 1 /N\ncorresponds to terms with b1̸=c1̸=b2...(all indices are distinct), namely,\nCn≡X\nb1c1b2c2...1\nNn\u0000\ng2\nsb1c1g2\nsb2c2...\u0001D\nTτn\nc†\nb1(τi1)cc1(τi1)c†\nc1(τi2)cb1(τi2)oE\n(A4)\n×D\nTτn\nc†\nb2(τi1)cc2(τi3)c†\nc2(τi4)cb2(τi4)oE\n...+ permutations + O\u00121\nN\u0013\n. (A5)\nThe particle-hole pairs obey the bosonic Wick’s theorem since the couplings gabcobey it. Therefore,\nCn=˜J(τi1−τi2)˜J(τi3−τi4)...+ permutations (A6)\nwhere we have denoted\n˜J(τi1−τi2)≡g2G(τi1−τi2)G(τi2−τi1) (A7)\nAs mentioned earlier, the fact that the bath is Ohmic follows from spatial randomness of the couplings, which translates\nto the TLS being coupled to the local particle-hole correlators, with G(τ) =´\nkG(τ,k). Upon analytical continuation,\nthe spectral function of the bath is given by\n˜J(ω) =g2ρ2\nF\n2πω≡ηω, (A8)\nsuch that the dimensionless coupling strength (using the conventions of Ref. [37]) is given by\nα≡2\nπη=g2ρ2\nF\nπ2. (A9)\nGeneralizing the derivation to the xyz−model is straightforward.\nAppendix B: Details of RG flow of 1bSB\nWe present here the calculation of the renormalized scale hRin the different regimes I −III in the 1bSB. As a\nreminder, the flow equations are, to order ˜h2:\ndα\ndℓ=−˜h2α (B1)\nd˜h\ndℓ= (1−α)˜h (B2)\nWe allow the couplings to flow until there is only one energy scale in the problem, i.e. until the cutoff Λ = ωc∗exp(−ℓ)\nand TLS energy hR=˜hΛ′become equal, which is given by ˜h(ℓ∗) = 1.\nWe start with the regime where, (1 −α)≫˜h2, so that the flow of αis much slower than the flow of ˜h. We can thus\nsolve the flow of ˜hwhile treating αas a constant, and the weak change in αat the end of the flow (when ˜happroaches\n1) will only change the result by a multiplicative factor, which is absorbed into the definition of the prefactor cαin\nEq. (B4). Therefore, allowing ˜hto flow until it reaches the value 1 we find that\n(1−α)ℓ∗= logωc\nh(B3)\nInserting Eq. (B3) into the definition of hR, we find that\nhR=cαωc\u0012h\nωc\u00131\n1−α\n(B4)\nAs mentioned in the main text, this prefactor cαcannot be determined merely from the RG flow. However, it can be\nextracted using exact techniques such as bosonization or Bethe ansatz [43–46], and is given by\ncα= (Γ(1 −α) exp ( αlogα+ (1−α) log(1 −α)))1\n1−α (B5)25\nwhich satisfies the two known limits c0= 1, c1/2=π/4. We now study the regime near the BKT transition, α≈1.\nIn this regime, we define J= 1−αsuch that |J| ≪1. The RG equations then approximately become:\ndJ\ndℓ=˜h2(B6)\nd˜h\ndℓ=˜hJ (B7)\nNote that the combination x0≡˜h2−J2obeys:\n1\n2dx0\ndℓ=˜hd˜h\ndℓ−JdJ\ndℓ= 0 (B8)\nsox0is constant along flow lines. Using this relation the equations can thus be solved easily\nd˜h\ndℓ=˜hq\n˜h2−x0 (B9)\nℓ∗=ˆ1\n˜h0d˜h\n˜hp˜h2−x0. (B10)\nwhere ˜h0=h/ωc. We now separate to the cases where x0>0 and x0<0. If x0<0, we have that\nℓ∗=asinh\u0010√−x0\n˜h0\u0011\n−asinh (√−x0)\n√−x0(B11)\nhR=ωc\n√−x0\n˜h0+q\n1−x0\n˜h2\n0√−x0+√1−x0\n−1√\n−x0\n(B12)\nIf˜h2\n0≪Jthis expression simplifies to the power law given earlier. On the other hand, for x0→0 this expression\nbecomes the familiar Kondo scale of the isotropic Kondo model hR∝ωcexp(−1/˜h0). For x0>0, we obtain that\nℓ∗=atan\u0010q\n1−x0\nx0\u0011\n−sgn(J)atan\u0012q\n˜h2\n0−x0\nx0\u0013\n√x0. (B13)\nTaking x0→0 also gives the isotropic Kondo result ℓ= 1/˜h0. Setting J= 0 we find that ℓ=π/2˜h0. Therefore,\nin this regime we can approximately think of the renormalized scale as taking the form hR∝ωcexp(−b(J,˜h0)/˜h0)\nwith bbeing a slowly varying function of order 1. However, when the flow approaches the BKT line J=−˜hthis\napproximation does not hold, and instead the renormalized scale is set by the distance from the transition:\nhR∝ωcexp\u0012\n−π√x0\u0013\n. (B14)\nAppendix C: Sum rules for the 1bSB\nThis is based on a short analysis first derived in [48]. We define the correlation function\n⟨σx(t)σx(0)⟩=ˆ\nωeiωtAx(ω) (C1)\nThis is related to the dynamical susceptibility by the fluctuation dissipation theorem [39]:\nAx(ω) =\u0012\n1 + coth\u0012βω\n2\u0013\u0013\nχ′′\nx(ω) (C2)26\nAdditionally, the following equation of motion derives from the Hamiltonian Eq. (23):\nidσx\ndt=−2hσy (C3)\nThus, by Fourier transforming ⟨σx(t)σx(0)⟩,\ndσx\ndt(t)σx(0)\u000b\nand\ndσx\ndt(t)dσx\ndt(0)\u000b\n, setting t= 0 and using the antisym-\nmetry of χ′′\nx(ω) we obtain the three sum rules:\n1 =ˆ\nωχ′′\nx(ω)coth\u0012βω\n2\u0013\n(C4)\n2h⟨σz⟩=ˆ\nωωχ′′\nx(ω) (C5)\n4h2=ˆ\nωω2χ′′\nx(ω)coth\u0012βω\n2\u0013\n(C6)\nAppendix D: Explicit calculation of χ′′xfor the x−model\nThe average (imaginary part of the) susceptibility is given by\nχ′′x=ˆhc\n0χ′′\nx(ω, h)Pβ(h)dh (D1)\n=ˆhc,R\n01\nωfα\u0012ω\nhR\u0013\nPr(hR)dhR (D2)\n= sgn( ω)ˆ∞\n|ω|/hc,Rfα(x)\nx2Pr(|ω|/x)dx (D3)\nThe result of this integral thus depends on the renormalized distribution Pr.\n1.α <1\nStarting with hR=cαωc(h/ωc)1/(1−α)⇒h= (hR/cα)1−αω−α\nc:\nP(hR) =P(h)\u0012dh\ndhR\u0013\n(D4)\n=Nhβ(1−α)−α\nR (D5)\nSince the distribution is cut off at hc,R=hR(hc), the normalization constant must be N=γ/hγ\nc,R, with γ=\nβ(1−α)−α+ 1 = (1 + β)(1−α). Inserting this into the averaged susceptibility gives:\nχ′′x= sgn( ω)ˆ∞\n|ω|/hc,Rfα(x)\nx2γ|ω|γ−1\nhγ\nc,Rxγ−1dx (D6)\n=1\nω\f\f\f\fω\nhc,R\f\f\f\fγ\n×γˆ∞\n|ω|/hc,Rfα(x)\nxγ−1dx (D7)\nUsing that fact that at long times χ(t)∝1/t2, we see that χ′′\nx(ω/hR)∝ω⇒fα(x≪1)∝x2. Near the lower\nintegration limit the integrand is ∝1/xγ−3. Ifγ < 2 then the integral converges when taking ω/hc,R→0, and\ncan thus be considered as a constant. (If γ >2 then the integral diverges and the resulting frequency dependence\nisχ′′x∝ω. This is because the averaged susceptibility cannot decay faster than the susceptibility of the TLSs with\nhighest h.)27\n2.α≈1\nIn this case we use hR=cαωcexp(−bωc/h)⇒h=bωc\n2 log\u0010\ncαωc\nhR\u0011, which gives the renormalized distribution:\nP(hR) =N\nhRlog2+β\u0010\ncαωc\nhR\u0011 (D8)\nand the normalization can be found to be N= (1 + β) log1+β(ωc/hc,R). We neglect for simplicity the factor of\ncα∼ O(1) inside the logarithm. The averaged susceptibility is then given by\nχ′′x= sgn( ω)(1 + β) log1+β(ωc/hc,R)ˆ∞\n|ω|/hc,Rfα(x)\nxlog2+β(xωc/|ω|)dx (D9)\nIn order to simplify the integral, we rely on the fact that fα(x≫1)∝1/x4−2α[48] and f(x≪1)∝x2, such that\nmost of the weight of the integral is around x∼ O(1), for which |log(x)| ≪log(ωc/|ω|). Thus we may neglect the\nxdependence inside the log, giving the form of the susceptibility presented in the main text (using the sum rule\nEq. (C4) for´∞\n0f(x)/xdx = 1/2).\n3. Through the BKT transition ( 1< α < 1 +hc/EF)\nThe behavior around the BKT transition is slightly more convoluted, since when x0→0−the dependence of\nhRon˜his slightly different. However, if we work close enough to the transition, we can just change variables to\ny0(˜h) =√−x0, and use the form (B14) which explicitly depends only on y0. Changing variables we thus find that:\nP(y0) =1 +β\nh1+β\ncy0\u0000\ny2\n0+J2\u0001−1+β\n2(D10)\nwith the cutoff yc=p\n(hc/wc)2−J2, and note that the range y∈(0, yc) covers only the range h∈(|J|, hc), since the\nTLSs with h <|J|are in the localized phase. Thus, if yc≪Jwe can approximate P(y0)∝y0for any β, and we will\ntherefore find that the distribution of P(hR) will be identical to (D8) with β= 1. Thus, while for α= 1 the exponent\nof the log will be 1 + β, it will change smoothly to 2 near the end of the transition. We evaluate this numerically\nfor any value of 1 < α < 1 +hc/ωcusing the form given in (B13), and for the sake of the computation using the\nsimplification χ′′(ω) =δ(|ω| −2hR) (since the results should not depend on the actual function f(x) but rather on\nthe form of the distribution Pr). The results, confirming the analysis presented in this subsection and the previous\none, are shown in Fig. 7.\nFigure 7. Averaged susceptibility of TLSs around the BKT transition, for β= 0,1,2 and varying values of 1 < α < 1 +hc/ωc.\nAs expected, for α= 1 ( J= 0) the susceptibility is ∝1/ωlogβ+2(ωc/ω), while as α→1 +hc/ωc(J→hc/ωc) this changes\nsmoothly into ∝1/ωlog3(ωc/ω). Note that the change in the cutoff of the values in the xaxis with increasing Jis due to the\nlowering of hc,R.28\n4. Localized phase ( α >1 +hc/EF)\nIn the localized phase, where hR= 0, most of the weight of χ′′\nx(ω) lies in a delta function at zero frequency. However,\nthere are still weak residual quantum fluctuations. The form of these fluctuations can be found using a simple scaling\nanalysis: we write the susceptibility at finite frequency as some function χ′′\nx(ω) =1\nωF(h/E F,|ω|/EF). Reducing the\ncutoff to EF→EF/b, the field rescales to h/E F→h/E F/b(1−a). Since the result must be independent of b, we can\nsetb=|ω|/EFand find that\nχ′′\nx=1\nωF \nh\nEF\u0012EF\n|ω|\u00131−α\n,1!\n=1\nωF\u0012h(ω)\n|ω|,1\u0013\n(D11)\nwith h(ω) =h(|ω|/EF)αthe frequency dependant energy scale. Since h(ω)≪ω, we may expand to second order\nusing Fermi’s golden rule, and find\nχ′′\nx(ω)∝1\nω\u0012h(ω)\nω\u00132\n∝sign(ω)h2\nE2α\nF|ω|3−2α(D12)\nThe constant of proportionality may be set using the sum rule Eq. (C6), and then averaging over hwe obtain Eq. (42)\nof the main text.\n5. Biased case\nFor the biased case, the ⟨σx⟩has an equilibrium value, so that ⟨σx(t→ ∞ )σx(0)⟩ → ⟨ σx⟩2. We therefore decompose\nχ′′\nx(ω) =⟨σx⟩2δ(ω) +χ′′\ninel(ω) (D13)\nThe equilibrium value, which contributes to the elastic scattering rate, is given by [51]\n⟨σx⟩=2\nπatan\u0012hx\nhR\u0013\n(D14)\nSince hc/hc,R∝(ωc/hc)α/(1−α)≫1, when averaging over hx, hRwill not have much effect, and thus\n⟨σx⟩2= 1− O\u0012hc,R\nhc\u0013\n(D15)\nThe inelastic contribution χ′′\ninelwill now have a two-parameter scaling form:\nχ′′\ninel(ω) =1\nωfα\u0012ω\nhR,hx\nhR\u0013\n(D16)\nForα <1, the distributions are P(hR) =γ\nhγ\nc,Rhγ−1\nR,P(ϵ) =1+βx\nh1+βxchβxx.Thus:\nχ′′inel(ω) =ˆ\nP(hR, hx)1\nωfα\u0012ω\nhR,hx\nhR\u0013\ndhRdhx\n= (1 + βx)γ1\nhγ\nc,Rhβx+1\ncˆ\nhγ+βx\nRyβxfα\u0012ω\nhR, y\u0013dhR\nωdy\n= (1 + βx)γωγ+βx\nhγ\nc,Rhβx+1\ncˆ∞\nω/hc,Rdx\nxγ+βx+2ˆhcx/ω\n0dyf(x, y)\nSince hc/hc,R≫1, the upper limit of the yintegral is large for any value of x > ω/h c,R. Therefore defining\n˜fα(x) =´∞\n0fα(x, y)dy, we can rewrite the susceptibility in a form similar to earlier:\nχ′′\ninel(ω)≈(1 +βx)γωγ+βx\nhγ\nc,Rh1+βxcAα29\nAα=ˆ∞\n0˜fα(x)\nxγ+βx+2dx\nHere we have assumed that the upper limit of the yintegral and the lower limit of the xintegral can be continued\nto∞safely. In this case, the validity of this assumption is not as clear as it was in the unbiased case. We verify this\nby an explicit calculation at the TP. There, the scaling function is given exactly by [Ulrich Maura]:\nf(x, y) =4\nπ1\nx2+ 4\nxatan ( x+y) +xatan ( x−y) + ln\n\u0010\n1 + (x+y)2\u0011\u0010\n1 + (x−y)2\u0011\n(1 +y2)2\n\n (D17)\nFor large y,f(x, y)∝1\ny2, so the integral over yindeed converges (this should generically be the case since the\nsingle-TLS susceptibility is an analytic symmetric function of ywhich vanishes for y→ ∞ ). In this case the integral\ncan be evaluated exactly, and we find that:\n˜f(x) =4x2\nx2+ 4(D18)\nWe confirm that ˜f(x≪1)∝x2, just as in the 1bSB, and our approximation is justified as long as γ+βx≤1.\nNear the critical point α→1, the splitting distribution takes the form P(hR)∝1\nhR(logωc/hR)2+βz. When integrating\novery, the effective distribution will change to P(hR)∝hβx\nR\n(logωc/hR)2+βz. Therefore the self energy will be of the form:\nΣ′′\ninel(ω)∝ω1+βx\n(logωc/hR)2+βz(D19)\nNote that for the physical case βx=βz= 0 this will result in MFL-like behavior around α≈1.\nAppendix E: RG flow of 2bSB\nWe now discuss the details of the RG flow of the 2-bath SB model. We will mainly consider the region of interest\nwhich is analogous to the α <1 region in the 1bSB, where the effect of hon the flow of the couplings is negligible\nand the renormalization of hRis a power law. Therefore we begin by examining the effect of the two couplings on\neach other. For example, for the xy−model the RG equations will be (as in Eq. (55))\ndαx\ndℓ=dαy\ndℓ=−2αxαy+O(˜h2). (E1)\nWe can simplify these equations by using the constant of flow δα=αx−αy, which is approximately conserved along\nflow. We assume δα > 0 without loss of generality. We can then simply integrate the equations:\n2ℓ=ˆα0\nx\nαx(ℓ)dαx\nαx(αx−δα)=log\u0010\nr\n1−δα/α x\u0011\nδα(E2)\nwhere r=α0\ny/α0\nx, and α0\naare the bare couplings. We thus find that:\nαx(ℓ) =δα\n1−re−2δαℓ(E3)\nαy(ℓ) =rδα\ne2δαℓ−r(E4)\nAssuming that the initial h/ωc≪1 is small enough, the flow will reach δαℓ≪1, at which point the dominant coupling,\nwhich in this case is αx, saturates at the value αx,R=δα, while the subleading coupling continues to decrease:\nαy,R=rδα(Λ′/ωc)δα. Since the flow stops when hR= Λ′then in the low energy theory αy,R=rδα(hR/ωc)δα. Once\nthis point has been reached, we can examine the beta function of ˜h:\nd˜h\ndℓ= (1−αx−αy)˜h (E5)30\nAs mentioned above, after some “time” ℓ∼δα−1(which importantly does not depend on the initial value of h/ωc),\nαxwill saturate, while αybecomes negligible. Thus at this point the flow is identical to the flow of the 1bSB, with\nα=δα. We therefore find that if δα > 1 the tunneling flows to zero and the TLS becomes localized, while for\nδα < 1 the renormalized tunneling assumes the familiar form hR∝ωc(h/ωc)1/(1−δα). Note that in this case the\nproportionality constant will depend on the time it took αxto saturate, which is a quantity which depends on α0\nx, α0\ny\nand not on h, ωc. This can be found exactly by insetring αx,y(ℓ) into (E5) and integrating:\nlogωc\nh=ˆℓ∗\n0\u0012\n1 +δα1 +re−2δαℓ\n1−re−2δαℓ\u0013\ndℓ= (1−δα)ℓ∗−log\u00121−re−2δαℓ∗\n1−r\u0013\n(E6)\n⇒hR∝(1−r)2\n1−δαωc\u0012h\nωc\u0013 1\n1−δα\n(E7)\nThe flow of the couplings in the xz−model is identical. However, if the dominant coupling is αzthen after ℓ≳δα−1\nthe flow of hwill slow down, and thus hwill only by renormalized by a multiplicative factor. We find in this case\n(analogous to only inserting αy(ℓ) into (E5)):\nlogωc\nh=ˆℓ∗\n0\u0012\n1 +δαre−2δαℓ\n1−re−2δαℓ\u0013\ndℓ=ℓ∗−log\u00121−re−2δαℓ∗\n1−r\u0013\n(E8)\n⇒hR=δα\nαzh (E9)\nAppendix F: Subleading corrections in xyz−model\nFollowing the methods of Ref. [48], we characterize the magnitude of the different subleading corrections in the\nmulti-bath case. There are two types of subleading corrections: the susceptibilities of the subdominant baths which\nappear in the self energy, and perturbative corrections to the susceptiblity of the dominant bath due to the weak\ncoupling to the losing baths. We will study these in the xy−model and in the xz−model, and the generalization\nto the xyz−model is straightforward since the couplings to the subleading baths are perturbative in the low energy\ntheory.\n1.xy−model\nAs usual we will assume without loss of generality that αx> α y. As mentioned above, there are two types of\ncorrections to the self energy. We start with that due to perturbative corrections to χx. As presented in [48], if only\nthexbath was present after integrating out the high energy modes, we could expand the ground and excited states\nas (performing perturbation theory in the low-energy modes):\n|g⟩0=\f\f\f˜↓E\n+ϕx\nhR\f\f\f˜↑E\n+1\n2\u0012ϕx\nhR\u00132\f\f\f˜↓E\n+··· (F1)\n|ωi, x⟩0=b†\nx,i|g⟩+··· (F2)\nwhere ϕα=P\ni√\nΠa(ωi)\nhR(b†\na,i+ba,i), Πa(ω)∝αr\naandb†\na,i, ba,iare respectively the bath operator, bath spectral function,\nand boson creation and annihilation operators of the abath. Importantly, the states\f\f\f˜↑,˜↓E\n= 1/√\n2\u0000\f\f˜+\u000b\n±\f\f˜−\u000b\u0001\nare\nsuperpositions of the high-frequency-moded dressed xstates\f\f˜±\u000b\n. This gives the expected χx(ω≪hR)∝αx,Rω/h2\nRat\nlow frequencies, but for general frequencies should be treated in a non-perturbative manner in αx,R. However, since\nαy,Ris small, we can add it perturbatively only to first order:\n|g⟩ ≈ | g⟩0−iϕy\nhR \n1 +\u0012ϕx\nhR\u00132\n+···!\f\f\f˜↑E\n(F3)\nand the relevant excited states will involve insertions of one yboson with multiple xbosons. Using the spectral\ndecomposition for χx:\nχx(ω) =X\nn|⟨n|σx|g⟩|2δ(En−ω) (F4)31\nwe find that the leading correction will come from matrix elements of the form ⟨ω1,···ω2k, x;ωj, y|σx|g⟩. While\nthe summation over the many orders of ϕxis nontrivial, we know that it must produce a scaling function that only\ndepends on ω/hR, and we may thus write:\nχ′′\nx(ω, αy,R) = χ′′\nx(ω,0) +αy,R1\nω˜fαx,R\u0012ω\nhR\u0013\n(F5)\n=1\nωfαx,R\u0012ω\nhR\u0013\n+αy\u0012hR\nωc\u0013αx,R1\nω˜fαx,R\u0012ω\nhR\u0013\n(F6)\nwhere in the second line we inserted the expression for αy. While averaging over the first term will give the usual\ncontribution, in the second term we can treat the distribution as effectively having an increased exponent ˜Pr∼\nhγ−1+αx,R\nR , which will in turn produce a term with a subleading frequency dependence in the averaged susceptibility\n∝ωγ−1+αx,R.\nWe now consider the susceptibility χy, whose spectral decomposition is\nχy(ω) =X\nn|⟨n|σy|g⟩|2δ(En−ω). (F7)\nUsing the fact that the bare σyflips the TLS without properly adjusting the high-energy bosons, we have that\nD\n˜↑|σy|˜↓E\n∝hR\nh∝\u0012hR\nωc\u0013αx,R\n. (F8)\nFor small frequencies we may use the perturbative form (F2) and find that χ′′\ny∝\u0010\nhR\nωc\u00112αx,R\nχ′′\nx, which will naively\ntranslate into a frequency dependence ωγ−1+2αx,Rin the averaged susceptibility. However, the averaged susceptibility\ndepends on the full χ′′\ny, and since for intermediate and high frequencies this perturbation theory is not applicable, we\ncannot fully determine the nonuniversal prefactor, and can only argue that χ′′y∝ωγ−1+ϵwith ϵ >0.\n2.xz−model\nWe begin by studying the similar case where αx> αz. The susceptibility χ′′\nxwill now acquire similar corrections\ndue to αz. However, the matrix elements with single insertions of σzϕzvanish, and we must instead go to second\norder in σzϕz. This means that the corresponding correction to the averaged susceptibility will be ∝ωγ−1+2αx,R.\nThe susceptibility χ′′\nzwill be nonuniversal due to considerations identical to (F8), and will thus be suppressed by\na prefactor\u0010\nhR\nωc\u00112αx,R\nat low frequencies, although we do not know the generalization of it to higher frequencies.\nHowever, in addition this susceptibility includes a static delta function peak due to the equilibrium value of ⟨σz⟩∞.\nFor small αz≪1 this can be calculated using the sum rule Eq. (C5):\n⟨σz⟩=1\nhˆωc\n0f(ω/hR)dω=\u0012hR\nωc\u0013αx\n× ˆωc/hR\n0f(x)dx!\n(F9)\nWe must therefore find if this integral converges or diverges when the upper limit is taken to ωc/hR→ ∞ . Ref. [48]\nshows that f(x≫1)∝1/x2−2α, so that for α <1/2 this integral converges to some constant = Aαwhile for α >1/2\nthis integral diverges as a power law = Aα(ωc/hR)2α−1(forα= 1/2 it diverges logarithmicaly ∝log(ωc/hR)). Thus\nwe can write\nχ′′\nz≈δ(ω)×Aαx\u0012hc,R\nωc\u0013min(αx,1−αx)\n+··· (F10)\nwith the dots referring to the subleading frequency-dependent terms.\nWe now turn to the case where αz> αx. Here, with no αxthe ground state is exactly a coherent state of all the\nbosons centered around the location corresponding to |↓⟩. Acting with the high-frequency mode dressed ˜ σxwill only\nagitate the low energy modes, and thus we can writeD\n˜↑|˜σx|˜↓E\n=sx∼ O(1). Therefore incorporating the effects of\nαx,Rperturbatively will modify the ground state as:\n|g⟩= \n1 +1\n2s2\nx\u0012ϕx\nhR\u00132!\f\f\f˜↑E\n+sxϕx\nhR\f\f\f˜↓E\n+··· (F11)32\nIn terms of the resulting modification to χ′′\nz, we can easily see that the elastic peak will decrease by a small amount\nproportional to αx,R, and the inelastic part will be modified by a term proportional to ( αx,R)2ω/h2\nR(assuming that\nω/hR≪1/αx,R), which in turn will give a correction to χ′′zproportional to ωγ−1+2αz,R. The susceptilibty χ′′\nxis\nsimply a delta function time the factor corresponding to (F8):\nχ′′\nx(ω) =\u0012hR\nωc\u00132αz,R\nδ(ω±2hR) (F12)\nχ′′x(ω)∝ωγ−1+2αz\nx (F13)\nAnd we can thus conclude that the frequency dependence of the self energy in this case will be\nΣ′′(ω)−Σ′′(0)∝|ω|γ+2αz,R\nhγ\nc,Rω2αz,Rc(F14)\nNote that when averaging over hRwe neglect the contribution from TLS whose splitting obeys ω/hR≫(ωc/hR)αz,R→\nhR≪ω(ω/ω c)αz,R/(1−αz,R), for which the perturbation theory breaks down.\nAppendix G: Derivation of specific heat from internal energy\nHere we derive Eq. (90) that enables us to obtain the specific heat of the model following Ref. [92]. Considering\nthe Hamiltonian (1):\nH=X\nαεαc†\nαcα+X\nαβγ,lgl\nαβγc†\nαcβσl\nγ+X\nγhγσz\nγ. (G1)\nThe single-particle quantum numbers stand for combinations of momenta and flavor indices and we also supress factors\nofMandNfor brevity. It is useful to introduce an arbitrary retarded fermionic Green’s function\n⟨⟨A, B⟩⟩r\nt≡ −iθ(t)\n[A(t), B]+\u000b\n(G2)\nand its Fourier transform ⟨⟨A, B⟩⟩r\nω. Here [ A, B]+is the anti-commutator. The advanced Green’s function is given\nby⟨⟨A, B⟩⟩a\nω= (⟨⟨A, B⟩⟩r\nω)∗where ( ·)∗denotes complex conjugation. We use the fact that the retarded and advanced\nGreen’s functions both obey the equation of motion:\nω⟨⟨A, B⟩⟩ω=\n[A, B]+\u000b\nω+\n\n[A, H]−, B\u000b\u000b\nω. (G3)\nWe use Eq. (G3) to obtain the equation for motion of the retarded/advanced fermionic Green’s function:\n(ω−εα)\n\ncα, c†\nα\u000b\u000b\nω= 1 +X\nαβγ,lgl\nαβγ\n\ncβσl\nγ, c†\nα\u000b\u000b\nω. (G4)\nWe proceed to consider the internal energy U≡ ⟨H⟩:\nU=X\nαεα\nc†\nαcα\u000b\n+X\nαβγ,lgl\nαβγ\nc†\nαcβσl\nγ\u000b\n+X\nγhγ\nσz\nγ\u000b\n(G5)\nUsing the identity (that follows from the spectral representation)\n⟨AB⟩=iˆ∞\n−∞dω\n2π⟨⟨A, B⟩⟩r\nω− ⟨⟨A, B⟩⟩a\nω\neβω+ 1, (G6)\nand Eq. (G4) we may express the first two terms in Uas\nX\nαεα\nc†\nαcα\u000b\n+X\nαβγ,lgl\nαβγ\nc†\nαcβσl\nγ\u000b\n=−ˆdω\nπX\nαIm\n\ncα, c†\nα\u000b\u000br\nω. (G7)\nThe right-hand-side can be written in terms of the electronic spectral function:´\nkAk(ω) =−1\nπP\nαIm\n\ncα, c†\nα\u000b\u000br\nω.\nInserting this form to the expression for the internal energy, we obtain the form given in Eq. (90).33\nAppendix H: Superconductivity\n1. Effective action\nWe generalize our model to spin-1 /2 fermions by altering the interaction term in Eq. (1) to\nHint=1\nNX\nr,s,ijlgijl,r·σl,rc†\nirscjrs. (H1)\nHere, c†\nikαis the fermionic creation operator for momentum k,spins=↑,↓, and flavor index i= 1,···, N/2 (so that\nthe total number of electron flavors remains N), and the rest of the definitions are identical to the case in the main\ntext.\nAfter averaging over the coupling constants via replica trick, introducing the bilocal fields (including the new pairing\nfieldF)\nGr,r′s(τ, τ′) =1\nNX\ni¯cirα(τ)cir′s(τ′),\nFr,r′(τ, τ′) =1\nNX\nicir↓(τ)cir′↑(τ′),\nχa,r(τ, τ′) =1\nMX\nlσa\nl,r(τ)σa\nl,r(τ′), (H2)\nand integrating over the fermions, we obtain the effective action:\nS=−Ntr log\u0010\nˆG−1\n0−ˆΣ\u0011\n−NˆX\nr,sGr,r′s(τ, τ′) Σr′,rs(τ′, τ)\n−NˆX\nr,σ\u0010\nF†\nr,r′(τ, τ′) Φr′,r(τ′, τ) +Fr,r′(τ, τ′) Φ†\nr′,r(τ′, τ)\u0011\n−MX\naX\nrg2\naˆ\nχa,r(τ, τ′)\n×\"X\nsGr,rs(τ, τ′)Gr,rs(τ′, τ)−(−1)δa,y2F†\nr,r(τ, τ′)Fr,r(τ′, τ)#\n+MˆX\nrsχa,r(τ, τ′) Πa,r(τ′, τ) +X\nrMX\nl=1STLS[σl,r], (H3)\nwhere\nSTLS[σ] = SBerry [σ]−ˆ\ndτhl,r·σ(τ)\n−ˆ\ndτdτ′X\naΠa,r(τ′−τ)σa(τ)σa(τ′) (H4)\nis the action of a spin-boson problem with multiple baths.\nIn the first term we use a 2 ×2 Nambu-Gor’kov formulation, sufficient for singlet pairing\nˆΣrr′(τ, τ′) =\u0012Σrr′↑(τ, τ′) Φrr′(τ, τ′)\nΦ†\nrr′(τ, τ′)−Σr′r↓(τ′, τ)\u0013\n. (H5)\nGeneralizations to triplet pairing are straightforward but can only play a role for odd-frequency pairing. We use a\nsimilar expression for the propagator\nˆGrr′(τ, τ′) =\u0012Grr′↑(τ, τ′)Frr′(τ, τ′)\nF†\nrr′(τ, τ′)−Gr′r↓(τ′, τ)\u0013\n. (H6)34\nThe bare propagator in frequency and momentum space is\nˆG0k(iω)−1=\u0012\niω−εk 0\n0 iω+εk\u0013\n. (H7)\nIn the limit of large MandN, with fixed ratio M/N , we can analyze the saddle point limit. We consider a saddle\npoint that does not break time-reversal symmetry Grr′↑(τ, τ′) =Grr′↓(τ, τ′) and drop the spin index. Performing\nthe variation with respect to ˆΣ gives\nˆGr,r′(iω) =\u0010\nˆG−1\n0(iω)−ˆΣ (iω)\u0011−1\f\f\f\f\nr,r′. (H8)\nThe variation w.r.t. GandFyield\nΣr,r′(τ) = δr,r′M\nNX\nag2\naGr,r(τ)χa,r(τ) (H9)\nΦr,r′(τ) =−δr,r′M\nNX\na(−1)δa,yg2\naFr,r(τ)χa,r(τ) (H10)\nThese two equations resemble the ones that occur for electrons that couple to bosonic modes with propagator χa,r(τ)\nvia a Yukawa coupling. The stationary point that follows from the variation with respect to χis\nΠa,r(τ) =−2g2\na\u0002\nGr,r(τ)Gr,r(−τ)−(−1)δa,yF†\nr,r(τ)Fr,r(−τ)\u0003\n, (H11)\nan expression that is also analogous to the self energy of a bosonic problem.\nThe TLS-correlation functionD\nσa\nl,r(τ)σa\nl,r(τ′)E\nis determined from the solution of the spin boson problem.\n2. Linearized Eliashberg equations\nAs long as we are only interested in the onset of pairing and the superconducting phase transition is of second order\nwe can focus on the linearized gap equation. In this case we can neglect the feedback of superconductivity on the\nohmic bath. The solution of the spin-boson problem then yields the local propagator χa(ω). The equation for the\nmomentum-independent normal self energy is:\nΣ (iω) =M\nNTX\nω′,ag2\naG(iω′)χa(iω−iω′)\n=M\nNTX\nω′,ag2\naρFˆ\ndεk1\niω′−εk−Σ (ω′)χa(iω−iω′). (H12)\nThe linearized equation for the s-wave anomalous self energy is (assuming particle-hole symmetry for simplicity, we\nuse the fact that iΣ (iω′) is real):\nΦ (iω) =−M\nNTX\nω′,ag2\na(−1)δa,yF(iω′)χa(iω−iω′)\n=M\nNTX\nω′,aˆ\ndεkρFg2\na(−1)δa,yΦ (iω′)χa(iω−iω′)\n(iω′−εk−Σ (iω′)) (−iω′−εk+ Σ ( iω′))\n=M\nNTX\nω′,aˆ\ndεkρFg2\na(−1)δa,yΦ (iω′)χa(iω−iω′)\n(iω′+iΣ (iω′))2+ε2\nk. (H13)\nIf we perform the integration over εk, we get for both self energies the Eliashberg equations.\nΣ (iω) =−iTX\nω′sign ( iω′)DΣ(iω−iω′),35\nΦ (iω) = TX\nω′Φ (iω′)\n|ω′+iΣ (iω′)|DΦ(iω−iω′), (H14)\nwith\nDΣ(iω) =M\nNX\naρFg2\naχa(iω),\nDΦ(iω) =M\nNX\na(−1)δa,yρFg2\naχa(iω). (H15)\nThe contribution due to the coupling gyis pair breaking and sufficiently large gycan partially or fully destroy\nsuperconductivity.\nFor the solution of the linearized gap equation we introduce\n∆ (iωn) =ωnΦ (iωn)\nωn+iΣ(iωn), (H16)\nwhich yields a closed equation\n∆ (iωn) =TX\nn′sign(iωn′)\u0012∆ (iωn′)\nωn′DΣ(iωn−iωn′)−∆ (iωn)\nωnDΦ(iωn−iωn′))\u0013\n. (H17)\nOne nicely finds that if DΣ(iω) =DΦ(iω)≡D(iω) ( i.e. gy= 0) the zeroth bosonic Matsubara frequency does\nnot contribute to the solution of the coupled equation. Static fluctuations are irrelevant for the pairing problem, in\nagreement with Anderson’s theorem. From now on we will assume that gy= 0.\nHence, in what follows, we can just skip n′=nin the sum. Then we do not have any problem with a potentially\ndivergent D(0) for γ≤1.\n∆ (iωn) =TX\nn′̸=n\u0012\n∆ (iωn′)−ωn′\nωn∆ (iωn)\u0013D(iωn−iωn′)\n|ωn′|. (H18)\nFor even frequency pairing we have ∆ ( iωn) = ∆ ( −iωn). Hence we can write (we only consider ωn>0)\n∆ (iωn) = TX\nn′≥0∆ (iωn′)(1−δn,n′)D(iωn−ωn′) +D(iωn+iωn′)\nωn′\n−∆ (iωn)TX\nn′≥0(1−δn,n′)D(iωn−iωn′)−D(iωn+iωn′)\nωn(H19)\nNext we introduce\nΨn=∆ (iωn)\n|ωn|1/2(H20)\nand obtain\nΨn=TX\nn′≥0Ψn′(1−δn,n′)D(iωn−iωn′) +D(iωn+iωn′)√ωnωn′\n−ΨnTX\nn′≥0(1−δn,n′)D(iωn−iωn′)−D(iωn+iωn′)\nωn(H21)\nWe can write this as a matrix equation, where Nmaxis the maximum number of Matsubara frequencies included:\nΨn=∞X\nn′=0Kn,n′Ψ′\nn (H22)\nKn,n′=δn,n′Kdiag\nn+ (1−δn,n′)D(iωn−ωn′) +D(iωn+ωn′)\nπp\n(2n+ 1)(2 n′+ 1)(H23)36\nKdiag\nn =∞X\nm=0(1−δn,m)D(iωn−ωm) + (1 + δn,m)D(iωn+ωm)\nπ(2n+ 1)(H24)\nGiven this matrix equation, we first consider the behavior of Tcfor a power law form of the pairing propagator (with\nsome high energy cutoff Λ, so that the Matsubara sum runs up to Nmax= Λ/T)\nD(iωn) =A\f\f\f\fΛ\nωn\f\f\f\fa\n≡A\u0012Λ\nT\u0013a\ndn (H25)\nwhere we introduced the rescaled propagator dn=1\n(2πn)a. Since Kn,n′is linear in D, we may also define its rescaled\nversion Kn,n′=A\u0000Λ\nT\u0001akn,n′where kn,n′is defined by replacing D(iωn) with dnin the definition of Kn,n′. Defining\nκ(a, N max) as the largest eigenvalue of kn,n′, the Eliashberg equation simplifies to:\n1 =A\u0012Λ\nTc\u0013a\nκ(a, N max) (H26)\nThe qualitative behavior of the eigenvalue can be estimated by inspecting the diagonal elements of k, namely, a\nsimple power counting indicates whether the sum converges or not, which determines the depedence on the cutoff\nNmax. In particular, for a= 0, κ≈κ0log(Nmax) and we find the BCS solution Tc∝Λ exp( −1/Aκ 0). For a >0\nthe series converges for n, n′→ ∞ andκ=κadepends only on the exponent a, giving the critical temperature Tc=\nΛ(Aκa)1/a. A numerical calculation of κais shown in Fig. 8. Finally, consider the case where D(iωn) =Alog(Λ /|ωn|)\n(corresponding to a MFL behavior of the electrons). Using dn= log( Nmax/|n|) one can show numerically that\nκ≈κ′\n0log2(Nmax) such that Tc∝Λ exp( −1/√Aκ0). An analytic derivation of this result using the Eliashberg\nequation is given in [93].\n3. Analytic continuation of the bosonic propagator\nWe start with the imaginary part of the bosonic propagator, assuming a power-law form:\nD′′(ω) = sign( ω)A\f\f\f\fΛ\nω\f\f\f\fa\n(H27)\nAnalytically continuing to Matsubara frequencies (for a <2), we obtain\nD(iΩ) = −1\nπˆ∞\n−∞dω\niΩ−ωD′′(ω) (H28)\n=2\nπˆ∞\n0ωdω\nΩ2+ω2D′′(ω) (H29)\n=2AΛa\nπˆΛ\n0ω1−adω\nΩ2+ω2(H30)\n=2A\n(2−a)πΛ2\n|Ω|22F1\u0012\n1,−a\n2,1−a\n2,−Λ2\nΩ2\u0013\n(H31)\nwhere 2F1is the Gaussian hypergeometric function. In order to obtain anayltical results, we approximate:\nD(iΩ)≈2AΛa\nπ ˆΩ\n0ω1−a\nΩ2dω+ˆΛ\nΩω−1−adω!\n(H32)\n=2A\nπ\u00121\n2−a\f\f\f\fΛ\nΩ\f\f\f\fa\n−1\na\u0012\n1−\f\f\f\fΛ\nΩ\f\f\f\fa\u0013\u0013\n(H33)\nwhich coincides with the appropriate limits of the function 2F1. Note that the leading behavior for |Ω| ≪Λ is of the\nform\nD(iΩ)≈2A\nπ×\n\n1\n|a|a <0\nlog\u0010\nΛ\n|Ω|\u0011\na= 0\n2\na(2−a)\f\fΛ\nΩ\f\fa0< a < 2(H34)\nWe now use the above to obtain Tcin the x-model.37\nFigure 8. Numerical calculation of the largest eigenvalue κ(a, N max) as a function of a(top) and for the special cases a= 0\n(bottom left) and a= 0′(i.e. logarithmic DΦ) (bottom right). For any finite value of athis approaches a constant as Nmax→ ∞ ,\nwhile for a= 0,0′it scales as log( Nmax),log2(Nmax) respectively at large Nmax.\n4. Detailed analysis of the x−model\nWe analyze the scaling of Tcfor two parameter regimes: “weak coupling”, for which Tc≪hc,R, and “strong\ncoupling” for which Tc≫hc,R. We will find approximate solutions by analytically continuing the TLS susceptibility\nχxand identifying the most singular contribution to DΦ(iΩ), from which we obtain Tcusing Eq. (H26).\na. Weak coupling\nHere we study the susceptibility for T, ω≪hc,R. From dimensional considerations, we write the averaged suscep-\ntibility as a scaling function:\nχ′′\nx(ω, T) =Aαγ\nω\f\f\f\fω\nhc,R\f\f\f\fγ\nmin\u0012\n1,\u0010ω\nbT\u0011δ\u0013\n(H35)\nwith some δ >0 and b∼ O(1). Analyzing the susceptibility at weak coupling and at the Toulouse point suggests that\nδ= 1, although as we will see the exact value of δdoes not qualitatively change Tc. The corresponding Matsubara38\nfrequency correlator is given by\nχx(iΩn)≈2γAα\nπhc,R\u0012\n1\nγ−1−2\nγ2−1\f\f\fΩ\nhc,R\f\f\fγ−1\n−δ\n(γ+1)(γ+1+δ)Tγ\nΩ2hγ−1\nc,R\u0013\n(H36)\nThe most singular contribution to DΦis therefore (the temperature dependent term is not significant and can be\nignored, since Ω ≳T)\nDΦ(iΩ)≈2πAαλ×\n\nγ\nγ−1γ >1\nlog\u0010\nhc,R\n|Ω|\u0011\nγ= 1\n2γ\n1−γ2\f\f\fhc,R\nΩ\f\f\f1−γ\nγ <1(H37)\nUsing Eq.(H26) we find that\nTc/hc,R∝\n\nexp\u0010\n−γ−1\n2πAαγκ0λ\u0011\nγ >1 +O(√\nλ)\nexp\u0012\n−1√\n2πAακ′\n0λ\u0013\nγ= 1\n\u0010\n4πAαγκγ−1\n1−γ2λ\u00111\n1−γγ <1− O(√\nλ)(H38)\nwhere the requirement |γ−1|>O(√\nλ) in the BCS and quantum critical regimes is necessary for self consistency.\nAdditionally, demanding that Tc≪hc,R, which is assumed in taking the low- Tform of the TLS-susceptibility, requires\nλ≪1, i.e.\n1≪αM\nNEF\nhc,R∝αM\nN\u0012EF\nhc\u00131\n1−α\n. (H39)\nThis condition will always break down at some α <1. For M/N ∼ O(1) this will happen at very small values of α:\nα∝hc/EF, while in the limit where TLSs are extremely sparse:M\nN≪hc\nEF, this occurs at α≈1−log\u0010EF\nhc\u0011\nlog(N\nM).\nb. Strong coupling\nWe now turn to the regime T≫hc,R. Note that for α >1 this is always the case since hc,R= 0. For frequencies\nω≪Tthe TLS correlation function decays exponentially with rate [94],[35] (Eq 5.29),\nΓ =ch2\nT\u0012T\nEF\u00132α\n(H40)\nwith csome αdependent prefactor. Note that T≫Γ for T≫hR. In this regime, the TLS-susceptibility can be\napproximated as\nχ′′\nx(ω) =1\n2πΓ\nTω\nΓ2+ω2(H41)\n(The prefactor Γ /Tis due to the sum rule Eq. (C4)). For ω≫T, hRthe analysis of App. D 4 can be extended for\nα̸= 1 (or more precisely |α−1|> hc/EF). Overall for ω≫hc,Rone finds that\nχ′′\nx(ω, T) = 4 αh2\nE2α\nF(max ( ω, bT ))2−2α\nω(H42)\nwith b∼ O(1). Analytically continuing and seperating the different frequency regimes, we define\nχx(iΩ) =2\nπ ˆΓ\n0+ˆbT\nΓ+ˆΩ\nbT+ˆEF\nΩ!\ndωωχ′′\nx(ω)\nω2+ Ω2≡χ1+χ2+χ3+χ4. (H43)39\nThus, for Ω > aT≫Γ\nχ1≈1\n3π2Γ2\nTΩ2(H44)\nχ2≈1\nπ2Γ\nΩ2\u0012\n1−Γ\nbT\u0013\n(H45)\nχ3≈2α\nπh2\nE2α\nF|Ω|2α−3\n2α−1 \n1−\u0012bT\n|Ω|\u00132α−1!\n(H46)\nχ4≈2α\nπh2\nE3\nF1\n2α−3 \n1−\u0012|Ω|\nEF\u00132α−3!\n(H47)\nThe most singular contribution to DΦis given by\nDΦ(iΩ) = 2 πϵα2bβ×\n\n2α\n1−2αE3−2α\nF\nT1−2αΩ2 α <1/2\nE2\nF\nΩ2log\u0010\n|Ω|\nT\u0011\nα= 1/2\n2\n(2α−1)(3−2α)\f\fEF\nΩ\f\f3−2α1/2< α < 3/2\nlog\u0010\nEF\n|Ω|\u0011\nα= 3/2\n1\n2α−3α >3/2(H48)\nwhere bβ=\u0010\n1+β\n3+β\u0011\ncomes from averaging over h2. Inserting these into Eq. (H26), we obtain\nTc/EF∝\n\n\u0010\n4πα2κ3−2αbβ\n(2α−1)(3−2α)ϵ\u00111\n3−2αα <3/2− O(√ϵ)\nexp\u0012\n−1\nα√\n2πbβκ′\n0ϵ\u0013\nα= 3/2\nexp\u0010\n−2α−3\n2πα2βκ0ϵ\u0011\nα >3/2 +O(√ϵ)(H49)\nNote that for α≤1/2 the prefactor in the parentheses changes, according to the corresponding expression in Eq. (H48).\nHowever, the dependence on the small parameter ϵremains ϵ1/(3−2α)for all α <3/2.\nOnce again, for α <1 consistency requires that Tc≫hc,R, which translates into\n1≪αM\nN\u0012EF\nhc\u00131\n1−α\n⇐⇒ λ≫1 (H50)\nwhich is complementary to the requirement for weak coupling." }, { "title": "1908.04099v2.Low_energy_type_II_Dirac_fermions_and_spin_polarized_topological_surface_states_in_transition_metal_dichalcogenide_NiTe__2_.pdf", "content": "Low-energy type-II Dirac fermions and spin-polarized topological surface states in\ntransition-metal dichalcogenide NiTe 2\nBarun Ghosh,1Debashis Mondal,2Chia-Nung Kuo,3Chin Shan Lue,3Jayita\nNayak,1Jun Fujii,2Ivana Vobornik,2,\u0003Antonio Politano,4,yand Amit Agarwal1,z\n1Department of Physics, Indian Institute of Technology - Kanpur, Kanpur 208016, India\n2Istituto O\u000ecina dei Materiali (IOM)-CNR, Laboratorio TASC,\nin Area Science Park, S.S.14, Km 163.5, I-34149 Trieste, Italy.\n3Department of Physics, National Cheng Kung University, 1 Ta-Hsueh Road 70101 Tainan, Taiwan\n4Dipartimento di Scienze Fisiche e Chimiche (DSFC),\nUniversit\u0012 a dell'Aquila, Via Vetoio 10, I-67100 L'Aquila, Italy\nUsing spin- and angle- resolved photoemission spectroscopy (spin-ARPES) together with ab initio\ncalculations, we demonstrate the existence of a type-II Dirac semimetal state in NiTe 2. We show\nthat, unlike PtTe 2, PtSe 2, and PdTe 2, the Dirac node in NiTe 2is located in close vicinity of the\nFermi energy. Additionally, NiTe 2also hosts a pair of band inversions below the Fermi level along\nthe \u0000\u0000Ahigh-symmetry direction, with one of them leading to a Dirac cone in the surface states.\nThe bulk Dirac nodes and the ladder of band inversions in NiTe 2support unique topological surface\nstates with chiral spin texture over a wide range of energies. Our work paves the way for the\nexploitation of the low-energy type-II Dirac fermions in NiTe 2in the \felds of spintronics, THz\nplasmonics and ultrafast optoelectronics.\nThe discovery of topological semimetals has ushered\nin a new era of exploration of massless relativistic quasi-\nparticles in crystalline solids[1{4]. These arise as emer-\ngent quasi-particles in crystals with linearly dispersing\nbands in vicinity of a degenerate band crossing point (ei-\nther accidental or symmetry-enforced) and are protected\nby crystalline symmetries[5]. Double, triple and quadru-\nple degeneracy of the band crossing leads to topologi-\ncally protected Weyl[6{14], triple point[15{21] and Dirac\nfermions[22{30], respectively. In contrast to their high\nenergy counter-parts, these emergent quasi-particles are\nnot protected by Lorentz symmetry, and can also occur\nin a tilted form, giving rise to type-I and type-II Dirac\nfermions. Speci\fcally, Na 3Bi[26, 29] and Cd 3As2[25, 30]\nare type-I Dirac semimetal (DSM), while the transition-\nmetal dichalcogenides (TMDs) PtTe 2[31{33], PtSe 2[34],\nand PdTe 2[32, 35, 36] are type-II DSM.\nIn group X Pd- and Pt- based dichalcogenides, the\nbulk Dirac node lies deep below the Fermi level ( \u00180.6,\n\u00180.8 and\u00181.2 eV in PdTe 2, PtTe 2, and PtSe 2, respec-\ntively) [31{36], hindering their successful exploitation in\ntechnology. In contrast, NiTe 2has been predicted to host\ntype-II Dirac fermions in vicinity of the Fermi energy [37].\nThe so far performed experimental studies on NiTe 2have\nprimarily focused on its crystal structure, and transport\nproperties while its topological band structure remains\nunexplored [37{44]. Motivated by this, we explored the\nelectronic band structure of NiTe 2by means of spin- and\nangle-resolved photoemission spectroscopy (ARPES) in\ncombination with density functional theory (DFT).\nOur spin-resolved ARPES measurements explicitly\ndemonstrate the existence of a pair of type-II Dirac nodes\nin NiTe 2along the C3rotation axis, lying just above\n(within 20 meV) the Fermi energy. Additionally, we\nshow that NiTe 2also hosts a series of inverted band-gaps\nFIG. 1. (a) The side view and (b) hexagonal crystal structure\nof NiTe 2with theC3rotation axis. Layers of Ni are stu\u000bed\nbetween two Te layers. (c) The LEED pattern of (0001)-\noriented NiTe 2single crystals, acquired at a primary electron\nbeam energy of 84 eV, clearly indicates its purity and the six-\nfold symmetry along along the (001) direction. (d) The bulk\nand the (001) surface Briluoin zone (BZ) of NiTe 2.\n(IBG). Especially, one of the IBG below the Fermi level\nsupports a Dirac cone in the surface states. Together, the\nbulk Dirac node and the pair of IBG in NiTe 2give rise\nto topological spin-polarized surface states over a wide\nrange of energies. This non-trivial band morphology in\nNiTe 2originates primarily from the 5 p-orbital manifold\nof the Te atoms modi\fed by the intra-layer hybridization,\ntrigonal crystal \feld splitting and spin-orbit coupling.arXiv:1908.04099v2 [cond-mat.mes-hall] 13 Aug 20192\nFIG. 2. (a) Band structure of NiTe 2(including SOC) clearly showing the tilted type-II Dirac node along the \u0000 \u0000Adirection\n(at D). The irreducible representation of the bands (close to Fermi energy) at the \u0000 and Apoints are also marked. (b) The\nFermi surface of NiTe 2originating from the crossing Dirac bands. The type-II Dirac points appear at the touching points of the\nelectron and hole pockets. (c) kzdispersion along \u0000 \u0000Adeduced from the h\u0017-dependent data measured along \u0016\u0000\u0000\u0016Kdirection\n(shown in part in (d-h)); red dashed lines represent the DFT calculations; (d)-(h) The measured band dispersion along the\n\u0016K\u0000\u0016\u0000\u0000\u0016Kdirection for di\u000berent kzvalues, with the red dashed lines indicating the bulk DFT band structure. The blue arrows\nin panel (f) mark the surface states. Data were taken at photon energies of 17, 19, 21, 23 and 25 eV, respectively. Panel (f)\ncorresponding to kz= 0:34c\u0003, is closest to the location of the Dirac point ( kz= 0:35c\u0003in our DFT calculations). Note that,\nfor matching with the experimental data, we have shifted the DFT band structure downward by 100 meV.\nBulk NiTe 2crystallizes in the CdI 2type trigonal struc-\nture (space group P\u00163m1, number 164). It has a layered\nstructure with individual monolayer stacked together via\nweak van der Walls force. As shown in Fig. 1(a)-(b),\neach monolayer has three sub-layers, with the central Ni\nlayer being sandwiched between two adjacent Te layers\n(Ni-Te bond length 2.60 \u0017A). The observation of sharp\nspots in the low-energy di\u000braction pattern (LEED) in\nFig. 1(c) con\frms the high quality of the NiTe 2crystals\ncleaved along the (001) direction, along with the pres-\nence of six-fold symmetry. Surface cleanliness of the as-\ncleaved samples was checked by high-resolution electron\nenergy loss spectroscopy and X-ray photoelectron spec-\ntroscopy. The details of crystal preparation and charac-\nterization, ARPES measurements and DFT calculations\nare presented in Sec. S1, S2 and S3, of the Supplemen-\ntary material (SM) [45].\nThe electronic band structure of NiTe 2including spin-\norbit coupling is shown in Fig. 2(a). It clearly depicts\nthe presence of a pair of tilted band crossings along the\u0000\u0000Adirection. The presence of inversion and time-\nreversal symmetry mandates these bands to be doubly\ndegenerate. Furthermore, the \u0000 \u0000Ahigh-symmetry di-\nrection is the invariant subspace of the three fold rotation\n(C3) symmetry, and a symmetry analysis reveals that the\ncrossing bands have opposite rotation character. This\nprevents their hybridization, resulting in a pair of gapless\nquadruply degenerate type-II Dirac points. The type-II\nnature of the DSM phase is also con\frmed by the fact\nthat the Dirac point appears at the touching point of the\nelectron and hole pocket, as highlighted in Fig. 2(b).\nOur photon-energy dependent ARPES data, in part\npresented in Fig. 2(d-h), results in the kzdispersion pre-\nsented in Fig. 2(c). Our experimental results are con-\nsistent with the DFT-based bulk band structure calcu-\nlations. Extrapolating the \ftted DFT band structure,\nwe \fnd that the Dirac cone is located just above ( \u001820\nmeV) the Fermi energy. in contrast to other TMD-\nbased type-II DSMs like PtTe 2[31{33], PtSe 2[34], and\nPdTe 2[32, 35, 36]. Our attempt to electronically dope3\nFIG. 3. (a) The evolution of the Te 5p orbitals in the formation of Dirac-cone states in NiTe 2. Step (I) shows the the creation of\nbonding and anti-bonding orbitals. Step (II) shows the e\u000bect of the strong trigonal crystal \feld which separates the pzorbitals\nfrom thepx;yorbitals. In step (III), we show the splitting of these states into the jJ;jmJjistates in the presence of SOC. In step\n(IV), we demonstrate the e\u000bect of out of plane dispersion and the formation of the Dirac point. (b) The orbital-resolved band\nstructure and various band inversions along the \u0000 Ahigh-symmetry direction is shown along with the irreducible representations\nof the bands. In panels (c) and (d) we show the ARPES data (for ~\u0017= 24 eV) along \u0016K\u0000\u0016\u0000\u0000\u0016Kand \u0016M\u0000\u0016\u0000\u0000\u0016M(for~\u0017= 30\neV) directions of the (001) surface BZ, respectively. ARPES results are consistent with the DFT predictions in panels (e) and\n(f). To match the ARPES results with DFT, we have used a surface potential of \u00000:14 eV.\nthe sample via alkali metal (potassium) deposition (see\nFig. S6 in the SM [45]), and shift the Fermi energy above\nthe Dirac point, revealed that only the surface states in\nNiTe 2are impacted by surface deposition. Bulk doing is\nneeded to shift the bulk bands.\nExtended energy range ARPES spectra for the two\nhigh-symmetry directions are shown in Fig. 3(c) and\n(d). These spectra are mainly dominated by the surface\nstates, as seen from the comparison with the calculated\nband structure in Fig. 3 (e) and (f). In order to under-\nstand their origin, we note that there are two symmetry\ninequivalent Te atoms (Te1and Te2) and a single Ni atom\nin an unit cell of NiTe 2. The electronic con\fguration of\nNi is 3 s23d8and that of Te is 4 d105p4. We \fnd that sim-\nilar to other group-X TMDs [32, 46], the Te 5 porbital\nmanifold in NiTe 2, aided by the interplay between intra-\nlayer hopping, crystal \feld splitting, and SOC strength,\ngives rise to most of the bulk Dirac nodes and multiple\ninverted band-gaps. To highlight this, we show the evolu-\ntion of the p-orbital manifold of the Te atoms in Fig. 3(a).\nTo start with (step I in Fig. 3(a)), strong intra-layer hy-bridization between the Te1and Te2porbitals results\nin bonding and anti-bonding states. These orbitals are\nfurther split (in step II), due to a strong trigonal crystal\n\feld generated by the layered crystal structure of NiTe 2,\nseparating pzfrom the px; pyorbitals. Inclusion of SOC\n(step III) further splits the orbitals into jJ;jmJjistates.\nStep IV of Fig. 3(a) highlights the e\u000bect of the dispersion\nalong the \u0000\u0000Adirection, and the formation of the bulk\ntype-II Dirac point along with multiple band-inversions\nin the valance band.\nThe irreducible representation of some of these states\nat the at \u0000 and Apoints and along the \u0000 \u0000Ahigh-\nsymmetry line is shown in Fig. 3(b). The bulk Dirac\npoint originates from the crossing of the \u0001 4and \u0001 5;6\nstates along the \u0000 \u0000Adirection. As discussed earlier,\nthe doubly degenerate \u0001 4and \u0001 5;6bands have opposite\nrotation characters (+1 and \u00001, respectively) and, there-\nfore, the Dirac point is protected from gap opening by\nany perturbation which respects the C3symmetry. Ad-\nditionally, Fig. 3(b) also highlights the existence of a pair\nof IBGs in the valance band at the Apoint. However,4\nFIG. 4. Measured and calculated spin texture for the bands along \u0016\u0000\u0000\u0016K(a, b) and \u0016\u0000\u0000\u0016M(f, g) in the BZ. Experiments\nwere performed in the same conditions as in Figure 3; (c-e) spin-polarized spectra and spin polarization for the points along\n\u0016\u0000\u0000\u0016Kmarked by black arrows in (a); in all \fgures red/blue indicate positive/negative polarization perpendicular to the\nhigh-symmetry direction. Ebindenotes the binding energy.\nin comparison to PdTe 2, the parity of the crossing bands\nat the Apoint for NiTe 2is di\u000berent and only the lower\nIBG supports a Dirac node in the surface states. See\nSec. S5 in the SM [45] for a more detailed discussion and\ncomparison of the topological band structure and surface\nstates with PdTe 2.\nThe Dirac-like conical crossing in the surface states of\nNiTe 2(at -1.4eV) is evident in the ARPES data taken\nalong the two high-symmetry directions \u0016K\u0000\u0016\u0000\u0000\u0016Kand\n\u0016M\u0000\u0016\u0000\u0000\u0016Mas shown in Figs. 3(c) and (d). The domi-\nnant surface bands are indicated by the red arrows. In\naddition to the Dirac cone, several other surface states\nare present in NiTe 2, owing to several band inversions\nbelow and above the Fermi level. Along the \u0016M\u0000\u0016\u0000\u0000\u0016M\ndirection, the surface states near the Fermi energy has its\norigin from a band inversion above the Fermi energy (see\nFig. 3(b), and Fig. S3 in SM [45]) and the correspond-\ning surface Dirac cone lies above the Fermi level. While\nits Dirac-like nature is signi\fcantly altered far away from\nthe\u0016\u0000 point, we demonstrate its topological origin by dis-\nplaying its chiral spin texture in Fig. 4. Similar surface\nstates have also been observed in other Te-based TMDs\nlike PtTe 2and PdTe 2, while they are absent in the Se\nbased compounds like PtSe 2. This is a consequence of\nthe avoided band inversion in PtSe 2resulting from the\nreduced interlayer hopping [31, 47].\nSince these surface states have a topological origin,\nwe now focus on the spin polarization of the bands us-\ning spin-polarized ARPES. In Fig. 4(a) we display the\nspin-resolved data superimposed directly onto the spin-\nintegrated band structure shown earlier in Fig. 3(c) and\n(d). The measured spin polarization matches reason-\nably well with the calculated spin textures reported in\nFig. 4(b) and (g). In the present dataset, the spin com-\nponent is always perpendicular to the dispersion direc-tion. The most prominent feature in Fig. 4(a)-(b) is\nthe crossover of two opposite spin polarizations of almost\nequal magnitude for the surface state bands crossing at\nthe \u0000 point at a binding energy of \u0018\u00001:4 eV. This con-\n\frms the helical nature of the spin-momentum locking\nin vicinity of the surface Dirac point, resulting from the\nIBG with the Z2= 1 topological order. In Fig. 4(c)-\n(e), we display the spin-polarized spectra and the spin\npolarization for the points marked by black arrows in\nFig. 4(a). The measured polarization perpendicular to\n\u0016\u0000\u0000\u0016Kreaches almost 50%.\nIn the case of the bands along \u0016M\u0000\u0016\u0000\u0000\u0016Mdirection of\nthe surface BZ [Fig. 4 (f) and (g)], the polarization was\nmeasured for the electron-pocket-like surface states close\nto the Fermi energy. As discussed previously, although\ntheir shape is considerably di\u000berent than a usual topo-\nlogical surface states, the clear spin polarization demon-\nstrates their topological origin (see Fig. S3 of SM [45]).\nThese indeed appear to be the most prominent spin-\npolarized features also in the calculated spin polarization\nperpendicular to the \u0016\u0000\u0000\u0016Mdirection. The high values of\nthe measured and calculated spin polarization indicates\nthat NiTe 2belongs to the recently identi\fed topological-\nladder family of Pt/PdTe2 [32, 46].\nTo summarize, we have established the existence\nof type-II DSM phase in NiTe 2single crystals using\nspin-resolved ARPES measurements in conjunction with\nDFT-based ab-initio calculations. We show that, in con-\ntrast to similar class of materials like PtTe 2, PdTe 2, and\nPtSe 2, where the Dirac point is buried deep in the valence\nband, the Dirac point in NiTe 2is located in vicinity of\nthe Fermi energy. In addition to the bulk Dirac node, the\nTep-orbital manifold in NiTe 2also gives rise to a series\nof IBGs with non-trivial Z2topological orders. Together,\nthese give rise to topological Dirac nodes in the surface5\nstates characterized by the particular spin texture over\na wide range of energies. Our \fndings establish NiTe 2\nas a prime candidate for exploration of Dirac fermiology\nand applications in TMD-based spintronic devices and\nultrafast optoelectronics.\nACKNOWLEDGEMENTS\nA.A. acknowledges funding support by Dept. of Sci-\nence and Technology, Government of India, via DST\nGrant No. DST/NM/NS/2018/103(G), and from SERB\nGrant No. CRG/2018/002440. B.G. acknowledges CSIR\nfor senior research fellowship. A. A. and B. G. acknowl-\nedges HPC- IIT Kanpur for its computational facilities.\nThis work has been partly performed in the framework of\nthe nanoscience foundry and \fne analysis (NFFA-MIUR\nItaly, Progetti Internazionali) facility.\nNote: B. G. and D. M. contributed equally to this work\nas \frst authors.\n\u0003ivana.vobornik@elettra.trieste.it\nyantonio.politano@iit.it\nzamitag@iitk.ac.in\n[1] A. Bansil, Hsin Lin, and Tanmoy Das, \\ Colloquium :\nTopological band theory,\" Rev. Mod. Phys. 88, 021004\n(2016).\n[2] M. Z. Hasan and C. L. Kane, \\Colloquium: Topological\ninsulators,\" Rev. Mod. Phys. 82, 3045{3067 (2010).\n[3] N. P. Armitage, E. J. Mele, and Ashvin Vishwanath,\n\\Weyl and dirac semimetals in three-dimensional solids,\"\nRev. Mod. Phys. 90, 015001 (2018).\n[4] Desheng Kong and Yi Cui, \\Opportunities in chemistry\nand materials science for topological insulators and their\nnanostructures,\" Nat. Chem. 3, 845 (2011).\n[5] Barry Bradlyn, Jennifer Cano, Zhijun Wang, M. G.\nVergniory, C. Felser, R. J. Cava, and B. Andrei Bernevig,\n\\Beyond dirac and weyl fermions: Unconventional quasi-\nparticles in conventional crystals,\" Science 353, 6299\n(2016).\n[6] Su-Yang Xu, Ilya Belopolski, Nasser Alidoust, Mad-\nhab Neupane, Guang Bian, Chenglong Zhang, Raman\nSankar, Guoqing Chang, Zhujun Yuan, Chi-Cheng Lee,\nShin-Ming Huang, Hao Zheng, Jie Ma, Daniel S. Sanchez,\nBaoKai Wang, Arun Bansil, Fangcheng Chou, Pavel P.\nShibayev, Hsin Lin, Shuang Jia, and M. Zahid Hasan,\n\\Discovery of a weyl fermion semimetal and topological\nfermi arcs,\" Science 349, 613{617 (2015).\n[7] Shin-Ming Huang, Su-Yang Xu, Ilya Belopolski, Chi-\nCheng Lee, Guoqing Chang, BaoKai Wang, Nasser Ali-\ndoust, Guang Bian, Madhab Neupane, Chenglong Zhang,\nShuang Jia, Arun Bansil, Hsin Lin, and M. Zahid Hasan,\n\\A weyl fermion semimetal with surface fermi arcs in the\ntransition metal monopnictide TaAs class,\" Nature Com-\nmunications 6, 7373 (2015).\n[8] B. Q. Lv, H. M. Weng, B. B. Fu, X. P. Wang, H. Miao,\nJ. Ma, P. Richard, X. C. Huang, L. X. Zhao, G. F. Chen,\nZ. Fang, X. Dai, T. Qian, and H. Ding, \\Experimen-tal discovery of weyl semimetal TaAs,\" Phys. Rev. X 5,\n031013 (2015).\n[9] Su-Yang Xu, Nasser Alidoust, Ilya Belopolski, Zhu-\njun Yuan, Guang Bian, Tay-Rong Chang, Hao Zheng,\nVladimir N. Strocov, Daniel S. Sanchez, Guoqing Chang,\nChenglong Zhang, Daixiang Mou, Yun Wu, Lunan\nHuang, Chi-Cheng Lee, Shin-Ming Huang, BaoKai\nWang, Arun Bansil, Horng-Tay Jeng, Titus Neupert,\nAdam Kaminski, Hsin Lin, Shuang Jia, and M. Za-\nhid Hasan, \\Discovery of a weyl fermion state with fermi\narcs in niobium arsenide,\" Nature Physics 11, 748 (2015).\n[10] Chen Fang, Matthew J. Gilbert, Xi Dai, and B. Andrei\nBernevig, \\Multi-weyl topological semimetals stabilized\nby point group symmetry,\" Phys. Rev. Lett. 108, 266802\n(2012).\n[11] Hongming Weng, Chen Fang, Zhong Fang, B. Andrei\nBernevig, and Xi Dai, \\Weyl semimetal phase in noncen-\ntrosymmetric transition-metal monophosphides,\" Phys.\nRev. X 5, 011029 (2015).\n[12] Bahadur Singh, Ashutosh Sharma, H. Lin, M. Z. Hasan,\nR. Prasad, and A. Bansil, \\Topological electronic struc-\nture and weyl semimetal in the TlBiSe 2class of semicon-\nductors,\" Phys. Rev. B 86, 115208 (2012).\n[13] Tay-Rong Chang, Su-Yang Xu, Daniel S. Sanchez, Wei-\nFeng Tsai, Shin-Ming Huang, Guoqing Chang, Chuang-\nHan Hsu, Guang Bian, Ilya Belopolski, Zhi-Ming Yu,\nShengyuan A. Yang, Titus Neupert, Horng-Tay Jeng,\nHsin Lin, and M. Zahid Hasan, \\Type-ii symmetry-\nprotected topological dirac semimetals,\" Phys. Rev. Lett.\n119, 026404 (2017).\n[14] Su-Yang Xu, Nasser Alidoust, Guoqing Chang, Hong Lu,\nBahadur Singh, Ilya Belopolski, Daniel S. Sanchez, Xiao\nZhang, Guang Bian, Hao Zheng, Marious-Adrian Hu-\nsanu, Yi Bian, Shin-Ming Huang, Chuang-Han Hsu, Tay-\nRong Chang, Horng-Tay Jeng, Arun Bansil, Titus Neu-\npert, Vladimir N. Strocov, Hsin Lin, Shuang Jia, and\nM. Zahid Hasan, \\Discovery of lorentz-violating type II\nweyl fermions in LaAlGe,\" Science Advances 3(2017),\n10.1126/sciadv.1603266.\n[15] B. Q. Lv, Z. L. Feng, Q. N. Xu, X. Gao, J. Z. Ma, L. Y.\nKong, P. Richard, Y. B. Huang, V. N. Strocov, C. Fang,\nH. M. Weng, Y. G. Shi, T. Qian, and H. Ding, \\Ob-\nservation of three-component fermions in the topologi-\ncal semimetal molybdenum phosphide,\" Nature 546, 627\n(2017).\n[16] Wenshuai Gao, Xiangde Zhu, Fawei Zheng, Min Wu, Jin-\nglei Zhang, Chuanying Xi, Ping Zhang, Yuheng Zhang,\nNing Hao, Wei Ning, et al. , \\A possible candidate\nfor triply degenerate point fermions in trigonal layered\nPtBi 2,\" Nature communications 9(2018).\n[17] Ziming Zhu, Georg W. Winkler, QuanSheng Wu, Ju Li,\nand Alexey A. Soluyanov, \\Triple point topological met-\nals,\" Phys. Rev. X 6, 031003 (2016).\n[18] Jiabin Yu, Binghai Yan, and Chao-Xing Liu, \\Model\nhamiltonian and time reversal breaking topological\nphases of antiferromagnetic half-heusler materials,\"\nPhys. Rev. B 95, 235158 (2017).\n[19] I. C. Fulga and Ady Stern, \\Triple point fermions in a\nminimal symmorphic model,\" Phys. Rev. B 95, 241116\n(2017).\n[20] Hao Yang, Jiabin Yu, Stuart S. P. Parkin, Claudia Felser,\nChao-Xing Liu, and Binghai Yan, \\Prediction of triple\npoint fermions in simple half-heusler topological insula-\ntors,\" Phys. Rev. Lett. 119, 136401 (2017).6\n[21] Jianfeng Wang, Xuelei Sui, Wujun Shi, Jinbo Pan,\nShengbai Zhang, Feng Liu, Su-Huai Wei, Qimin\nYan, and Bing Huang, \\Prediction of ideal topolog-\nical semimetals with triply degenerate points in the\nNaCu 3Te2family,\" Phys. Rev. Lett. 119, 256402 (2017).\n[22] S. M. Young, S. Zaheer, J. C. Y. Teo, C. L. Kane, E. J.\nMele, and A. M. Rappe, \\Dirac semimetal in three di-\nmensions,\" Phys. Rev. Lett. 108, 140405 (2012).\n[23] Steve M. Young and Charles L. Kane, \\Dirac semimetals\nin two dimensions,\" Phys. Rev. Lett. 115, 126803 (2015).\n[24] Benjamin J. Wieder, Youngkuk Kim, A. M. Rappe, and\nC. L. Kane, \\Double dirac semimetals in three dimen-\nsions,\" Phys. Rev. Lett. 116, 186402 (2016).\n[25] Z. K. Liu, B. Zhou, Y. Zhang, Z. J. Wang, H. M. Weng,\nD. Prabhakaran, S.-K. Mo, Z. X. Shen, Z. Fang, X. Dai,\nZ. Hussain, and Y. L. Chen, \\Discovery of a three-\ndimensional topological dirac semimetal, Na 3Bi,\" Science\n343, 864{867 (2014).\n[26] Z. K. Liu, J. Jiang, B. Zhou, Z. J. Wang, Y. Zhang, H. M.\nWeng, D. Prabhakaran, S-K. Mo, H. Peng, P. Dudin,\nT. Kim, M. Hoesch, Z. Fang, X. Dai, Z. X. Shen, D. L.\nFeng, Z. Hussain, and Y. L. Chen, \\A stable three-\ndimensional topological dirac semimetal Cd 3As2,\" Na-\nture Materials 13, 677 (2014).\n[27] Madhab Neupane, Su-Yang Xu, Raman Sankar, Nasser\nAlidoust, Guang Bian, Chang Liu, Ilya Belopolski, Tay-\nRong Chang, Horng-Tay Jeng, Hsin Lin, Arun Bansil,\nFangcheng Chou, and M. Zahid Hasan, \\Observation of\na three-dimensional topological dirac semimetal phase in\nhigh-mobility Cd 3As2,\" Nature Communications 5, 3786\n(2014).\n[28] Sergey Borisenko, Quinn Gibson, Danil Evtushinsky,\nVolodymyr Zabolotnyy, Bernd B uchner, and Robert J.\nCava, \\Experimental realization of a three-dimensional\ndirac semimetal,\" Phys. Rev. Lett. 113, 027603 (2014).\n[29] Zhijun Wang, Yan Sun, Xing-Qiu Chen, Cesare Fran-\nchini, Gang Xu, Hongming Weng, Xi Dai, and Zhong\nFang, \\Dirac semimetal and topological phase transitions\nin A 3Bi (a = Na, K, Rb),\" Phys. Rev. B 85, 195320\n(2012).\n[30] Zhijun Wang, Hongming Weng, Quansheng Wu, Xi Dai,\nand Zhong Fang, \\Three-dimensional dirac semimetal\nand quantum transport in Cd 3As2,\" Phys. Rev. B 88,\n125427 (2013).\n[31] Mingzhe Yan, Huaqing Huang, Kenan Zhang, Eryin\nWang, Wei Yao, Ke Deng, Guoliang Wan, Hongyun\nZhang, Masashi Arita, Haitao Yang, Zhe Sun, Hong Yao,\nYang Wu, Shoushan Fan, Wenhui Duan, and Shuyun\nZhou, \\Lorentz-violating type-ii dirac fermions in tran-\nsition metal dichalcogenide PtTe 2,\" Nature Communica-\ntions8, 257 (2017).\n[32] M. S. Bahramy, O. J. Clark, B. J. Yang, J. Feng, L. Baw-\nden, J. M. Riley, I. Markovi\u0013 c, F. Mazzola, V. Sunko,\nD. Biswas, S. P. Cooil, M. Jorge, J. W. Wells, M. Lean-\ndersson, T. Balasubramanian, J. Fujii, I. Vobornik, J. E.\nRault, T. K. Kim, M. Hoesch, K. Okawa, M. Asakawa,\nT. Sasagawa, T. Eknapakul, W. Meevasana, and P. D. C.\nKing, \\Ubiquitous formation of bulk dirac cones and\ntopological surface states from a single orbital manifold in\ntransition-metal dichalcogenides,\" Nature Materials 17,\n21 EP { (2017).\n[33] Antonio Politano, Gennaro Chiarello, Barun Ghosh, Kr-\nishanu Sadhukhan, Chia-Nung Kuo, Chin Shan Lue, Vit-\ntorio Pellegrini, and Amit Agarwal, \\3d dirac plasmonsin the type-ii dirac semimetal PtTe 2,\" Phys. Rev. Lett.\n121, 086804 (2018).\n[34] Kenan Zhang, Mingzhe Yan, Haoxiong Zhang, Huaqing\nHuang, Masashi Arita, Zhe Sun, Wenhui Duan, Yang\nWu, and Shuyun Zhou, \\Experimental evidence for type-\nii dirac semimetal in PtSe 2,\" Phys. Rev. B 96, 125102\n(2017).\n[35] Han-Jin Noh, Jinwon Jeong, En-Jin Cho, Kyoo Kim, B. I.\nMin, and Byeong-Gyu Park, \\Experimental realization\nof type-ii dirac fermions in a PdTe 2superconductor,\"\nPhys. Rev. Lett. 119, 016401 (2017).\n[36] Fucong Fei, Xiangyan Bo, Rui Wang, Bin Wu, Juan\nJiang, Dongzhi Fu, Ming Gao, Hao Zheng, Yulin Chen,\nXuefeng Wang, Haijun Bu, Fengqi Song, Xiangang Wan,\nBaigeng Wang, and Guanghou Wang, \\Nontrivial berry\nphase and type-ii dirac transport in the layered material\nPdTe 2,\" Phys. Rev. B 96, 041201 (2017).\n[37] Chunqiang Xu, Bin Li, Wenhe Jiao, Wei Zhou, Bin Qian,\nRaman Sankar, Nikolai D. Zhigadlo, Yanpeng Qi, Dong\nQian, Fang-Cheng Chou, and Xiaofeng Xu, \\Topologi-\ncal type-ii dirac fermions approaching the fermi level in\na transition metal dichalcogenide NiTe 2,\" Chemistry of\nMaterials 30, 4823{4830 (2018).\n[38] Enji Uchida and Hisamoto Kondoh, \\Magnetic properties\nof nickel telluride,\" Journal of the Physical Society of\nJapan 11, 21{27 (1956).\n[39] M. Ettenberg, K.L. Komarek, and E. Miller, \\Thermo-\ndynamic properties of nickel-tellurium alloys,\" Journal of\nSolid State Chemistry 1, 583 { 592 (1970).\n[40] G Y Guo and W Y Liang, \\Study of the electronic\nstructures of ni-group metal ditellurides: NiTe 2, PdTe 2\nand PtTe 2by the self-consistent LMTO-ASA method,\"\nJournal of Physics C: Solid State Physics 19, 5365{5380\n(1986).\n[41] P J Orders, J Liesegang, R C G Leckey, J G Jenkin,\nand J D Riley, \\Angle-resolved photoemission from the\nvalence bands of NiTe 2, PdTe 2and PtTe 2,\" Journal of\nPhysics F: Metal Physics 12, 2737{2753 (1982).\n[42] Joo Frederico Haas Leandro Monteiro, Mariana Braga\nMarciniak, Alcione Roberto Jurelo, Ezequiel Costa\nSiqueira, Fbio Teixeira Dias, and Jorge Luiz Pimentel\nJnior, \\Synthesis and microstructure of NiTe 2,\" Journal\nof Crystal Growth 478, 129 { 133 (2017).\n[43] Bei Zhao, Weiqi Dang, Yuan Liu, Bo Li, Jia Li, Jun Luo,\nZhengwei Zhang, Ruixia Wu, Huifang Ma, Guangzhuang\nSun, Yu Huang, Xidong Duan, and Xiangfeng Duan,\n\\Synthetic control of two-dimensional NiTe 2single crys-\ntals with highly uniform thickness distributions,\" Jour-\nnal of the American Chemical Society 140, 14217{14223\n(2018).\n[44] Qianqian Liu, Fucong Fei, Bo Chen, Xiangyan Bo,\nBoyuan Wei, Shuai Zhang, Minhao Zhang, Faji Xie,\nMuhammad Naveed, Xiangang Wan, Fengqi Song, and\nBaigeng Wang, \\Nontopological origin of the planar hall\ne\u000bect in the type-ii dirac semimetal NiTe 2,\" Phys. Rev.\nB99, 155119 (2019).\n[45] Supplementary material detailing 1) Sample prepara-\ntion and characterization, 2) computational details, 3)\nARPES measurements, 4) Topological band inversions\nand 5) ARPES measurements with potassium doping, is\navailable at https://bit.ly/2YWZf1o .\n[46] O J Clark, F Mazzola, I Markovi\u0013 c, J M Riley, J Feng,\nB-J Yang, K Sumida, T Okuda, J Fujii, I Vobornik, T K\nKim, K Okawa, T Sasagawa, M S Bahramy, and P D C7\nKing, \\A general route to form topologically-protected\nsurface and bulk dirac fermions along high-symmetry\nlines,\" Electronic Structure 1, 014002 (2019).\n[47] O. J. Clark, M. J. Neat, K. Okawa, L. Bawden,\nI. Markovi\u0013 c, F. Mazzola, J. Feng, V. Sunko, J. M. Ri-\nley, W. Meevasana, J. Fujii, I. Vobornik, T. K. Kim,M. Hoesch, T. Sasagawa, P. Wahl, M. S. Bahramy, and\nP. D. C. King, \\Fermiology and superconductivity of\ntopological surface states in PdTe 2,\" Phys. Rev. Lett.\n120, 156401 (2018)." }, { "title": "1304.3661v1.Electric_Field_Tuning_of_the_Surface_Band_Structure_of_Topological_Insulator_Sb2Te3_Thin_Films.pdf", "content": " \n1 \n Electric Field Tuning of the Surface Band Structure of Topological \nInsulator Sb 2Te3 Thin Films \n \nTong Zhang1,2, Jeonghoon Ha1,2,3, Niv Levy1,2, Young Kuk3, Joseph Stroscio1* \n \n1Center for Nanoscale Science and Technology, NIST, Gaithersburg, MD 20899, USA \n2Maryland NanoCenter, University of Maryland, College Park, MD 20742, USA \n3Department of Physics and Astronomy, Seoul National University, Seoul 151 -747, Korea \n \nWe measured the response of the surface state spectrum of epitaxial Sb 2Te3 thin films to \napplied gate electric fields by low temperature scanning tunneling microscopy. The gate \ndependen t shift of the Fermi level and the screening effect from bulk carrie rs vary as a function \nof film thickness. We observe d a gap opening at the Dira c point for films thinner than four \nquintuple layers, due to the coupling of the top and bottom surfaces . Moreover, the top surface \nstate band gap of the three quintuple layer films was found to be tunable by back gate , indicating \nthe possibility of observing a topological phase transition in this system . Our result s are well \nexplained by an effective model of 3D topological insulator thin films with structure inversion \nasymmetry , indicati ng that three quintuple layer Sb2Te3 films are topologically nontrivial and \nbelong to the quantum spin Hall insulator class . \n \n2 \n Topological insulators (TI) represent a novel state of quantum matter that has \ntopologically protected edge or surface states [1,2] . Theory and experimental verification of TI \nmatter rapidly increased since the first predictions of these states of matter in 2 -dimensions (2D) . \n2D TIs are equivalent to quantum spin Hall (QSH) insulators which host 1-dimensional (1D) \nspin polarized edge states [3,4] . The concept of QSH insulators has been generalized to three -\ndimensional (3D) TIs, which are 3D ban d insulators surrounded by 2D metallic surface states \nwith helical spin texture [1,2] . These surface stat es are protected by time reversal symmetry and \nare not eliminated by scattering from weak non-magnetic disorder. Recently, t he family of \nbismuth chalcogenide materials (Bi 2Se3, Bi 2Te3 and Sb 2Te3) have been confirmed to be 3D TIs \nwith a single Dirac cone at the Γ point [5–11]. In sufficiently thin 3D TI films the top and \nbottom surface states hybridize , disrupt ing the Dirac cone and creat ing an energy gap at k=0, \nallowing a range of topologically interesting phases [12–21]. In the ultrathin film geometry, the \nlow-lying physics is described by two degenerate Dirac hyperbolas. Each massive band has a k-\ndependent spin configuration determined by the energy gap and spin -orbit coupling. As a result, \nultrathin 3D TI film s can be considered a 2D QSH system with 1D spin helical edge states if the \nsystem is topologically nontrivial. Calculations show that the surface band gap of the ultrathin \n3D TIs displays an oscillating dependence on thickness [13–21], and the system may alterna te \nbetween being a topologically trivial or no ntrivial insulator . More interestingly, the band \ntopology is predicted to be tunable by external electric fields , which cause structure inversion \nasymmetry (SIA), leading to the possibility of controlling a ph ase transition between \ntopologically trivial and nontrivial phases [15–21]. For practical applications it is important to \nstudy the surface state properties of different phases. This may enable control of the QSH edge \nstates with potential applications in future spintronic devices. \n3 \n In this letter , we stud y the gate tunable electron ic structure of the 3D TI Sb2Te3 using low \ntemperature scanning tunneling microscopy and spectroscopy (STM/STS) on in-situ epitaxiall y \ngrown thin films on back gated SrTiO 3 substrates . We find that t he field effect gating of the thin \nfilms varies strongly with film thickness . This was determined by measuring the shift of the thin \nfilm quantum well states with back gate potential. For films below four quintuple layers (QL) \nthickness a band gap is observed in the surface state spectrum, which we attribute to the \nhybridization of the top and bottom surfaces. A t a film thickness of 3 QL the field effect is able \nto reduce the gap with incr eased gate field . Comparing an effective Hamiltonian model with \nstructure inversion asymmetry to the gap dependence on gating implies that 3 QL Sb 2Te3 film is \ntopological ly non-trivial , and suggest s that a topological phase transition may occur if the \nelectric field applied to the film is large enough. \nDespite many transport measurement s on gated TI devices [22–26], making a gate \ntunable TI film that is accessible to low temperature STM is challenging, mainly because \nchalcogenide TI’s surfaces are environmentally sensitive and degrade easily through exposure to \nair during standard lithography proce sses. We overcome this problem by epitaxially growing TI \nfilms on pre -patterned SrTiO 3 (111) (STO) substrates mounted on special sample holders [27]. \nThe sample s were transferred in ultra-high vacuum into an STM right after growth, which \navoided any ex-situ processing. The sample bias is applied through two top electrodes and a gate \nvoltage is applied to a back contact on the STO. STM/STS and two -terminal transport \nmeasurements are both carried out in-situ, in a homemade STM operating at 5 K, which is \nconnected to the molecul ar beam epitaxy (MBE) system [28]. The sample s studied here are \nSb2Te3 thin film s grow n by co -depositing Sb and Te at a substrate temperature of 200 °C. \nSamples were characterized with reflection -high-energy -electron diffraction during growth. \n4 \n Fig. 1(a) shows the typ ical topography of a nominally 5 QL thick Sb 2Te3 film. Terraces \nwith thick nesses between 2 QL to 5 QL can be found on the same film , and their height \ndetermined by finding pinholes to the STO substrate level. Figure 1(b) shows atomic resolution \nof the Te terminated (111) surface with lattice constant of 0.42 nm. The differential tunneling \nconductance , dI/dV, is measured by standard lock -in technique s, with a modulation frequency of \n≈500 Hz. Typical dI/dV spectra on different layer thickness es are shown in Fig. 1(c). The \npronounced peaks (see arrows in Fig. 1(c)) are attributed to quantum well states (QWS) from \nbulk bands that undergo quantum confinement in a thin film geometry. [29]. A closer inspection \nof the surface state region, obtained at lower junction impedance, shows a hybridized gap a t the \nsample bias ≈VB=0.25 V on 2 QL and 3 QL terraces [ Fig. 1(d)], with a width of ≈160 meV and \n80 meV, respectively. The gap is absent for the 4 QL film and leaves a “V” shaped density of \nstates that indicates an intact Dirac cone. Similar results were reported on Sb 2Te3 film grown on \ngraphitized SiC [30]. \nOur experiment al setup allows , for the first time , for a gate voltage VG to be applied to the \nSb2Te3 film during STM measurement s, without any ex-situ processing and exposure of the film \nto atmosphere . To measure the gating characteristics , we charged the sample (which can be \nthought of as a TI-STO -gate capacitor) with a current source while measuring the change of VG. \nFig. 1 (e) shows the total charge density n and displacement field versus VG by integrating the \ncurrent over time. n reaches about 4.5×1013/cm2 at VG = 200 V, which indicates a large gating \nability of STO at low temperature. A two-terminal film resistance measurement versus VG is \nshown in Fig. 1(f), which displays a rapid increase at positive VG, as expected for gating a p-\ndoped semiconductor. \n5 \n The gating effect on the local density of states (LDOS) is measured using the thin film \nQWS as a fiduciary mark in the tunneling spectra [ Fig. 1(c)]. The overall gati ng dependence of \nthe LDOS is shown in Figs. 2(a) and 2(b) for 2 QL and 3 QL films by focusing on a large sample \nbias range ( -0.4 V to 0.4 V) . Increasing gate voltage is accompanied by the whole spectrum \nmoving to lower energies, wh ich is a clear signature of the shifting of the Fermi energy EF due to \nthe gate induced doping of the Sb2Te3 film. Note in the tunneling experiment EF is at VB = 0 V, \nand therefore the change in doping is observed by the shift of the electronic bands rela tive to \nzero sample bias . The relative shifts of EF (with respect to VG=0) as a function of displacement \nfield are plotted in Fig. 2(c) for film thicknesses of 2 QL, 3 QL, and 4 QLs. One can see that t he \ngating tu nability decrease s fast with increasing film thickness , approximately as \n2d [Fig. 2(d)] . \nFor the 2 QL film, the thinnest case, we estimate the change in surface carrier density , \ncorresponding to the measured shift in EF, only reach es to ¼ of the total charge density induced \nby the gate. Therefore the majority of the carriers are expected to be in the bulk of the film , \nwhich screens the electric field reaching the top surface. Subsequently, in these thin films an \noverall Fermi level shift and band bending will coexist through the film providing a potential \nasymmetry between the top and bottom surfaces [31]. Nevertheless, th e gate tunability is still \nsufficient to deduce the topological character of the 3 QL f ilm, as described below. \nNow we focus on the hybridization gaps that open due to the coupling of the top and \nbottom surfaces , and their response to applied fields . Figures 3(a) and 3(b) show the tunneling \nspectra of the surface state band gaps for 2 QL an d 3 QL terraces, respectively, as a fu nction of \nVG. The gap size is measured by the peak positions in the second derivative d2I/dV2, which \nindicate the inflection points on either side of the surface state gap [see dashed line in Fig. 3(a)] . \n6 \n A noticeable feature is that the surface states gap is rather constant at 2 QL, but varies \nconsiderably with VG for 3 QL . Differences in the response of the gap to fields , as observed \nbetween the 2 QL and 3 QL films, can occur due changes in the topology of the system . Such \nchanges can arise with variations as small as 1 QL in film thickness [18,19] . The gap for the 3 \nQL film shows a linear dependence on displacement field [Fig. 3(d )]. This is reminiscent of bi -\nlayer graphene where the asymmetric potential controls the gap at the Dirac point [32,33] . The \ndifference here is that in a TI thin film, the inter -layer and spin orbit coupling play important \nroles. \nTo understand how an electric field affects the surface band structure, we refer to an \neffective Hamiltonian model that describes thin 3D TI films as [17,18] , \n \n22\n0ˆ ˆ ˆ ˆ ˆ ()2eff F y x x y z z xH k E Dk v k k Bk U \n (1) \nThe first three terms in Eq. (1) account for isolated TI surface states. The fourth term describes \nthe inter -surface coupling, w hich induces a hybridized gap \n and parabolic term \n2()Bk . \nˆ and \nˆ\n are Pauli matrices of electron spin and isospin (i.e. bonding/anti -bonding states of the top and \nbottom surfaces) , respectively . The asymmetric potential leading to structure inversion \nasymmetry (SIA) is introduced by adding an effective potential energy \nU to \neffH . The \nparameters \n0, , , ,F E D B v and \n are material and thickness dependent , and \nU is the potential \ndifference between the top and bottom surfaces induced by the gate electric field . Solving for \nthe energy eigenvalues for t he effective Hamiltonian in Eq. (1) gives rise to four surface state \nenergy bands [17,18] , \n7 \n \n2\n222\n102F E E Dk Bk U v k \n (2) \n \n2\n222\n202F E E Dk Bk U v k \n (3) \n \nwhere the \n() sign stands for the conduction (valence ) band, and the 1 (2) stands for the inner \n(outer) branches of the bands . Without the SIA term \n( 0)U these states consist of spin \ndegenerate conduction and valence bands separated by a hybridization gap [Fig. 4(a)]. The \npresence of applied potential U leads to a Rashba -like splitting o f conduction and valence bands \n[Figs. 4(b)]. Moreover, from Eq s. (2) and (3), the actual gap size also var ies with potential \nasymmetry [see Figs. 4, (a) -(d)]. This is directly related to our observations of the 3 QL Sb 2Te3 \ngap decreasing with increasing applied gate potential [Figs. 3, (b ) and (d) ], which is examined \nbelow . \nAs discussed in R efs. [17] and [18], the sign of \n/B and the value of the SIA term \ndetermine if the system is a topologically trivial or non -trivial insulator . If \n22BD and ∆ and B \nhave the same sign, i.e. \n/0B , the system is in the quantum spin hall state. The SIA term \nU\n reduce s the energy gap [ Fig. 4(b)] and eventually close it at the critical potential difference\n/2CFU v B\n [Fig. 4(c)] and then reopen it with increasing potential [Fig. 4(d)], leading to \na topological phase transition as a function of applied electric field [17,19] . If ∆ and B have \nopposite sign, the system remain s gapped as a function of applied electric field. Fig. 4(f) \ncompares these two cases, which we use below to determine the topological character of the 3 \nQL film. We note that \nU contributes to Eq. (2) and (3) through its absolute value , so the gap \n8 \n variation should be symmetric with respect to the gate voltage. However, in Fig. 3(d) the gap \ndisplays a monotonic dependence on displacement field within the range investigated . We \nexpect t his lack of symmetry is due to an initial band bending which exists even at VG = 0 [see \nschematic in Fig. 4(e)] and is common for hetero -junctions . \nWe can estimate th e initial bending direction from the thickness dependence of the \nposition of the Dirac point (or mid -gap energy position) . As shown by the dashed line in Fig. \n1(d), the middle of the gap moves closer to EF with incre asing thickness from 2 QL to 4 QL, \nwhich implies the top surface of the film is less doped at increased thicknesses. The increased \ndoping for thinner films is probably due to an increased number of defects at the TI/STO \ninterface . These defects are very effective acceptors [29]. The higher hole doping at the \ninterface gives an upward band bending through the film . The band bending is present at zero \napplied gate potential [Fig. 4(e)] and will affect the surface band gap as if a gate field where \napplied [31]. Applying a negative gate potential leads to hole accumulation and further \nenhances the band bending intensity , and hence the potential asymmetry [Fig. 4(e) top panel]. \nWe expect that the potential difference between the top and botto m 3 QL film is no less than the \ndifference of the EF shifts of the top surfaces of the 2 QL and 3 QL films, i.e. 45 meV at VG = \n200 V. Therefore, i f the system is non -trivial , applying a negative gate potential leads to a \ndecreas ed gap according to the above analysis [Fig. 4(f)] . This is what we observed in Fig 3(b) \nfor 3 QL . We model the spectra in Fig. 3(b ) for 3 QL by calculating the density of states from \nEqs. (2) and ( 3), taking \n2.5Fv eV Å, B = 15 eV Å2, and D = -5 eV Å2 from fits to \nphotoemission spectroscopy and density functional calculation results in Ref. [34]. The model \nDOS curves are convoluted with a Gaussian ( with standard deviation σ=20 meV) to account for \n9 \n instrumental broadening and are show n in Fig. 3(c). The fitting parameters used in the DOS \ncalculation are: = 0.16 eV, with E0 varying from 0.24 eV to 0.27 eV, and varying from 0.07 \neV to 0.12 eV, as VG changes from 200 V to -200 V . The variation of is within the range of \ntotal potential as ymmetry we achieved by gating (≈ 50 meV). The simulated spectra in Fig. 3(c) \nare in good agreement with experiment [Fig. 3(b)]. The lack of an observed maximum in the \nexperimental gap measurements indicates that the initial band bending at VG = 0 is so strong that \neven VG = 200 V cannot flatten it. From the model calculation in Fig. 3(c), we observe the gap \ndecreasing with increased potential asymmetry in agreement with the experimental observations . \nAs shown in Fig. 3(d), t he band gap s determined from the calculation show a trend quantitative ly \nsimilar to the experiment for the range of potential asymmetry chosen for the model calculation . \nWe note that the experimental spectra have a different background than the model DOS, which is \nlikely from contributions from bulk states. \nFrom the model parameters, ∆ and B have the same sign, \n/0B , which indicates a \ntopologically nontrivial QSH phase. We determine that the 3 QL Sb2Te3 system is in a nontriv ial \ntopological state by noting that the band gap decreas es with negative applied gate voltage . For \nthe initial p -type doping, such a decrease can only occur if ∆ and B have the same sign [see Fig. \n4(f)]. The topological character of the system is typically characterized by the Z2 invariant , \nwhere Z2=1 signifies the topological nontrivial phase [1,2] . Figure 4( g) summarizes the phase \ndiagram in terms of the expected Z2 invariant deduced from our results; at VG=0, 3 QL Sb2Te3 is \nin a topologically nontrivial phase and th us should undergo a phase transition with increasing \ngate field. This conclusion is similar to what has been predicted for 4 QL Sb 2Te3 thin films [19]. \nGapless edge states are supposed to exist in this regime [13,17,19] and s earching for these edge \n10 \n states will be a n interesting subject for further studies. We expect that the topological phase \ntransition should also be observable in future studies with more insulating material to obtain \nhigher electric fields . \n \nAcknowledgements : We thank Mark Stiles, Paul Haney, Minsu ng Kim and Jisoon Ihm \nfor useful discussions . We thank Steve Blankenship and Alan Band for technical assistance. N. \nL., T. Z., and J. H. acknowledge support under the Cooperative Research Agreement between the \nUniversity of Maryland and the National Inst itute of Standards and Technology Center for \nNanoscale Science and Technology, Grant No. 70NANB10H193, through the University of \nMaryland. J. H. and Y. K. are partly supported by Korea Research Foundation through Grant No. \nKRF -2010 -00349. \n \n11 \n Figure Captions \nFigure 1: Characterization of Sb 2Te3 thin films. (a) STM topographic image, 150 nm × 150 nm, \nof nominally five -QL Sb 2Te3 film grown by MBE on SrTiO 3(111) . (b) Atom ic resolution STM \nimage, 10 nm × 10 nm of Sb 2Te3 film. The STM topographic height is shown in a color scale \nfrom dark to bright covering a range of 5 nm for (a) and 0.14 nm for (b). Tunneling parameters: \nI=30 pA, VB=1.5 V for (a) and VB =0.2 V for (b). (c) -(d) dI/dV spectra as a function of Sb 2Te3 \nfilm thickness for VG=0 V , for wide (c) and n arrow (d) energy ranges . The narrow energy range \nspectra are obtained with lower tunneling impedance to boost signal to noise in the band gap \nregion . The curves are offset vertically for clarity. The dashed line in (d) indicates the shift of \nthe mid -gap position toward the Fermi level with increased thickness. (e) Total charge density \nand displacement field versus gate voltage for a 3 QL Sb 2Te3/SrTiO 3 device obtained by \nmeasuring the induced voltage as the device was charged at constant current. (f) 3 QL Sb 2Te3 \nfilm two terminal resistance versus gate voltage. \n \nFigure 2: In-situ measur ement of electric field gating of MBE grown Sb2Te3 thin films. dI/dV \nversus VB spectra as a function of the indicated back gate voltage for a (a) 2 QL thick film, and \n(b) 3 QL thick film. A shift of the QW peaks to higher energy is seen with increased gate \npotential, as indicated by the dashed lines which are guides to the eye. (c) The relative shift of \nthe QW peak positions versus displacement fie ld as the gate potential is varied for 2, 3, and 4 QL \nfilms. Error bars are one standard deviation uncertainties in determin ing the QW peak pos itions. \nThe solid lines are linear fits to the data points. (d) The gating tunability versus film thicknes s \nobtained from the slopes of the linear fits of the QW peak shifts in (c). The error bars are one \nstandard deviation uncertainties from the linear fits. The solid line is a polynomial fit to the data \npoints . \n \nFigure 3: Electric field effect of surface state band gaps in MBE grown Sb2Te3 thin films. dI/dV \nversus VB spectra of the surface state gap as a function of the indicated back gate voltage for (a) \n2 QL Sb2Te3 film, and (b) 3 QL Sb2Te3 film. The curves are offset for clarity. (c) Calculated \nsurface state DOS using the effective Hamiltonian in Eq. (1) and resulting dispersions in Eqs. (2) \nand (3) as a function of applied potential. See text for model parameters. (d) Measured surface \nstate band gap f rom the spectra in (b) of 3 QL Sb2Te3 thin films as a function of displacement \nfield (blue round symbols), and surface state band gap from the model DOS calculation in (c) \nversus γ U (red square symbols). The band gaps are determined from the peak minimum and \nmaximum in the 2nd deriv ative spectra (see green dashed curve in (a)) for both the experiment \nand model calculation . The error bars are one standard deviation uncertainty determined from \nthe uncertainty in fitting the peaks. The solid line is a linear fit to the data with a slo pe of 17.3 ± \n2.5 meV/V Å-1. The error is one standard deviation uncertainty determined from the linear fit. \n \nFigure 4: Topological phase transitions in applied electric field. (a) -(d) Surface state energy \nversus momentum dispersion calculated using Eqs. (2) and (3) showing the variation of the band \ngap as a function of increasing structure inversion asymmetry potential. See text for model \nparameters. At the potential γ U=0.18 =γ UC the gap closes indicating a topological phase \n12 \n transition. (e) Schematic o f band bending through the TI film. Th e diagram shows the \nconduction and valence bands going through the film. At each interface the surface state \nhyperbolic dispersions are indicated. At VG=0 there is initial band bending indicating an initial \npotential asymmetry γ U0 which increases with negative gate voltage. (f) Surface state b and gap s \ncalculated from the 3D TI model using Eqs. (2) and (3) with Δ=0.16 eV, D=-5 eV Å2, vF=2.5 \neV Å , and Δ/B> 0 (blue; B=15 eV Å2), and Δ/B< 0 (red; B=-15 eV Å2). (g) Schematic of the Z 2 \ntopological invariant for 3 QL Sb2Te3 thin films as deduced from the analysis in this paper. The \nschematic shows the initial potential asymmetry position, and the topological phase transition \nthat is approached with increasing negati ve gate voltage leading to a decrease of the surface state \nband gap as seen in (a) to (c) and in the experimental data in Figs. 3, (b) and (d). The grey bar \nindicates the approximate range of potential asymmetry probed by the current experiment [see \nFig. 3(d)] . \n \n13 \n \nReference s \n* Author to whom cor respondence should be addressed. joseph.stroscio@nist.gov \n[1] M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010). \n[2] X.-L. Qi and S. -C. Zhang, Rev. Mod. Phys . 83, 1057 (2011). \n[3] C. L. Kane and E. J. Mele, Phys. Rev. Lett. 95, 226801 (2005). \n[4] B. A. Bernevig and S. -C. Zhang, Phys. Rev. Lett. 96, 106802 (2006). \n[5] H. Zhang, C. -X. Liu, X. -L. Qi, X. Dai, Z. Fang, and S. -C. Zhang, Nature Physics 5, 438 \n(2009). \n[6] Y. Xia, D. Qian, D. Hsieh, L. Wray, A. Pal, H. Lin, A. Bansil, D. Grauer, Y. S. Hor, R. J. \nCava, and M. Z. Hasan, Nat Phys 5, 398 (2009). \n[7] D. Hsieh, Y. Xia, D. Qian, L. Wray, J. H. Dil, F. Meier, J. Osterwalder, L. Patthey, J. G. \nCheckelsky, N . P. Ong, A. V. Fedorov, H. Lin, A. Bansil, D. Grauer, Y. S. Hor, R. J. Cava, \nand M. Z. Hasan, Nature 460, 1101 (2009). \n[8] Y. L. Chen, J. G. Analytis, J. -H. Chu, Z. K. Liu, S. -K. Mo, X. L. Qi, H. J. Zhang, D. H. Lu, \nX. Dai, Z. Fang, S. C. Zhang, I. R. Fi sher, Z. Hussain, and Z. -X. Shen, Science 325, 178 \n(2009). \n[9] P. Roushan, J. Seo, C. V. Parker, Y. S. Hor, D. Hsieh, D. Qian, A. Richardella, M. Z. \nHasan, R. J. Cava, and A. Yazdani, Nature 460, 1106 (2009). \n[10] T. Zhang, P. Cheng, X. Chen, J. -F. Jia, X. Ma, K. He, L. Wang, H. Zhang, X. Dai, Z. Fang, \nX. Xie, and Q. -K. Xue, Phys. Rev. Lett. 103, 266803 (2009). \n[11] Z. Alpichshev, J. G. Analytis, J. -H. Chu, I. R. Fisher, Y. L. Chen, Z. X. Shen, A. Fang, and \nA. Kap itulnik, Phys. Rev. Lett. 104, 016401 (2010). \n[12] B. Zhou, H. -Z. Lu, R. -L. Chu, S. -Q. Shen, and Q. Niu, Phys. Rev. Lett. 101, 246807 \n(2008). \n[13] J. Linder, T. Yokoyama, and A. Sudbø, Phys. Rev. B 80, 205401 (2009). \n[14] C.-X. Liu, H. Zhang, B. Yan, X. -L. Qi, T. Frauenheim, X. Dai, Z. Fang, and S. -C. Zhang, \nPhys. Rev. B 81, 041307 (2010). \n[15] J. Li and K. Chang, Applied Physics Letters 95, 222110 (2009). \n[16] H.-Z. Lu, W. -Y. Shan, W. Yao, Q. Niu, and S. -Q. Shen, Phys. Rev. B 81, 115407 (2010). \n[17] W.-Y. Shan, H. -Z. Lu, and S. -Q. Shen, New Journal of Physics 12, 043048 (2010). \n[18] H. Li, L. Sheng, D. N. Sheng, and D. Y. Xing, Phys. Rev. B 82, 165104 (2010). \n[19] M. Kim, C. H. Kim, H. -S. Kim, and J. Ihm, PNAS 109, 671 (2012). \n[20] H. Li, L. Sheng, and D. Y. Xing, Phys. Rev. B 85, 045118 (2012). \n[21] G. Liu, G. Zhou, and Y. -H. Chen, Applied Physics Letters 101, 223109 (2012). \n[22] J. Chen, H. J. Qin, F. Yang, J. Liu, T. Guan, F. M. Qu, G. H. Zhang, J. R. Shi, X. C. Xie, C. \nL. Yang, K. H. Wu, Y. Q. Li, and L. Lu, Phys. Rev. Lett. 105, 176602 (2010). \n[23] J. G. Checkelsky, Y. S. Hor, R. J. Cava, and N. P. Ong, Phys. Rev. Lett. 106, 196801 \n(2011). \n[24] H. Steinberg, D. R. Gardner, Y. S. Lee, and P. Jarillo -Herrero, Nano Lett. 10, 5032 (2010). \n[25] D. Kong, Y. Chen, J. J. Cha, Q. Zhang, J. G. Analytis, K. Lai, Z. Liu, S. S. Hong, K. J. \nKoski, S. -K. Mo, Z. Hussain, I. R. Fisher, Z. -X. Shen, and Y. Cui, Nat Nano 6, 705 (2011). \n14 \n [26] D. Kim, S. Cho, N. P. Butch, P. Syers, K. Kirshenbaum, S. Adam, J. Pag lione, and M. S. \nFuhrer, Nat Phys 8, 459 (2012). \n[27] T. Zhang, N. Levy, J. Ha, Y. Kuk, and J. A. Stroscio, Phys. Rev. B 87, 115410 (2013). \n[28] Y. J. Song, A. F. Otte, V. Shvarts, Z. Zhao, Y. Kuk, S. R. Blankenship, A. Band, F. M. \nHess, and J. A. Stroscio, Review of Scientific Instruments 81, 121101 (2010). \n[29] Y. Jiang, Y. Y. Sun, M. Chen, Y. Wang, Z. Li, C. Song, K. He, L. Wang, X. Chen, Q. -K. \nXue, X. Ma, and S. B. Zhang, Phys. Rev. Lett. 108, 066809 (2012). \n[30] Y. Jiang, Y. Wang, M. Chen, Z. Li, C. Song, K. He, L. Wang, X. Chen, X. Ma, and Q. -K. \nXue, Phys. Rev. Lett. 108, 016401 (2012). \n[31] D. Galanakis and T. D. Stanescu, Phys. Rev. B 86, 195311 (2012). \n[32] E. McCann, Phys. Rev. B 74, 161403 (2006). \n[33] T. Ohta, A. Bostwick, T. Seyller, K. Horn, and E. Rotenberg, Science 313, 951 (2006). \n[34] C. Pauly, G. Bihlmayer, M. Liebmann, M. Grob, A. Georgi, D. Subramaniam, M. R. \nScholz, J. Sá nchez -Barriga, A. Varykhalov, S. Blügel, O. Rader, and M. Morgenstern, Phys. \nRev. B 86, 235106 (2012). \n (a)\n(b)(c)\nQW\n(d)(e)\n(f)\nFigure 1. Zhang et al.2 QL 3 QL(a) (b)\n(c) (d)\nFigure 2: Zhang et al.(a)\nVG = –100 V\nVG = 0 V\nVG = 30 V\nVG = 60 V\nVG = 100 VVG = –100 V\nVG = -50 V\nVG = 0 V\nVG = 50 V\nVG = 200 VVG = 100 VVG = 20 VVG = –20 V(d) (c) (b) 2 QL Sb2Te3\nGapd2I/d2V3 QL Sb2Te3Sb2Te3 Model DOS \nFigure 3. Zhang et al.γU = –0.07 eVγU = –0.077 eVγU = –0.084 eVγU = –0.091 eVγU = –0.099 eVγU = –0.106 eVγU = –0.113 eVγU = –0.12 eVEnergy\nDistanceVG > 0VG < 0\nVG = 0\nBottom Surface\nSb2Te3STO InterfaceTop Surface\nSb2Te3 Vacuum InterfaceC\nV(a)\n(e) (g) (f)(d) (c) (b)\nγUcγU = 0 γU = –0.1 eV γU = –0.3 eV γU = –0.18 eV\nVG<0 VG>0γU0Experiment\nFigure 4. Zhang et al." }, { "title": "1410.1425v1.Competing_Ultrafast_Energy_Relaxation_Pathways_in_Photoexcited_Graphene.pdf", "content": "arXiv:1410.1425v1 [cond-mat.mes-hall] 3 Oct 2014Competing Ultrafast Energy Relaxation Pathways in Photoex cited Graphene\nS.A. Jensen,1,2Z. Mics,1I. Ivanov,1H.S. Varol,1D. Turchinovich,1\nF.H.L. Koppens,3,∗M. Bonn,1,∗and K.J. Tielrooij3,∗\n1Max Planck Institute for Polymer Research, Ackermannweg 10 , 55128 Mainz, Germany\n2FOM Institute AMOLF, Amsterdam, Science Park 104, 1098 XG Am sterdam, Netherlands\n3ICFO - Institut de Ci´ encies Fot´ oniques, Mediterranean Te chnology Park, Castelldefels (Barcelona) 08860, Spain\nFor most optoelectronic applications of graphene a thoroug h understanding of the processes that\ngovern energy relaxation of photoexcited carriers is essen tial. The ultrafast energy relaxation in\ngraphene occurs through two competing pathways: carrier-c arrier scattering – creating an elevated\ncarrier temperature – and optical phonon emission. At prese nt, it is not clear what determines the\ndominating relaxation pathway. Here we reach a unifying pic ture of the ultrafast energy relaxation\nby investigating the terahertz photoconductivity, while v arying the Fermi energy, photon energy,\nand fluence over a wide range. We find that sufficiently low fluenc e (<∼4µJ/cm2) in conjunction\nwith sufficiently high Fermi energy ( >∼0.1 eV) gives rise to energy relaxation that is dominated\nby carrier-carrier scattering, which leads to efficient carr ier heating. Upon increasing the fluence\nor decreasing the Fermi energy, the carrier heating efficienc y decreases, presumably due to energy\nrelaxation that becomes increasingly dominated by phonon e mission. Carrier heating through\ncarrier-carrier scattering accounts for the negative phot oconductivity for doped graphene observed\nat terahertz frequencies. We present a simple model that rep roduces the data for a wide range of\nFermi levels and excitation energies, and allows us to quali tatively assess how the branching ratio\nbetween the two distinct relaxation pathways depends on exc itation fluence and Fermi energy.\nGraphene is a promising material for, amongst others,\nphoto-sensing and photovoltaic applications [1], owing to\nits broadband absorption [2, 3], its high carrier mobility\n[4, 5] and the ability to create a photovoltage from\nheated electrons or holes [6]. It furthermore uniquely\nallows for electrical control of the carrier density and\npolarity [4]. To establish the potential and limitations\nof graphene-based optoelectronic devices, a thorough\nunderstanding of the ultrafast (sub-picosecond) primary\nenergy relaxation dynamics of photoexcited carriers is\nessential. For undoped graphene (with Fermi energy\nEF≈0), ultrafast energy relaxation through interband\ncarrier-carrier scattering was predicted [7] and observed\n[8, 9] to lead to multiple electron-hole pair excitation.\nFor doped graphene (with Fermi energy |EF|>0),\nultrafast energy relaxation through carrier-carrier in-\nteraction also plays an important role, with intraband\nscattering leading to carrier heating [10–13]. In addition,\nthe ultrafast energy relaxation was ascribed to optical\nphonon emission [14, 15], which reduces the carrier\nheating efficiency. Closely related, for undoped graphene\nthe sign of the terahertz (THz) photoconductivity is\npositive (see e.g. [16–19]), whereas for intrinsically\ndoped graphene the sign is negative, meaning that\nphotoexcitation gives rise to an apparent decrease of\nconductivity [11, 16, 17, 20–22]. This negative photo-\nconductivity was attributed to stimulated THz emission\n[20], and to a reduction of the intrinsic conductivity by\nenhanced scattering with optical phonons [21, 22] or\nby carrier heating [11, 16, 17]. Since the experimental\n∗Electronicaddress: klaas-jan.tielrooij@icfo.es, bonn@ mpip-mainz.mpg.de, frank.koppens@icfo.esparameters of all these studies differ strongly, different\nconclusions were drawn regarding the dominating energy\nrelaxation pathway and the origin of the sign of the THz\nphotoconductivity.\nHere, we reach a unifying picture of the energy\nrelaxation of photoexcited carriers in graphene by ex-\nperimentally studying the ultrafast energy relaxation of\nphotoexcited carriers for a large parameter space, where\nwe vary the Fermi energy and the fluence. We compare\nthe data with a simple model of carrier heating, which\nquantitatively reproduces the frequency-dependent THz\nphotoconductivity for a large range of Fermi energies\nand fluences with a single free fit parameter. This\nfit parameter is the carrier heating efficiency, i.e. the\nfraction of absorbed energy from incident light that\nis transferred to the electron system. Although the\nTHz photoconductivity is not a direct probe of the\ncarrier temperature, and the model parameters carry\nsome uncertainty, this approach allows us to iden-\ntify qualitatively how the ultrafast energy relaxation\nprocesses depend on excitation fluence and Fermi energy.\nWe study the ultrafast carrier-energy relaxation using\ntime-resolved THz spectroscopy experiments, where an\nexcitation pulse with a wavelength of 400, 800 or 1500\nnm (corresponding to a photon energy of E0= 3.10,\n1.55 or 0.83 eV, respectively) excites electron-hole pairs,\nand a low-energy terahertz (THz, f= 0.3−2 THz)\nprobe pulse interrogates the sample, where it interacts\nwith mobile carriers. An optical delay line controls the\nrelative time delay between excitation and probe pulses,2\nmaking it possible to determine the photo-induced\nchange in conductivity (the photoconductivity ∆ σ) as a\nfunction of time with a time resolution of ∼150 fs [23].\nWe use two different samples: The first sample, where\nthe graphene has a controllable Fermi energy, contains\nChemical Vapor Deposition (CVD) grown graphene\nwith an area of 1 cm2transferred onto a substrate that\nconsists of doped silicon, covered by a 300 nm thick\nlayer of SiO 2. The silicon serves as the backgate for\nelectrical control of the Fermi energy (carrier doping) of\nthe graphene sheet from EF= 0.3 eV down to ∼0.06\neV (the width of the neutrality region is ∼0.06 eV, see\nSupp. Info). We use weakly doped silicon (resistivity\nof 10-20 Ωcm) and sufficiently low pump energy (0.83\neV) to ensure that the photo-induced signal from the\nsilicon in the substrate (without graphene) is minimized\n(See Supp. Info). Two silver-pasted electrical contacts\nto the graphene sheet enable resistance measurements\nto retrieve the Fermi energy through the capacitive\ncoupling of the backgate (See Supp. Info). The second\nsample contains CVD grown graphene with an area\nof a square inch transferred onto SiO 2with a fixed\nFermi level of <0.15 eV (see Supp. Info). We carefully\nmeasure the fluence Fas described in Ref. [24] and\nuse the experimentally determined optical absorption of\ngraphene on SiO 2using standard UV-Vis spectrometry\nto obtain the absorbed fluence Fabs, and from this the\nnumber of absorbed photons Nexc=Fabs/E0, which is\nequal to the number of primary excited carriers.\nTo investigate the ultrafast (sub-picosecond) energy\nrelaxation dynamics of photoexcited carriers, we first\nexamine the temporal evolution of the photoconduc-\ntivity. In Fig. 1a we show the dynamics for a range\nof Fermi energies using the sample with controllable\nFermi energy, and in Fig. 1b for a range of fluences\nusing the sample with fixed Fermi energy. We find that\nall traces, except the one with Fermi energy within\nthe neutrality region width ( EF<0.06 eV), show\nnegative photoconductivity with a sub-picosecond rise,\nfollowed by a picosecond decay. Similar pump-probe\ndynamics have been observed before [11, 16–22] and\ncan be understood as follows: During the rise of the\nsignal three main processes take place: (i)the creation\nof initial electron-hole pairs; and subsequent ultrafast\nenergy relaxation through two competing relaxation\nchannels, namely (ii)carrier-carrier scattering and\n(iii)optical phonon emission. Thus, the peak signal\ncorresponds to a ’hot state’ with an elevated carrier\ntemperature Teland/or more energy in optical phonons\n[8, 10–15]. During the subsequent picosecond decay,\nthe ’hot state’ cools down to the same state as before\nphotoexcitation. The fraction of absorbed energy that\n– after the initial ultrafast energy relaxation – ends up\nin the electron system or in the phonon system depends\non the timescales associated with carrier heating and\nphonon emission, respectively.The ultrafast energy relaxation takes place during the\nfirst few hundred femtoseconds after photoexcitation,\ni.e. during the rise of the conductivity change. Figs. 1c\nand 1d show the normalized photoconductivity signals\nfor this time window. We will discuss the evolution\nof the signal amplitude as a function of Fermi energy\nand fluence later. First we note that the rise dynamics\nexhibit an intriguing effect: upon decreasing the Fermi\nenergy (i.e. the density of intrinsic carriers Nint) or\nincreasing the fluence (i.e. the density of primary excited\ncarriersNexc) the signal peak is reached at increasingly\nlater times. To quantify these results, we describe the\ndynamics using two rise times and an exponential decay\ntime. The two rise times allow for part of the conductiv-\nity change to occur with the (fixed) experimental time\nresolution and part with a (free) slower time scale. We\nthen examine the effective rise time τrise, which is the\namplitude-weighted average of the two (see Methods).\nThe insets of Fig. 1c and Fig. 1d show τriseas a function\nofNintandNexc, respectively. Indeed, for decreasing\nNintand increasing Nexcthe effective rise time increases\nfrom below 200 fs (limited by the experimental time\nresolution) up to 400 fs (for EF<0.1 eV).\nThe slowing down of the ultrafast energy relaxation\nof photoexcited carriers with decreasing Fermi energy\nis consistent with energy relaxation through intraband\ncarrier-carrier scattering (see bottom right inset in Fig.\n1c). The microscopic picture of this scattering process\nis shown for two different Fermi energies in the top left\ninset of Fig. 1c. Photoexcited carriers relax by exchang-\ning energy with intrinsic conduction band carriers that\nthus heat up. The amount of energy that is exchanged\nbetween photoexcited carriers and intrinsic conduction\nband carriers in each intraband carrier-carrier scattering\nevent is ∼EF[10]. Therefore, if EFdecreases, more\nenergy-exchangeevents are required for the photoexcited\ncarriers to complete their energy relaxation cascade and\ntherefore the relaxation time will increase (see bottom\nright inset of Fig. 1c). If the energy relaxation through\ncarrier-carrier scattering would slow down in such a\nway that the relaxation rate becomes comparable to the\nrate of other relaxation channels (e.g. optical phonon\nemission), we expect these channels to start contributing\nto the overall energy relaxation. This would lead to a\ndecrease in the fraction of energy that is transferred\nto the electron system, i.e. a reduced carrier heating\nefficiency.\nWe quantify the fraction of absorbed energy that leads\nto carrier heating by comparing our THz photoconduc-\ntivity data with the results of a simple thermodynamic\nmodel. In short, carrier heating leads to a broader\ncarrier distribution (higher Tel), which – in combination\nwith an energy-dependent scattering time [25, 26] –\nleads to the photo-induced change in THz conductivity\n(see Methods). The carrier heating is governed by the\namount of absorbed energy Fabsand the electronic heat3\ncapacity of graphene, which for a degenerate electron\ngas is given by Cel=αTel[27]. Here α=2πEF\n3¯h2v2\nFk2\nB, with\n¯h,vFandkBthe reduced Planck constant, the Fermi\nvelocity and Boltzmann’s constant, respectively. The\npossibility of controlling the Fermi energy of graphene\nthus allows for tunability of the heat capacity (see\nFig. 2a), which in turn determines the ’hot’ carrier\ntemperature T′\nelthat the system reaches. Figure 2b\nshows that T′\nelis equivalently determined by both the\nfluence and the Fermi energy.\nTo calculate the ’hot’ carrier temperature we\nuse a basic numerical approach, which produces a\nheat capacity in the degenerate regime that corre-\nsponds well with the analytical heat capacity (see\nFig. 2a), while remaining valid for non-degenerate\nelectron temperatures ( kBTel> EF). The numerical\napproach is based on the concept that before pho-\ntoexcitation there is a known amount of energy in\nthe electronic system: E1=/integraltext∞\n0dǫD(ǫ)ǫF(EF,Tel);\nand a known number of carriers in the conduction\nband:Nint=/integraltext∞\n0dǫD(ǫ)F(EF,Tel), where D(ǫ) is the\nenergy-dependent density of states and F(EF,Tel) is\nthe Fermi-Dirac distribution that depends on Fermi\nenergyEFand carrier temperature Tel. Due to optical\nexcitation, an amount of energy Fabsis absorbed in the\ngraphene and a fraction ηof this energy ends up in\nthe electronic system trough intraband carrier-carrier\nscattering. After intraband heating is complete, the\nsystem is then described by the following set of equa-\ntions:E2=E1+ηFabs=/integraltext∞\n0dǫD(ǫ)ǫF(E′\nF,T′\nel), and\n(conserving the number of conduction band carriers)\nNint=/integraltext∞\n0dǫD(ǫ)F(E′\nF,T′\nel). Here E′\nFandT′\nelare the\nchemical potential and the carrier temperature in the\n’hot state’, respectively. Carrier heating thus alters the\ncarrierdistribution, where we find by numerically solving\nthe equations for E2andNintthat the carrier temper-\nature increases and the chemical potential decreases\nby photoexcitation (see inset of Fig. 2c and Methods).\nThe photo-induced increase of carrier temperature,\nand the associated decrease in chemical potential were\nexperimentally confirmed recently [13].\nThe hot carrier distribution, with E′\nFandT′\nelcalcu-\nlated using the carrier heating model, directly leads to\nnegative THz photoconductivity (see Methods), which\nscales linearly with carrier temperature up to ∼2000\nK and then shows some saturation behavior (Fig. 2c).\nWe test the validity of our carrier heating model by\ncomparing its predictions for the frequency-resolved\nphotoconductivity with our experimental results for\nthe sample with fixed Fermi energy. In Fig. 3a we\nshow this comparison for a fluence of ∼12µJ/cm2\n(pump wavelength 800 nm, Nexc= 1·1012absorbed\nphotons/cm2). We compare the data (at the time\ndelay that corresponds to the signal peak) with the\nmodel result for a ground state Fermi energy of 0.11\neV and a scattering time proportionality constant of200 fs/eV (extracted from the Raman spectrum, THz\nconductivity measurements on the same sample without\nphotoexcitation, and the saturation value of the THz\nphotoconductivity at high fluence; see also Supp. Info)\nand find good agreement with a heating efficiency of η=\n0.75. The small discrepancies between data and model\ncan be ascribed to artifacts that arise from the temporal\nchange of the photoconductivity during the interaction\nwith the THz pulse [28], although we largely avoid these\nby moving the optical pump delay line simultaneously\nwith the THz probe delay line.\nThe overall agreement between data and model shows\nthat the observed negative THz photoconductivity of\nintrinsically doped graphene [11, 16, 17, 20–22] can be\nfully reproduced by considering intraband carrier heat-\ning, which reduces the thermally averaged conductivity\nof the intrinsic carriers. Despite the simplicity of the\nmodel, it can also explain the experimental results in\nRef. [21] using their experimental parameters, as well\nas the results in Ref. [20], by letting the environmental\ngas change the Fermi energy. These observations lead us\nto conclude that, despite some uncertainty in the deter-\nmination of the Fermi energy and the scattering time,\nthe model is suitable for obtaining reliable qualitative\nindications on how the carrier heating efficiency depends\non the Fermi energy and the fluence.\nWe now examine the ’high fluence’ regime using\nNexc= 8·1012absorbed photons/cm2(a fluence of\n∼100µJ/cm2) in Fig. 3b. This corresponds to the\nregime where the energy relaxation is slower than\nthe experimental time resolution (see Fig. 1d). Here\nwe find that we can only describe the data with a\nsignificantly reduced carrier heating efficiency of η≈0.5\n(keeping the ground state Fermi energy and scattering\ntime proportionality constant the same as in the low\nfluence regime). Combined, these results show that at\nsufficiently low fluence, a large fraction of the absorbed\nenergy ends up in the electron system, i.e. the ultrafast\nenergy relaxation occurs through efficient (and fast)\ncarrier-carrier scattering. However, upon increasing\nthe fluence (i.e. the carrier temperature), the relative\namount of energy transferred to the electron system\ndecreases, which means that carrier-carrier scattering\nbecomes less efficient (and slower) and other relaxation\nprocesses start to contribute.\nTo determine in more detail how the carrier heating\nefficiency depends on fluence, we study the peak pho-\ntoconductivity of the sample with fixed Fermi energy\nfor a large range of excitation powers, for both 800 nm\nand 400 nm excitation. We show the photoconductivity\nat the peak (when the ultrafast energy relaxation is\ncomplete) in Fig. 4a-b, together with the results of the\ncarrier heating model for the same parameters as in Fig.\n3a and a frequency of 0.7 THz. For low fluences (up to\n∼4µJ/cm2, corresponding to Nexc∼0.3·1012absorbed4\nphotons/cm2, 800 nm excitation, ∼2% absorption) the\nexperimental data are in agreement with the heating\nmodel with a fixed heating efficiency of η= 1. We notice\nthat the calculated photoconductivity shows saturation\nbehavior with Nexceven for constant η. This is related\nto the nonlinear dependence of the THz photoconduc-\ntivity on carrier temperature (Fig. 2c). Interestingly,\nat fluences above ∼4µJ/cm2the experimental photo-\nconductivity starts saturating and the model is only\nin agreement for a heating efficiency that gradually\ndecreases to ∼50% for the highest fluences applied here\n(Fig. 4a). These observations suggest that once a certain\ncarrier temperature ( ∼4000 K, see inset Fig. 4b) is\nreached, the heating efficiency decreases. Interestingly,\nthe experimental data for excitation with 400 nm light\nstart deviating from the model (with efficient heating)\natNexc∼0.15·1012absorbed photons/cm2(Fig. 4b),\ninstead of ∼0.3·1012absorbed photons/cm2in the case\nof excitation with 800 nm light. This is because each\n400 nm photon has twice the energy of a 800 nm photon.\nThus, in both cases the carrier heating efficiency starts\ndecreasing around the same carrier temperature.\nWe now determine how the heating efficiency depends\non the Fermi energy by measuring the peak photo-\nconductivity for the sample with controllable Fermi\nenergy as a function of both excitation power and Fermi\nenergy. The combined results (for excitation at 1500\nnm) are represented in Fig. 5a, where we show the\npeak photoconductivity as a function of gate voltage\nfor five different excitation powers. In Fig. 5b we show\nthe peak photoconductivity as a function of Nexcfor\nthree distinct gate voltages, corresponding to Nint≈\n1, 2 and 3 ·1012carriers/cm2, and compare these to\nthe results of the heating model. We find that for a\ndoping of Nint= 3·1012carriers/cm2(EF∼0.2 eV)\nthe data is in good agreement with the carrier heating\nmodel, using a scattering time proportionality constant\nof∼50 fs/eV and a carrier heating efficiency of η=1.\nHowever, for the lowest Nint, which corresponds to a\nFermi energy of ∼0.1 eV, we find a carrier heating\nefficiency of ∼20%, using the same energy-dependent\nscattering time. These results are in good agreement\nwith the observed slowdown of the rise dynamics with\ndecreasing Fermi energy in Fig. 1b.\nComparing the data and the heating model leads to\nthe following physical picture of the ultrafast energy re-\nlaxation in graphene: Until a certain carrier temperature\nis reached ( ∼4000 K), the ultrafast energy relaxation is\ndominated by carrier-carrier scattering, which leads to\nefficient and fast ( <150 fs) carrier heating. Once this\ncarriertemperature is reached, the relaxationslowsdown\nand the carrier heating efficiency decreases, as ultrafast\nenergy relaxation occurs through additional pathways\ninvolving optical phonon emission [14]. The reduction\nin heating efficiency that follows from the macroscopic\nheating model can be explained using the microscopicpicture of intraband carrier-carrier scattering, as put\nforward in Refs. [10, 11]. At increased electron temper-\natures, the (quasi-equilibrium) Fermi energy decreases\n(see Methods), which means that the electronic heat ca-\npacity decreases. It furthermore implies that the amount\nof energy that is exchanged in intraband carrier-carrier\nscattering events ( ∼EF) decreases. Therefore, energy\nrelaxation of a photoexcited carrier requires an increas-\ning number of intraband carier-carrierscattering cascade\nsteps. Thus for an increasingcarriertemperature, energy\nrelaxation through intraband carrierheating slows down.\nThe physical picture of carrier-temperature dependent\nultrafast energy relaxation of photoexcited carriers in\ngraphene unites the conclusions of a large fraction of\nthe existing literature on this topic. For example in Ref.\n[14], with excitation in the ’high fluence’ regime ( ∼1014\nabsorbed photons/cm2at 800 nm), it was concluded\nfrom the experimentally measured carrier temperature\nthat only part of the absorbed light energy ends up\nin the electronic system, whereas the rest couples to\noptical phonons. An ultrafast optical pump-probe study\nemploying a fluence of 200 µJ/cm2[15], also in the ’high\nfluence’ regime, similarly demonstrated optical phonon\nmediated relaxation, in addition to carrier heating. By\nmeasuring in the ’low fluence’ regime (a few µJ’s), Ref.\n[11] concluded that the ultrafast energy relaxation was\ndominated by carrier heating. Furthermore, two very\nrecent optical pump - THz probe studies [16, 17] both\nused anexcited carrierdensityof Nexc≈1·1012absorbed\nphoton/cm2at 800 nm and ascribe their observed nega-\ntiveTHzphotoconductivity(partially)tocarrierheating.\nIn conclusion, we provide a unifying explanation of\nthe ultrafast energyrelaxationofphotoexcited carriersin\ngraphene. For sufficiently low excitation power and suf-\nficiently high Fermi energy, the relaxation is dominated\nby carrier-carrier scattering, which leads to efficient gen-\neration of hot carriers. This regime typically persists up\nto a fluence of ∼4µJ/cm2(orNexc= 0.3·1012absorbed\n800 nm photons/cm2) for a Fermi energy of 0.1 eV. For\nlargerFermi energy, a higher fluence can be used without\nsignificant reduction in heating efficiency. In the case of\nlower Fermi energy and/or a higher fluence, the heating\nefficiency will decrease due to slower intraband carrier-\ncarrier scattering and additional energy relaxation chan-\nnels involving optical phonon emission. This opens up\nthe possibility to control the pathway of ultrafast energy\nrelaxation, i.e. the ability to tune the efficiency of energy\ntransfer from the primary excited carriers to electronic\nheat or to alternative degrees of freedom, such as lattice\nheat. Such tunability is useful for future applications, for\ninstance in the field of photodetection, where hot carriers\nare the dominant source of photocurrent generation [6].\nFinally, we note that the terrestrial solar radiation (on\nthe order of a pJ/cm2during a 10 ps timescale) corre-\nsponds to the ’low fluence’ regime with efficient carrier\nheating, which is therefore the relevant process to con-5\nsider for photovoltaic applications.\nMETHODS\nPhotoconductivity from carrier heating model\nWe use a numerical model based on carrier heating\nto calculate the complex photoconductivity of photoex-\ncited graphene. The frequency-dependent conductivity\nof graphene is generally given by [26]\nσ(ω) =e2v2\nF\n2/integraldisplay∞\n0dǫD(ǫ)τscatter(ǫ)\n1−iωτscatter(ǫ)dF(EF,Tel)\ndǫ,\n(1)\nwithethe elementary charge, vFthe Fermi velocity,\nD(ǫ) =2ǫ\nπ¯h2v2\nFthe density of states, ¯ hthe reduced Planck\nconstant, τscatter(ǫ) the energy-dependent momentum\nscattering time and F(EF,Tel) the Fermi-Dirac distribu-\ntion that is determined by the Fermi level and the carrier\ntemperature. For unexcited graphene we use the steady\nstate Fermi level EFand the ambient temperature Tel.\nWe furthermore use a scattering time that is determined\nby charged impurity scattering and increases linearly\nwith energy ǫ[25, 26]. After electron-hole pair excitation\nof graphene, photoexcited carriers can interact with the\nintrinsic carriers, leading to intraband thermalization\n(see also main text): the carrier temperature increases\ntoT′\neland the chemical potential decreases to E′\nF. The\nreason for the decrease of the chemical potential is the\nlinear scaling of the D(ǫ) with energy: a broader carrier\ndistribution (higher carrier temperature) would lead\nto an increased number of carriers in the conduction\nband, if the Fermi energy would be kept constant.\nTherefore, a higher carrier temperature leads to a lower\nFermi energy, as confirmed experimentally in Ref. [13].\nThe conductivity of photoexcited graphene σ′(ω) then\nfollows from Eq. 1, using the ’hot Fermi level’ and ’hot\ncarrier temperature’ with the photoconductivity given\nby ∆σ(ω) =σ′(ω)−σ(ω).\nOur model takes into account the effect of intraband\ncarrierheatingonthephotoconductivitythroughthecar-\nrier distribution, which is sufficient to explain the ob-\nservednegative photoconductivity [11, 16, 17, 20–22] and\nthe dependence on excitation power and Fermi energy.\nThe model does not explicitly include the effect of energy\nrelaxation to optical phonons on the photoconductivity,\nas proposed in Ref. [22]. This is justified, because for the\ngraphene used here (with an impurity-scattering-limited\nmobility below 2500 cm2/Vs) the effect of phonons on\nthe conductivity is negligibly small at low fluences [11].\nWe note that for very high fluences ( Nexc>1012ab-\nsorbed photons/cm2) and for phonon-scattering-limited\ngraphene with mobilities >10,000 cm2/Vs [29], this ef-\nfect likely plays a role, in addition to carrier heating.\nFurthermore, a more advanced model could include de-viations from linear scaling between the scattering time\nand the carrier energy [30], as well as changes in the\nDrude weight. The latter effect occurs when the valence\nand conduction band electrons no longer have separate\nthermal distributions and likely plays an important role\naround the Dirac point, where it correctly produces pos-\nitive photoconductivity [17].\nRise dynamics\nWe describe the time-resolved photoconductity with a\nphenomenological model that includes a rise step and an\nexponential decay step with time τdecay. In the low exci-\ntation power/high Fermi energy regime (highly efficient\ncarrier heating), the rise occurs within our instrument\nresponse function. For high excitation power/low Fermi\nenergythe risecontainsa veryfast component and aslow\ncomponent (reduced carrier heating efficiency). There-\nfore we model the rise dynamics with a ’fast’ rise compo-\nnent that is fixed at a value equal to the pulse duration\nofτpulse= 120–150fs (depending on the excitation wave-\nlength) and a variable ’slow’ rise component τslowwith\na longer pulse duration, and describe the time-resolved\nphotoconductivity by\n−∆σ(t) =A·Conv(e−t/τdecay,τpulse) +\nB·Conv(e−t/τdecay,τslow),(2)\nwhereAandBare the amplitudes correspond-\ning to the ’fast’ and the ’slow’ rise components and\nConv(y(t),τ) means taking the convolution of y(t) with\na Gaussian pulse of width τ. The effective rise time is\nthen given by\nτrise=A+B\nA/τpulse+B/τslow. (3)\nIn the fit, the free parameters are A,B,τslowand\nτdecay. The resulting effective rise time gives an indi-\ncation of how fast the initial energy relaxation of pho-\ntoexcited carriers takes place.\nACKNOWLEDGEMENTS\nWe would like to thank Justin Song, Leonid Levitov,\nSebastien Nanot and Enrique C´ anovas for useful discus-\nsions. KJT thanks NWO for a Rubicon fellowship. FK\nacknowledges support by the Fundacio Cellex Barcelona,\nthe ERC Career integration grant 294056 (GRANOP),\nthe ERC starting grant 307806 (CarbonLight) and\nsupport by the E. C. under Graphene Flagship (contract\nno. CNECT-ICT-604391).6\n[1] Bonaccorso et al., Graphene Photonics and Optoelec-\ntronics.Nature Phot. 4, 611 (2010)\n[2] J.M. Dawlaty et al. Measurement of the optical absorp-\ntion spectra of epitaxial graphene from terahertz to visi-\nble.Appl. Phys. Lett. 93, 131905 (2008)\n[3] L. Ren et al., Terahertz and Infrared Spectroscopy of\nGated Large-Area Graphene. Nano Lett. 12, 3711 – 3715\n(2012)\n[4] K.S. Novoselov et al., Electric field effect in atomically\nthin carbon films. Science306666 – 669 (2004)\n[5] L. Wang et al., One-Dimensional Electrical Contact to a\nTwo-Dimensional Material. Science342614-617 (2013)\n[6] N.M. Gabor et al., Hot Carrier-Assisted Intrinsic Pho-\ntoresponse in Graphene. Science334, 648-652 (2011)\n[7] Winzer, T. , Knorr, A.& Mali´ c, E. Carrier Multiplicatio n\nin Graphene. Nano Lett. 10, 4839–4843 (2010).\n[8] D. Brida et al., Ultrafast collinear scattering and carr ier\nmultiplication ingraphene. Nature Comm. 4, 1987 (2013)\n[9] T. Pl¨ otzing et al. Nano Lett. Article ASAP ,DOI\n10.1021/nl502114w (2014)\n[10] J.C.W. Song et al., Photoexcited carrier dynamics and\nimpact-excitation cascade in graphene. Phys. Rev. B 87\n,155429 (2013)\n[11] K.J. Tielrooij et al., Photoexcitation cascade and mul ti-\nple hot-carrier generation in graphene. Nature Phys. 9,\n248-252 (2013)\n[12] J.C. Johannsen et al., Direct view on the ultrafast carr ier\ndynamicsingraphene. Phys. Rev. Lett. 11, 027403 (2013)\n[13] I. Gierz et al., Snapshots of non-equilibrium Dirac car -\nrier distributions in graphene. Nature Mat. 12, 1119-1124\n(2013)\n[14] C.H. Lui et al., Ultrafast photoluminscence from\ngraphene. Phys. Rev. Lett. 105, 127404 (2010)\n[15] M. Breusing et al., Ultrafast nonequilibrium carrier d y-\nnamics in a single graphene layer. Phys. Rev. B 83,\n153410 (2011)\n[16] S.-F. Shi et al., Controlling Graphene Ultrafast Hot Ca r-\nrier Response from Metal-like to Semiconductor-like by\nElectrostatic Gating. Nano Lett. 14, 1578 – 1582 (2014)\n[17] A.J. Frenzel et al., Semiconducting-to-metallic phot ocon-\nductivity crossover and temperature-dependent Drude\nweight in graphene. Phys. Rev. Lett. 113, 056602 (2014)\n[18] J.H. Strait et al., Very Slow Cooling Dynamics of Pho-\ntoexcited Carriers in Graphene Observed by Optical-\nPump Terahertz-Probe Spectroscopy. Nano Lett. 11,\n4902-4906 (2011)\n[19] P.A. George et al., Ultrafast Optical-Pump Terahertz-\nProbe Spectroscopy of the Carrier Relaxation and Re-\ncombination Dynamics in Epitaxial Graphene. Nano\nLett.84248-4251 (2008)\n[20] C.J. Dochertyetal., Extremesensitivityofgraphenep ho-\ntoconductivity to environmental gases. Nature Comm. 3,\n1228 (2012)\n[21] A.J. Frenzel et al., Observation of suppressed teraher tz\nabsorption in photoexcited graphene. Appl. Phys. Lett.\n102, 113111 (2013)\n[22] G. Jnawali et al., Observation of a Transient Decrease\nin Terahertz Conductivity of Single-Layer Graphene In-\nduced by Ultrafast Optical Excitation. Nano Lett. 13,\n524–530 (2013)\n[23] R. Ulbricht et al., Carrier dynamics in semiconductorsstudied with time-resolved terahertz spectroscopy. Rev.\nMod. Phys. 83, 543 (2011)\n[24] J.J.H. Pijpers et al., Assessment of carrier-multipli cation\nefficiency in bulk PbSe and PbS. Nature Phys. 5, 811\n(2009).\n[25] T. Ando, Screening Effect and Impurity Scattering in\nMonolayer Graphene. J. Phys. Soc. Jpn. 75, 074716\n(2006)\n[26] S. Das Sarma, S. Adam, E. H. Hwang, and E. Rossi,\nElectronic transport in two-dimensional graphene. Rev.\nMod. Phys. 83, 407 (2011)\n[27] C. Kittel, Introduction to Solid State Physics , Wiley and\nSons (2005)\n[28] E. Hendry et al., Interchain effects in the ultrafast pho to-\nphysics of a semiconducting polymer: THz time-domain\nspectroscopy of thin films and isolated chains in solution.\nPhys. Rev. B 71, 125201 (2005)\n[29] V. Perebeinos and P. Avouris, Inelastic scattering and\ncurrent saturation in graphene. Phys. Rev. B 81, 195442\n(2010)\n[30] E. H. Hwang, S. Adam, and S. Das Sarma, Carrier Trans-\nport in Two-Dimensional Graphene Layers. Phys. Rev.\nLett.98, 186806 (2007)7\nI. FIGURES\nTHz\n-0.5 0 0.501a\nc\nTime (ps)/c68/c115(norm.)THzTime (ps)/c45/c68/c115(e /h)2Decreasing Fermi Energy (eV) EF\ngatedrain\nsourceSiO2\nSigraphene\n-1 0 1 2 3\n/c116rise(ps)E (eV)F\n0.1 0.2\n0.20.5-10123\n-0.5 0 0.501b\ndIncreasing Fluence (10 /cm ) Nexc12 2\nTime (ps)/c68/c115(norm.)-1 0 1 2 3\nTime (ps)SiO\n2graphene\nN (10 /cm )exc12 2\n0 3 6 9\n0.10.304812\n/c45/c68/c115(e /h)2\n/c116rise(ps)0.267.6\n<0.060.21\nFIG. 1: Energy relaxation dynamics - dependence on Fermi energy and fluence. a)The time-resolved photoconductivity for\nnine different gate voltages ( V−VD= -70, -60, -50, -40, -30, -20, -15, -10 and 0 V), i.e. for decre asing Fermi energy EF\nor intrinsic carrier concentration Nint(sample with controllable Fermi energy). The photoconduct ivity is extracted from the\nmeasured changeinTHz transmission after photoexcitation with1500 nmlight (excitationpower corresponds to Nexc= 0.8·1012\nabsorbed photons/cm2), using the thin film approximation as in Ref. [11]. The inset shows the geometry of the sample, as\nexplained in the text. The photoconductivity is negative fo r all traces, except for the one closest to the Dirac point, wh ere\nit is positive. The solid lines are fits with an effective rise t imeτriseand an exponential decay time (see Methods). b)The\ntime-resolved photoconductivity for excitation with 800 n m light at six different excitation powers, corresponding to absorbed\nphoton densities of Nexc= 0.26, 0.50, 1.2, 2.5, 5.0 and 7.6 ·1012photons/cm2(sample with fixed Fermi energy). The geometry\nof the sample with a fixed Fermi energy of EF∼0.1 eV (see Supp. Info), is shown in the inset and explained in the text. The\nphotoconductivity is negative and increases for increasin g excitation power, as the carrier temperature increases. T he solid lines\nare fits with an effective rise time τriseand an exponential decay time (see Methods). c)The photoconductivity normalized\nto the peak of the signal (sample with controllable Fermi ene rgy), showing that decreasing the Fermi energy leads to slow er\nrise dynamics (i.e. slower energy relaxation). The fitted va lues forτriseas a function of Nintare shown in the inset, together\nwith the theoretically predicted energy relaxation time ba sed on Ref. [10] (scaled by a factor 1.8) and the experimental time\nresolution (horizontal dashed line). The schematics in the top left show the ultrafast energy relaxation of a photoexci ted carrier\nfor two different Fermi energies, according to the carrier he ating process described in Refs. [10, 11]. d)The photoconductivity\nnormalized to the peak of the signal (sample with fixed Fermi e nergy), showing that increasing the excitation power leads to\nslower rise dynamics. The fitted values for τriseas a function of Nexcare shown in the inset together with the experimental\ntime resolution (horizontal dashed line).8\nPump fluence ( J/cm ) /c1092Fermi energy (eV)a\nb\n0 0.5 1 1.5 2 2.5600 K900 K1200 K1500 K2200 K\nEF E’F\nTel= 300 K\nT’el= 2000 K\nT’el- (K)Tel2000 4000 6000 00THz\n/c45/c68/c115(a.u.)cC = /c97el el T\nHeating\nmodel\n0 0.1 0.2 0.300.20.4\nHeat capacity C ( J/m K)el/c1092\nFermi energy E (eV)F\n0.05\n0.15\n0.25\n50010002000\n150025003000T’ (K)el\nFIG. 2:Carrier heating model. a)The electronic heat capacity according to the analytical th eory (red line) and our numerical\nmodel (blue line), showing linear scaling with Fermi energy .b)Contour plot of the temperature of the carrier population,\nshowing thatdecreasing theFermi energy andincreasing the fluencehavean equivalenteffect onthe temperatureof theele ctrons\nafter photoexcitation and intraband thermalization. The c arrier temperature is calculated through a simple thermody namic\nmodel (see text). c)The calculated negative THz photoconductivity as a functio n of carrier temperature, showing a linear\ndependence up to ∼2000 K. The top left inset shows schematically the process of ultrafast energy relaxation of a photoexcited\ncarrier, which leads to an increase of the carrier temperatu re fromTeltoT′\nel. The inset on the bottom right shows the Fermi-\nDirac distribution of the conduction band carriers, where t he temperature increase leads to a decrease of the Fermi ener gy (to\nkeep the total number of conduction band carriers constant) .9\na\nb0.4 0.8 1.2 1.6/c68/c115(e /h)2\nReal\nFrequency (THz)Imag.\n-8-4048E = 1.55 eV0\n/c104= 0.75\nN = 110 /cmexc. 12 2\nE = 1.55 eV0\n/c104= 0.5\nN = 810 /cmexc. 12 2\n0.4 0.8 1.2 1.6/c68/c115(e /h)2\nFrequency (THz)-8-4048RealImag.\nFIG. 3: Comparison of experimental and theoretical photoconducti vity.a)The complex photoconductivity of the sample with\nfixed Fermi energy as a function of frequency for an excitatio n fluence corresponding to Nexc= 1·1012carriers/cm2, together\nwith the model result with a carrier heating efficiency of η= 0.75. The model (see also main text and Methods) describes\nthe photoconductivity after photoexcited carriers have th ermalized through intraband carrier-carrier scattering a nd directly\nyields the observed negative photoconductivity. b)The complex photoconductivity of the sample with fixed Fermi energy\nas a function of frequency for an excitation fluence correspo nding to Nexc= 8·1012carriers/cm2, together with the model\nresult with a carrier heating efficiency of η= 0.5. This shows that in this regime other energy relaxation channels contribute\nto the ultrafast energy relaxation. The insets in panels aandbschematically show the process of carrier heating through\nintraband scattering and the corresponding carrier distri butions for the ambient carrier temperature Tel(blue) and the elevated\ncarrier temperature T′\nelafter photoexcitation and carrier thermalization (red). T he error bars in panels aandbrepresent the\nexperimental (non-systematic) error bars (95 % confidence i nterval).10\na\nb0 5 10 15048/c104= 1/c104= 0.5\nN (10 /cm )exc12 2/c45/c68/c115(e /h)2\n0 0.4 0.8048\nN (10 /cm )exc12 2/c45/c68/c115(e /h)2\nE = 1.55 eV0E = 3.1 eV00 0.4 0.85000\n/c68T (K)E = 1.55 eV0\nN (10 /cm )exc12 20.1 1 10\n048\n/c104= 0.5\n/c104= 1N (10 /cm )exc12 2\n/c45/c68/c115(e /h)2\n0\nFIG. 4: Carrier heating efficiency - dependence on fluence. a)The peak photoconductivity of the sample with fixed Fermi\nenergy for a large range of excitation powers and an excitati on wavelength of 800 nm together with the model with η= 1\n(short dashed line) and a heating efficiency of η= 0.5 (long dashed line). The inset shows the same data on a log arithmic\nhorizontal scale. b)The peak photoconductivity of the sample with fixed Fermi ene rgy as a function of excitation power ( Nexc)\nfor excitation with 800 nm light (corresponding to an energy of the primary excited carrier of E0/2 = 0.75 eV) and 400 nm\nlight (E0/2 = 1.5 eV). The dashed lines correspond to the model with the s ame parameters (Fermi energy and scattering time\nproportionality constant) as in panel a, usingη= 1. The model describes the data well up to Nexc= 0.3·1012carriers/cm2for\n800 nm excitation and up to Nexc= 0.15·1012carriers/cm2for 400 nm excitation. The inset shows the carrier temperatu re\nthat is reached after excitation and thermalization by carr ier-carrier scattering, reaching ∼4000 K before heating becomes less\nefficient, for both excitation wavelengths. The symbol size o f the data in panels aandbrepresents the error.11\na\n/c45/c68/c115(e /h)2\nN (10 /cm )exc12 2bGate Voltage (V-V )D0 20 40 60-20246\n/c45/c68/c115(e /h)2\n0 2 4 60246\n/c104= 1\n/c104= 0.6\n/c104= 0.20 0.11 0.16 0.2Fermi energy (eV)\nEF~ 0.2 eV\nEF~ 0.17 eV\nEF~ 0.1 eVNexc= 5.7\nNexc= 3.8\nNexc= 1.9\nNexc= 0.24\nNexc= 0.14\nFIG. 5: Carrier heating efficiency – dependence on Fermi energy. a)The peak photoconductivity of the sample with\ncontrollable Fermi energy as a function of gate voltage for fi ve different excitation powers at excitation wavelength 150 0 nm.\nThe corresponding excitation density Nexcis given in units of 1012absorbed photons/cm2. At gate voltages away from the\nDirac point the photoconductivity is negative, while it cha nges sign upon approaching the Dirac point, in agreement wit h very\nrecent observations [16, 17]. b)The peak photoconductivity as a function of excitation powe r (Nexc) for three gate voltages,\ncorresponding to Nint= 1, 2 and 3 ·1012carriers/cm2(light to dark blue). The solid lines show the results of the m odel described\nin the text and the Methods section with a fixed heating efficien cy. For a decreasing number of intrinsic carriers (lower Fer mi\nenergy) the carrier heating efficiency decreases, in agreeme nt with the slowing down of the rise with decreasing Fermi ene rgy.\nThe symbol size of the data in panels aandbrepresents the error.12\nSUPPLEMENTARY MATERIAL\nA. Characterization of the sample with controllable Fermi e nergy\nIn order to translate the applied gate voltage Vginto the Fermi energy of the graphene sheet, we determine the\ncapacitive coupling of the weakly doped silicon backgate to the graph ene sheet. For this, we use a sample with the\nsame substrate as the sample used for the EF-dependent measurements, however with graphene shaped as a H all\nbar (see bottom right inset of Fig. S2a). We apply a current of I= 1µA in the x-direction and measure the Hall\nvoltageVHbetween two contacts in the y-direction. We measure the Hall voltage as a function of backgate v oltage\nfor az-directed magnetic field of both +0.6 T and -0.6 T, and use the differen ce between these two scans to account\nfor offsets (from device asymmetry, for example). From the Hall v oltage we extract the carrier density vs.backgate\nvoltage (see Fig. S1a) using\nNint=I·B\ne·VH, (S1)\nwithethe elementarycharge. This allowsus to extract the capacitanceo fthe backgate, wherewe find C= 6.3·10−5\nF/m2, which is somewhat smaller than the theoretical value for 300 nm oxid e:Cth= 11.6·10−5F/m2. The reason\nfor this could be the low density of carriers in the silicon or the large siz e of the graphene flake. We use the obtained\ncapacitive coupling to obtain the Fermi level corresponding to the a pplied gate voltages of the sample. It also allows\nus to obtain the device mobility, contact resistance and width of the neutrality point through the device resistance\nas a function of gate voltage. We find a contact resistance of 3.3 kΩ , a neutrality point width of ∆ = 58 meV, and a\n(lower bound) of the mobility of ∼300 cm2/Vs (see Fig. S1b). This is in reasonable agreement with the ∼700 cm2/Vs\nfound from the photoconductivity fits.\nGate Voltage (V )gNint(10 /cm12 2/c41\n-60 -40 -20 0 20 40 60-8-6-4-2\nx\nyz\n20 60 100 1400.411.6\nG (10 S/m)-4\nGate Voltage (V-V )Da b\nSample with controllable\nFermi energy\nC = 6.310 F/m. -5 2Hall bar sample\nC = 6.310 F/m. -5 2Fermi energy (eV)\n0.11 0.20 0.25 0.30\nFIG. S1: a)The number of intrinsic carriers as a function of gate voltag e, extracted from Hall measurements on a similar\ndevice as the one used in the Fermi-level dependent measurem ents, equipped with additional contacts in the Hall geometr y.b)\nThe device resistance as a function of ( Vg-VD) together with a fit using the determined capacitance, which yields the device\nmobility, contact resistance and neutrality region width.\nB. Characterization of the sample with fixed Fermi energy\nWe characterize the sample with fixed Fermi energy using Raman spe ctroscopy with pump wavelenth 532 nm (see\nan exemplary Raman trace in Fig. S2a). The width of the 2D peak of ∼33 cm−1shows that the graphene, grown by\nchemical vapor deposition, is predominantly monolayer. We also extr act the Fermi energy, by analyzing the G peak\nlocation and the ratio between the 2D and the G peak ( ∼1585 cm−1) and>2, respectively, for the trace shown in\nFig. S2). The Raman spectra of more than 100 traces on different lo cations on the sample yield an average G peak\nposition of ∼1585 cm−1, which indicates a Fermi energy of <0.15 eV [S1, S2]. To get the most realistic estimate for\nthe Fermi energy that corresponds to our optical pump - THz pro be experiment, the Raman spectra are taken under13\nPhoton counts (a.u.)\nRaman shift (cm )-11600 2000 2400 2800100020003000\n1560 1580 1600G peak:\n1585 cm-1\n2640 2680 27202D peak\nwidth:\n33 cm-1a b\n0.4 0.8 1.2 1.6 2048121620\nFrequency (THz)/c115(e /h)2Sample with fixed Fermi energy\n/c116/c47E\nEF\nF= 200 fs/eV\n= 0.155 eVSample with fixed Fermi energy\nFIG. S2: a)Raman spectrum of the sample with fixed Fermi energy. The inse ts show the G peak and the 2D peak. b)The\nsteady state conductivity spectrum of the sample with fixed F ermi energy\n(without photoexcitation), together with a Drude conducti vity fit (black solid line).\nvery similar environmental conditions (nitrogen flushing).\nWe further characterize the sample using THz conductivity measur ements, where we alternatively measure the\nsubstrate with and without graphene in the THz focus. This allows us to obtain the steady state conductivity\n(without photoexcitation). The results are shown in Fig. S2b toget her with a conductivity fit as explained in the\nMethods section. We obtain good agreement between data and mod el for a Fermi level of EF= 0.155 eV and a\nscattering time proportionality constant of 200 fs/eV, correspo nding to a mobility of ∼2300 cm2/Vs. We use this\nscattering time proportionality constant and a slightly lower Fermi e nergy (EF= 0.11 eV) to compare all optical\npump - terahertz probe data measured on this sample to the carrie r heating model. The slightly lower Fermi energy\nfollows from the photoconductivity at high fluence in Fig. 4a (given th e scattering time proportionality constant\nof 200 fs/eV), and could be due to spatial variation of the Fermi en ergy, photo-cleaning during the course of the\nexperiment, or due to a modified humidity during the experiment.\nC. Positive photoconductivity at the Dirac point and the Sil icon contribution\n/c68T/T (%)\nN (10 /cm )exc12 20 2 4 6-0.6-0.4-0.20\nGraphene,\nV = VDSilicon\nt = tpeak\n/c68T/T (%)\nGate voltage V-Vd0 20 40 60-0.6-0.4-0.20 a b\nSilicon\nt = tpeak\nFIG. S3: Pump-probe signal recorded at the pump-probe delay corresponding to the graphene signal peak, measured at the\nDirac point (blue) with negative change in transmission (po sitive conductivity) as a function of absorbed fluence (in th e\ngraphene sheet). The same measurement on the substrate with out graphene at the same pump-probe delay gives a quadratic\nsignal due to two photon absorption (red).\nWe use an excitation wavelength of 1500 nm for our pump-probe mea surements on the gated graphene sample to\navoid exciting electron-hole pairs in the silicon, which would obscure ou r signal. However, there is a small contribution\nof two-photon absorption that leads to THz photoconductivity in t he silicon, in addition to the photoconductivity of14\nthe graphene sheet. We find that for the lowest fluences the silicon signal is negligible. However due to the quadratic\nincrease with fluence, the silicon signal is not negligible at the highest fl uences, in cases where the Fermi energy,\nand therefore the graphene signal, is very low. In Fig. S3a we show t he pump probe signal that originates from the\nsubstrate without graphene and the pump probe signal for the su bstrate with graphene at very low Fermi energy. It\nis noteworthy to point out that the silicon contribution to the combin ed graphene-silicon signal is insignificant at gate\nvoltages far away from the Dirac point and that it does not change w ith gate voltage (see Fig. S3b). Only around\nthe Dirac point, and at the highest fluences, does part of the posit ive photoconductivity come from the graphene and\npart from the silicon. Importantly, at low fluence, the positive phot oconductivity that we observe stems completely\nfrom the graphene, showing the capability of tuning the Fermi ener gy close enough to the Dirac point to change the\nsign of the graphene THz photoconductivity. And even at the highe st employed fluence, the signal from graphene at\nthe Dirac point is still larger than the signal from the silicon substrat e. We corrected the fluence dependent results\npresented in Fig. 5 of the main paper for this small substrate contr ibution.\nSupplementary References\n[S1] H. Yan et al., Infrared spectroscopy of wafer-scale graphen e.ACS Nano 5, 9854–9860 (2011)\n[S2] A. Das et al., Monitoring dopants by Raman scattering in an electr ochemically top-gated graphene transistor.\nNature Nanotech. 3, 210–215 (2008)" }, { "title": "1810.13277v2.Many_body_filling_factor_dependent_renormalization_of_Fermi_velocity_in_graphene_in_strong_magnetic_field.pdf", "content": "arXiv:1810.13277v2 [cond-mat.mes-hall] 13 Feb 2019Many-body filling-factor dependent renormalization of Fer mi velocity in graphene\nin strong magnetic field\nAlexey A. Sokolik1,2and Yurii E. Lozovik1,2,3,∗\n1Institute for Spectroscopy, Russian Academy of Sciences, 1 42190 Troitsk, Moscow, Russia\n2National Research University Higher School of Economics, 1 09028 Moscow, Russia\n3Dukhov Research Institute of Automatics (VNIIA), 127055 Mo scow, Russia\nWe present the theory of many-body corrections to cyclotron transition energies in graphene in\nstrong magnetic field due to Coulomb interaction, considere d in terms of the renormalized Fermi\nvelocity. A particular emphasis is made on the recent experi ments where detailed dependencies of\nthis velocity on the Landau level filling factor for individu al transitions were measured. Taking\ninto account the many-body exchange, excitonic correction s and interaction screening in the static\nrandom-phase approximation, we successfully explained th e main features of the experimental data,\nin particular that the Fermi velocities have plateaus when t he 0th Landau level is partially filled and\nrapidly decrease at higher carrier densities due to enhance ment of the screening. We also explained\nthe features of the nonmonotonous filling-factor dependenc e of the Fermi velocity observed in the\nearlier cyclotron resonance experimentwith disordered gr aphene bytakingintoaccount thedisorder-\ninduced Landau level broadening.\nI. INTRODUCTION\nMassless Dirac electrons in single-layer graphene offer\nan opportunity to study condensed-matter counterparts\nof relativistic effects and to achieve new regimes in quan-\ntum many-bodysystems [1–3]. Low-energyelectronic ex-\ncitations in this material obey the Dirac equation and\nmove with the constant Fermi velocity vF≈106m/s. In\na strong perpendicular magnetic field B, quantization of\nan electron kinetic energy in graphene results in the rel-\nativistic Landau levels [4]\nEn= sgn(n)vF/radicalbig\n2|n|Be/planckover2pi1/c, n= 0,±1,±2,...(1)\nUnlike usual Landau levels for massive electrons, the rel-\nativistic ones are not equidistant En∝/radicalbig\n|n|, scale as a\nsquare root of magnetic field, En∝√\nB, and obey the\nelectron-holesymmetry, En=−E−n. Relativisticnature\nofgrapheneLandaulevelswasfirstconfirmedbythe half-\ninteger quantum Hall effect [2], and direct observations\nof these levels using the scanning tunneling spectroscopy\nhad followed (see the review of experiments in [5]).\nAnother way to study Landau levels in graphene is to\ninduce electron interlevel transitions by an electromag-\nnetic radiation, typically in the infrared range. The se-\nlection rules for photon absorption[6] require∆ |n|=±1,\nimplying the intraband −n−1→ −n,n→n+ 1, and\ninterband −n−1→n(which will be referred to as T−\nn+1)\nand−n→n+1 (referred to as T+\nn+1) transitions. The\ninterband transitions T±\nn+1, which are more widely stud-\nied, have the energies\nEn+1−E−n=vF/radicalbig\n2Be/planckover2pi1/c/parenleftbig√n+√\nn+1/parenrightbig\n(2)\nin the ideal picture of massless Dirac electrons (1) in the\nabsence of interaction and disorder.\n∗Electronic address: lozovik@isan.troitsk.ruIn a series of cyclotron resonance measurements,\nmainly on epitaxial graphene, transition energies in very\ngoodagreementwith Eq.(2) werereported(see[7, 8] and\nreferences therein). However, the other experiments [9–\n12] demonstrated deviations from Eq. (2) due to many-\nbody effects and, possibly, disorder. Similar deviations\nwere discovered in magneto-Raman scattering for both\ncyclotron T±\nn+1[13] and symmetric interband −n→n\n[14–16] transitions. Indeed, the Kohn’s theorem [17],\nwhichprotectscyclotronresonanceenergiesofusualmas-\nsiveelectronsagainstmany-bodyrenormalizations,isnot\napplicable to graphene [18–29]. The observed energies of\nT±\nn+1can be described by the counterpart of Eq. (2)\nΩ±\nn+1=v∗\nF/radicalbig\n2Be/planckover2pi1/c/parenleftbig√n+√\nn+1/parenrightbig\n(3)\nwith the bare Fermi velocity vFreplaced by the renor-\nmalized velocity v∗\nF. While the former one, vF, should be\nclose to 0.85×106m/s, as indicated by fitting theoreti-\ncalcalculationstovariousexperimentaldataongraphene\n(see, e.g., [29–32]), the latter one, v∗\nF, range from 106m/s\nto 1.4×106m/s depending on carrier density, magnetic\nfield and substrate material [9–16]. The existing the-\nory describes renormalization of Fermi velocity in mag-\nnetic field with reasonable accuracy in the Hartree-Fock\n[15, 22–24] and static random-phase [26, 29, 33] approx-\nimations.\nIn two very recent experiments [12, 13], the energies of\nthe T±\nn+1transitions were measured with high accuracy\nas functions of the Landau level filling factor ν, that may\nprovide an especially deep insight into the many-body\nphysics of graphene in magnetic field. Unlike graphene\nwithout magneticfield, where v∗\nFdivergeslogarithmically\nupon approach to the charge neutrality point [3, 31, 34],\nhere it saturates to a finite value at ν→0, and, in the\nmost cases, has even a broad plateau in the range −2<\nν <2.\nIn this article, we calculate the energies of the T±\nn+1\ntransitions as functions of the filling factor νwith taking2\ninto account many-body effects. Our approach, which\nis described in Sec. II and Appendices A, B, and C,\ntakes into account the screening of the Coulomb inter-\naction as one of the key points, that contrasts with the\nmost calculations on this subject [15, 18–25, 28] based on\ntheHartree-Fockapproximationwithunscreenedinterac-\ntion. The screening allowed us to describe experimental\ndata on both Landau levels [35] and interlevel transition\nenergies [29] earlier, and provides improved understand-\ning to the filling-factor dependence of the observed v∗\nFin\nthis work as well.\nIn Sec. III we analyze the electron-hole asymmetry of\ntransition energies and the presence of plateaus at −2<\nν <2, following from the propertiesofinteraction matrix\nelements. In Sec. IV we present the results of numerical\ncalculations, which reproduce the main features of the\nexperimental the vF(ν) dependencies from Refs. [12, 13]:\na) the plateausin v∗\nFat−2<ν <2 when the 0th Landau\nlevelispartiallyfilled, b) the rapiddecreaseof v∗\nFat|ν|>\n2 with increasing carrier density, c) the decrease of v∗\nFat\nν= const at increasing magnetic field. We have found\ngood agreement between the experiments and the theory\nusing the bare Fermi velocity vF= 0.85×106m/s and\nrealisticvaluesofthedielectricconstant ε. Theintraband\ntransitionsn→n+1 and−n−1→nwerealsoanalyzed,\nandwepredicttheV-shapeddependence oftheirenergies\nonν.\nAdditionally, we have considered the nonmonotonous\ndependence v∗\nF(ν) for the T±\n1transition observed in [11]\nwith the maximum at ν= 0 and minima at ν=±2.\nTaking into account a disorder-induced broadening of\nLandau levels, we have explained this dependence with\ngood accuracy in Sec. V. Our conclusions are presented\nin Sec. VI.\nII. THEORETICAL APPROACH\nDynamical conductivity of graphene can be calculated\nusing the Kubo formula [36]\nσαβ(q,ω) =1\n/planckover2pi1ωS∞/integraldisplay\n0dtei(ω+iδ)t∝an}bracketle{t[jα(q,t),jβ(−q,0)]∝an}bracketri}ht,(4)\nwherejα(q,t) is theα-axis projection of the Fourier\ncomponent of the current density operator jα(q) =\nevF/integraltext\ndrΨ+(r)σαΨ(r)e−iqrevolving in time in the\nHeisenberg representation, Ψ( r) is the two-component\nfield operator for Dirac electrons, Sis the system area,\nandδ→+0.\nDiagrammatic representation of the conductivity,\nshown in Fig. 1(a), allows its calculation in terms of the\ncurrent vertex matrix Γ β, which would be equal to σβin\nthe absenceofinteractionand disorder. To find it, we use\nthe mean field approximation, where the excitonic ladder\n[Fig. 1(b)] for the vertex Γ βand the one-loop self-energy\ncorrections [Fig. 1(c)] for the single-particle Green func-\ntionsGare taken into account. Using the interaction,(a) σαβ=\n(b) = +\n(c) = +\n(d) = +\nFIG. 1: (a) Diagrammatic relationship (C2) between the cur-\nrent Green function and the vertex. (b), (c) Equations for,\nrespectively, the vertex function and the electron Green fu nc-\ntion in the mean-field approximation. (d) Coulomb interac-\ntion screening in the random-phase approximation.\nwhich is statically screened in the random-phase approx-\nimation [Fig. 1(d)], greatly simplify the calculations. If\nwe additionally neglect the mixing of different pairs of\nelectron and hole Landau levels, appearing in the exci-\ntonic ladder (which was shown to be weak under typical\nconditions with using the screened interaction [29]), the\noptical conductivity σαβ(ω)≡σαβ(0,ω) is (see the de-\ntails of calculations in Appendix C):\nσαβ(ω) =ie2v2\nF\nω/summationdisplay\nn1n2fn2−fn1\n/planckover2pi1ω−Ωn1n2+iδ\n×Tr[Φn1n2(0)σα]Tr/bracketleftbig\nΦ+\nn1n2(0)σβ/bracketrightbig\n.(5)\nHerefnis the occupation number (0 /lessorequalslantfn/lessorequalslant1) of the\nnth Landau level, and the matrix Φ n1n2(0), which is de-\nfined by (C5) and (A2), determines the selection rules\n|n1|=|n2|±1 for eachn2→n1transition. The resonant\ntransition energy Ω n1n2, whereσαβhas a pole, consists\nof the difference between the bare Landau level energies\nEn1−En2, the difference between electron self-energies\nΣn1−Σn2, and the excitonic correction ∆ E(exc)\nn1,n2(see the\nsimilar formula in [25]):\nΩn1n2=En1−En2+Σn1−Σn2+∆E(exc)\nn1n2.(6)\nIn the mean field approximation, the self-energy\nΣn=−/summationdisplay\nn′fn′∝an}bracketle{tnn′|V|n′n∝an}bracketri}ht, (7)\nas shown in Appendix B, is a sum of the exchange matrix\nelements\n∝an}bracketle{tnn′|V|n′n∝an}bracketri}ht= 2δn0+δn′0−2l2\nH\n2π/integraldisplay\ndqV(q)\n×/vextendsingle/vextendsinglesnsn′φ|n|−1,|n′|−1(aq)+φ|n||n′|(aq)/vextendsingle/vextendsingle2(8)3\nof the screened Coulomb interaction V(q) between the\nnthandallfilled n′thLandaulevels[15,18,19],wherethe\nfunctionsφnkare defined in (A2), and aq≡ −l2\nH[ez×q].\nThe excitonic correction\n∆E(exc)\nn1n2=−(fn2−fn1)∝an}bracketle{tn1n2|V|n1n2∝an}bracketri}ht(9)\nis the direct interaction matrix element\n∝an}bracketle{tn1n2|V|n1n2∝an}bracketri}ht= 2δn10+δn20−2l2\nH\n2π/integraldisplay\ndqV(q)\n×/braceleftBig\nφ∗\n|n1|−1,|n1|−1(aq)+φ∗\n|n1||n1|(aq)/bracerightBig\n(10)\n×/braceleftbig\nφ|n2|−1,|n2|−1(aq)+φ|n2||n2|(aq)/bracerightbig\nwith the minus sign, weighted with the difference of oc-\ncupation numbers of the final and initial levels.\nThe dynamically screened interaction in the random-\nphase approximation is [see Fig. 1(d)]\nV(q,iω) =vq\n1−vqΠ(q,iω), (11)\nwherevq= 2πe2/εqis the bare Coulomb interaction\nweakened by the surrounding medium with the dielec-\ntric constant ε, and\nΠ(q,iω) =g/summationdisplay\nnn′Fnn′(q)fn−fn′\niω+En−En′,(12)\nisthe polarizability(ordensityresponsefunction) ofnon-\ninteracting Dirac electrons [4, 37–41]. Here\nFnn′(q) = 2δn0+δn′0−2\n×/vextendsingle/vextendsinglesnsn′φ|n|−1,|n′|−1(aq)+φ|n||n′|(aq)/vextendsingle/vextendsingle2(13)\nis the form-factor of Landau level wave functions and\ng= 4 is the degeneracy of electron states by valleys and\nspin. The staticallyscreened interaction V(q) is obtained\nfrom (11), (12) by taking iω= 0.\nIn our model, there are three mechanisms leading to\ndependence of Ω n1n2on the filling factor νvia the occu-\npation numbers\nfn=\n\n0, ifν≤4n−2,\n(ν−4n+2)/4,if 4n−2<ν <4n+2,\n1, ifν≥4n+2,(14)\ni.e.: through exchange energies (7), excitonic corrections\n(9), and polarizability (12).\nNotethatthesum(7)overthefilledLandaulevels n′in\nthe valence band diverges at n′→ −∞, so we impose the\ncutoffn′/greaterorequalslant−ncto obtain finite results. The physical rea-\nson of thus cutoff is a finite actual number of Landau lev-\nels in the valence band, which can be found from the to-\ntal electron density: nc= 2π/planckover2pi1c/√\n3a2eH≈39600/B[T],\nwherea≈2.46˚A [29, 35].-6 -4 -2 0 2 4 61.051.101.151.20\n-4 -2 0 2 4 6v¤\nFv¤\nF\n(106m/s) (106m/s)\nº º º º0!10!1¡1!0¡1!0\n¡1!2¡1!2¡2!1¡2!1(a) (b) \nFIG. 2: Renormalized Fermi velocities v∗\nFfor (a) the T±\n1and\n(b) theT±\n2transitions calculated with the screened interac-\ntion atvF= 0.85×106m/s,ε= 3.27,B= 8T. Solid lines\nshow the velocities found from the weighted transition ener -\ngies (20).\nIII. ELECTRON-HOLE ASYMMETRY AND\nPLATEAUS AT −2< ν <2\nThe selection rule |n1|=|n2| ±1 for the interband\nn2→n1transitions implies n1,n2=n+1,−n(the T+\nn+1\ntransition)or n1,n2=n,−n−1(theT−\nn+1transition). In\nthe idealized Dirac model without interactions, the ener-\ngies of these transitions (2) are equal. However this is no\nlongerthecasewhenexchangeself-energiesaretakeninto\naccount. Any nonzero doping ν∝ne}ationslash= 0 introduces an asym-\nmetrybetweenΩ+\nn+1andΩ−\nn+1, atleast,inthemean-field\napproximation. Looking at (6) and taking into account\nthat∝an}bracketle{tn+1,−n|V|n+1,−n∝an}bracketri}ht=∝an}bracketle{t−n−1,n|V|−n−1,n∝an}bracketri}ht,\nwe have:\nΩ+\nn+1−Ω−\nn+1= Σn+1+Σ−n−1−Σn−Σ−n\n+(fn+fn+1−f−n−f−n−1)∝an}bracketle{tn+1,−n|V|n+1,−n∝an}bracketri}ht.(15)\nThe electron-hole asymmetry in graphene, which is in-\nduced by the exchange interaction in the absence of mag-\nnetic field and is similar in scale to our case, was found\nin [42].\nThe first line of (15) is a contribution of exchange self-\nenergies to the asymmetry. Let us separate the occu-\npation numbers fn′=f(0)\nn′+ ∆fn′on those of undoped\ngraphenef(0)\nn′and the doping-induced part ∆ fn′, and de-\nfine Σ(0)\nn=−/summationtext\nn′f(0)\nn′∝an}bracketle{tnn′|V|n′n∝an}bracketri}ht. Using (8) and (A6),\nand neglecting a difference of small matrix elements at\nn′≈ −nc, we get Σ(0)\nn+1+Σ(0)\n−n−1−Σ(0)\nn−Σ(0)\n−n= 0. Thus\nthe exchange energy contribution to (15) arises only at\nnonzero doping ν∝ne}ationslash= 0.\nThe second line of (15) corresponding to excitonic ef-\nfects is nonzeroonlywhen either ±nth or±(n+1)thlevel\nis partially filled, i.e. at 4 n−2<|ν|<4n+ 6. Since\nthe polarizability (12) and hence the screened interaction\nV(q) are even functions of ν, both parts of (15) change\nsign atν→ −ν, so\nΩ+\nn+1(ν) = Ω−\nn+1(−ν). (16)4\nThis is illustrated in Fig. 2, where the typical calculated\nv∗\nFare shown as functions of ν.\nThe caseof n= 0is the specialone. The explicit struc-\nture of the wave functions (A2) imply the following re-\nlationships connecting the matrix elements of direct and\nexchange interaction (valid even for non-Coulomb poten-\ntials):\n∝an}bracketle{t±1,0|V|±1,0∝an}bracketri}ht+∝an}bracketle{t±1,0|V|0,±1∝an}bracketri}ht=∝an}bracketle{t00|V|00∝an}bracketri}ht.(17)\nIn result, the doping-induced changes of exchange and\nexcitonic parts of (6) due to f0cancel each other at\n−2<ν <2, when the 0th level is partially filled. Addi-\ntionally, the polarizability (12) and hence V(q) are also\nunchanged in this range of ν, thus we expect plateaus in\nboth Ω±\n1:\nΩ+\n1(ν) = Ω−\n1(−ν) = const at −2<ν <2,(18)\nas seen in Fig. 2(a). For n∝ne}ationslash= 0 this is no longer the\ncase, although variations of Ω±\nn+1(ν) at−2< ν <2 are\ntypically very small [see Fig. 2(b)].\nIn experiments, Ω+\nn+1and Ω−\nn+1can be separated by\nobserving cyclotron resonant absorption of light with op-\nposite circular polarizations. Using linear polarization,\none can observe a mixture of these transitions with rela-\ntiveintensities I+\nn+1=f−n−fn+1andI−\nn+1=fn−f−n−1,\nequal to occupation number differences in final and ini-\ntial states. Assuming that experimental apparatus does\nnot resolve the individual lines Ω+\nn+1and Ω−\nn+1, we will\ncalculate the weighted transition energy\n∝an}bracketle{tΩn+1∝an}bracketri}ht=Ω+\nn+1I+\nn+1+Ω−\nn+1I−\nn+1\nI+\nn+1+I−\nn+1(19)\nand compare it with the experiments in the next section.\nFrom the particle-hole symmetry relationship fn(−ν) =\n1−fn(ν) we see that ∝an}bracketle{tΩn+1∝an}bracketri}htis even function of ν. At\n−2< ν <2,∝an}bracketle{tΩn+1∝an}bracketri}htis linear (via f0) and at the same\ntime even function of ν, so\n∝an}bracketle{tΩn+1∝an}bracketri}ht= const at −2<ν <2, (20)\nas seen in Figs. 2(a,b). Thus our model predicts plateaus\nin all weighted transition energies ∝an}bracketle{tΩn+1∝an}bracketri}htat−2<ν <2.\nSimilar conclusions about existence of the electron-hole\nasymmetry (15) and the conjugation property (16) were\nmade in the recent theoretical work [25], which considers\ntransition energies in the Hartree-Fock approximation.\nIV. CALCULATION RESULTS\nFirst we compare our calculations of the renormalized\nFermi velocities\nv∗\nF=∝an}bracketle{tΩn+1∝an}bracketri}ht/radicalbig\n2Be/planckover2pi1/c/parenleftbig√n+√n+1/parenrightbig (21)with the data of Ref. [12] where Ω 1...Ω6as functions of\nνwere measured at three magnetic fields B= 5,8,11T.\nWe fit the experimental points in three approximations:\n1) Hartree-Fock approximation, where the unscreened\nCoulomb potential vqis used in all calculations.\n2) Static random-phase approximation, where the po-\ntentialV(q) is screened (11) with using the polarizability\nof noninteracting electron gas in magnetic field.\n3) Self-consistent screening approximation, where the\npolarizability is multiplied by vF/v∗\nFto take into ac-\ncount weakening of the screening caused by many-body\nincrease of the energy differences En′−Enin denomi-\nnators of (12). Similarly to our previous works [29, 35],\nthis semi-phenomenological model is aimed to achieve a\nself-consistency between many-body renormalizations of\ntransitionenergiesandscreening. Usingtheiterativepro-\ncedure, we take v∗\nF, obtained on each step, to renormalize\nthe screening when calculating new v∗\nFon the next step.\nAbout 5-6 iterations are usually sufficient to achieve a\nconvergence.\nCalculations in our approach depend only on two pa-\nrameters: the bare Fermi velocity vFand the dielectric\nconstant of a surrounding medium ε. In principle, both\nvFandεcan be treated as fitting parameters. However\nvariation of vFin the range (0 .8÷0.95)×106m/s allows\ntoachievealmostequallygoodagreementwith theexper-\nimental data at slightly different ε, so a simultaneous fit-\nting of both parameters does not provide reliable results.\nTherefore we choose a specific value vF= 0.85×106m/s,\nwhichwasconcludedtobethemostsuitableonebasedon\ntheoretical fits of several experimental data on graphene\nboth in presence [29, 30] and absence [31, 32] of mag-\nnetic field. After that, the optimal dielectric constant\nof the surrounding medium εis the only adjustable pa-\nrameter in our model, and we find it by performing the\nleast square fitting of the experimental points for all n\nandBsimultaneously. Nevertheless it should be kept in\nmind that our fitting results can slightly change quanti-\ntatively with different choice of vF(although qualitative\nconclusions will be the same), and it could be promis-\ning to implement a renormalization-groupscheme for the\nLandau level data on graphene where all unobservable\nvariables like vFcan be excluded from the model.\nThe first line of Table I shows the optimal εused to\nfit the cyclotron resonance data of Ref. [12] where high-\nmobility graphene samples were encapsulated from both\nsides in hexagonal boron nitride monolayers and placed\nTABLE I: Dielectric constants of surrounding medium ε,\nwhich provide the best least-square fittings of the experime n-\ntal data from Refs. [12, 13] at vF= 0.85×106m/s in the three\napproximations for the interaction listed in Sec. IV.\nUnscreened Screened Self-consistent\nExperiment interaction interaction screening\nRussell et al. [12] 7.72 3.27 4.36\nSonntag et al. [13] 5.50 1.05 2.555\n-8 -4 0 4 8 -8 -4 0 4 81.051.101.151.20\n-8 -4 0 4 8 -8 -4 0 4 8 -8 -4 0 4 8 -8 -4 0 4 8\n-8 -4 0 4 8 -8 -4 0 4 81.051.101.151.20\n-8 -4 0 4 8 -8 -4 0 4 8 -8 -4 0 4 8 -8 -4 0 4 8\n-8 -4 0 4 81.051.101.151.20\n-8 -4 0 4 8 -8 -4 0 4 8 -8 -4 0 4 8 -8 -4 0 4 8 -8 -4 0 4 8(a)B=5T (a)B=5T\n(b)B=8T (b)B=8T\n(c)B=11T (c)B=11Tº º\nº º\nº ºT1T1 T2T2 T3T3 T4T4 T5T5 T6T6\nT1T1 T2T2 T3T3 T4T4 T5T5 T6T6\nT1T1 T2T2 T3T3 T4T4 T5T5 T6T6v¤\nFv¤\nF\n(106m/s) (106m/s)v¤\nFv¤\nF\n(106m/s) (106m/s)v¤\nFv¤\nF\n(106m/s) (106m/s)\n2 4 6 800.010.020.030.04¢(106m/s) ¢(106m/s)(d)(d)\n\" \"\nFIG. 3: Renormalized Fermi velocities v∗\nFat (a)B= 5, (b) 8, and (c) 11T for the set of the Tn+1transitions ( −n→\nn+1/−n−1→n), taken from the experiment [12] (crosses), and calculated theoretically in the Hartree-Fock approximation\n(dashed lines), with taking into account the interaction sc reening (solid lines) and with the self-consistent screeni ng (dotted\nlines). The dielectric constants ε, used in each calculation, are listed in the first line of Tabl e I. Root mean square deviations\n(22) between the calculated and experimental v∗\nFare also shown (d) as functions of εin the three approximations.\non an oxidized silicon. Fig. 3 shows the experimental\npoints together with our calculations at these εin the\nthree approximations described above. The calculation\nwith the unscreened Coulomb interaction (Hartree-Fock\napproximation) demonstrates two significant drawbacks.\nFirst, the dielectric constant ε≈7.72 is unrealistically\nhigh, because in this approximation it should imitate the\ninteraction screening by Dirac electrons in graphene in\naddition to the screening by an external medium. Sec-\nond, thefalloffof v∗\nFat|ν|>2turnsouttobeinsufficient,\nbecause the increase of the screening strength (and, con-\nsequently, suppression of the upward renormalization of\nthe Fermi velocity) following the carrier density, is ab-\nsent here. For the T1transitions, the calculated v∗\nFeven\nincreases at |ν|>2, in contradiction with the experi-\nment, because the excitonic correction (9), which nor-\nmally decrease v∗\nF, become suppressed due to partial fill-\ning of 1th or −1th Landau level. The similar drawbacks\nof the Hartree-Fock approximations were mentioned in\nour previous works [29, 35].\nThe screening allows us to achieve much better agree-\nment with the experimental points at more realistic ε≈\n3.27, and the falloff of v∗\nFat|ν|>2 is reproduced very\nwell. The iterative calculations with the self-consistent\nscreening provide almost the same curves, but at some-\nwhat higher ε≈4.36. This distinction arises because the\nhigherεisneededtocompensatethescreeningweakeningcaused by an upward renormalization of energy denomi-\nnators in (12).\nOurcalculationswithtakingintoaccountthescreening\nare thus able to fit the data of Ref. [12] at three different\nBand for six resonances Tn+1simultaneously with the\nsingle adjustable parameter ε. We can explain both the\ndecrease of v∗\nFatν= 0 asBgets higher, the plateaus\nat|ν|<2, and the rapid falloff of v∗\nFat|ν|>2,n/greaterorequalslant1\ndue to increase of the screening strength. The exceptions\nare some inconsistencies of v∗\nFat specific resonances (T 2\nand T 6atB= 5T, T 5and T 6atB= 8T) and the\nlocal maxima at ν= 0 for the T 1transitions. Moreover,\nthe local minima of v∗\nFfor T2atν=±4, when the 1th\nor−1th Landau level is half-filled, which occur only at\nB= 8T and are absent in other fields, are not predicted\nby our approach.\nTo characterize an accuracy of our fitting, we present\nin Fig. 3(d) the root mean square deviation\n∆ =/radicaltp/radicalvertex/radicalvertex/radicalbtN/summationdisplay\ni=1[(v∗\nF)calc\ni−(v∗\nF)exp\ni]2\nN(22)\nof calculated renormalized Fermi velocities ( v∗\nF)calc\nifrom\nNexperimentalvalues( v∗\nF)exp\ni(hereit meansforallfields\nand all resonances at once, 260 points in total). As func-\ntions of the fitting parameter ε, ∆ reach rather sharp6\n-28 -24 -20 -16 -12 -8 -4 0 4 8 12 16 20 24 2811.11.21.3\nv¤\nFv¤\nF\n(106m/s) (106m/s)\nº ºB=8TB=8T\n1)21)22)32)33)43)44)54)55)65)6-2)-1 -2)-1-3)-2 -3)-2-4)-3 -4)-3-5)-4 -5)-4 -6)-5 -6)-5\nFIG. 4: Renormalized Fermi velocities v∗\nFfor the intraband n→n+1 and −n−1→ −ntransitions calculated at B= 8T in\nthe Hartree-Fock approximation (dashed lines) and with tak ing into account the interaction screening (solid lines). T he self-\nconsistent iterative calculations are not shown because th eir results are close to those with the non-self-consistent screening.\nThe dielectric constants εare taken from the first line of Table I.\nminima at optimal εin each approximation. The mini-\nmal ∆ about 0 .015×106m/s are comparable to the ex-\nperimental uncertainties of ( v∗\nF)exp\ni[12], so the fitting can\nbe considered to be sufficiently accurate.\nFor completeness of the analysis, we can also consider\nthe intraband transitions n→n+1 and −n−1→ −n.\nSeveral examples calculated in the conditions of the ex-\nperiment [12] are presented in Fig. 4. Each n→n+ 1\n(−n−1→ −n) transition exists in the range 4 n−2<\nν <4n+6 (−4n−6< ν <−4n+2) of the filling fac-\ntors,andthetransitionenergiesareminimalat ν= 4n+2\n(ν=−4n−2). These minima arecaused by the excitonic\ncorrection (9), which is maximal when the initial Landau\nlevel is completely filled, and the final level is completely\nempty. We can also note that v∗\nFagain decreases with\nincreasing the doping level due to enhancement of the\nscreening, while the Hartree-Fock approximation misses\nthis effect and greatly overestimates the variations of v∗\nF\nvs.ν. The electron-hole asymmetry for these transitions\nis negligible.\nAnother experiment we analyze is Ref. [13] where\ngrapheneis suspended 160nmaboveoxidizedsilicon, and\nthe filling-factor dependence of the T 2transition energy\nwas measured at B= 3T by observing its avoided cross-\ningwiththephononenergyinRamanspectrum. InFig.5\nweplottheresultsofourcalculationsforthistransitionin\nthethreeapproximationsatoptimal εlistedinthesecond\nline of Table I. We observe the same regularities as in the\npreviouscase. TheHartree-Fockapproximationsrequires\noverestimated εand cannot explain the rapid falloff of v∗\nF\nat|ν|>2. At|ν|>6 we see even slight increase of v∗\nF\ndue to suppression of the excitonic correction when the\n2nd or−2nd Landau level start to be partially filled.\nIn contrast, with taking into account the screening we\nobtain the realistic εfor graphene suspended above the\noxidized silicon, and the falloff is well reproduced. Nev-\nertheless, the experimental points demonstrate an addi-\ntional maximum at ν= 0. This is not described by our\napproach,whichpredictsplateausat |ν|<2, asdiscussed\nin Sec. III.\nThe root mean square deviations (22), calculated for31 experimental points, are shown in the inset to Fig. 5\nand demonstrate pronounced minima at the optimal ε.\nThe minimal ∆ about 0 .015×106m/s, achieved with the\nscreened interaction, are comparable to the experimental\nuncertainties (0 .01÷0.05)×106m/s [13].\nV. LANDAU LEVEL BROADENING\nOne more experiment where the filling-factor depen-\ndent transition energy was measured is Ref. [11]. In this\nearlier work, graphene layer lied directly on an oxidized\nsilicon substrate and carrier mobility was one-two orders\n-8 -6 -4 -2 0 2 4 6 81.201.251.301.351.40\nv¤\nFv¤\nF\n(106m/s) (106m/s)\nº ºT2; B= 3T T2; B= 3T\n0 2 4 600.010.020.030.04\n\" \"¢ (106cm/s ) ¢ (106cm/s )\nFIG. 5: Renormalized Fermi velocity v∗\nFfor theT2transition\n(−1→2/−2→1) atB= 3T, taken from the experiment\n[13] (squares) and calculated theoretically in theHartree -Fock\napproximation (dashed lines), with taking into account the\ninteraction screening (solid lines) and with the self-cons istent\nscreening (dotted lines). The dielectric constants ε, used in\neach approximation, are listed in the second line of Table I.\nInset shows root mean square deviations (22) between the\ncalculated and experimental v∗\nFas functions of εin the three\napproximations.7\nof magnitude lower than in the aforementioned works\n[12, 13] due to charged impurities in the substrate. The\nT1cyclotron resonance was studied at B= 18T and the\nunusual W-shaped form of the transition energy vs. ν\nwas found with the local maximum at ν= 0 and two\nminima at integer Landau level fillings ν=±2.\nTo explain these results, we need to take into account\ndisorder, because at mobilities of several thousands of\ncm2/V·s, reported in [11], the disorder-induced Landau\nlevel widths ∼20meV become comparable with the en-\nergy scalee2/εlHof Coulomb interaction effects. The\nmain mechanism of disorder effect on the transition en-\nergies is the following. Assume that Landau levels are\nbroadened giving rise to Gaussian mini-bands in the den-\nsity of states, as shown in Fig. 6. At partial filling of\neach level, its mini-band is partially filled, so the aver-\nage energy of the filled (empty) electron states is lower\n(higher) than the band center where the unperturbed\nLandau level energy would be located. As a result, the\naverage transition energy increases due to Landau level\nbroadening in addition to interaction effects when either\ninitial or final level is partially filled ( ν∝ne}ationslash=±2 in our\ncase). The similar effect was discussed in [44] for a two-\ndimensional gas of massive electrons in the framework of\nself-consistent Born approximation.\nTo describe this effect, we assume the Gaussian spec-\ntral density ρn(E) = (√\n2πΓn)−1exp[−(E−En)2/2Γ2\nn]\nfor eachnth partially filled broadened level. Integrat-\ning it up to the Fermi level µand assuming low tem-\nperature, we find the occupation number fn, and, us-\ning (14), we get the relationship between νandµ:\nν= 4n+2Φ([µ−En]/√\n2Γn), where Φ is the error func-\ntion. The disorder-induced correction ∝an}bracketle{t∆Ωn∝an}bracketri}htto the tran-\nsition energy is a difference between the average energies\n(relative to the band centers) of empty states on a fi-\nnal Landau level and filled states on an initial level. It\nshould be additionally weighted according to (20), when\n−2< ν <2 and thus both transitions T±\n1are present,\nresulting in:\n∝an}bracketle{t∆Ωn∝an}bracketri}ht=\n\n/radicalbigg\n2\nπΓ−1e−(µ−E−1)2\n2Γ2\n−1\n3+ν/2,if−6<ν <−2,\n/radicalbigg\n2\nπΓ0e−(µ−E0)2\n2Γ2\n0,if−2<ν <2,\n/radicalbigg\n2\nπΓ1e−(µ−E1)2\n2Γ2\n1\n3−ν/2,if 2<ν <6.(23)\nThis dependence has a maximum at ν= 0 and minima\natν=±2 in accordance with the experiment [11].\nAnother effect of the disorder is the presence of inter-\nlevel transitions when any nth Landau level is partially\nfilled, which provide an extra contribution to the screen-\ning. In the simplest approximation, they lead to the po-\nlarizability of the Thomas-Fermi kind\nΠTF\nn(q) =−gFnn(q)ρn(µ), (24)E E\n¡1¡10 01 1\nT¡\n1T¡\n1¡6<º<¡2 ¡6<º<¡2\nE E\n¡1¡10 01 1\nT¡\n1T¡\n1¡2<º<2 ¡2<º<2\nT+\n1T+\n1E E\n¡1¡10 01 12<º<62<º<6\nT+\n1T+\n1\nFIG. 6: Broadened Landau levels n= 0,±1 (not in scale)\nand cyclotron transitions between them when these levels ar e\npartially filled.\nwhich was used in [43] to study Landau level broadening\nin graphene.\nWe use the self-consistent Born approximation for a\npolarizability in magnetic field, which was originally de-\nvelopedin[45,46]foratwo-dimensionalelectrongaswith\nshort-rangeimpurities. Inourwork,weassumethedisor-\ndertobelong-ranged,becausethemainoriginofdisorder\nin graphene on a SiO 2substrate are long-range charged\nimpurities [47]. Introducing the mean square ∝an}bracketle{tU2∝an}bracketri}htof the\nslowly varying disorder potential U(r), we get the follow-\ning polarizability of disordered graphene (see the similar\nformulas in [45, 46] obtained by summing an impurity\nladder in a polarization loop):\nΠD(q,iω) =g/summationdisplay\nnn′Fnn′(q)\n×T/summationdisplay\nǫGD\nn′(iǫ+iω)GD\nn(iǫ)\n1−∝an}bracketle{tU2∝an}bracketri}htGD\nn′(iǫ+iω)GDn(iǫ),(25)\nwhereGD\nn(iǫ) =/integraltext\ndEρn(E)/(iω−E+µ) is the Green\nfunction of electron on the nth Landau level in the pres-\nence of disorder. Instead of a half-elliptic spectral den-\nsity [43], which is known to be an artefact of the self-\nconsistent Born approximation [48], we use, as above,\nthe Gaussian function ρn(E).\nTaking the static limit iω→0 and switching in (25)\nfrom the frequency summation to an integration along\nthe branch cut at Im( iǫ) = 0, we get in the limit T→0:\nΠD(q,0) =−g\nπ/summationdisplay\nnn′Fnn′(q)\n×0/integraldisplay\n−∞ImGD\nn′(z+iδ)GD\nn(z+iδ)\n1−∝an}bracketle{tU2∝an}bracketri}htGD\nn′(z+iδ)GDn(z+iδ).(26)\nThis polarizability consists of two physically distinct\nparts. The first one is the contribution of interlevel tran-\nsitions with n∝ne}ationslash=n′. It does not differ too much from\nthan in a clean system (12) if the widths of Landau lev-\nels Γnare much smaller than interlevel separations. The\nsecond one is the contribution of intralayer transitions\nn=n′arising when the nth layer is partially filled. Tak-\ning the disorder strength to be equal to the Landau level8\nwidth/radicalbig\n∝an}bracketle{tU2∝an}bracketri}ht= Γn, as follows from calculations of GD\nwith the long-range disorder, we get the static polariz-\nability of disordered graphene:\nΠD(q,0)≈Π(q,0)−gFnn(q)0/integraldisplay\n−∞Im[GD\nn(z+iδ)]2\n1−Γ2n[GDn(z+iδ)]2(27)\nand use it in the following calculations.\nFig. 7 shows the examples of static polarizabilities cal-\nculated at half-fillings of 0th and ±1th Landau levels.\nIn a clean graphene, Π( q,0)∝q2atq→0 since the\nsystem becomes insulating in magnetic field, and the\nonly source of the screening are gapped interlevel tran-\nsitions. Disorder makes ΠD(q,0) nonzero at q→0 due\nto intralevel transitions. The Thomas-Fermi approxima-\ntion, by taking into account only the latter, provides a\nwrong short-wavelength asymptotic of the polarizability\nΠTF\nn(q), which should tend to the polarizability of un-\ndoped graphene Π( q,0) =−gq/16/planckover2pi1vF[1].\nWe calculated the renormalized Fermi velocity, corre-\nsponding to the weighted energy (20) of the T1transi-\ntion with taking into account the correction (23) and the\nscreening (27) in the disordered system. For comparison,\nwecarriedout the samecalculationsforthe clean system,\nas did in the previous section. The results of the fitting\nof experimental points from Ref. [11] are shown in Fig. 8,\nand the calculation parameters are listed in Table II.\nFor the values of ε, we observe the same regularities\nas noted in the previous section. These values are close\nto those obtained in our earlier analysis [29] of cyclotron\nresonance data for graphene on SiO 2. However the most\ndrastic effects come from inclusion of disorder: while in\n0 1 2 3 4 500.511.500.511.52\n~¦(q;0) ~¦(q;0)\n(b)º=§4 (b)º=§4\nqlHqlH~¦(q;0) ~¦(q;0)(a)º=0 (a)º=0\nFIG. 7: Dimensionless static polarizability of graphene in\nmagnetic field ˜Π(q,0) =−(2πvFlH/g)Π(q,0), where lH=/radicalbig\n/planckover2pi1c/eH, calculated (a) when the 0th Landau level is half-\nfilled,ν= 0, (b) when the 1th or −1th level is half-filled,\nν=±4. Solid lines: clean graphene (12), dashed lines: disor-\ndered graphene (27), dotted lines: the Thomas-Fermi approx -\nimation (24). Calculation parameters are vF= 0.85×106m/s,\nB= 18T, Γ 0= Γ±1= 20meV.v¤\nFv¤\nF\n(106m/s) (106m/s)\nº º(a)(a)\n(b)(b)T1; B= 18T T1; B= 18T\nT1; B= 18T T1; B= 18T\n-6 -4 -2 0 2 4 61.051.101.151.051.101.151.20\nv¤\nFv¤\nF\n(106m/s) (106m/s)\nFIG. 8: Renormalized Fermi velocity v∗\nFfor theT1transition\n(0→1/−1→0) atB= 18T, taken from the experiment\n[11] (squares) and calculated theoretically for (a) clean a nd\n(b) disordered graphene. The calculations are carried out i n\nthe Hartree-Fock approximation (dashed lines), with takin g\ninto account the interaction screening (solid lines) and wi th\ntheself-consistent screening(dottedlines). Thedielect riccon-\nstantsεused in each calculation are listed in Table II.\nthecleansystem v∗\nFhasthe plateauat |ν|<2andremain\nthe same (or slightly increases due to suppression of the\nexcitonic correction) at |ν|>2, in the disordered system\nit hasthe parabolic-likemaximumat ν= 0andthe sharp\nminima at ν=±2, just as the experiment shows. The\nvalues of Landau level widths Γ nobtained via the fitting\nprocedure (15 −25meV) look realistic, since they are\nclose to typical widths of spectral lines observed in the\nsame experiment [11] and in other works on graphene\non a SiO 2substrate [49]. The minimal value of the root\nmean square deviation (22) is about 0 .009×106m/s in\nthis case.\nTABLE II: First two lines: dielectric constants of surround ing\nmedium ε, which provide the best least-square fittings of the\nexperimental data from Ref. [11] at vF= 0.85×106m/s in\nthe three approximations for the interaction listed in Sec. IV\nfor clean or disordered system. For disordered system, the\nwidths of 0th and ±1st Landau levels are also specified in the\nlast two lines.\nUnscreened Screened Self-consistent\nSystem interaction interaction screening\nClean 7.26 2.82 3.85\nDisordered 9.24 4.95 5.74\nΓ0(meV) 22 25 23\nΓ±1(meV) 12 19 179\nVI. CONCLUSIONS\nWe present detailed calculations of the inter-Landau\nlevel cyclotron transition energies in graphene in strong\nmagnetic fields taking into account Coulomb interaction\nbetweenmasslessDiracelectrons. Calculatingtheoptical\nconductivity and solvingthe vertex equation in the static\nrandom-phase approximation with the excitonic ladder,\nwe found the many-body correctionsto the transition en-\nergies coming from the self-energy and excitonic effects.\nWe show that the cyclotron transition lines can be split\nin doped graphene for opposite circular polarizations be-\ncause of the electron-hole asymmetry of exchange self-\nenergies, although this splitting may be unobservable if\nthese lines are sufficiently wide or either a linearly po-\nlarized or unpolarized light is used. By this reason, we\ncalculate the weighted transition energy for both polar-\nizations at once and convert it to the renormalized Fermi\nvelocityv∗\nFfor each transition.\nThe dependence of v∗\nFon the Landau level filling fac-\ntorνisanalyzed. In the mean-field approximation, v∗\nF(ν)\nhas a plateau at −2< ν <2 due to a partial cancela-\ntion of the self-energy and excitonic effects and rapidly\ndecreases at |ν|>2 due to enhancement of the screen-\ning. Our calculations, carried out with the bare Fermi\nvelocityvF= 0.85×106m/s and with the dielectric con-\nstant ofsurroundings ε, treated as an adjustable parame-\nter, showed good agreement with two recent experiments\n[12, 13] on high-mobility graphene samples, when the\nscreening by graphene electrons is taken into account.\nThe obtained phenomenological εdescribe the external\ndielectric screening not only by an underlayingsubstrate,\nbut also by adjacent hexagonal boron nitride layers. The\nHartree-Fock approximation, which neglects the density-\ndependent screening by graphene electrons, fails to ex-\nplain the observed rapid decrease of v∗\nFat|ν|>2.\nOur calculations for the intraband transitions n→n+\n1 and−n−1→ −npredict the V-like dependence v∗\nF(ν)\nwith the minima at, respectively, ν= 4n+ 2 andν=\n−4n−2 caused by the excitonic effects. Existence of\nthese minima can be verified experimentally, although\nan accurate detection of the interband transition lines\ncan be challenging (but possible [6]) due to their much\nlower energies: even for the the highest magnetic fields\n20-30 T these energies are below 100 meV.We also describe the data of the earlier cyclotron res-\nonance experiment [11] with graphene sample on SiO 2,\nwhere carriermobility is much lower. In this casewe take\nintoaccountlong-rangedisorder,whichbroadensLandau\nlevels and thus shifts the resonant energy upward when\ninitial or final level is partially filled, and induces the in-\nterleveltransitionscontributingtothescreening. Assum-\ning the Gaussian spectral density for the 0th and ±1th\nbroadened Landau levels, we achieved good agreement\nwith the experiment and explained the main features of\nthev∗\nF(ν) dependence: the parabolic-like maximum at\nν= 0 and the sharp minima at ν=±2.\nAs shown, the combined action of exchange interac-\ntion, excitonic effects, interaction screening and disorder\nshould be taken into account when considering graphene\nin strong magnetic field. Our approach takes into con-\nsideration these factors and thus allowed us to explain\nmain features of the filling-factor dependent experimen-\ntal data [11–13], which would be hardly possible within\nthe Hartree-Fock approximation [15, 18–25, 28] where\nthe screening and Landau level broadeningare neglected.\nHowever, some issues remain to be clarified. In partic-\nular, the mean-field approach does not describe the Λ-\nshaped maxima of v∗\nFatν= 0 observed in [12, 13] for\nT1transitions, the minima at ν=±4 observed for T 2\natB= 8T in [12], and a possible splitting of the T 1\ntransition line observed in [12]. All these features go be-\nyond the mean-field theory for massless Dirac electrons\nand can be attributed to some unaccounted role of disor-\nder, finite size effects, Moire superlattice potential from\nadjacent boron nitride layers [50], Landau level splitting\n[4, 51] or electron dynamics on a partially filled level [52].\nNotethatassumptionofasubstrate-inducedbandgapal-\nlowed to explain some features of the experimental data\nof[12] in the recent work[25], so a further analysisin this\ndirection with considering possible symmetry breakings\nandgapformationin asystemofDiracelectronstogether\nwith the interaction, screening and disorder seems to be\npromising.\nAcknowledgments\nThe work was supported by the grant No. 17-12-01393\nof the Russian Science Foundation.\nAppendix A: Electron wave functions\nSimilarlyto[4,38–40], wedescribesingle-particlestatesofmasslesse lectronsinmagneticfield Husingthesymmetric\ngaugeA=1\n2[H×r]. In the absence of a valley splitting or intervalley transitions, it is suffi cient to consider the\nelectrons only in the Kvalley, where the Dirac Hamiltonian is\nH=vF/parenleftBig\np−e\ncA/parenrightBig\n·σ=/planckover2pi1vF√\n2\nlH/parenleftBigg\n0a\na+0/parenrightBigg\n. (A1)\nHerelH=/radicalbig\n/planckover2pi1c/|e|His the magnetic length (we assume e <0 in this section), a=lHp−//planckover2pi1−ir−/2lHanda+=\nlHp+//planckover2pi1+ir+/2lHare, respectively, lowering and raising operators obeying the comm utation relation [ a,a+] = 1, and10\np±= (px±ipy)/√\n2,r±= (x±iy)/√\n2.\nIntroducing the complementary set of ladder operators [4] b=lHp+//planckover2pi1−ir+/2lH,b+=lHp−//planckover2pi1+ir−/2lH, which\nobey [b,b+] = 1 and commute with a,a+, we can construct the states of a two-dimensional oscillator |φnk∝an}bracketri}ht=\n(a+)n(b+)k|φ00∝an}bracketri}ht/√\nn!k! with the wave functions in polar coordinates:\nφnk(r,ϕ) =i|n−k|\n√\n2πlH/radicalBigg\nmin(n,k)!\nmax(n,k)!e−r2/4l2\nH/parenleftbiggr√\n2lH/parenrightbigg|n−k|\nei(n−k)ϕL|n−k|\nmin(n,k)/parenleftbiggr2\n2l2\nH/parenrightbigg\n, (A2)\nwhereLm\nn(x) are the associated Laguerre polynomials. The eigenfunctions of t he Hamiltonian (A1) are [4, 18–21, 37–\n39, 41]\nψnk= (√\n2)δn0−1/parenleftBigg\nsnφ|n|−1,k\nφ|n|,k/parenrightBigg\n, (A3)\nand eigenvalues are (1). Here n= 0,±1,±2,...is the Landau level number, k= 0,1,2,...is the guiding center index\nresponsible for Landau levels degeneracy, sn≡sign(n), and we assume that φnk= 0 ifnorkis negative.\nThe bare electron Green function in the Matsubara representatio nG(r,r′,τ) =−∝an}bracketle{tTτΨ(r,τ)Ψ+(r′,0)∝an}bracketri}htcan be\nconstructed from (A3):\nG0(r,r′,iǫ) =/summationdisplay\nnkψnk(r)ψ+\nnk(r′)\niǫ−En+µ, (A4)\nwhereµis the chemical potential; note G0is the (2 ×2) matrix in the sublattice space.\nUsing the table integral Eq. 2.20.16.10 from [53], we can present (A2) in Cartesian coordinates as\nφnk(x,y) =in−k\n√\n2π/integraldisplay+∞\n−∞dteityϕn(lHt+x/2lH)ϕk(lHt−x/2lH), (A5)\nwhereϕn(x) =e−x2/2Hn(x)//radicalbig\n2nn!√πare the dimensionless eigenfunctions of quantum one-dimensional h armonic\noscillator,Hn(x) are Hermite polynomials. Then, using (A5), and orthonormality and completeness of the basis\n{ϕn(x)}, a lot of useful transformation rules for φnkcan be obtained, for example, the summation formula\n∞/summationdisplay\nk=0φn1k(r1)φ∗\nn2k(r2) =ei(r1r2ez)/2l2\nH\n√\n2πlHφn1n2(r1−r2) (A6)\nand the form-factor of Landau level wave functions (see also [40])\n/integraldisplay\ndreiqrφ∗\nn1n2(r)φn3n4(r) = 2πl2\nHφ∗\nn1n3(aq)φn2n4(aq),aq≡ −l2\nH[ez×q]. (A7)\nAppendix B: Exchange self-energies\nExchange self-energy acquired by an electron in the state ψnkis given by the usual Fock expression\nΣexch\nnk=−/summationdisplay\nn′k′fn′k′/integraldisplay\ndr1dr2v(r1−r2)ψ+\nnk(r1)ψn′k′(r1)ψ+\nn′k′(r2)ψnk(r2). (B1)\nAfter the Fourier transform of the Coulomb interaction v(r) = (2π)−2/integraltext\ndqvqeiqrand using (A6)–(A7), we get\nΣexch\nnk=−/summationdisplay\nn′fn′∝an}bracketle{tnn′|v|n′n∝an}bracketri}ht (B2)\nwith the exchange matrix elements of Coulomb interaction defined as\n∝an}bracketle{tnn′|v|n′n∝an}bracketri}ht= 2δn0+δn′0−2l2\nH\n2π/integraldisplay\ndqvq/vextendsingle/vextendsinglesnsn′φ|n|−1,|n′|−1(aq)+φ|n||n′|(aq)/vextendsingle/vextendsingle2. (B3)11\nWe assumed that the occupation numbers do not depend on k′,fn′k′≡fn′, and the resulting Σexch\nnkturns out to\nbe also independent on k, so the Landau level degeneracy is preserved. By replacing vqwith the statically screened\ninteraction V(q), as depicted in Fig. 1(c), we get the screened exchange energy ( 7)–(8), and the bare electron Green\nfunction (A4) becomes “dressed” with the interaction and turns in to\nG(r,r′,iǫ) =/summationdisplay\nnkψnk(r)ψ+\nnk(r′)\niǫ−En−Σn+µ. (B4)\nAppendix C: Vertex equation\nIntroducing the Green function for currents Gj\nαβ(r,r′,τ) =−∝an}bracketle{tTτΨ+(r,τ)σαΨ(r,τ)Ψ+(r′,0)σβΨ(r′,0)∝an}bracketri}ht, we can\nwrite the conductivity (4) as\nσαβ(q,ω) =ie2v2\nF\n/planckover2pi1ωS/integraldisplay\ndrdr′e−iq(r−r′)Gj\nαβ(r,r′,/planckover2pi1ω+iδ), (C1)\nwhereGj\nαβcan be calculated, as shown in Fig. 1(a), from the (2 ×2) vertex matrix:\nGj\nαβ(r,r′,iω) =T/summationdisplay\nε/integraldisplay\ndr1dr2Tr[σαG(r,r1,iǫ+iω)Γβ(r1,r2,r′,iǫ,iω)G(r2,r,iǫ)]. (C2)\nHere the sum is taken over the fermionic Matsubara frequencies ǫ=πT(2n+1).\nThe vertex equation in the mean-field (or ladder) approximation, de picted in Fig. 1(b), is written analytically as\nΓβ(r1,r2,r′,iǫ,iω) =δ(r1−r′)δ(r2−r′)σβ−T/summationdisplay\nǫ′/integraldisplay\ndr′\n1dr′\n2V(r1−r2,iǫ−iǫ′)\n×G(r1,r′\n1,iǫ′+iω)Γβ(r′\n1,r′\n2,r′,iǫ′,iω)G(r′\n2,r2,iǫ′). (C3)\nTo solve it, we can use the basis of magnetoexcitonic states of Dirac electrons in the symmetric gauge, which were\ndescribed earlier in [38] in slightly different notation:\nΨPn1n2(r1,r2) =1\n2πeiP(r1+r2)/2+i(r1r2ez)/2l2\nHΦn1n2(r1−r2−aP). (C4)\nHerePis the conserved magnetic momentum of the electron-hole pair and\nΦn1n2(r) =√\n2δn10+δn20−2/parenleftBigg\nsn1sn2φ|n1|−1,|n2|−1(r)sn1φ|n1|−1,|n2|(r)\nsn2φ|n1|,|n2|−1(r)φ|n1|,|n2|(r)/parenrightBigg\n(C5)\nis the matrix wave function of relative motion of electron and hole writ ten in the basis of their sublattices A,B. Using\n(A5), the unitary transformations between the magnetoexciton ic states and the states (A3) of individual electron and\nhole can be derived:\nψn1k1(r1)ψ+\nn2k2(r2) =l2\nH/integraldisplay\ndPφ∗\nk1k2(aP)ΨPn1n2(r1,r2),ΨPn1n2(r1,r2) =l2\nH/summationdisplay\nk1k2φk1k2(aP)ψn1k1(r1)ψ+\nn2k2(r2).(C6)\nProjecting the vertex matrix Γ βon the magnetoexcitonic states\nΓβ,Pn1n2(r′,iǫ,iω) =/integraldisplay\ndr1dr2Tr/bracketleftbig\nΨ+\nPn1n2(r1,r2)Γβ(r1,r2,r′,iǫ,iω)/bracketrightbig\n(C7)\nand using (B4), (C6), we get (C3) in the electron-hole pair (or magn etoexcitonic) representation:\nΓβ,Pn1n2(r′,iǫ,iω) = Γ(0)\nβ,Pn1n2(r′)−T/summationdisplay\nǫ′n′\n1n′\n2/integraldisplay\ndP′∝an}bracketle{tΨPn1n2|V(iǫ−iǫ′)|ΨP′n′\n1n′\n2∝an}bracketri}htΓβ,P′n′\n1n′\n2(r′,iǫ′,iω)\n(iǫ′+iω−En′\n1−Σn′\n1+µ)(iǫ′−En′\n2−Σn′\n2+µ).(C8)\nHere the bare vertex is Γ(0)\nβ,Pn1n2(r′) = (e−iPr′/2π)Tr/bracketleftbig\nΦ+\nn1n2(aP)σβ/bracketrightbig\n.12\nTo solve Eq. (C8), we use the static approximation V(r,iǫ−iǫ′) =V(r) and neglect the mixing of different\nelectron-hole pairs in the ladder diagrams, assuming n′\n1=n1,n′\n2=n2. Therefore the vertex matrix turns out to be\nindependent on a relative energy of electron and hole ǫ:\nΓβ,Pn1n2(r′,iω) =e−iPr′\n2πTr/bracketleftbig\nΦ+\nn1n2(aP)σβ/bracketrightbig/braceleftbigg\n1+∝an}bracketle{tn1n2|VP|n1n2∝an}bracketri}htfn2−fn1\niω+En2+Σn2−En1−Σn1/bracerightbigg−1\n.(C9)\nThe average interaction energies of magnetoexcitons are ∝an}bracketle{tn1n2|VP|n1n2∝an}bracketri}ht=/integraltext\ndrV(r−aP)Tr/bracketleftbig\nΦ+\nn1n2(r)Φn1n2(r)/bracketrightbig\n,\ntheir counterparts in usual 2D electron gas were extensively stud ied earlier [54]. Making the Fourier transform\nV(r) = (2π)−2/integraltext\ndqV(q)eiqrand using (A7), we obtain\n∝an}bracketle{tn1n2|VP|n1n2∝an}bracketri}ht=l2\nH\n2π/integraldisplay\ndqV(q)e−iqaPTr/bracketleftbig\nΦ+\nn1n1(aq)/bracketrightbig\nTr[Φn2n2(aq)]. (C10)\nThe Green function for currents (C2) can by found using (B4), (C 4), (C6), (C7), and (C9):\nGj\nαβ(r,r′,iω) =/summationdisplay\nn1n2/integraldisplaydP\n(2π)2eiP(r−r′)Tr[Φn1n2(aP)σα]Tr/bracketleftbig\nΦ+\nn1n2(aP)σβ/bracketrightbig\n(fn2−fn1)\niω+En2+Σn2−En1−Σn1+(fn2−fn1)∝an}bracketle{tn1n2|VP|n1n2∝an}bracketri}ht.(C11)\nSubstituting it in (C1) and taking P= 0 for optical transitions in (C10), we finally obtain (5)–(10).\n[1] A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S.\nNovoselov, and A. K. Geim, The electronic properties of\ngraphene, Rev. Mod. Phys. 81, 109 (2009).\n[2] K.S.Novoselov, A.K.Geim, S.V.Morozov, D.Jiang, M.\nI. Katsnelson, I. V. Grigorieva, S. V. Dubonos, and A. A.\nFirsov, Two-dimensional gas of massless Dirac fermions\nin graphene, Nature 438, 197 (2005).\n[3] V. N. Kotov, B. Uchoa, V. M. Pereira, F. Guinea, and\nA. H. Castro Neto, Electron-Electron Interactions in\nGraphene: Current Status and Perspectives, Rev. Mod.\nPhys.84, 1067 (2012).\n[4] M. O. Goerbig, Electronic properties of graphene in a\nstrong magnetic field, Rev. Mod. Phys. 83, 1193 (2011).\n[5] L.-J. Yin, K.-K. Bai, W. Wang, S.-Y. Li, Y. Zhang, and\nL.He, LandauquantizationofDiracfermions ingraphene\nand its multilayers, Front. Phys. 12, 127208 (2017).\n[6] M.L. Sadowski, G. Martinez, M. Potemski, C. Bergert,\nand W.A. De Heer, Magneto-spectroscopy of epitaxial\ngraphene, Int. J. Mod. Phys. B 21, 1145 (2007).\n[7] M. Orlita and M. Potemski, Dirac electronic states in\ngraphene systems: optical spectroscopy studies, Semi-\ncond. Sci. Technol. 25, 063001 (2010).\n[8] D. N. Basov, M. M. Fogler, A. Lanzara, F. Wang, and Y.\nZhang, Colloquium: Graphene spectroscopy, Rev. Mod.\nPhys.86, 959 (2014).\n[9] Z. Jiang, E. A. Henriksen, L. C. Tung, Y.-J. Wang, M.\nE. Schwartz, M. Y. Han, P. Kim, and H. L. Stormer,\nInfraredspectroscopy ofLandaulevels ofgraphene, Phys.\nRev. Lett. 98, 197403 (2007).\n[10] Z. Jiang, E. A. Henriksen, P. Cadden-Zimansky, L.-C.\nTung, Y.-J. Wang, P. Kim, and H. L. Stormer, Cyclotron\nresonance near the charge neutrality point of graphene,\nAIP Conf. Proc. 1399, 773 (2011).\n[11] E. A. Henriksen, P. Cadden-Zimansky, Z. Jiang, Z. Q.\nLi, L.-C. Tung, M. E. Schwartz, M. Takita, Y.-J. Wang,P. Kim, and H. L. Stormer, Interaction-induced shift of\nthe cyclotron resonance of graphene using infrared spec-\ntroscopy, Phys. Rev. Lett. 104, 067404 (2010).\n[12] B. J. Russell, B. Zhou, T. Taniguchi, K. Watanabe, and\nE. A. Henriksen, Many-Particle Effects in the Cyclotron\nResonance of Encapsulated Monolayer Graphene, Phys.\nRev. Lett. 120, 047401 (2018).\n[13] J. Sonntag, S. Reichardt, L. Wirtz, B. Beschoten, M. I.\nKatsnelson, F. Libisch, C. Stampfer, Impact of Many-\nBody Effects on Landau Levels in Graphene, Phys. Rev.\nLett.120, 187701 (2018).\n[14] S. Berciaud, M. Potemski, andC. Faugeras, Probingelec -\ntronic excitations inmono- topentalayergraphene bymi-\ncro magneto-Raman spectroscopy, Nano Lett. 14, 4548\n(2014).\n[15] C. Faugeras, S. Berciaud, P. Leszczynski, Y. Henni, K.\nNogajewski, M. Orlita, T. Taniguchi, K. Watanabe, C.\nForsythe, P. Kim, R. Jalil, A. K. Geim, D. M. Basko,\nand M. Potemski, Landau level spectroscopy of electron-\nelectron interactions in graphene, Phys. Rev. Lett. 114,\n126804 (2015).\n[16] C. Faugeras, M. Orlita, and M. Potemski, Raman scat-\ntering of graphene-based systems in high magnetic fields,\nJ. Raman Spectrosc. 49, 146 (2018).\n[17] W. Kohn, Cycloron resonance and de Haas-van Alphen\noscillations of an interacting electron gas, Phys. Rev.\n123, 1242 (1961).\n[18] A. Iyengar, J. Wang, H. A. Fertig, and L. Brey, Excita-\ntions from filled Landau levels in graphene, Phys. Rev.\nB75, 125430 (2007).\n[19] Yu. A. Bychkov and G. Martinez, Magnetoplasmon ex-\ncitations in graphene for filling factors ν/lessorequalslant6, Phys. Rev.\nB77, 125417 (2008).\n[20] R. Rold´ an, J.-N. Fuchs, and M. O. Goerbig, Spin-flip\nexcitations, spinwaves, andmagnetoexcitons ingraphene13\nLandau levels at integer filling factors, Phys. Rev. B 82,\n205418 (2010).\n[21] Y. Barlas, W.-C. Lee, K. Nomura, and A. H. MacDonald,\nRenormalized Landau levels and particle-hole symmetry\nin graphene, Int. J. Mod. Phys. B 23, 2634 (2009).\n[22] K. Shizuya, Many-body corrections to cyclotron reso-\nnance in monolayer and bilayer graphene, Phys. Rev. B\n81, 075407 (2010).\n[23] K. Shizuya, Many-body corrections to cyclotron reso-\nnance in graphene, J. Phys.: Conf. Ser. 334, 012046\n(2011).\n[24] L. A. Chizhova, J. Burgdorfer, and F. Libisch, Magneto-\noptical response of graphene: Probing substrate interac-\ntions, Phys. Rev. B 92, 125411 (2015).\n[25] K. Shizuya, Many-body effects on Landau-level spectra\nand cyclotron resonance in graphene, Phys. Rev. B 98,\n115419 (2018).\n[26] E. V. Gorbar, V. P. Gusynin, V. A. Miransky, and I. A.\nShovkovy,Coulomb interaction and magnetic catalysis in\nthe quantum Hall effect in graphene, Phys. Scr. T146,\n014018 (2012).\n[27] N. Menezes, V. S. Alves, and C. M. Smith, The influence\nof a weak magnetic field in the Renormalization-Group\nfunctionsof(2+1)-dimensional Diracsystems, Eur.Phys.\nJ. B89, 271 (2016).\n[28] K. Shizuya, Direct-exchange duality of the Coulomb in-\nteraction and collective excitations in graphene in a mag-\nnetic field, Int. J. Mod. Phys. B 31, 1750176 (2017).\n[29] A. A. Sokolik, A. D. Zabolotskiy, and Yu. E. Lozovik,\nMany-bodyeffectsofCoulombinteractiononLandaulev-\nels in graphene, Phys. Rev. B 95, 125402 (2017).\n[30] G. L. Yu, R. Jalil, B. Belle, A. S. Mayorov, P. Blake, F.\nSchedin, S. V. Morozov, L. A. Ponomarenko, F. Chiap-\npini, S. Wiedmann, U. Zeitler, M. I. Katsnelson, A. K.\nGeim, K. S. Novoselov, and D. C. Elias, Interaction phe-\nnomena in graphene seen through quantum capacitance,\nProc. Natl. Acad. Sci. USA 110, 3282 (2013).\n[31] D. C. Elias, R. V. Gorbachev, A. S. Mayorov, S. V. Mo-\nrozov, A. A. Zhukov, P. Blake, L. A. Ponomarenko, I. V.\nGrigorieva, K. S. Novoselov, F. Guinea, and A. K. Geim,\nNature Phys. 7, 701 (2011).\n[32] L. Yang, J. Deslippe, C.-H. Park, M. L. Cohen, and S.\nG. Louie, Excitonic Effects on the Optical Response of\nGraphene and Bilayer Graphene, Phys. Rev. Lett. 103,\n186802 (2009).\n[33] I. A. Shovkovy and L. Xia, Generalized Landau level rep-\nresentation: Effect of static screening in the quantum\nHall effect in graphene, Phys. Rev. B 93, 035454 (2016).\n[34] J. Gonz´ alez, F. Guinea, and M. A. H. Vozmediano, Non-\nFermi liquid behavior of electrons in the half-filled honey-\ncomb lattice (A renormalization group approach), Nucl.\nPhys. B 424, 595 (1994).\n[35] A. A. Sokolik and Yu. E. Lozovik, Many-body renor-\nmalization of Landau levels in graphene due to screened\nCoulomb interaction, Phys. Rev. B 97, 075416 (2018).\n[36] V. P. Gusynin, S. G. Sharapov, and J. P. Carbotte, AC\nconductivity of graphene: from tight-binding model to\n2+1-dimensional quantum electrodynamics, Int. J. Mod.\nPhys. B 21, 4611 (2007).\n[37] R. Rold´ an, J.-N. Fuchs, and M. O. Goerbig, Collec-\ntive modes of doped graphene and a standard two-\ndimensional electron gas in a strong magnetic field:\nLinear magnetoplasmons versus magnetoexcitons, Phys.\nRev. B80, 085408 (2009).[38] Yu. E. Lozovik and A. A. Sokolik, Influence of Landau\nlevel mixing on the properties of elementary excitations\ningrapheneinstrongmagnetic field, Nanoscale Res.Lett.\n7, 134 (2012).\n[39] R. Rold´ an, M. O. Goerbig, and J.-N. Fuchs, The mag-\nnetic field particle-hole excitation spectrum in doped\ngraphene andin astandard two-dimensional electron gas,\nSemicond. Sci. Technol. 25, 034005 (2010).\n[40] P. K. Pyatkovskiy and V. P. Gusynin, Dynamical polar-\nization of graphene in a magnetic field, Phys. Rev. B 83,\n075422 (2011).\n[41] G. Gumbs, A. Balassis, D. Dahal, and M. L. Glasser,\nThermal smearing and screening in a strong magnetic\nfield for Dirac materials in comparison with the two di-\nmensional electron liquid, Eur. Phys. J. B 89, 234 (2016).\n[42] A. Kretinin, G. L. Yu, R. Jalil, Y. Cao, F. Withers,\nA. Mishchenko, M. I. Katsnelson, K. S. Novoselov, A.\nK. Geim, and F. Guinea, Quantum capacitance mea-\nsurements of electron-hole asymmetry and next-nearest-\nneighbor hopping in graphene, Phys. Rev. B 88, 165427\n(2013).\n[43] C. H. Yang, F. M. Peeters, and W. Xu, Landau-\nlevel broadening due to electron-impurity interaction in\ngraphene in strong magnetic fields, Phys. Rev. B 82,\n075401 (2010).\n[44] T. Ando, Theory of Cyclotron Resonance Lineshape in\na Two-Dimensional Electron System, J. Phys. Soc. Jpn.\n38, 989 (1975).\n[45] T. AndoandY.Uemura, ScreeningEffect inaDisordered\nElectron System I. General Consideration of Dielectric\nFunction, J. Phys. Soc. Jpn. 30, 632 (1971).\n[46] T. Ando and Y. Uemura, Theory of Oscillatory gfactor\nin a MOS Inversion Layer under Strong Magnetic Fields,\nJ. Phys. Soc. Jpn. 37, 1044 (1974).\n[47] S. Das Sarma, S. Adam, E. H. Hwang, and E. Rossi,\nElectronic transport in two-dimensional graphene, Rev.\nMod. Phys. 83, 407 (2011).\n[48] A. H. MacDonald, H. C. A. Oji, and K. L. Liu, Ther-\nmodynamic properties of an interacting two-dimensional\nelectron gas in a strong magnetic field, Phys. Rev. B 34,\n2681 (1986).\n[49] A. Luican, G. Li, and E. Y. Andrei, Quantized Landau\nlevel spectrum and its density dependence in graphene,\nPhys. Rev. B 83, 041405(R) (2011).\n[50] Z.-G. Chen, Z. Shi, W. Yang, X. Lu, Y. Lai, H. Yan,\nF. Wang, G. Zhang, and Z. Li, Observation of an in-\ntrinsic bandgap and Landau level renormalization in\ngraphene/boron-nitride heterostructures, Nature Com-\nmun.5, 4461 (2014).\n[51] Y. J. Song, A. F. Otte, Y. Kuk, Y. Hu, D. B. Torrance, P.\nN. First, W. A. de Heer, H. Min, S. Adam, M. D. Stiles,\nA. H. MacDonald, and J. A. Stroscio, High-resolution\ntunnelling spectroscopy of a graphene quartet, Nature\n467, 185 (2010).\n[52] D. M. Basko, P. Leszczynski, C. Faugeras, J. Binder, A.\nA. L. Nicolet, P. Kossacki, M. Orlita, and M. Potem-\nski, Multiple magneto-phonon resonances in graphene,\n2D Mater. 3, 015004 (2016).\n[53] A. P. Prudnikov, Yu. A. Brychkov, and O. I. Marichev,\nIntegrals and Series. Volume 2. Special Functions, Gor-\ndon and Breach Science Publishers, 1986.\n[54] L. V. Butov, C. W. Lai, D. S. Chemla, Yu. E. Lozovik, K.\nL. Campman, and A. C. Gossard, Observation of Mag-\nnetically Induced Effective-Mass Enhancement of Quasi-14\n2D Excitons, Phys. Rev. Lett. 87, 216804 (2001)." }, { "title": "2205.01349v1.Observation_of_the_noise_driven_thermalization_of_the_Fermi_Pasta_Ulam_Tsingou_recurrence_in_optical_fibers.pdf", "content": "Observation of the noise-driven thermalization of the Fermi-Pasta-Ulam-Tsingou\nrecurrence in optical \fbers\nGuillaume Vanderhaegen,1Pascal Szriftgiser,1Alexandre Kudlinski,1\nMatteo Conforti,1Andrea Armaroli,1and Arnaud Mussot1\n1University of Lille, CNRS, UMR 8523-PhLAM-Physique des Lasers Atomes et Mol\u0013 ecules, F-59000 Lille, France\n(Dated: May 4, 2022)\nWe report the observation of the thermalization of the Fermi-Pasta-Ulam-Tsingou recurrence\nprocess in optical \fbers. We show the transition from a reversible regime to an irreversible one,\nrevealing a spectrally thermalized state. To do so, we actively compensate the \fber loss to make the\nobservation of several recurrences possible. We inject into the \fber a combination of three coherent\ncontinuous waves, which we call Fourier modes, and a random noise. We enhance the noise-driven\nmodulation instability process against the coherent one by boosting the input noise power level to\nspeed up the evolution to the thermalization. The distributions of the Fourier modes power along\nthe \fber length are recorded thanks to a multi-heterodyne time-domain re\rectometer. At low input\nnoise levels, we observe up to four recurrences. Whereas, at higher noise levels, the Fourier modes\nfade into the noise-driven modulation instability spectrum revealing that the process reached an\nirreversible thermalized state.\nI. INTRODUCTION\nThe study of the Fermi Pasta Ulam Tsingou (FPUT)\nparadox in the 50s [1] marked the birth of numerical sim-\nulations and of nonlinear physics [2]. These numerical ex-\nperiments aimed at studing the energy transfers between\nthe eigenmodes of a chain of oscillators. A linear coupling\nmakes the excited mode preserving its energy, while a\nnonlinear coupling enables energy exchanges. Fermi and\nco-workers were expecting an equidistribution of the en-\nergy into all the modes. But, against all odds, the system\nexhibited a reversible behavior, and, after a certain time,\nreturned to its initial state.\nThis recursive energy transfer is rather general and\ncan be observed in many other \felds of physics, where a\nweak nonlinear coupling exists between the system eigen-\nmodes. It is now conventionally denoted as FPUT re-\ncurrence process. Such behaviors were especially inves-\ntigated in focusing cubic media through the nonlinear\nstage of modulation instability (MI), as demonstrated\nin hydrodynamics [3, 4], bulk crystals optics [5], planar\nwaveguides [6], \fber optics [7{14] and also magnetic feed-\nback rings [15]. A recurrent regime is obtained by trig-\ngering these systems with a weak coherent seed aside the\npump located inside the MI gain band. During the prop-\nagation, the pump transfers energy to this input signal\nand to the symmetric idler wave generated by four wave\nmixing, the Fourier modes of the systems. Higher order\nFourier modes, also called harmonics, are then generated\nby four-wave mixing process up to a speci\fc distance be-\nfore the energy transfers reversed back and the system\nreturns to the initial excitation state [7{14].\nA major question was to \fgure out if these systems\nbeyond a single recurrence period could reach a thermal-\nized state, as \frstly expected by Fermi, Pasta, Ulam and\nTsingou. The limited performances of their MANIAC\ncomputer [2] didn't allow to pursue the study until this\nirreversible \fnal state. Indeed, the thermalization is along-process of energy equipartition between the eigen-\nmodes of the mechanical oscillators requiring lengthy cal-\nculations. Fortunately, the computational power of mod-\nern computers enabled the achievement of these complex\ncalculations. Hence, they revealed that after several re-\ncurrences, pseudo-recurrences appear before the system\neventually reaches a thermalized state [16{19]. By anal-\nogy, in \fber optics, the irreversible evolution towards an\nenergy distribution among the whole spectrum (not only\nharmonics of the signal, but any spectral components in-\ncluding noisy ones) after the recurrences breakdown, was\ncalled thermalization too [20]. In that case, the noise-\ndriven MI process enters into competition with the seeded\none, which becomes eventually the dominant e\u000bect [20].\nAside the FPUT recurrence process, this competition\nwas deeply investigated in the context of supercontinuum\ngeneration and rogue waves because MI is the triggering\nmechanism of these complex nonlinear processes [21{28].\nA coherent seed was used to either increase supercontin-\nuum coherence or inhibit rogue wave formation [29, 30].\nWhile the nonlinear stage of MI through the prism of\nFPUT recurrences in either noise-driven systems [25, 26]\nor coherently seeded ones [7{13] had already been stud-\nied, their competition and the route to a thermalized\nstate had never been reported experimentally. Indeed, for\na while, the recurrence number was limited to a single cy-\ncle [7{10, 12] or two periods [11, 13] due to the \fber loss.\nIndeed dissipation, while weak in optical \fbers, dramat-\nically increases the period of recurrence. This prevents\nthe study of the thermalization process that is expected\nto appear after many periods [16{18]. Recently, it was\npossible to overcome this issue in \fber optics thanks to\nan active loss compensation scheme through Raman am-\npli\fcation [11, 14, 31] or the use of a recirculating loop\n[26].\nIn this work, we investigate the competition between\nnoise-driven and seeded MI processes in the framework\nof multiple FPUT recurrences to report the observation\nof the thermalization process. As in Refs. [11, 14, 31],arXiv:2205.01349v1 [nlin.PS] 3 May 20222\nwe compensate almost perfectly the loss through Raman\nampli\fcation to get several recurrences periods and we\naccelerate the route to thermalization by increasing the\nnoise power level at the \fber input. It enables to \fnely\ncontrol the balance between both MI processes to clearly\nobserve the transition from a perfectly recurrent regime\nto an irreversible spectrally thermalized state. The pa-\nper is organized as follows. First, we provide numerical\nsimulations to illustrate the noise and coherently driven\nMI processes and an example of competition between\nboth. Second, we introduce our experimental setup,\nmade of a multi-heterodyne optical time domain re\rec-\ntometer (HODTR) system combined with an active loss\ncompensation scheme. Finally, the experimental results\nare displayed, compared to the NLSE numerics.\nII. NUMERICAL SIMULATION OF\nSPONTANEOUS AND SEEDED MI\nNumerical simulations rely on the integration of the\nnonlinear Schr odinger equation (NLSE):\n@E\n@z=\u0000\f2\n2@2E\nT2+i\rjEj2E (1)\nwhereEis the electrical \feld, zis the distance along\nthe \fber and Tis the retarded time of the frame moving\nat the group velocity. \f2is the group velocity dispersion\nwith\f2=\u000021 ps2km\u00001and\ris the nonlinear coe\u000ecient\nwith\r= 1:3 W\u00001km\u00001. This equation is solved numer-\nically using the split-step Fourier method [32]. The evo-\nlution of the noise-driven MI process as a function of the\n\fber length is illustrated in Fig. 1 with a typical example.\nThe input spectrum is shown in Fig. 1. (a). A monochro-\nmatic pump is surrounded by a white noise \roor with a\npower spectral density PSD =\u0000112 dBm/Hz. In all the\nnumerical simulation, the real and the imaginary part of\nthe input noise spectrum at each frequency bin are sam-\nples of two independent Gaussian random variables with\nzero mean and unit variance. The noise \roor is then ad-\njusted at the desired level by multiplying the amplitude\nof the noise bins with a constant factor.\nThe evolution of the spectrum along the propagation\ndistance is plotted in Fig. 1. (b) for a single shot of noisy\ninitial condition. At the beginning (up to 5 km), only the\nspectral components within the MI gain band (the cuto\u000b\nfrequency is !C=p\n4\rPP=j\f2j[32]) are ampli\fed. Then\nthe spectrum broadens due to four-wave mixing (FWM)\nbetween the unstable frequencies. Looking at the input\ntime pro\fle (Fig. 1. (c)) and its evolution along the \fber\nlength (Fig. 1. (d)), we notice the irregular appearance\nof high power pulses. We can even distinguish solitonic\nstructures similar to the Akhmediev breather (AB), the\nKuznetsov-Ma soliton (KM) and the Peregrine soliton\n(PS) as previously reported in [24{26]. To highlight the\ncontinuous spectral broadening during the propagation,\n00.51Power (W)\nSSE-100-500Rel. power\n|g12|\n0 5 10 15 20\nDistance (km)\n-40-30-20-1001020(a)\n00.20.40.6PP (W)(c)\n1234567Relative power (dBm)\nPower (W)\n(e)\n02468(50 dB/div)\n-150 -100 -50 0 50 100\nTime (ps)150\n-300 -200 -100 0 100 200\nFrequency (GHz)3000\n5\n10\n15\n20Distance (km)0\n5\n10\n15\n20Distance (km)(b) (d)FIG. 1. Numerical simulations of the dynamic of spontaneous\nMI. (a) Spectrum at the \fber input, and (b) power spectral\ndistribution over the whole \fber length. (c) Power time pro-\n\fles at the \fber input, and (d) temporal power distribution\nalong the \fber length. (e) Shannon spectral entropy SSE and\npower of the initial pump excitation as function of the \fber\ndistance. Parameters: group velocity dispersion \f2=\u000021\nps2km\u00001, nonlinear coe\u000ecient \r= 1:3 W\u00001km\u00001, input\npump power PP= 500 mW, noise PSD =\u0000112 dBm/Hz.\nwe calculate the Shannon spectral entropy (SSE) [33]:\nSSE =\u0000X\n!p!loge(p!) (2)\nwherep!is the fraction of total power contained in the\nfrequency bin !. The evolution of the SSE (solid blue\nline) together with the pump power along the \fber length\n(solid orange line) are plotted in Fig. 1. (e). The SSE\nis initially close to 1 and increases from about 5 km too,\nwhen spectral components are generated outside the MI\nband, to reach a maximum value around 7. Then the\nentropy slightly oscillates around this maximum during\nthe rest of the propagation. After dropping to almost\nzero, the pump power oscillates too, which is a signature\nof spontaneous FPUT recurrences [26]. We notice the\nSSE and the pump power are antiphased. This shows\nthe system entropy is maximum when the pump power\nis minimum and vice versa.\nA numerical simulation of coherently driven MI is pre-\nsented in Fig. 2. We used a set of three waves, a pump\nand two symmetric sidebands located close to the maxi-\nmum of the MI gain band [32] (Fig. 2. (a)). Fig. 2. (b)\nshows the evolution of the spectrum along the propaga-\ntion distance. On a \frst stage, the seeds are ampli\fed by\nthe pump and the spectrum is broadened with equidis-\ntant peaks into a triangular shaped frequency comb, at3\n00.51Power (W)\n0 5 10 15 20\nDistance (km)00.20.40.6PP + 2PS (W)\n0.511.522.53\n-40-30-20-1001020-100-500\nRelative power (dBm)\nPower (W)\nSSE\n02468Rel. power\n(50 dB/div)\n-100 -50 0 50 100 150\nTime (ps)0\n5\n10\n15\n20Distance (km)\n-200 -100 0 100 200 300\nFrequency (GHz)0\n5\n10\n15\n20Distance (km)(a)\n(b)(c)\n(d)\n(e)\nFIG. 2. Same as Fig. 1 but with seededed MI. The modulation\nwave is set at 12 dB below the pump wave. On (e) is plotted\nthe total power of initial 3 waves excitation. PSD\u0019 \u0000400\ndBm/Hz (numerical noise level).\nthe expense of the pump wave. New waves located out-\nside the gain band, called harmonics, are generated out-\nside the gain band due to multiple four-wave mixing\n(FWM) processes. Then the system reaches the max-\nimum pump saturation stage around 2 :5 km and the\nenergy transfer to the higher order modes is not pos-\nsible anymore. Therefore, the energy transfer direction\nreverses and \rows back to the pump wave before the\nsystem returns almost exactly to its initial state. This\ncorresponds to a \frst FPUT recurrence cycle. Then, this\ndynamics repeats itself periodically during the propaga-\ntion till the end of the \fber. In this case, we can distin-\nguish 4 recurrences. The temporal evolution along the\n\fber length in the time domain is depicted in Fig. 2.\n(d). The system is triggered with a weakly modulated\nCW (Fig. 2. (c)). We notice the emergence of maximum\ncompression points coinciding with the spectral broaden-\ning maxima, whose positions are accurately predicted by\nthe theory [34, 35]. The \u0019-shift between two consecutive\nrecurrences is due to the initial phase relation between\nthe initial three waves [11, 36]. As with spontaneous\nMI, we plot the SSE of the system in Fig. 2. (e) (solid\nblue line). The entropy is weak ( <2), oscillates periodi-\ncally, at the the FPUT period but doesn't expand. The\npower contained in the initial waves (orange solid line),\nthe pump and the modulation sidebands, is also periodic.\nThe minima of this curve corresponds to maximal spec-\ntral broadenings because a signi\fcant part of the total\npower is contained in the Fourier modes of order n>2.III. NUMERICAL SIMULATION OF\nNOISE-INDUCED THERMALIZATION\nWe now look at the competition between spontaneous\nand seeded MI, when the CW is initially modulated by\na coherent sinusoidal modulation with noise (Figs. 3. (a)\nand (c)). The spectral and time evolutions are presented\n-300 -200 -100 0 100 200\nFrequency (GHz)-100 -50 0 50 100 150\nTime (ps)0\n5\n10\n20Distance (km)\n15-100-500\n-50-40-30-20-1001020\n0 5 10 15 20\nDistance (km)00.20.40.6PP + 2PS (W)\n1234567 0\n5\n10\n20Distance (km)\n1500.51Power (W)(a)\n(b)(c)\n(d)Relative power (dBm)\nPower (W)\n0246SSE(e)8Rel. power\n(50 dB/div)\nFIG. 3. Same as in Fig. 1 and Fig. 2. Same 3 waves input pa-\nrameters as the seeded MI numerics but with the spontaneous\nMI numerics noise \roor, PSD =\u0000112 dBm/Hz.\nin Figs. 3. (b) and (d) respectively. At the beginning of\nthe propagation in the \fber, the dynamics is very similar\nto the one illustrated Fig. 2 with two clear recurrences.\nThis means the dynamics of the system is dominated by\nthe coherently driven MI process. However, by further\npropagating within the \fber, the noise level of spectral\ncomponent located within the MI gain band increases and\nthe noise driven MI becomes signi\fcant at the expense of\nthe Fourier modes. Higher Fourier orders progressively\ndisappear into the noise \roor. The dynamics is then sim-\nilar to the one illustrated in Fig. 1, with the appearance\nof irregular temporal pulses and the disappearance of the\nfourth recurrence. The evolutions of the SSE and the to-\ntal power contained in the central 3 waves (Fig. 3. (e))\nalso reveal this transition toward an irreversible state.\nThe SSE keeps increasing to 7, and the total power tends\nto zero, as for the spontaneous MI example (Fig. 1). A\nfully thermalized state showing a complete disappearance\nof the coherent waves could be asymptotically reached by\npropagating in a much longer \fber (not shown here for\nthe sake of clarity). The transition speed from sponta-\nneous and noise driven MI can be controlled by tuning\nthe input noise level. We will exploit experimentally this\nfeature for the observation of the of the system evolving\ntoward an irreversible thermalized state.4\nIV. EXPERIMENTAL SETUP\nThe setup is similar to the one used in [11, 31] and\na simpli\fed version is presented in Fig. 4. A 1550 nm\nFIG. 4. Simpli\fed sketch of the setup with \"spectrum boxes\".\nf1is the frequency of the main laser and f2of the local oscil-\nlator, phase-locked at 800 MHz from f1.fmis the modulation\nfrequency, set here at 38 :2 GHz. SM \fber: single mode \fber.\nWaveshaper: amplitude and phase programmable optical \fl-\nter.\nCW signal is generated from a single frequency laser\ndiode. To excite the modulation instability regime co-\nherently, the signal is \frst phase-modulated at the fre-\nquencyfm= 38:2 GHz, very close of the maximum gain\nfrequency ( fmax= 38:4 GHz). The signal is then time\nshaped with an intensity modulator into 50 ns square\npulses, which are short enough to avoid stimulated Bril-\nlouin scattering and allow optical time domain re\rectom-\netry (OTDR). These pulses are long enough to be con-\nsidered as a CW wave for the modulation instability time\nscale. To shape the signal into the desired 3 waves input,\nthe relative power between the Fourier modes are tuned\nwith a Waveshaper. The signal is then combined with\nsynchronized additional white noise pulses, from a power\ntunable white noise source. It is followed by an ampli-\n\fcation stage through an erbium doped \fber ampli\fer\n(EDFA) to tune the nonlinear length ( PP= 470 mW).\nAn acousto-optic modulator, which has a extinction ra-\ntio much higher than electro-optic modulators, is then\ncascaded to reject the inter-pulses noise and to control\nthe pulses sequence. The signal is injected into a 16 :8\nkm long single mode \fber (G-652) in which nonlinear\ninteractions will occur.\nEven if the dissipation value is low ( \u000b= 0:2 dB/km),\non tens of kilometers, the nonlinear dynamic is highly\nimpacted [31]. In order to keep the signal power almost\nconstant over the whole \fber length, allowing us to pre-\nserve the integrable NLSE dynamics, an active loss com-\npensation scheme was implemented. At the \fber output,\na backward 1450 nm laser is injected to amplify the sig-\nnal by Raman e\u000bect. The Raman pump is located 13 :2THz away from the ampli\fed signal to bene\ft from the\nmaximum Raman gain [32]. For these tens of ns duration\npulses, we can neglect the saturation of the ampli\fer and\nconsider \rat ampli\fed pulses [37].\nThe power distribution along the \fber of the main\nFourier modes are obtained by an OTDR combined with\na multi-heterodyne technique [11, 31]. At each location\nin the \fber, a tiny part of the signal is Rayleigh backscat-\ntered and combined with a frequency comb with repeti-\ntion frequency fm, shifted by 600 MHz from the main\nlaser (the AO modulator induces a 200 MHz shift on f1).\nThe independant recording of the beating between the\nmain Fourier modes allows to recover the relative evolu-\ntion of the power along the all \fber length.\nThe noise power is increased progressively to study the\nin\ruence of the input noise level on the thermalization of\nthe FPUT recurrences. For each initial noise level value,\nwe record the power distribution of the two main Fourier\nwaves, the pump and the signal (the initial excitation\nbeing symmetric, we consider that the idler wave evolves\nidentically to the signal) and the spectrum at the \fber\nend. Much details about the experimental setup can be\nfound in Ref. [31].\nV. EXPERIMENTAL RESULTS\nWe varied the initial noise power spectral density\n(PSD) from\u0000121:3 to\u000091:9 dBm/Hz. We recorded the\npump and signal waves power evolutions along the \fber\nlength as well as the output spectra. These results are\npresented on the left panel of Fig. 5. (a-g). Fig. 5. (a)\ndisplays the output spectra as a function of the initial\nnoise PSD. At low amplitude noise \roors, we clearly dis-\ncern the Fourier modes: the pump, the seeds (at \u000638:2\nGHz) and the \frst harmonics (at \u000676:4 GHz). Between\nthese sidebands, MI also ampli\fes noisy components. We\nnote spectral holes around signal and idler waves. This\nis due to multiple four-wave mixing processes involving\nnoise components and these waves [38] (see Fig. 5. (e)).\nBy increasing the input noise PSD, the higher order har-\nmonics intensity decreases at the expense of the inter-\nbands lobes. The \frst harmonics (at \u000676:4 GHz) are\neven not visible anymore from the ampli\fed noise \roor\nfrom about\u0000105 dBm/Hz. The seed (at \u000638:2 GHz) are\nalso lost in the noise around \u000095 dBm/Hz. At the same\ntime, the amplitude of the noise \roor at the \fber out-\nput increases with the initial noise level. Indeed, with a\nhigher input noise level, the pump saturates at a shorter\n\fber length. This happens at the expense of the coherent\ndynamic process of seeded MI. The pump is not power-\nful enough to pursue the power transfer with the other\nFourier modes anymore, the spontaneous MI process is\ntaking the lead.\nIn Figs. 5. (b-d) are represented the pump (solid blue\nline), the seed (dotted blue line) and the relative total\npower of the three central waves distributions along the\n\fber length for noise PSD of \u0000121:3,\u0000101:8 and\u000092:45\n-100 -50 0 50 100\nFrequency (GHz)\nRelative power (5 dB/div)\n-100 -50 0 50 100\nFrequency (GHz)-120\n-115\n-110\n-105\n-100\n-95-120\n-115\n-110\n-105\n-100\n-95Noise power spectral density (dBm/Hz)(a)Noise power spectral density (dBm/Hz)Experiments Numerics\nRelative power (10 dB/div)00.20.40.6Power (W)\n00.20.40.6Power (W)\n0 5 10 15\nDistance (km)00.20.40.6Power (W)\n010101\n00.20.40.6Power (W)\n00.20.40.6Power (W)\n0 5 10 15\nDistance (km)00.20.40.6Power (W)\n010101\nPP+ 2PS (rel.) PP + 2PS (rel.) PP + 2PS (rel.)PP+ 2PS (rel.) PP + 2PS (rel.) PP + 2PS (rel.)(b)\n(c)\n(d)\n-100 -50 0 50\nFrequency (GHz)Rel. power (10 dB/div) -100 -50 0 50\nFrequency (GHz)-100 -50 0 50 100\nFrequency (GHz)-100 -50 0 50\nFrequency (GHz)Rel. power (10 dB/div)(e) (f) (g)\n-100 -50 0 50\nFrequency (GHz)-100 -50 0 50\nFrequency (GHz)100(a') (b')\n(c')\n(d')\n(e') (f') (g')\nFIG. 5. (a), (a') Spectrum at the \fber end as a function of the initial power spectral density of noise. (b), (c), (d) and (b'),\n(c'), (d') Evolution of the pump (solid blue line), the signal (dotted blue line) and the sum of the central 3 waves PP+ 2PS\n(solid orange line) along the \fber distance. (e), (f), (g) and (e'), (f'), (g') Output spectra. The initial power spectral densities\nare respectively ( \u0000121:3;\u0000101:8;\u000092:4) dBm/Hz and ( \u0000121:3;\u0000106:4;\u000097:3) dBm/Hz. The left panel is corresponding to the\nexperimental recordings while the right panel to the numerics.\ndBm/Hz respectively. For the \frst case (PSD = \u0000121:3\ndBm/Hz, see Fig. 5 (b)), no additional noise has been\nadded and the noise level is the intrinsic noise of the\nsetup (lowest value). We observe 3 :5 recurrences with a\nvery small decrease of the 3 waves total power, as it has\nbeen observed in [14]. This process is then still highly\ncoherent and the seeded MI is dominating the nonlin-\near power transfers. When the initial noise \roor is in-\ncreased at\u0000101:8 dBm/Hz (Fig. 5 (c)), the power of the\nFourier modes decreases during the propagation. It in-\ndicates that a part of the power is located elsewhere in\nthe spectrum, mainly in between coherent components.\nFor instance at z= 7:4 km, (after 1.5 recurrences), only\nhalf of the power is located in the Fourier modes and\nit reaches 17% at the \fber end. We see in the output\nspectrum depicted in Fig. 5 (f) that only the three main\nFourier modes remain visible. The seeded MI process is\nnot longer e\u000bective and the spontaneous MI is not neg-\nligible anymore. In the last case, with a noise PSD of\n\u000092:4 dBm/Hz (Fig. 5 (d)), we can only distinguish one\nand half recurrences and the power of the central three\nwaves becomes lower than 50% before the system reaches\nthe \frst half recurrence. After one and half recurrence\n(about 5 km), the power of both the pump and the seed\nwaves drops drastically during the propagation to reach\nweak values, comparable to the noise \roor. At the \fber\noutput, the spectrum in Fig. 5. (g) has a triangular shape\nfrom which it is impossible to distinguish any coherent\nFourier modes. This shape is typical from parametrically\ndriven systems [39] and is the signature of the thermal-ization of the FPUT process in optical \fbers. With this\nhigh noise input level, all coherent states are hidden in\nthe noise components including the pump. The conse-\nquence is that a comeback to a coherent initial state is\nimpossible, the evolution being thus irreversible. This\nunambiguously proves the system reached a thermalized\nstate of the FPUT recurrence process in optical \fbers.\nThe results from numerical simulations are plotted in\nthe right panel of Fig. 5. The agreement between experi-\nments and numerics is very good. The experiments follow\nthe same dynamics than the one predicted through nu-\nmerics, the progressive coherence loss with the increase\nof the input noise. Remarkably, the disappearance of co-\nherent structures and the triangular shape of the output\nspectra for high input noise levels is very similar to ex-\nperiments (Fig. 5 (g) and (g')). This con\frms the system\nreached an irreversible thermalized state.\nVI. CONCLUSION\nWe reported the experimental observation of the ther-\nmalization of the FPUT process in optical \fbers. We\nshowed that the FPUT recurrences irreversibly disap-\npear due to a competition between noise driven and co-\nherently driven MI processes leading to an increase of\nthe system entropy. Our experiments were realized by\nmeans of a multi-HOTDR system to monitor the evolu-\ntion of the power along the \fber length. The observation\nof the route to the thermalisation is made possible with6\nan active loss compensation scheme to observe up to 4\nrecurrences and by adding noise at the \fber input to ac-\ncelerate the process. At high input noise level, we showed\nthat the system reaches an irreversible thermalized state,\nwith an output spectrum only made of noisy components.\nExperimental results were con\frmed by numerical simu-\nlations with an excellent agreement. These experimental\nworks revealed the evolution of a nonlinear system from a\ncoherent dynamics between its nonlinear modes toward\nan irreversible thermalized state. This behavior, origi-\nnally expected by Fermi, Pasta, Ulam and Tsingou in a\nnonlinearly coupled masses system, had never been re-\nported experimentally so far. We anticipate these resultsto be generalizable to all focusing cubic media and to\ncontribute to a better understanding of the dynamics of\nnoisy and coherently driven nonlinear systems and the\nformation of rogue waves.\nACKNOWLEDGMENTS\nThe present research was supported by the agence\nNationale de la Recherche (Programme Investissements\nd'Avenir, I-SITE VERIFICO); Ministry of Higher Edu-\ncation and Research; Hauts de France Council; European\nRegional Development Fund (Photonics for Society P4S)\nand the CNRS (IRP LAFONI).\n[1] E. Fermi, P. Pasta, S. Ulam, and M. Tsingou,\nCollected Papers of Enrico Fermi, edited by E. ed. Segr\u0013 e,\nVol. 2 (Univ. Chicago Press, Chicago, Illinois, 1955) pp.\n977{988.\n[2] M. Porter, N. Zabusky, B. Hu, and D. Campbell, Amer-\nican Scientist 97(2009), 10.1511/2009.78.214.\n[3] B. M. Lake, H. C. Yuen, H. Rungaldier, and W. E.\nFerguson, Journal of Fluid Mechanics 83, 49{74 (1977).\n[4] O. Kimmoun, H. C. Hsu, H. Branger, M. S. Li, Y. Y.\nChen, C. Kharif, M. Onorato, E. J. R. Kelleher, B. Ki-\nbler, N. Akhmediev, and et al., Scienti\fc Reports 6,\n28516 (2016).\n[5] D. Pierangeli, M. Flammini, L. Zhang, G. Marcucci, A. J.\nAgranat, P. G. Grinevich, P. M. Santini, C. Conti, and\nE. DelRe, Physical Review X 8, 041017 (2018).\n[6] C. Cambournac, H. Maillotte, E. Lantz, J. M. Dudley,\nand M. Chauvet, Journal of the Optical Society of Amer-\nica B19, 574 (2002).\n[7] G. Van Simaeys, P. Emplit, and M. Haelterman, Physi-\ncal Review Letters 87, 033902 (2001).\n[8] G. Van Simaeys, P. Emplit, and M. Haelterman, Journal\nof the Optical Society of America B 19, 477 (2002).\n[9] K. Hammani, B. Wetzel, B. Kibler, J. Fatome, C. Finot,\nG. Millot, N. Akhmediev, and J. M. Dudley, Optics Let-\nters36, 2140 (2011).\n[10] A. Bendahmane, A. Mussot, A. Kudlinski, P. Szriftgiser,\nM. Conforti, S. Wabnitz, and S. Trillo, Optics Express\n23, 30861 (2015).\n[11] A. Mussot, C. Naveau, M. Conforti, A. Kudlinski,\nF. Copie, P. Szriftgiser, and S. Trillo, Nature Photonics\n12, 303{308 (2018).\n[12] X. Hu, W. Chen, Y. Lu, Z. Yu, M. Chen, and Z. Meng,\nIEEE Photonics Technology Letters 30, 47{50 (2018).\n[13] J.-W. Goossens, H. Hafermann, and Y. Jaou en, Scienti\fc\nReports 9, 18467 (2019).\n[14] G. Vanderhaegen, P. Szriftgiser, A. Kudlinski, M. Con-\nforti, S. Trillo, M. Droques, and A. Mussot, Optics Ex-\npress28, 17773 (2020).\n[15] M. Wu and C. E. Patton, Physical Review Letters 98,\n047202 (2007).\n[16] Y. V. Lvov and M. Onorato, Physical Review Letters\n120, 144301 (2018).\n[17] G. Benettin, H. Christodoulidi, and A. Ponno, Journal\nof Statistical Physics 152, 195{212 (2013).[18] M. Onorato, L. Vozella, D. Proment, and Y. V. Lvov,\nProceedings of the National Academy of Sciences 112,\n4208{4213 (2015).\n[19] S. D. Pace and D. K. Campbell, Chaos: An Interdisci-\nplinary Journal of Nonlinear Science 29, 023132 (2019).\n[20] S. Wabnitz and B. Wetzel, Physics Letters A 378,\n2750{2756 (2014).\n[21] N. Akhmediev, J. Soto-Crespo, and A. Ankiewicz,\nPhysics Letters A 373, 2137{2145 (2009).\n[22] J. M. Dudley, F. Dias, M. Erkintalo, and G. Genty,\nNature Photonics 8, 755{764 (2014).\n[23] S. Toenger, T. Godin, C. Billet, F. Dias, M. Erkintalo,\nG. Genty, and J. M. Dudley, Scienti\fc Reports 5, 10380\n(2015).\n[24] P. Suret, R. E. Koussai\f, A. Tikan, C. Evain, S. Ran-\ndoux, C. Szwaj, and S. Bielawski, Nature Communica-\ntions7, 13136 (2016).\n[25] M. N arhi, B. Wetzel, C. Billet, S. Toenger, T. Sylvestre,\nJ.-M. Merolla, R. Morandotti, F. Dias, G. Genty, and\nJ. M. Dudley, Nature Communications 7, 13675 (2016).\n[26] A. E. Kraych, D. Agafontsev, S. Randoux, and P. Suret,\nPhysical Review Letters 123, 093902 (2019).\n[27] A. Gelash, D. Agafontsev, V. Zakharov, G. El, S. Ran-\ndoux, and P. Suret, Physical Review Letters 123, 234102\n(2019).\n[28] J. M. Dudley, G. Genty, F. Dias, B. Kibler, and\nN. Akhmediev, Optics Express 17, 21497 (2009).\n[29] D. R. Solli, C. Ropers, and B. Jalali, Physical Review\nLetters 101, 233902 (2008).\n[30] J. M. Dudley, G. Genty, and B. J. Eggleton, Optics\nExpress 16, 3644 (2008).\n[31] C. Naveau, G. Vanderhaegen, P. Szriftgiser, G. Mar-\ntinelli, M. Droques, A. Kudlinski, M. Conforti, S. Trillo,\nN. Akhmediev, and A. Mussot, Frontiers in Physics 9,\n25 (2021).\n[32] G. Agrawal, Nonlinear Fiber Optics, 4th edition (Aca-\ndemic Press, 2007).\n[33] C. E. Shannon, The Bell System Technical Journal 27,\n379{423 (1948).\n[34] P. Grinevich and P. Santini, Physics Letters A 382,\n973{979 (2018).\n[35] C. Naveau, P. Szriftgiser, A. Kudlinski, M. Conforti,\nS. Trillo, and A. Mussot, Optics Letters 44, 5426 (2019).7\n[36] G. Vanderhaegen, P. Szriftgiser, C. Naveau, A. Kudlinski,\nM. Conforti, S. Trillo, N. Akhmediev, and A. Mussot,\nOptics Letters 45, 3757 (2020).[37] G. Vanderhaegen, P. Szriftgiser, M. Conforti, A. Kudlin-\nski, M. Droques, and A. Mussot, Optics Letters 46, 5019\n(2021).\n[38] K. Inoue and T. Mukai, Optics Letters 26, 869 (2001).\n[39] N. Akhmediev, A. Ankiewicz, J. Soto-Crespo, and J. M.\nDudley, Physics Letters A 375, 775{779 (2011)." }, { "title": "1012.5395v2.Decoherence_of_the_two_level_system_coupled_to_a_fermionic_reservoir_and_the_information_transfer_from_it_to_the_environment.pdf", "content": "arXiv:1012.5395v2 [cond-mat.mes-hall] 1 Feb 2011Decoherence of the two-level system coupled to a fermionic r eservoir and the\ninformation transfer from it to the environment\nPei Wang∗\nWangcunlvtukuang, 255311 Zibo, P. R. China\n(Dated: November 8, 2018)\nWe study the evolution of reduced density matrix of an impuri ty coupled to a Fermi sea after the\ncoupling is switched on at time t= 0. We find the non-diagonal elements of the reduced density\nmatrix decay exponentially, and the decay constant is the im purity level width Γ. And we study the\ninformation transfer rate between the impurity and the Ferm i sea, which also decays exponentially.\nAnd the decay constant is kΓ withk= 2∼4. Our results reveal the relation between information\ntransfer rate and decoherence rate.\nPACS numbers: 03.65.Yz, 03.67.Mn, 73.23.-b\nI. INTRODUCTION\nCoherence is the fundamental characteristics of quan-\ntum systems. The state of an isolated system is repre-\nsented by a vector moving in Hilbert space. The state\nvector decides the mean value and probability distribu-\ntion of observables. But in reality truly isolated sys-\ntems do not exist. The systems we can observe are all\nopen systems, which will exchange energy, matter and\ninformation with the rest of the world. The state of an\nopen system is generally a mixed state and should be\nrepresented by a density matrix. Open systems show\nmany different features from isolated ones. One of the\nmost important is decoherence [1–3], i.e., the nondiago-\nnal elements of the density matrix in pointer state basis\nrapidly decay to zero. The decoherence can be modelled\nas a non-unitary process, which can be described by the\nmaster equation with a Lindblad decohering term [4, 5].\nThe decoherence marks the boundary between quantum\nworldandclassicalworld[6], andhasbeenintenselystud-\nied because of its role in quantum measurement theory\nand quantum computation [6, 7]. The two level sys-\ntems coupled to spin bath [12] and oscillator bath [3]\nare good examples for studying decoherence. Plenty of\nworks have been contributed to this subject (for a review\nsee Ref. [2, 6, 12]). Recently the two level system coupled\nto fermionic bath [13, 14] emerged before the explorers.\nIn this paper we consider a simple fermionic bath: the\nresonant level model.\nAdditionally there exists information transfer between\nan open system and its environment. The quantum mea-\nsurement is realized due to the information transfer from\nthe system to apparatus. Zurek noticed the relation\nbetween decoherence and information transfer from the\nsystem to the observer and environment [6]. This rela-\ntion is familiar to quantum mechanics community, as the\nwell known state vector collapse in measurement. And\nthe which-path experiments, which are designed to verify\ncomplementarityprinciple, giveastraightforwardpicture\n∗Electronic address: pei.wang@live.combetween decoherence and information about particle po-\nsitions. Generally speaking, any one unavoidably cause\ndecoherence of a system when trying to get the informa-\ntion of it.\nThe relation between information transfer and deco-\nherence can be understood in the language of state vec-\ntor. The system and environment compose an isolated\nsystem which is supposed to be in a pure state. If the\ncomposite system is all the time in a product state, the\nstate of the system we concern is always pure and its\nevolution can be described by a unitary transformation.\nAnd the system and environment share zero informa-\ntion. Otherwise the composite system evolves into an\nentangled state. The evolution of system we concern is\nnon-unitary,indicatingdecoherenceprocess. Atthesame\ntime the system and environment exchange information,\nincreasing their mutual information into a finite value.\nAn open system is all the time transfering informa-\ntion outside and suffering decoherence. The dynamics of\nentropy entanglement (the half mutual information) has\nbeen studied in different one dimensional systems [8–11].\nWhile thereis still little study about the relationbetween\ndecoherence rate and the information transfer rate in a\ndynamicalsystem. In this paper, wetryto usean exactly\nsolvable model to demonstrate this relation.\nII. TWO LEVEL SYSTEM COUPLED TO\nFERMIONIC RESERVOIR\nThe resonant level model is an exactly solvable model\nwhich has been widely used to describe the transport\nthroughquantum dot. The Hamiltonian ofresonantlevel\nmodel is\nˆH=/summationdisplay\nkǫkˆc†\nkˆck+ǫiˆd†ˆd+V/summationdisplay\nk(ˆc†\nkˆd+h.c.),(1)\nwhere ˆckandˆdare the single electron annihilation op-\nerators in the lead and at the impurity site respectively.\nWe set the Fermi energy of the lead to be zero. The\nresonant level model can be regarded as a two level sys-\ntem coupled to Fermi sea, if we use |1∝angbracketrightto represent the2\nstate when the impurity site is occupied by an electron\nand|0∝angbracketrightto represent the state when the impurity site is\nempty. We suppose at initial time the state of the impu-\nrity is (α|0∝angbracketright+β|1∝angbracketright). The decoherence will happen after\nwe switch on the coupling between the impurity and the\nFermi sea at time t= 0. The pointer states are |0∝angbracketrightand\n|1∝angbracketright. The non-diagonal elements of the reduced density\nmatrix of the impurity will decay to zero exponentially.\nThe simple way to get the elements of the reduced\ndensity matrix at arbitrary time is to calculate the ob-\nservables which can be related to them. We suppose the\nreduced density matrix is expressed as\nˆρi(t) =ρ00(t)|0∝angbracketright∝angbracketleft0|+ρ11(t)|1∝angbracketright∝angbracketleft1|\n+ρ01(t)|0∝angbracketright∝angbracketleft1|+ρ10(t)|1∝angbracketright∝angbracketleft0|.(2)\nAt time t= 0, we have ρ00(0) =|α|2,ρ11(0) =|β|2\nandρ01(0) =ρ∗\n10(0) =αβ∗. We notice the annihilation\noperator at the impurity site can be expressed as ˆd=\n|0∝angbracketright∝angbracketleft1| ⊗1F, where 1Fis the identity operator on the\nsubspace of the Fermi sea. Then it is easy to know\n∝angbracketleftˆd(t)∝angbracketright=ρ10(t). (3)\nSimilarly we have ˆd†=|1∝angbracketright∝angbracketleft0|⊗1Fandˆd†ˆd=|1∝angbracketright∝angbracketleft1|⊗1F,\nwhich indicate ρ01(t) =∝angbracketleftˆd†(t)∝angbracketrightandρ11(t) =∝angbracketleftˆd†(t)ˆd(t)∝angbracketright\nrespectively.\nWe could calculate the expectation value of ˆd(t) and\nˆd†(t)ˆd(t) by non-equilibrium Green’s function method\nor by diagonalizing the Hamiltonian (1) (for details see\nRef. [15]). Here wefollowthe latter way, and first express\nthe Hamiltonian in hybridization basis as\nˆH=/summationdisplay\nsǫsˆc†\nsˆcs, (4)\nwhere ˆcs=/summationtext\nkV\nǫs−ǫkBsˆck+Bsˆd. The transformation\ncoefficient is Bs=V√\n(ǫs−ǫi)2+Γ2, where Γ = ρπV2is\nthe linewidth of the impurity level and ρthe density of\nstates of the Fermi sea. The inverse transformation is\nˆd=/summationtext\nsBsˆcsand through this the time-dependent anni-\nhilation operator evaluates\nˆd(t) =e−iǫit−Γtˆd+V/summationdisplay\nke−iǫkt−e−iǫit−Γt\nǫk−ǫi+iΓˆck.(5)\nFor simplicity we set /planckover2pi1= 1 throughout the paper. Con-\nsidering the initial condition we then have ρ10(t) =\nα∗βe−iǫit−Γt,ρ01(t) =αβ∗eiǫit−Γt, and\nρ11(t) =V2/summationdisplay\nk|ei(ǫk−ǫi)t−e−Γt|2\n(ǫk−ǫi)2+Γ2θ(−ǫk)\n+e−2Γt|β|2. (6)\nThe non-diagonal elements ρ01(t) andρ10(t) decay ex-\nponentially, and the decay rated|ρ01(t)|\ndt=d|ρ10(t)|\ndt∝\nΓe−Γt. This is the effect of decoherence. Here the Fermi\nsea plays the role of environment.-8-40\n 0 0.5 1 1.5 2 2.5ln(T/Γ)\nTime t (1/ Γ)2\n1.5\n1\n0.5\n0T/Γ\nFIG. 1: Information transfer rate between the impurity and\nthe Fermi sea. The impurity level is at the resonant position ,\ni.e.,ǫi= 0. The top figure shows the time-dependentinforma-\ntion transfer rate. The bottom figure shows the logarithmic\ninformation transfer rate. We see clearly the exponential d e-\ncay ofT.\nIII. INFORMATION TRANSFER RATE\nNext we study the information transfer between the\nimpurity and the Fermi sea. In an open system, such\nas the impurity in resonant level model, the information\nwill broadcast to the environment. This is a stumbling\nblock in quantum computation, while a necessary pro-\ncess in quantum measurement where our purpose is to\nget the information of the system. The mutual informa-\ntion between the system and the environment Imeasures\nhow much information is shared between them. Its time\nderivative is a good measure of the information transfer\nrate, which is defined in this paper as\nT=1\n2dI\ndt. (7)\nHere we take the factor1\n2because the information trans-\nfer both from the impurity to the Fermi sea and vice\nversa. The composite system of the impurity and the\nFermi sea is in a pure state. According to Schmidt de-\ncomposition [7] we know the mutual information Iis two\ntimes of the entropy of the impurity, which can be got\nfrom the eigenvalues of the reduced density matrix. We\nfind at arbitrary time the information transfer rate is\nT=dλ+\ndt(lnλ−−lnλ+), (8)\nwhereλ±=1±/radicalbig\n1−4(ρ00ρ11−ρ01ρ10)\n2.\nThe non-diagonal elements of the reduced density ma-\ntrix decay exponentially, so does the information transfer\nrateT(see Fig. 1). The information transfer rate gradu-\nally decreases when the mutual information between the\nimpurity and the Fermi sea increases to maximum. We3\n-8-6-4-2 0 2\n 0 0.5 1 1.5 2ln(T/Γ)\nTime t (1/ Γ)|α|2=1\n|α|2=0\n|α|2=0.81\n|α|2=0.19\n|α|2=0.64\n|α|2=0.36\nFIG.2: Information transferrateatdifferent |α|2whenǫi= 0.\nThe decay constant of the information transfer rate depends\non|α|2.\nemploy Γ as the unit of informationtransfer rateand 1 /Γ\nthe unit of time. The decay rate of the function T(t) de-\npends on Γ. At large time scale we approximately have\nT∝Γe−kΓt, wherekis a dimensionless constant. Recall-\ning the fact that the decay rate of non-diagonal density\nmatrix elements is proportional to Γ e−Γt, we get the re-lation between decoherencerate and information transfer\nrate: they both decay exponentially, and their decay con-\nstants are proportional to each other.\nWe find at large time scale the information transfer\nratedoes notdepend on the position ofthe impurity level\nǫi, but only depends on the initial condition |α|2. Fig. 2\nshows the information transfer rate at different |α|2. The\ninformationtransferrateisproportionaltoΓ e−kΓt, where\nkdecreasesfrom 4 to 2 when |α|2decreases from 1 to 0 .5,\nand increases back to 4 when |α|2decreases from 0 .5 to\n0.\nIV. SUMMARY AND CONCLUSION\nWe study the decoherence of an impurity coupled to a\nFermi sea after the coupling is switched on at time t= 0.\nThe exponential decay of non-diagonal elements in the\nreduced density matrix of the impurity is observed. And\nthe decay constant is found to be Γ. We also study the\ninformation transfer rate between the impurity and the\nFermi sea. At large time scale the information transfer\nrate decays exponentially, and the decay constant is kΓ\nwithk= 2∼4 depending on the initial state of the im-\npurity (or |α|2). Our study reveals the relation between\ninformation transfer rate and decoherence rate.\n[1] W. H. Zurek, Physics Today 44, 36 (1991).\n[2] M. Schlosshauer, Rev. Mod. Phys. 76, 1267 (2004).\n[3] A. J. Leggett, S. Chakravarty, A. T. Dorsey, M. P. A.\nFisher, A. Garg, W. Zwerger, Rev. Mod. Phys. 59, 1\n(1987).\n[4] V. Gorini, A. Kossakowski, and E. C. G. Sudarshan, J.\nMath. Phys. (N.Y.) 17, 821 (1976)\n[5] G. Lindblad, Commun. Math. Phys. 48, 119 (1976).\n[6] W. H. Zurek, Rev. Mod. Phys. 75, 715 (2003).\n[7] M. A. Nielsen, and I. L. Chuang, Quantum Computa-\ntion and Quantum Information (Cambridge Univ. Press,\nCambridge, 2000).\n[8] V. Eisler, and I. Peschel, J. Stat. Mech., P06005 (2007).\n[9] G. De Chiara, S. Montangero, P. Calabrese, R. Fazio, J.Stat. Mech., P03001 (2006).\n[10] J. Eisert, and T. J. Osborne, Phys. Rev. Lett. 97, 150404\n(2006).\n[11] P. Calabrese, and J. Cardy, J. Stat. Mech., P04010\n(2005).\n[12] N. Prokof’ev, and P. Stamp, Rep. Prog. Phys. 63, 669\n(2000).\n[13] R. M. Lutchyn, L. Cywinski, C. P. Nave, S. Das Sarma,\nPhys. Rev. B 78, 024508 (2008).\n[14] N. Yamada, A. Sakuma, H. Tsuchiura, J. Appl. Phys.\n101, 09C110 (2007).\n[15] P. Wang, M. Heyl, S. Kehrein, J. Phys.: Condens. Matter\n22, 275604 (2010)." }, { "title": "2101.10849v1.The_Central_Engines_of_Fermi_Blazars.pdf", "content": "DRAFT VERSION JANUARY 27, 2021\nTypeset using L ATEXtwocolumn style in AASTeX63\nThe Central Engines of Fermi Blazars\nVAIDEHI S. P ALIYA ,1, 2A. D OM´INGUEZ ,3, 4M. A JELLO ,5A. O LMO -GARC´IA,6AND D. H ARTMANN5\n1Aryabhatta Research Institute of Observational Sciences (ARIES), Manora Peak, Nainital 263001, India\n2Deutsches Elektronen Synchrotron DESY, Platanenallee 6, 15738 Zeuthen, Germany\n3Institute of Particle and Cosmos Physics (IPARCOS), Universidad Complutense de Madrid, E-28040 Madrid, Spain\n4Department of EMFTEL, Universidad Complutense de Madrid, E-28040 Madrid, Spain\n5Department of Physics and Astronomy, Clemson University, Kinard Lab of Physics, Clemson, SC 29634-0978, USA\n6Departamento de F ´ısica de la Tierra y Astrof ´ısica, Universidad Complutense de Madrid (UCM, Spain) and Instituto de F ´ısica de Part ´ıculas y del Cosmos\n(IPARCOS)\n(Received January 27, 2021; Revised January 27, 2021; Accepted January 27, 2021)\nSubmitted to ApJS\nABSTRACT\nWe present a catalog of central engine properties, i.e., black hole mass ( MBH) and accretion luminosity\n(Ldisk), for a sample of 1077 blazars detected with the Fermi Large Area Telescope. This includes broad\nemission line systems and blazars whose optical spectra lack emission lines but dominated by the absorption\nfeatures arising from the host galaxy. The average MBHfor the sample ishlogMBH;allM\fi= 8:60and there\nare evidences suggesting the association of more massive black holes with absorption line systems. Our results\nindicate a bi-modality of Ldiskin Eddington units ( Ldisk/LEdd) with broad line objects tend to have a higher\naccretion rate ( Ldisk/LEdd>0.01). We have found that Ldisk/LEddand Compton dominance (CD, the ratio\nof the inverse Compton to synchrotron peak luminosities) are positively correlated at >5\u001bconfidence level,\nsuggesting that the latter can be used to reveal the state of accretion in blazars. Based on this result, we propose\na CD based classification scheme. Sources with CD >1 can be classified as High-Compton Dominated or HCD\nblazars, whereas, that with CD .1 are Low-Compton Dominated (LCD) objects. This scheme is analogous to\nthat based on the mass accretion rate proposed in the literature, however, it overcomes the limitation imposed\nby the difficulty in measuring LdiskandMBHfor objects with quasi-featureless spectra. We conclude that the\noverall physical properties of Fermi blazars are likely to be controlled by the accretion rate in Eddington units.\nThe catalog is made public at http://www.ucm.es/blazars/engines and Zenodo.\nKeywords: methods: data analysis — gamma rays: general — galaxies: active — galaxies: jets\n1.INTRODUCTION\nBlazars are a subclass of jetted active galactic nuclei\n(AGN) family that host relativistic jets closely aligned to\nthe line of sight to the observer. Due to their peculiar ori-\nentation, the emitted radiation is relativistically amplified\nthereby making blazars observable at cosmological distances\n(z >5). Typically, blazars are hosted by elliptical galaxies\nand powered by massive (10\u00188\u000010M\f) black holes (e.g.,\nUrry et al. 2000; Shaw et al. 2012). The multi-wavelength\nspectral energy distribution (SED) of a blazar exhibits a typ-\nCorresponding author: Vaidehi S. Paliya\nvaidehi.s.paliya@gmail.comical double hump structure. The low-energy bump peaks at\nradio-to-X-rays and is well explained with the synchrotron\nmechanism. On the other hand, the inverse Compton process\nis found to satisfactorily reproduce the high-energy bump lo-\ncated in the MeV-TeV enetgy range, considering the leptonic\nradiative models (e.g., Dermer et al. 2009; van den Berg et al.\n2019). The properties of the central engine, i.e., black hole\nmass (MBH) and accretion disk luminosity ( Ldisk), are found\nto significantly correlate with the broadband SED of blazars\n(Ghisellini et al. 2014; Paliya et al. 2017b, 2019, 2020a),\nindicating a close-connection between the accretion process\nand relativistic jets.\nHistorically, the beamed AGN have been classified based\non the appearance of the broad emission lines in their op-\ntical spectra. Blazars exhibiting strong and broad emissionarXiv:2101.10849v1 [astro-ph.HE] 26 Jan 20212 P ALIYA ET AL .\nlines (rest-frame equivalent width of EW >5˚A) are known\nas flat spectrum radio quasars or FSRQs. The objects show-\ning quasi-featureless spectra (EW <5˚A), on the other hand,\nbelong to the category of BL Lacertae or BL Lac sources\n(Stickel et al. 1991). It was predicted that the lack of broad\nemission lines in the optical spectra of the latter might be\ndue to Doppler boosted continuum swamping out any spec-\ntral lines even if they exist. However, stellar absorption fea-\ntures originated from the host galaxy are observed in many\nnearby (z<1) BL Lac sources, thus indicating that emission\nlines are intrinsically weak (Plotkin et al. 2011). To make\nthe case more complex, broad emission lines have been de-\ntected in the low jet activity states of many BL Lac objects\n(e.g., Vermeulen et al. 1995). Therefore, a blazar classifica-\ntion scheme based on the EW does not reveal the physical\ndistinction between FSRQs and BL Lac sources.\nGhisellini et al. (2011) proposed that the criterion to dis-\ntinguish FSRQs and BL Lac objects should be based on the\nbroad line region (BLR) luminosity ( LBLR). FSRQs are\nsources with a luminous BLR ( LBLR&10\u00003times Edding-\nton luminosity or LEdd) which, in turn, suggests a radiatively\nefficient accretion process ( Ldisk&1%LEdd). BL Lac\nsources, on the other hand, host a radiatively inefficient ac-\ncretion flow which fails to photo-ionize the BLR clouds (see\nalso, Sbarrato et al. 2014). The presence of broad and strong\nemission lines in the optical spectra of FSRQs and absence\nin BL Lacs, thus, can be explained in this more physically\nintuitive classification. However, one of the key parameters\nin this scheme, LEdd, requires the knowledge of the mass of\nthe central black hole. Moreover, an estimation of Ldisk, or\nequivalently mass accretion rate, remains a challenge for BL\nLac objects.\nIn the context of the \r-ray emitting beamed AGN, the re-\nsearch works focused on the central engine properties have\nso far remained concentrated on the broad emission lines\nblazars (cf. Ghisellini et al. 2014). Apart from a few individ-\nual works (e.g., Becerra Gonz ´alez et al. 2020), a population\nstudy on the central engine of BL Lac sources is still lacking.\nIn this work, we have attempted to address this outstanding\nissue by carrying out a detailed optical spectroscopic analy-\nsis of a large sample of blazars present in the Fermi Large\nArea Telescope (LAT) fourth source catalog data release 2\n(4FGL-DR2; Abdollahi et al. 2020). Furthermore, we have\nalso determined that Compton dominance (CD, the ratio of\nthe inverse Compton to synchrotron peak luminosities) can\nbe considered as a good proxy for the accretion rate in Ed-\ndington units. This parameter can be crucial for objects\nwhose optical spectroscopic analyses are tedious either due\nto telescope constraints, intrinsic source faintness, or lack of\nemission lines in their optical spectrum. As discussed above,\nthe latter could be due to a radiatively inefficient accretion\nprocess and/or Doppler boosted jet radiation swamping outTable 1. TheFermi blazar sample studied in this work.\nSample Classification Number of Sources\nEmission line blazars 258 (SDSS)\n416 (literature)\n674 (all)\nAbsorption line blazars 200 (SDSS)\n146 (literature)\n346 (all)\nBlazars with bulge magnitude from literature 47\nBlazars with MBHandLdiskfrom literature 10\nemission lines and thus does not allows us to infer the intrin-\nsic nature of the quasar. The CD, therefore, could be the key\nto connect the FSRQ-BL Lac dichotomy and a classification\nscheme based on this parameter may fully explain the diverse\nphysical properties of Fermi blazars.\nIn Section 2, we describe the Fermi blazar sample and the\ndetails of the optical spectroscopic data analysis procedures\nare provided in Section 3. We elaborate various techniques\nto determine MBHandLdiskin Section 4 and discuss them\nin Section 5. Section 6 and 7 are devoted to the physical\ninterpretation of the estimated central engine parameters and\nCD. We summarize our findings in Section 8. Throughout,\nwe adopt the flat cosmology parameters: h= \n \u0003= 0:7.\n2.SAMPLE\nWe have considered all \r-ray emitting blazars or blazar\ncandidates present in the 4FGL-DR2 catalog to prepare the\nbase sample of 5250 objects. Since the main goal of this\nwork is to derive MBHandLdiskfrom the optical spectrum,\nwe carried out an extensive search for the optical spectro-\nscopic information for as many blazars as possible. This was\ndone by: (i) cross-matching the 4FGL-DR2 catalog with the\n16th data release of the Sloan Digital Sky Survey (SDSS-\nDR16; Ahumada et al. 2020), (ii) searching the published\noptical spectrum of all remaining blazars in the literature us-\ning NASA Extragalactic Database and SIMBAD Astronom-\nical Database, and (iii) searching the published MBHand\nLdiskvalues in the literature for objects leftover after com-\npleting steps (i) and (ii). In all of the cases, we used the posi-\ntional information of the low-frequency counterpart of the \r-\nray source given in the 4FGL-DR2 catalog or in other works\nand visually examine the quality of the spectra. This exercise\nled to a collection of 458 SDSS spectra including 258 emis-\nsion line objects and 200 blazars showing prominent absorp-\ntion lines. From research articles, we were able to collect op-\ntical spectra of 562 blazars either in tabular format or simplyBLAZAR ’SCENTRAL ENGINE 3\nas plots which were digitized using WebPlotDigitizer1\n(Rohatgi 2020). Among them, 416 objects are broad emis-\nsion line systems and the optical spectra of 146 are domi-\nnated by absorption lines arising from the host galaxy. Fur-\nthermore, we were able to find host galaxy bulge magnitude\nfor 47 sources and MBHand/orLdiskvalues for 10 blazars\nfrom the literature. Altogether, our final sample consists of\n1020 objects with available optical spectra and 57 others with\nbulge magnitude or MBHandLdiskdirectly taken from the\nliterature. Note that emission line blazars mainly belong to\nFSRQ class of AGN, whereas BL Lacs dominate the ab-\nsorption line sample. However, since many of the BL Lac\nobjects have exhibited broad emission lines in their optical\nspectra taken during a low-jet activity state, i.e., akin to FS-\nRQs, we divide the whole sample in emission and absorption\nline systems, respectively, rather than distinguishing them as\nFSRQ/BL Lacs. Table 1 summarizes our blazar sample.\nThe redshift distribution of the sample is shown in Fig-\nure 1. Instead of directly using the redshift information from\nthe fourth catalog of the Fermi -LAT detected AGN (4LAC;\nAjello et al. 2020), we have visually inspected the optical\nspectrum of all of the sources to determine the accurate\nsource redshift. Though most of our measurements agree\nwith redshifts published in the 4LAC catalog, a small frac-\ntion (<5%) of sources do show differences (Figure 2, left\npanel). This is expected since 4LAC itself is a compilation of\ninformation taken from various databases. In the right panel\nof Figure 2, we show the optical spectrum of the blazar 4FGL\nJ1205.8+3321 as an illustration. The 4LAC catalog reports\nthe redshift of this object as z= 2:085which is likely to be\nphotometric in nature and probably adopted from Richards\net al. (2009). On the other hand, the SDSS-DR16 spectrum\nclearly shows many broad emission lines securing the spec-\ntroscopic redshift as z= 1:007.\nWe acquired optical spectra of a few ( <50) sources from\nthe 6-degree Field Galaxy Survey (Jones et al. 2009). Since\nthese spectra are not flux calibrated, we adopted the wave-\nlength dependent conversion factor reported in Chen et al.\n(2018) to convert counts spectra in energy flux units.\n3.OPTICAL SPECTROSCOPIC ANALYSIS\nA good fraction of sources in our sample has optical spec-\ntrum taken from the SDSS-DR16, however, we do not use the\nemission/absorption line measurements obtained with the au-\ntomatic SDSS data reduction pipeline. A more sophisticated\napproach was needed to decompose the broad and narrow\ncomponents of the emission lines, to model and subtract the\nhost galaxy emission, and to consider the Fe complexes since\nthese are likely not to be properly taken into account by the\nSDSS pipeline. Moreover, a significant number of blazars do\n1https://automeris.io/WebPlotDigitizer/\n0.03 0.1 0.3 12345\nRedshift0306090Number of sourcesemission line\nabsorption line\nallFigure 1. The redshift distribution of Fermi blazars (black) present\nin our sample. Red solid and blue dotted histograms refer to objects\nwhose optical spectra are dominated by broad emission lines and\nhost galaxy absorption features, respectively.\nnot have SDSS spectra and a separate spectroscopic analysis\nwas required. During the data analysis, every optical spec-\ntrum was individually analyzed rather than by running any\nautomatic pipeline.\n3.1. Emission Line Spectrum\nThere are 674 \r-ray sources in our sample whose opti-\ncal spectra exhibit at least one of the broad emission lines\nH\u000b, H\f, Mg II, and C IV. To derive the spectral parameters\nassociated with these lines, e.g., full-width at half-maximum\nor FWHM, and continuum luminosities at 1350 ˚A, 3000 ˚A,\nand 5100 ˚A (L1350,L3000, andL5100, respectively), we have\nadopted the publicly available software PyQSOFit2(Guo\net al. 2018). The tool applies the spectral models and tem-\nplates to data following a \u001f2-based fitting technique (see\nalso, Guo et al. 2019; Shen et al. 2019). Here we briefly\ndescribe the adopted steps.\nThe quasar spectrum was brought to rest-frame and cor-\nrected for Galactic reddening following the extinction curve\nfrom Cardelli et al. (1989) and dust map of Schlegel et al.\n(1998). This correction was applied only to the SDSS data.\nFor spectra taken from the published articles, we do not ap-\nply Galactic extinction correction since it is usually done as a\npart of the data reduction prior to the publication. The spec-\ntrum was decomposed into the quasar and host galaxy com-\nponents following the principal component analysis method\npresented in Yip et al. (2004a,b). We have considered a\npower-law and a third-order polynomial along with optical\n2https://github.com/legolason/PyQSOFit4 P ALIYA ET AL .\n00.5 11.5 22.5 33.5 44.5\nredshift (4LAC)00.511.522.533.544.5redshift (this work)\n4000 5000 6000 7000 8000 9000 10000\nWavelength ( ˚A)246810Fλ(10−17erg cm−2s−1˚A−1)\nMg IIC III\nHβ4FGL J1205.8+3321 ( z= 1.007)\nFigure 2. Left: A comparison of the redshift values reported in the 4LAC catalog and this work. The dashed line denotes the equality of the\nplotted quantities. Right: The SDSS-DR16 spectrum of the \r-ray source 4FGL J1205.8+3321 whose redshift reported in the 4LAC catalog is\nz= 2:085. However, based on the detection of various broad emission lines, its correct redshift is z= 1:007.\n4000 4500 5000 5500 6000 6500 7000\nWavelength ( ˚A, rest-frame)200030004000500060007000Fλ(10−17erg s−1cm−2˚A−1)\nHα+NII\nSII6718,6732Hβ\n[OIII]HeII4687HγCaII39344FGL J1229.0+0202 ( z= 0.158)\ndata\nline\nline br\nline na\nconti\n4500 5000 5500 6000 6500 7000\nWavelength ( ˚A, rest-frame)140160180200220Fλ(10−17erg s−1cm−2˚A−1)\nHα+NII\nSII6718,6732Hβ\n[OIII]HeII46874FGL J2202.7+4216 ( z= 0.0686)\ndata\nline\nline br\nline na\nconti\nFigure 3. The optical spectra of prototype blazars 3C 273 or 4FGL J1229.0+0202 (left) and BL Lacertae or 4FGL J2202.7+4216 modeled with\nPyQSOFit . The spectral data is shown with the black line. Broad and narrow components of the emission line are represented by red and\ngreen lines and the modeled continuum is plotted with the orange line. Blue line is the sum of all the components. Horizontal grey dahses at the\ntop of the plots denote the line-free wavelength regions selected to model the continuum emission. The data are adopted from Torrealba et al.\n(2012, for 3C 273) and Vermeulen et al. (1995, for BL Lacertae).\nand UV Fe IItemplates (Boroson & Green 1992; Vestergaard\n& Wilkes 2001) to fit the line-free continuum over the en-\ntire spectrum. The best fitted continuum was then subtracted\nfrom the spectrum to acquire a line-only spectrum which was\nthen used to extract the spectral properties of H \u000b, H\f, Mg II,\nand C IVemission lines.\nThe H\u000bline was fitted in the wavelength range\n[6400, 6800] ˚A. The broad component of H \u000bwas modeled\nwith 3 Gaussians (FWHM >1200 km s\u00001; Hao et al. 2005),\nwhereas, we adopted a single Gaussian with the FWHM up-\nper limit of 1200 km s\u00001to reproduce the narrow H \u000bline.\nFurthermore, the flux ratio of the [N II]\u0015\u00156549, 6585 doublet\nwas fixed to 3 (Shen et al. 2011).\nH\fline fitting was carried out in the wavelength range\n[4640\u00005100] ˚A. We used 2 Gaussians (FWHM >1200\nkm s\u00001) and a single Gaussian (FWHM <1200 km s\u00001) fit\nto model the broad and narrow components of the H \fline,\nrespectively. We also fitted the narrow [O III]\u0015\u00154959, 5007lines with 2 Gaussians and their flux ratio was allowed to\nvary. However, we tied the velocity and width of the narrow\nH\fline to the core [O III] component.\nWe fitted Mg IIand C IVemission lines in the wavelength\nrange [2700\u00002900] ˚A and [1500\u00001700] ˚A, respectively. The\nbroad component of both lines were modeled with 2 Gaus-\nsians (FWHM >1200 km s\u00001) and we used a single Gaus-\nsian function (FWHM <1200 km s\u00001) to fit their narrow\nemission line components.\nPyQSOFit determines the uncertainties in the derived pa-\nrameters employing a Monte-Carlo technique. In particular,\na random Gaussian fluctuation with a zero mean and disper-\nsion equals to the uncertainty measured at the given pixel\nwas added to the observed spectral flux at every pixel and\n50 mock spectra were created and fitted with the same strat-\negy as that adopted to model the real data. This exercise\nwas repeated 20 times and uncertainties were calculated asBLAZAR ’SCENTRAL ENGINE 5\nTable 2. The spectral parameters associated with the\nH\u000bemission line.\n4FGL name redshift FWHM LH\u000b\n[1] [2] [3] [4]\nJ0001.5+2113 0.439 1629 \u000684 42.942 \u00060.016\nJ0006.3 \u00000620 0.347 8991 \u00065494 42.782 \u00060.189\nJ0013.6+4051 0.256 2049 \u00061739 41.277 \u00060.205\nJ0017.5 \u00000514 0.227 2073 \u0006982 42.926 \u00060.090\nJ0049.6 \u00004500 0.121 4298 \u00063904 41.658 \u00060.117\nNOTE—Column information are as follows: Col.[1]: 4FGL\nname; Col.[2]: redshift; Col.[3]: FWHM of H \u000bline, in\nkm s\u00001; and Col.[4]: log-scale H \u000bline luminosity, in\nerg s\u00001. (This table is available in its entirety in a machine-\nreadable form in the online journal. A portion is shown here for\nguidance regarding its form and content.)\nTable 3. The spectral parameters associated with the H \femission line.\n4FGL name redshift L5100 ˚AFWHM LH\f\n[1] [2] [3] [4] [5]\nJ0001.5+2113 0.439 44.577 \u00060.003 2114 \u0006483 42.029 \u00060.054\nJ0006.3 \u00000620 0.347 44.757 \u00060.004 7051 \u00064699 42.004 \u00060.227\nJ0010.6+2043 0.598 44.867 \u00060.003 2593 \u0006207 43.047 \u00060.048\nJ0014.3 \u00000500 0.791 44.951 \u00060.008 1619 \u0006653 42.446 \u00060.060\nJ0017.5 \u00000514 0.227 44.347 \u00060.005 3294 \u0006487 42.302 \u00060.115\nNOTE—Column information are as follows: Col.[1]: 4FGL name; Col.[2]: redshift;\nCol.[3]: log-scale continuum luminosity at 5100 ˚A; Col.[4]: FWHM of H \fline,\ninkm s\u00001; and Col.[5]: log-scale H \fline luminosity, in erg s\u00001. (This table is\navailable in its entirety in a machine-readable form in the online journal. A portion\nis shown here for guidance regarding its form and content.)\nthe semi-amplitude of the range covering the 16th \u000084th per-\ncentiles of the parameter distribution from the trials.\nThe tool PyQSOFit is primarily developed to analyze the\nSDSS data and hence requires the pixel scale of the optical\nspectrum to be same as that of SDSS, i.e., 10\u00004in log-space3.\nSince a significant fraction of the optical spectra in our sam-\nple is collected from the literature and does not have the req-\nuisite binning, we rebinned them using a simple linear inter-\npolation to have the pixel scale same as that of the SDSS data.\nFurthermore, since we do not have the flux uncertainty mea-\nsurements in such cases, we conservatively assumed an un-\ncertainty of 25% of the measured fluxes, a typical value asso-\nciated with the ground-based optical observations (cf. Healey\net al. 2008; Shaw et al. 2012).\n3https://www.sdss.org/dr16/spectro/spectro basics/Table 4. The spectral parameters associated with the Mg IIemission line.\n4FGL name redshift L3000 ˚AFWHM LMgII\n[1] [2] [3] [4] [5]\nJ0001.5+2113 0.439 44.706 \u00060.002 1719 \u0006165 42.503 \u00060.032\nJ0004.4 \u00004737 0.880 45.345 \u00060.003 2965 \u0006913 42.885 \u00060.099\nJ0010.6+2043 0.598 44.799 \u00060.005 2222 \u000699 43.027 \u00060.017\nJ0011.4+0057 1.491 45.589 \u00060.003 3382 \u0006245 43.365 \u00060.045\nJ0013.6 \u00000424 1.076 44.898 \u00060.003 2383 \u0006221 42.818 \u00060.080\nNOTE—Column information are as follows: Col.[1]: 4FGL name; Col.[2]: red-\nshift; Col.[3]: log-scale continuum luminosity at 3000 ˚A; Col.[4]: FWHM of\nMgIIline, in km s\u00001; and Col.[5]: log-scale Mg IIline luminosity, in erg s\u00001.\n(This table is available in its entirety in a machine-readable form in the online\njournal. A portion is shown here for guidance regarding its form and content.)\nTable 5. The spectral parameters associated with the C IVemission line.\n4FGL name redshift L1350 ˚AFWHM LCIV\n[1] [2] [3] [4] [5]\nJ0004.3+4614 1.810 45.615 \u00060.001 2652 \u0006300 44.126 \u00060.031\nJ0011.4+0057 1.491 45.558 \u00060.001 4430 \u0006437 43.901 \u00060.014\nJ0016.2 \u00000016 1.577 45.445 \u00060.001 4648 \u0006216 43.809 \u00060.073\nJ0016.5+1702 1.721 45.304 \u00060.000 5722 \u0006275 43.849 \u00060.021\nJ0028.4+2001 1.553 45.499 \u00060.001 2631 \u0006128 43.927 \u00060.012\nNOTE—Column information are as follows: Col.[1]: 4FGL name; Col.[2]: red-\nshift; Col.[3]: log-scale continuum luminosity at 1350 ˚A; Col.[4]: FWHM of\nCIVline, in km s\u00001; and Col.[5]: log-scale C IVline luminosity, in erg s\u00001.\n(This table is available in its entirety in a machine-readable form in the online\njournal. A portion is shown here for guidance regarding its form and content.)\nIn Figure 3, we show an example of the emission line\nfitting done on the optical spectra of the prototype blazars\n3C 273 (4FGL J1229.0+0202, z= 0:158) and BL Lacertae\n(4FGL J2202.7+4216, z= 0:0686 ). The spectral parame-\nters derived for H \u000b, H\f, Mg II, and C IVemission lines are\nprovided in Table 2, 3, 4, and 5, respectively.\n3.2. Absorption Line Spectrum\nThe optical spectra of 346 \r-ray blazars show prominent\nabsorption lines, e.g., Ca H&K doublet, primarily originat-\ning from the stellar population in the host galaxy. It has been\nshown in various works that the mass of the central black\nhole significantly correlates with the stellar velocity disper-\nsion (\u001b\u0003, cf. Ferrarese & Merritt 2000; G ¨ultekin et al. 2009;\nKormendy & Ho 2013). Therefore, we have used the penal-\nized PiXel Fitting tool ( pPXF ; Cappellari & Emsellem 2004)\nto derive\u001b\u0003for blazars present in the sample. This software\nworks in pixel space and uses a maximum penalized like-\nlihood approach to calculate the line-of-sight velocity dis-6 P ALIYA ET AL .\n3040Fλ(10−17erg cm−2s−1)4FGL J1132.2-4736 ( z= 0.23) [σ∗= 287.99±6.97 km s−1]\n3500 3750 4000 4250 4500 4750 5000 5250 5500\nWavelngth ( ˚A, rest-frame)−101\n152025Fλ(10−17erg cm−2s−1)4FGL J2220.5+2813 ( z= 0.15) [σ∗= 261.88±10.52 km s−1]\n3500 4000 4500 5000 5500 6000 6500 7000\nWavelngth ( ˚A, rest-frame)−101\nFigure 4. The optical spectra of two Fermi -LAT detected blazars (black) along with the results of the stellar population synthesis done using\npPXF (red). In both plots, the bottom panel refers to the residual of the fit. The wavelength regions excluded from the fit to mask emission\nlines are highlighted with the grey shaded columns. The spectroscopic data of 4FGL J1132.2 \u00004736 is taken from Pe ˜na-Herazo et al. (2017),\nwhereas, that of 4FGL J2220.5+2813 is adopted from SDSS-DR16 (Ahumada et al. 2020).\ntribution (LOSVD) from kinematic data (Merritt 1997). A\nlarge set of stellar population synthesis models (adopted from\nVazdekis et al. 2010) with spectral resolution of FWHM = 2.5\n˚A and the wavelength coverage of [3525, 7500] ˚A (S ´anchez-\nBl´azquez et al. 2006) were used in this work. The pPXF code\nfirst creates a template galaxy spectrum by convolving the\nstellar population models with the parameterized LOSVD.\nTo mimic the non-thermal power-law contribution from the\ncentral nucleus, we further added a fourth-order Legendre\npolynomial to the model. The regions of bright emission\nlines were masked prior to the fitting. The model was then\nfitted on the rest-frame galaxy spectrum and \u001b\u0003and associ-\nated 1\u001buncertainty were derived from the best-fit spectrum.\nThe whole analysis was carried out on a case-by-case basis\nto achieve the best results. This is because the optical spectra\nof many sources reveal strong telluric lines which may over-\nlap with the absorption lines arising from the host galaxy and\nthus needed to be avoided. Therefore, although we attempted\na fit in the full [3525, 7500] ˚A wavelength range, it was not\nalways possible. The examples of this analysis are shown in\nFigure 4 and the derived \u001b\u0003values are provided in Table 6.\nUnlike broad line blazars where emission line luminosi-\nties can be used to infer the BLR luminosity, it is not pos-\nsible estimate the latter for absorption line systems since\nbroad emission lines are not detected. Therefore, we have\nderived 3\u001bupper limit in the H \f(or Mg II, depending on the\nsource redshift and wavelength coverage) line luminosity by\nadopting the following steps. The observed spectrum was\nfirst brought to the rest-frame and the host galaxy component\nwas subtracted using PyQSOFit . In the wavelength range\n[4700\u00005000] ˚A or [2650\u00002950] ˚A for H\for Mg II, respec-\ntively, we then fitted a power-law (reproducing continuum)\nplus a Gaussian function (mimicking the emission line) with\na variable luminosity while keeping the FWHM fixed to 4000\nkm s\u00001, a value typically observed in blazars (e.g., Shaw\net al. 2012). Based on the derived \u001f2, we considered the up-\nper limit to the line luminosity when \u001f2> \u001f2(99.7%), i.e.,\nat 3\u001bconfidence level. We show an example of the adopted\nmethod in Figure 5 and report the upper limits in Table 6.Table 6. The stellar velocity dispersion and line luminosity up-\nper limit measurements for 346 blazars.\n4FGL name redshift \u001b\u0003Lline;UL\n[1] [2] [3] [4]\nJ0003.2+2207 0.100 197.61 \u00067.87 40.34\nJ0006.4+0135 0.787 385.99 \u000661.79 42.41\nJ0013.9\u00001854 0.095 459.46 \u000619.50 40.87\nJ0014.2+0854 0.163 296.52 \u000610.53 40.97\nJ0015.6+5551 0.217 467.26 \u000622.05 41.65\nNOTE—Column information are as follows: Col.[1]:\n4FGL name; Col.[2]: redshift; Col.[3]: stellar velocity\ndispersion (\u001b\u0003) inkm s\u00001; and Col.[4]: 3 \u001bupper limit\nin H\fline luminosity (log-scale, in erg s\u00001), except for\nsources 4FGL J0006.4+0135, 4FGL J0204.0-3334, and\n4FGL J1146.0-0638, for which we quote the Mg IIline\nluminosity upper limit. (This table is available in its en-\ntirety in a machine-readable form in the online journal.\nA portion is shown here for guidance regarding its form\nand content.)\n4.BLACK HOLE MASS AND DISK LUMINOSITY\nMEASUREMENTS\n4.1. Emission Line Black Hole Mass\nWe have derived MBHof emission line blazars from the\nsingle-epoch optical spectra assuming that BLR is virialized,\nbroad line FWHM represents the virial velocity, and contin-\nuum luminosity can be considered as a proxy for the BLR\nradius. The virial MBHcan be calculated using the follow-\ning equation (Shen et al. 2011):\nlog\u0012MBH\nM\f\u0013\n=\u000b+\flog\u0012\u0015L\u0015\n1044erg s\u00001\u0013\n+2 log\u0012FWHM\nkm s\u00001\u0013\n(1)BLAZAR ’SCENTRAL ENGINE 7\n4700 4750 4800 4850 4900 4950 5000\nλ(Hz, rest-frame)036912Fλ(10−17erg cm−2s−1˚A−1)\n40.8 41 41.2\nlogLHβ[erg s−1]0246810 χ2−χ2\nmin4FGL J0837.3+1458 ( z= 0.1523)\nFigure 5. A demonstration of the technique adopted to derive the 3 \u001b\nupper limit on the emission line luminosity. Blue thin and black dot-\nted lines refer to the observed spectrum and power-law continuum,\nrespectively. We also plot Gaussian functions with variable line lu-\nminosity. The right panel shows the derived \u001f2as a function of the\nline (H\fin this case) luminosity. The vertical line corresponds to\nthe line luminosity beyond which \u001f2>\u001f2(99.7%).\nwhere\u0015L\u0015is the continuum luminosity at 5100 ˚A (for H\f),\n3000 ˚A (for Mg II), and 1350 ˚A (for C IV). The calibration co-\nefficients\u000band\fare taken from McLure & Dunlop (MD04;\n2004) for H\fand Mg IIand Vestergaard & Peterson (VP06;\n2006) for C IVlines and have the following values\n(\u000b;\f) =8\n><\n>:(0:672;0:61);H\f(MD04)\n(0:505;0:62);MgII(MD04)\n(0:660;0:53);CIV(VP06); (2)\nShen et al. (2011) also provided the following empirical\nrelation to estimate MBHfrom the H\u000bline FWHM and lu-\nminosity:\nlog\u0012MBH\nM\f\u0013\nH\u000b= 0:379 + 0:43 log\u0012LH\u000b\n1042erg s\u00001\u0013\n+ 2:1 log\u0012FWHM H\u000b\nkm s\u00001\u0013\n; (3)\nwhereLH\u000bis the total H \u000bline luminosity. For objects with\nmore than one MBHmeasurements, we have taken the geo-\nmetric mean.\nThe optical spectral continuum of blazars could be signif-\nicantly contaminated from the non-thermal jetted emission\nwhich can affect the MBHestimation. To address this issue,\nwe also derived MBHvalues using emission line parameters\nalone. This was done by replacing continuum luminosity in\nEquation 1 with the line luminosity and adopting the cali-\nbration coefficients ( \u000b;\f) as (1.63, 0.49), (1.70, 0.63), and\n(1.52, 0.46) for H \f, Mg II, and C IVlines, respectively, fol-\nlowing Shaw et al. (2012). We compared the masses cal-\nculated from two approaches in Figure 6. There are indi-\ncations that MBHcomputed from virial approach is slightly\nhigher than that calculated from H \fline alone (Figure 6, left\npanel) which could be due to contamination from the host\ngalaxy emission as also suggested by Shaw et al. (2012). ForMgIIand C IVlines, the overall impact of the jetted emis-\nsion on theMBHmeasurement is negligible. Therefore, the\nMBHvalues for emission line blazars computed from the tra-\nditional virial technique (Equation 1) are used in this work as\nthis may also allow a comparison study with other non-blazar\nclass of AGN. We report them in Table 7.\n4.2. Absorption Line Black Hole Mass\nWe used the following empirical relation to compute MBH\nfrom the measured stellar velocity dispersion (G ¨ultekin et al.\n2009)\nlog\u0012MBH\nM\f\u0013\n= (8:12\u00060:08)+(4:24\u00060:41) log\u0010\u001b\u0003\n200 km s\u00001\u0011\n;\n(4)\nThe estimated masses of 346 blazars are provided in Ta-\nble 8.\n4.3. Bulge Luminosity Black Hole Mass\nFrom the literature, we were able to obtain appar-\nent/absolute R- orK-band magnitudes of the host galaxy\nbulge for 47 blazars. The following equations waere used\nto deriveMBHfrom the bulge luminosity (Graham 2007)\nlog\u0012MBH\nM\f\u0013\n=(\n(\u00000:38\u00060:06)(MR+ 21) + (8:11\u00060:11)\n(\u00000:38\u00060:06)(MK+ 24) + (8:26\u00060:11)\n(5)\nwhereMRandMKare the absolute magnitudes of the host\ngalaxy bulge in R- andK-bands, respectively. The derived\nMBHvalues are reported in Table 9.\n4.4. Accretion Disk Luminosity\nFrom the emission line luminosities or 3 \u001bupper limits, we\ncomputed the BLR luminosity (or 3 \u001bupper limits) as fol-\nlows: we assigned a reference value of 100 to Ly \u000bemission\nand summed the line ratios (with respect to Ly \u000b) reported in\nFrancis et al. (1991) and Celotti et al. (1997, for H \u000b) giv-\ning the total BLR fraction hLBLRi= 555:77\u00185:6Ly\u000b.\nThe BLR luminosity can then be derived using the following\nequation\nLBLR=Lline\u0002hLBLRi\nLrel:frac:(6)\nwhereLlineis the emission line luminosity and Lrel:frac:is\nthe line ratio, 77, 22, 34, and 63 for H \u000b, H\f, Mg II, and\nCIVlines, respectively (Francis et al. 1991; Celotti et al.\n1997). When more than one line luminosity measurements\nwere available, we took their geometric mean to derive the\naverageLBLR and then calculated Ldiskfrom theLBLR as-\nsuming 10% BLR covering factor. The computed Ldiskval-\nues and 3\u001bupper limits are provided in Table 7 and 8, re-\nspectively.8 P ALIYA ET AL .\n7 8 9 1078910log Mass from lines alone (M ⊙)Hβ\n7 8 9 10\nlog Traditional virial mass (M ⊙)Mgii\n7 8 9 10Civ\nFigure 6. A comparison of the MBHvalues derived from the traditional virial technique using line FWHM and continuum luminosity with that\nestimated from the emission line alone. The dashed line represents the equality of the plotted quantities.\nTable 7. The central engine and SED properties of emission line blazars.\n4FGL name MBH Ldisk\u0017peak\nsyn\u0017Fpeak\n\u0017;syn\u0017peak\nIC\u0017Fpeak\n\u0017;IC CD\n[1] [2] [3] [4] [5] [6] [7] [8]\nJ0001.5+2113 7.54 \u00060.07 44.65\u00060.02 13.81 -11.97 20.64 -10.48 30.90\nJ0004.3+4614 8.36 \u00060.10 46.07\u00060.03 12.35 -12.49 21.35 -11.70 6.17\nJ0004.4\u00004737 8.28\u00060.27 45.10\u00060.10 13.01 -11.62 21.37 -11.24 2.40\nJ0006.3\u00000620 8.93\u00060.40 44.52\u00060.15 12.92 -11.09 19.66 -12.04 0.11\nJ0010.6+2043 7.86 \u00060.04 45.35\u00060.03 12.42 -12.12 22.60 -11.97 1.41\nNOTE—Column information are as follows: Col.[1]: 4FGL name; Col.[2]: log-scale mass\nof the central black hole, in M\f; Col.[3]: log-scale accretion disk luminosity, in erg s\u00001;\nCol.[4] and [5]: log-scale synchrotron peak frequency and corresponding flux, in Hz and\nerg cm\u00002s\u00001, respectively; Col.[6] and [7]: log-scale inverse Compton peak frequency\nand corresponding flux, in Hz and erg cm\u00002s\u00001, respectively; and Col.[8]: Compton dom-\ninance. (This table is available in its entirety in a machine-readable form in the online\njournal. A portion is shown here for guidance regarding its form and content.)\nTable 9 . The mass of the central black holes for 47\nblazars derived from the host galaxy bulge luminos-\nity.\n4FGL name redshift MBH\n[1] [2] [3]\nJ0037.8+1239 0.089 8.64 \u00060.14\nJ0050.7\u00000929 0.635 8.85 \u00060.18\nTable 9 continuedBLAZAR ’SCENTRAL ENGINE 9\nTable 8. The central engine and SED properties of absorption line blazars.\n4FGL name MBHLdisk\u0017peak\nsyn\u0017Fpeak\n\u0017;syn\u0017peak\nIC\u0017Fpeak\n\u0017;IC CD\n[1] [2] [3] [4] [5] [6] [7] [8]\nJ0003.2+2207 8.10 \u00060.11 42.74 15.15 -12.23 22.16 -12.91 0.21\nJ0006.4+0135 9.33 \u00060.33 44.62 15.96 -12.49 22.93 -12.65 0.69\nJ0013.9\u00001854 9.65\u00060.19 43.27 17.43 -11.39 23.95 -12.47 0.08\nJ0014.2+0854 8.85 \u00060.12 43.37 15.64 -12.21 22.25 -12.58 0.43\nJ0015.6+5551 9.68 \u00060.19 44.05 17.11 -11.61 24.89 -12.07 0.35\nNOTE—Column information are same as in Table 7 except Column 3 where 3 \u001b\nupper limits on the Ldiskare reported. (This table is available in its entirety in a\nmachine-readable form in the online journal. A portion is shown here for guidance\nregarding its form and content.)\nTable 9 (continued)\n4FGL name redshift MBH\n[1] [2] [3]\nJ0109.1+1815 0.145 9.09 \u00060.23\nJ0123.1+3421 0.272 8.68 \u00060.14\nJ0159.5+1046 0.195 8.51 \u00060.14\nJ0202.4+0849 0.629 8.85 \u00060.16\nJ0208.6+3523 0.318 8.62 \u00060.14\nJ0214.3+5145 0.049 8.55 \u00060.13\nJ0217.2+0837 0.085 8.40 \u00060.13\nJ0238.4\u00003116 0.233 8.73 \u00060.15\nJ0303.4\u00002407 0.266 8.95 \u00060.17\nJ0319.8+1845 0.190 8.66 \u00060.14\nJ0340.5\u00002118 0.233 8.29 \u00060.11\nJ0349.4\u00001159 0.188 8.60 \u00060.13\nJ0422.3+1951 0.512 8.57 \u00060.14\nJ0424.7+0036 0.268 8.76 \u00060.16\nJ0507.9+6737 0.416 9.00 \u00060.18\nJ0509.6\u00000402 0.304 8.88 \u00060.16\nJ0617.7\u00001715 0.098 8.83 \u00060.20\nJ0623.9\u00005259 0.513 8.85 \u00060.16\nJ0712.7+5033 0.502 8.49 \u00060.15\nJ0757.1+0956 0.266 8.53 \u00060.13\nJ0814.6+6430 0.239 8.23 \u00060.14\nJ1103.6\u00002329 0.186 9.01 \u00060.18\nTable 9 continuedTable 9 (continued)\n4FGL name redshift MBH\n[1] [2] [3]\nJ1136.4+7009 0.045 8.62 \u00060.14\nJ1217.9+3007 0.130 8.75 \u00060.15\nJ1257.2+3646 0.530 8.92 \u00060.19\nJ1315.0\u00004236 0.105 8.56 \u00060.13\nJ1359.8\u00003746 0.334 8.77 \u00060.16\nJ1501.0+2238 0.235 8.81 \u00060.16\nJ1517.7+6525 0.702 10.11 \u00060.31\nJ1535.0+5320 0.890 9.64 \u00060.24\nJ1548.8\u00002250 0.192 8.69 \u00060.16\nJ1643.5\u00000646 0.082 8.43 \u00060.14\nJ1728.3+5013 0.055 8.28 \u00060.11\nJ1748.6+7005 0.770 10.02 \u00060.30\nJ1757.0+7032 0.407 8.78 \u00060.18\nJ1813.5+3144 0.117 8.02 \u00060.11\nJ2005.5+7752 0.342 8.77 \u00060.16\nJ2009.4\u00004849 0.071 8.90 \u00060.17\nJ2039.5+5218 0.053 8.11 \u00060.11\nJ2042.1+2427 0.104 8.76 \u00060.16\nJ2055.4\u00000020 0.440 8.38 \u00060.15\nJ2143.1\u00003929 0.429 8.68 \u00060.15\nJ2145.7+0718 0.237 8.70 \u00060.15\nJ2252.0+4031 0.229 8.72 \u00060.15\nTable 9 continued10 P ALIYA ET AL .\nTable 9 (continued)\n4FGL name redshift MBH\n[1] [2] [3]\nJ2359.0\u00003038 0.165 8.67 \u00060.14\nNOTE—Column information are as follows:\nCol.[1]: 4FGL name; Col.[2]: redshift; and\nCol.[3]: log-scale black hole mass, in M\f.\n4.5. Literature Collection\nThere are 10 blazars in our sample whose optical spectra\ncould not be retrieved, however, their emission line pa-\nrameters or MBHvalues have been published (e.g., Bald-\nwin et al. 1981; Chen et al. 2015). We were able to col-\nlect/estimate both MBHandLdisk values for 10 such ob-\njects and report them in Table 7. These objects are: 4FGL\nJ0243.2\u00000550, 4FGL J0509.4+0542, 4FGL J0501.2 \u00000158,\n4FGL J0525.4\u00004600, 4FGL J0836.5 \u00002026, 4FGL\nJ1214.6\u00001926, 4FGL J1557.9 \u00000001, 4FGL J1626.0 \u00002950,\n4FGL J2056.2\u00004714, and 4FGL J2323.5 \u00000317.\nFinally, we tabulate references of all the research articles\nused to collect optical spectra or spectral parameters for 1077\nblazars in Appendix (Table 10).\n5.PROPERTIES OF THE CENTRAL ENGINE\nIn the left panel of Figure 7, we show the distribu-\ntion of the MBHvalues estimated from different meth-\nods described above. The average MBHfor the whole\npopulation ishlogMBH;allM\fi= 8:60. Considering\nemission and absorption line systems separately, we get\nhlogMBH;emiM\fi= 8:48andhlogMBH;absM\fi= 8:80,\nrespectively. The dispersion for all three distributions is sim-\nilar,\u00180.6, when fitted with a Gaussian function. Though the\nspread is large, there are tentative evidences hinting the as-\nsociation of more massive black holes with the absorption\nline systems, i.e., objects lacking broad emission lines or BL\nLacs. To understand it further, we compared MBHfor a sub-\nsample of blazars with 0:3\u0014z\u00140:7. This was done to\navoid any redshift dependent selection effect and to com-\npare a similar number of sources from emission (139) and\nabsorption line (110) blazar samples. In the middle panel\nof Figure 7, we show the histograms of MBHvalues derived\nfor the two populations in the opted redshift range. The av-\nerage masses for emission and absorption line blazars are\nhlogMBH;M\fi= 8:13 and 8:70, respectively. Though the\nspread is large ( \u001b\u00180.55), this finding hints that BL Lac ob-\njects do tend to host more massive black holes than broad\nline FSRQs. A Kolmogorov-Smirnov test was carried out to\ndetermine whether two populations are distinctively differ-\nent. The derived p-value is 1:9\u000210\u00008, suggesting that thenull-hypothesis of both samples belonging to the same source\npopulation can be rejected at >5\u001bconfidence level.\nThe observation of more massive black holes residing in\nBL Lac objects is likely to be connected with the cosmic\nevolution of blazars (B ¨ottcher & Dermer 2002; Cavaliere &\nD’Elia 2002; Ajello et al. 2014). According to the proposed\nevolutionary sequence, high-power ( Lbol:>1046erg s\u00001)\nblazars, primarily FSRQs, evolve to their low-power counter-\nparts, mainly BL Lacs, over cosmological timescales. Such\na transition is likely caused by a gradual depletion of the\ncentral environment by accretion onto the black hole. In\nother words, luminous broad emission line blazars evolve\nto low-luminosity objects exhibiting quasi-featureless spec-\ntra as their accretion mode changes from being radiatively ef-\nficient to advection dominated accretion flow (Narayan et al.\n1997). If so, one would expect the black holes hosted in the\nlatter to be more massive than that found in the former since\nthey would keep growing as more and more mass is dumped\nby accretion.\nWe show the variation of MBHas a function of the blazar\nredshift in the right panel of Figure 7. There is a little overlap\nsince all of the absorption line objects are located below z=\n0:85, whereas, a major fraction of the emission line sources\nare above it (see also Figure 1). There is a trend of more\nmassive black holes being located at higher redshifts among\nthe broad line sources. This is likely a selection effect since\nwe expect to detect only the most luminous systems, hence\nthe most massive black holes, at cosmological distances.\nThe left panel of Figure 8 shows the Ldiskdistribution for\nbroad line objects. For a comparison, we also plot the his-\ntogram of the 3 \u001bupper limits on the Ldisk measured for\nblazars lacking emission lines. As expected, strong emission\nline systems have more luminous accretion disks. The dif-\nference between the two populations becomes clearer when\nwe derive the Ldisk values in Eddington units. The re-\nsults are shown in the right panel of Figure 8 and reveal\na bi-modality. A major fraction of broad line objects have\nLdisk/LEdd&0:01, whereas,Ldiskof absorption line sys-\ntems have values .1% ofLEddwhich is expected to be even\nlower since the plotted quantity corresponds to 3 \u001bluminos-\nity upper limit. Interestingly, a low-level of accretion activ-\nity (Ldisk/LEdd.0:01) is noticed from a few emission line\nblazars. These are objects typically classified as BL Lacs,\nhowever, have revealed faint broad emission lines in their\noptical spectra taken during a low jet activity state (cf. Ver-\nmeulen et al. 1995).\n6.A COMPTON DOMINANCE BASED\nCLASSIFICATION\nA blazar classification based on Ldiskin units ofLEddis\nphysically intuitive (Ghisellini et al. 2011), however, extend-\ning it to lineless sources remained a challenge. If the opticalBLAZAR ’SCENTRAL ENGINE 11\n7 8 9 10\nlogMBH[M⊙]04080120160Number of sourcesemission line\nabsorption line\nall\n6 7 8 9 10\nlogMBH[M⊙, 0.3≤z≤0.7]081624Number of sourcesemission line\nabsorption line\n0.03 0.1 0.3 1 2345\nRedshift678910logMBH[M⊙]\nabsorption line\nemission line bulge luminosity\nFigure 7. Left: The histograms of the MBHfor broad emission line blazars (red filled), absorption line systems (blue dotted), and the whole\nsample (black empty) including MBHderived from the host galaxy bulge luminosity. Middle: The distributions of MBHfor blazars in the\nredshift range 0:3\u0014z\u00140:7. Right: The redshift dependence of the MBHderived from various methods as labeled.\n41 42 43 44 45 46 47 48\nlogLdisk[erg s−1]0306090120150180Number of sourcesemission line\nabsorption line (UL)\n-5 -4 -3 -2 -1 0\nlogLdisk/LEdd0306090120Number of sourcesemission line\nabsorption line (UL)\nFigure 8. Left: The luminosity of the accretion disk for emission line blazars is shown with the red filled histogram. The blue dotted histogram\nfor absorption line objects, on the other hand, refers to 3 \u001bupper limit in Ldisk. Right: Same as left but with Ldiskplotted in Eddington units.\nspectrum of a FSRQ is dominated by the non-thermal jet-\nted radiation which swamps out broad emission lines, it may\nnot be possible to characterize its state of accretion even if it\nhosts a luminous accretion disk. Therefore, it is necessary to\nidentify some other observational features which can indicate\nthe state of accretion or, in other words, show a correlation\nwithLdisk/LEdd. Below we show that Compton dominance\nor CD can be considered as one such parameter to reveal the\nphysics of the central engine in beamed AGN.\nAccording to the canonical picture of the blazar leptonic\nemission models, radio-to-optical-UV radiation is dominated\nby the synchrotron emission, though the big blue bump aris-\ning from the accretion disk has also been observed. If the\naccretion process is radiatively efficient, it will illuminate\nthe BLR and dusty torus enabling a photon-rich environ-\nment surrounding the jet. In such objects, the high-energy\nX- and\r-ray emission is mainly powered by the externalCompton mechanism (see, e.g., Sikora et al. 1994). Due\nto an additional beaming factor associated with the external\nCompton process (Dermer 1995), the bolometric jet emis-\nsion will be dominated by the high-energy \r-ray emission\nleading to the observation of a Compton dominated SED\n(e.g., Paliya 2015). On the other hand, the overall SED of\na blazar with an intrinsically low-level of accretion activity\n(Ldisk/LEdd<0:01), will be synchrotron dominated. This\nis because, the \r-ray emission in such objects originates pri-\nmarily via synchrotron self Compton mechanism (SSC; cf.\nFinke et al. 2008) since the environment surrounding the jet\nlacks the external photons needed for the external Compton\nprocess. To summarize, one can get a fair idea of the ac-\ncretion level of a blazar by estimating CD from its multi-\nfrequency SED. Having measured the Ldisk in Eddington12 P ALIYA ET AL .\n108101110141017102010231026\nν(Hz)10−1510−1410−1310−1210−11νFν(erg cm−2s−1)\n4FGL J0532.6+0732 ( z= 1.254)\n108101110141017102010231026\nν(Hz)10−1510−1410−1310−1210−11νFν(erg cm−2s−1)\n4FGL J0602.0+5315 ( z= 0.0522)\nFigure 9. The multi-frequency SEDs of two Fermi -LAT detected blazars (grey data points) when fitted with second-degree polynomials (left)\nand a synchrotron-SSC model (right). The fitted functions are shown with the blue dotted lines. In the right panel, the bump seen at IR-optical\nwavelengths arises from the host galaxy.\n0.01 0.1 1 10 100\nCompton dominance020406080100120Number of sourcesemission line\nabsorption line\nFigure 10. The Compton dominance histograms of the emission\n(red filled) and absorption line (blue dotted) blazars.\nunits for a large sample of blazars, we, therefore, next at-\ntempted determining their CD as described below.\n6.1. Compton Dominance Measurement\nWe collected broadband spectral data of blazars present in\nour sample from Space Science Data Center (SSDC) SED\nbuilder tool4. While doing so, we did not consider United\nStates Naval Observatory data and flux values reported by the\nCatalina Real-Time Transient Survey since their fluxes are\n4https://tools.ssdc.asi.it/SED/often outside the range of the other data taken at the same fre-\nquency. To improve the data coverage at X- and \r-ray bands,\nwe included the flux measurements from the 2ndSwift X-\nray Point Source catalog (2SXPS; Evans et al. 2020) and the\n4FGL-DR2 catalog (Abdollahi et al. 2020). The synchrotron\nand inverse Compton peak frequencies and corresponding\nflux values were then estimated by fitting a second-degree\npolynomial to both SED peaks using the built-in function\nprovided in the SSDC SED builder tool. In many blazars,\nthe soft X-ray spectrum is still dominated by the synchrotron\nemission and forms the tail-end of the low-energy peak. In\nsuch objects, it may not be possible to accurately constrain\nthe high-energy peak with only Fermi -LAT data. The multi-\nwavelength SEDs of such blazars, therefore, were fitted with\na synchrotron-SSC model assuming a log-parabolic electron\nenergy distribution (Tramacere et al. 2009). Furthermore,\nthe SEDs of powerful FSRQs often exhibit the accretion\ndisk bump at optical-UV energies and many high-frequency\npeaked BL Lacs show a host galaxy emission at IR-optical\nwavelengths. Such features were avoided during the fit. The\nexamples of this analysis are illustrated in Figure 9. The CD\nwas derived by taking the ratio of the high- and low-energy\npeak luminosities, which is equivalent to the ratio of their\npeak fluxes since it is essentially a redshift independent quan-\ntity. The estimated values are provided in Table 7 and 8 and\nwe plot the CD histograms for both emission and absorption\nline blazars in Figure 10. As can be seen, a bi-modality ap-\npears with the broad emission line blazars have an average\nCD>1 (hCDemii= 5:85), whereas, absorption line objects\nhave less Compton dominated SED ( hCDabsi= 0:46). This\nfinding is in line with the hypothesis of broad line blazars be-\ning more Compton dominated sources compared to lineless\nBL Lac objects as discussed above.BLAZAR ’SCENTRAL ENGINE 13\n-4 -3 -2 -1 0\nlogLdisk/LEdd0.1110100Compton dominance\nabsorption line\nemission line\nFigure 11. The Compton dominance as a function of the Ldiskin Eddington units. For the sake of clarity, uncertainties associated with the\nMBHvalues were not propagated while calculating Ldisk/LEdd.\n6.2. Compton Dominance and Ldisk/LEdd\nIn Figure 11, we show the distribution of CD as a function\nofLdiskin Eddington units. A positive correlation is appar-\nent with highly accreting objects tend to have larger CD. To\nquantify the strength of the correlation, we adopted the astro-\nnomical survival analysis package (ASURV; Lavalley et al.\n1992) which takes into account the upper/lower limits in the\ndata (Isobe et al. 1986). The derived Spearman’s correlation\ncoefficient is \u001a= 0:76with probability of no correlation or\nPNC<1\u000210\u000010indicating a positive correlation with high\n(>5\u001b) confidence. This strengthens our argument that CD\ncan be considered as a good proxy for the accretion rate in\nblazars. Furthermore, a linear regression analysis carried out\nusing ASURV software resulted in the following empirical\nrelation connecting the two variables\nlog CD = (1 :16\u00060:03)+(0:74\u00060:02) log\u0012Ldisk\nLEdd\u0013\n:(7)Figure 11 reveals a clear distinction of two groups in re-\ngions defined by CD>1,Ldisk/LEdd>0.01 for broad line\nblazars and CD<1,Ldisk/LEdd<0.01 for absorption line\nsystems. The two populations overlap smoothly for interme-\ndiate values of CD and Ldiskin Eddington units. Based on\nthis finding, we propose that blazars can be classified as high-\nCompton dominated (HCD) with CD >1 and low-Compton\ndominated (LCD) sources for CD <1. Moreover, blazars\nhaving CD>1,Ldisk/LEdd>0.01 should be identified as\nFSRQs and those with CD<1,Ldisk/LEdd<0.01 as BL\nLacs. One good example is the blazar 4FGL J0238.6+1637\nor AO 0235+164 ( z= 0:94) which is historically classified\nas a BL Lac object (Spinrad & Smith 1975). For this object,\nwe have found Ldisk/LEdd= 0.04 and CD = 2.69, thus in-\ndicating an underlying radiatively efficient accretion. There-\nfore, this source can be identified as a FSRQ and as a HCD\nblazar (see also Ghisellini et al. 2011).\nThe above mentioned classification scheme is analogous\nto that based on LBLR/LEddproposed by Ghisellini et al.14 P ALIYA ET AL .\n(2011). However, it overcomes the limitation imposed by\nthe difficulty in measuring LBLR andMBHfor objects with\nquasi-featureless spectra. Another advantage of a CD based\nclassification is that one can easily get an idea about the\nintrinsic physical nature of blazars just by measuring the\nrelative dominance of the high-energy peak in their multi-\nfrequency SEDs. The proposed scheme is useful keep-\ning in mind the ongoing and next-generation surveys, e.g.,\nVery Large Array Sky Survey, Dark Energy Survey, e-\nROSITA , All-sky MeV Energy Gamma-ray Observatory, and\nCherenkov Telescope Array, which will provide unprece-\ndented broadband coverage of the blazar SEDs.\nBased on Figure 11, we argue that the region of CD.1,\nLdisk/LEdd>0.01 could be populated with strong line AGN\nhaving misaligned jets, e.g., broad line radio galaxies. This\nis due to the fact that the external Compton mechanism pro-\nducing X- to \r-ray emission is highly sensitive to the Doppler\nboosting. Therefore, as the viewing angle increases, the de-\ncrease in the high-energy peak flux is much more rapid than\nthe synchrotron peak flux, thereby effectively reducing CD.\nHowever, since the BLR emission is likely to be isotropic,\ntheLdiskof the source does not change. At the same time,\nthe region of CD&1,Ldisk/LEdd<0.01 may remain largely\nforbidden due to low-level of accretion and hence a jet envi-\nronment starved of the seed photons needed for the external\nCompton process.\nRecently, a minority of sources, so-called masquerading\nBL Lacs, have been identified challenging our current under-\nstanding of the FSRQ/BL Lac division. It has been proposed\nthat these objects are intrinsically FSRQs, i.e., radiatively ef-\nficient accreting systems. However, since their optical spec-\ntra lack emission lines due to strongly amplified jetted radi-\nation contamination, they are classified as BL Lac sources\n(Giommi et al. 2013; Padovani et al. 2019). Furthermore,\nmany of such objects exhibit high-frequency peaked SEDs\n(Padovani et al. 2012). This was explained by arguing the \r-\nray emitting region to be located outside BLR where a rela-\ntively weak cooling environment allows jet electrons to attain\nvery high-energies (Ghisellini et al. 2012). In the context of\nour work, physics of such blue FSRQs can be explained as\nfollows: the Doppler boosting effect that swamps out broad\nemission lines by enhancing synchrotron radiation should\namplify the inverse Compton emission by an even larger fac-\ntor if the primary mechanism to produce \r-rays is external\nCompton. If so, the ratio of the SED peak luminosities, i.e.,\nCD, cannot be significantly lower than one (see, e.g., Ghis-\nellini et al. 2012). This supports our claim that radiatively ef-\nficient accreting systems should have a Compton dominated\nSED. Based on Figure 11, it can be understood that sources\nlying around CD\u00191 andLdisk/LEdd\u00190.01 may belong to\nthe family of masquerading BL Lacs.\n12 13 14 15 16 17 18\nlogνpeak\nsyn[Hz]−5−4−3−2−10logLdisk/LEdd\nabsorption line\nemission lineFigure 12. This diagram illustrates the variation of the Ldisk in\nEddington units as a function of the synchrotron peak frequency for\nFermi blazars studied in this work.\nWe stress that our physical interpretation of the obtained\nCD andLdisk/LEdd correlation may not be extendable to\nblazar flares which can be extremely complex and diverse\n(e.g., Rajput et al. 2020). This is because we used multi-\nwavelength SEDs generated using all archival observations\nto compute CD hence reflect the average activity state of\nsources. Similarly, optical spectra too were mostly taken dur-\ning average/low activity state of sources and may not be dur-\ning any specific flaring episodes. Therefore, the derived cor-\nrelation of CD and Ldisk/LEddrefers to the average physical\nproperties of the blazar population. Extending this work to\nconsider blazar flares is beyond the scope of this paper since\neven one source behaves differently during different flares\n(e.g., Chatterjee et al. 2013; Paliya et al. 2015, 2016).\n7.A SEQUENCE OF Ldisk/LEdd\nThe observed anti-correlation between the synchrotron\npeak luminosity and corresponding peak frequency, i.e.,\nblazar sequence, has remained one of the crucial topics of\nresearch in jet physics. This phenomenon has been ar-\ngued to have a physical origin (e.g., Ghisellini et al. 1998)\nand also criticized due to possible selection effects of miss-\ning high-frequency peaked, luminous blazars (cf. Padovani\n2007). A few other works have proposed the Doppler boost-\ning/viewing angle to be the driving factor of the observed\nsequence (e.g., Nieppola et al. 2008; Meyer et al. 2011; Fan\net al. 2017; Keenan et al. 2020). Recently, the identification\nof the firstz > 3BL Lac object (Paliya et al. 2020c) has\nalso suggested that a population of luminous, high-frequency\npeaked blazars may exist which is against the prediction of\nthe blazar sequence (Ghisellini et al. 2017).\nIn Figure 12, we show the variation of the MBHnormal-\nizedLdiskas a function of the synchrotron peak frequency for\nobjects studied in this work. Note that Ldiskvalues are de-BLAZAR ’SCENTRAL ENGINE 15\nrived from the optical spectroscopic emission line fluxes or\nupper limits and therefore are free from the Doppler boost-\ning effects. In other words, Figure 12 provides a glimpse of\nthe intrinsic physical behavior of beamed AGN population.\nA strong anti-correlation of the plotted quantities is evident\nwhich is confirmed from the Spearmann’s test done using\nASURV giving \u001a=\u00000:75with PNC<1\u000210\u000010. This\nanti-correlation can be understood as follows.\nPopulation studies focused on the average activity state\nof blazars have revealed the blazar-zone to lie at a distance\nof a few hundreds/thousands of Schwarzschild radii from\nthe central black hole (e.g., Paliya et al. 2017a). Consid-\nering the BLR/torus radius- Ldiskrelationship (e.g., Tavec-\nchio & Ghisellini 2008), strong line blazars hosting a lu-\nminous accretion disk have a large BLR/torus and there-\nfore the emission region is typically located within it. The\nrelativistic electrons present in the jet would loose energy\nmainly by interacting with ambient photons before they could\nreach to high-energies, leading to the observation of a low-\nfrequency peaked SED. In radiatively inefficient systems, on\nthe other hand, Ldiskhas a low-value implying a relatively\nsmall BLR/torus radius and hence the emission region would\nbe located farther out from it. In the absence of a strong pho-\nton field, jet electrons can attain high-energies making the\nobserved SED to be high-frequency peaked. Moreover, Fig-\nure 12 also suggests that the shift of the SED peaks to higher\nfrequencies occurs with the depletion of the central environ-\nment by accretion, i.e., with the evolution of HCD blazars to\ntheir LCD counterparts (B ¨ottcher & Dermer 2002; Cavaliere\n& D’Elia 2002).\nFinally, since Ldisk/LEdd positively correlates with CD\n(Figure 11), one would also expect an anti-correlation of the\nlatter with the synchrotron peak frequency similar to Fig-\nure 12. Such an anti-correlation has already been reported\nby Finke (2013).\n8.SUMMARY\nIn this work, we have presented a catalog of the central\nengine properties, i.e., MBHandLdisk, for a sample of 1077\n\r-ray detected blazars. We summarize our results below.\n1. The average MBHfor the whole blazar population is\nhlogMBH;allM\fi= 8:60. There are evidences indi-cating black holes residing in absorption line systems\nto be more massive compared to emission line blazars.\n2. The distribution of Ldisk/LEdd reveals a clear bi-\nmodality with emission line sources tend to have a\nlarger accretion rate ( Ldisk/LEdd>0.01) compared to\nblazars whose optical spectra are dominated by the ab-\nsorption lines arising from the host galaxy. Similar re-\nsults were obtained for CD where emission line blazars\nexhibit a more Compton dominated SED than absorp-\ntion line sources.\n3. There is strong positive correlation between\nLdisk/LEdd and CD, suggesting that the latter can\nbe used to determine the state of accretion activity in\nbeamed AGNs.\n4. Based on their position in Ldisk/LEdd-CD plane,\nblazars can be classified as High-Compton Domi-\nnated (HCD, CD >1) and Low-Compton Dominated\n(LCD, CD<1) objects. Our results suggests that\nthis scheme is analogous to that predicted based on\nLBLR/LEdd, however, can be extended to objects\nlacking high-quality optical spectroscopic measure-\nments. We also propose that blazars having CD>1,\nLdisk/LEdd>0.01 should be identified as FSRQs and\nCD<1,Ldisk/LEdd<0.01 as BL Lacs.\n5.Fermi blazars show a significant anti-correlation be-\ntweenLdisk/LEdd and synchrotron peak frequency.\nBeing free from Doppler boosting related effects, the\nobserved trend has a physical origin.\n6. The overall findings reported in this work directly\npoints to a scenario in which the average physical prop-\nerties of blazars are mainly controlled by the accretion\nrate in Eddington units.\nWe make the catalog publicly available online at Zenodo\n(doi: 10.5281/zenodo.4456365)5.\nSoftware: ASURV (Lavalley et al. 1992), Astropy (As-\ntropy Collaboration et al. 2013, 2018), pPXF (Cappellari &\nEmsellem 2004), PyQSOFit (Guo et al. 2018).\nREFERENCES\nAbdollahi, S., Acero, F., Ackermann, M., et al. 2020, ApJS, 247,\n33, doi: 10.3847/1538-4365/ab6bcb\n5The fits file is also available at http://www.ucm.es/blazars/enginesAcosta-Pulido, J. A., Agudo, I., Barrena, R., et al. 2010, A&A,\n519, A5, doi: 10.1051/0004-6361/200913953\nAfanas’Ev, V . L., Dodonov, S. N., Moiseev, A. V ., et al. 2005,\nAstronomy Reports, 49, 374, doi: 10.1134/1.1923547\n—. 2006, Astronomy Reports, 50, 255,\ndoi: 10.1134/S106377290604001916 P ALIYA ET AL .\nAhumada, R., Allende Prieto, C., Almeida, A., et al. 2020, ApJS,\n249, 3, doi: 10.3847/1538-4365/ab929e\nAjello, M., Romani, R. W., Gasparrini, D., et al. 2014, ApJ, 780,\n73, doi: 10.1088/0004-637X/780/1/73\nAjello, M., Ghisellini, G., Paliya, V . S., et al. 2016, ApJ, 826, 76,\ndoi: 10.3847/0004-637X/826/1/76\nAjello, M., Angioni, R., Axelsson, M., et al. 2020, ApJ, 892, 105,\ndoi: 10.3847/1538-4357/ab791e\nAliu, E., Aune, T., Beilicke, M., et al. 2011, ApJ, 742, 127,\ndoi: 10.1088/0004-637X/742/2/127\n´Alvarez Crespo, N., Massaro, F., Milisavljevic, D., et al. 2016a, AJ,\n151, 95, doi: 10.3847/0004-6256/151/4/95\n´Alvarez Crespo, N., Masetti, N., Ricci, F., et al. 2016b, AJ, 151,\n32, doi: 10.3847/0004-6256/151/2/32\nAstropy Collaboration, Robitaille, T. P., Tollerud, E. J., et al. 2013,\nA&A, 558, A33, doi: 10.1051/0004-6361/201322068\nAstropy Collaboration, Price-Whelan, A. M., Sip ˝ocz, B. M., et al.\n2018, AJ, 156, 123, doi: 10.3847/1538-3881/aabc4f\nBade, N., Beckmann, V ., Douglas, N. G., et al. 1998, A&A, 334,\n459. https://arxiv.org/abs/astro-ph/9803204\nBade, N., Fink, H. H., Engels, D., et al. 1995, A&AS, 110, 469\nBaker, J. C., Hunstead, R. W., Kapahi, V . K., & Subrahmanya,\nC. R. 1999, ApJS, 122, 29, doi: 10.1086/313209\nBaldwin, J. A., Wampler, E. J., & Burbidge, E. M. 1981, ApJ, 243,\n76, doi: 10.1086/158568\nBaldwin, J. A., Wampler, E. J., & Gaskell, C. M. 1989, ApJ, 338,\n630, doi: 10.1086/167224\nBecerra Gonz ´alez, J., Acosta-Pulido, J. A., & Clavero, R. 2020,\nMNRAS, 494, 6036, doi: 10.1093/mnras/staa1144\nBechtold, J., Dobrzycki, A., Wilden, B., et al. 2002, ApJS, 140,\n143, doi: 10.1086/342489\nBoisse, P., & Bergeron, J. 1988, A&A, 192, 1\nBoroson, T. A., & Green, R. F. 1992, ApJS, 80, 109,\ndoi: 10.1086/191661\nB¨ottcher, M., & Dermer, C. D. 2002, ApJ, 564, 86,\ndoi: 10.1086/324134\nBruni, G., Panessa, F., Ghisellini, G., et al. 2018, ApJL, 854, L23,\ndoi: 10.3847/2041-8213/aaacfb\nButtiglione, S., Capetti, A., Celotti, A., et al. 2009, A&A, 495,\n1033, doi: 10.1051/0004-6361:200811102\nCappellari, M., & Emsellem, E. 2004, PASP, 116, 138,\ndoi: 10.1086/381875\nCarangelo, N., Falomo, R., Kotilainen, J., Treves, A., & Ulrich,\nM. H. 2003, A&A, 412, 651, doi: 10.1051/0004-6361:20031519\nCardelli, J. A., Clayton, G. C., & Mathis, J. S. 1989, ApJ, 345, 245,\ndoi: 10.1086/167900\nCavaliere, A., & D’Elia, V . 2002, ApJ, 571, 226,\ndoi: 10.1086/339778\nCelotti, A., Padovani, P., & Ghisellini, G. 1997, MNRAS, 286, 415Chai, B., Cao, X., & Gu, M. 2012, ApJ, 759, 114,\ndoi: 10.1088/0004-637X/759/2/114\nChatterjee, R., Fossati, G., Urry, C. M., et al. 2013, ApJL, 763,\nL11, doi: 10.1088/2041-8205/763/1/L11\nChen, S., Berton, M., La Mura, G., et al. 2018, A&A, 615, A167,\ndoi: 10.1051/0004-6361/201832678\nChen, Y . Y ., Zhang, X., Zhang, H. J., & Yu, X. L. 2015, MNRAS,\n451, 4193, doi: 10.1093/mnras/stv658\nChomiuk, L., Strader, J., Landt, H., & Cheung, C. C. 2013, The\nAstronomer’s Telegram, 4777, 1\nChu, Y .-q., Zhu, X.-f., & Butcher, H. 1986, ChA&A, 10, 196,\ndoi: 10.1016/0275-1062(86)90005-6\nColless, M., Dalton, G., Maddox, S., et al. 2001, MNRAS, 328,\n1039, doi: 10.1046/j.1365-8711.2001.04902.x\nde Menezes, R., Amaya-Almaz ´an, R. A., Marchesini, E. J., et al.\n2020, Ap&SS, 365, 12, doi: 10.1007/s10509-020-3727-5\nDermer, C. D. 1995, ApJL, 446, L63, doi: 10.1086/187931\nDermer, C. D., Finke, J. D., Krug, H., & B ¨ottcher, M. 2009, ApJ,\n692, 32, doi: 10.1088/0004-637X/692/1/32\nDesai, A., Marchesi, S., Rajagopal, M., & Ajello, M. 2019, ApJS,\n241, 5, doi: 10.3847/1538-4365/ab01fc\ndi Serego Alighieri, S., Danziger, I. J., Morganti, R., & Tadhunter,\nC. N. 1994, MNRAS, 269, 998, doi: 10.1093/mnras/269.4.998\nDrinkwater, M. J., Webster, R. L., Francis, P. J., et al. 1997,\nMNRAS, 284, 85\nDunlop, J. S., Peacock, J. A., Savage, A., et al. 1989, MNRAS,\n238, 1171, doi: 10.1093/mnras/238.4.1171\nEvans, I. N., & Koratkar, A. P. 2004, ApJS, 150, 73,\ndoi: 10.1086/379649\nEvans, P. A., Page, K. L., Osborne, J. P., et al. 2020, ApJS, 247, 54,\ndoi: 10.3847/1538-4365/ab7db9\nFalomo, R., Kotilainen, J., Pursimo, T., et al. 1997, A&A, 321, 374\nFalomo, R., Scarpa, R., Treves, A., & Urry, C. M. 2000, ApJ, 542,\n731, doi: 10.1086/317044\nFalomo, R., Treves, A., Scarpa, R., Paiano, S., & Land oni, M.\n2017, MNRAS, 470, 2814, doi: 10.1093/mnras/stx1411\nFan, J. H., Yang, J. H., Xiao, H. B., et al. 2017, ApJL, 835, L38,\ndoi: 10.3847/2041-8213/835/2/L38\nFerrarese, L., & Merritt, D. 2000, ApJL, 539, L9,\ndoi: 10.1086/312838\nFinke, J. D. 2013, ApJ, 763, 134,\ndoi: 10.1088/0004-637X/763/2/134\nFinke, J. D., Dermer, C. D., & B ¨ottcher, M. 2008, ApJ, 686, 181,\ndoi: 10.1086/590900\nFoschini, L., Berton, M., Caccianiga, A., et al. 2015, A&A, 575,\nA13, doi: 10.1051/0004-6361/201424972\nFrancis, P. J., Hewett, P. C., Foltz, C. B., et al. 1991, ApJ, 373, 465,\ndoi: 10.1086/170066\nFricke, K. J., Kollatschny, W., & Witzel, A. 1983, A&A, 117, 60BLAZAR ’SCENTRAL ENGINE 17\nGhisellini, G., Celotti, A., Fossati, G., Maraschi, L., & Comastri,\nA. 1998, MNRAS, 301, 451,\ndoi: 10.1046/j.1365-8711.1998.02032.x\nGhisellini, G., Righi, C., Costamante, L., & Tavecchio, F. 2017,\nMNRAS, 469, 255, doi: 10.1093/mnras/stx806\nGhisellini, G., Tavecchio, F., Foschini, L., & Ghirlanda, G. 2011,\nMNRAS, 414, 2674, doi: 10.1111/j.1365-2966.2011.18578.x\nGhisellini, G., Tavecchio, F., Foschini, L., et al. 2012, MNRAS,\n425, 1371, doi: 10.1111/j.1365-2966.2012.21554.x\nGhisellini, G., Tavecchio, F., Maraschi, L., Celotti, A., & Sbarrato,\nT. 2014, Nature, 515, 376, doi: 10.1038/nature13856\nGioia, I. M., Wolter, A., Mullis, C. R., et al. 2004, A&A, 428, 867,\ndoi: 10.1051/0004-6361:20041426\nGiommi, P., Padovani, P., & Polenta, G. 2013, MNRAS, 431, 1914,\ndoi: 10.1093/mnras/stt305\nGorham, P. W., van Zee, L., Unwin, S. C., & Jacobs, C. 2000, AJ,\n119, 1677, doi: 10.1086/301289\nGraham, A. W. 2007, MNRAS, 379, 711,\ndoi: 10.1111/j.1365-2966.2007.11950.x\nG¨ultekin, K., Richstone, D. O., Gebhardt, K., et al. 2009, ApJ, 698,\n198, doi: 10.1088/0004-637X/698/1/198\nGuo, H., Liu, X., Shen, Y ., et al. 2019, MNRAS, 482, 3288,\ndoi: 10.1093/mnras/sty2920\nGuo, H., Shen, Y ., & Wang, S. 2018, PyQSOFit: Python code to fit\nthe spectrum of quasars. http://ascl.net/1809.008\nHalpern, J. P., Chen, V . S., Madejski, G. M., & Chanan, G. A.\n1991, AJ, 101, 818, doi: 10.1086/115725\nHalpern, J. P., Eracleous, M., & Forster, K. 1997, AJ, 114, 1736,\ndoi: 10.1086/118602\nHalpern, J. P., Eracleous, M., & Mattox, J. R. 2003, AJ, 125, 572,\ndoi: 10.1086/345796\nHao, L., Strauss, M. A., Tremonti, C. A., et al. 2005, AJ, 129,\n1783, doi: 10.1086/428485\nHealey, S. E., Romani, R. W., Cotter, G., et al. 2008, ApJS, 175,\n97, doi: 10.1086/523302\nHenstock, D. R., Browne, I. W. A., Wilkinson, P. N., & McMahon,\nR. G. 1997, MNRAS, 290, 380, doi: 10.1093/mnras/290.2.380\nIsobe, T., Feigelson, E. D., & Nelson, P. I. 1986, ApJ, 306, 490,\ndoi: 10.1086/164359\nJones, D. H., Read, M. A., Saunders, W., et al. 2009, MNRAS,\n399, 683, doi: 10.1111/j.1365-2966.2009.15338.x\nJunkkarinen, V . 1984, PASP, 96, 539, doi: 10.1086/131374\nKeenan, M., Meyer, E. T., Georganopoulos, M., Reddy, K., &\nFrench, O. J. 2020, arXiv e-prints, arXiv:2007.12661.\nhttps://arxiv.org/abs/2007.12661\nKlindt, L., van Soelen, B., Meintjes, P. J., & V ¨ais¨anen, P. 2017,\nMNRAS, 467, 2537, doi: 10.1093/mnras/stx218\nKormendy, J., & Ho, L. C. 2013, ARA&A, 51, 511,\ndoi: 10.1146/annurev-astro-082708-101811Koss, M., Trakhtenbrot, B., Ricci, C., et al. 2017, ApJ, 850, 74,\ndoi: 10.3847/1538-4357/aa8ec9\nKotilainen, J. K., Hyv ¨onen, T., & Falomo, R. 2005, A&A, 440,\n831, doi: 10.1051/0004-6361:20042548\nLandoni, M., Falomo, R., Treves, A., et al. 2013, AJ, 145, 114,\ndoi: 10.1088/0004-6256/145/4/114\nLandoni, M., Paiano, S., Falomo, R., Scarpa, R., & Treves, A.\n2018, ApJ, 861, 130, doi: 10.3847/1538-4357/aac77c\nLandt, H., Padovani, P., Perlman, E. S., et al. 2001, MNRAS, 323,\n757, doi: 10.1046/j.1365-8711.2001.04269.x\nLaurent-Muehleisen, S. A., Kollgaard, R. I., Ciardullo, R., et al.\n1998, ApJS, 118, 127, doi: 10.1086/313134\nLavalley, M. P., Isobe, T., & Feigelson, E. D. 1992, in Bulletin of\nthe American Astronomical Society, V ol. 24, Bulletin of the\nAmerican Astronomical Society, 839–840\nLawrence, C. R., Zucker, J. R., Readhead, A. C. S., et al. 1996,\nApJS, 107, 541, doi: 10.1086/192375\nLondish, D., Croom, S. M., Heidt, J., et al. 2007, MNRAS, 374,\n556, doi: 10.1111/j.1365-2966.2006.11165.x\nMarcha, M. J. M., Browne, I. W. A., Impey, C. D., & Smith, P. S.\n1996, MNRAS, 281, 425, doi: 10.1093/mnras/281.2.425\nMarchesini, E. J., Andruchow, I., Cellone, S. A., et al. 2016, A&A,\n591, A21, doi: 10.1051/0004-6361/201527632\nMarchesini, E. J., Pe ˜na-Herazo, H. A., ´Alvarez Crespo, N., et al.\n2019, Ap&SS, 364, 5, doi: 10.1007/s10509-018-3490-z\nMarziani, P., Sulentic, J. W., Zamanov, R., et al. 2003, ApJS, 145,\n199, doi: 10.1086/346025\nMasetti, N., Mason, E., Landi, R., et al. 2008, A&A, 480, 715,\ndoi: 10.1051/0004-6361:20078901\nMassaro, F., Landoni, M., D’Abrusco, R., et al. 2015, A&A, 575,\nA124, doi: 10.1051/0004-6361/201425119\nMcLure, R. J., & Dunlop, J. S. 2004, MNRAS, 352, 1390,\ndoi: 10.1111/j.1365-2966.2004.08034.x\nMerritt, D. 1997, AJ, 114, 228, doi: 10.1086/118467\nMeyer, E. T., Fossati, G., Georganopoulos, M., & Lister, M. L.\n2011, ApJ, 740, 98, doi: 10.1088/0004-637X/740/2/98\nMezcua, M., Chavushyan, V . H., Lobanov, A. P., & Le ´on-Tavares,\nJ. 2012, A&A, 544, A36, doi: 10.1051/0004-6361/201117724\nMorton, D. C., Savage, A., & Bolton, J. G. 1978, MNRAS, 185,\n735, doi: 10.1093/mnras/185.4.735\nMurdoch, H. S., Hunstead, R. W., & White, G. L. 1984,\nProceedings of the Astronomical Society of Australia, 5, 341,\ndoi: 10.1017/S1323358000017173\nNarayan, R., Garcia, M. R., & McClintock, J. E. 1997, ApJL, 478,\nL79, doi: 10.1086/310554\nNieppola, E., Valtaoja, E., Tornikoski, M., Hovatta, T., & Kotiranta,\nM. 2008, A&A, 488, 867, doi: 10.1051/0004-6361:200809716\nNilsson, K., Pursimo, T., Heidt, J., et al. 2003, A&A, 400, 95,\ndoi: 10.1051/0004-6361:2002186118 P ALIYA ET AL .\nNkundabakura, P., & Meintjes, P. J. 2012, MNRAS, 427, 859,\ndoi: 10.1111/j.1365-2966.2012.21953.x\nOlgu ´ın-Iglesias, A., Le ´on-Tavares, J., Kotilainen, J. K., et al. 2016,\nMNRAS, 460, 3202, doi: 10.1093/mnras/stw1208\nPadovani, P. 2007, Ap&SS, 309, 63,\ndoi: 10.1007/s10509-007-9455-2\nPadovani, P., Giommi, P., & Rau, A. 2012, MNRAS, 422, L48,\ndoi: 10.1111/j.1745-3933.2012.01234.x\nPadovani, P., Oikonomou, F., Petropoulou, M., Giommi, P., &\nResconi, E. 2019, MNRAS, 484, L104,\ndoi: 10.1093/mnrasl/slz011\nPaiano, S., Falomo, R., Franceschini, A., Treves, A., & Scarpa, R.\n2017a, ApJ, 851, 135, doi: 10.3847/1538-4357/aa9af4\nPaiano, S., Falomo, R., Treves, A., Franceschini, A., & Scarpa, R.\n2019, ApJ, 871, 162, doi: 10.3847/1538-4357/aaf6e4\nPaiano, S., Falomo, R., Treves, A., & Scarpa, R. 2020, MNRAS,\n497, 94, doi: 10.1093/mnras/staa1840\nPaiano, S., Landoni, M., Falomo, R., Treves, A., & Scarpa, R.\n2017b, ApJ, 844, 120, doi: 10.3847/1538-4357/aa7aac\nPaiano, S., Landoni, M., Falomo, R., et al. 2017c, ApJ, 837, 144,\ndoi: 10.3847/1538-4357/837/2/144\nPaliya, V . S. 2015, ApJ, 804, 74, doi: 10.1088/0004-637X/804/1/74\nPaliya, V . S., Ajello, M., Cao, H. M., et al. 2020a, ApJ, 897, 177,\ndoi: 10.3847/1538-4357/ab9c1a\nPaliya, V . S., B ¨ottcher, M., Olmo-Garc ´ıa, A., et al. 2020b, ApJ,\n902, 29, doi: 10.3847/1538-4357/abb46e\nPaliya, V . S., Diltz, C., B ¨ottcher, M., Stalin, C. S., & Buckley, D.\n2016, ApJ, 817, 61, doi: 10.3847/0004-637X/817/1/61\nPaliya, V . S., Marcotulli, L., Ajello, M., et al. 2017a, ApJ, 851, 33,\ndoi: 10.3847/1538-4357/aa98e1\nPaliya, V . S., Sahayanathan, S., & Stalin, C. S. 2015, ApJ, 803, 15,\ndoi: 10.1088/0004-637X/803/1/15\nPaliya, V . S., Stalin, C. S., Ajello, M., & Kaur, A. 2017b, ApJ, 844,\n32, doi: 10.3847/1538-4357/aa77f5\nPaliya, V . S., Koss, M., Trakhtenbrot, B., et al. 2019, ApJ, 881,\n154, doi: 10.3847/1538-4357/ab2f8b\nPaliya, V . S., Dom ´ınguez, A., Cabello, C., et al. 2020c, ApJL, 903,\nL8, doi: 10.3847/2041-8213/abbc06\nPe˜na-Herazo, H. A., Marchesini, E. J., ´Alvarez Crespo, N., et al.\n2017, Ap&SS, 362, 228, doi: 10.1007/s10509-017-3208-7\nPe˜na-Herazo, H. A., Massaro, F., Chavushyan, V ., et al. 2019,\nAp&SS, 364, 85, doi: 10.1007/s10509-019-3574-4\nPe˜na-Herazo, H. A., Amaya-Almaz ´an, R. A., Massaro, F., et al.\n2020, A&A, 643, A103, doi: 10.1051/0004-6361/202037978\nPerlman, E. S., Padovani, P., Giommi, P., et al. 1998, AJ, 115,\n1253, doi: 10.1086/300283\nPerlman, E. S., Stocke, J. T., Schachter, J. F., et al. 1996, ApJS,\n104, 251, doi: 10.1086/192300\nPietsch, W., Bischoff, K., Boller, T., et al. 1998, A&A, 333, 48.\nhttps://arxiv.org/abs/astro-ph/9801210Piranomonte, S., Perri, M., Giommi, P., Landt, H., & Padovani, P.\n2007, A&A, 470, 787, doi: 10.1051/0004-6361:20077086\nPita, S., Goldoni, P., Boisson, C., et al. 2014, A&A, 565, A12,\ndoi: 10.1051/0004-6361/201323071\nPlotkin, R. M., Markoff, S., Trager, S. C., & Anderson, S. F. 2011,\nMNRAS, 413, 805, doi: 10.1111/j.1365-2966.2010.18172.x\nPunsly, B., & Kharb, P. 2017, MNRAS, 468, L72,\ndoi: 10.1093/mnrasl/slx024\nPursimo, T., Ojha, R., Jauncey, D. L., et al. 2013, ApJ, 767, 14,\ndoi: 10.1088/0004-637X/767/1/14\nRaiteri, C. M., Villata, M., Capetti, A., et al. 2007, A&A, 464, 871,\ndoi: 10.1051/0004-6361:20066599\nRajput, B., Stalin, C. S., & Sahayanathan, S. 2020, MNRAS, 498,\n5128, doi: 10.1093/mnras/staa2708\nRector, T. A., Stocke, J. T., Perlman, E. S., Morris, S. L., & Gioia,\nI. M. 2000, AJ, 120, 1626, doi: 10.1086/301587\nRicci, F., Massaro, F., Landoni, M., et al. 2015, AJ, 149, 160,\ndoi: 10.1088/0004-6256/149/5/160\nRichards, G. T., Myers, A. D., Gray, A. G., et al. 2009, ApJS, 180,\n67, doi: 10.1088/0067-0049/180/1/67\nRohatgi, A. 2020, Webplotdigitizer: Version 4.3.\nhttps://automeris.io/WebPlotDigitizer\nS´anchez-Bl ´azquez, P., Peletier, R. F., Jim ´enez-Vicente, J., et al.\n2006, MNRAS, 371, 703,\ndoi: 10.1111/j.1365-2966.2006.10699.x\nSandrinelli, A., Treves, A., Falomo, R., et al. 2013, AJ, 146, 163,\ndoi: 10.1088/0004-6256/146/6/163\nSbarrato, T., Padovani, P., & Ghisellini, G. 2014, MNRAS, 445,\n81, doi: 10.1093/mnras/stu1759\nSbarufatti, B., Ciprini, S., Kotilainen, J., et al. 2009, AJ, 137, 337,\ndoi: 10.1088/0004-6256/137/1/337\nSbarufatti, B., Falomo, R., Treves, A., & Kotilainen, J. 2006a,\nA&A, 457, 35, doi: 10.1051/0004-6361:20065455\nSbarufatti, B., Treves, A., Falomo, R., et al. 2005, AJ, 129, 559,\ndoi: 10.1086/427138\n—. 2006b, AJ, 132, 1, doi: 10.1086/503031\nSchlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJ, 500,\n525, doi: 10.1086/305772\nShaw, M. S., Filippenko, A. V ., Romani, R. W., Cenko, S. B., & Li,\nW. 2013a, AJ, 146, 127, doi: 10.1088/0004-6256/146/5/127\nShaw, M. S., Romani, R. W., Cotter, G., et al. 2012, ApJ, 748, 49,\ndoi: 10.1088/0004-637X/748/1/49\n—. 2013b, ApJ, 764, 135, doi: 10.1088/0004-637X/764/2/135\nShen, Y ., Richards, G. T., Strauss, M. A., et al. 2011, ApJS, 194,\n45, doi: 10.1088/0067-0049/194/2/45\nShen, Y ., Hall, P. B., Horne, K., et al. 2019, ApJS, 241, 34,\ndoi: 10.3847/1538-4365/ab074f\nSikora, M., Begelman, M. C., & Rees, M. J. 1994, ApJ, 421, 153,\ndoi: 10.1086/173633BLAZAR ’SCENTRAL ENGINE 19\nSmith, P. S., Montiel, E., Rightley, S., et al. 2009,\narXiv:0912.3621. https://arxiv.org/abs/0912.3621\nSowards-Emmerd, D., Romani, R. W., & Michelson, P. F. 2003,\nApJ, 590, 109, doi: 10.1086/374981\nSowards-Emmerd, D., Romani, R. W., Michelson, P. F., &\nUlvestad, J. S. 2004, ApJ, 609, 564, doi: 10.1086/421239\nSpinrad, H., & Smith, H. E. 1975, ApJ, 201, 275,\ndoi: 10.1086/153883\nSteidel, C. C., & Sargent, W. L. W. 1991, ApJ, 382, 433,\ndoi: 10.1086/170732\nStickel, M., Fried, J. W., & Kuehr, H. 1988, A&A, 191, L16\n—. 1989, A&AS, 80, 103\n—. 1993a, A&AS, 98, 393\nStickel, M., & Kuehr, H. 1993, A&AS, 100, 395\n—. 1994, A&AS, 105, 67\nStickel, M., Kuehr, H., & Fried, J. W. 1993b, A&AS, 97, 483\nStickel, M., & Kuhr, H. 1993, A&AS, 101, 521\nStickel, M., Padovani, P., Urry, C. M., Fried, J. W., & Kuehr, H.\n1991, ApJ, 374, 431, doi: 10.1086/170133\nTavecchio, F., & Ghisellini, G. 2008, MNRAS, 386, 945,\ndoi: 10.1111/j.1365-2966.2008.13072.x\nThompson, D. J., Djorgovski, S., & de Carvalho, R. 1990, PASP,\n102, 1235, doi: 10.1086/132758\nTitov, O., Jauncey, D. L., Johnston, H. M., Hunstead, R. W., &\nChristensen, L. 2011, AJ, 142, 165,\ndoi: 10.1088/0004-6256/142/5/165\nTitov, O., Pursimo, T., Johnston, H. M., et al. 2017, AJ, 153, 157,\ndoi: 10.3847/1538-3881/aa61fd\nTitov, O., Stanford, L. M., Johnston, H. M., et al. 2013, AJ, 146,\n10, doi: 10.1088/0004-6256/146/1/10\nTorrealba, J., Chavushyan, V ., Cruz-Gonz ´alez, I., et al. 2012,\nRMxAA, 48, 9. https://arxiv.org/abs/1107.3416Tramacere, A., Giommi, P., Perri, M., Verrecchia, F., & Tosti, G.\n2009, A&A, 501, 879, doi: 10.1051/0004-6361/200810865\nUrry, C. M., Scarpa, R., O’Dowd, M., et al. 2000, ApJ, 532, 816,\ndoi: 10.1086/308616\nvan den Berg, J. P., B ¨ottcher, M., Dom ´ınguez, A., & L ´opez-Moya,\nM. 2019, ApJ, 874, 47, doi: 10.3847/1538-4357/aafdfd\nVandenbroucke, J., Buehler, R., Ajello, M., et al. 2010, ApJL, 718,\nL166, doi: 10.1088/2041-8205/718/2/L166\nVazdekis, A., S ´anchez-Bl ´azquez, P., Falc ´on-Barroso, J., et al. 2010,\nMNRAS, 404, 1639, doi: 10.1111/j.1365-2966.2010.16407.x\nVermeulen, R. C., Ogle, P. M., Tran, H. D., et al. 1995, ApJL, 452,\nL5, doi: 10.1086/309716\nVermeulen, R. C., Taylor, G. B., Readhead, A. C. S., & Browne,\nI. W. A. 1996, AJ, 111, 1013, doi: 10.1086/117847\nVestergaard, M., & Peterson, B. M. 2006, ApJ, 641, 689,\ndoi: 10.1086/500572\nVestergaard, M., & Wilkes, B. J. 2001, ApJS, 134, 1,\ndoi: 10.1086/320357\nWallace, P. M., Halpern, J. P., Magalh ˜aes, A. M., & Thompson,\nD. J. 2002, ApJ, 569, 36, doi: 10.1086/339286\nWhite, R. L., Becker, R. H., Gregg, M. D., et al. 2000, ApJS, 126,\n133, doi: 10.1086/313300\nWilkes, B. J., Wright, A. E., Jauncey, D. L., & Peterson, B. A.\n1983, Proceedings of the Astronomical Society of Australia, 5, 2\nWolter, A., Ruscica, C., & Caccianiga, A. 1998, MNRAS, 299,\n1047, doi: 10.1046/j.1365-8711.1998.01834.x\nYip, C. W., Connolly, A. J., Szalay, A. S., et al. 2004a, AJ, 128,\n585, doi: 10.1086/422429\nYip, C. W., Connolly, A. J., Vanden Berk, D. E., et al. 2004b, AJ,\n128, 2603, doi: 10.1086/42562620 P ALIYA ET AL .\nTable 10. References used in this work.\n4FGL name Reference 4FGL name Reference 4FGL name Reference\nJ0001.5+2113 Ahumada et al. (2020) J0833.4-0458 Paliya et al. (2020a) J1417.9+2543 Ahumada et al. (2020)\nJ0003.2+2207 Ahumada et al. (2020) J0833.9+4223 Ahumada et al. (2020) J1417.9+4613 Ahumada et al. (2020)\nJ0004.3+4614 Sowards-Emmerd et al. (2003) J0836.5-2026 Fricke et al. (1983) J1418.4+3543 Ahumada et al. (2020)\nJ0004.4-4737 Shaw et al. (2012) J0836.9+5833 Ahumada et al. (2020) J1419.5+3821 Ahumada et al. (2020)\nJ0006.3-0620 Stickel et al. (1989) J0837.3+1458 Ahumada et al. (2020) J1419.8+5423 Ahumada et al. (2020)\nNOTE—(This table is available in its entirety in a machine-readable form in the online journal. A portion is shown here for guidance\nregarding its form and content.)\nReferences —Acosta-Pulido et al. (2010); Afanas’Ev et al. (2005, 2006); Ahumada et al. (2020); Ajello et al. (2016); Aliu et al. (2011);\n´Alvarez Crespo et al. (2016a,b); Bade et al. (1995, 1998); Baker et al. (1999); Baldwin et al. (1981, 1989); Becerra Gonz ´alez et al.\n(2020); Bechtold et al. (2002); Boisse & Bergeron (1988); Bruni et al. (2018); Buttiglione et al. (2009); Carangelo et al. (2003);\nChai et al. (2012); Chomiuk et al. (2013); Chu et al. (1986); Colless et al. (2001); Desai et al. (2019); Drinkwater et al. (1997);\nDunlop et al. (1989); Evans & Koratkar (2004); Falomo et al. (1997, 2000, 2017); Foschini et al. (2015); Fricke et al. (1983); Gioia\net al. (2004); Gorham et al. (2000); Halpern et al. (1991, 1997, 2003); Healey et al. (2008); Henstock et al. (1997); Jones et al.\n(2009); Junkkarinen (1984); Klindt et al. (2017); Koss et al. (2017); Kotilainen et al. (2005); Landoni et al. (2013, 2018); Landt et al.\n(2001); Laurent-Muehleisen et al. (1998); Lawrence et al. (1996); Londish et al. (2007); Marcha et al. (1996); Marchesini et al. (2016,\n2019); Marziani et al. (2003); Masetti et al. (2008); Massaro et al. (2015); Mezcua et al. (2012); Morton et al. (1978); Murdoch et al.\n(1984); Nilsson et al. (2003); Nkundabakura & Meintjes (2012); Olgu ´ın-Iglesias et al. (2016); Padovani et al. (2019); Paiano et al.\n(2017a,b,c, 2019, 2020); Paliya et al. (2019, 2020a,b); Pe ˜na-Herazo et al. (2017, 2019, 2020); Perlman et al. (1996, 1998); Pietsch\net al. (1998); Piranomonte et al. (2007); Pita et al. (2014); Punsly & Kharb (2017); Pursimo et al. (2013); Raiteri et al. (2007); Rector\net al. (2000); Ricci et al. (2015); Sandrinelli et al. (2013); Sbarufatti et al. (2005, 2006a,b, 2009); Shaw et al. (2012, 2013a,b); Smith\net al. (2009); Sowards-Emmerd et al. (2003, 2004); Steidel & Sargent (1991); Stickel & Kuehr (1993, 1994); Stickel & Kuhr (1993);\nStickel et al. (1988, 1989, 1993a,b); Thompson et al. (1990); Titov et al. (2011, 2013, 2017); Torrealba et al. (2012); Urry et al. (2000);\nVandenbroucke et al. (2010); Vermeulen et al. (1995, 1996); Wallace et al. (2002); White et al. (2000); Wilkes et al. (1983); Wolter\net al. (1998); de Menezes et al. (2020); di Serego Alighieri et al. (1994)\nAPPENDIX\nACKNOWLEDGMENTS\nWe are grateful to the journal referee for constructive crit-\nicism. VSP is thankful to A. Desai, R. de Menezes, T. Pur-\nsimo, and J. Strader for providing optical spectra of a few\nblazars in tabular format and H. Guo and M. Cappellari for\nuseful discussions over the application of PyQSOFit and\npPXF tools, respectively. VSP’s work was supported by the\nInitiative and Networking Fund of the Helmholtz Associa-\ntion. A.D. acknowledges the support of the Ram ´on y Ca-\njal program from the Spanish MINECO. A.O.G. acknowl-\nedges financial support from the Spanish Ministry of Science,\nInnovation and Universities (MCIUN) under grant numbers\nRTI2018-096188-B-I00, and from the Comunidad de Madrid\nTec2Space project S2018/NMT-4291.\nFunding for the Sloan Digital Sky Survey IV has been pro-\nvided by the Alfred P. Sloan Foundation, the U.S. Depart-ment of Energy Office of Science, and the Participating Insti-\ntutions. SDSS-IV acknowledges support and resources from\nthe Center for High-Performance Computing at the Univer-\nsity of Utah. The SDSS web site is www.sdss.org.\nSDSS-IV is managed by the Astrophysical Research Con-\nsortium for the Participating Institutions of the SDSS Col-\nlaboration including the Brazilian Participation Group, the\nCarnegie Institution for Science, Carnegie Mellon Univer-\nsity, the Chilean Participation Group, the French Participa-\ntion Group, Harvard-Smithsonian Center for Astrophysics,\nInstituto de Astrof ´ısica de Canarias, The Johns Hopkins\nUniversity, Kavli Institute for the Physics and Mathemat-\nics of the Universe (IPMU) / University of Tokyo, the Ko-\nrean Participation Group, Lawrence Berkeley National Lab-\noratory, Leibniz Institut f ¨ur Astrophysik Potsdam (AIP),\nMax-Planck-Institut f ¨ur Astronomie (MPIA Heidelberg),BLAZAR ’SCENTRAL ENGINE 21\nMax-Planck-Institut f ¨ur Astrophysik (MPA Garching), Max-\nPlanck-Institut f ¨ur Extraterrestrische Physik (MPE), National\nAstronomical Observatories of China, New Mexico State\nUniversity, New York University, University of Notre Dame,\nObservat ´ario Nacional / MCTI, The Ohio State University,\nPennsylvania State University, Shanghai Astronomical Ob-\nservatory, United Kingdom Participation Group,Universidad\nNacional Aut ´onoma de M ´exico, University of Arizona, Uni-\nversity of Colorado Boulder, University of Oxford, Univer-\nsity of Portsmouth, University of Utah, University of Vir-\nginia, University of Washington, University of Wisconsin,\nVanderbilt University, and Yale University.\nThis research has made use of data obtained through the\nHigh Energy Astrophysics Science Archive Research Cen-\nter Online Service, provided by the NASA/Goddard Space\nFlight Center. This research has made use of the NASA/IPAC\nExtragalactic Database (NED), which is operated by the Jet\nPropulsion Laboratory, California Institute of Technology,\nunder contract with the National Aeronautics and Space Ad-\nministration. Part of this work is based on archival data, soft-\nware or online services provided by the Space Science Data\nCenter (SSDC). This research has made use of NASA’s As-\ntrophysics Data System Bibliographic Services.\nThis research made use of Astropy,6a community-\ndeveloped core Python package for Astronomy (Astropy Col-\nlaboration et al. 2013, 2018).\n6http://www.astropy.org" }, { "title": "2107.07950v1.Band_gap_measurements_of_monolayer_h_BN_and_insights_into_carbon_related_point_defects.pdf", "content": "Band gap measurements of monolayer h-BN and\ninsights into carbon-related point defects\nRicardo Javier Pe~ na Rom\u0013 an,\u0003,yF\u0013 abio J. R. Costa ,yAlberto Zobelli,zChristine\nElias,{Pierre Valvin,{Guillaume Cassabois,{Bernard Gil,{Alex Summer\feld,x\nTin S. Cheng,xChristopher J. Mellor,xPeter H. Beton,xSergei V. Novikov,xand\nLuiz F. Zagonel\u0003,y\nyInstitute of Physics \\Gleb Wataghin\", Department of Applied Physics, State University of\nCampinas-UNICAMP, 13083-859, Campinas, Brazil\nzUniversit\u0013 e Paris-Saclay, CNRS, Laboratoire de Physiques des Solides, 91405, Orsay,\nFrance\n{Laboratoire Charles Coulomb, UMR5221 CNRS-Universit\u0013 e de Montpellier, 34095\nMontpellier, France\nxSchool of Physics and Astronomy, University of Nottingham, Nottingham, NG7 2RD, UK\nE-mail: rikrdopr@i\f.unicamp.br; zagonel@unicamp.br\nThis is the version of the article before peer review or editing, as submitted by an author\nto IOP 2D Materials.\nFind the published version at: https://doi.org/10.1088/2053-1583/ac0d9c\nCite as: Ricardo Javier Pe~ na Rom\u0013 an et al 2021 2D Mater. 8 044001\n1arXiv:2107.07950v1 [cond-mat.mes-hall] 16 Jul 2021Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nAbstract\nBeing a \rexible wide band gap semiconductor, hexagonal boron nitride (h-BN) has\ngreat potential for technological applications like e\u000ecient deep ultraviolet (DUV) light\nsources, building block for two-dimensional heterostructures and room temperature sin-\ngle photon emitters in the UV and visible spectral range. To enable such applications,\nit is mandatory to reach a better understanding of the electronic and optical properties\nof h-BN and the impact of various structural defects. Despite the large e\u000borts in the\nlast years, aspects such as the electronic band gap value, the exciton binding energy\nand the e\u000bect of point defects remained elusive, particularly when considering a single\nmonolayer.\nHere, we directly measured the density of states of a single monolayer of h-BN epi-\ntaxially grown on highly oriented pyrolytic graphite, by performing low temperature\nscanning tunneling microscopy (LT-STM) and spectroscopy (STS). The observed h-BN\nelectronic band gap on defect-free regions is (6 :8\u00060:2) eV. Using optical spectroscopy\nto obtain the h-BN optical band gap, the exciton binding energy is determined as being\nof (0 :7\u00060:2) eV. In addition, the locally excited cathodoluminescence and photolumi-\nnescence show complex spectra that are typically associated to intragap states related\nto carbon defects. Moreover, in some regions of the monolayer h-BN we identify, us-\ning STM, point defects which have intragap electronic levels around 2.0 eV below the\nFermi level.\nIntroduction\nHexagonal boron nitride (h-BN) is a layered compound that is isomorphous with graphite.\nIn its bulk form, h-BN is formed from monolayers composed of boron and nitrogen atoms\nin a hexagonal sp2covalent lattice that are organized vertically by van der Waals (vdW)\ninteractions.1{3With an optical band gap of about 6 eV, and an indirect-to-direct band\ngap crossover in the transition from bulk to monolayer,4{6h-BN shows very bright deep\nultraviolet (DUV) emission7{9and defect mediated emission from the DUV all the way\nPage 2Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nto the near-infrared.10{14In particular, point defects have been observed to act as single-\nphoton sources.15{23Such properties place bulk h-BN and its monolayer form in the spotlight\nfor many potential applications, including DUV light emitting devices,24{26dielectric layers\nfor two-dimensional (2D) heterostructures27{29and room temperature (RT) single photon\nemitters (SPEs) for quantum technologies.30{34Similarly to other technologically-relevant\nsemiconductors, the successful application of h-BN depends on the understanding and control\nof its electronic and optical properties. However, partially due to its wide band gap and\nchallenging large area synthesis, many fundamental electronic and optical properties of h-\nBN remain elusive. Moreover, the morphology, the electronic properties and the optical\nemission of structural point defects are complex and are currently poorly understood.\nGiven the wide band gap of h-BN and its insulating character, studying its morpho-\nlogical, electronic and optical properties can be tricky, particularly for the case of a single\nmonolayer. The band structure of bulk and monolayer h-BN have been reported by Angle\nResolved Photoemission Spectroscopy (ARPES).35,36However, ARPES cannot resolve the\nconduction band structure. Electron Energy Loss spectroscopy (EELS) has already been\napplied to measure the band gap on h-BN nanotubes and h-BN monolayers, but EELS mea-\nsures the optical band gap, similarly to optical absorption.37,38Therefore, scanning tunneling\nspectroscopy (STS) becomes an appropriate approach to probe both the valence band and\nconduction band edges as well as for the determination of the electronic band gap.39,40Still,\nso far, STS has not provided a clear measurement of the electronic band gap for monolayer\nh-BN or intra-gap states related to defects. The structure of individual defects in h-BN has\nbeen explored in previous works by means of scanning tunneling microscopy (STM)41and\nhigh resolution transmission electron microscopy,42but, despite these e\u000borts, a clear correla-\ntion between electronic levels and light emission related to defects has not been established\nyet. Optically, the properties of few-layer h-BN samples have recently been investigated by\ncathodoluminescence (CL) measurements. However, the thinnest h-BN samples for which\na CL signal could be collected were six43and three44monolayers \rakes. Consequently, the\nPage 3Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nelectronic band gap value and the exciton binding energy are unaddressed experimentally\neven now, alongside the CL emission in the monolayer and electronic signatures of defect\nstates. Finally, the direct character of the band gap and optical transitions in the DUV in\nmonolayer h-BN have been recently reported,6but the optical transition in the near UV and\nvisible spectral range have not yet been extensively explored.\nHere, we show that monolayer h-BN epitaxially grown on highly oriented pyrolytic\ngraphite (HOPG) is a model system for the study of the morphological, electronic, and\noptical properties of h-BN monolayers. The weak sample-substrate interaction enables, for\nthe \frst time, the determination of the electronic band gap value by means of STS measure-\nments. Correlation between tunneling spectroscopy with DUV optical spectroscopy enables\na value for the exciton binding energy in monolayer h-BN to be estimated. Additionally,\nstructural point defects are observed by STM images. Luminescence due to defects in a\nbroad emission range is observed for the \frst time using CL in monolayer h-BN. Moreover,\ninsights on the optical signatures of defects are obtained from the interesting contrast be-\ntween in situ photoluminescence (PL) and CL. These results indicate that monolayer h-BN\ndoes not form signi\fcant interface states with HOPG, so that h-BN exhibits its fundamental\nelectronic and optical properties.\nResults and discussion\nSample and methods descriptions\nIn Figure 1, a general description of the h-BN sample investigated in this work is displayed.\nAs shown in Figure 1(a), h-BN and HOPG have an in-plane hexagonal structure with similar\nlattice parameters. The small lattice mismatch makes them highly compatible and appro-\npriate for the epitaxial growth of vertical heterostructures. Additionally, the surfaces of\nthese kinds of materials are naturally passivated without any dangling bonds. Therefore, in\nvdW epitaxial heterostructures, an atomically sharp interface is obtained, where chemical\nPage 4Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nbonds are absent and only vdW interactions are present between the sample and the sub-\nstrate.45{49The region at the interface is usually referred to as vdW gap, such as illustrated\nin Figure 1(a). Since the vdW interaction is weak, it is expected that each material in the\nheterostructure preserves most of its electronic properties. This was revealed in recent works\nby ARPES measurements. Sediri et al.50proved that when a monolayer h-BN is epitaxially\ngrown on graphene (Gr), the electronic structure of Gr remains una\u000bected under the pres-\nence of h-BN. In the same way, Pierucci et al.36demonstrated in monolayer h-BN/HOPG\nthat the electronic properties of h-BN are not perturbed signi\fcantly by the substrate. This\nmeans that in vdW heterostructures, the sample and the substrate are electronically decou-\npled, and there are no doping e\u000bects or charge transfer. All these e\u000bects make the vdW\nheterostructure of monolayer h-BN on HOPG an important model system for the study of\nthe fundamental properties of h-BN, and for applications when conductivity is relevant as\nin STM/STS and light emitting diodes.\nFigure 1: (a) Top view of the in-plane atomic structure of h-BN and HOPG. Side view\nof the monolayer h-BN/HOPG van der Waals heterostructure. AFM (b) topographic and\n(c) phase images of the epitaxial monolayer h-BN/HOPG. (d) Height pro\fle of a single\nmonolayer h-BN.\nPage 5Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nThe h-BN layers were grown on HOPG substrates by high-temperature plasma-assisted\nmolecular beam epitaxy (PA-MBE).6,51{53Figures 1(b) and (c) show large area topographic\nand phase atomic force microscopy (AFM) images, respectively, of the h-BN layer on the\nHOPG substrate, acquired with the amplitude-modulated tapping mode in ambient con-\nditions. As observed, the HOPG surface is almost totally covered by h-BN, where the\nwhite regions in the phase image correspond to uncovered HOPG surface areas. The AFM\ntopographic image in Figure 1(b) shows that the sample is predominantly composed by\nlarge terraces of monolayer h-BN with some small bilayer regions and some thick regions\naround the HOPG grain boundaries and surface steps, which appear as bright (dark) fea-\ntures in the topography (phase) zoomed image (see also Figure S1). The height pro\fle in\nFigure 1(d), taken along the white line on the h-BN/HOPG surface of Figure 1(b), shows\nthat the step height is approximately 0.4 nm, in agreement with the expected thickness of\na single monolayer h-BN. This morphology is consistent with previous reported results on\nsimilar samples.6,51{53\nThe surface morphology and electronic properties of the monolayer h-BN were investi-\ngated locally by means of STM and STS measurements performed under ultra-high vacuum\n(UHV) conditions at cryogenic temperatures using a modi\fed RHK PanScan FlowCryo mi-\ncroscope. An STM image revealing the smooth sample surface is shown in Figure 2(a), as\nobserved previously by AFM (Figure S1). Considering only the STM image, it is di\u000ecult\nto know if the imaged area corresponds to the h-BN or to the HOPG surface because they\nare morphologically very similar. Despite being isomorphic, h-BN and HOPG have strik-\ning di\u000berent electronic structures (wide band gap semiconductor and zero-gap conductor,\nrespectively) and therefore very di\u000berent local density of states (LDOS) to be measured by\nSTS. The gray curve in Figure 2(b) is a typical STS spectrum obtained in the region of\nFigure 2(a). This dI/dV curve shows an electronic band gap of \u00186 eV, de\fned by the low\ndi\u000berential conductance range from \u0018-3 V to\u0018+3V, con\frming that the scanned region\ncorresponds to a h-BN covered surface. Considering the general morphology of the h-BN\nPage 6Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nFigure 2: (a) STM image of the monolayer h-BN/HOPG surface (0.8 nA, 0.2 V, 80 K).\nScale bar of 30 nm. (b) STS curves at 80 K (The feedback loop was disabled at -6.2 V, -150\npA.). Schematic illustration of the experimental setup for (c) in-situ PL/Raman and (d)\nSTM-CL measurements in the STM con\fguration. The spectra shown here are uncorrected\nraw data.\ngrown on HOPG observed by AFM and the fact the the majority of the sample is covered by\nmonolayers, the h-BN thickness in smooth regions as that one in Figure 2(a) is a considered\nto be a monolayer. In other regions with similar surface morphology, the STS curves exhibit\nthe characteristic parabolic shape around the Fermi level of the LDOS of graphite,54,55as in\nthe blue plot in Figure 2(b).\nThe STM was adapted to include a high numerical aperture light collection and injection\nsystem with optimized transmission. This system uses an o\u000b-axis parabolic mirror inserted\naround the tunneling junction and in situ PL and Raman spectroscopies can be performed\nusing the light injection and collection system as illustrated in Figure 2(c). In this exper-\nimental con\fguration, the STM tip is retracted from the sample surface and the light of a\ngreen (532 nm) laser diode is injected into the STM junction under UHV conditions. Addi-\ntionally, STM-induced light emission or the CL response of the sample can be investigated\nPage 7Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nusing the setup, see Figure 2(d). For recording the CL signal, the STM is operated in \feld\nemission mode using an external voltage source of (0-500) V for the local excitation of the\nsample by electron bombardment (STM-CL). The PL/Raman and CL spectra, shown in\nFigure 2(c) and (d), correspond to the spectroscopic raw data that need to be corrected for\nthe proper interpretation, as explained in the Supplementary materials (SM). This setup has\nalso been previously used to investigate WSe 2and MoS 2monolayers.56,57\nElectronic band gap, optical band gap and exciton binding energy\nSince monolayer h-BN is an atomically thin and a wide band gap material, access to its\nelectronic structure is always a highly challenging experimental and theoretical task. Within\nthe last twenty years, the h-BN electronic band gap has been calculated a number of times\nwithin the GW approximation, while the optical response has been calculated using the\nBethe-Salpeter equation.58{61In this approximation, the Green's function for the quasi-\nparticles are evaluated within the screened Coulomb potential in a way to consider many-\nbody problems of the interacting electrons.62However, non self-consistent GW calculations\n(G0W0), which start from DFT orbitals and perform the calculation only once, typically\nprovide underevaluated values for the electronic band gap and, in the case of h-BN, a rigid\nshift of the optical spectra is required for a direct comparison with experiments. Only\nvery recently, self-consistent GW calculations, which update interactively both the wave\nfunctions and the eigenvalues, managed a good agreement with synchrotron ellipsometry\nexperiments for bulk h-BN.63In the case of free-standing monolayer h-BN and using the\nGW 0approach, which updates interactively the eigenvalues only in the Green's function, the\nelectronic band gap has been evaluated as 8.2 eV and the related optical gap at 6.1 eV.60,64\nSlightly higher values might be expected within the full self-consistent GW approach. On the\nother hand, at the moment, there is no experimental evidence with e\u000bective measurements\nof the electronic band gap in both bulk or monolayer h-BN. A particular reason for the lack\nof such results is the technical challenge regarding the accomplishment of experiments for\nPage 8Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nDUV wavelengths. The electronic properties and band structure of epitaxial monolayer h-\nBN on graphite have been studied recently by ARPES measurements.36This study reported\nimportant information about band alignment, Fermi level and Valence Band (VB) positions.\nHowever, it is noteworthy that ARPES measurements only resolve \flled states,35,36,65,66and\ntherefore the description of the whole band structure and electronic band gap of monolayer\nh-BN are still open questions.\nOne technique able to probe both \flled and empty states, with a complete information\nabout the LDOS, doping e\u000bects and charge transfer is scanning tunneling spectroscopy or\nSTS.40,67{71Previous STM/STS measurements on h-BN have not given a clear answer for\nthe electronic band gap. The main reason for that is because most of the STM/STS studies\nhave been carried out on h-BN samples grown directly on metallic substrates.72Nowadays,\nit is well documented that the growth of a 2D material on a metallic substrate leads to a\nband gap renormalization due to strong interactions with the substrate, which is related to\nthe dielectric screening by the metal and/or the formation of additional interface electronic\nstates, including hybridization among others.57,73{76These e\u000bects reduce the electronic band\ngap value to about 3 eV for h-BN on Ru(0001)75and on Re(0001)77surfaces, to\u00184 eV\nfor h-BN on Au(001),78and to\u00185 eV for the case of h-BN on Rh(111),79,80Ir(111),81and\nCu(111)82substrates. In addition, performing scanning tunneling experiments on h-BN can\nbe complex due to its expected insulating character. For instance, Wong, et al used Gr\nas conducting layer on top of bulk h-BN.41Using the Gr as conducting electrode a\u000bects\nconsiderably STM images and make STS spectra di\u000ecult to interpret, particularly with\nrespect to the h-BN band gap. Interestingly, when h-BN is grown on Gr/Cu(111), electronic\nstates of the metallic substrate are observed on the Gr and also on the h-BN layer.76In this\ncase, intragap states due to the copper support are measured by STS and the h-BN band\ngap is not observed. For vdW epitaxial heterostructures, the sample-substrate interaction is\nnot an issue because, as mentioned above, there is a sharp interface between the monolayer\nh-BN and the HOPG substrate governed by weak vdW bonds (vdW gap). Therefore, h-BN\nPage 9Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\non HOPG can be considered as being electronically decoupled from the substrate, and is\nnot expected to demonstrate intragap states coming from graphite, but only e\u000bect of the\ndielectric environment surrounding the monolayer.\nFigure 3(a) and (b) shows STM and STS results acquired at 80 K in the monolayer h-\nBN/HOPG. A STM image of a smooth and defect-free region of the sample surface together\nwith an insert showing a moir\u0013 e structure with with a periodicity of approximately 2.4 nm are\ndisplayed in Figure 3(a). This moir\u0013 e pattern corresponds to a local rotation of approximately\n5.8 °between the monolayer h-BN and the HOPG surface in a particular region, as recently\nobserved on similar samples.51The apparent height pro\fle of the moir\u0013 e is presented in\nFigure 3(b). Detailed studies of the moir\u0013 e structure in this system were reported in recent\nworks.51,83\nThe LDOS of the monolayer h-BN has been investigated by performing STS measure-\nments on the defect-free region at di\u000berent tip positions. The tunneling processes that probe\nboth empty and \flled states by STS in the monolayer h-BN are illustrated in Figure 3(c)\nand (d), respectively. The alignment between the Fermi level in h-BN and HOPG at 0 V\nwas demonstrated in the results depicted in Figure 2(b). Thus, when a positive sample bias\nranging from 0 V to +6 V is applied, the Fermi level in the tip is shifted with respect to the\nFermi level in h-BN/HOPG, and, as a consequence, electrons tunnel from the tip to empty\nstates of the sample, such as shown in Figure 3(c). The red curve in the \fgure represents the\nmeasured dI/dV for positive biases, which is a direct measurement of the local empty DOS.\nIt can be observed that the LDOS increases from biases above \u00183.0 V, which means that\nelectrons are tunneling only for states in the Conduction Band (CB) of h-BN. Below \u00183.0\nV the curve is \rat (the tunneling current is equal to \u00180 pA), because there are no available\nelectronic states inside the h-BN electronic band gap for electrons to tunnel to. The same\nprocess happens when the polarity of the bias is changed to negative values in order to\nprobe \flled states, as illustrated in Figure 3(d). It is worthwhile to note that, during typical\nSTS measurements, the voltage ramps while the current is measured but the tip-sample dis-\nPage 10Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nFigure 3: (a) STM image of a smooth and defect-free monolayer h-BN region (0.66 nA, 3.8\nV, 80 K). Scale bar of 30 nm. Inset: moir\u0013 e structure (-0.3 nA, -1.0 V, 16 K). Scale bar of\n5 nm. (b) Height pro\fle associated to the moir\u0013 e. Schematic illustration of the tunneling\nprocess for (c) positive and (d) negative biases. The red curve corresponds to the LDOS\nmeasured by STS as dI/dV. The blue arrow indicates the tunneling direction. (e) I-V and\ndI/dV typical curves obtained on a smooth and defect-free monolayer h-BN region and at 80\nK (The feedback loop was disabled at -6.2 V, -150 pA). The electronic band gap and band\nedges are indicated with the values obtained of a statistical analysis of 480 individual curves.\nGlobal measurements of the (f) optical absorption at RT and (g) PL at 10 K. (h) Energy\nlevels diagram proposed for defect-free monolayer h-BN. In the \fgure VBM, E g, Eoptand\nEbcorrespond to the valence band maximum, the electronic band gap, the optical band gap\nand the exciton binding energy, respectively.\ntance is not adjusted. Therefore, the tip-sample distance is set by the tunneling current and\nsample bias prior to disabling the tip-sample distance feedback loop and starting the STS\nPage 11Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nmeasurement (stable tunneling parameters). Here, we used typically -150 pA at -6.2 V (see\nMethods section for more details) as the stable condition prior to starting the acquisition\nof the scanning tunneling spectra in order to ensure reproducible and reliable spectra. If\nthe tunneling current is set to slightly higher or lower values (from -50 to -300 pA) before\ndisabling the feedback loop, the obtained spectra are compatible to those measured using\n-150 pA, see Figure S2.\nIt is relevant to mention that direct tunneling from or to the HOPG is possible even\nthrough the monolayer h-BN. In fact, this e\u000bect can be observed when STS curves were\nrecorded stabilizing the tunneling junction at high tunneling currents ( \u0015600 pA) with -6.2\nV of sample bias, as discussed in the SM, see Figure S2. A possible explanation for this\nobservation is that for a \fxed sample bias and higher stabilization currents, the tip-sample\ndistance becomes shorter, which means that the tunneling barrier is narrower, and then the\nprobability of the direct tunneling from or to the substrate is not negligible. This leads\nto the measurement of an apparent reduced band gap, because states of the HOPG near\nthe band edges of h-BN are observed in the STS curve for h-BN, as can be seen in Figure\nS2(b). The direct tunneling from/to the HOPG substrate for short tip-sample distance\nin STS measurements on vdW epitaxy MoS 2/HOPG and WSe 2/HOPG samples have been\ndemonstrated by Chiu et al.84In that work, it was also demonstrated that the LDOS of\nHOPG can be measured by tunneling electrons through the TMD layer if the stabilizing\nbias for the STS acquisition lays inside the band gap region of the TMD. For wide band\ngap materials such as h-BN, using appropriate tunnel junction conditions before starting the\nSTS is therefore crucial.\nFigure 3(e) shows typical STS results obtained from a smooth and defect-free h-BN\nsample surface region. The top panel of the \fgure shows the I-V curve plotted together with\nthe dI/dV tunneling spectrum. The electronic band gap of the h-BN sample is de\fned by the\nbias range at which the current and dI/dV are close to their background values. The band\nedges and band gap values were extracted from the dI/dV curve in logarithmic scale,68,85\nPage 12Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nas represented in the bottom panel of Figure 3(e). The values shown in the \fgure were\nobtained following the method and the statistical analysis explained in the SM, Figures S3\nand S4, which considers a total number of 480 individual dI/dV curves similar to Figure 3(e)\ntaken at di\u000berent tip positions. The \fnal results are a Valence Band Maximum (VBM)\nand a Conduction Band Minimum (CBM) positioned at ( \u00003:4\u00060:1) eV and (3 :4\u00060:1)\neV, respectively, for monolayer h-BN on HOPG. This results in an electronic band gap of\n(6:8\u00060:2) eV. Also, the Fermi level is in the middle of the band gap, indicating that the\ndefect-free monolayer h-BN regions, as shown in Figure 3(a), are undoped and therefore\nthere is no charge transfer with the HOPG. ARPES found the VBM position close to the\n\u0018-2.8 eV on a similar h-BN sample.36This ARPES result for the VBM is slightly di\u000berent\nthan the value measured by STS (-3.4 eV). This di\u000berence is possibly caused by some degree\nof p-type doping on the sample and/or some level of hole excess coming from a insu\u000ecient\ncompensation of the photo-emitted electrons. On 3-monolayer thick h-BN samples, ARPES\ndetermined the VBM at -3.2 eV, which, considering the band structure for 3-monolayer thick\nh-BN and the expected electronic band gap opening for a monolayer h-BN, agrees well with\nthe results presented here.86From the experimental point of view, STS is not sensitive to the\ndirect or indirect nature of the electronic band gap. Nonetheless, recently, it was con\frmed\nthe recombination of direct excitons in monolayer h-BN,6which suggests that this is a direct\nband gap material.\nThe electronic band gap value found here for monolayer h-BN is larger, as expected from\nsimulations,61than values for bulk single crystal and thin \flms of h-BN, as obtained indi-\nrectly by photocurrent spectroscopy, which results in an electronic band gap of about 6.45\neV.87,88Regarding theoretical predictions for free-standing monolayer h-BN, GW 0calcula-\ntions indicate a band gap of 8.2 eV at the K point of the Brillouin zone with an exciton\nbinding energy of 2.1 eV, giving an optical band gap of 6.1 eV.60,64This value is close to the\noptical onset of 6.3 eV measured by EELS on a free-standing monolayer.38Simulations for\nmonolayer h-BN on top of graphene have shown a strong renormalization of the electronic\nPage 13Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nband gap, of the order of 1 eV, due to strong screening e\u000bects of the substrate.89,90The\nsimulations also show that the electronic band gap renormalization of monolayer h-BN is\nbasically the same on top of graphene and on top of graphite.89This last scenario is the\nperfect model for our system of monolayer h-BN on HOPG and gives a strong support for\nthe band gap value measured here by STS.\nOptical absorption and PL were performed in the DUV to determine the optical (ex-\ncitonic) band gap of monolayer h-BN on HOPG (see details in SM). Figure 3(f) and (g)\nshow measurements of the optical absorption and PL (spot-size of 200 \u0016m and 50\u0016m, re-\nspectively) obtained at RT and at 10 K, respectively. The optical absorption spectrum was\ncalculated from spectroscopic ellipsometry data. The PL in the DUV employs a 6.4 eV\nlaser obtained from the fourth harmonic of a Ti:Sa oscillator. A systematic study on the\ncomplex DUV absorption and emission of high-temperature PA-MBE h-BN on HOPG can\nbe found in recent reports.6,52The spectrum in Figure 3(f) shows an abrupt change in the\nabsorption coe\u000ecient that suggests an optical band gap of above 5.7 eV. Furthermore, the\nPL in Figure 3(g) shows the direct fundamental exciton transition at 6.1 eV (optical band\ngap) in monolayer h-BN, besides other emission lines associated with transitions in few-layer\nand defective regions around 5.9 eV and 5.6 eV, respectively. These results are in excellent\nagreement with previous experiments by Elias et al.6Moreover, this value for the optical\nband gap is consistent with recent simulation and indicate, as expected, an invariance of the\nexciton emission.60,64,68,90\nThe knowledge of both the electronic band gap and the optical band gap is essential\nto the design of applications in optoelectronics and light emitting devices employing large\nexciton binding energy materials, like the monolayer h-BN. Considering the h-BN electronic\nband gap found by STS and the energy observed for the exciton recombination, the obtained\nexciton binding energy is (0 :7\u00060:2) eV. The comparison between this experimental evaluation\nand the theoretical predicted value of 2.1 eV for the free-standing monolayer, suggest a\nstrong renormalization, of about 1.4 eV, of the exciton binding energy due to substrate\nPage 14Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nscreening e\u000bects.64Moreover, the observed renormalization of the exciton binding energy is\nexpected to be the same on the electronic band gap, which nearly leads to an invariance\nof the optical onset.68,90This renormalization is in close agreement with simulations for\nmonolayer h-BN on top of graphite.89The similar shift on the electronic band gap and on the\nexciton biding energy agrees well with simulations for monolayer h-BN and with experiments\nand simulations for monolayer MoSe 2.68,90Based on all the spectroscopic data, Figure 3(h)\nsummarizes the energy levels determined here for a defect-free region of monolayer h-BN on\nHOPG. In Figure 3(h), the h-BN electronic band gap is indicated as (6 :8\u00060:2) eV, while\nthe h-BN optical band gap is 6.1 eV. This leads to a Frenkel exciton with a binding energy\nestimated as (0 :7\u00060:2) eV.\nElectronic Structure and Light Emission related to point defects\nThe wide electronic band gap in h-BN allows the observation of several optical transitions\ninvolving di\u000berent intragap states associated with defects.10,12,13,91Therefore, light emission\nin h-BN may be dominated by structural defects. Point defects have a particular interest\ndue to their characteristic quantum emission at RT in a broad range of wavelengths.16{18,20\nDi\u000berent types of point defects have been proposed as being the sources of the single photon\nemission observed in h-BN.14,91,92However, their morphological and electronic signatures\nhave not been measured directly. Furthermore, a direct correlation with the optical emission\nis missing. Even though STM is a tool able to identify atomic defects in 2D materials,56,93{96\nSTM/STS measurements of point defects in monolayer h-BN have not been reported yet.\nThe properties of individual point defects have been investigated only by simulated STM\nimages in recent theoretical works.97{100Here some results obtained on defective regions of\nthe sample are presented in order to help acquiring a better understanding of the impact of\npoint defects in the properties of monolayer h-BN.\nFigure 4(a) shows a STM image of the monolayer h-BN/HOPG surface, where point\ndefects are revealed. These defects are bright spots of about 1 to 2 nm of diameter, according\nPage 15Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nFigure 4: (a) STM image of bright point defects in monolayer h-BN (0.8 nA, 0.8 V, 80 K).\nScale bar of 30 nm. (b) Heigth pro\fle of some individual defects. (c) Typical STS curve\nobtained on the defective region at 80 K (The feedback loop was disabled at -6.2 V, -150\npA). (d) In-situ PL and STM-CL spectra in which Phonon Side-Bands (PSB) are observed\nnear Zero Phonon Loss (ZPL) peaks. (e) In-situ Raman spectrum at RT. Inset: Lorenztian\n\ftting of the peak at 1350 cm\u00001.\nto the full width at half maximum in the height pro\fles of Figure 4(b), see also Figure S5.\nDefects in h-BN on HOPG were already observed by Summer\feld et al.51using conductive\nAFM and STM measurements, and were attributed to defects in the HOPG substrate,\ncreated by the active nitrogen plasma irradiation damage during the sample growth process.\nIn another study, bright point defects relating to possible carbon impurities were imaged\nby STM in Gr capped bulk h-BN.41In order to inspect the electronic signature of the\npoint defects found here, STS curves were acquired. Figure 4(c) displays a typical tunneling\nspectrum recorded when the STM tip is paced close to such a defect. By plotting the STS\ncurve in logarithmic scale, it can be noted that the bands onset are not linear, in contrast\nPage 16Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nto the ones observed in defect-free regions (see Figure 3(e) and Figure S3(c)). The curve\nin Figure 4(c) presents three main resonances at \u0018\u00003:5 V,\u0018\u00002:0 V and\u0018+3:5 V. By\ncomparing with the results for defect-free regions, the resonances at \u00063:5 V are located on\ntop of the CB and VB edges, respectively, indicating that such defects have energy levels\nclose to the band edges. Besides the spectra on Figure 4(c), the electronic level near -2.0\nV was observed in several other tunneling spectra obtained on regions with point defects,\nsee Figure S5. Therefore, this energy level is attributed to the observed defects. The fact\nthat the STS curves show roughly the expected band gap for h-BN even on defective regions\nsuggests that the scanned regions corresponds to a h-BN covered area and the observed point\ndefects are in the monolayer h-BN.\nA good starting point to interpret the imaged defects and the STS results is to consider\nthat in the system of monolayer h-BN on HOPG, the atoms of Boron (B), Nitrogen (N)\nand Carbon (C) are present. Thus, substitutional defects such as C Band C N, carbon dimer\ndefects like C BCN, or carbon anti-sites vacancies of the types V BCN, and V NCBare possi-\nble, and also are the main point defect candidates to be considered. Indeed, the random\ndistribution of C atoms in the h-BN lattice is energetically possible,101and on the other\nhand, the high temperature and plasma assisted growth process can create vacancies during\nthe sample preparation.19,51From the electronic point of view, band structure calculations\nshow that for the LDOS in monolayer h-BN the states near the VBM are concentrated in\nthe N sites, while the states near the CBM are in the B sites.101,102Impurities occupying\nN or B sites could act as acceptor-like or donor-like impurities, respectively. There are sev-\neral theoretical reports exploring the doping e\u000bects of C impurities in single layers of h-BN.\nIn those works, it is demonstrated that C Band C Ndefects induce n-doping and p-doping,\nrespectively.97,99,103,104In particular, C Ndefects introduce an acceptor state localized at 2.0\neV below the Fermi level and between 1-2 eV above the VBM.97,103,105The density of this\ndefect state depends on the C concentration,101and the energy position can be shifted due\nto local strain e\u000bects,97,106,107and also by band bending e\u000bects.93More recently, it also has\nPage 17Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nbeen predicted that the C BCNdimer introduces a stable and neutral intragap state localized\nat\u00180.8 eV above the VBM.108,109\nIt is important to point out that our STM results are consistent with the predictions\nmade by Fujimoto et al.97and by Haga et al.99,100in their works on simulated STM images of\nindividual point defects in h-BN and Gr vertical heterostructures. In those reports, the bright\nappearance of C Npoint defects were simulated for negative tunneling biases. The calculated\nSTM images of C-doped monolayer h-BN show that C atoms produce a redistribution of\nthe local electron density around the defects, being the electron density in C Nhigher than\nin C B, which results in STM images where C Ndefects look as bright spots when compared\nwith the image of C B.97,99Moreover, the local electron density, and as a consequence, the\nsize of the bright spots related to C Ndefects can be a\u000bected by the stacking and moir\u0013 e\npattern between the h-BN and Gr.99,100Defects imaged with negative bias can be observed\nin Figure S5. These theoretical considerations indicate that the defects identi\fed by STM\nare consistent with defects involving C on N sites, i. e. C N, CBCNand V BCN, even if other\nkinds of defect can not be excluded. Nonetheless, it is noteworthy that a detailed study of the\nmorphological and electric structure of individual point defects in h-BN is required, which\ncan be achieved by performing STM/AFM imaging and STS measurements at cryogenic\ntemperatures, such as what has been done in the case of point defects in single layers of\ntransition metal dichalcogenides.110{113\nIn order to obtain more insights on the observed defects, in situ PL and STM-CL ex-\nperiments were carried out at 100 K and 300 K, using the experimental setup described in\nFigure 2. The raw luminescence data were presented in Figure 2(c) and (d), but for the\nproper interpretation, the spectra were corrected following the procedure explained in SM\nand the results are presented in Figure 4(d). The PL spectra show a main peak at 2.10\neV, which is best resolved at 100 K, and two shoulders around 1.96 eV and 1.81 eV. These\ntransitions are related to the zero-phonon line (ZPL) and two phonon replicas or phonon\nside bands (PSB) of carbon-related defects, which behave as single photon sources.22Simu-\nPage 18Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nlations suggest V BCNor C BVNas defects at the origin of this emission.22,106The width of\nthe PL peaks is larger than that typically observed in thick h-BN \rakes, but in agreement\nwith reports on monolayers.15Also in the PL spectra, two sharp peaks are observed and\nlabeled with a red star symbol. These peaks correspond to the Raman response depicted\nfully in Figure 4(e), where the stronger peaks around 1580 cm\u00001and 2700 cm\u00001are the G\nand 2D Raman bands of HOPG, respectively.114In addition, an asymmetric and weak peak\nis observed at 1350 cm\u00001, and as shown in the inset \fgure, this peak has two contributions.\nOne contribution at 1341 cm\u00001associated to the D Raman mode of HOPG and another one\nat 1362 cm\u00001corresponding to the E 2gRaman mode of monolayer h-BN, in agreement with\nrecent reported results.53\nThe CL spectra in the bottom panel of Figure 4(d) show a broad luminescence that\nranges from the near infrared to the UV. As in PL, the peaks are better resolved at 100\nK than at 300 K. Again, the optical transitions associated to carbon-related defects are\nobserved with the ZPL at 2.10 eV and the PSB at 1.81 eV, respectively, as seen in PL.\nAt higher energies, the CL spectra present a series of peaks at 3.90 eV, 3.70 eV and 3.43\neV, which have been reported as being phonon replicas of a deep well-known carbon-related\ndefect level at 4.1 eV.105,115Unfortunately, the emission at 4.1 eV has not been resolved here\ndue to the transmission function of the setup. The spectra only could be corrected for the\ninstrument response function up to 300 nm (4.1 eV). However, the PSB were observed, which\nmeans that the transitions at 4.1 eV were also present in the emitted spectra. Interestingly,\nthe light emission at 4.1 eV has been reported as being an optical transition related to\nthe single photon emission.16Recent literature suggests that this emission is associated\nwith C BCNdefects, carbon dimers, but other defects like C Nare also considered in some\ncases.108,109,116,117The most intense features observed in the CL spectra of Figure 4(d) are\nthe peaks at 3.08, 2.65 and 2.37 eV, with shifts of 430 meV and 280 meV between them.\nSuch peaks possibly have distinct origins related to the several possible defects, particularly\nin the presence of carbon.116Emissions in this energy range have been observed in bulk\nPage 19Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nh-BN by performing CL and PL experiments, and are associated to carbon21,105impurities\nor nitrogen vacancy type-centers.118,119\nConclusions\nWe used low temperature UHV STM together with an optimized light collection and injection\ndevice to study the morphological, electronic, and optical properties of a monolayer h-BN\nepitaxially grown on HOPG. The STM images reveal h-BN regions free of defects and regions\nwith point defects. An electronic band gap of (6 :8\u00060:2) eV was determined by performing\nSTS measurements on defect-free monolayer h-BN. This result, combined with the h-BN\noptical band gap of 6.1 eV from the exciton transition, leads to a binding energy for the\nFrenkel exciton of (0 :7\u00060:2) eV. Additionally, bright point defects were observed in h-\nBN by STM imaging. The STS indicates an acceptor level around -2 eV related to the\npresence of the observed defects. PL and CL have shown the emission typically associated\nto carbon-related defects at 2.1 eV. Besides that, emissions at 3.08 eV and photon side band\npossibly associated to an emission at 4.1 eV were observed by CL on monolayer h-BN. These\nresults indicate the simultaneous presence of more than one kind of carbon-related defect.\nWe consider that the \fndings presented in this work could help in the understanding of the\nfundamental properties of monolayer h-BN, as well as in the identi\fcation at the atomic level\nof sources responsible for the SPEs in h-BN samples. Moreover, h-BN on HOPG represents\nan excellent platform to study individual defects with respect to their morphology, electronic\nand optical properties.\nPage 20Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nMethods\nSample preparations\nMonolayer h-BN was grown on HOPG substrate by the high-temperature plasma-assisted\nmolecular beam epitaxy (PA-MBE) method.6,51{53This sample preparation method allows to\nproduce monolayer and few-layer h-BN with atomically \rat surfaces and monolayer control\nof the sample thickness. The h-BN thickness and coverage can be controlled by substrate\ntemperature, boron:nitrogen \rux ratio and growth time. In particular, the sample investi-\ngated here was grown at a substrate temperature of about 1390\u000eC with a high-temperature\ne\u000busion Knudsen cell for boron and a standard Veeco radio-frequency plasma source for ac-\ntive nitrogen. More details about the sample growth conditions and the MBE system can\nbe found elsewhere.53,120{122\nSTM/STS\nSTM and STS measurements were performed under ultra-high vacuum (UHV) conditions\nat low-temperatures using a modi\fed RHK PanScan FlowCryo microscope. This STM was\nadapted to receive a high numerical aperture light collection and injection system with\noptimized transmission. In this setup, imaging can be associated to electronic and optical\nspectroscopies as illustrated in Figure 2.\nPrior to STM/STS measurements, the h-BN sample was annealed at 773 K under a base\npressure of 1.9x10\u000010mbar for about six hours, in order to eliminate surface impurities and\ncontaminants. STM images were acquired in the constant-current mode using a grounded\ntungsten tip and a bias voltage applied to the h-BN sample. Tungsten tips were prepared\nby electrochemical etching with a NaOH solution. After preparation, the tips were inserted\nin the UHV chamber to be subjected to thermal annealing and argon sputtering. This\nprocedure allows the removal tungsten oxide layers. For the acquisition of STS curves,\ndi\u000berential conductance (dI/dV) measurements at 80 K were carried out using the lock-in\nPage 21Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\ntechnique, with a bias modulation of 80 mV of amplitude and 800 Hz of frequency.\nTo perform STS measurements, the tunnel junction was stabilized at a sample bias outside\nthe electronic band gap of the monolayer h-BN. In most measurements, we stabilized the\ntip at -150 pA and -6.2 V on a speci\fc tip position before disabling the tip-sample distance\nfeedback loop. Using these parameters, the tunnel current is e\u000bectively related to states on\nthe h-BN. Each tunneling spectrum was recorded by turning o\u000b the tip distance feedback\nloop and ramping the bias from -6.0 V to 6.0 V in steps of 50 mV and a dwell time of 50 ms.\nThis dwell time is needed to avoid overestimating the electronic band gap due to low signal\nto noise ratio near the band edges. A detailed description of the STS analysis is presented\nin the SM, Figures S3 and S4.\nIn situ PL, Raman and CL\nPL and Raman were excited with a laser diode of 532 nm. The spot size of the light focused\nby the mirror on the sample surface is \u00182\u0016m. The PL/Raman signal emitted by the sample\nis collected by the mirror and sent as parallel rays towards an optical setup outside the UHV\nSTM chamber.\nCL, also called STM-CL, was performed by operation of the STM in \feld emission mode.\nIn this case, the STM tip was retracted by approximately 150 nm from the h-BN sample\nsurface. A high bias voltage, between 150-180 V, was applied to the sample, which caused\na \feld emission currents of 5-10 \u0016A. In this STM operation mode, the spatial (lateral)\nresolution is roughly similar to the tip-to-sample distance.45Therefore, the obtained CL\nspectra refer to a region of about one hundred nanometers wide of the monolayer h-BN.\nAcknowledgement\nThis work was supported by the Funda\u0018 c~ ao de Amparo \u0012 a Pesquisa do Estado de S~ ao Paulo\n(FAPESP) Projects 14/23399-9 and 18/08543-7. This work at Nottingham was supported by\nPage 22Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nthe Engineering and Physical Sciences Research Council UK (Grant Numbers EP/K040243/1\nand EP/P019080/1). We also thank the University of Nottingham Propulsion Futures Bea-\ncon for funding towards this research. PHB thanks the Leverhulme Trust for the award\nof a Research Fellowship (RF-2019-460). This work was \fnancially supported by the net-\nwork GaNeX (ANR-11-LABX-0014), the ZEOLIGHT project (ANR-19-CE08-0016), and the\nBONASPES project (ANR-19-CE30-0007).\nSupplementary materials available\nAFM measurements and sample surface morphology. Tunneling parameters and STS. Elec-\ntronic band gap determination and statistical analysis. Optical absorption. STM/STS of\npoints defects. Correction in the luminescence spectra.\nReferences\n(1) Adachi, S. Optical Constants of Crystalline and Amorphous Semiconductors: Numer-\nical Data and Graphical Information ; Springer US: Boston, MA, 1999; pp 127{136.\n(2) Alem, N.; Erni, R.; Kisielowski, C.; Rossell, M. D.; Gannett, W.; Zettl, A. Atomically\nthin hexagonal boron nitride probed by ultrahigh-resolution transmission electron mi-\ncroscopy. Phys. Rev. B 2009 ,80, 155425.\n(3) Bhimanapati, G.; Glavin, N.; Robinson, J. In 2D Materials ; Iacopi, F., Boeckl, J. J.,\nJagadish, C., Eds.; Semiconductors and Semimetals; Elsevier, 2016; Vol. 95; pp 101 {\n147.\n(4) Watanabe, K.; Taniguchi, T.; Kanda, H. Direct-bandgap properties and evidence for\nultraviolet lasing of hexagonal boron nitride single crystal. Nature Materials 2004 ,3,\n404{409.\nPage 23Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\n(5) Cassabois, G.; Valvin, P.; Gil, B. Hexagonal boron nitride is an indirect bandgap\nsemiconductor. Nature Photonics 2016 ,10, 262{266.\n(6) Elias, C.; Valvin, P.; Pelini, T.; Summer\feld, A.; Mellor, C. J.; Cheng, T. S.; Eaves, L.;\nFoxon, C. T.; Beton, P. H.; Novikov, S. V.; Gil, B.; Cassabois, G. Direct band-gap\ncrossover in epitaxial monolayer boron nitride. Nature Communications 2019 ,10,\n2639.\n(7) Watanabe, K.; Taniguchi, T.; Miya, K.; Sato, Y.; Nakamura, K.; Niiyama, T.;\nTaniguchi, M. Hexagonal boron nitride as a new ultraviolet luminescent material and\nits application|Fluorescence properties of hBN single-crystal powder. Diamond and\nRelated Materials 2011 ,20, 849{852.\n(8) Kubota, Y.; Watanabe, K.; Tsuda, O.; Taniguchi, T. Deep Ultraviolet Light-Emitting\nHexagonal Boron Nitride Synthesized at Atmospheric Pressure. Science 2007 ,317,\n932{934.\n(9) Valvin, P.; Pelini, T.; Cassabois, G.; Zobelli, A.; Li, J.; Edgar, J. H.; Gil, B. Deep\nultraviolet hyperspectral cryomicroscopy in boron nitride: Photoluminescence in crys-\ntals with an ultra-low defect density. AIP Advances 2020 ,10, 075025.\n(10) Museur, L.; Feldbach, E.; Kanaev, A. Defect-related photoluminescence of hexagonal\nboron nitride. Phys. Rev. B 2008 ,78, 155204.\n(11) Bourrellier, R.; Amato, M.; Galv~ ao Tizei, L. H.; Giorgetti, C.; Gloter, A.; Heggie, M. I.;\nMarch, K.; St\u0013 ephan, O.; Reining, L.; Kociak, M.; Zobelli, A. Nanometric Resolved\nLuminescence in h-BN Flakes: Excitons and Stacking Order. ACS Photonics 2014 ,\n1, 857{862.\n(12) Schu\u0013 e, L.; Stenger, I.; Fossard, F.; Loiseau, A.; Barjon, J. Characterization methods\ndedicated to nanometer-thick hBN layers. 2D Materials 2016 ,4, 015028.\nPage 24Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\n(13) Vuong, T. Q. P.; Cassabois, G.; Valvin, P.; Ouerghi, A.; Chassagneux, Y.; Voisin, C.;\nGil, B. Phonon-Photon Mapping in a Color Center in Hexagonal Boron Nitride. Phys.\nRev. Lett. 2016 ,117, 097402.\n(14) Gil, B.; Cassabois, G.; Cusco, R.; Fugallo, G.; Artus, L. Boron nitride for excitonics,\nnano photonics, and quantum technologies. Nanophotonics 2020 ,9, 3483{3504.\n(15) Tran, T. T.; Bray, K.; Ford, M. J.; Toth, M.; Aharonovich, I. Quantum emission from\nhexagonal boron nitride monolayers. Nature Nanotechnology 2016 ,11, 37{41.\n(16) Bourrellier, R.; Meuret, S.; Tararan, A.; St\u0013 ephan, O.; Kociak, M.; Tizei, L. H. G.;\nZobelli, A. Bright UV Single Photon Emission at Point Defects in h-BN. Nano Letters\n2016 ,16, 4317{4321, PMID: 27299915.\n(17) Tran, T. T.; Elbadawi, C.; Totonjian, D.; Lobo, C. J.; Grosso, G.; Moon, H.; En-\nglund, D. R.; Ford, M. J.; Aharonovich, I.; Toth, M. Robust Multicolor Single Photon\nEmission from Point Defects in Hexagonal Boron Nitride. ACS Nano 2016 ,10, 7331{\n7338, PMID: 27399936.\n(18) Tran, T. T.; Zachreson, C.; Berhane, A. M.; Bray, K.; Sandstrom, R. G.; Li, L. H.;\nTaniguchi, T.; Watanabe, K.; Aharonovich, I.; Toth, M. Quantum Emission from\nDefects in Single-Crystalline Hexagonal Boron Nitride. Phys. Rev. Applied 2016 ,5,\n034005.\n(19) Xu, Z.-Q. et al. Single photon emission from plasma treated 2D hexagonal boron\nnitride. Nanoscale 2018 ,10, 7957{7965.\n(20) Tan, Q.-H.; Xu, K.-X.; Liu, X.-L.; Guo, D.; Xue, Y.-Z.; Ren, S.-L.; Gao, Y.-F.;\nDou, X.-M.; Sun, B.-Q.; Deng, H.-X.; Tan, P.-H.; Zhang, J. Ultraviolet to Near-\ninfrared Single Photon Emitters in hBN. arXiv:1908.06578 [cond-mat, physics:physics]\n2019 , arXiv: 1908.06578.\nPage 25Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\n(21) Hayee, F.; Yu, L.; Zhang, J. L.; Ciccarino, C. J.; Nguyen, M.; Marshall, A. F.;\nAharonovich, I.; Vu\u0014 ckovi\u0013 c, J.; Narang, P.; Heinz, T. F.; Dionne, J. A. Revealing\nmultiple classes of stable quantum emitters in hexagonal boron nitride with correlated\noptical and electron microscopy. Nature Materials 2020 ,19, 534{539.\n(22) Mendelson, N. et al. Identifying carbon as the source of visible single-photon emission\nfrom hexagonal boron nitride. Nature Materials 2020 ,20, 321{328.\n(23) Mart\u0013 \u0010nez, L. J.; Pelini, T.; Waselowski, V.; Maze, J. R.; Gil, B.; Cassabois, G.;\nJacques, V. E\u000ecient single photon emission from a high-purity hexagonal boron ni-\ntride crystal. Phys. Rev. B 2016 ,94, 121405.\n(24) Watanabe, K.; Taniguchi, T.; Niiyama, T.; Miya, K.; Taniguchi, M. Far-ultraviolet\nplane-emission handheld device based on hexagonal boron nitride. Nature Photonics\n2009 ,3, 591{594.\n(25) Watanabe, K.; Taniguchi, T. Hexagonal Boron Nitride as a New Ultraviolet Lumines-\ncent Material and Its Application. International Journal of Applied Ceramic Technol-\nogy2011 ,8, 977{989.\n(26) Jiang, H. X.; Lin, J. Y. Hexagonal boron nitride for deep ultraviolet photonic devices.\nSemiconductor Science and Technology 2014 ,29, 084003.\n(27) Hui, F.; Pan, C.; Shi, Y.; Ji, Y.; Grustan-Gutierrez, E.; Lanza, M. On the use of two\ndimensional hexagonal boron nitride as dielectric. Microelectronic Engineering 2016 ,\n163, 119{133.\n(28) Wang, J.; Ma, F.; Sun, M. Graphene, hexagonal boron nitride, and their heterostruc-\ntures: properties and applications. RSC Adv. 2017 ,7, 16801{16822.\n(29) Wu, L. et al. InSe/hBN/graphite heterostructure for high-performance 2D electronics\nand \rexible electronics. Nano Research 2020 ,13, 1127{1132.\nPage 26Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\n(30) Wrachtrup, J. Single photons at room temperature. Nature Nanotechnology 2016 ,11,\n7{8.\n(31) Aharonovich, I.; Englund, D.; Toth, M. Solid-state single-photon emitters. Nature\nPhotonics 2016 ,10, 631{641.\n(32) Aharonovich, I.; Toth, M. Quantum emitters in two dimensions. Science 2017 ,358,\n170{171.\n(33) Kim, S.; Duong, N. M. H.; Nguyen, M.; Lu, T.-J.; Kianinia, M.; Mendelson, N.;\nSolntsev, A.; Bradac, C.; Englund, D. R.; Aharonovich, I. Integrated on Chip Platform\nwith Quantum Emitters in Layered Materials. Advanced Optical Materials 2019 ,7,\n1901132.\n(34) Lukishova, S. G. Room-Temperature Single-Photon Sources: State of the Art. OSA\nQuantum 2.0 Conference. 2020; p QW6B.3.\n(35) Henck, H.; Pierucci, D.; Fugallo, G.; Avila, J.; Cassabois, G.; Dappe, Y. J.; Silly, M. G.;\nChen, C.; Gil, B.; Gatti, M.; Sottile, F.; Sirotti, F.; Asensio, M. C.; Ouerghi, A. Direct\nobservation of the band structure in bulk hexagonal boron nitride. Phys. Rev. B 2017 ,\n95, 085410.\n(36) Pierucci, D.; Zribi, J.; Henck, H.; Chaste, J.; Silly, M. G.; Bertran, F.; Le Fevre, P.;\nGil, B.; Summer\feld, A.; Beton, P. H.; Novikov, S. V.; Cassabois, G.; Rault, J. E.;\nOuerghi, A. Van der Waals epitaxy of two-dimensional single-layer h-BN on graphite\nby molecular beam epitaxy: Electronic properties and band structure. Applied Physics\nLetters 2018 ,112, 253102.\n(37) Arenal, R.; St\u0013 ephan, O.; Kociak, M.; Taverna, D.; Loiseau, A.; Colliex, C. Electron\nEnergy Loss Spectroscopy Measurement of the Optical Gaps on Individual Boron\nNitride Single-Walled and Multiwalled Nanotubes. Phys. Rev. Lett. 2005 ,95, 127601.\nPage 27Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\n(38) Liu, Z.; Tizei, L. H. G.; Sato, Y.; Lin, Y.-C.; Yeh, C.-H.; Chiu, P.-W.; Terauchi, M.;\nIijima, S.; Suenaga, K. Postsynthesis of h-BN/Graphene Heterostructures Inside a\nSTEM. Small 2015 ,12, 252{259.\n(39) Feenstra, R. M. Scanning tunneling spectroscopy. Surface Science 1994 ,299-300 ,\n965{979.\n(40) Feenstra, R. M. Tunneling spectroscopy of the (110) surface of direct-gap III-V semi-\nconductors. Phys. Rev. B 1994 ,50, 4561{4570.\n(41) Wong, D.; Velasco, J.; Ju, L.; Lee, J.; Kahn, S.; Tsai, H.-Z.; Germany, C.;\nTaniguchi, T.; Watanabe, K.; Zettl, A.; Wang, F.; Crommie, M. F. Characterization\nand manipulation of individual defects in insulating hexagonal boron nitride using\nscanning tunnelling microscopy. Nature Nanotechnology 2015 ,10, 949{953.\n(42) Jin, C.; Lin, F.; Suenaga, K.; Iijima, S. Fabrication of a Freestanding Boron Nitride\nSingle Layer and Its Defect Assignments. Phys. Rev. Lett. 2009 ,102, 195505.\n(43) Schu\u0013 e, L.; Berini, B.; Betz, A. C.; Pla\u0018 cais, B.; Ducastelle, F.; Barjon, J.; Loiseau, A.\nDimensionality e\u000bects on the luminescence properties of hBN. Nanoscale 2016 ,8,\n6986{6993.\n(44) Hern\u0013 andez-M\u0013 \u0010nguez, A.; L ahnemann, J.; Nakhaie, S.; Lopes, J. M. J.; Santos, P. V.\nLuminescent Defects in a Few-Layer h-BN Film Grown by Molecular Beam Epitaxy.\nPhys. Rev. Applied 2018 ,10, 044031.\n(45) S\u0013 aenz, J. J.; Garc\u0013 \u0010a, R. Near \feld emission scanning tunneling microscopy. Applied\nPhysics Letters 1994 ,65, 3022{3024.\n(46) Koma, A. Van der Waals epitaxy for highly lattice-mismatched systems. Journal of\nCrystal Growth 1999 ,201-202 , 236{241.\nPage 28Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\n(47) Novoselov, K. S.; Mishchenko, A.; Carvalho, A.; Castro Neto, A. H. 2D materials and\nvan der Waals heterostructures. Science 2016 ,353.\n(48) Walsh, L. A.; Hinkle, C. L. van der Waals epitaxy: 2D materials and topological\ninsulators. Applied Materials Today 2017 ,9, 504{515.\n(49) Choi, J.-H.; Cui, P.; Chen, W.; Cho, J.-H.; Zhang, Z. Atomistic mechanisms of van der\nWaals epitaxy and property optimization of layered materials. WIREs Computational\nMolecular Science 2017 ,7, e1300.\n(50) Sediri, H.; Pierucci, D.; Hajlaoui, M.; Henck, H.; Patriarche, G.; Dappe, Y. J.;\nYuan, S.; Toury, B.; Belkhou, R.; Silly, M. G.; Sirotti, F.; Boutchich, M.; Ouerghi, A.\nAtomically Sharp Interface in an h-BN-epitaxial graphene van der Waals Heterostruc-\nture. Scienti\fc Reports 2015 ,5, 16465.\n(51) Summer\feld, A.; Kozikov, A.; Cheng, T. S.; Davies, A.; Cho, Y.-J.; Khlobystov, A. N.;\nMellor, C. J.; Foxon, C. T.; Watanabe, K.; Taniguchi, T.; Eaves, L.; Novoselov, K. S.;\nNovikov, S. V.; Beton, P. H. Moir\u0013 e-Modulated Conductance of Hexagonal Boron Ni-\ntride Tunnel Barriers. Nano Letters 2018 ,18, 4241{4246, PMID: 29913062.\n(52) Vuong, T. Q. P.; Cassabois, G.; Valvin, P.; Rousseau, E.; Summer\feld, A.; Mel-\nlor, C. J.; Cho, Y.; Cheng, T. S.; Albar, J. D.; Eaves, L.; Foxon, C. T.; Beton, P. H.;\nNovikov, S. V.; Gil, B. Deep ultraviolet emission in hexagonal boron nitride grown by\nhigh-temperature molecular beam epitaxy. 2D Materials 2017 ,4, 021023.\n(53) Cho, Y.-J.; Summer\feld, A.; Davies, A.; Cheng, T. S.; Smith, E. F.; Mellor, C. J.;\nKhlobystov, A. N.; Foxon, C. T.; Eaves, L.; Beton, P. H.; Novikov, S. V. Hexagonal\nBoron Nitride Tunnel Barriers Grown on Graphite by High Temperature Molecular\nBeam Epitaxy. Scienti\fc Reports 2016 ,6, 34474.\n(54) Zhou, S. Y.; Gweon, G.-H.; Graf, J.; Fedorov, A. V.; Spataru, C. D.; Diehl, R. D.;\nPage 29Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nKopelevich, Y.; Lee, D.-H.; Louie, S. G.; Lanzara, A. First direct observation of Dirac\nfermions in graphite. Nature Physics 2006 ,2, 595{599.\n(55) Castellanos-Gomez, A.; Wojtaszek, M.; Arramel,; Tombros, N.; van Wees, B. J. Re-\nversible Hydrogenation and Bandgap Opening of Graphene and Graphite Surfaces\nProbed by Scanning Tunneling Spectroscopy. Small 2012 ,8, 1607{1613.\n(56) Pe~ na Rom\u0013 an, R. J.; Auad, Y.; Grasso, L.; Alvarez, F.; Barcelos, I. D.; Zagonel, L. F.\nTunneling-current-induced local excitonic luminescence in p-doped WSe2 monolayers.\nNanoscale 2020 ,12, 13460{13470.\n(57) do Amaral, G. M.; da Costa Tonon, I.; Pe~ na Rom\u0013 an, R. J.; de Oliveira Plath, H.;\nTaniguchi, T. M.; de Lima, L. H.; Zagonel, L. F.; Landers, R.; de Siervo, A. Epitaxial\ngrowth, electronic hybridization and stability under oxidation of monolayer MoS2 on\nAg(111). Applied Surface Science 2021 ,538, 148138.\n(58) Blase, X.; Rubio, A.; Louie, S. G.; Cohen, M. L. Quasiparticle band structure of bulk\nhexagonal boron nitride and related systems. Physical Review B 1995 ,51, 6868{6875.\n(59) Arnaud, B.; Leb\u0012 egue, S.; Rabiller, P.; Alouani, M. Huge Excitonic E\u000bects in Layered\nHexagonal Boron Nitride. Physical Review Letters 2006 ,96.\n(60) Wirtz, L.; Marini, A.; Rubio, A. Excitons in Boron Nitride Nanotubes: Dimensionality\nE\u000bects. Physical Review Letters 2006 ,96.\n(61) Paleari, F.; Galvani, T.; Amara, H.; Ducastelle, F.; Molina-S\u0013 anchez, A.; Wirtz, L.\nExcitons in few-layer hexagonal boron nitride: Davydov splitting and surface local-\nization. 2D Materials 2018 ,5, 045017.\n(62) Aryasetiawan, F.; Gunnarsson, O. TheGWmethod. Reports on Progress in Physics\n1998 ,61, 237{312.\nPage 30Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\n(63) Art\u0013 us, L.; Feneberg, M.; Attaccalite, C.; Edgar, J. H.; Li, J.; Goldhahn, R.; Cusc\u0013 o, R.\nEllipsometry Study of Hexagonal Boron Nitride Using Synchrotron Radiation: Trans-\nparency Window in the Far-UVC. Advanced Photonics Research 2021 ,2, 2000101.\n(64) Hunt, R. J.; Monserrat, B.; Z\u0013 olyomi, V.; Drummond, N. D. Di\u000busion quantum Monte\nCarlo andGW study of the electronic properties of monolayer and bulk hexagonal\nboron nitride. Phys. Rev. B 2020 ,101, 205115.\n(65) Yang, H.; Liang, A.; Chen, C.; Zhang, C.; Schroeter, N. B. M.; Chen, Y. Visualiz-\ning electronic structures of quantum materials by angle-resolved photoemission spec-\ntroscopy. Nature Reviews Materials 2018 ,3, 341{353.\n(66) Zhang, Y. et al. Direct observation of the transition from indirect to direct bandgap\nin atomically thin epitaxial MoSe 2. Nature Nanotechnology 2014 ,9, 111{115.\n(67) Zhang, C.; Johnson, A.; Hsu, C.-L.; Li, L.-J.; Shih, C.-K. Direct Imaging of Band\nPro\fle in Single Layer MoS2 on Graphite: Quasiparticle Energy Gap, Metallic Edge\nStates, and Edge Band Bending. Nano Letters 2014 ,14, 2443{2447, PMID: 24783945.\n(68) Ugeda, M. M.; Bradley, A. J.; Shi, S.-F.; da Jornada, F. H.; Zhang, Y.; Qiu, D. Y.;\nRuan, W.; Mo, S.-K.; Hussain, Z.; Shen, Z.-X.; Wang, F.; Louie, S. G.; Crommie, M. F.\nGiant bandgap renormalization and excitonic e\u000bects in a monolayer transition metal\ndichalcogenide semiconductor. Nature Materials 2014 ,13, 1091{1095.\n(69) Huang, Y. L.; Chen, Y.; Zhang, W.; Quek, S. Y.; Chen, C.-H.; Li, L.-J.; Hsu, W.-\nT.; Chang, W.-H.; Zheng, Y. J.; Chen, W.; Wee, A. T. S. Bandgap tunability at\nsingle-layer molybdenum disulphide grain boundaries. Nature Communications 2015 ,\n6, 6298.\n(70) Zhang, C.; Chen, Y.; Johnson, A.; Li, M.-Y.; Li, L.-J.; Mende, P. C.; Feenstra, R. M.;\nShih, C.-K. Probing Critical Point Energies of Transition Metal Dichalcogenides: Sur-\nPage 31Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\nprising Indirect Gap of Single Layer WSe2. Nano Letters 2015 ,15, 6494{6500, PMID:\n26389585.\n(71) Krane, N.; Lotze, C.; Franke, K. J. Moir\u0013 e structure of MoS2 on Au(111): Local\nstructural and electronic properties. Surface Science 2018 ,678, 136{142, Surface\nStructure and Dynamics { in Honor of Karl-Heinz Rieder.\n(72) Auw arter, W. Hexagonal boron nitride monolayers on metal supports: Versatile tem-\nplates for atoms, molecules and nanostructures. Surface Science Reports 2019 ,74,\n1{95.\n(73) Bruix, A.; Miwa, J. A.; Hauptmann, N.; Wegner, D.; Ulstrup, S.; Gr\u001cnborg, S. S.;\nSanders, C. E.; Dendzik, M.; Grubi\u0014 si\u0013 c \u0014Cabo, A.; Bianchi, M.; Lauritsen, J. V.; Khaje-\ntoorians, A. A.; Hammer, B.; Hofmann, P. Single-layer MoS 2on Au(111): Band gap\nrenormalization and substrate interaction. Phys. Rev. B 2016 ,93, 165422.\n(74) Dendzik, M.; Bruix, A.; Michiardi, M.; Ngankeu, A. S.; Bianchi, M.; Miwa, J. A.;\nHammer, B.; Hofmann, P.; Sanders, C. E. Substrate-induced semiconductor-to-metal\ntransition in monolayer WS 2.Phys. Rev. B 2017 ,96, 235440.\n(75) Zhang, Q.; Chen, Y.; Zhang, C.; Pan, C.-R.; Chou, M.-Y.; Zeng, C.; Shih, C.-K.\nBandgap renormalization and work function tuning in MoSe2/hBN/Ru(0001) het-\nerostructures. Nature Communications 2016 ,7, 13843.\n(76) Pan, M.; Liang, L.; Lin, W.; Kim, S. M.; Li, Q.; Kong, J.; Dresselhaus, M. S.;\nMeunier, V. Modi\fcation of the electronic properties of hexagonal boron-nitride in\nBN/graphene vertical heterostructures. 2D Materials 2016 ,3, 045002.\n(77) Qi, Y.; Han, N.; Li, Y.; Zhang, Z.; Zhou, X.; Deng, B.; Li, Q.; Liu, M.; Zhao, J.; Liu, Z.;\nZhang, Y. Strong Adlayer{Substrate Interactions \\Break\" the Patching Growth of h-\nBN onto Graphene on Re(0001). ACS Nano 2017 ,11, 1807{1815, PMID: 28110522.\nPage 32Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\n(78) Zhang, Z. et al. Direct Chemical Vapor Deposition Growth and Band-Gap Character-\nization of MoS2/h-BN van der Waals Heterostructures on Au Foils. ACS Nano 2017 ,\n11, 4328{4336, PMID: 28333441.\n(79) Brihuega, I.; Michaelis, C. H.; Zhang, J.; Bose, S.; Sessi, V.; Honolka, J.; Alexander\nSchneider, M.; Enders, A.; Kern, K. Electronic decoupling and templating of Co nan-\nocluster arrays on the boron nitride nanomesh. Surface Science 2008 ,602, L95{L99.\n(80) Herrera-Reinoza, N.; dos Santos, A. C.; de Lima, L. H.; Landers, R.; de Siervo, A.\nAtomically Precise Bottom-Up Synthesis of h-BNC: Graphene Doped with h-BN Nan-\noclusters. Chemistry of Materials 2021 ,33, 2871{2882.\n(81) Liu, M.; Li, Y.; Chen, P.; Sun, J.; Ma, D.; Li, Q.; Gao, T.; Gao, Y.; Cheng, Z.;\nQiu, X.; Fang, Y.; Zhang, Y.; Liu, Z. Quasi-Freestanding Monolayer Heterostructure\nof Graphene and Hexagonal Boron Nitride on Ir(111) with a Zigzag Boundary. Nano\nLetters 2014 ,14, 6342{6347, PMID: 25268563.\n(82) Li, Q.; Zou, X.; Liu, M.; Sun, J.; Gao, Y.; Qi, Y.; Zhou, X.; Yakobson, B. I.; Zhang, Y.;\nLiu, Z. Grain Boundary Structures and Electronic Properties of Hexagonal Boron\nNitride on Cu(111). Nano Letters 2015 ,15, 5804{5810.\n(83) Thomas, J.; Bradford, J.; Cheng, T. S.; Summer\feld, A.; Wrigley, J.; Mellor, C. J.;\nKhlobystov, A. N.; Foxon, C. T.; Eaves, L.; Novikov, S. V.; Beton, P. H. Step-\row\ngrowth of graphene-boron nitride lateral heterostructures by molecular beam epitaxy.\n2D Materials 2020 ,7, 035014.\n(84) Chiu, M.-H.; Zhang, C.; Shiu, H.-W.; Chuu, C.-P.; Chen, C.-H.; Chang, C.-Y. S.;\nChen, C.-H.; Chou, M.-Y.; Shih, C.-K.; Li, L.-J. Determination of band alignment\nin the single-layer MoS 2 /WSe 2 heterojunction. Nature Communications 2015 ,6,\n7666.\nPage 33Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\n(85) Bradley, A. J.; M. Ugeda, M.; da Jornada, F. H.; Qiu, D. Y.; Ruan, W.; Zhang, Y.;\nWickenburg, S.; Riss, A.; Lu, J.; Mo, S.-K.; Hussain, Z.; Shen, Z.-X.; Louie, S. G.;\nCrommie, M. F. Probing the Role of Interlayer Coupling and Coulomb Interactions\non Electronic Structure in Few-Layer MoSe2 Nanostructures. Nano Letters 2015 ,15,\n2594{2599, PMID: 25775022.\n(86) Zribi, J.; Khalil, L.; Avila, J.; Chaste, J.; Henck, H.; Oehler, F.; Gil, B.; Liu, S.;\nEdgar, J. H.; Giorgetti, C.; Dappe, Y. J.; Lhuillier, E.; Cassabois, G.; Ouerghi, A.;\nPierucci, D. Structural and electronic transitions in few layers of isotopically pure\nhexagonal boron nitride. Physical Review B 2020 ,102.\n(87) Museur, L.; Brasse, G.; Pierret, A.; Maine, S.; Attal-Tretout, B.; Ducastelle, F.;\nLoiseau, A.; Barjon, J.; Watanabe, K.; Taniguchi, T.; Kanaev, A. Exciton optical\ntransitions in a hexagonal boron nitride single crystal. physica status solidi (RRL) {\nRapid Research Letters 2011 ,5, 214{216.\n(88) Uddin, M. R.; Li, J.; Lin, J. Y.; Jiang, H. X. Probing carbon impurities in hexagonal\nboron nitride epilayers. Applied Physics Letters 2017 ,110, 182107.\n(89) H user, F.; Olsen, T.; Thygesen, K. S. Quasiparticle GW calculations for solids,\nmolecules, and two-dimensional materials. Phys. Rev. B 2013 ,87, 235132.\n(90) Guo, C.; Xu, J.; Ping, Y. Substrate e\u000bect on excitonic shift and radiative lifetime of\ntwo-dimensional materials. Journal of Physics: Condensed Matter 2021 ,33, 234001.\n(91) Sajid, A.; Ford, M. J.; Reimers, J. R. Single-photon emitters in hexagonal boron\nnitride: a review of progress. Reports on Progress in Physics 2020 ,83, 044501.\n(92) Zhang, J.; Sun, R.; Ruan, D.; Zhang, M.; Li, Y.; Zhang, K.; Cheng, F.; Wang, Z.;\nWang, Z.-M. Point defects in two-dimensional hexagonal boron nitride: A perspective.\nJournal of Applied Physics 2020 ,128, 100902.\nPage 34Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\n(93) Edelberg, D. et al. Approaching the Intrinsic Limit in Transition Metal Diselenides\nvia Point Defect Control. Nano Letters 2019 ,19, 4371{4379, PMID: 31180688.\n(94) Addou, R.; Colombo, L.; Wallace, R. M. Surface Defects on Natural MoS2. ACS\nApplied Materials & Interfaces 2015 ,7, 11921{11929, PMID: 25980312.\n(95) Ziatdinov, M.; Dyck, O.; Li, X.; Sumpter, B. G.; Jesse, S.; Vasudevan, R. K.;\nKalinin, S. V. Building and exploring libraries of atomic defects in graphene: Scan-\nning transmission electron and scanning tunneling microscopy study. Science Advances\n2019 ,5.\n(96) Zhussupbekov, K.; Ansari, L.; McManus, J. B.; Zhussupbekova, A.; Shvets, I. V.;\nDuesberg, G. S.; Hurley, P. K.; Gity, F.; \u0013O Coile\u0013 ain, C.; McEvoy, N. Imaging and\nidenti\fcation of point defects in PtTe2. npj 2D Materials and Applications 2021 ,5,\n14.\n(97) Fujimoto, Y.; Saito, S. E\u000bects of strain on carbon donors and acceptors in hexagonal\nboron nitride monolayers. Phys. Rev. B 2016 ,93, 045402.\n(98) Fujimoto, Y. Formation and Physical Properties of h-BN Atomic Layers: A First-\nPrinciples Density-Functional Study. Advances in Materials Science and Engineering\n2017 ,2017, 1{6.\n(99) Haga, T.; Fujimoto, Y.; Saito, S. STM visualization of carbon impurities in sandwich\nstructures consisting of hexagonal boron nitride and graphene. Japanese Journal of\nApplied Physics 2019 ,58, SIIB03.\n(100) Haga, T.; Fujimoto, Y.; Saito, S. Electronic structure and scanning tunneling mi-\ncroscopy images of heterostructures consisting of graphene and carbon-doped hexag-\nonal boron nitride layers. Phys. Rev. B 2019 ,100, 125403.\nPage 35Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\n(101) Jamr\u0013 oz, A.; Majewski, J. A. Morphology, Ordering, Stability, and Electronic Struc-\nture of Carbon-Doped Hexagonal Boron Nitride. physica status solidi (b) 2019 ,256,\n1800554.\n(102) Galvani, T.; Paleari, F.; Miranda, H. P. C.; Molina-S\u0013 anchez, A.; Wirtz, L.; Latil, S.;\nAmara, H.; Ducastelle, F. m. c. Excitons in boron nitride single layer. Phys. Rev. B\n2016 ,94, 125303.\n(103) Berseneva, N.; Gulans, A.; Krasheninnikov, A. V.; Nieminen, R. M. Electronic struc-\nture of boron nitride sheets doped with carbon from \frst-principles calculations. Phys.\nRev. B 2013 ,87, 035404.\n(104) Attaccalite, C.; Bockstedte, M.; Marini, A.; Rubio, A.; Wirtz, L. Coupling of excitons\nand defect states in boron-nitride nanostructures. Phys. Rev. B 2011 ,83, 144115.\n(105) Katzir, A.; Suss, J. T.; Zunger, A.; Halperin, A. Point defects in hexagonal boron ni-\ntride. I. EPR, thermoluminescence, and thermally-stimulated-current measurements.\nPhys. Rev. B 1975 ,11, 2370{2377.\n(106) Sajid, A.; Thygesen, K. S. VNCB defect as source of single photon emission from\nhexagonal boron nitride. 2D Materials 2020 ,7, 031007.\n(107) Grosso, G.; Moon, H.; Lienhard, B.; Ali, S.; Efetov, D. K.; Furchi, M. M.; Jarillo-\nHerrero, P.; Ford, M. J.; Aharonovich, I.; Englund, D. Tunable and high-purity room\ntemperature single-photon emission from atomic defects in hexagonal boron nitride.\nNature Communications 2017 ,8, 705.\n(108) Mackoit-Sinkevi\u0014 cien_ e, M.; Maciaszek, M.; Van de Walle, C. G.; Alkauskas, A. Carbon\ndimer defect as a source of the 4.1 eV luminescence in hexagonal boron nitride. Applied\nPhysics Letters 2019 ,115, 212101.\nPage 36Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\n(109) Jara, C.; Rauch, T.; Botti, S.; Marques, M. A. L.; Norambuena, A.; Coto, R.;\nCastellanos- \u0013Aguila, J. E.; Maze, J. R.; Munoz, F. First-Principles Identi\fcation of\nSingle Photon Emitters Based on Carbon Clusters in Hexagonal Boron Nitride. The\nJournal of Physical Chemistry A 2021 ,125, 1325{1335, PMID: 33554602.\n(110) Zheng, Y. J.; Chen, Y.; Huang, Y. L.; Gogoi, P. K.; Li, M.-Y.; Li, L.-J.; Tre-\nvisanutto, P. E.; Wang, Q.; Pennycook, S. J.; Wee, A. T. S.; Quek, S. Y. Point\nDefects and Localized Excitons in 2D WSe2. ACS Nano 2019 ,13, 6050{6059, PMID:\n31074961.\n(111) Schuler, B. et al. How Substitutional Point Defects in Two-Dimensional WS2 Induce\nCharge Localization, Spin{Orbit Splitting, and Strain. ACS Nano 2019 ,13, 10520{\n10534, PMID: 31393700.\n(112) Barja, S. et al. Identifying substitutional oxygen as a proli\fc point defect in monolayer\ntransition metal dichalcogenides. Nature Communications 2019 ,10, 3382.\n(113) Schuler, B.; Cochrane, K. A.; Kastl, C.; Barnard, E. S.; Wong, E.; Borys, N. J.;\nSchwartzberg, A. M.; Ogletree, D. F.; de Abajo, F. J. G.; Weber-Bargioni, A. Elec-\ntrically driven photon emission from individual atomic defects in monolayer WS2.\nScience Advances 2020 ,6.\n(114) Z\u0013 olyomi, V.; Koltai, J.; K urti, J. Resonance Raman spectroscopy of graphite and\ngraphene. physica status solidi (b) 2011 ,248, 2435{2444.\n(115) Pelini, T.; Elias, C.; Page, R.; Xue, L.; Liu, S.; Li, J.; Edgar, J. H.; Dr\u0013 eau, A.;\nJacques, V.; Valvin, P.; Gil, B.; Cassabois, G. Shallow and deep levels in carbon-\ndoped hexagonal boron nitride crystals. Phys. Rev. Materials 2019 ,3, 094001.\n(116) Weston, L.; Wickramaratne, D.; Mackoit, M.; Alkauskas, A.; Van de Walle, C. G.\nNative point defects and impurities in hexagonal boron nitride. Phys. Rev. B 2018 ,\n97, 214104.\nPage 37Pena Roman et al 2021 2D Material 8 044001 doi:10.1088/2053-1583/ac0d9c\n(117) Linder alv, C.; Wieczorek, W.; Erhart, P. Vibrational signatures for the identi\fcation\nof single-photon emitters in hexagonal boron nitride. Phys. Rev. B 2021 ,103, 115421.\n(118) Shevitski, B.; Gilbert, S. M.; Chen, C. T.; Kastl, C.; Barnard, E. S.; Wong, E.;\nOgletree, D. F.; Watanabe, K.; Taniguchi, T.; Zettl, A.; Aloni, S. Blue-light-emitting\ncolor centers in high-quality hexagonal boron nitride. Phys. Rev. B 2019 ,100, 155419.\n(119) Berzina, B.; Korsaks, V.; Trinkler, L.; Sarakovskis, A.; Grube, J.; Bellucci, S. Defect-\ninduced blue luminescence of hexagonal boron nitride. Diamond and Related Materials\n2016 ,68, 131{137.\n(120) Cheng, T.; Summer\feld, A.; Mellor, C.; Khlobystov, A.; Eaves, L.; Foxon, C.; Be-\nton, P.; Novikov, S. High-Temperature Molecular Beam Epitaxy of Hexagonal Boron\nNitride with High Active Nitrogen Fluxes. Materials 2018 ,11, 1119.\n(121) Cheng, T. S.; Summer\feld, A.; Mellor, C. J.; Davies, A.; Khlobystov, A. N.; Eaves, L.;\nFoxon, C. T.; Beton, P. H.; Novikov, S. V. High-temperature molecular beam epitaxy\nof hexagonal boron nitride layers. Journal of Vacuum Science & Technology B 2018 ,\n36, 02D103.\n(122) Wrigley, J.; Bradford, J.; James, T.; Cheng, T. S.; Thomas, J.; Mellor, C. J.;\nKhlobystov, A. N.; Eaves, L.; Foxon, C. T.; Novikov, S. V.; Beton, P. H. Epitaxy\nof boron nitride monolayers for graphene-based lateral heterostructures. 2D Materials\n2021 ,8, 034001.\nPage 38" }, { "title": "1205.3600v1.Ab_initio_Low_Energy_Model_of_Transition_Metal_Oxide_Heterostructure_LaAlO3_SrTiO3.pdf", "content": "arXiv:1205.3600v1 [cond-mat.str-el] 16 May 2012Typeset with jpsj2.cls Full Paper\nAb-initio Low-Energy Model of Transition-Metal-Oxide Heterostruct ure\nLaAlO 3/SrTiO 3\nMotoaki Hirayama∗,1,3TakashiMiyake,2,3and Masatoshi Imada1,3\n1Department of Applied Physics, University of Tokyo, 7-3-1 H ongo, Bunkyo-ku, Tokyo 113-8656, Japan\n2Nanosystem Research Institute, AIST, Tsukuba 305-8568, Ja pan\n3CREST, JST, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-8656, Japan\nWe develop the multi-scale ab-initio scheme for correlated electrons (MACE) for transition-\nmetal-oxide heterostructures, and determine the paramete rs of the low-energy effective model.\nBy separating Ti t2gbands near the Fermi level from the global Kohn-Sham (KS) ban ds\nof LaAlO 3/SrTiO 3which are highly entangled with each other, we are able to cal culate the\nparameters of the low-energy effective model of the interfac e with the help of constrained\nrandom phase approximation (cRPA). The on-site energies of the Tit2gorbitals in the 1st-\nlayer is about 650 meV lower than those in the 2nd-layer. In th e 1st-layer, the transfer integral\nof the Ti t2gorbital is nearly the same as that of the bulk SrTiO 3, while the effective screened\nCoulomb interaction becomes about 10 percent larger than th at of the bulk SrTiO 3. The\ndifferences of the parameters from the bulk SrTiO 3reduce rapidly with increasing distance\nfrom the interface. Our present versatile method makes it po ssible to derive effective ab-initio\nlow-energy models and allows studying interfaces of strong ly correlated electron systems from\nfirst principles.\nKEYWORDS: first-principles calculation, effective Hamilto nian, downfolding, constrained RPA method,\ncorrelated-electron systems, heterostructure, interfac e, two-dimensional electron systems\n1. Introduction\nIn recent years, interfaces of strongly correlated\nelectrons have been under intense investigations. Es-\npecially, transition-metal-oxide heterostructure SrTiO 3\n(STO)/LaAlO 3(LAO) has received a lot of attention,\nbecause of its remarkable transport properties.1The in-\nterface of the SrTiO 3/LaAlO 3shows metallic conductiv-\nity, although the bulk materials of each transition-metal-\noxide, SrTiO 3and LaAlO 3, are band insulators. The\nTiO2-terminated ( n-type) interfaces show the metallic\nconductivity,whenthe thicknessofdepositedLAOlayers\nis thicker than four unit cells,2but the SrO-terminated\n(p-type)interfacesareinsulatingforanyLAOthickness.3\nThe origin of this conductivity at the n-type interface\nremains a debated issue. An intrinsic electronic effect,\nnamely the polar discontinuity, and an extrinsic atomic\neffect, atomic vacancies, were proposed to play key roles\nat high carrier concentrations. When the polar LAO lay-\nersaredepositedontheTiO 2-terminatedsubstrateSTO,\nan electric potential along the [001] direction diverges as\nthe LAO thickness increases. The instability of this elec-\ntric potential is suppressed, if 1 /2 doped electron or a\ncorresponding atomic charge vacancy per unit cell ex-\nist at the interface.3,4Besides such transport properties,\nthe existence of superconductivity and magnetic order\nare also reported at the SrTiO 3/LaAlO 3interfaces.5,6\nThese unique transport properties are expected to offer\na useful functionality in the possible applications.\nThe Local Density Approximation (LDA) calculations\nfor transition-metal-oxide heterostructures have already\nbeen performed by several groups. Park et al.inves-\ntigated LaAlO 3/SrTiO 3for three kinds of superlattice\n∗E-mail: hirayama@solis.t.u-tokyo.ac.jpstructures with n-type,p-type, and both types of in-\nterfaces.7By taking into account the relaxation of lat-\ntice, Ishibashi and Terakura investigated the influence of\nLaAlO 3thicknesson the carrierdensity at the interface.8\nThe LDA, however, often fails to capture correlation ef-\nfectsintransition-metal-oxides.Tostudynovelelectronic\nphases in transition-metal-oxide heterostructures, in ad-\ndition to the carrier doping from the electronic recon-\nstruction, we should treat the correlation effects beyond\nthe LDA, because the correlation effect is, on general\ngrounds, expected to be enhanced at interfaces because\nof the effective reduction of the spatial dimensionality.\nThe LDA+U is useful for the strongly correlated ma-\nterials, especially for insulators. Pentcheva and Pickett\ntried to explain the insulating behavior of the p-type\nLAO/STO interface by introducing the on-site U, since\nthe LDAgivesametallic state.9However,in the LDA+U\nmethod, there is no established ab-initio way to estimate\nU, althoughthe value of Ustronglyaffects the calculated\nresults of the low-energy physics. Furthermore, owing to\nthe experimental difficulties, Uhas to be cited from the\nexperimental values of the bulk, not of the interface, es-\npecially in the case where thick LAO layersare on a STO\nsubstrate. Furthermore, the single-particle theory often\ncollapses in strongly correlated materials, even with the\nsuitableU.\nTo overcome such serious problems, a combined\nmethod of the multi-scale ab-initio scheme for corre-\nlated electrons (MACE) is very useful.10With the con-\nstrained random phase approximation (cRPA), we can\ncalculate not only the transfer integral but also the\nscreened Coulomb interaction without relying on any ex-\nperimentalparameters.Correlationeffectscanbetreated\naccurately by solving the obtained ab-initio effective\n12 J. Phys. Soc. Jpn. Full Paper Author Name\nmodel using low-energysolverssuch asvariationalMonte\nCarlo,11dynamical mean field theory (DMFT)12,13\nand path-integral renormalization group (PIRG).14,15\nThe MACE has already been applied to a wide vari-\nety of materials; semiconductor,16transition-metal,17–19\ntransition-metal-oxide,19–23molecular organic conduc-\ntors24and Fe-based layered superconductor.25,26So far,\nthe MACE has been applied to the bulk material. Since\ninterfaces have opened a new avenue of research, it is\nhighly desired to develop a methodology which is able to\ntreatthe interfacefrom the samefootingofthe MACEto\nunderstand the electron correlation effect from the first\nprinciples.\nIn this paper, we apply the ab-initio downfolding\nmethod to transition-metal-oxide heterostructures, and\ndetermine an ab-initio low-energy effective Hamiltonian\nof the SrTiO 3/LaAlO 3. We derive the effective model\nfrom the conduction bands near the Fermi level origi-\nnatedmainlyfromthe Ti t2gorbital.The on-siteenergies\nof the Tit2gorbitals in the 1st-layer is about 650 meV\nlower than those in the 2nd-layer. In the 1st-layer, the\ntransfer integral is nearly the same as that of the bulk\nSrTiO 3, while the effective screened Coulomb interaction\nbecomes about 10 percent larger than that of the bulk\nSrTiO 3. The parameters of the 2nd-layer are similar to\nthose of the bulk SrTiO 3.\nIn Sec.2 we describe our method. Section 3 describes\nthe band structure and the derived effective model of the\nSrTiO 3/LaAlO 3. We also present an effective model of\nSrTiO 3forcomparison.Section 4 is devoted to summary.\n2. Methods\nTo construct the effective low-energy model of the\nheterostructure with well defined parameters, we treat\nthe screening by the high-energy parts working on the\nlow-energy parts starting from the band structure of\nLAO/STO, and renormalize the high-energy parts into\nthe low-energy parts. This downfolding procedure is in-\ntroduced by Aryasetiawan et al.27and Solovyev et al.17\nThe low-energy effective model obtained from the down-\nfolding procedure offers a starting point for studies on\nlow-energy physics.\nWe consider an extended Hubbard Hamiltonian de-\nscribing low-energy electronic properties of the interface\nof LAO/STO,\nH=/summationdisplay\nσ/summationdisplay\nij/summationdisplay\nnmtmn(Ri−Rj)aσ†\ninaσ\njm\n+1\n2/summationdisplay\nσρ/summationdisplay\nij/summationdisplay\nnm/braceleftbigg\nUmn(Ri−Rj)aσ†\ninaρ†\njmaρ\njmaσ\nin\n+Jmn(Ri−Rj)/parenleftbig\naσ†\ninaρ†\njmaρ\ninaσ\njm+aσ†\ninaρ†\ninaρ\njmaσ\njm/parenrightbig/bracerightbigg\n,\n(1)\nwhereaσ†\nin(aσ\nin) is a creation (annihilation) operator of\nan electron with spin σin then-th orbital, which is de-\nfinedbyamaximallylocalizedWannierfunctioncentered\nat theRi-th unit cell.28,29Especially, in this paper, we\nconstruct a model for all Ti t2gorbitals, not only at theinterface but also in the bulk STO region, in the super-\ncell of LAO/STO. With thick LAO and STO layers, this\n3-dimensional model corresponds to the semi-infinite 3-\ndimensional model for (vacuum)-(LaAlO 3)N-(SrTiO 3)∞\nwhere the LAO layers are deposited on the substrate\nbulk STO, because the inter-supercell screening effect is\nrather weaker than the intra-one. Therefore, the param-\neters in this calculation may be used not only for super-\nlattices but also for a semi-infinite interface LAO/STO.\nThe parameters of the Ti t2gorbitals away from the in-\nterface recover the bulk STO nature in both the super-\ncell and semi-infinite interface, because the polarizations\naway from the bulk STO region contribute little to the\nscreening to such t2gorbitals. In fact, in the present cal-\nculation, we show that the parameters in the 2nd STO\nlayer from the interface nearly converge to those of the\nbulk STO. With thick LAO layers, the parameters of\nthis 3-dimensional model are also similar to those of a 2-\ndimensional model for LAO/STO superlattice, where all\nof the inter-supercell screening effects are renormalized.\nWe derive the effective model in the following way:\nFirst, we calculate the whole band structure of the\ntransition-metal-oxide heterostructure in the framework\nof density functional theory (DFT). We then choose\nthe target low-energy band around the Fermi level and\nconstruct the maximally-localized Wannier functions28\nin the low-energy Hilbert space. The transfer integral,\nwhich definesthe one-bodypartofthe low-energymodel,\nisobtainedasthematrixelementoftheKohn-Sham(KS)\nHamiltonian HKS,\ntmn(R) =/angbracketleftφ0m|HKS|φRn/angbracketright, (2)\nwhereφRnis then-th Wannier function centered at the\ncellR.\nNext, we renormalize the effect from the high-energy\nspace (rspace) into the target low-energy space ( d\nspace), and evaluate effective electron interaction by us-\ning the cRPA. We divide the polarization\nP(r,r′;ω) =occ/summationdisplay\niunocc/summationdisplay\njψi(r)ψ∗\ni(r′)ψ∗\nj(r)ψj(r′)\n×/bracketleftBig1\nω−ǫj+ǫi+iδ−1\nω+ǫj−ǫi−iδ/bracketrightBig\n(3)\ninto the polarization Pdthat includes only the d-dtran-\nsitions and the rest of the polarization Pr. The screened\nCoulomb interaction Wis given by\nW= [1−vP]−1v\n= [1−WrPd]−1Wr,(4)\nwhere we define the partially screened Coulomb interac-\ntionWrthat does not include the screening arising from\nthe polarization from the d-dtransitions as\nWr= [1−vPr]−1v. (5)\nThe screened Coulomb matrix is defined by\nWr(R1n,R2n′,R3m,R4m′;ω)J. Phys. Soc. Jpn. Full Paper Author Name 3\n=/integraldisplay\nd3rd3r′φ†\nR1n(r)φR2n′(r)\n×Wr(r,r′;ω)φ†\nR3m(r′)φR4m′(r′).(6)\nThe effective Coulomb interaction Uand the exchange\ninteraction Jare given by\nUnm(R) =Wr(0n,0n,Rm,Rm;0),(7)\nJnm(R) =Wr(0n,0m,Rm,Rn;0).(8)\nIn LaAlO 3/SrTiO 3, the Tit2gstates are entangled\nwith the La 4 fstates. We disentangle the Ti t2gbands\nusing the recently developed disentangling technique.30\nIn the LDA, the La 4 flevel in LAO/STO is located\nnear the Fermi level. However, in the real material, the\nLa 4fbands are expected to be at higher energy, and, as\naconsequence,donotstronglyscreentheinteractionsbe-\ntween the electrons in the t2gbands. To calculate the pa-\nrametersofthe low-energymodel with ahigher accuracy,\na better way is to take into account the correlation effect\nof the La 4fbands by the GW approximation (GWA) as\na preconditioning before determining the effective model\nfor Ti 3d t2gorbitals. The detail of the GWA for the\ndisentangled 4 fbands is explained in Appendix.\nComputational conditions are as follows. We calcu-\nlate the band structures of the bulk STO and LAO,\nand the LAO/STO heterostructure based on the DFT-\nLDA.31,32The calculations are carried out with the pro-\ngrambasedonthe full-potentiallinearmuffin-tin orbitals\n(FP-LMTO) method.33Because the LAO are deposited\nto fit the substrate STO, the lattice constants of the de-\nposited LAO layers in the direction horizontal to the in-\nterfacearesameasthoseofthesubstrateSTO,whilethat\nin the direction perpendicular to the interface slightly\nchanges depending on the carrier density at the inter-\nface. Therefore, in this study, the structure is fixed as\ncubic, and the lattice constants for the bulk and the het-\nerostructureperpendiculartothe[001]stackingdirection\nare fixed to 3 .905˚A,which correspondsto the experimen-\ntal lattice constant of the bulk SrTiO 3. In the LDA cal-\nculations, 8 ×8×8k-point sampling is employed for the\nbulk, and 8 ×8×2k-point sampling is employed for the\nLAO/STO to represent electronic structures of the sys-\ntem. The muffin-tin (MT) radii are: RMT\nTi= 2.50 bohr,\nRMT\nSr= 2.10bohr,RMT\nAl= 1.90 bohr,RMT\nLa= 1.60 bohr,\nandRMT\nO= 1.6 bohr. The angular momentum cut off is\ntaken atl= 4 for all the sites. In the cRPA and the GW\ncalculations, 3 ×3×3k-point sampling is employed for\nthe bulk, and 3 ×3×1k-point sampling is employed for\nthe LAO/STO.\n3. Results\n3.1 band structure and density of state\nFirst, we show the band structure and the density of\nstates of the insulators, bulk SrTiO 3and LaAlO 3. The\nupper panels of Figs. 1 and 2 show the band structure\nand the density of states , respectively. At room temper-\nature, SrTiO 3has the cubic perovskite structure, and\nbecomes tetragonal below 105 K. SrTiO 3has a high di-\nelectric constant at low temperatures because of the na-\nture of quantum paraelectricity.34In this calculation, theFig. 2. (color online) Density of states of bulk STO and LAO,\nand LAO/STO obtained by LDA. Energy is measured from the\nFermi level.\nstructure is fixed as cubic, and the lattice parameters\nare fixed at 3 .905˚A of the bulk SrTiO 3. The lower three\nconduction bands are derived from the t2gorbital of Ti\nsites, where the octahedral crystal field of O2−partially\nbreaks the 5-fold symmetry of the 3 dorbitals into the\nlower orbitals of the t2gand the higher orbital eg. The\ncalculated band gap is 2 .1 eV and the band width of\nthet2gband is 2.8 eV. In the experiment, the band gap\nis 3.3 eV.35In the strongly correlated materials, such\nunderestimation of the band gap causes to overestimate\nthe screening effect from the high-energy bands to the\nlow-energy bands in the cRPA. In the LAO/STO, the\nunderestimation of the energy levels of La 4 fbands is a\nmajor problem which are entangled with Ti t2gbands in\nthe LDA as we will show. LaAlO 3has the rhombohedral\nperovskite structure at room temperature, and becomes\ncubicabove821K.TocomparewiththeLAOlayerinthe\nheterostructure, the structure and the lattice parameters\nare fixed at those of the cubic SrTiO 3in this calculation.\nThe lowernarrowconductionbandsofLAOat ∼4eVare\nderived from the 4 forbital of La sites. The La 5 dbands\nare hybridized with the La 4 fbands at the Γ-point. In\nthe experiment, the band gap between the La 4 fand the\nO 2pis 5.6 eV.36In this calculation, however, the band\ngap is 3.4 eV because the LDA underestimates the value\nof the band gap. This band-gap problem in the LDA is4 J. Phys. Soc. Jpn. Full Paper Author Name\nFig. 1. Electronic band structures obtained by the LDA. The z ero energy corresponds to the Fermi level. Upper left panel: Electronic\nband of STO. Upper right panel: Electronic band of LAO. Lower panels: Electronic bands of LAO1 .5STO3.5. The right one is the\nenlarged view around the Fermi level. Wave functions of the s tates indicated by arrows are displayed in Fig. 4.\nFig. 3. (color online) Schematic picture of the model of the\ntransition-metal-oxide heterostructure LAO1 .5STO3.5. The sys-\ntem is modeled by a periodically repeated supercell contain ing\ntwo interfaces. The both interfaces are LaO-TiO 2. Owing to neu-\ntrality of the electrons and atoms, these systems are formal ly\ncharged by −eelectron per supercell.\nimproved dramatically with the GWA. We will show a\nresult in the GWA later.\nNext,weshowthe LDAresultsofthetransition-metal-\noxide heterostructure LAO1 .5STO3.5. We refer to the\nTiO2layer at the interface as the 1st-layer and the TiO 2\nlayerinthe bulk regionofthe SrTiO 3asthe 2nd-layer.Inthis paper, we refer to the heterostructure -(LaAlO 3)1-\nLaO/TiO 2-(SrTiO 3)3- as LAO1 .5STO3.5 (see Fig. 3).\nThis heterostructure has two crystallographically equiv-\nalentn-type interface, and has about 1 /2 carrierelectron\nat eachn-type interface, because LAO1 .5+1STO3.50has\na positive charge in the ionic limit. In terms of the po-\nlar discontinuity, these models of the heterostructures\nare the cases where the instabilities of the potential di-\nvergence are completely suppressed.3,4The lower pan-\nels of Fig. 1 show the band structure of the n-type\nLAO1.5STO3.5 heterostructure. The energy bands are\nrather degenerate because there are two crystallograph-\nically equivalent n-type interfaces. The LAO1 .5STO3.5\nhas the Fermi surface around the Γ point. The density of\nstates ofthe LAO/STOcomes from nearly superimposed\nstates of the bulk LAO and STO (see Fig. 2).\nFigures 4 shows the isosurface contours of selected\nBloch functions of the conduction bands at the Γ-point.\nThe panel (a) of Fig. 4 is the isosurface contour of the\nlowest conduction band (see Fig. 1 (d)). This band is\noriginated mainly from the Ti dxyin the 1st-layer. The\npanel (b) is the isosurface contour of the Ti dyzband,\nwhich spreads in the direction horizontal to the inter-\nface. The panel (c) and (d) are the isosurface contours\nof the Tidxyband in the 2nd-layer and the La 4 fband,\nrespectively. The Ti t2gand La 4forbitals are spatially\nclose and, in the LDA level, energetically close, so thatJ. Phys. Soc. Jpn. Full Paper Author Name 5\nFig. 4. (color online)Isosurface contours of the Bloch func tions of\nLAO1.5STO3.5 conduction bands at Γ-point[000]. The structure\ncorresponds to Fig. 3, for example, the panel A represents th e\nTidxyat the interface. Value of the isosurface is set to ±0.05\nbohr−3/2. Each energy level is measured from the Fermi level\nand is depicted as arrows in Fig. 1.\nthe Tit2gand La 4forbitals are hybridized at the in-\nterface (see Fig. 4). The Bloch functions transfer to the\nbulk region with increasing those energy levels. There is\na positive crystal field from the polar perovskite LaAlO 3\nto the non-polar substrate SrTiO 3as compared with the\ncase with only Sr2+for the bulk SrTiO 3. This crystal\nfield vanishes in the bulk region of SrTiO 3, because the\nnegative field from the doped electrons compensates this\npositive field. From these reasons, the energy levels of\nthe orbitals in the 1st-layer are lower than that in the\n2nd-layer, and the gap between the valence and conduc-\ntion bands in LAO1 .5STO3.5 is smaller than that of the\nbulk STO. Such effect of compensation and confinement\nbecomes strong as the polar perovskite LaAlO 3becomes\nthicker. This tendency is also seen in Fig. 5 which shows\nthe partial density of states at the Ti sites of dorbitals\nand the O sites of porbitals in the TiO 2layer of the\nbulk STO and LAO/STO obtained by the LDA. Here-Fig. 5. (color online) Partial densities of states of Ti dand Opat\nTiO2layers of bulk STO and LAO/STO obtained by the LDA.\nEnergy is measured from the Fermi level.\nafter, for instance, a Wannier orbital of the xyorbital in\nthe 1st-layer is denoted by 1 xy. Pronounced peak shifts\nof DOS to the lower energy are found for the sites in the\n1st-layer.\nConsidering the experimental band gap of the bulk\nLaAlO 3, we note that the location of the energy of the\nLa 4fbands in the LaAlO 3/SrTiO 3is too low in the\nLDA. The La 4 fbands screen and hybridize with the Ti\nt2gbands weaker in the real material. To calculate the\nparameter of the low-energymodel with a high accuracy,\na better way is to take into account the correlation effect\nof La separately by the GWA. To calculate the self ener-\ngies of the La 4 fbands, we first construct the maximally\nlocalized Wannier function of La 4 ffrom a linear com-\nbination of the target low-energy KS-bands. We choose\nthe energy window from −1 eV to 3 eV for the Wannier\nfunctions. We find, however, that the screened Coulomb\ninteraction of the low-energy bands derived from the Ti\nt2gis not sensitive to the choice of the energy window if\nthe window exceeds a certain width but is not too wide.\nNext, we calculate the self-energy corrections of the La\n4fbands by the 1-shot GW scheme. Figure 6 shows the\nband structure after the 1-shot GW corrections for the6 J. Phys. Soc. Jpn. Full Paper Author Name\nFig. 6. (color online) Electric band structures with 1-shot GW\nself-energy for 4 fbands. Energy is measured from the Fermi\nlevel.\nFig. 7. (color online) Disentangled d-bands having strong t2g\ncharacter ((blue) dashed line) and diagonalized r-bands (solid\nline).\n4fbands.The energylevel of4 fbands israisedbyabout\n1.5 eV with GW. This value is consistent with the ex-\nperimental result ofthe LAO.36In the following sections,\nwe show this self-energy effect on the effective model pa-\nrameters.\n3.2 Wannier function and transfer integral\nWe calculate the transfer integral tof the Tit2gor-\nbital to determine the 1-body part of low-energy effec-\ntive Hamiltonian for the heterostructure. First, we con-\nstruct 6×2 maximally localized Wannier functions hav-\ning strong Ti t2gcharacters, from the linear combination\nof the target low-energy KS-bands. We choose −1-3 eV\nas the energy window for the Wannier functions. Figure\n7 shows the disentangled t2gbands and the rest bands,\nand Fig. 8 shows the isosurfaces of Wannier functions of\nthedxyanddyzorbitals.\nWe show in Table I the transfer integrals tof the bulk\nSTO calculated in the LDA. In the tables and this sub-\nsection,mandnspecify symmetries of the Ti t2gor-\nbitals. The values of on-site energies are listed in the\ncolumn for ( Rx,Ry,Rz) = (0,0,0). The values of the\non-site energies are the same for all the t2gorbitals be-\ncause of the cubic crystal symmetry. The major values\nof the nearest hopping between the same symmetries\nFig. 8. (color online) Isosurface of the maximally localize d Wan-\nnier functions ±0.05 a.u. for the dxyanddyzorbitals in the\n1st-layer.\n(Rx,Ry,Rz) = (1,0,0) are 299 meV. These values are\nconsistent with the values in the literature in nearest-\nneighbor tight-binding models of the transition-metal-\noxide interfaces ( t∼0.3 eV).37The nearest hopping for\nthe perpendicular directions of the orbitals symmetry is\n−39 meV, which is 13 percent of these of the main direc-\ntions. The next nearest hopping ( Rx,Ry,Rz) = (1,1,0)\nis 35 percent of the nearest neighbor hopping.\nNext, we show in Table II the transfer integrals tof\nthe LAO1.5STO3.5 calculated in the LDA and the GW,\nrespectively. The values of on-site energy in the 1st-layer\nt1m,1m(0,0,0) are about 650 meV lower than those in\nthe 2nd-layer t2m,2m(0,0,0), mainly because the dipole\nmoment of LaO1+-AlO1−\n2stabilizes the energy of the t2g\norbitals in the 1st-layer. The hybridization between the\nLa 4fand Tit2gorbitals also slightly stabilizes the en-\nergy of the t2gorbitals. These layer dependent potential\nlocalizes carriers at the interface. Actually, in the exper-\niment, the transition into the 2D superconducting state,\nnamely the Berezinskii-Kosterlitz-Thouless transition is\nseen at the interface of the LAO/STO.5The value of\nthe on-site energy of 1 xy,t1xy,1xy(0,0,0), is lower than\nthat of the other 1 t2gdue to the crystal field and the\nhybridization with the LAO layer. In the GWA, this dif-\nference of the on-site energy at the 1st-layer is 16 meV\nlarger than that in the LDA, because the hybridizations\nof the La 4forbitals with the Ti t2gorbitals in the 1st-\nlayer, especially with 1 yzand 1zx, become weaker. In\nthe 2nd-layer, the dipole moment of LaO1+-AlO1−\n2have\nlittle effect, and the values of on-site energy partially\nrecover the bulk STO nature. The hopping parameters\nare similar to those of the bulk STO. The major val-\nues of the nearest hopping between the same symmetries\nt1m,1m(1,0,0) are about 0 .3 eV. The nearest hopping of\nthe 1xyis smaller than that of the bulk STO. The near-\nest hopping of the 1 yzand 1zxare nearly the same as\nthose of the bulk STO. The main difference of the hop-\nping parametersfrom the bulk STOis seen in the nearest\nhopping for the perpendicular directions of the orbitals\nsymmetryt1yz,1yz(1,0,0). Such hoppings are −59 meV,\nwhich are about twice as large as those of the bulk STO.J. Phys. Soc. Jpn. Full Paper Author Name 7\nIn the LAO/STO, therefore, the carriers tend to spread\nin the direction horizontal to the interface compared to\nthe bulk STO. As with the on-site energy, the hopping\nparameters in the 2nd-layer t2m,2m(1,0,0) are nearly the\nsame as those of the bulk STO. In the GWA, the hy-\nbridization between the La 4 fandt2gorbitals in the\n1st-layer, especially 1 yzand 1zx, becomes weaker. As a\nresult, the splitting of on-site energies of t2gorbitals be-\ncomes larger, and the hopping parameter between 1 zx-\n2zxbecomes smaller.\n3.3 screened Coulomb interaction\nNext, we calculate the on-site screened Coulomb in-\nteractionUmn(0,0,0) and the on-site screened exchange\ninteraction Jmn(0,0,0) of the Ti t2gorbital to determine\nthe 2-body part of the low-energy effective Hamiltonian\nfor the heterostructure. The screened Coulomb interac-\ntion is computed with the matrix elements in the max-\nimally localized Wannier basis in the framework of the\ncRPA (see eq. (6)).\nIn the top of Table III, we show the effective on-site\nCoulomb interaction UandJfor the bulk STO. As with\nthe transfer integrals, mandndenote symmetries of the\nt2gorbitalsofTiinthetablesandthissubsection.Theef-\nfective on-site Coulomb interaction Ubetween the same\norbitals is 3 .76 eV, while the bare on-site Coulomb in-\nteraction is 14 .27 eV. The on-site Coulomb interaction is\nreduced about to 1 /4 by the polarizations without the d-\ndcontributions. The effective on-site screened exchange\ninteraction Jis 0.46 eV.\nWe show the effective on-site Coulomb interaction U\nandJof LAO/STO calculated with cRPA in the mid-\ndle and bottom of the Table III. Here, Ubetween the\nsame orbitals are 3 .4-3.6 eV from the LDA and 3 .7-4.0\neV from the GWA, while that of the bulk STO is 3 .76\neV. In both the LDA and GWA, U1xy,1xyis the largest\namong on-site screened Coulomb interactions in the 1st-\nlayer, andU1yz,1yzandU1zx,1zxare 4-5 percent smaller\nthanU1xy,1xy. In the GWA, UandJat the 1st-layer\nbecome larger than those of the bulk STO, while these\nare smaller in the LDA. This is because the La bands\nare raised away from the Fermi level by the self-energy\ncorrection,thepolarizationsbetweenLa4 fandTit2gor-\nbitals become weaker compared to the LDA bands, and,\nas a result, the screening effect becomes weaker. The on-\nsite screened Coulomb interaction Umnand the on-site\nscreened exchange interaction Jmnsatisfy the equation\nUmn=Umm−2Jmnin thebulk STO.Onthe otherhand,\nat the interface of the LAO/STO, these parameters do\nnot satisfy this equation because of the inversion sym-\nmetry breaking. In the 2nd-layer, UandJrecover the\nbulk STO nature. Although, similarly in the 1st-layer,\nU2xy,2xyis the largest on-site screened Coulomb inter-\naction in the 2nd-layer, the difference between U2xy,2xy\nand the others of t2gis only about 1 percent. Because the\nelectrons confined at the interface have low dimensional-\nity, the correlation becomes effectively stronger than the\nbulk STO, even if the transfer integrals at the interface\nare similar to those of the bulk STO.4. Summary\nIn this paper, we have determined the parameters of\nthe low-energy effective model of LaAlO 3/SrTiO 3by\nthe MACE. As with many interfaces, LaAlO 3/SrTiO 3\nhas a complex band structure where the bands of both\nthe interface and bulk regions are highly entangled with\neach other. By the disentangling scheme using the max-\nimally localized Wannier function,30we disentangle the\nlow-energy part having the strong characters of Ti t2g\norbital from the global KS-bands, and thus enable to\ncalculate the parameters in the effective Hamiltonian of\nLaAlO 3/SrTiO 3by the cRPA. The parameters in this\nstudy offer not only the superlattice model but also\nthe semi-infinite interface model (vacuum)-(LaAlO 3)N-\n(SrTiO 3)∞, because the screening effect from the inter-\nsupercell rapidly decreases with distance and thus the\nsuperlattice model corresponds to the semi-infinite in-\nterface model in the limit of thick LAO and STO. The\nparameters have anisotropies and a layer dependence.\nThe on-site energies in the 1st-layer are 650 meV higher\nthan those in the 2nd-layer, which causes localization of\nthe carriers at the interface. In the 1st-layer, while the\ntransfer integral of the t2gorbital is similar to that of\nthe bulk SrTiO 3, the screened Coulomb interaction Uof\nthet2gorbital becomes 10 percent larger than that of\nthe bulk SrTiO 3. In the bulk region of LaAlO 3/SrTiO 3,\nthe parameters of the t2gorbitals recover the values\nand the symmetry of the bulk SrTiO 3. The obtained\n3-dimensional parameters constitute the low-energy ef-\nfective models of LaAlO 3/SrTiO 3, either in semi-infinite\nstructures in one of the directions or in supercell struc-\ntures. The resultant low-energy effective model offers a\nfirm and quantitative basis when one wishes to solve the\neffective model by using accurate low-energy solvers in\nthe future.\nRecently, an ab-initio dimensional downfolding\nscheme, which downfolds a 3-dimensional model to\na lower-dimensional model in real space, has been\nformulated.39By applying the dimensional downfolding\nscheme to the 3-dimensionalmodel ofLAO/STO,we can\nalso obtain a 2-dimensional model of a single LAO/STO\nsupercell. In the dimensional downfolding, one is able\nto expect that the weakness of inter-layer (inter-chain)\ncouplings justify the RPA type perturbative treat-\nment.10In the present case, however, the interface layer\nis not particularly weakly coupled with the other layers.\nNevertheless, the layers far away from the interface\ndo not join in the low-energy excitations if the bulk is\ninsulating as in the present case. Therefore, the gapped\nexcitations in the bulk part can be safely downfolded\ninto the metallic low-energy excitation near the interface\nin the same spirit of the cRPA. This enables to derive\neffective low-energy models solely for the interface part.\nOf course, we may need to keep the bands of not one\nbut several layers near the interface in the low-energy\nmodels. This dimensional downfolding is a challenging\nfuture issue.8 J. Phys. Soc. Jpn. Full Paper Author Name\nTable I. Transfer integrals for the t2gorbitals of the Ti sites in the bulk STO, tmn(Rx,Ry,Rz), where mandndenote symmetries of\nt2gorbitals. Units are given in meV.\nSTO\nPPPPPP(m,n)R/bracketleftbig\n0,0,0/bracketrightbig /bracketleftbig\n1,0,0/bracketrightbig /bracketleftbig\n1,1,0/bracketrightbig\n(xy,xy) 2430 −299−109\n(xy,yz) 0 0 0\n(xy,zx) 0 0 0\n(yz,yz) 2430 −39 7\n(yz,zx) 0 0 9\n(zx,zx) 2430 −299 7\nTable II. Transfer integrals for t2gorbitals of Ti sites in LAO1 .5STO3.5 calculated in LDA and GWA level, tmn(Rx,Ry,Rz), where m\nandndenote layers and symmetries of t2gorbitals. Units are given in meV. The parameters calculated from the LDA band are listed\nas “LDA”, and the parameters calculated from the LDA band wit h the self-energy correction of the La 4 fare listed as “GW”.\nLAO/STO LDA GW\nPPPPPP(m,n)R/bracketleftbig\n0,0,0/bracketrightbig /bracketleftbig\n1,0,0/bracketrightbig /bracketleftbig\n1,1,0/bracketrightbig/bracketleftbig\n0,0,0/bracketrightbig /bracketleftbig\n1,0,0/bracketrightbig /bracketleftbig\n1,1,0/bracketrightbig\n(1xy,1xy)2606 −289−101 2607 −288−104\n(1xy,1yz) 0−7−1 0−4 2\n(1xy,1zx) 0 0 −1 0 0 2\n(1xy,2xy) −42 6 −9−42 6 −9\n(1xy,2yz) 0 10 6 0 9 6\n(1xy,2zx) 0 0 6 0 0 6\n(1yz,1yz)2617 −57 −22634 −59 −5\n(1yz,1zx) 0 0 −6 0 0 −9\n(1yz,2xy) 0 8 8 0 8 8\n(1yz,2yz)−296 3 −6−299 4 −5\n(1yz,2zx) 0 0 8 0 0 8\n(1zx,1zx)2617 −302 −22634 −298 −5\n(1zx,2xy) 0 0 8 0 0 8\n(1zx,2yz) 0 0 8 0 0 8\n(1zx,2zx)−296 −98 −6−299 −98 −5\n(2xy,2xy)3268 −298−109 3264 −298−109\n(2xy,2yz) 0−3 0 0−3 0\n(2xy,2zx) 0 0 0 0 0 0\n(2yz,2yz)3262 −39 6 3262 −39 6\n(2yz,2zx) 0 0 8 0 0 7\n(2zx,2zx)3262 −297 6 3262 −297 6\nAcknowledgment\nMH would like to thank Kazuma Nakamura, Takahiro\nMisawa, Youhei Yamaji, Hiroshi Shinaoka, and Ryota\nWatanabeforusefuladvicesandfruitfuldiscussions.This\nwork has been supported by Grant-in-Aid for Scientific\nResearch from MEXT Japan under the grant numbers\n22104010 and 22340090. This work has also been finan-\ncially supported by MEXT HPCI Strategic Programsfor\nInnovative Research (SPIRE) and Computational Mate-\nrials Science Initiative (CMSI).Appendix: GWA for disentangled band\nThe quasiparticle energies and wave functions are ob-\ntained by solving\n(T+Vext+VH)ψnk(r)+/integraldisplay\ndr′Σ(r,r′;Enk)ψnk(r′)\n=Enkψnk(r),\n(A·1)\nwhereTis the kinetic energy operator, and Vextis the\nexternal potential, VHis the Hartree potential. If we con-\nsider only the diagonal parts of the self-energy Σ, eq.\n(A·1) is reduced to\nEnk=ǫLDA\nnk−/angbracketleftnk|VLDA\nex|nk/angbracketright+/angbracketleftnk|Σ(Enk)|nk/angbracketright.(A·2)\nThe self-energy operator must be estimated at the quasi-\nparticle energy Enk. This is done by expanding the ma-J. Phys. Soc. Jpn. Full Paper Author Name 9\nTable III. Effective Coulomb interaction ( U)/exchange ( J) interactions between the two electrons for all the combina tions of Ti- t2g\norbitals in the bulk STO and LAO1 .5STO3.5, respectively (in eV). Especially, for the combinations b etween the orbitals in the same\nTi sites, the matrix elements represent the effective on-sit e Coulomb interaction. The band structures of LAO1 .5STO3.5 are calculated\nin the LDA and GWA.\nSTO U J\nxy yz zx xy yz zx\nxy 3.76 2.81 2.81 xy 0.46 0.46\nyz 2.81 3.76 2.81 yz0.46 0.46\nzx 2.81 2.81 3.76 zx0.46 0.46\nLAO/STO(LDA) U J\n1xy1yz1zx2xy2yz2zx 1xy1yz1zx2xy2yz2zx\n1xy 3.65 2.63 2.63 0.49 0.55 0.55 1 xy 0.46 0.46 0.00 0.00 0.00\n1yz 2.63 3.48 2.58 0.55 0.65 0.63 1 yz0.46 0.44 0.00 0.01 0.00\n1zx 2.63 2.58 3.48 0.55 0.63 0.65 1 zx0.46 0.44 0.00 0.00 0.01\n2xy 0.49 0.55 0.55 3.62 2.67 2.67 2 xy0.00 0.00 0.00 0.46 0.46\n2yz 0.55 0.65 0.63 2.67 3.59 2.65 2 yz0.00 0.01 0.00 0.46 0.46\n2zx 0.55 0.63 0.65 2.67 2.65 3.59 2 zx0.00 0.00 0.01 0.46 0.46\nLAO/STO(GW) U J\n1xy1yz1zx2xy2yz2zx 1xy1yz1zx2xy2yz2zx\n1xy 4.00 2.98 2.98 0.72 0.79 0.79 1 xy 0.46 0.46 0.00 0.00 0.00\n1yz 2.98 3.83 2.93 0.78 0.89 0.87 1 yz0.46 0.44 0.00 0.01 0.00\n1zx 2.98 2.93 3.83 0.78 0.87 0.89 1 zx0.46 0.44 0.00 0.00 0.01\n2xy 0.72 0.78 0.78 3.80 2.84 2.84 2 xy0.00 0.00 0.00 0.46 0.46\n2yz 0.79 0.89 0.87 2.84 3.76 2.83 2 yz0.00 0.01 0.00 0.46 0.46\n2zx 0.79 0.87 0.89 2.84 2.83 3.76 2 zx0.00 0.00 0.01 0.46 0.4610 J. Phys. Soc. Jpn. Full Paper Author Name\nFig. A·1. Real and imaginary parts of the matrix elements of the\nself-energy Σ for a 4 fWannier band in LAO1 .5STO3.5 at Γ-\npoints. Solid and dashed lines represent the real and imagin ary\nparts of the Σ, respectively.\ntrix elements ofthe self-energy operatorto the first order\nin the energy around ǫnk. Then the quasiparticle energy\nis obtained explicitly;\nEnk=ǫnk+Znk(ǫnk)(ǫLDA\nnk−ǫnk+∆Σnk(ǫnk)),(A·3)\nwhere the self-energy is given by\n∆Σnk= Σnk−VLDA\nex (A·4)\nandZnkis the renormalization factor\nZnk= (1−∂∆Σnk(ǫnk)\n∂ω)−1. (A·5)\nIfǫnkcoincides with ǫLDA\nnk, then eq. (A ·3) is simplified;\nEnk=ǫLDA\nnk+Znk(ǫLDA\nnk)∆Σnk(ǫLDA\nnk).(A·6)\nWe show the real and imaginary parts of the self-\nenergy for a 4 fWannier band calculated by the 1-shot\nGW approximation in Fig. A ·1. The real and imaginary\npartsofthe self-energystronglyoscillateabovethe Fermi\nlevelEF= 0 eV, especially at 2-3 eV near the La 4 f\nbands in the LDA level. This lack of smoothness of the\nself-energy Σ is obtained in the “1-shot” correction. The\nself-energy Σ should hopefully be calculated in a self-\nconsistent procedure of the Hedin’s set of coupled equa-\ntions.38In this calculation, we have calculated the self-\nenergy correction by one shot without iteration to save\nthe computational cost. Considering the unstable behav-\nioroftheself-energyneartheLa4 fbands,weexpandeq.\n(A·2) aroundEF= 0 eV and then quasiparticle energies\nare approximated in the first order as follows;\nEnk=Znk(0)(ǫLDA\nnk+∆Σnk(0)).(A·7)\nThe approximation eq. (A ·6) is equivalent to eq. (A ·7),\nwhich is justified if the linearity of self-energy Σ in\nthe low-energy region is eventually recovered in the\nself-consistent accurate estimate. The smoothness of Σ\nshould be eventually obtained after the self-consistent\ncalculation of the GWA. In this calculation, we average\nZnk(0) and ∆Σ nk(0) in terms of the band index and thek-points;\nEnk=/angbracketleftZnk(0)/angbracketrightnk(ǫLDA\nnk+/angbracketleft∆Σnk(0)/angbracketrightnk).(A·8)\nSince we are interested in the La 4f level measured\nfrom the Fermi level, the self-energy correction to the\nLa 4f level is corrected by subtracting the correction to\nthe Fermi level. The latter is approximately evaluated by\nthe GW self-energy for the bottom of the t2gband us-\ning eq.(A ·7). Figure 6 shows the band structure thus ob-\ntained after the 1-shot GW corrections for the 4 fbands.\n1) A. Ohtomo and H. Y. Hwang: Nature (London) 427(2004)\n423.\n2) S. Thiel, G. Hammer, A. Schmehl, C. W. Schneider, and J.\nMannhart: Science 313(2006) 1942.\n3) N. Nakagawa, H. Y. Hwang, and D. A. Muller: Nat. Mater. 5\n(2006) 204.\n4) M. Hirayama and M. Imada: J. Phys. Soc. Jpn. 79(2010)\n034704.\n5) N. Reyren, S. Thiel, A. D. Caviglia, L. Fitting Kourkoutis ,\nG. Hammerl, C. Richter, C. W. Schneider, T. Kopp, A.-S.\nR¨ uetschi, D.Jaccard, M.Gabay, D.A.Muller, J.-M.Triscon e,\nand J. Mannhart: Science 317(2007) 1196.\n6) A. Brinkman, M. Huijben, M. V. Zalk, J. Huijben, U. Zeitler ,\nJ. C. Maan, W. G. Van Der Wiel, G. Rijders, D. H. A. Blank,\nand H. Hilgenkamp: Nat. Mater. 6(2007) 493.\n7) M. S. Park, S. H. Rhim, and A. J. Freeman: Phys. Rev. B 74\n(2006) 205416.\n8) S. Ishibashi and K. Terakura: J. Phys. Soc. Jpn. 77(2008)\n104706.\n9) R. Pentcheva and W. E. Pickett: Phys. Rev. B 74(2006)\n035112.\n10) M.Imada, and T.Miyake: J.Phys.Soc.Jpn. 79(2010) 112001.\n11) D.Tahara and M.Imada: J.Phys.Soc.Jpn. 77(2008) 114701.\n12) W. Metzner and D. Vollhardt: Phys. Rev. Lett. 62 (1989) 32 4.\n13) A.Georges, G.Kotliar, W.Krauth, and M.J.Rosenberg: Re v.\nMod. Phys. 68(1996) 13.\n14) M.Imada, and T.Kashima: J.Phys.Soc.Jpn. 69(2000) 2723.\n15) T. Mizusaki and M. Imada: Phys. Rev. B 74(2006) 014421.\n16) K. Nakamura, Y. Yoshimoto, R. Arita, S. Tsuneyuki, and M.\nImada: Phys. Rev. B 77(2008) 195126.\n17) I.V.Solovyev, and M.Imada: Phys.Rev.B 71(2005) 045103.\n18) I. V. Solovyev: Phys. Rev. B 73(2006) 155117.\n19) T.Miyake, and F. Aryasetiawan: Phys. Rev. B 77(2008)\n085122.\n20) E. Pavarini, S. Biermann, A. Poteryaev, A. I. Lichtenste in,\nA. Georges, and O. K. Andersen: Phys. Rev. Lett. 92(2004)\n176403.\n21) Y. Imai, I. V. Solovyev, and M. Imada: Phys. Rev. Lett. 95\n(2005) 176405.\n22) Y. Imai and M. Imada: J. Phys. Soc. Jpn. 75(2006) 094713.\n23) Y.Otsuka and M.Imada: J.Phys.Soc.Jpn. 75(2006) 124707.\n24) K. Nakamura, Y. Yoshimoto, T. Kosugi, R. Arita, and M.\nImada: J. Phys. Soc. Jpn. 78(2009) 083710.\n25) K. Nakamura, R. Arita, and M. Imada: J. Phys. Soc. Jpn. 77\n(2008) 093711.\n26) T. Miyake, K. Nakamura, R. Arita, and M. Imada: J. Phys.\nSoc. Jpn. 79(2010) 044705.\n27) F. Aryasetiawan, M. Imada, A. Georges, G. Kotliar, S. Bie r-\nmann, and A. I. Lichtenstein: Phys. Rev. B 70(2004) 195104.\n28) I.Souza, N.Marzari,and D.Vanderbilt: ibid65(2001) 035109.\n29) N. Marzari and D. Vanderbilt: Phys. Rev. B 56(1997) 12847.\n30) T. Miyake, F. Aryasetiawan, and M. Imada: Phys. Rev. B 80\n(2009) 155134.\n31) P. Hohenberg and W. Kohn: Phys. Rev. 136(1964) B864.\n32) W. Kohn and L. S. Sham: Phys. Rev. 140(1965) A1133 .\n33) M. Methfessel, M. van Schilfgaarde, and R. A. Casali; in L ec-\nture Notes in Physics, edited by H. Dreysse (Springer-Verla g,\nBerlin, 2000), Vol. 535.J. Phys. Soc. Jpn. Full Paper Author Name 11\n34) K. A. Muller and H. Burkard: Phys. Rev. B 19(1979) 3593.\n35) K.van Benthem, C.Elsasser, and R.H.French: J.Appl.Phy s.\n90(2001) 6156.\n36) S.-G. Lim, S. Kriventsov, T. N. Jackson, J. H. Haeni, D. G.\nSchlom, A. M. Balbashov, R. Uecker, P. Reiche, J. L. Freeouf,and G. Lucovsky: J. Appl. Phys. 91(2002) 4500.\n37) S. Okamoto, and A. J. Millis: Phys. Rev. B 70(2004) 075101.\n38) L. Hedin: Phys. Rev. 139(1965) A796.\n39) K. Nakamura, Y. Yoshimoto, Y. Nohara, and M. Imada: J.\nPhys. Soc. Jpn. 79(2010) 123708." }, { "title": "1804.09389v2.Nonsymmorphic_cubic_Dirac_point_and_crossed_nodal_rings_across_the_ferroelectric_phase_transition_in_LiOsO__3_.pdf", "content": "arXiv:1804.09389v2 [cond-mat.mtrl-sci] 30 Apr 2018Nonsymmorphic cubic Dirac point and crossed nodal rings acr oss the ferroelectric\nphase transition in LiOsO 3\nWing Chi Yu∗,1XiaotingZhou∗,1,2,†Feng-ChuanChuang,3ShengyuanA. Yang,4,‡Hsin Lin,1,5,6,§and Arun Bansil7\n1Center for Advanced 2D Materials and Graphene Research Cent er,\nNational University of Singapore, Singapore 117546, Singa pore\n2Physics Division, National Center for Theoretical Science , Hsinchu 30013, Taiwan\n3Department of Physics, National Sun Yat-Sen University, Ka ohsiung 804, Taiwan\n4Research Laboratory for Quantum Materials, Singapore Univ ersity of Technology and Design, Singapore 487372, Singapo re\n5Department of Physics, National University of Singapore, S ingapore 117546, Singapore\n6Institute of Physics, Academia Sinica, Taipei 11529, Taiwa n\n7Department of Physics, Northeastern University, Boston, M assachusetts 02115, USA\nCrystalline symmetries can generate exotic band-crossing features, which can lead to unconven-\ntionalfermionic excitationswithinterestingphysicalpr operties. WeshowhowacubicDiracpoint—a\nfour-fold-degenerate band-crossing point with cubic disp ersion in a plane and a linear dispersion in\nthe third direction—can be stabilized through the presence of a nonsymmorphic glide mirror sym-\nmetry in the space group of the crystal. Notably, the cubic Di rac point in our case appears on a\nthreefold axis, even though it has been believed previously that such a point can only appear on\na sixfold axis. We show that a cubic Dirac point involving a th reefold axis can be realized close\nto the Fermi level in the non-ferroelectric phase of LiOsO 3. Upon lowering temperature, LiOsO 3\nhas been shown experimentally to undergo a structural phase transition from the non-ferroelectric\nphase to the ferroelectric phase with spontaneously broken inversion symmetry. Remarkably, we\nfind that the broken symmetry transforms the cubic Dirac poin t into three mutually-crossed nodal\nrings. There also exist several linear Dirac points in the lo w-energy band structure of LiOsO 3, each\nof which is transformed into a single nodal ring across the ph ase transition.\nI. INTRODUCTION\nIn the past decade, the study of topological phases\nof matter has become one of the most active areas of\nresearch in condensed matter physics [ 1–4]. Extending\nthe earlier work on gapped topological phases, such as\ntopological insulators [ 5–16] and topological supercon-\nductors [ 17–27], the focus has shifted recently to gap-\nless phases, especially the so-called topological semimet-\nals (TSMs) [ 28–33].\nTSMs arecharacterizedby protected band-crossingsin\nmomentum space, which can be zero-dimensional nodal\npoints or one-dimensional nodal lines. Around these\nband-crossings, electrons behave drastically differently\nfrom the conventional Schr¨ odinger fermions. For exam-\nple, the low-energy electrons in three-dimensional (3D)\nDirac [34–38] and Weyl semimetals [ 39–42] resemble the\nrelativistic Dirac and Weyl fermions with linear disper-\nsion, allowing investigation of related exotic phenomena,\nwhich have previously belonged to the domain of high-\nenergy physics, in a desktop materials setting. In fact,\nthe family of nodal-line semimetals [ 43–55] harbors an\neven richer topological structure, and supports nodal\nchains [56,57], crossing nodal lines [ 58–60] as well as\nHopf links [ 61–66], which have no counterparts in high-\n∗These authors contributed equally to this work.\n†Electronic address: physxtzhou@gmail.com\n‡Electronic address: shengyuan yang@sutd.edu.sg\n§Electronic address: nilnish@gmail.comenergy physics. In addition to these novel bulk fermionic\nexcitations, TSMs also possess exotic topological surface\nstates in the form of Fermi arcs in Dirac/Weyl semimet-\nals and drumhead surface states in nodal-loop semimet-\nals [27,32,39].\nNonsymmorphic crystalline symmetries, i.e. sym-\nmetries which involve a fractional lattice translation,\ncan generate TSMs with exotic type of band crossing\nfeatures. For example, a glide mirror or a two-fold\nscrew axis can give rise to an hourglass dispersion [ 67],\nhourglass loop/chain [ 54–57], and 2D spin-orbit Dirac\npoints [68,69], which are robust against spin-orbit cou-\npling (SOC) effects.\nHere, we show that the glide mirror symmetry, com-\nbined with symmorphic three-fold rotation, inversion,\nand time-reversalsymmetries, can generatea cubic Dirac\npoint. This cubic Diracpointis four-folddegeneratewith\nan associated band dispersion, which is cubic in a plane\nand linear in the third direction normal to the plane.\nNotably, as a result of the nonsymmorphic symmetry,\nthe cubic Dirac point here appears on a threefold axis,\neven though it has been believed previously that such\na point can only appear on a sixfold axis [ 70,71]. We\nfurther show that when the inversion symmetry is bro-\nken, this cubic Dirac point transforms into three crossed\nnodal rings.\nThrough first-principles electronic structure computa-\ntions, weshowthatLiOsO 3providesamaterialsplatform\nfor realizing the novel topological physics outlined in the\npreceding paragraph. LiOsO 3is the first example of a\nferroelectric metal [ 72]. In a recent experiment, Shi et\nal.report that LiOsO 3undergoes a ferroelectric phase2\ntransition at a critical temperature T∗≈140 K [73], at\nwhich the material transforms from a centrosymmetric\nR¯3cstructure to a non-centrosymmetric R3cstructure,\nwhile remaining metallic. LiOsO 3thus realizes the sce-\nnario proposed in Ref. [ 72]. Notably, a related material,\nHgPbO 3, has been suggested recently to host a ferroelec-\ntric Weyl semi-metal phase [ 74].\nOur analysis shows that the non-ferroelectric R¯3c\nstructure satisfies the symmetry requirements for hosting\na cubic Dirac point. Also, when the inversion symmetry\nbreaks at the ferroelectric transition, we find that the\ncubic Dirac point gives rise to three crossed nodal rings\nin the ferroelectric R3cstructure. The low-energy band\nstructure of LiOsO 3further shows the presence of several\nlinear Dirac nodes, and we show that these nodes turn\ninto nodal loops during the ferroelectricphase transition.\nOur study thus not only discovers a new mechanism for\ngeneratingacubicDiracpoint, butitalsooffersapromis-\ning new materials platform for exploring the interplay\nbetween structural transitions, ferroelectricity, and novel\ntopological fermions.\nII. COMPUTATIONAL METHODS\nOurab-initio calculations were performed by using\ntheall-electronfull-potentiallinearizedaugmentedplane-\nwave code WIEN2k [ 75]. Generalized gradient ap-\nproximation (GGA) of Perdew, Burke and Ernzerhof\n(PBE) [76] was employed for the exchange-correlation\npotential. Experimental lattice constants [ 73] were used\nin the calculations and the internal structure was op-\ntimized. The self-consistent iteration process was re-\npeated until the charge, energy and force converged to\nless than 0.0001e, 0.00001 Ry, and 1 mRy/a.u, respec-\ntively. SOC was included by using the second variational\nprocedure [ 77]. The muffin-tin radii were set to 1.47,\n1.84, and 1.51 a.u. for the Li, Os, and O atoms, respec-\ntively.Rmin\nMTKmax= 7 and ak-point mesh of 4000 in the\nfirst Brillouin zone were used. An effective tight-binding\nmodel was constructed via projection onto the Wannier\norbitals [ 78–82] for carrying out topological analysis of\nthe electronic spectrum.We used the Os dorbitals with-\nout performing the maximizing localization procedure.\nSurface states were obtained using the surface Green’s\nfunction technique [ 83] applied on a semi-infinite slab.\nIII. CRYSTAL STRUCTURE AND SYMMETRY\nThenon-ferroelectricphaseofLiOsO 3hasarhombohe-\ndral structure with space group R¯3c(No.167) [Fig. 1(a)].\nThe Os atom is located at the center between two Li\natomsandalsoatthesymmetricpositionoftwoOplanes,\nwhich preserves inversion symmetry P. Upon lowering\nthe temperature below T∗, the two Os atoms are dis-\nplaced along the [111] direction, resulting in the loss of\nX\nL\n\u0001T\n\u0000\nKM(111) surface\nQOsLi\nO\n(c)\n (b) (a)\nFIG.1: CrystalstructuresofLiOsO 3(a)Thenon-ferroelectric\nphase and (b) the ferroelectric phase. The insets are\nschematic drawings of the oxygen planes (red lines) and the\nlocation of the Os and Li atoms relative to these planes. (c)\nBrillouin zone (BZ) ofLiOsO 3and theprojected (111) surface\nBZ.\ninversion center [Fig. 1(b)] [73]. The crystal symmetry is\nthus lowered to R3c(No.161) in the ferroelectric phase.\nWe emphasize that the key physics underlying the cu-\nbic Dirac point and its transformations in LiOsO 3is con-\ntrolledbysymmetry, andnotby thedetails ofthe specific\nmaterial involved. The important symmetries are as fol-\nlows. TheR¯3cspace group of the non-ferroelectric phase\ncan be generated by three symmetry elements: the inver-\nsionP, the three-fold rotation along [111]-direction ( z-\naxis)c3z, and the glide mirror with respect to the ( ¯110)\nplane (xz-plane)/tildewiderMy={My|1\n21\n21\n2}, which is a nonsym-\nmorphic symmetry. In the ferroelectric phase, when Pis\nbroken, the generators of the remaining R3cspace group\nreduce toc3zand/tildewiderMyonly. Lattices in both phases are\nbipartite: there are two sublattices (denoted as A and B)\nwith a relative displacement of (1\n2,1\n2,1\n2).Pand/tildewiderMymap\nbetween the two sublattices ( A↔B), whereasC3zmaps\nwithin each sublattice ( A(B)↔A(B)). The system pre-\nserves the time-reversal symmetry T, since no magnetic\nordering has been observed in either phase [ 73].\nIV. BAND STRUCTURES WITHOUT SOC\nWithout the SOC, the band structures of non-\nferroelectric and ferroelectric phases of LiOsO 3are sim-\nilar as seen in Fig. 2. Both phases exhibit two-fold de-\ngenerate (four-fold if spin-degeneracy is counted) nodal\nlines passing through the L and T points of the BZ per-\npendicular to the glide mirror planes. These features of\nthe band structure can be understood by noting that the\nL and T are time-reversal invariant momentum (TRIM)\npoints, which also reside on the glide mirror planes /tildewiderMi\n(i= 1,2,3)(the othertwoglidemirrorsarerelatedto /tildewiderMy\nbyC3z), see Fig. 1(c). It is easily seen that the line pass-\ning through an L point and normal to the corresponding\nmirror plane /tildewiderMiis invariant under the anti-unitary sym-3\n(b)\n(f)(a)\n(c)\n(d)\n(e)\nFIG. 2: Electronic band structure of LiOsO 3in the absence\nof SOC for (a) the non-ferroelectric phase and, (b) the ferro -\nelectric phase. (c)-(f) Zoom-in images of the band crossing s\nat the L and T symmetry points marked by red circles in (a)\nand (b).\nmetry operation T/tildewiderMi, so that\n(T/tildewiderMi)2=e−ikL\nz=−1, (1)\nwhere we have used T2= +1 for a spinless system and\nkL\nzis thekz-component of L. The T/tildewiderMisymmetry thus\nguarantees a two-fold Kramers-like degeneracy on this\nline. A similar analysis applies to the T point: Since\nT lies on the intersection of three glide mirrors, three\nsuch nodal lines pass through T. Moreover, since /tildewiderMi+1=\nC3z/tildewiderMiC−1\n3z, [C3z,/tildewiderMy]/ne}ationslash= 0,whichleadsto[ /tildewiderMi,/tildewiderMj/negationslash=i]/ne}ationslash= 0.\nFor an eigenstate |ψ0/an}bracketri}htat T, the three states |ψi=1,2,3/an}bracketri}ht=\nT/tildewiderMi|ψ0/an}bracketri}htmust be orthogonal to each other, indicating\na four-fold (eight-fold if counting spin) degeneracy at T.\nThis analysis is in accord with our DFT band structure\nresults.V. BAND STRUCTURES WITH SOC\nA. Non-ferroelectric phase: linear and cubic Dirac\npoints\nThe band structure including SOC is of greater inter-\nest because the low-energy states mainly arise from the\nOs-5dorbitalswith strongSOCeffects. When theSOCis\nturnedon, theband-crossingsinthenon-ferroelectric R¯3c\nphase evolve from 1D nodal lines into 0D Dirac points,\neachwithfour-folddegeneracy. Fig. 3(a)showsthatclose\nto Fermi level, now we have a linear Dirac node at each\nL point (there are three inequivalent L points in the BZ)\n[Fig.3(c) and Fig. 4(a,b)]. The original eight-fold degen-\neracy at T splits under SOC into two Dirac points, where\none is linear [Fig. 3(d)] and the other is cubic [Fig. 3(e)\nand Fig. 4(d,e)]. Hence, we may call this phase as a\nmulti-type Dirac semimetal.\nThe cubic Dirac point, around which the dispersion\nis linear along one direction ( kz-axis) and cubic along\nthe other two directions, has been rarely reported in real\nmaterials[ 70]. Inpreviouswork,itwasbelievedthatsuch\na Dirac point can only appear on a sixfold axis [ 70,71]\nand is possible only for two space groups (No. 176 and\nNo. 192) [ 70]. Our results clearly demonstrate that it\ncan also occur on a threefold axis in the presence of the\nadditional nonsymmorphic (glide mirror) symmetry. In\nthe following, we shall show that the cubic Dirac point\nis indeed protected by the glide mirror together with the\nsymmorphic crystal symmetries.\nThe following points may be noted in connection with\nsymmetry considerations. (1) In the non-ferroelectric\nphase, the presence of both TandPsymmetries forces\nthe two-fold spin-degeneracy of each band. Hence the\ncrossingbetween the bandsmust be at least 4-folddegen-\nerate (i.e. Dirac type). (2) All states at T and L symme-\ntry points must be degenerate quadruplets. This reason\nis that Kramers degeneracy at TRIM points requires the\neigenvalues of Pto be paired as (1 ,1) or (−1,−1). In\ncontrast, at T and L,\n[P,/tildewiderMy]/ne}ationslash= 0, (2)\nimplying that the degeneratestates generatedby /tildewiderMywill\nhavePeigenvalues paired as (1 ,−1), which guarantees\nthe four-fold degeneracy at T and L points. (3) The na-\nture of the Dirac point (linear or cubic) strongly relies on\nthe additional rotational symmetry C3z, which is present\nat T but not at L, as we will explicitly demonstrate be-\nlow.\nAs we already pointed out, the low-energy bands are\nmainly derived from the Os-5 dorbitals. Under trigonal\nprismatic coordination, the Os d-orbitals split into two\ngroups:A1g(dz2) andEg{(dx2−y2,dxy),(dxz,dyz)}. Un-\nder SOC, these states can be combined into the follow-\ning spin-orbit-coupled symmetry-adapted basis, keeping\nin mind that each unit cell contains two Os atoms, one4\n0.050.10.150.550.60.65-0.075-0.07-0.0650.060.070.080.09-0.12-0.1-0.08\n0.540.560.58\n(b)(f)\n(g)\n(h)\n( \u0002 \u0003\n\u0004 \u0005 \u0006\n(d)\n(e)\nFIG. 3: Electronic band structure for (a) the non-ferroelec tric\nphase and (b) the ferroelectric phase where SOC is included.\n(c)-(h) Zoom-in images of the band crossings at the L and T\npoints marked by red circles in (a) and (b).\nin each sublattice:\nφτ\n0,s=|dτ\nz2,s/an}bracketri}ht, (3)\nφτ\n±,s=1√\n2(sinλ|dτ\nxz,s±isdτ\nyz,s/an}bracketri}ht+cosλ|dτ\nx2−y2,s∓is2dτ\nxy,s/an}bracketri}ht),\n(4)\nwhereτlabels the two sublattices A and B, slabels the\nspin, andλis a normalization coefficient.\nAt the T point, Kramers degeneracy requires that the\nfour basis functions φτ\ni,swith a fixed i= 0,±are cou-\npled together into a quadruplet. Consider first the two\nquadruplet basis Ψ 0/−= (φA\n0/−,↑,φA\n0/+,↓,φB\n0/−,↑,φB\n0/+,↓),\nfor which the symmetry operations at T take the follow-\ning representations:\nT=τ3⊗s2K ,P=−τ2,\n/tildewiderMy=e−iqz/2τ1⊗s2, C3z=τ0⊗e∓iπ\n3s3.(5)\nHereτandsare Pauli matrices acting on the sublattice\nand spin spaces, respectively, and qis the wave-vector\nmeasured from T. These expressions fix the k·pHamil-\ntonian at T expressed using Γ matrices, which we define\nhere asγ1=τ3s1,γ2=τ3s2,γ3=τ1s0,γ4=τ2s0, and\nγ5=τ3s3. Then, to the lowest order, we find that\nHlinear\nT(q) =α1(qxγ1±qyγ2)+α2qzγ3+α3qzγ5,(6)\n(g)(a)\u0007 \b \t\n\n \u000b \f\n(d) (e)\r \u000e \u000f\nlinear + cubic Dirac\nlinear Dirac\nT\nLT\nL(h)\nFIG. 4: Top and middle panels show the dispersions around\nthe L and T points. (a), (b), (d) and (e) are for the non-\nferroelectric phase, while (c) and (f) are for the ferroelec tric\nphase. (g, h) Schematic diagrams showing the band-crossing s\nin the non-ferroelectric (g) and ferroelectric (h) phases.\nwhereα’s arethe expansioncoefficients, andthe sign ±is\nfor the two basis Ψ 0/−. This model describes the linear\nDirac points at T, which can be viewed as consisting\nof two Weyl points with Chern numbers 1 and −1. As\nexpected, the total Chern number for a closed surface\nsurrounding the Dirac points vanishes.\nAs for the other quadruplet basis, Ψ +=\n(φA\n+,↑,φA\n−,↓,φB\n+,↑,φB\n−,↓), the representations of the\nsymmetry operations are the same as in Eq. ( 5) except\nthatC3z=−τ0⊗s0. Due to this different transformation\nbehavior of the basis under C3z, the related effective\nHamiltonian is different from ( 6), where the diagonal\nterms proportional to τ0s0have been dropped:\nHcubic\nT(q) =/parenleftbiggh11h12\nh†\n12−h11/parenrightbigg\n, (7)\nwhere\nh11= [c1(q3\n++q3\n−)+b1(q+q−qz)+a1qz]s1\n+ic2(q3\n+−q3\n−)s2\n+[c3(q3\n++q3\n−)+b2(q+q−qz)+a2qz]s3,\nh12= [c4(q3\n++q3\n−)+b3(q+q−qz)+a3qz]s0.\nHereq±=qx±iqy,andai,bi,andciaretheexpansionco-\nefficients. We find that the band-crossing at T described\nby Eq. (7) to be a cubic Dirac point, with cubic disper-\nsion in theqx-qyplane. The diagonal blocks describe two\ntriple-Weyl fermions with Chern numbers ±3. The cubic\nDirac point can be viewed as being composed of the two\ntriple-Weyl points.5\n(b)\n(f)(a)\n(c)\n(d)\n(e)\nFIG. 5: Accidental band crossing points along the Γ-T sym-\nmetry line. (a) Band structure with SOC for the non-\nferroelectric phase. (b) Band structure with SOC for the fer -\nroelectric phase. (c,d) Zoom-in images of the band-crossin g\npoints marked by green circles in (a). (e,f) Zoom-in images o f\nthe band-crossing points in (b). Each band in (c,d) is two-fo ld\ndegenerate, so that the crossing points are four-fold degen er-\nate. The crossing points in (e,f) lie between a non-degenera te\nanda two-fold degenerate band, so thatthese points are trip ly\ndegenerate.\nWe emphasize that when onlysymmorphic symmetries\nare considered, cubic Dirac points require the presence\nof a six-fold axis. [ 70,71] Our case, however, involves a\nnonsymmorphicglidemirror,whichplaysacrucialrolein\nrealizing the cubic Dirac point on a three-fold axis. Our\nanalysis indicates that a sufficient condition for realizing\na cubic Diracpoint is the presenceofP, T, aglide mirror,\nand a c3-axis within the mirror. P, T, and the glide\nmirror only protect a four-fold degeneracy; the presence\nof an additional three-fold rotation then yields the cubic\nin-plane dispersion.\nA similar analysis for the L point at (0 ,0,1\n2) leads to\nthe following effective Hamiltonian:\nHL(q) =/summationdisplay\ni=1,3,5(βx,iqx+βz,iqz)γi+βyqyγ2.(8)\nHereqis measured from L, and the β’s are the expansion\ncoefficients. This demonstrates that the crossings at L\nare linear Dirac points.\nThe linear and cubic Dirac points discussed above are\nessential in the sense that their appearance at the high-symmetryTRIMpointsTandLismandatedbythesym-\nmetry operations of the system. In addition, we note the\npresenceofaccidentalDiracpoints, whichappearinpairs\nalongtheprimaryrotationaxisΓ-T.Therearethreesuch\npairs around the Fermi level, one of these is seen to be\nof type-II with an over-tilted dispersion. Zoom-in images\nof these Dirac points are shown in Figs. 5(a,c,d).\nB. Ferroelectric phase: crossed nodal rings\nIn going across the ferroelectric R3cphase transition,\nthe loss of inversion symmetry induces profound changes\nin the band structure, see Fig. 3(b). The spin-degeneracy\nof the bands gets lifted, and the Dirac points in the non-\nferroelectric phase become unstable. We find that each\nlinear Dirac point at L develops into a nodal ring on the\nglide mirror plane, which encloses the L point [Figs. 3(f)\nand4(c)]. Also, the cubicDiracpoint atTistransformed\ninto three mutually-crossed nodal rings, each lying on a\nglide mirror plane [Figs. 3(h) and4(f)]. This topological\nphase maythus be called a crossed-nodal-ringsemimetal.\nThe transformations of these band-crossings across the\nphase transition are illustrated in Fig. 4(g,h).\nThe occurrence of the preceding nodal rings in the\nband structure is also essential in that it is solely de-\ntermined by symmetry considerations. Note that the\nglide eigenvalues are ±iat Γ1∈ {Γ,X}, and±1 at\nΓ2∈ {T,L}. The presence of There requires that the\nKramers pairs at the above TRIM points carry complex-\nconjugated eigenvalues, i.e.(i,−i) at Γ 1; and (1,1) or\n(−1,−1)at Γ 2. Hence, alonganypathconnecting Γ 1and\nΓ2in the glide plane, the evolution of the glide eigen-\nvalues drives a switching of Kramers partners. During\nthis switching, two bands with opposite glide eigenvalues\nmust produce a crossing. As this argument holds for any\nin-plane path, a nodal loop separating Γ 1and Γ 2must\nappear. This symmetry analysis highlights the impor-\ntance of the nonsymmorphic glide mirror in producing\nthese essential band-crossings.\nThe transformations described above can also be cap-\ntured in the effective models. For example, at the T\npoint, starting from the model of Eq. 7for the cubic\nDirac point, the leading order perturbation δτ1s2when\nthePis broken, so that a minimal model may be ex-\npressed as\nHCNR\nT(q) =Hcubic\nT(q)+δτ1s2. (9)\nIt is straightforward to verify that this model gives three\nmutually-crossed nodal rings like the DFT calculations.\nSimilarly, the nodal ring around the L point can be de-\nrived from the model of Eq. 8as follows\nHNR\nL(q) =HL(q)+δ′τ1s2. (10)\nNote that the linear Dirac points at T are still pre-\nserved [Figs. 3(g)], which can be attributed to the non-\ncommutativity of C3zand/tildewiderMyin the correspondingbasis.6\nThe three pairs of accidental Dirac points residing on the\nprimary rotational axis are, however, turned into triply-\ndegenerate nodes as shown in Fig. 5(b,e,f).\nUnlike Weyl points, the essential Dirac points may or\nmay not give rise to nontrivial surface states [ 71]. In\nFig.6, we plot the surface spectrum of the (111)-surface\nfor both the non-ferroelectric and ferroelectric phases.\nBothphasesareseentosupportsurfacebandsconnecting\nthe two surface-projected L points.\n(b)\n(d)\n(a) \n(c)\nFIG. 6: Surface states of LiOsO 3in the non-ferroelectric (up-\nper panel) and ferroelectric (lower panel) phases in the pro -\njected spectra of the (111) surface. SOC is included. (a) and\n(c) Constant energy surfaces at the band-crossing energy at\nL. (d) and (e) Dispersions along the kx-direction at ky= 0.5.\nVI. CONCLUSION\nWe emphasize a number of points in closing. Firstly,\nthe Dirac nodes at the TRIM points and the multiple\nnodal rings surrounding these points discussed here are\nessential band-crossings in the sense that the presence of\nthese features is solely dictated by the nonsymmorphic\nspace group symmetries (plus T) of the system. Their\nexistence is thus guaranteed as long as these symmetries\nare preserved. Our analysis of symmetry considerations\nunderlying cubic Dirac points and crossed nodal rings\nwill provide a useful guide in searching for exotic band\ncrossings in realistic material systems more generally.\nSecondly, our analysis clearly indicates that LiOsO 3\nwill provide a useful platform for exploring exotic topo-\nlogical phases and their interplay with ferroelectric or-\ndering. The unique advantages of LiOsO 3are as follows.\n(i) The essential band-crossings are close to the Fermi\nlevel, so that the associated topological physics will be\nreflected in various electronic properties. (ii) The twotopological phases that coexist in the same material are\nconnected via a ferroelectric phase transition, which is\ntunable by varying temperature [ 73]. And, (iii) the ma-\nterial has already been realized experimentally and its\nferroelectric phase transition has been observed. The in-\nterestingtransformationsintheband-crossingsacrossthe\nphase transition, which we have predicted here, could be\ndetected by ARPES experiments.\nThirdly, we have identified the first case of an ex-\nperimentally realized material, which harbors a cubic\nDirac point in a 3D material. Such a Dirac point has\nbeen predicted previously only in quasi-1D molybdenum\nmonochalcogenide compounds [ 70], where an experimen-\ntal verification is still lacking. Our work offers a new\nrouteforrealizingcubicDiracpoints,extendingtherange\nof materials in search of cubic Dirac fermions.\nFinally, our study opens a new pathway for exploring\na variety of novel effects associated with the various non-\ntrivial band-crossings. For example, a linear Dirac point\nmayexhibitnegativemagnetoresistance[ 84,85], aspecial\nmagnetic oscillation frequency driven by surface Fermi\narcs [86], and an artificial gravity field via strain modu-\nlation[87]. AcubicDiracpointcanexhibitunusualquan-\ntuminterferencecontributionstomagneto-transport[ 88],\nstronger screening of interactions, and possible presence\nof continuous quantum phase transitions driven by inter-\nactions [89]. A nodal ring may yield strong anisotropy in\nelectricaltransport[ 47], unusualopticalresponse[ 90]and\ncircular dichroism [ 91], and possible surface magnetism\nand superconductivity [ 92].\nACKNOWLEDGMENTS\nThe workat National University ofSingapore wassup-\nported by the Singapore National Research Foundation\nunder the NRF fellowship Award No. NRF-NRFF2013-\n03. The work at Northeastern University was supported\nby the US Department of Energy (DOE), Office of Sci-\nence, Basic Energy Sciences grant number DE-FG02-\n07ER46352, and benefited from Northeastern Univer-\nsity’s Advanced Scientific Computation Center and the\nNational Energy Research Scientific Computing Center\nthroughDOE grantnumberDE-AC02-05CH11231. FCC\nand XZ acknowledge support from the National Cen-\nter for Theoretical Sciences. Work at Singapore Uni-\nversity of Technology and Design is supported by Sin-\ngapore MOE Academic Research Fund Tier 2 (Grant\nNo. MOE2015-T2-2-144). FCC also acknowledges sup-\nport from the Ministry of Science and Technology of\nTaiwan under Grants Nos. MOST-104-2112-M-110-002-\nMY3 and the support under NSYSU-NKMU JOINT RE-\nSEARCH PROJECT #105-P005 and #106-P005. He is\nalsogratefultotheNationalCenterforHigh-performance\nComputing for computer time and facilities.7\n[1] A. Bansil, H. Lin, and T. Das, Rev. Mod. Phys. 88,\n021004(2016).\n[2] X. L. Qi and S. C. Zhang, Rev. Mod. Phys. 83, 1057\n(2011).\n[3] M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045\n(2010).\n[4] B. Yan and C. Felser, Annu. Rev. Condens. Matter Phys.\n8, 337 (2017).\n[5] Kane, C. L., and E. J. Mele, Phys. Rev. Lett. 95, 226801\n(2005).\n[6] Kane, C. L., and E. J. Mele, Phys. Rev. Lett. 95, 146802\n(2005).\n[7] B. A. Bernevig, T. L. Hughes, and S. C. Zhang, Science\n314, 1757 (2006).\n[8] L. Fu, C. L. Kane, and E. J. Mele, Phys. Rev. Lett. 98,\n106803 (2007).\n[9] J. E. Moore, and L. Balents, Phys. Rev. B 75, 121306\n(2007).\n[10] R. Roy, Phys. Rev. B 79, 195322 (2009).\n[11] M. Konig, S. Wiedmann, C. Brune, A. Roth, H. Buh-\nmann, L.W. Molenkamp, X.-L. Qi, and S.-C. Zhang, Sci-\nence318, 766 (2007).\n[12] D.Hsieh, D.Qian, L. Wray,Y.Xia, Y.S.Hor, R.J. Cava,\nand M. Z. Hasan, Nature (London) 452, 970 (2008).\n[13] D. Hsieh et al., Nature (London) 460, 1101 (2009).\n[14] D. Hsieh et al., Phys. Rev. Lett. 103, 146401 (2009).\n[15] Y. L. Chen et al., Science 325, 178 (2009).\n[16] Y. Xia et al., Nat. Phys. 5, 398 (2009).\n[17] A. P. Schnyder, S. Ryu, A. Furusaki, and A. W. W. Lud-\nwig, Phys. Rev. B 78, 195125(2008).\n[18] A. Kitaev, AIP Conf. Proc. 1134, 22 (2008).\n[19] S. B. Chung, and S. C. Zhang, Phys. Rev. Lett. 103,\n235301 (2009).\n[20] X. L. Qi, T. L. Hughes, S. Raghu, and S.-C. Zhang, Phys.\nRev. Lett. 102, 187001 (2009).\n[21] H. Zhang, C.-X. Liu, X.-L. Qi, X. Dai, Z. Fang, and S.-C.\nZhang, Nature Phys. 5, 438 (2009).\n[22] L. Fu and C. L. Kane, Phys. Rev. Lett. 100, 096407\n(2008).\n[23] X. L. Qi, T. L. Hughes, and S.-C. Zhang, Phys. Rev. B\n82, 184516 (2010).\n[24] L. Santos, T. Neupert, C. Chamon, and C. Mudry, Phys.\nRev. B81, 184502 (2010).\n[25] J. D. Sau, R. M. Lutchyn, S. Tewari, and S. Das Sarma,\nPhys. Rev. Lett. 104, 040502 (2010).\n[26] A. Das et al., Nature Phy. 8, 887 (2012).\n[27] S. A. Yang, H. Pan, and F. Zhang, Phys. Rev. Lett. 113,\n046401 (2014).\n[28] M. Z. Hasan, S.-Y. Xu, and G. Bian, Phys. Scripta 2015,\n014001 (2015).\n[29] L. Balents, Physics 4, 36 (2011).\n[30] A. M. Turner and A. Vishwanath, Beyond band insula-\ntors: Topology of semimetals and interacting phases (El-\nsevier, New York, 2013).\n[31] S. A. Yang, SPIN 6, 1640003 (2016).\n[32] C. Fang, H. Weng, X. Dai, and Z. Fang, Chin. Phys. B\n25, 117106 (2016).\n[33] A. A. Burkov, arXiv:1704.06660.\n[34] S. M. Young, S. Zaheer, J. C. Y. Teo, C. L. Kane, E. J.\nMele, and A. M. Rappe, Phys. Rev. Lett. 108, 140405\n(2012).[35] Z. Wang, Y. Sun, X.-Q. Chen, C. Franchini, G. Xu, H.\nWeng, X. Dai, and Z. Fang, Phys. Rev. B 85, 195320\n(2012).\n[36] Z. Wang, H. Weng, Q. Wu, X. Dai, and Z. Fang, Phys.\nRev. B88, 125427 (2013).\n[37] Z. K. Liu, B. Zhou, Y. Zhang, Z. J. Wang, H. M. Weng,\nD. Prabhakaran, S.-K. Mo, Z. X. Shen, Z. Fang, X. Dai,\nZ. Hussain, and Y. L. Chen, Science 343, 864 (2014).\n[38] T. R. Chang et al., Phys. Rev. Lett. 119, 026404 (2017).\n[39] X. Wan, A. M. Turner, A. Vishwanath, and S. Y.\nSavrasov, Phys. Rev. B 83, 205101 (2011).\n[40] B. Singh, A. Sharma, H. Lin, M. Z. Hasan, R. Prasad,\nand A. Bansil, Phys. Rev. B 86, 115208 (2012).\n[41] H. Weng, C. Fang, Z. Fang, B. A. Bernevig, and X. Dai,\nPhys. Rev. X 5, 011029 (2015).\n[42] S.-M. Huang, S.-Y. Xu, I. Belopolski, C.-C. Lee, G.\nChang, B. Wang, N. Alidoust, G. Bian, M. Neupane,\nC. Zhang, S. Jia, A. Bansil, H. Lin, and M. Z. Hasan,\nNat. Commun. 6, 7373 (2015).\n[43] A. A. Burkov, M. D. Hook, and L. Balents, Phys. Rev.\nB84, 235126 (2011).\n[44] M. Phillips and V. Aji, Phys. Rev. B 90, 115111 (2014).\n[45] C.-K. Chiu and A. P. Schnyder, Phys. Rev. B 90, 205136\n(2014).\n[46] H. Weng, Y. Liang, Q. Xu, R. Yu, Z. Fang, X. Dai, and\nY. Kawazoe, Phys. Rev. B 92, 045108 (2015)\n[47] K. Mullen, B. Uchoa, and D. T. Glatzhofer, Phys. Rev.\nLett.115, 026403 (2015).\n[48] Y. Chen, Y. Xie, S. A. Yang, H. Pan, F. Zhang, M. L.\nCohen, and S. Zhang, Nano Lett. 15, 6974 (2015).\n[49] C. Fang, Y. Chen, H.-Y. Kee, and L. Fu, Phys. Rev. B\n92, 081201 (2015).\n[50] R. Yu, H. Weng, Z. Fang, X. Dai, and X. Hu, Phys. Rev.\nLett.115, 036807 (2015).\n[51] Y. Kim, B. J. Wieder, C. L. Kane, and A. M. Rappe,\nPhys. Rev. Lett. 115, 036806 (2015).\n[52] R. Li, H. Ma, X. Cheng, S. Wang, D. Li, Z. Zhang, Y. Li,\nand X.-Q. Chen, Phys. Rev. Lett. 117, 096401 (2016).\n[53] S. Li, Z.-M. Yu, Y. Liu, S. Guan, S.-S. Wang, X. Zhang,\nY. Yao, and S. A. Yang, Phys. Rev. B 96, 081106 (2017).\n[54] Y. Chen, H.-S. Kim, and H.-Y. Kee, Phys. Rev. B 93,\n155140 (2016).\n[55] S. Li, Y. Liu, S.-S. Wang, Z.-M. Yu, S. Guan, X.-L.\nSheng, Y. Yao, and S. A. Yang, Phys. Rev. B 97, 045131\n(2018).\n[56] T. Bzdusek, Q. Wu, A. Ruegg, M. Sigrist, and A. A.\nSoluyanov, Nature 538, 75 (2016).\n[57] S.-S. Wang, Y. Liu, Z.-M. Yu, X.-L. Sheng, and S. A.\nYang, Nat. Commun. 8, 1844 (2017).\n[58] M. Zeng, C. Fang, G. Chang, Y.-A. Chen, T. Hsieh, A.\nBansil, H. Lin, and L. Fu, arXiv:1504.03492.\n[59] S. Kobayashi, Y. Yamakawa, A. Yamakage, T. Inohara,\nY. Okamoto, and Y. Tanaka, Phys. Rev. B 95, 245208\n(2017).\n[60] X.-L. Sheng, Z.-M. Yu, R. Yu, H. Weng, and S. A. Yang,\nJ. Phys. Chem. Lett. 8, 3506 (2017).\n[61] G. Chang, S.-Y. Xu, X. Zhou, S.-M. Huang, B. Singh, B.\nWang, I. Belopolski, J. Yin, S. Zhang, A. Bansil, H. Lin,\nand M. Z. Hasan, Phys. Rev. Lett. 119, 156401 (2017).\n[62] C. Zhong, Y. Chen, Z.-M. Yu, Y. Xie, H. Wang, S. A.\nYang, and S. Zhang, Nat. Commun. 8, 15641 (2017).8\n[63] W. Chen, H.-Z. Lu, and J.-M. Hou, Phys. Rev. B 96,\n041102 (2017).\n[64] Z. Yan, R. Bi, H. Shen, L. Lu, S.-C. Zhang, and Z. Wang,\nPhys. Rev. B 96, 041103 (2017).\n[65] P.-Y. Chang and C.-H. Yee, Phys. Rev. B 96, 081114\n(2017).\n[66] M. Ezawa, Phys. Rev. B 96, 041202 (2017).\n[67] Z. Wang, A. Alexandradinata, R. J. Cava, and B. A.\nBernevig, Nature 532, 189 (2016).\n[68] S. M. Young and C. L. Kane, Phys. Rev. Lett. 115,\n126803 (2015).\n[69] S. Guan, Y. Liu, Z.-M. Yu, S.-S. Wang, Y. Yao, and S.\nA. Yang, Phys. Rev. Materials 1, 054003 (2017).\n[70] Q. Liu and A. Zunger, Phys. Rev. X 7, 021019 (2017).\n[71] B.-J. Yang and N. Nagaosa, Nat. Commun. 5, 4898\n(2014).\n[72] P. W. Anderson and E. I. Blount, Phys. Rev. Lett. 14,\n217 (1965); Phys. Rev. Lett. 14, 532(1965).\n[73] Y. Shi et al., Nat. Mater. 12, 1024 (2013).\n[74] R. Li, Y. Xu, J. He, S. Ullah, J. Li, J.-M. Liu, D. Li, C.\nFranchini, H. Weng, and X.-Q. Chen, arXiv:1610.07142.\n[75] K. Schwarz and P. Blaha, Comput. Mater. Sci. 28, 259\n(2003).\n[76] J. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett.\n77, 3865 (1996).\n[77] A. H. MacDonald, W. E. Pickett, and D. D. Koelling, J.\nPhys. C: Solid State Phys. 13, 2675 (1980).\n[78] A. A. Mostofi, J. R. Yates, G. Pizzi, Y. S. Lee, I. Souza,\nD. Vanderbilt, and N. Marzari, Comput. Phys. Commun.\n185, 2309(2014)\n[79] N. Marzari and D. Vanderbilt, Phys. Rev. B 56,22(1997).\n[80] I. Souza, N. Marzari, and D. Vanderbilt, Phys. Rev. B\n65, 035109 (2001).\n[81] N. Marzari, A. A. Mostofi, J. R. Yates, I. Souza, and D.\nVanderbilt, Rev. Mod. Phys. 84, 1419 (2012)\n[82] J. Kuneˇ s, R. Arita, P. Wissgott, A. Toschi, H. Ikeda, an d\nK. Held, Comput. Phys. Commun. 181, 1888 (2010).\n[83] M. P. L. Sancho, J. M. L. Sancho, J. Rubio, J. Phys. F\nMet. Phys. 15, 851(2000).\n[84] J. Xiong, S. K. Kushwaha, T. Liang, J. W. Krizan, M.\nHirschberger, W. Wang, R. Cava, and N. P. Ong, Science\n350, 413 (2015).\n[85] Y. Gao, S. A. Yang, and Q. Niu, Phys. Rev. B 95, 165135\n(2017).\n[86] P. J. W. Moll, N. L. Nair, T. Helm, A. C. Potter, I.\nKimchi, A. Vishwanath, and J. G. Analytis, Nature 535,\n266 (2016).\n[87] S. Guan, Z.-M. Yu, Y. Liu, G.-B. Liu, L. Dong, Y. Lu,\nY. Yao, and S. A. Yang, npj Quantum Materials 2, 23\n(2017).\n[88] X. Dai, H.-Z. Lu, S.-Q. Shen, and H. Yao, Phys. Rev. B\n93, 161110 (2016).\n[89] B. Roy, P. Gaswami, and V. Juriˇ ci´ c, Phys. Rev. B 95,\n201102 (2017).\n[90] S. Ahn, E. J. Mele, and H. Min, Phys. Rev. Lett. 119,\n147402 (2017).\n[91] Y. Liu, S. A. Yang, and F. Zhang, Phys. Rev. B 97,\n035153 (2018).\n[92] T. T. Heikkila and G. E. Volovik, JETP Lett. 93, 59\n(2011)." }, { "title": "1004.2934v2.Low_Energy__Coherent__Stoner_like_Excitations_in_CaFe___2__As___2__.pdf", "content": "arXiv:1004.2934v2 [cond-mat.mtrl-sci] 29 Jun 2010Low Energy, Coherent, Stoner-like Excitations in CaFe 2As2\nLiqin Ke1, M. van Schilfgaarde1, J.J.Pulikkotil2, T. Kotani3, and V.P.Antropov2\n1School of Materials, Arizona State University\n2Ames Laboratory, IA 50011 and\n3Tottori University, Tottori, Japan\n(Dated: October 9, 2018)\nAbstract\nUsing linear-response density-functional theory, magnet ic excitations in the striped phase of\nCaFe2As2are studied as a function of local moment amplitude. We find a n ew kind of excitation:\nsharp resonances of Stoner-like (itinerant) excitations a t energies comparable to the N´ eel temper-\nature, originating largely from a narrow band of Fe dstates near the Fermi level, and coexist with\nmore conventional (localized) spin waves. Both kinds of exc itations can show multiple branches,\nhighlighting the inadequacy of a description based on a loca lized spin model.\n1Magnetic interactions are likely to play a key role in mediating supercon ductivity in\nthe recently discovered family of iron pnictides; yet their characte r is not yet well under-\nstood. In particular, whether the system is best described in term s of large, local magnetic\nmoments centered at each Fe site, in which case elementary excitat ions are collective spin\nwaves (SWs) called magnons, or is itinerant (elementary excitations characterized by sin-\ngle particle electron-hole transitions) is a subject of great debate . This classification also\ndepends on the energy scale of interest. The most relevant energ y scale in CaFe 2As2and\nother pnictides ranges to about twice the N´ eel temperature, 2 TN≈40 meV. Unfortunately,\nneutron scattering experiments have focused on the character of excitations in the 150-200\nmeV range [1, 2], much larger than energy that stabilizes observed m agnetism or supercon-\nductivity. Experiments in Refs. [1, 2], while very similar, take complet ely different points\nof view concerning the magnetic excitations they observe. There is a similar dichotomy in\ntheoretical analyses of magnetic interactions[3, 4]. Model descrip tions usually postulate a\nlocal-moments picture [5]. Most ab initio studies start from the local spin-density approxi-\nmation (LSDA) to density functional theory. While the LSDA traditio nally favors itinerant\nmagnetism (weak on-site Coulomb correlations), practitioners str ongly disagree about the\ncharacter of pnictides; indeed the same results are used as a proo f of both localized and\nitinerant descriptions [3, 4].\nThe dynamic magnetic susceptibility (DMS), is the central quantity t hat uniquely char-\nacterizes magnetic excitations. It can elucidate the origins of magn etic interactions and\ndistinguish between localized and itinerant character. However, th e dynamical linear re-\nsponse is very difficult to carry out computationally; studies to date been limited to a few\nsimple systems. Here we adapt an all-electron linear response techn ique developed recently\n[6, 7] to calculate the transverse DMS, χ(q,ω). Besides SW excitations seen in neutron\nscattering, we find low energy particle-hole excitations at low q, and also at high q. In\nstark contrast to conventional particle-hole excitations in the St oner continuum, they can\nbe sharply peaked in energy (resonances) and can be measured.\nThe full transverse DMS χ(r,r′,q,ω) is a function of coordinates randr′(confined to\nthe unit cell). It is obtained from the non-interacting susceptibility χ0(r,r′,q,ω) via the\nstandard relation [8]\nχ=χ0[1−χ0I]−1(1)\nIis the exchange-correlation kernel. When computed within the time- dependent LSDA\n2−1−0.5 00.5 112344\nM=0.4\nM=1.1(a)N(E)\nE (eV) 0.2 0.6 11.400.511.522.5N(EF)\nM(µB)(b)\nlocalized itinerant\n00.51.01.502468M=0.4\nM=0.8\nM=1.1\n(c)Imχ0(q=0)\nω (eV)\nFIG. 1: ( a)N(E), in units of eV−1per cell containing one Fe atom. Data are shown for M=0.4µB\nand 1.1µB. Both kinds of transitions are reflected in peaks in χ0(q=0,ω), shown in panel ( c) for\nM=0.4, 0.8, and 1.1 µB. (b)N(E) at the Fermi level EFas a function of moment M. The blue\nvertical bar denotes the experimental moment, and also appr oximately demarcates the transition\nfrom itinerant to localized behavior. N(EF) in the nonmagnetic case is shown as a dotted line. ( c)\nbare susceptibility χ0(q=0,ω) in the same units. The text discusses the significance of the arrows\nin panels ( a) and (c).\n(TDLDA) Iis local: I=I(r)δ(r−r′) [8].χ0can be obtained from the band structure\nusing the all-electron methodology we developed [6, 7]. To reduce the computational cost\nwe calculate Im χ0and obtain Re χ0from the Kramers Kronig transformation [6]. Im χ0\noriginates from spin-flip transitions between occupied states at kand unoccupied states at\nk+q: it is a k-resolved joint density of states Ddecorated by products Pof four wave\nfunctions [7]\nD(k,q,ω) =f(ǫ↑\nk)(1−f(ǫ↓\nq+k))δ(ω−ǫ↓\nq+k+ǫ↑\nk) (2)\nImχ+−\n0=/integraldisplay\ndωd3kP(r,r′,k,q)×D(k,q,ω) (3)\nEven with the Kramers-Kronig transformation, Eq. (3) poses a hu ge computational burden\nfor the fine frequency and kresolution required here. We make a simplification, map-\npingχ0onto the local magnetization density which is assumed to rotate rigid ly. The full\nχ0(r,r′,q,ω) simplifies to the discrete matrix χ0(R,R′,q,ω) associated with pairs of mag-\nnetic sites ( R,R′) in the unit cell; and I(r) simplifies to a diagonal matrix IRR. In Ref. [7]\n3we show that we need not compute Iexplicitly but can determine it from a sum rule. I\ncan be identified with the Stoner parameter in models. We have found that for Fe and Ni\nthese approximations yield results in rather good agreement with th e full TDLDA results,\nand expect similar agreement here. We essentially follow the procedu re described in detail\nin Ref. [7], and obtain χ(q,ω) as a 4×4 matrix corresponding to the four Fe sites in the unit\ncell. To make connection with neutron experiments, spectra are ob tained from the matrix\nelement/summationtext\nR,R′< eiq·R|χ(R,R′,q,ω)|eiq·R′>. For brevity we omit indices RR′henceforth.\nAswe will see, thecharacter of χ0(q,ω)can largelybeunderstood fromstates near EF, so\nwe begin by analyzing the ground state band and magnetic structur e of the low-temperature\n(striped) phase of CaFe 2As2within the LSDA. Results for tetragonal and orthorhombic\nstriped structures are very similar, suggesting that the slight diffe rence in measured aand\nblattice parameters plays a minor role in the description of magnetic pr operties. Experi-\nmental lattice parameters were employed[9]. CaFe 2As2has an internal parameter zwhich\ndetermines the Fe-As distance RFe−As. Here we treat zas a parameter whos main effect is\nto control the magnetic moment M, which in turn strongly affects the character of magnetic\ninteraction, as we will show.\nIn Fig. 1( a), the density of states N(E) is shown over a 2 eV energy window for a small\nand large moment case ( M=0.4µBand 1.1µB). Particularly of note is a sharp peak near\nEF, of width ∼50 meV. This narrow band is found to consist almost entirely of major ity-\nspin Fedxyanddyzorbitals. The peak falls slightly below EFforM=0.4µBand slightly\nabove for M=1.1µB. Thus, as RFe−Asis smoothly varied so that Mchanges continuously,\nthis peak passes through EF. As a result N(EF) reaching a maximum around M=0.8µB\n(Fig. 1(b)). This point also coincides with the LSDA minimum-energy value of RFe−As.\nThat this unusual dependence originates in the magnetic part of th e Hamiltonian can be\nverified by repeating the calculation in the nonmagnetic case. As Fig. 1(b) shows, the\nnonmagnetic N(EF) is large, and approximately independent of RFe−As. In summary, the\nmagnetic splitting produces a pseudogap in N(E) for large M; the pseudogap shrinks as\nMdecreases and causes a narrow band of Fe dstates to pass through EF, creating a sharp\nmaximum in N(EF) nearM=0.8µB. As we will show below that moment demarcates a\npoint of transition from itinerant to localized behavior (see also Ref. [13]).\nNext we turn to magnetic excitations. In the standard picture of m agnons in metals,\nImχ0(ω) is significant only for frequencies exceeding the magnetic (Stoner ) splitting of the d\n4bands,ǫ↓\nd−ǫ↑\nd=IM(cfEq. 2). In such cases Im χ0is small at low qand low energies and well\ndefined magnons appear at energies near |1−IReχ0|= 0 (cfEq. 1). As Im χ0increases, it\ninitially broadens the (formerly sharp) SW spectrum ¯ ω(q); as it becomes large the spectrum\ncan become incoherent, or (Stoner) peaks can arise from Im χ0, possibly enhanced by small\n1−IReχ0. Fig. 1(c) shows Im χ0(q=0,ω) on a broad energy scale. The picture we developed\nforN(E) leads naturally to a classification of Stoner transitions into three m ain types.\n(1) Excitations between the (largely) Fe dxzstates centered near −0.5 eV to dstates\ncentered near 1 eV (for M=1.1µB). The magnetic splitting of these states matches well\nwith the usual Stoner splitting IMin localized magnets, and scales with M. (I≈1 eV in\n3dtransition metals.) These high-energy transitions are depicted by a large red arrow in\nFig. 1(a) forM=1.1µB, and by vertical arrows in Fig. 1( c) showing χ0(q=0,ω) forM=0.4,\n0.8, and 1.1 µB. They are too high in energy to be observed by neutron measureme nts.\n(2) Excitations from dxztodxyanddyzstates just above the psueudogap, depicted by\na small red arrow in Fig. 1( a), and slanting arrows in Fig. 1( d). These are the excita-\ntions probably detected in neutron measurements [1, 2]. This pseud ogap is well defined for\nM≥1.1µB, but is modified as Mdecreases, which leads to the following:\n(3) Near M=0.8µB, the narrow dband passes through EF, opening up channels, not\npreviously considered, for low-energy, particle-hole transitions w ithin this band. When M\nreaches 0.4 µB, this band has mostly passed through EFand the pseudogap practically\ndisappears.\nWhat makes the pnictide systems so unusual is that Im χ0is already large at very low\nenergies ( ∼10 meV) once the sharp peak in N(E) approaches EF. One of our central find-\nings is that this system undergoes a transition from localized to a coexistence of localized\nand itinerant behavior asMdecreases from M/greaterorsimilar1.1µBtoM≈0.8µB. Moreover, the itiner-\nant character is of an unusual type: elementary excitations are m ostly single particle-hole\nlike: they can be well defined in energy and q. Those represent coherent excitations. The\ndependence of N(EF) onMis not only responsible for them, but also may explain the\nunusual linear temperature-dependence of paramagnetic susce ptibility, and the appearance\nof a Lifshitz transition with Co doping [10].\nFig. 2 focuses on the AFM line, q=[H00]2π/a. Panel ( a) shows the full Im χ(ω) for\nM=1.1µBfor several q-points spanning the entire line, 0 0.8 the peaks are strongly broadened,\nespecially for small M. ¯ωis in good agreement with neutron data of Ref.[2], except that\nneutron data are apparently smaller than large Mcalculations predict [11].\nOne experimental measure of the validity of the local-moment pictur e is the ratio of half-\nwidth at half-maximum (HWHM) Γ to ¯ ω, shown in Fig. 2( d). For large Mand most of q,\nΓ/¯ω∼0.15-0.18. For intermediate and small M, Γ/¯ω≥0.2 for a wide diapason of q. This\nis significant: it reflects the increasing Stoner character of the ele mentary excitations. Were\nthere an abrupt transition into a conventional Stoner continuum a s argued in Ref. [2], it\nwould be marked by an abrupt change in Γ /¯ω. This is not observed; yet damping appears\nto increase with energy and q, reflecting normal metallic behavior.\nSpectra for H=0.9 (Fig. 2( c)) adumbrate two important findings of this work. When\nM=0.8µB, a sharp peak in χ(ω) appears near 10 meV. There is a sharp peak in χ0(ω) at\n¯ω≈10 meV also, classifying this as a particle-hole excitation originating fr om the narrow d\nband depicted in Fig. 1. Being well defined in energy it is coherent, ana logous to a SW,\n6only with a much larger Γ /¯ω. Yet it is strongly enhanced by collective interactions, since\n|1−IReχ0|ranges between 0.1 and 0.2 for ω<100meV. This new kind of itinerant excitation\nwill be seen at many values q, typically at small q. The reader many note the sharp rise\nand fall in Γ /¯ωat small qin Fig. 2( d). This anomaly reflects a point where a SW and a\nparticle-hole excitation coalesce to the same ¯ ω.\nReturning to H=0.9 when M=1.1, there is a standard (broadened) SW at 200 meV; cf\ncontours in Fig. 2( b). Asecond, low-energy excitation can be resolved near 120 meV. This\nis no corresponding zero in the denominator in Eq. (1), but no stron g peak in χ0(ω), either.\nThis excitation must be classified as a hybrid intermediate between St oner excitations and\nSWs. Only a single peak remains when M=0.4µBand the peak in N(E) has mostly passed\nthroughEF(Fig. 1).\n010020030040013\n5 7911\nω(meV)(a)\n00.20.40.60.80100200300\nKω\n(b)\nFIG. 3: Im χ(q,ω) along the FM axis, q=[0,K,0]2π/a. (a) Imχ(ω) atK= 1/12, 3/12, 5/12, 7/12,\n9/12, 11/12, for M=0.8µB. Strong Stoner excitations can be seen for K=1/12 and 3/12 near\n¯ω∼10-20 meV. ( b): Contours |1−IReχ0|= 0 in the ( ω,K) plane, for M=1.1µBand 0.8µB,\nanalogous to Fig. 2( b), and dominant peak positions ¯ ωobtained by a nonlinear least-squares fit of\none or two gaussian functions to χ(ω) over the region where peaks occur.\nAlong the FM line q=(0,K,0)2π/a,χ(ω) is more complex, and more difficult to interpret.\nAtM=1.1µBsharp, well defined collective excitations are found at low q, and broaden with\nincreasing q. There is a reasonably close correspondence with the zeros of |1−IReχ0|and\npeaks in χ, as Fig. 3( b) shows. The M=0.8µBcase is roughly similar, except that for K>0.4\nexcitations cannot be described by a single peak. Note that for fixe dq, propagating spin\nfluctuations (characterized by peaks at ¯ ω) can exist at multiple energies — the magnetic\nanalog of the dielectric function passing through zero and sustainin g plasmons at multiple\nenergies. Peaks in χ(ω) can broaden as a consequence of this; the K=7/12, 9/12 and 11/12\ndata of Fig. 3( a) are broadened in part by this mechanism as distinct from the usual one\n(intermixing of Stoner excitations). Second, consider how ∂¯ω/∂qchanges with MforK>0.5\n7(Fig. 3(b)). When M=1.1µB, ¯ωincreases monotonically with K. ForM=0.8µB, ¯ωhas a\ncomplex structure but apparently reaches a maximum before Kreaches 1. That ∂¯ω/∂q\nchanges sign is significant: it marks the disappearance of magnetic f rustration between the\nferromagnetically aligned spins at lowmoments, andthe emergence o f stable FMorder along\n[010], characteristic of local-moment behavior (see also Ref.[12]). Ex perimentally, Ref. [2]\nreports∂¯ω/∂q>0 forK>0.5. While our calculations provide a clear physical interpretation,\nwe note significant differences in the moment where we observe this e ffect (M=0.8-1.1µB)\nand the effective spin ( S=0.2) used in Ref.[2].\n0501001500.1\n0.2\n0.3\n0.4\n0.5\nω(meV)(a)\n00.10.20.30.4Lω\n050100(b)\nFIG. 4: Im χ(q,ω) along the caxis,q=[0,0,L]2π/c. (a) Imχ(ω) at values of Llisted in the key,\nforM=0.8µB. WhenM=1.1µB(not shown) the spectrum is well characterized by a single sh arp\npeak at any L. WhenM=0.8 or 0.4 µBthe SW peak remains, but a low-energy Stoner excitation\ncoexists with it. The latter are most pronounced for larger Lbut can be resolved at every L. Peak\npositions (single for M=1.1µB, double for M=0.8 and 0.4 µB) are shown in ( b).\nElementary excitations along the caxis,q=(0,0,L)2π/c, bring into highest relief the\ntransition from pure local-moment behavior (collective excitations) , to one where coherent\nitinerant Stoner and collective excitations coexist. Collective excita tions are found for all\nLand allM. ForM=1.1µB, a single peak is found: excitations are well described by the\nHeisenberg model with weak damping. Comparing Fig. 4( b) to Figs. 2( b) and 3( b), it\nis apparent that ¯ ωrises much more slowly along Lthan along HorK, confirming that\ninterplane interactions are weak. When Mdrops to 0.8 µB, a second-low energy peak at\n¯ω′emerges at energies below 20 meV, for small qalong [0,K,0], and for all qalong [0,0, L],\ncoexisting with the collective excitation. Why these transitions are a bsent for M=1.1µBand\nare so strong at M=0.8µBcan be understood in terms of roughly cylindrical Fermi surface\natk=(1/2,0,kz). Single spin-flip transitions between occupied states ǫ↑\nkand an unoccupied\nstatesǫ↓\nk+qseparated by ∼10 meV are responsible for this peak (see Eq. 2). They originate\nfrom “hot spots” where N(ǫ↑\nk) andN(ǫ↓\nk+q) are both large.\n8From the SW velocities , (∂¯ω/∂q)q=0, anisotropy of the exchange couplings can be de-\ntermined [4, 12, 14]. We find strong in-plane and out-of-plane anisot ropies, and pre-\ndictvb/va=0.55 and find vc/va=0.35, where va=490meV ·˚A. Neutron scattering experi-\nments [14, 15] have measured with vc/va−bto be∼0.2-0.5.\nIn summary, we broadly confirm the experimental findings of Refs. [1, 2], that there is\na spectrum of magnetic excitations of the striped phase of CaFe 2As2, which for the most\npart are weakly damped at small qand more strongly damped at large q. A new kind of\nexcitation was found, which originates from single particle-hole tran sitions within a narrow\nband of states near EF, renormalized by a small denominator |1−IReχ0|(Eq. (1)). They\nappear when the Fe moment falls below a threshold, at which point the narrow band passes\nthroughEF. The character of itineracy is novel: excitations occur at low-ener gy, at energy\nscales typically below TN, and can be sharply peaked, which should make them accessible\nto experimental studies. The distinction between the two kinds of e xcitations is particularly\nobservable in the anomalous dependence of Γ /¯ωonqin Fig. 2( d). Collective spin-wave-like\nexcitations also are unusual: multiple branches are found at some q. Finally, at higher q\nwe find that the SW velocity changes sign, which is also observed expe rimentally at similar\nvalues of M.\nOverall a picture emerges where localized and itinerant magnetic car riers coexist and\ninfluence each other. This description falls well outside the framewo rk of a local-moments\nmodel such as the Heisenberg hamiltonian. The low-energy particle- hole excitations are\nstrongly affected by Fe-As distance (and thus by lattice vibrations ) while the SWs are much\nless so. While phonons and Stoner excitations are generally consider ed separately, these\nfindings suggest that they strongly influence each other.\nThis work was supported by ONR, grant N00014-07-1-0479 and by DOE contract DE-\nFG02-06ER46302. Work at the Ames Laboratory was supported b y DOE Basic Energy\nSciences, Contract No. DE-AC02-07CH11358.\n[1] S.O. Diallo et al., Phys. Rev. Lett. 102, 187206 (2009).\n[2] J. Zhao et al., Nature Physics 5, 555 (2009).\n[3] M. D. Johannes, I.I. Mazin. Phys. Rev. B 79, 220510 (2009) ; I.I. Mazin, J. Schmalian. Physica\n9C, 469, 614 (2009).\n[4] Z. P. Yin et al., Phys. Rev. Lett. 101, 047001 (2008).\n[5] Chen Fang et al., Phys. Rev. B 77, 224509 (2008); F. Kruger , S. Kumar, J. Zaanen and J.\nvan den Brink. Phys. Rev.B 79, 054504 (2009); C. Lee, W. Lin an d W. Ku. arXiv.0905.3782;\nR. R. P. Singh. arxiv.0903.4408.\n[6] T. Kotani, M. van Schilfgaarde, S. V. Faleev, Phys. Rev. B 76, 165106 (2007).\n[7] T. Kotani and M. van Schilfgaarde, J. Phys. Cond. Matt. 20 , 295214 (2008).\n[8] R. Martin, Electronic Structure , Cambridge University Press (Cambridge, 2004).\n[9] A.I. Goldman et al., Phys. Rev. B 78, 100506 (2008).\n[10] N. Ni et al., Phys. Rev. B 80, 024511 (2009); E. D. Mun, S. L . Bud’ko, N. Ni, P. C. Canfield.\narXiv:0906.1548.\n[11] The origin of the discrepancy in ¯ ωwith experiment is not clear. Our calculation contains\nseveral important approximations (local potential, linea r response, rigid spin).\n[12] J J Pulikkoti et. al, Supercond. Sci. Technol. 23, 054012 (2010).\n[13] G. Samolyuk and V.P.Antropov. Phys. Rev. B 79, 052505 (2 009).\n[14] R. J. McQueeney et al., Phys. Rev. Lett. 101, 227205 (200 8).\n[15] J. W. Lynn and P. Dai. Physica C469, 469 (2009).\n10" }, { "title": "1508.00624v1.High_field_response_of_gated_graphene_at_THz_frequencies.pdf", "content": "High \feld response of gated graphene at THz frequencies\nHadi Razavipour, Wayne Yang, Abdeladim Guermoune, Michael Hilke, and David G. Cooke\u0003\nDepartment of Physics, McGill University, Montr\u0013 eal, Qu\u0013 ebec, Canada\nIbraheem Al-Naib and Marc M. Dignam\nDepartment of Physics, Engineering Physics and Astronomy, Queen's University, Kingston, Ontario, Canada\nFran\u0018 cois Blanchard\nD\u0013 epartement de g\u0013 enie \u0013Electrique, \u0013Ecole de technologie sup\u0013 erieure, Montr\u0013 eal, Qu\u0013 ebec, Canada\nHassan A. Hafez, Xin Chai, Denis Ferachou, and Tsuneyuki Ozaki\nINRS-EMT, Advanced Laser Light Source, Varennes, Qu\u0013 ebec, Canada\nPierre L\u0013 evesque and Richard Martel\nD\u0013 epartement de Chimie, Universit\u0013 e de Montr\u0013 eal, Montr\u0013 eal, Qu\u0013 ebec, Canada\nWe study the Fermi energy level dependence of nonlinear terahertz (THz) transmission of gated\nmulti-layer and single-layer graphene transferred onto sapphire and quartz substrates. The two\nsamples represent two limits of low-\feld impurity scattering: short-range neutral and long-range\ncharged impurity scattering, respectively. We observe an increase in the transmission as the \feld\namplitude is increased due to intraband absorption bleaching starting at \felds above 8 kV/cm. This\ne\u000bect arises from a \feld-induced reduction in THz conductivity that depends strongly on the Fermi\nenergy. We account for intraband absorption using a free carrier Drude model that includes neutral\nand charged impurity scattering as well as optical phonon scattering. We \fnd that although the\nFermi-level dependence in the monolayer and \fve-layer samples is quite di\u000berent, both exhibit a\nstrong dependence on the \feld amplitude that cannot be explained on the basis of an increase in\nthe lattice temperature alone. Our results provide a deeper understanding of transport in graphene\ndevices operating at THz frequencies and in modest kV/cm \feld strengths where nonlinearities\nexist.\nPACS numbers: 78.67.Wj, 72.20.Ht, 78.67.-n, 72.80.Vp\nI. INTRODUCTION\nThe well known linear electronic dispersion of graphene\nnaturally gives rise to unusual optoelectronic properties,1\npotentially useful for high-speed modulators and the\nother active components. While it has been touted as\nan important material for terahertz (THz) frequency\ndevices,2charge transport properties at THz frequencies\nunder \feld conditions commonly found in devices, on the\norder 10's kV/cm, are poorly understood. This is in part\ndue to a lack of control over the extrinsic properties of the\ndevices and graphene's sensitivity to its environment.3\nThe electrodynamic response of doped graphene has pre-\nviously been studied in the low \feld, linear regime under\ngated conditions4for frequencies ranging from near dc5\nto THz,6{10mid IR11and optical regimes.12Under in-\ntense \felds, however, charge carriers can acquire signi\f-\ncant momentum and for Fermi energies ( Ef) close to the\ncharge neutrality point (CNP), a complicated interplay\nbetween intra- and inter-band transitions can occur.13A\nbetter understanding of these nonlinear interactions and\ntheir density dependence is needed for the future device\ndesign, which is the aim of this study. For this purpose,\nwe study the THz electrodynamic response of graphene\nunder high \feld conditions and controlled Fermi level\nthrough a gate voltage VGapplied to an ionic gel topgate which is transmissive to THz light (see Fig. 1).\nThe development of e\u000ecient nonlinear optical methods\nfor the generation of intense THz pulses14has made it\npossible to time resolve the high \feld response of charge\nand spin excitations in materials.15In graphene, due to a\nstrong interband dipole matrix element that diverges at\nthe charge neutrality (Dirac) point, the THz light-matter\ninteraction can be easily pushed into the nonperturba-\ntive regime even by modest electric \felds on the order\nof several 10's kV/cm.16,17As the thermalization time of\naccelerated hot carriers is noted to be exceedingly short\nin graphene,18these nonlinearities can be quanti\fed as\nan elevated temperature of the electronic subsystem.23\nRecent works have demonstrated a strong THz \feld-\ninduced transparency in room-temperature graphene19in\nboth doped19{22and optically excited graphene.24This\nphenomenon has been described by ultrafast intra-band\ncarrier thermalization on the order of 30 fs,18leading\nto a transient elevated carrier temperature on the order\nof 2000 K and subsequent optical phonon emission dur-\ning cooling. In this model, the non-equilibrium heating\nand generation of optical phonons subsequently enhances\nthe carrier scattering, suppressing the \feld-driven current\nand thus THz absorption in graphene. Other sources of\ncarrier scattering that can potentially play an important\nrole in the THz response of graphene include long-rangearXiv:1508.00624v1 [cond-mat.mes-hall] 3 Aug 20152\nFIG. 1. (a) Schematic of the graphene sample on a substrate\nwith gold four-point-probe contacts and a side gate contact.\nThe broadband THz pulses traverse the sample at normal\nincidence. (b) The mechanism for ionic gating of graphene is\nshown in the inset, illustrating the accumulated Debye layer\nunder a gate voltage V G.\nimpurity scattering, due to the presence of the charged\nimpurities in the substrate, short-range scattering, due\nto neutral impurities or disorder in the graphene itself,\nand carrier-carrier scattering. While impurity scattering\ndominates at low lattice temperatures, optical phonon\nscattering becomes more dominant at higher tempera-\ntures (T > 700 K).25\nSaturable absorption due to interband contributions\nhas also been observed in the visible regime due to the\nphase-space \flling e\u000bects as well as carrier heating. At\nTHz frequencies, the threshold for an interband response\nis determined by the carrier density, as the carriers must\nbe excited at twice the quasi-Fermi energy to undergo\na transition. In this work, we study the Fermi energy\ndependence of the nonlinear THz transmission of gated\ngraphene and show that although the nonlinear response\nis similar in di\u000berent samples with di\u000berent doping lev-\nels, it does not appear that this e\u000bect can be described\nby heating of the lattice alone. The paper is organized\nas follows. In Sec. II, we discuss the possible scattering\nmechanisms and their e\u000bects on THz transmission as a\nbasis for the model used to describe the data. The results\nfor both \fve-layer and single-layer graphene samples are\npresented in Sec. III. Finally, the conclusions are sum-\nmarized in Sec. IV.II. THEORETICAL MODEL\nIn this section, we discuss the model that we use to\naccount for the e\u000bects of tuning the Fermi energy on\nthe transmission of THz light. Using this model, we\nextract the dependence of the scattering time on the\nFermi energy and the lattice and electron temperature.\nThere are four potentially important sources of carrier\nscattering in graphene: short-range impurity scattering,\nlong-range impurity scattering, phonon scattering, and\ncarrier-carrier scattering.\nStrictly speaking the carrier dynamics should be mod-\neled using a full dynamic model that includes the nonlin-\near response arising from the linear band dispersion and\ninterband transitions.13We \fnd however, in agreement\nwith our previous work,24that for the samples and \feld\nstrengths considered here, there is essentially no intrin-\nsically nonlinear response and a linear Drude model is\nsu\u000ecient. The nonlinear response of the graphene is in-\ncorporated in the model by employing a scattering time\nthat depends on the THz \feld amplitude. Thus, in what\nfollows, we employ the Drude model.\nFirst, we need to relate the transmission level to the\nconductivity of graphene. It can be easily shown that the\ntransmission through the graphene sample normalized to\nthe transmitted \feld through the substrate is given by6\nT(!) =TG+S\nTS=1 +n\n1 +n+Zo\u001b(!)N; (1)\nwhereTG+SandTSare the transmission through the\nsample and the substrate only, respectively, nis the re-\nfractive index of the substrate, \u001b(!) is the complex ac\nsheet conductivity of a single layer of graphene and Nis\nthe number of graphene layers in the sample. The con-\nductivity, dominated by the intra-band response at THz\nfrequencies is given by a simple Drude model11\n\u001b(!) =D\n(1=\u001c\u0000i!); (2)\nwhere!is the angular frequency, \u001cis the total carrier\nmomentum scattering time, and Dis the Drude weight\ngiven by9,10\nD=2e2kBTe\n(\u0019~2)ln\u0014\n2cosh\u0016f\n2kBTe\u0015\n; (3)\nwhereeis the electron charge, kBis the Boltzmann\nconstant,Teis the electron temperature, ~is the re-\nduced Planck's constant, and \u0016fis the chemical poten-\ntial. For simplicity, the chemical potential is considered\nto be equal to the Fermi energy at room temperature.\nHowever, as the electron temperature rises, we adjust\nthe chemical potential so that the net charge on the\ngraphene is unchanged as the temperature rises. The\nadjusted chemical potential along with the correspond-\ning temperature are used then to calculate the Drude3\nweight. Using Eqs. (1-3), we can extract the total ef-\nfective scattering time \u001cfor any measured transmission\nlevel for each gate voltage and THz \feld amplitude. The\ngate voltage and \feld amplitude dependence of the scat-\ntering time can therefore be directly obtained from the\nTHz transmission data once the electron temperature is\ndetermined.\nWe now turn to the model that we employ for the\nscattering time used in the Drude model. We present\na simple model of the e\u000bects of the lattice temperature\nand the carrier density on the scattering times based on\nthree scattering mechanisms: short-range neutral impu-\nrity scattering, charged long-range impurity scattering\nand scattering due to absorption of optical phonons. Al-\nthough there can be contributions to the scattering rate\ndue to interactions with acoustic phonons, these are usu-\nally considered to be relatively small at the temperatures\nconsidered here.29We do not include carrier-carrier scat-\ntering due to the di\u000eculties in accurately modelling its\ne\u000bects and because, for similar Fermi energies and THz\n\feld amplitudes, it has been shown that carrier-carrier\nscattering can be neglected without qualitatively chang-\ning the results.26We will return to this issue when dis-\ncussing our \feld-dependent results.\nThe expression for the total scattering rate is thus\ngiven by\n1\n\u001c=1\n\u001cimp+1\n\u001cop; (4)\nwhere\u001cimpis the impurity scattering time \u001copis the opti-\ncal phonon scattering time. The impurity scattering rate\nis given by\n1\n\u001cimp=1\n\u001csr+1\n\u001cCoul+1\n\u001cml: (5)\nIn this expression, 1 =\u001csris the short-range neutral impu-\nrity scattering rate, which has been shown theoretically\nto be linearly dependent on the energy of the carriers,27;\nin our model, we take this rate to be proportional to the\naverage energy per carrier with respect to the Dirac point\nK, which is given by\nEav=P\nkf\u001aee(K+k) +\u001ahh(K+k)g~vFk\nP\nkf\u001aee(K+k) +\u001ahh(K+k)g;(6)\nwhere\u001aee(\u001ahh) is the free electron (hole) population den-\nsity, andvFis the Fermi velocity. The second term in the\nimpurity scattering rate is, \u001cCoul, which is due to charged\nlong-range impurities. This rate has been shown to be\ninversely proportional to the square root of the carrier\ndensity (due to carrier screening)28. Finally,\u001cmlis an\nadditional gate-voltage-independent scattering time that\nwe \fnd is necessary in order to model the bias depen-\ndence of the scattering time in our multi-layer sample.\nThe physical origin of this term is not clear, but it may\nbe due to scattering from the ionic gel or is perhaps arisesdue to a correction term in the relationship between the\ngate voltage and the Fermi energy. Thus, the \fnal ex-\npression for the impurity scattering rate is\n1\n\u001cimp=csrEav+cCoul\nEeff+1\n\u001cml; (7)\nwherecsrandcCoulare constants that are experimentally\ndetermined and are dependent on the impurity densities,\nandEeffis the e\u000bective Fermi energy that corresponds\nto the square root of the carrier density and is given by\nD\u0019~2=e2.\nWe now turn to the scattering due to optical phonons.\nThere are two potential types of optical phonons that can\ncontribute to scattering: optical phonons in the graphene\nitself and polar surface phonons in the substrate.41,42\nWe \fnd, using the parameters in Ref. 41 that, due to\nthe weak coupling strength, the scattering rate contribu-\ntion from the surface phonons in both substrate materials\nis much smaller than those due to the graphene optical\nphonons. Thus, in what follows, we neglect the phonons\nin the substrate and include only the contribution from\nthe phonons in the graphene itself.29\nPrevious studies have shown that energy transfer from\nelectrons to phonons takes place on a timescale of about\n100 fs.30,31Thus, in principle, if there are electrons\nthat have been driven by the \feld more than an optical\nphonon energy above any vacant states, one would ex-\npect emission of optical phonons over the pulse duration,\nwhich would increase the lattice temperature. However,\nwithout performing a detailed simulation that tracks the\nenergies of individual electrons, it is very di\u000ecult to accu-\nrately include this contribution to the scattering. Thus,\nfor simplicity, we only include the process of phonon ab-\nsorption, but not phonon emission.\nThe optical phonon scattering rate due to optical\nphonon emission in the graphene is given by\n1\n\u001cop=D2\no\n(2~2v2\nF\u001bm!o)(Eav+~!o)\n(e~!o=kBTl\u00001); (8)\nwhere ~!o= 147 meV is the graphene optical phonon\nenergy,\u001bm= 7:6\u000210\u00008g/cm2is the 2D mass density\nof graphene29,Do= 5\u0002109eV/cm is the deformation\npotential for the optical phonons32andTlis the lattice\ntemperature. Various values for the optical phonon en-\nergy and deformation potential have been reported in the\nliterature29,32{34. We employ those found in Ref. 24, al-\nthough, as we discuss in the following sections, we have\nconsidered the e\u000bect of modifying the deformation po-\ntential on the results. We \fnally note that in all cases we\nassume equality of the carrier and lattice temperatures\nthat appear respectively in the Drude weight (Eq. (3))\nand the optical phonon scattering rate (Eq. (8)).4\nIII. EXPERIMENTAL RESULTS AND\nDISCUSSION\nIn order to investigate the e\u000bect of the substrates and\nnumber of graphene layers on the transport properties\nof graphene to the intense THz \feld, we fabricated two\ndi\u000berent samples treated in the next two sections. The\n\frst is a \fve layer graphene sample, fabricated by trans-\nferring sequentially monolayer graphene onto a sapphire\nsubstrate, and the second a monolayer sample on a quartz\nsubstrate. The substrates were patterned with 5/100 nm\nthickness of Cr/Au (prior to transfer) for the electrical\nfour-point-probe measurements as shown in Fig. 1. The\ngate was created by placing the \ffth contact at the edge\nof the substrate so that it is in contact with the ionic gel\nfor electrochemical doping. The ionic gel was fabricated\nusing the method proposed by Ref. 35 and transferred to\nthe sample under an inert glovebox environment. A uni-\nform layer of ionic gel with a thickness of less than 50 \u0016m\nwas spin coated on the graphene sample as well as the\nbare substrate to serve as a reference. Upon applying the\ngate voltage, mobile ions are transferred to the surface of\nthe graphene to form a Debye layer with a thickness on\nthe order of 1 nm.36This type of solid polymer electrolyte\ngate is much thinner than the conventional 300 nm SiO 2\nback gate, thus a much higher gate capacitance can be\nachieved. With the electrolyte ionic gating, high doping\nlevels on the order of 5 \u00021013cm\u00002are possible.36The\ngating response of the graphene is measured in terms of\nthe electrical current driven from the source to the drain\nelectrode through the graphene channel as a function of\nVG. In order to estimate the Fermi energy as a function of\ngate voltage relative to the charge neutrality point volt-\nage (VCNP), we use the following equation from Ref. 10\nfor a similar geometry and ionic gel, where\nEf= 0:346p\n(VG\u0000VCNP) [eV]: (9)\nThe case of VG=VCNP needs to be treated separately\nbecause, in agreement with other researchers,37we \fnd\nthat due to spatial inhomogeneity of the charge density\nacross the graphene sample an e\u000bective Efmust be de-\ntermined at the nominal CNP gate voltage, in a range of\n\u0006100 meV.10\nThe high-\feld time-domain THz spectrometer was de-\nveloped based on tilted-pulse front optical recti\fcation in\nlithium niobate,14with peak THz \feld amplitudes lim-\nited to 70 kV/cm in this work. The bandwidth of the\nTHz pulses was\u00183 THz and was detected using electro-\noptic sampling in a 300 \u0016m thick (110)-cut GaP crystal.\nThe emitted THz pulses were collected and focused by a\nset of o\u000b-axis parabolic mirrors to a 450 \u0016m spot size as\ndetermined by an uncooled microbolometer camera. The\ngraphene sample was positioned carefully at the THz fo-\ncus for high-\feld transmission measurements at normal\nincidence. The amplitude of the THz pulses was varied\nusing a pair of wire-grid polarizers prior to the sample in\nthe collimated section of the beam. The measurements\nwere performed by sweeping the graphene gate voltage\nFIG. 2. (a) Four-point-probe measurements of the 5-layer\nsample dc resistance with the charge neutrality point indi-\ncated by the resistance peak. (b) The Raman data (with\nthe G and 2D peaks) of single layer of CVD grown graphene\nbefore stacking to form the 5 layers.\nand registering the value of the transmitted peak THz\nelectric \feld. No THz pulse reshaping was observed as\nthe gate voltage varies from \u00002:5V(below the Dirac\npoint) to 2 :5V, thus all information on the nonlinear\ntransmission can be quanti\fed by monitoring the THz\npulse peak \feld transmission.\nA. Five-layer sample\nLarge-area, single-layer graphene was fabricated by\nchemical vapor deposition (CVD) on Cu-foil using stan-\ndard techniques. Poly(methyl methacrylate) (PMMA) is\nused to transfer the graphene to a c-cut sapphire sub-\nstrate with refractive index of 3.31 and the PMMA layer\nwas removed. The process was repeated \fve times so that\nthe \fve-layer graphene sample was obtained, with the\nlayers electrically decoupled from one another. Fig. 2(a)\nshows the sheet resistance of graphene as a function of\ngate voltage. The CNP is speci\fed as the resistance\nmaximum which occurs in this case at VG=\u00000:8V.\nWe note that hysteresis was observed in these resistance\nmeasurements when sweeping the gate voltage, which is\ncommonly observed in ionic gel gates.38,39We therefore\nonly perform measurements on the positive sweep of the\ngate voltage for repeatability. The Debye absorption of5\nFIG. 3. For the \fve-layer sample: (a) The THz pulse peak\n\feld transmission through the gated graphene sample for the\n\feld strengths indicated, normalized to the transmission at\nthe Dirac point at VG\u0000VCNP = 0. (b) The \feld dependence\nof the THz peak \feld transmission jtj=EG+S\nESat various\ngate voltages. Eincis the incident THz electric \feld at the\ngraphene sample position.\nthe ionic gel is located at lower frequencies than THz,\nproducing a transparent top gated graphene sample. All\nmeasurements in this work are performed at room tem-\nperature. Fig. 2(b) shows the Raman spectroscopy of one\nof the \fve layers of graphene before stacking. The ratio\nbetween the 2D and G peak values is 2.8 which indicates\nhigh quality single layer graphene.40\nFig. 3(a) shows the variation of the transmitted peak\nTHz electric \feld (normalized to its maximum value at\nthe Dirac point) as a function of gate voltage sweep\nfor di\u000berent incident peak electric \felds in the range\nof 2 kV/cm to 70 kV/cm. The THz modulation in-\nduced by the gate voltage sweep (Fermi level change in\ngraphene) is a strongly nonlinear function of the peak\nelectric \feld value. At low \felds, gate-induced THz mod-\nulation is more than 31% whereas for higher \feld ampli-\ntudes the modulation decreases and saturates at a value\nof 7:4% above 60 kV/cm. As in previous work on un-\ngated graphene, a saturable power transmission function\ncan be used to phenomenologically describe the behavior\nobserved in Fig. 3(b).19The power transmission is de-\n\fned asjTj=\u0010\nEsamp\nEref\u00112\n, whereEsamp is the peak THz\nelectric \feld transmitted through the ionic gel, 5 layers\nFIG. 4. For the \fve layer sample: (a) THz power transmission\nde\fned asjTj=E2\nsamp\nE2\nref, curve \ftted by Eq. 10 for two\ndi\u000berent gate voltages, one at the Dirac point ( VG\u0000VCNP =\n0V, red line) and the other one at a n-doped Fermi level\ncorresponding to ( VG\u0000VCNP =\u00002:4V, blue line). (b-d) Fit\nparameters Esat,Tlin, andTnsextracted from Eq. (10) for\npower transmission curves as a function of gate voltage.\nof graphene and the substrate, and Erefis the peak \feld\ntransmitted through the ionic gel and the substrate only.\nThe model is given by:19\nT(Einc) =Tnsln[1 +Tlin\nTns(eEinc2=Esat2\u00001)]\nEinc2=Esat2(10)\nwhereEinc,Esat,TlinandTnsare the incident peak elec-\ntric \feld, the saturation electric \feld, the linear power\ntransmission coe\u000ecient and the nonsaturable power\ntransmission coe\u000ecient, respectively.\nIn Fig. 4(a) we present the measured transmitted\npower as a function of the incident \feld along with the\n\ft using Eq. (10) for VG=VCNP (corresponding to\nthe CNP, red line) and VG\u0000VCNP =\u00002:4V(i.e.\nhighly the n-doped case, blue line). The saturation on-\nset of the power transmission, jTj, occurs at a lower\nEinc= 9:7 kV/cm at the CNP relative to what occurs\nfor higher doping, as demonstrated by Fig. 4(b). Tlin\n[Fig. 4(c)] and Tns[Fig. 4(d)] are higher at the CNP\nthan at other Fermi levels, overall expected due to the\nlarger conductivity of the doped graphene. The satura-\ntion behavior for the curves in Fig. 4(a) is explained by a\ndecrease in carrier mobility as the THz excitation redis-\ntributes carriers within the conduction band.19There-\nfore, the decrease in Esatwhen the gate voltage ap-\nproaches the Dirac point can be attributed to an increase\nin the scattering rate when the Fermi level is tuned to the\nCNP, as we shall discuss shortly.\nWe now use the model presented in Section II to ex-\ntract the scattering time as a function of the Fermi energy\nand the incident THz \feld. While the Fermi energy may6\nnot be precisely the same in all 5 layers of the graphene\nsample, for simplicity we use the same value for the \fve\nlayers. Similarly, we assume that the scattering time is\nthe same in all layers. We estimate Effrom gate voltages\nother than the CNP voltage using Eq. 9. To determine\nthe Fermi energy for VG=VCNP, we \frst use the model\nof Section II to extract the \ftting parameters, csr,cCoul\nand\u001cmlin Eq. 7 for the transmission data for the lowest\nTHz \feld amplitude of 2 kV/cm. Taking the lattice and\nelectron temperatures at this lowest \feld to be 300 K, we\nobtaincsr=49 ps\u00001eV\u00001,cCoul = 0 and1\n\u001cml=7.3 ps\u00001.\nThe data as well as the \ft are shown in Fig. 5(a). We now\nextract an e\u000bective Fermi energy at the VG=VCNP of\napproximately 90 meV by extrapolating the \ftting curve.\nThis value is in good agreement with other studies that\nused the same mechanism to tune the Fermi energy.10We\n\fnd from our \ft that the contribution of charged impu-\nrities is negligible and that at this temperature and \feld\nstrength, the contribution of the optical phonons is also\nvery small. Thus, the dominant contribution to scatter-\ning at this lowest \feld appears to arise from short-range\nimpurities, which results in the approximately hyperbolic\nrelationship between \u001candEfseen in the \fgure.27\nWe now turn to the dependence of the scattering time\non the incident \feld for di\u000berent Fermi energies. Rather\nthan using our temperature-dependent model of scatter-\ning, we \frst assume that the electron and phonon tem-\nperatures do not change with the incident THz \feld am-\nplitude and simply extract the net scattering time as a\nfunction of the incident \feld strength to two di\u000berent\nFermi levels. This is plotted in Fig. 5(b). As can be\nseen, for both Fermi energies, the scattering time de-\ncreases dramatically with increasing \feld amplitude. We\nsee that the decrease in the scattering time with \feld is\nmuch larger for the low Fermi energy case. This would\nbe expected if the increased scattering is due to a mech-\nanism that has a rate that is roughly proportional to\nthe Fermi energy and that would increase at high \felds.\nThere are two likely candidates: optical phonon scatter-\ning and carrier-carrier scattering.\nWe \frst consider optical phonon scattering. When the\n\feld is higher, it heats the graphene up considerably; this,\nin turn results in a larger optical phonon population and\na shorter scattering time. To test the viability of the\noptical phonon mechanism, we employed our model that\nincludes the e\u000bects of electron and lattice temperature.\nUsing the impurity scattering rates found at the lowest\n\feld and requiring that the electron and phonon temper-\natures are equal, we were able to obtain good agreement\nwith the experimental data for the transmission by using\nthe temperature as a \ftting parameter. We found that at\nthe highest \felds, the temperature required was approx-\nimately 1550 K. To determine if such temperatures are\nphysically sensible, we compared the experimental result\nfor the energy absorbed per unit area to the increase in\nthe electron and phonon populations per unit area. We\nfound that the required increase in the system energy\nwas more than an order of magnitude greater than theabsorbed THz energy for the 70 kV/cm \feld. Most of\nthe increase in the energy of the system when the tem-\nperature increases to such levels resides in the optical\nphonons, not the electrons. Because there is not uni-\nversal agreement as the values of the optical phonon en-\nergy and deformation potential, we tried adjusting these\nparameters. However, we could not \fnd any values of\nthese parameters that yielded anything approaching en-\nergy balance at all \felds and Fermi energies for one set of\nparameters. We conclude from this, that for this sample\nat least, the decrease in the scattering time with increas-\ning \feld does not arise entirely from an increase in the\noptical phonon population due to an increase in the lat-\ntice temperature.\nAnother possible mechanism for the \feld dependent\nscattering is carrier-carrier scattering. This scattering\nrate increase rapidly as the carrier density increases (as\nobserved). Moreover, it also increases with the energy of\nthe carriers, which will occur when the \feld amplitude\nis high. Modeling carrier-carrier scattering with energy\ndependence would require a sophisticated Monte-Carlo\nor density matrix simulation. Thus, although we cannot\nrule out this mechanism, it is beyond the scope of the\ncurrent work.\nWe now examine the results for our monolayer\ngraphene sample to see how its transmission depends on\nthe Fermi energy and the THz \feld amplitude.\nB. Single-layer sample\nA single-layer graphene sample was prepared by chem-\nical vapor deposition on Cu-foil and transferred onto a\nz-cut quartz substrate with a refractive index n= 1:96.\nThe sheet resistance of this sample as a function of VG\nwas determined by a two terminal measurement shown in\nFig. 6(a) with the CNP indicated again by the resistance\npeak. Electron (hole) doping is induced by changing VG\nabove (below) the CNP voltage VCNP and Eq. 9 is used\nto calculate the Fermi level energy. Fig. 6(b) shows the\nRaman spectra for the single layer graphene sample, with\na 2D/G peak ratio of 1.78 con\frming high quality mono-\nlayer graphene.\nFig. 7(a) shows the normalized THz peak electric \feld\namplitude transmitted through the graphene when the\nFermi-level is swept from -1 to 0.5 V. Again, our data\ndoes not show any phase change in the THz waveform\ntransmitted through the graphene by increasing the in-\ncident THz \feld and so the conductivity change can be\ncompletely quanti\fed by monitoring the peak transmis-\nsion. For the lowest THz peak electric \feld strength of 1.5\nkV/cm, a signi\fcant 20% reduction in the peak transmis-\nsion was observed at V G- VCNP =\u00001 V or Ef=\u0000340\nmeV. As expected, the increased doping leads to higher\nconductivity and therefore absorption of the THz pulse,\nhowever, increasing the incident THz \feld leads to a re-\nduction in the doping-induced absorption as in the 5-\nlayer sample.21,24As in the 5-layer case, a continuous7\nFIG. 5. (a) The impurity scattering time ( \u001cimp) versus Fermi\nenergy, and (b) total scattering time versus incident \feld am-\nplitude at the CNP and at the highest Fermi level of 442 meV\nfor the \fve-layer sample. Note that in (a), the value for Ef\nthe \frst point is determined via a \ft to the curve.\nreduction in \feld-induced transmission modulation oc-\ncurs as the gate voltage approaches the charge neutrality\npoint. This behavior is best seen in Fig. 7(b), showing\nthe peak transmission as a function of the THz peak \feld\nstrength for various applied V G. The THz \feld-induced\ntransparency (the percentage increase in transmission)\nis more pronounced for high doping levels. As with the\n\fve-layer sample, we \fnd that the induced nonlinearity is\nmainly due to an increase in the carrier scattering when\nthe THz \feld is increased.21,24\nWe extract the scattering behavior by using the model\nof Section II. We extract the low-\feld scattering parame-\nters in the same way as was done in the last section taking\nthe temperature to be room temperature. The transmis-\nsion data at the lowest \feld amplitude can be well repro-\nduced by the theory only if we consider both short-range\nneutral and long-range charged impurity scattering in ad-\ndition to optical phonon scattering. We \fnd that the best\n\ft is obtained for csr=30ps\u00001eV\u00001,cCoul=3.63ps\u00001eV\nand and\u001cml=0. The data as well as the \ft are shown in\nFig. 8(a). Note that the increase in the scattering time\nwith Fermi energy indicates that long range charged im-\nFIG. 6. (a) Two terminal measurements of the resistance\nand (b) the Raman data (with the G and 2D peaks) of the\nmonolayer sample.\npurity scattering is dominant at this \feld. The presence\nof long range scattering in this sample may be linked\nwith the direct contact of the graphene with the SiO 2\nsubstrate as opposed to the 5 layer sample, where 4 of\nthe layers are physically removed from the substrate and\nthe e\u000bects of charged impurities on carriers in the up-\nper layers are partially screened by carriers in the lower\nones. Finally, using the extrapolation of the \ft to low\nFermi energy, we extract and Fermi energy of 88 meV\nfor the bias V G= VCNP.\nWe now use the same constant-temperature model of\nthe previous section to determine the net scattering time\nas a function of the incident \feld strength for the \fve dif-\nferent Fermi levels and plot these times in Fig. 8(b). As\nin the \fve-layer case, the scattering time decreases dra-\nmatically with increasing \feld amplitude for all Fermi en-\nergies. However, in contrast to the results of the previous\nsection, we \fnd that the decrease in the scattering time\nwith the \feld is much smaller when V G= VCNP. One\nreason for this di\u000berence is that in the monolayer, the\nlow-\feld scattering time is the shortest when the Fermi\nenergy is the lowest, thus the e\u000bect of any additional scat-\ntering channels is smaller. At the highest \felds, just as\nin the \fve-layer sample, the scattering time is the largest\nwhen VG= VCNP. This seems to indicate again that8\nFIG. 7. (a) The normalized peak transmission as a function of\nthe gate voltage for various THz peak electric \feld strengths\nand (b) the dependence of the peak transmission on the THz\npeak electric \feld for various doping levels for the monolayer\nsample.\nthe \feld-dependent scattering rate is larger when the car-\nrier density (or perhaps average carrier energy) is larger.\nHowever, the trend is not the same for the intermediate\nFermi energies, where at the highest \feld, the scattering\nrate is larger when the Fermi energy is smaller (with the\nexception of the CNP result). It appears that the result\nat the CNP is somewhat of an anomaly in this sample.\nAlthough, the Fermi energy dependence of the \feld-\ndependent scattering component is not as clear as in\nthe \fve-layer sample, we again examined where optical-\nphonon scattering is a possible mechanism. We modeled\nthe optical phonon scattering by taking the temperature\nto be a \ftting parameter. Again, although we could re-\nproduce the observed transmission curves, the tempera-\ntures required were very large. The temperatures at the\nhighest \feld ranged from 1300 K for the highest Fermi en-\nergy to almost 3000 K at the CNP. Again, such huge tem-\nperatures require that the energy deposited in the system\nis a factor of 3 to 10 larger than the energy that is experi-\nmentally found to be absorbed by the graphene. As in the\n\fve-layer case, we tried adjusting the phonon energy and\ndeformation potential, but for all combinations that we\nFIG. 8. (a) The impurity scattering time ( \u001cimp) versus Fermi\nenergy at the lowest incident THz \feld of 1.5 kV/cm, and (b)\nthe THz \feld dependence of (a) the carrier scattering rate.\nNote that in (a), the value for Efthe \frst point is determined\nvia a \ft to the curve.\ntried we could not achieve anything approaching energy\nbalance at all \felds and Fermi energies. We thus conclude\nthat for this sample as well, optical phonons cannot be\nthe major contributor to the increase in the scattering\nrate with increasing \feld. In this sample, however, due\nto the more complex dependence of the \feld-dependent\nscattering contribution on the Fermi energy, it is not as\nclear that carrier-carrier scattering is a likely candidate.\nIV. CONCLUSION\nWe have studied nonlinear THz light-matter interac-\ntion for both monolayer graphene on quartz and a stack\nof 5 monolayers of graphene on sapphire under the con-\ntrolled Fermi level conditions. The two samples exhibited\nopposite dependencies on the Fermi energy for low THz\n\felds. The monolayer sample scattering rate was domi-\nnated by charged impurities, while the \fve layer sample\nscattering was neutral impurity dominated. This di\u000ber-\nence is likely explained as arising from di\u000berent prepara-\ntion recipes as well as di\u000berences in the substrates.\nWe also observed a strong increase in the THz trans-9\nmission with increasing \feld that occurs for THz peak\nelectric \felds as low as 8 kV/cm. This increase appears\nto result from a decrease in the carrier scattering times\nwith increasing \feld. Our analyses indicate that opti-\ncal phonon absorption is not likely the main source of\nthis \feld-induced increase in the scattering rate. The\nsource of this increase remains uncertain. For the \fve\nlayer sample, the most promising mechanism is perhaps\ncarrier-carrier scattering, but for the monolayer sample\nthe mechanism is even more unclear.\nThis work points to the importance of future modelling\nof full carrier and lattice dynamics in graphene excited by\nintense THz pulses to determine the main source of the\nstrongly nonlinear response. The interaction of high \feldTHz pulses with graphene is signi\fcantly dependent on\nFermi energy. Moreover, given the variability in process-\ning of graphene \flms, we show that it is entirely possible\nthat di\u000berent impurity scattering mechanisms dominate\nin di\u000berent samples, even though the \feld dependence is\nsimilar for a \fxed Fermi energy. Thus any comprehen-\nsive study must be under controlled gated conditions. We\nanticipate that our \fndings will be important for future\ndesigns of graphene devices operating at THz frequencies,\nin\ruencing performance due to nonlinearities arising at\nthe kV/cm operating \felds.\nThe authors gratefully acknowledge funding support\nfrom NSERC, FRQNT and CFI.\n1K. S. A. Novoselov, A. K. Geim, S. V. B. Morozov,\nD. Jiang, M. I. V. Grigorieva, S. V. Dubonos, A. A. Firsov,\nNature 438, 197-200 (2006).\n2T. Otsuji, S. A. B. Tombet, A. Satou, H. Fukidome,\nM. Suemitsu, E. Sano, V. Popov, M. Ryzhii, V. Ryzhii,\nJ. Phys. D: Appl. Phys. 45, 303001 (2012).\n3C. J. Docherty, C. T. Lin, H. J. Joyce, R. J. Nicholas,\nL. M. Herz, L. J. Li, M. B. Johnston, Nat. Commun. 3,\n1228 (2012).\n4F. Wang, Y. Zhang, C. Tian, C. Girit, A. Zettl, M. Crom-\nmie, Y. R. Shen, Sience 320, 206-209 (2008).\n5Y. M. Lin, K. A. Jenkins, A. Valdes-Garcia, J. P. Small,\nD. B. Farmer, P. Avouris, Nano Lett. 9, 422-426 (2009).\n6J. M. Dawlaty, S. Shivaraman, J. Strait, P. George,\nM. V. S. Chandrashekhar, F. Rana, M. G. Spencer, D. Vek-\nsler, Y. Chen, App. Phys. Lett. 93, 131905 (2008).\n7J. D. Buron, D. H. Petersen, P. B\u001cggild, D. G. Cooke,\nM. Hilke, J. Sun, E. Whiteway, P. F. Nielsen, O. Hansen,\nA. Yurgens, Nano Lett. 12, 5074-5081 (2012).\n8J. L. Tomaino, A. D. Jameson, J. W. Kevek, M. J. Paul,\nA. M. van der Zande, R. A. Barton, P. L. McEuen,\nE. D. Minot, Y. S. Lee, Opt. Expr. 19, 141-146 (2011).\n9A. J. Frenzel, C. H. Lui, Y. C. Shin, J. Kong, N. Gedik,\nPhys. Rev. Lett. 113, 056602 (2014).\n10S. F. Shi, T. T. Tang, B. Zeng, L. Ju, Q. Zhou, A. Zettl,\nF. Wang, Nano Lett. 14, 1578-1582 (2014).\n11J. Horng, C. F. Chen, B. Geng, C. Girit, Y. Zhang, Z. Hao,\nH. A. Bechtel, M. Martin, A. Zettl, M. F. Crommie, Y. Ron\nShen, F. Wang, Phys. Rev.B 83, 165113 (2011).\n12K. F. Mak, M. Y. Sfeir, Y. Wu, C. H. Lui, J. A. Misewich,\nT. F. Heinz, Phys. Rev. Lett. 101, 196405 (2008).\n13I. Al-Naib, J. E. Sipe, M. M. Dignam, Phys. Rev. B 90,\n245423 (2014).\n14H. Hirori, A. Doi, F. Blanchard, K. Tanaka, Appl. Phys.\nLett. 98, 091106 (2011).\n15A. Pashkin, A. Sell, T. Kampfrath, R. Huber, New J. Phys.\n15, 065003 (2013).\n16P. Bowlan, E. Martinez-Moreno, K. Reimann, T. Elsaesser,\nM. Woerner, Phys. Rev. B. 89, 041408 (2014).\n17J. A. Crosse, X. Xu, M. S. Sherwin, R. B. Liu, Nat. Com-\nmun. 5, 4854 (2014).\n18J. C. Johannsen, S. Ulstrup, F. Cilento, A. Crepaldi,\nM. Zacchigna, C. Cacho, I. .C. E. Turcu, E. Springate,\nF. Fromm, C. Raidel, T. Seyller, F. Parmigiani, M. Gri-oni, P. Ho\u000bmann , Phys. Rev. Lett. 111, 027403 (2013).\n19H. Y. Hwang, N. C. Brandt, H. Farhat, A. L. Hsu, J. Kong,\nK. A. Nelson, J. Phys. Chem. B 117, 15819-15824 (2013).\n20H. A. Hafez, I. Al-Naib, K. Oguri, Y. Sekine, M. M. Dig-\nnam, A. Ibrahim, D. G. Cooke, S. Tanaka, F. Komori,\nH. Hibino, T. Ozaki, AIP Advances 4, 117118 (2014).\n21M. J. Paul, Y. C. Chang, Z. J. Thompson, A. Stickel,\nJ. Wardini, H. Choi, E. D. Minot, B. Hou, J. A. Nees,\nT. B. Norris, Y. S. Lee, New J. Phys. 15, 085019 (2013).\n22M. J. Paul, B. Lee, J. L. Wardini, Z. J. Thompson,\nA. D. Stickel, A. Mousavian, H. Choi, E. D. Minot,\nY. S. Lee, Appl. Phys. Lett. 105, 221107 (2014).\n23Z. Mics, K. J. Tielrooij, K. Parvez, S. A. Jensen, I. Ivanov,\nX. Feng, K. M ullen, M. Bonn, D. Turchinovich, Nat. Com-\nmun. 6:7655 doi: 10.1038/ncomms8655 (2015).\n24H. A. Hafez, I. Al-Naib, M. M. Dignam, Y. Sekine,\nK. Oguri, F. Blanchard, D. G. Cooke, S. Tanaka, F. Ko-\nmori, H. Hibino, T. Ozaki, Phys. Rev. B 91, 035422 (2015).\n25W. S. Bao, S. Y. Liu, X. L. Lei, C. Wang, J. Phys. Condens.\nMatter. 21, 305302 (2009).\n26I. Al-Naib, M. Poschmann, M. M. Dignam, Phys. Rev. B\n91, 205407 (2015).\n27E. H. Hwang, S. D. Sarma, Phys. Rev. B 77, 195412 (2008).\n28S. Das Sarma, S. Adam, E. H. Hwang, E. Rossi, Rev. Mod.\nPhys. 83, 407 (2011).\n29T. Fang, A. Konar, H. Xing, D. Jena, Phys. Rev. B 84,\n125450 (2011).\n30M. Breusing, C. Ropers, T. Elsaesser, Phys. Rev. Lett.\n102, 086809 (2009).\n31C. H. Lui, K. F. Mak, J. Shan, T. F. Heinz, Phys. Rev.\nLett. 105, 127404 (2010).\n32N. Sule, I. Knezevic, J. Appl. Phys. 112, 053702 (2012).\n33J. H. Strait, H. Wang, S. Shivaraman, V. Shields,\nM. Spencer, F. Rana, Nano Lett. 11, 4902-4906, (2011).\n34E. H. Hwang, S. D. Sarma, Phys. Rev. B 79, 165404 (2009).\n35C. F. Chen, C. H. Park, B. W. Boudouris, J. Horng,\nB. Geng, C. Girit, A. Zettl, M. F. Crommie, R. A. Segal-\nman, S. G. Louie, F. Wang, Nature 471, 617-620 (2011).\n36A. Das, S. Pisana, B. Chakraborty, S. Piscanec, S. K. Saha,\nU. V. Waghmare, K. S. Novoselov, H. R. Krishnamurthy,\nA. K. Geim, A. C. Ferrari, A. K. Sood, Nature Nanotech.\n3, 210-215 (2008).\n37J. Martin, N. Akerman, G. Ulbricht, T. Lohmann,\nJ. H. Smet, K. Von Klitzing, A. Yacoby, Nat. Phys. 4,10\n144-148, (2008).\n38S. K. Lee, H. Y. Jang, S. Jang, E. Choi, B. H. Hong, J. Lee,\nS. Park, J. H. Ahn, Nano Lett. 12, 3472-3476 (2012).\n39J. Ye, M. F. Craciun, M. Koshino, S. Russo, S. Inoue,\nH. Yuan, H. Shimotani, A. F. Morpurgo, Y. Iwasa, Proc.\nNatl. Acad. Sci. 108, 13002-13006 (2011).40D. Graf, F. Molitor, K. Ensslin, C. Stampfer, A. Jungen,\nC. Hierold, L. Wirtz, Nano Lett. 7, 238-242 (2007).\n41K. Aniruddha, F. Tian, J. Debdeep, Phys. Rev. B 82,\n115452 (2010).\n42S. Berciaud, M. Y. Han, K. F. Mak, L. E. Brus, P. Kim,\nT. F. Heinz, Phys. Rev. Lett. 104, 227401 (2010)." }, { "title": "0912.3700v1.Magnetic_impurity_transition_in_a__d_s__wave_superconductor.pdf", "content": "arXiv:0912.3700v1 [cond-mat.supr-con] 18 Dec 2009physica statussolidi, 10 November2018\nMagnetic impurity transition in a\nd+s- wave superconductor\nL. S.Borkowski*\nQuantum Physics Division, Facultyof Physics,A.Mickiewic z University, Umultowska 85, 61-614 Poznan, Poland\nPACS74.20.Rp,74.25.Dw,74.62.Dh,74.72.-h\n∗Corresponding author: e-mail lech.s.borkowski@gmail.com\nWe considerthe superconductingstate of d+ssymme-\ntry with finite concentration of Anderson impurities in\nthe limit ∆s/∆d≪1. The model consists of a BCS-\nlike term in the Hamiltonian and the Andersonimpurity\ntreatedintheself-consistentlarge- Nmeanfieldapprox-\nimation.Increasing impurity concentration or lowering the ratio\n∆s/∆ddrives the system through a transition from a\nstate with two sharppeaksat low energiesandexponen-\ntially small density of states at the Fermi level to one\nwithN(0)≃(∆s/∆d)2. This transition is discontinu-\nous if the energyof the impurity resonanceis the small-\nest energyscale inthe problem.\nCopyrightlinewillbe provided by the publisher\n1 Introduction The order parameter symmetry of\nhigh-Tccompounds is believed to be predominantly of d-\nwave type. However a subdominant s-wave componentof\nthe order parameter may also occur, especially in materi-\nals with orthorombic distortions. One of such compounds\nis YBCO, exhibiting strong structural distortion and a\nsubstantial anisotropy in the London penetration depth in\nthea-bplane.[1] Raman scattering[2] provided evidence\nfor a5%admixture of s-wave component while thermal\nconductivity measurements in rotating magnetic field[3]\nplaced an upper limit of 10%. Angle-resolved photoemis-\nsion spectroscopy[4] on monocrystalline YBCO gave the\nratio of 1.5 for gap amplitudes in the aandbdirections in\ntheCuO2. Measurementsof a Josephson current between\nmonocrystalline YBa 2Cu3O7ands-wave Nb showed that\nthe obtained anisotropy could be explained by a 83%\nd-wave with a 17%s-wave component.[5] In another ex-\nperiment on YBCO/Nb junction rings the stodgap ratio\ninoptimallydopedYBCO wasestimatedtobe 0.1.[6]\nInelasticneutronscatteringonmonocrystallineandun-\ntwinned samples of YBCO lead to magnetic susceptibili-\ntieswithintensitiesandlineshapesbreakingthetetragon al\nsymmetry.[7,8,9,10] It was shown that these data may be\ninterpretedwithinananisotropicbandmodelwithanorder\nparameterofmixed dandssymmetry.[11]\nSuperconductingstates with mixed symmetry are also\nconsidered in other classes of compounds, e.g. in the re-\ncentlydiscoveredferropnictides.[12]The effects of dilute concentrations of magnetic and\nnonmagnetic point defects on a BCS superconductor of\npuresord-wavesymmetrywereintensivelystudiedinthe\npast andarewell known.Ina d-wave superconductorwith\nlines of order parameter nodes any amount of disorder in-\nducesanonzerodensityofstatesattheFermilevel.Inan s-\nwavesystemonlymagneticimpuritieschangetheresponse\nof the superconducting state. For sufficiently strong cou-\nplingbetweentheimpurityandtheconductionbandbound\nstatesmayappearin theenergygap.[13]\nIn an earlier work on nonmagneticimpuritiesin a d+\ns-wave superconductor it was shown that in the unitary\nlimit a nonzero density of states (DOS) at the Fermi level\nappears above certain critical impurity concentration, de -\npendingonthesizeofan s-wavecomponent.[14]However\nthelow-energyDOSismostlyfeaturelesssincethe s-wave\ncomponentpreventsabuildupofstatesduetononmagnetic\nscattering. In contrast the presence of magnetic impuritie s\nin asuperconductorwithan s-wavecomponentmayresult\ninsharppeaksinthelow-energyDOSforsmallconcentra-\ntion of defects, provided the energy scale associated with\ntheimpurityresonanceissmall.\n2 Model and Results We consider the order param-\neterofd+ssymmetryona cylindricalFermisurface.\n∆s+eiθ∆d(ˆk), (1)\nCopyrightlinewillbe provided by the publisher2 L. S.Borkowski: Magnetic impuritytransition ina d+s-wave superconductor\nwhere∆sand∆dareamplitudesof s-andd-wavecompo-\nnentrespectively.We assume θ= 0and∆s≪∆d.\nThesuperconductoristreatedinaBCSapproximation.\nThe magnetic scatterer is modelled as an Anderson impu-\nritytreatedwithintheslavebosonmeanfieldapproach.[15]\nThe low energy physics is dominated by the presence of\nstronglyscatteringimpurityresonance.Theself-consist ent\nself-energyequationsdescribingthe interplaybetweenth e\nsuperconductingandmagneticdegreesoffreedomhavethe\nfollowingform,\n/tildewideω=ω+nN\n2πN0Γ¯ω\n(−¯ω2+ǫ2\nf), (2)\n/tildewide∆=∆s+nN\n2πN0Γ¯∆\n(−¯ω2+ǫ2\nf),(3)\n¯ω=ω+Γ/angbracketleft/tildewideω\n/parenleftBig\n/tildewide∆2(k)−/tildewideω2/parenrightBig1/2/angbracketright,(4)\n¯∆=Γ/angbracketleft/tildewide∆\n/parenleftBig\n/tildewide∆2(k)−/tildewideω2/parenrightBig1/2/angbracketright, (5)\nwhere/tildewideω(¯ω),/tildewide∆(¯∆)is the renormalized frequency and or-\nder parameter of conduction electrons (impurity) respec-\ntively.\nWe assume ∆d/D= 0.01,where2Disthebandwidth\nof the conductionelectron band. In the equationsabove Γ\nis the hybridization energy between the impurity and the\nconduction band and ǫfis the resonant level energy. Here\nwe assume constant density of states in the normal state\nN0= 1/2Dand do the calculations for a nondegenerate\nimpurity, N= 2. Brackets denote average over the Fermi\nsurface.\nInitial results for the density of states of this system\nwere presented in an earlier paper.[16] For small impurity\nconcentration n,such that ∆sand∆dare notsignificantly\naffected, there are two peaks located symmetrically near\nthe gap center, provided ǫf≪Γ≪∆s. For larger n\nthe two peaksmerge into one peak centrally located at the\nFermienergy.\nFig. 1 shows the dependence of the density of states\nN(0)atEFas a function of impurity concentration.\nAt small n,N(0)is exponentially small, N(0)/N0∼\n(∆s/∆d)exp(−α(∆s/∆d)2/n), whereαis a numerical\nfactor. The critical concentration n0for the discontinuous\ntransition to N(0)/N0≃∆s/∆dis a quadratic func-\ntion of∆s/∆d. Forn > n 0,N(0)approaches DOS of a\nd-wavesuperconductorinboththemagnitudeanditsfunc-\ntional dependence on n. These relations are valid when\nǫf≪Γ,∆s.\nQualitatively similar scaling of N(0)as a function of\n∆s/∆din presence of nonmagnetic impurities in the uni-\ntarylimitwasobtainedinref.[14].Thedifferencebetween\nmagnetic and nonmagnetic defetcs in a superconductor-8-6-4-2 0\n-6 -5 -4 -3 -2log10(N(0)/N0)\nlog10n∆s+∆d(k)εf=10-4\n∆s/∆d=0.001\n0.01\n0.020.050.1\nFigure 1 Logarithm of the density of states at the Fermi\nlevel as a function of impurity concentration for several\nvaluesoftheratio ∆s/∆d0.Theresonantimpuritylevel ǫf\nisclose totheFermilevel. Γ/Disfixedat 0.001.\n-3-2-1\n-6-5-4-3-2-1log10(ωpeak)\nlog10n∆s+∆d(k)∆s/∆d=0.01\nεf=010-42x10-45x10-410-3\nFigure 2 Position of the peak in the conduction electron\nDOS as a functionof impurityconcentrationfor d+ssu-\nperconductor for several values of the resonant level en-\nergy. Energy is scaled by half of the conduction electron\nbandwidth DandΓ/D= 0.001. The suppression of ∆s\nand∆dwasnottakenintoaccount.\nwith∆s/∆d≪1is the character of the transition from\nthe exponentially small N(0)to finiteN(0)in the strong\nscattering limit. In contrast to nonmagneticimpurities, t he\nCopyrightlinewillbe provided by the publisherpss header willbeprovided by thepublisher 3\ntransition caused by resonantly scattering magnetic impu-\nrities is discontinuous. and there are two sharp peaks of\nN(ω)on both sides of the transition.[16] The position of\npeaks as a function of impurity concentration for differ-\nent values of ǫfis shown in Fig 2. These peaks strongly\nalter the low-energyand low-frequencyresponse and may\nbe detected in thermodynamicor transport measurements.\nThe increase of ∆sshifts the resonances towards EFand\nmakesthemmorenarrow.\nThed-wave component of the superconducting order\nparameter is more sensitive to pair breaking by magnetic\nimpuritiesin the limit T≪TK, whereTK=/radicalBig\nΓ2+ǫ2\nf,\nthanthes-wavepart.Dependingontherelativesize of ∆d\nand∆sthere maybeanotherimpuritytransitionforlarger\nn, when the order parameter nodes disappear due to van-\nishing of the d-wave component and a full gap opens up.\nTheimpuritypeakisthensplitand N(0)fallstozeroagain.\nHower a detailed description of this possibility requires a\ncareful analysis involving four energy scales: ∆d,∆s,Γ,\nandǫf.\nCan such transition be observed experimentally? The\ntransition may be tuned either by varying impurity con-\ncentrationor the ratio ∆s/∆d. The density of states of the\nnormal state at EF,N0, also has an effect. It appears in\nequations2 and3. One shouldbear in mind,however,that\nboth the superconducting transition temperature and the\nimpurityresonanceenergyscale TKareexponentiallysen-\nsitivetochangesof N0.Experimentsconductedatverylow\ntemperatures may give different signatures of low-energy\nbehavior depending on the location of the system on the\nphasediagramrelativetotheimpuritycriticalpoint.\nInad+s-superconductorinthelimitofhighconcentra-\ntionofpointdefectsthe d-wavecomponentvanisheswhile\nthes-wavepartoftheorderparameterremainsunaffected.\nThe situation is different in presence of magnetic scatter-\ners.If the s-wave componentissmall andthe energyscale\nof the resonance due to impurity scattering is at most of\ntheorder ∆s, increasingimpurityconcentrationmaydrive\n∆s→0while∆dwill be reducedandfinite. Thisfollows\nfrom the fact that the largest pair breaking occurs when\nthe energy scales of the impurity resonance and the order\nparameter are comparable. If ∆d≫TK, the rate of sup-\npressionof ∆dwithincreasing nissmall.\nThed+ssuperconductorwith a subdominant s-wave\ncomponent doped with magnetic impurities has a differ-\nent phase diagram in the T-nplane compared to the same\nsuperconductor with nonmagnetic defects. If the impurity\nresonance scale TKis comparable to ∆sand∆s≪∆d,\nthes-wavecomponentmayvanishat T≃TK.\n3 Conclusions TheT= 0impurity transition dis-\ncussedinthisworkmaybedetectedinthelowtemperature\nlimit. While the slave boson mean field formalism used in\nthispapercannotbeappliedattemperaturesexceeding TK,\nwe may also qualitatively describe the expected behavior\nof the system as a function of temperature. At finite butsmallimpurityconcentrationtheremaybeevenfourphase\ntransitionsasafunctionoftemperature:normalto d-wave,\nd-wave to d+s-wave,d+stod-wave, and d-wave to\nd+s-wave.Dueto thedifferenceofmagnitudesthesmall\n∆smay be driven to zero faster with increasing nthan\n∆d. Nonmonotonic behavior around T∼TKis a conse-\nquenceofstrongscatteringby theresonancestate forming\non the impurity site. As T→0, the impurity scattering\nbecomes weaker and the s-wave component may appear\nagain.Whetherthisparticularscenarioisrealized,depen ds\non the relative size of energyscales: ∆s/∆d,∆s/TK, and\nǫf/Γ.\nAcknowledgements Some of the computations were per-\nformed in the Computer Center of the Tri-city Academic Com-\nputer Network inGdansk.\nReferences\n[1] D. N. Basov, R. Liang, D. A. Bonn, W. A. Hardy,\nB. Dabrowski, M. Quijada, D. B. Tanner, J. P. Rice,\nD. M. Ginsberg, and T. Timusk, Phys. Rev. Lett. 74, 598\n(1995).\n[2] M. F. Limonov, A. I. Rykov, S. Tajima and A. Yamanaka,\nPhys.Rev. Lett. 80, 825 (1998)\n[3] H.Aubin, K. Behnia, M. Ribault, R.Gagnon, andL. Taille-\nfer,Phys.Rev. Lett. 782624.\n[4] D.H.Lu,D.L.Feng, N.P.Armitage,K.M.Shen, A.Dam-\nascelli, C. Kim, F. Ronning, Z.-X. Shen, D. A. Bonn,\nR. Liang, W. N. Hardy, A. I. Rykov, and S. Tajima, Phys.\nRev. Lett. 86, 4370 (2001).\n[5] H. J. H. Smilde, A. A. Golubov, Ariando, G. Rijnders,\nJ. M. Dekkers, S. Harkema, D. H. A. Blank, H. Rogalla,\nandH. Hilgenkamp, Phys. Rev. Lett. 95, 157001 (2005).\n[6] J. R. Kirtley, C. C. Tsuei, Ariando, C. J. M. Verwijs,\nS.Harkema, and H.Hilgenkamp, Nat.Phys. 2, 198 (2006).\n[7] H. A. Mook, P. C. Dai, F. Dogan, and R. D. Hunt, Nature\n404, 729 (2000).\n[8] C. Stock, W. J. L. Buyers, R. Liang, D. Peets, Z. Tun,\nD. Bonn, W. N. Hardy, and R. J. Birgeneau, Phys. Rev. B\n69, 014502 (2004).\n[9] C. Stock, W. J. L. Buyers, R. A. Cowley, P. S. Clegg,\nR. Coldea, C. D. Frost, R. Liang, D. Peets, D. Bonn,\nW. N. Hardy, and R. J. Birgeneau, Phys. Rev. B 71, 024522\n(2005).\n[10] V.Hinkov,S.Pailhes,P.Bourges,Y.Sidis,A.Ivanov,A .Ku-\nlakov, C. T. Lin, D. P. Chen, C. Bernhard, and B. Keimer,\nNature430, 650 (2004).\n[11] A.P.Schnyder, D.Manske, C.Mudry, andM.Sigrist,Phys .\nRev. B73, 224523 (2006).\n[12] V. Mishra, G. Boyd, S. Graser, T. Maier, P. J. Hirschfeld ,\nand D. J. Scalapino, Phys. Rev. B 79, 094512 (2009), and\nreferences therein.\n[13] L. S. Borkowski and P. J. Hirschfeld, J. Low Temp. Phys.\n96, 185 (1994), and references therein.\n[14] H.Kim,and E.J.Nicol, Phys.Rev. B 5213576 (1995)\n[15] L.S.Borkowski, Phys. Rev. B 78020507 (2008).\n[16] L.S.Borkowski, J.Phys.:Conf.Series 150,052023 (2009).\nCopyrightlinewillbe provided by the publisher" }, { "title": "1003.2452v1.Highlights_from_Fermi_GRB_observations.pdf", "content": "arXiv:1003.2452v1 [astro-ph.HE] 12 Mar 2010Highlights from Fermi GRB observations\nJonathan Granot\non behalf of the Fermi LAT and GBM collaborations\nCentre for Astrophysics Research, University of Hertfords hire,\nCollege Lane, Hatfield AL10 9AB, UK\nAbstract\nThe Fermi Gamma-Ray Space Telescope has more than doubled\nthe number of Gamma-Ray Bursts (GRBs) detected above 100 MeV\nwithin its first year of operation. Thanks to the very wide energy\nrange covered by Fermi’s Gamma-ray Burst Monitor (GBM; 8 keV to\n40 MeV) and Large Area Telescope (LAT; 25 MeV to >300 GeV)\nit has measured the prompt GRB emission spectrum over an unprece -\ndentedlylargeenergyrange(from ∼8 keVto ∼30 GeV). HereIbriefly\noutline some highlights from Fermi GRB observations during its first\n∼1.5 yr ofoperation, focusing onthe promptemission phase. Interes t-\ning new observations are discussed along with some of their possible\nimplications, including: (i) What can we learn from the Fermi-LAT\nGRB detection rate, (ii) A limit on the variation of the speed of light\nwith photon energy(for the firsttime beyond the Planckscale fora lin-\near energy dependence from direct time of arrival measurements ), (iii)\nLower-limits on the bulk Lorentz factor of the GRB outflow (of ∼1000\nfor the brightest Fermi LAT GRBs), (iv) The detection (or in other\ncases, lackthereof) ofa distinct spectralcomponent athigh (an d some-\ntimes also at low) energies, and possible implications for the prompt\nGRB emission mechanism, (v) The later onset (and longer duration)\nof the high-energy emission ( >100 MeV), compared to the low-energy\n(<∼1 MeV) emission, that is seen in most Fermi-LAT GRBs.\n1 Pre-Fermi high-energy GRB observations\nThe Energetic Gamma-Ray Experiment Telescope (EGRET) on-b oard the\nCompton Gamma-Ray Observatory (CGRO; 1991 −2000) was the first to\ndetect high-energy emission from GRBs. EGRET detected only five GRBs\nwith its Spark Chambers (20MeV to 30GeV) and a few GRBs with it s\nTotal Absorption Shower Counter (TASC; 1 −200 MeV). Nevertheless,\n1these events already showed diversity. The most prominent e xamples are\nGRB 940217, with high-energy emission lasting up to ∼1.5 hr after the\nGRB including an 18 GeV photon after ∼1.3 hr, [1] and GRB 941017 which\nhad adistinct high-energy spectral component [2] detected upto∼200 MeV\nwithνFν∝ν. This high-energy spectral component had ∼3 times more\nenergy and lasted longer ( ∼200 s) than the low-energy (hard X-ray to soft\ngamma-ray) spectral component (which lasted several tens o f seconds), and\nmay be naturally explained as inverse-Compton emission fro m the forward-\nreverse shock system that is formed as the ultra-relativist ic GRB outflow\nis decelerated by the external medium [3, 4]. Nevertheless, better data are\nneeded in order to determine the origin of such high-energy s pectral compo-\nnents more conclusively. The Italian experiment Astro-riv elatore Gamma a\nImmagini LEggero (AGILE; launched in 2007) has detected GRB 080514B\nat energies up to ∼300 MeV, and the high-energy emission lasted longer\n(>13 s) than the low-energy emission ( ∼7 s) [5]. Below are some highlights\nof Fermi GRB observations so far and what they have taught us.\n2 LAT GRB detection rate: what can it teach us?\nDuring its first 1 .5 yr of routine operation, from Aug. 2008 to Jan. 2010, the\nLAT has detected 14 GRBs, corresponding to a detection rate o f∼9.3 yr−1.\nTable 1 summarizes their main properties. While at least 13 o f the 14 LAT\nGRBs had ≥10 photons above 100 MeV, 4 were particularly bright in the\nLAT with ≥1 photon above 10 GeV, ≥10 photons above 1 GeV, and\n≥100 photons above 100 MeV. This corresponds to a bright LAT GR B\n(as defined above) detection rate of ∼2.7 GRB/yr (with a rather large\nuncertainty due to the small number statistics). There were also 11 GRBs\nwith≥1 photon above 1 GeV, corresponding to ∼7.3 GRB/yr. These\ndetection rates are compatible with pre-launch expectatio ns [6] based on a\nsample of bright BATSE GRBs for which the fit to a Band spectrum over\nthe BATSE energy range (20 keV to 2 MeV) was extrapolated into the LAT\nenergy range (see Fig. 1). The agreement is slightly better w hen excluding\ncases with a rising νFνspectrum at high energies (i.e. a high-energy photon\nindexβ >−2).1This suggests that, on average, there is no significant\nexcess or deficit of high-energy emission in the LAT energy ra nge relative to\nsuch an extrapolation from lower energies. As described in §5, however, in\nindividual cases we do have evidence for such an excess. The o bserved LAT\n1Such a hard high-energy photon index may be an artifact of the limited energy range\nof the fit to BATSE data, or may be affected by poor photon statis tics at>∼1 MeV.Figure1: LAT GRBdetection rates(colorellipses)superposedonto pofpre-launch\nexpected rates based on the extrapolation of a Band spectrum fit from the BATSE\nenergy range [6]. The ellipses’ inner color indicates the minimal photon energy\n(green, yellow and cyan correspond to 0.1, 1 and 10 GeV, respectiv ely), while their\nhight indicates the uncertainty ( ±N1/2/1.5yr)on the correspondingLAT detection\nrate (N/1.5yr) due to the small number ( N) of detected GRBs.\nGRB detection rate implies that, on average, only about ∼10−20% of the\nenergy that is radiated during the prompt GRB emission phase is channeled\ninto the LAT energy range, suggesting that in most GRBs the hi gh-energy\nradiative output does not significantly affect the total energ y budget. Short\nGRBs, however, appear to be different in this respect (see §7 and Fig. 3).\n3 Limits on Lorentz Invariance Violation\nSome quantum gravity models allow violation of Lorentz inva riance, and in\nparticular allow the photon propagation speed, vph, to depend on its energy,\nEph:vph(Eph)∝ne}ationslash=c, where c≡lim\nEph→0vph(Eph). The Lorentz invariance\nviolating (LIV) part in the dependence of the photon momentu m,pph, onlong number of HE emission extra highest\nGRB θLAT or events above starts lasts spec. energy z\nshort 0.1GeV 1GeV later longer comp. (GeV)\n080825C ∼60◦long ∼10 0 ? yes no ∼0.6 —–\n080916C 49◦long 145 14 yes yes ?∼13∼4.35\n081024B 21◦short ∼10 2 yes yes ? ∼3—–\n081215A ∼86◦long —– — ? ? — —– —–\n090217 ∼34◦long ∼10 0 no no no ∼1—–\n090323 ∼55◦long ∼20 >0? yes ? ? 3.57\n090328 ∼64◦long ∼20 >0? yes ? ?0.736\n090510 ∼14◦short >150 >20yes yes yes ∼310.903\n090626 ∼15◦long ∼20 >0? yes ? ? —–\n090902B 51◦long >200 >30yes yes yes ∼331.822\n090926 ∼52◦long >150 >50yes yes yes ∼202.1062\n091003A ∼13◦long ∼20 >0? ? ? ?0.8969\n091031 ∼22◦long ∼20 >0? ? ?∼1.2 —–\n100116A ∼29◦long ∼10 3 ? ? ?∼2.2 —–\nTable 1: Summary of the 14 GRBs detected by the LAT between August 2008 and\nJanuary 2010 – its first 1.5 years of routine operation follow ing Fermi’s launch on 11 June\n2008;θLATis the angle from the LAT boresight at the time of the GBM GRB tr igger.\nits energy, Eph, can be expressed as a power series,\np2\nphc2\nE2\nph−1 =∞/summationdisplay\nk=1sk/parenleftBigg\nEph\nMQG,kc2/parenrightBiggk\n, (1)\nintheratioof Ephandatypicalenergyscale MQG,kc2forthekthorder,which\nis expected to be up to the order of the Planck scale, MPlanck= (¯hc/G)1/2≈\n1.22×1019GeV/c2, wheresk∈ {−1,0,1}. Since we observe photons of\nenergy well below the Planck scale, the dominant LIV term is a ssociated\nwith the lowest order non-zero term in the sum, of order n= min{k|sk∝ne}ationslash= 0},\nwhich is usually assumed to be either linear ( n= 1) or quadratic ( n= 2).\nThe photon propagation speed is given by the corresponding g roup velocity,\nvph=∂Eph\n∂pph≈c/bracketleftBigg\n1−snn+1\n2/parenleftBigg\nEph\nMQG,nc2/parenrightBiggn/bracketrightBigg\n. (2)\nNote that sn= 1 correspondsto thesub-luminal case ( vph< cand apositive\ntime delay), while sn=−1 corresponds to the super-luminal case ( vph> c\nand a negative time delay). Taking into account cosmologica l effects [9], this\ninduces a time delay (or lag) in the arrival of a high-energy p hoton of energy\nEh, compared to a low-energy photon of energy El(emitted simultaneously\nat the same location), of\n∆t=sn(1+n)\n2H0(En\nh−En\nl)\n(MQG,nc2)n/integraldisplayz\n0(1+z′)n\n/radicalbig\nΩm(1+z′)3+ΩΛdz′.(3)Here we concentrate on our results for a linear energy depend ence (n= 1).\nWe have applied this formula to the highest energy photon det ected in\nGRB 080916C, with an energy of Eh= 13.22+0.70\n−1.54GeV, which arrived at t=\n16.54 s after the GRB trigger (i.e. the onset of the El∼0.1 MeV emission).\nSinceitis hardtoassociate thehighestenergy photonwitha particular spike\ninthelow-energy lightcurve, wehave madetheconservative assumptionthat\nit was emitted anytime after the GRB trigger, i.e. ∆ t≤t, in order to obtain\na limit for the sub-luminal case ( sn= 1):MQG,1>0.1MPlanck. This was\nthe strictest limit of its kind [7], at that time.\nHowever, the next very bright LAT GRB, 090510, was short and h ad\nvery narrow sharp spikes in its light curve (see Fig 2), thus e nabling us to\ndo even better [8]. Our main results for GRB 090510 are summar ized in\nTable 2. The first 4 limits are based on a similar method as desc ribed above\nfor GRB 080916C, using the highest energy photon, Eh= 30.53+5.79\n−2.56GeV,\nand assuming that its emission time thwas after the start of a relevant lower\nenergy emission episode: th> tstart. These 4 limits correspond to different\nchoices of tstart, which are shown by the vertical lines in Fig. 2. We con-\nservatively used the low end of the 1 σconfidence interval for the highest\nenergy photon ( Eh= 28 GeV) and for the redshift ( z= 0.900). The most\nconservative assumption of this type is associating tstartwith the onset of\nany detectable emission from GRB 090510, namely the start of the small\nprecursor that GBM triggered on, leading to ξ1=MQG,1/MPlanck>1.19.\nHowever, it is highly unlikely that the 31 GeV photon is indee d associated\nwith the small precursor. It is much more likely associated w ith the main\nsoft gamma-ray emission, leading to ξ1>3.42. Moreover, for any reasonable\nemission spectrum, the emission of the 31 GeV photon would be accompa-\nnied by the emission of a large number of lower energy photons , which\nwould suffer a much smaller time delay due to LIV effects, and woul d there-\nfore mark its emission time. We could easily detect such phot ons in energies\nabove 100 MeV, and therefore the fact that significant high-e nergy emission\nis observed only at later times (see Fig. 2) strongly argues t hat the 31 GeV\nphoton was not emitted before the onset of the observed high- energy emis-\nsion. One could choose either the onset time of the emission a bove 100 MeV\norabove1 GeV, whichcorrespondto ξ1>5.12, andξ1>10.0, respectively.2\n2We note that there is no evidence for LIV induced energy dispe rsion that might be\nexpected if indeed the 31 GeV photon was emitted near our choi ces fortstart, together\nwith lower energy photons, as can be expected for any reasona ble emission spectrum. This\nis evident from the lack of accumulation of photons along the solidcurves in panel (a) of\nFig. 2, at least for the first 3 tstartvalues, and provides support for these choices of tstart\n(i.e. that they can indeed serve as upper limits on a LIV induc ed energy dispersion).tstartlimit on Reason for choice of Elvalid lower limit on\n(ms)|∆t|(ms) tstartor limit on ∆ t (MeV) forsnMQG,1/MPlanck\n−30<859 start of any observed emission 0.1 1 >1.19\n530 <299 start of main <1MeV emission 0.1 1 >3.42\n630 <199 start of>100 MeV emission 100 1 >5.12\n730 <99 start of >1 GeV emission 1000 1 >10.0\n— <10 association with <1MeV spike 0.1 ±1 >102\n— <19 if 0.75GeVγis from 1stspike 0.1 ±1 >1.33\n|∆t\n∆E|<30ms\nGeVlag analysis of all LAT events — ±1 >1.22\nTable2: Lower-limits ontheQuantumGravity(QG)mass scale associa ted withapossible\nlinear (n= 1) variation ofthespeedoflight with photonenergy, thatw e canplace from the\nlack of time delay (of sign sn) in the arrival of high-energy photons relative to low-ener gy\nphotons, from our observations of GRB 090510 (from [8]).\nThe5thand 6thlimits inTable2aremorespeculative, as they relyon the\nassociation ofanindividualhigh-energyphotonwithapart icularspikeinthe\nlow-energy light curve, on top of which it arrives. While the se associations\nare not very secure (the chance probability is roughly ∼5−10%), they are\nstill most likely, making the corresponding limits interes ting, while keeping\nthis big caveat in mind. The allowed emission time of these tw o high-energy\nphotons, if these associations are real, is shown by the two t hin vertical\nshadedregions inFig. 2. Forthe31 GeV photonthisgives alim itofξ1>102\nfor either sign of sn.\nThe last limit in Table 2 is based on a different method, which is com-\nplementary and constrains both signs of sn. It relies on the highly variable\nhigh-energy light curve, with sharp narrow spikes, which wo uld be smeared\nout if there was too much energy dispersion (of either sign). We have used\nthe DisCan method [10] to search for linear energy dispersio n within the\nLAT data (the actual energy range of the photons used was 35 Me V to\n31 GeV)3during the most intense emission interval (0.5–1.45s). Thi s ap-\nproach extracts dispersion information from all detected L AT photons and\ndoes not involve binning in time or energy. Using this method we obtained\na robust lower limit of ξ1>1.22 (at the 99% confidence level).\nOur most conservative limits (the first and last limits in Tab le 2) rely on\nvery different and largely independent analysis, yet still gi ve a very similar\nlimit:MQG,1>1.2MPlanck. This lends considerable support to this result,\nandmakes itmorerobustandsecurethanforeachofthemethod sseparately.\n3We obtain similar results even if we use only photons below 3 G eV or 1 GeV.0 0.5 1 1.5 2Counts/bin \n050 100 150 \nCounts/s \n05000 10000 15000 -0.03 \nGBM NaIs \n0 0.5 1 1.5 2Counts/bin \n050 100 150 200 \nCounts/s \n05000 10000 15000 20000 GBM BGOs Counts/bin \n020 40 \nCounts/s \n02000 4000 LAT \n(All events) Counts/bin \n024\nCounts/s \n0200 400 LAT \n(> 100 MeV) \nTime since GBM trigger (May 10, 2009, 00:22:59.97 UT) (s) -0.5 0 0.5 1 1.5 2Counts/bin \n0123\nEnergy [GeV] 2\n1510 20 LAT \n(> 1 GeV) 0.63 0.73 0.73 Energy (MeV) \n10 210 310 410 \n0\n0(0.26–5 MeV) (8–260 keV) (a) \n(b) \n(c) \n(d) \n(e) \n(f) 0.53 \nFigure 2: Light curves of GRB 090510 at different energies (for det ails see [8]).4 Lower limits on the bulk Lorentz factor\nThe GRB prompt emission typically has very large isotropic e quivalent lu-\nminosities ( L∼1050−1053erg s−1), significant short time scale variability,\nand typical photon energies>∼mec2(in the source cosmological frame).\nThis would result in a huge optical depth to pair production ( γγ→e+e−)\nat the source, which would thermalize the spectrum and thus b e at odds\nwith the observed non-thermal spectrum, unless the emittin g material was\nmoving toward us relativistically, with a bulk Lorentz fact or Γ≫1. This\n“compactness” argument has been used to derive lower-limit s, Γmin, on the\nvalue of Γ, which were typically ∼102and in some cases as high as a few\nhundred (see [11] and references therein). However, the pho tons that pro-\nvided the opacity for these limits were well above the observ ed energy range,\nso there was no direct evidence that they actually existed in the first place.\nWith Fermi, however, we adopt a more conservative approach o f relying\nonly on photons within the observed energy range. Under this approach,\nΓmin<∼(1+z)Eph,max\nmec2≈200(1+z)/parenleftbiggEph,max\n100MeV/parenrightbigg\n, (4)\nwhereEph,maxis the highest observed photon energy, so that setting a larg e\nΓminrequires observing sufficiently high-energy photons.\nThe main uncertainty in deriving Γ minis usually the exact choice for\nthe variability time, tv. Other uncertainties arise from those on the spectral\nfit parameters, or on the degree of space-time overlap betwee n the high-\nenergy photon and lower energy target photons, in cases wher e there is\nmore than one spectral component without conclusive tempor al correlation\nbetween their respective light curves. Finally, the fact th at our limits rely\non a single high-energy photon also induces an uncertainty, as it might still\nescape from an optical depth of up to a few. However, in most ca ses the\nsecond or third highest-energy photons help to relax the affec t this has on\nΓmin(as the probability that multiple photons escape from τγγ>1 rapidly\ndecreases with the number of photons). Thus, we have derived reasonably\nconservative Γ minvalues for 3 of the brightest LAT GRBs: Γ min≈900 for\nGRB 080916C [7], Γ min≈1200 for GRB 090510 [12], and Γ min≈1000\nfor GRB 090902B [13]. This shows that short GRBs (such as 0905 10) are\nas highly relativistic as long GRBs (such as 080916C or 09090 2B), which\nwas questioned before the launch of Fermi [14]. Since our hig hest values of\nΓminare derived for the brightest LAT GRBs, they are susceptible to strong\nselection effects. It might be that GRBs with higher Γ tend to be brighter\nin the LAT energy range (e.g. by avoiding intrinsic pair prod uction [15]).5 Delayed onset and a distinct high-energy spec-\ntral component\nA delayed onset of thehigh-energy emission ( >100 MeV) relative tothelow-\nenergy emission (<∼1 MeV) appears to be a very common feature in LAT\nGRBs. It clearly appears in all 4 of the particularly bright L AT GRBs,\nwhile in dimmer LAT GRBs it is often inconclusive due to poor p hoton\nstatistics near the onset time. The time delay, tdelay, appears to scale with\nthe duration of the GRB ( tdelay∼several seconds in thelong GRBs 080916C\nand 090902B, while tdelay∼0.1−0.2 s in the short GRBs 090510 and\n081024B, though with a smaller significance for the latter).\nOnly 3 LAT GRBs so far have shown clear ( >5σ) evidence for a distinct\nspectral component. However, these GRBs are the 3 brightest in the LAT,\nwhile the next brightest GRB in the LAT (080916C) showed a hin t for\nan excess at high energies. This suggests that such a distinc t high-energy\nspectral component is probably very common, but we can clear ly detect it\nwith high significance only in particularly bright LAT GRBs, since a large\nnumberofLATphotonsisneededinordertodetectitwith >5σsignificance.\nThe distinct spectral component is usually well fit by a hard p ower-\nlaw that dominates at high energies. In GRB 090902B a single p ower-law\ncomponent dominates over the usual Band component both at hi gh energies\n(above∼100 MeV) and low energies (below ∼50 KeV; see lower panel\nof Fig. 3). There is also marginal evidence that the high-ene rgy power-law\ncomponent in GRB 090510, which dominates above ∼100 MeV, might also\nappear at the lowest energies (below a few tens of keV).\nBoth the delayed onset and distinct spectral component rela te to and\nmay help elucidate the uncertain prompt GRB emission mechan ism. The\nmain two competing classes of models are leptonic and hadron ic origin.\nLeptonic: the high-energy spectral component might be inverse-Compt on\nemission, and in particular synchrotron-self Compton (SSC ) if the usual\nBand component is synchrotron. In this case, however, it may be hard to\nproduce the observed tdelay> tv, wheretvis the width of individual spikes\nin the lightcurve ( tdelay<∼tvmight occur due to the build-up of the seed\nsynchrotron photon field in the emitting region over the dyna mical time).\nMoreover, the gradual increase in the photon index βof the distinct high-\nenergy power-law spectral component is not naturally expec ted in such a\nmodel, and the fact that it is different than the Band low-energ y photon\nindex as well as the excess flux (above the Band component) at l ow energies\nare hard to account for in this type of model.Hadronic: tdelaymight be identified with the acceleration time, tacc, of pro-\ntons(or heavier ions)toveryhighenergies(at whichtheylo osemuchoftheir\nenergy on a dynamical time, e.g. via proton synchrotron [16] , in order to\nhave a reasonable radiative efficiency). If the observed high -energy emission\n(and in particular thedistinct high-energy spectral compo nent) also involves\npaircascades (e.g. inverse-Compton emissionbysecondary e±pairs[17]pro-\nduced in cascades initiated by photo-hadronic interaction s) then it might\ntake some additional time for such cascades to develop. Such an origin for\ntdelay(∼tacc), however, requires the high-energy emission to originate from\nthe same physical region over times > tdelay, and implies high-energy emis-\nsion rise and variability times tv>∼tacc∼tdelay, due to the stochastic nature\nof the acceleration process (while tv< tdelayis usually observed). The grad-\nual increase in βis not naturally expected in hadronic models, though it\nmight be mimicked by a time-evolution of a high-energy Band- like spectral\ncomponent [16]. For GRB 090510 a hadronicmodel requires ato tal isotropic\nequivalent energy >102times larger than that observed in gamma-rays [17],\nwhich may pose a serious challenge for the progenitor of this short GRB.\nThe excess flux at low energies that is observed in GRB 090902B (and the\nhint for such an excess in GRB 090510) may be naturally explai n in this\ntype of model by synchrotron emission from secondary pairs [ 17, 13].\nAltogether, hadronicmodelsseemtofaresomewhatbetter, h owever both\nleptonic and hadronic models still face many challenges, an d do not yet\nnaturally account for all of the Fermi observations.\n6 Long-lived high-energy emission\nIn most LAT GRBs thehigh-energy ( >100 MeV) emission lasts significantly\nlonger than the low-energy (<∼1 MeV) emission. While the high-energy\nemissionusuallyshowssignificantvariabilityduringthep rompt(low-energy)\nemission phase, in some cases showing temporal correlation with the low-\nenergy emission, the longer lived emission is typically tem porally smooth\nand consistent with a power-law flux decay (of ∼t−1.2−t−1.5) with a LAT\nphoton index corresponding to a roughly flat νFν.\nIt is most natural to interpret the prompt high-energy emiss ion as the\nhigh-energy counterpart of the prompt soft gamma-ray emiss ion, from the\nsame emission region, especially when there is temporal cor relation between\nthe low and high-energy light curves, and sometimes even fro m the same\nspectral component (as appears to be the case for GRB 080916C ). However,\nwhen there is no such temporal correlation, an origin from a d ifferent emis-sion region is possible. The longer lived smooth power-law d ecay phase is\nmore naturally attributed to the high-energy afterglow, fr om the forward\nshock that is driven into the external medium. An afterglow o rigin has been\nsuggested in some cases for the whole LAT emission [20, 15], i ncluding dur-\ning the prompt soft gamma-ray emission stage, however in thi s scenario it is\ngenerally hard to explain the sharp spikes in the LAT lightcu rve during the\nprompt phase. It is easier to test the origin of this long live d high-energy\nemission when there is good multi-wavelength coverage of th e early after-\nglow emission (e.g., in X-ray and/or optical), such as for GR B 090510 [18].\nProducing particularly high-energy photons is challengin g for a synchrotron\norigin, both during the prompt emission [7], and even more so during the\nafterglow (e.g. [21]), as it requires a very high bulk Lorent z factor and up-\nstream magnetic field, in addition to a very efficient shock acc eleration (e.g.\na 33 GeV photon observed in GRB 090902B after 82 s, well after t he end of\nthe prompt emission [13], requires Γ >1500).\n7 High-energy emission of long versus short GRBs\nSo far, 2 (12) out of the 14 LAT GRBs are of the short (long) dura tion\nclass. This implies that ∼14% of LAT GRBs are short, with a large un-\ncertainty due to the small number statistics, which is consi stent with the\n∼20% short GRBs detected by the GBM. As can be seen from Table 1, the\nhigh-energy emission properties of short and long GRBs appe ar to be rather\nsimilar. They can both produce very bright emission in the LA T energy\nrange (090510 vs. 080916C, 090902B and 090926), with a corre spondingly\nhigh lower-limit on the bulk Lorentz factor (Γ min∼103), as well a distinct\nspectral component (090510 vs. 090902B and 090926). Both sh ow a delayed\nonsetandlonerlivedhigh-energyemission, comparedtothe low-energyemis-\nsion. However, the delay in the onset of the high-energy emis sion appears\nto roughly scale with the duration of the GRB, being ∼0.1−0.2 s for\nshort GRBs and several seconds for long GRBs. This is especia lly intrigu-\ning when comparing GRBs 080916C and 090510, which had a compa rable\nisotropic equivalent luminosity (of several 1053erg s−1), suggesting another\nunderlying cause for the difference in the time delay (e.g. [22 ]).\nAnotherinterestingpotential difference, whichstillneeds tobeconfirmed\n(as there are only 2 short LAT GRBs so far, and possible select ion effects),\nis that short GRBs appear to have a comparable energy output a t high and\nlow photon energies, while long GRBs tend to radiate a smalle r fraction of\ntheir energy output at high photon energies (see upper panel of Fig. 3).Figure 3: Top panel : the fluence at high (0.1–10 GeV) versus low (20keV –\n2MeV) energies (from [19]), for 4 long (080825C, 080916C, 090217 , 090902B) and\n2 short (081024B,090510)duration LAT GRBs. The diagonallines in dicate high to\nlow energy fluence ratios of 1%, 10%, and 100%. Bottom panel : the best fit time-\nintegrated νFνspectra for the same GRBs, two of which (090510, 090902B) show a\ndistinct spectral component, well described by a hard power-law, in addition to the\nusual Band spectral component. The colored shaded regions indic ate the energy\nranges used for calculating the fluences that are displayed in the top panel .JG gratefully acknowledges a Royal Society Wolfson Researc h Merit Award.\nReferences\n[1] Hurley, K., et al. 1994, Nature, 372, 652\n[2] Gon´ zalez, M. M., et al. 2003, Nature, 424, 749\n[3] Granot, J., & Guetta, D. 2003, ApJ, 598, L11\n[4] Pe’er, A. A., & Waxman, E. 2004, ApJ, 613, 448\n[5] Giuliani, A., et al. 2008, A&A, 491, L25\n[6] Band, D. L. , et al. 2009, ApJ, 701, 1673\n[7] Abdo, A. A., et al., 2009, Science, 323, 1688\n[8] Abdo, A. A., et al., 2009, Nature, 462, 331\n[9] Jacob, U., & Piran, T. 2008, JCAP, 01, 031\n[10] Scargle, J. D., Norris, J. P., & Bonnell, J. T. 2008, ApJ, 673, 972\n[11] Lithwick, Y., & Sari, R. 2001, ApJ, 555, 540\n[12] Abdo, A. A., et al., 2010, submitted to ApJ\n[13] Abdo, A. A., et al., 2009, 706, L138\n[14] Nakar, E. 2007, Phys. Rep., 442, 166\n[15] Ghisellini, G., Ghirlanda, G., Nava, L., & Celotti, A. 2 010, MNRAS in\npress (doi:10.1111/j.1365-2966.2009.16171.x; arXiv:09 10.2459)\n[16] Razzaque, S., Dermer, C. D., & Finke, J. D. 2009, preprin t\n(arXiv0908.0513)\n[17] Asano, K., Guiriec, S., & M´ esz´ aros, P. 2009, ApJ, 705, L191\n[18] De Pasquale, M., et al. 2010, ApJ, 709, L146\n[19] Abdo, A. A., et al. 2010, ApJ in press (arXiv:1002.3205)\n[20] Kumar, P., & Barniol Duran, R. 2009, MNRAS, 400, L75\n[21] Li, Z., & Waxman, E. 2006, ApJ, 651, 328\n[22] Toma, K., Wu, X. F., & Meszaros, P., 2010, preprint (arXi v:1002.2634)" }, { "title": "1912.01332v2.Electronic_and_Magnetic_Structure_of_Infinite_layer___textrm_NdNiO__2___Trace_of_Antiferromagnetic_Metal.pdf", "content": "arXiv:1912.01332v2 [cond-mat.str-el] 4 Dec 2019Electronic and Magnetic Structure of Infinite-layer NdNiO 2:\nTrace of Antiferromagnetic Metal\nZhao Liu,1Zhi Ren,2W. Zhu,2Z. F. Wang,1and Jinlong Yang3,∗\n1Hefei National Laboratory for Physical Sciences at the Micr oscale,\nCAS Key Laboratory of Strongly-Coupled Quantum Matter Phys ics,\nUniversity of Science and Technology of China, Hefei, Anhui 230026, China\n2Institute of Natural Sciences, Westlake Institution of Adv anced Study\nand School of Science, Westlake University, Hangzhou 30002 4, China\n3Hefei National Laboratory for Physical Sciences at the Micr oscale,\nSynergetic Innovation Center of Quantum Information and Qu antum Physics,\nUniversity of Science and Technology of China, Hefei, Anhui 230026, China\nThe recent discovery of Sr-doped infinite-layer nickelate NdNiO 2[D. Li et al. Nature 572, 624\n(2019)] offers an exciting platform for investigating uncon ventional superconductivity in nickelate-\nbased compounds. In this work, we present a first-principles calculations for the electronic and\nmagnetic properties of undoped parent NdNiO 2. Intriguingly, we found that: 1) the paramagnetic\nphase has complex Fermi pockets with 3D characters near the F ermi level; 2) by including electron-\nelectron interactions, 3d-electrons of Ni tend to form (π,π,π)antiferromagnetic ordering at low\ntemperatures; 3) with moderate interaction strength, 5d-electrons of Nd contribute small Fermi\npockets that could weaken the magnetic order akin to the self -doping effect. Our results provide\na plausible interpretation for the experimentally observe d resistivity minimum and Hall coefficient\ndrop. Moreover, we elucidate that antiferromagnetic order ing inNdNiO 2is relatively weak, arising\nfrom the small exchange coupling between 3d-electrons of Ni and also hybridization with 5d-electrons\nof Nd.\nSince the discovery of high-temperature (high- Tc)\nsuperconductivity in cuprates1, extensive effort has\nbeen devoted to investigate unconventional supercon-\nductors, ranging from non-oxide compounds2,3to iron-\nbased materials4,5. Exploring high- Tcmaterials could\nprovide a new platform to understand the fundamental\nphysics behind high- Tcphenomenon, thus is quite valu-\nable. Very recently, the discovery of superconductivity\nin Sr-doped NdNiO 26potentially raises the possibility\nto realize high- Tcin nickelate family7,8.\nOne key experimental observation for the infinite-\nlayerNdNiO 2is that its resistivity exhibits a minimum\naround 70 K and an upturn at a lower temperature6.\nAt the same time that the resistivity reaches minimum,\nthe Hall coefficient drops towards a large value, sig-\nnalling the loss of charge carriers6. Interestingly, no\nlong-range magnetic order has been observed in pow-\nder neutron diffraction on NdNiO 2when temperature\nis down to 1.7 K6. This greatly challenges the exist-\ning theories, since it is generally believed that mag-\nnetism holds the key to understand unconventional\nsuperconductivity9–12. Therefore, it is highly desirable\nto study the magnetic properties of undoped parent\nNdNiO 2and elucidate its experimental indications.\nIn this work, the electronic and magnetic properties\nofNdNiO 2are systemically studied by first-principles\ncalculations combined with classical Monte Carlo cal-\nculations. Firstly, the paramagnetic (PM) phase is\nstudied. Its Fermi surface includes one large sheet and\ntwo electron pockets at Γand A point, respectively.\nThis can be described by a three-band low-energy effec-\ntive model that captures the main physics of exchange\ncoupling mechanism. Then, the magnetic properties\nare studied by including Hubbard U and (π,π,π)anti-\nferromagnetic (AFM) ordering is confirmed to be the\nmagnetic ground state. Most significantly, the Fermi\nsurface of AFM phase is simpler than that of PMphase, demonstrating an interaction induced elimina-\ntion of Fermi pockets. Before NdNiO 2enters corre-\nlated insulator, it is a compensated metal with one\nsmall electron pocket formed by dxyorbital of Nd and\nfour small hole pockets formed by dz2orbital of Ni.\nThe estimated phase transition temperature ( TN) from\nPM phase to (π,π,π)AFM phase is 70 ∼90 K for\nmoderate interaction strength of U = 5 ∼6 eV.\nThrough these studies, we identify two key mes-\nsages that are distinguishable from the cuprates: 1)\nNdNiO 2is dominated by the physics of Mott-Hubbard\ninstead of charge-transfer; 2) effective exchange cou-\npling parameters are about one-order smaller than\nthose of cuprates. In this regarding, supposed that\nthe ground state is magnetic, our calculations demon-\nstrate(π,π,π)AFM ordering is energetically favor-\nable. Moreover, our results provide a natural under-\nstanding of two experimental observations. First, 3d-\nelectrons of Ni tend to form AFM ordering around 70\n∼90 K, coinciding with the minimum in resistivity and\nthe drop in Hall coefficient. Second, the (π,π,π)AFM\nordering could be weak (compared with cuprates), be-\ncause of the small effective exchange coupling and the\nhybridization with itinerant 5 d-electrons of Nd. This\ncould be the reason why AFM ordering is missing in\nprevious study, which calls for more careful neutron\nscattering measurements on NdNiO 2.\nThe first-principle calculations are carried out with\nthe plane wave projector augmented wave method\nas implemented in the Vienna ab initio simulations\npackage (VASP)13–15. The Perdew-Burke-Ernzerhof\n(PBE) functionals of generalized gradient approxima-\ntion (GGA) is used for PM phase16. To incorpo-\nrate the electron-electron interactions, DFT + U is\nused for AFM phase, which can reproduce correctly\nthe gross features of correlated-electrons in transi-\ntion metal oxides17–19. The 4 felectrons of Nd3+2\nFigure 1: (a) Atomic structure of tetragonal NdNiO 2and first Brillouin zone. The red and blue lines (labels) deno tes the\nfirst Brillouin zone of PM phase and (π,π,π)AFM phase, respectively. The high symmetric line for band ca lculation are\nΓ(0, 0, 0)-X(0.5, 0, 0)-M(0.5, 0.5, 0)- Γ-Z(0, 0, 0.5)-R(0.5, 0, 0.5)-A(0.5, 0.5, 0.5)-Z. (b) Orbita l resolved band structure of\nPM phase. The dxy,dz2,dxz+yzanddx2−y2of Nd (spin-up Ni, spin-down Ni) are marked by black, red, blu e and pick\nfilled circles. The pzandpx+yof O are marked by red, blue and pick filled circles. The size of circles represents the orbital\nweights. (c) Perspective view of Fermi surfaces. LEP and HEP denotes light and heavy electron pocket, respectively. (d)\nThe comparison between first-principles and Wannier-fittin g bands around the Fermi-level. (e) Top and side view of three\nmaximally localized Wannier functions.\nare expected to display the local magnetic moment\nasNd3+inNd2CuO420and are treated as the core-\nlevel electrons. The Hubbard U (0 ∼8 eV) term\nis added to 3 delectrons of Ni. The energy cut-\noff of 600 eV, and Monkhorst-Pack kpoint mesh of\n11×11×11and18×18×30is used for PM and\nAFM phase, respectively. The maximally localized\nWannier functions (WFs) are constructed by using\nWannier90 package23,24. The structure of infinite-layer\nNdNiO 2is shown in Fig. 1(a), including NiO2layers\nsandwiched by Nd, which can be obtained from the\nperovskite NdNiO 3with reduction of apical O atoms\nincdirection21,22. Due to apical O vacancies, the lat-\ntice constant in cdirection shrinks (smaller than adi-\nrection) and the space group becomes P4/mmm . The\nexperimental lattice constant a=b= 3.92 Å and c=\n3.28 Å are used in our calculations.\nFirstly, We present the band structure of PM phase\nwithout Hubbard U. The orbital resolved band struc-\nture of PM phase is shown in Fig. 1(b). Compar-\ning with typical cuprates CaCuO 225, two significant\ndifferences are noted: 1) there is a gap ∼2.5 eV be-\ntween2porbitals of O and 3dorbitals of Ni. According\nto Zaanen-Sawatzky-Allen classification scheme26, this\nindicates that the physics of NdNiO 2is close to Mott-\nHubbard rather than charge-transfer; 2) there are two\nbands crossing the Fermi level, in which one is mainly\ncontributed by dx2−y2orbital of Ni (called pure-band)\nand the other one has a complicated orbital composi-\ntions (called mixed-band). In kz= 0 plane, the mixed-\nband is mainly contributed by dz2orbital of Nd andNi. The dispersion around Γpoint is relatively small,\ncalled heavy electron pocket (HEP). In kz= 0.5 plane,\nthe mixed-band is mainly contributed by dxy(dxz,dyz\nanddz2) orbital of Nd (Ni). The dispersion around\nA point is relatively large, called light electron pocket\n(LEP). As a comparison, one notices that there is only\none pure-band crossing the Fermi level in CaCuO 225.\nThe Fermi surface of PM phase is shown in Fig. 1(c).\nThere is a large sheet contributed by the pure-band, as\nthe case in CaCuO 225. This Fermi surface is obviously\ntwo-dimensional (2D), because of the weak dispersion\nalongΓ-Z. In addition, there are two electron pock-\nets residing at Γand A point, respectively, showing a\nfeature of three-dimensional (3D) rather than 2D (see\nlabels HEP and LEP in Fig. 1(c)). Therefore, the\n3D metallic state will be hybridized with the 2D cor-\nrelated state in NiO2plane, suggesting NdNiO 2to be\nan \"oxide-intermetalic\" compound27,28.\nThe existence of mixed-band also reflects the inher-\nent interactions between Nd 5dand Ni3delectrons.\nTo explore the low energy physics of NdNiO 2, a three-\nband model consisting of Ni dx2−y2, Nddz2and Nd\ndxyorbitals is constructed by Wannier90 package. As\nshown Fig. 1(d), one can see the good agreement be-\ntween first-principles and Wannier-fitting bands near\nthe Fermi level. The corresponding three maximally\nlocalized WFs are shown in Fig. 1(e), demonstrat-\ning the main feature of dz2(WF1) and dxy(WF2)\norbital of Nd, and dx2−y2(WF3) orbital of Ni. How-\never, these WFs still have some derivations from stan-\ndard atomic orbitals, that is, WF1 and WF2 are mixed3\nFigure 2: (a) Illustration of six collinear spin configurati ons. The while and black ball represents local up and down spi n\nmoment, respectively. The four exchange coupling paramete rs are indicated by the blue arrows. (b) Energy comparison\nfor the six collinear spin configurations with different valu es of Hubbard U. Energy of (π,π,π)AFM is set to zero.\nwithdz2orbital of Ni, and WF3 is mixed with px/yor-\nbital of O in the NiO2plane. According to the classi-\ncal Goodenough-Kanamori-Anderson rules29–31, these\nderivations (or hybridizations) will give clues for the\nmagnetic properties.\nTo determine the magnetic ground state of NdNiO 2,\nsix collinear spin configurations are taken into account\nin a2×2×2supercell, that is, AFM1 with q =\n(π,π,π), AFM2 with q =(π,π,0), AFM3 with q\n=(0,0,π), AFM4 with q =(π,0,0), AFM5 with q\n=(π,0,π)and FM with q= (0, 0, 0), as shown in\nFig. 2(a). Within all Hubbard U ranges, we found\nthat AFM1 configuration always has the lowest en-\nergy, as shown in Fig. 2(b), indicating a stable (π,π,π)\nAFM phase with respect to electron-electron interac-\ntions and is in accordance with random phase approx-\nimation treatment32. This can be attributed to the\nspecial orbital distributions around the Fermi level.\nThe intralayer NN exchange coupling is the typical\n180◦typed Ni-O-Ni superexchange coupling, that is,\nthe coupling between dx2−y2orbital of Ni is mediated\nbypx/yorbital of O (see WF3), preferring a (π,π)\nAFM phase in NiO2plane. The interlayer NN ex-\nchange coupling is due to the superexchange between\nthe Nidz2orbitals mediated by Nd dz2orbital as shown\nin WF1, preferring a (π,π)AFM phase between NiO2\nplanes. Therefore, the superexchange coupling results\nin a stable (π,π,π)AFM phase in NdNiO 2. Moreover,\nthe magnetic anisotropy is further checked by includ-\ning the spin-orbit coupling (SOC). We found that the\nspin moment prefers along cdirection with the mag-\nnetic anisotropic energy of ∼0.5 meV/Ni. Thus, the\ntiny SOC effect can be safely neglected in the follow-\ning phase transition temperature calculations.\nIn cuprates, the Fermi surface is unstable with\nelectron-electron interactions, making its parent phase\nto be an AFM insulator. However, this is apparently\nnot the case in NdNiO 2, because of the extra electron\npockets and the inherent interaction between Nd 5d\nand Ni3delectrons. At U=0 eV, there are two elec-\ntron pockets at Γpoint and two hole pockets along\nX-R direction as shown in Fig. 3(a)-(b). Physically,the origin of these four pockets can be easily under-\nstood through the comparison of orbital resolved band\nstructures between PM phase (Fig. 1(b)) and (π,π,π)\nAFM phase (Fig. 4). Because of the Zeeman field on\nNi, its spin-up and -down bands are split away from\neach other. The original pure-band ( dx2−y2orbital of\nNi) in PM phase becomes partially occupied in spin-\nup channel (forming two hole pockets) and totally un-\noccupied in spin-down channel. Hence, the two hole\npockets in AFM phase are inherited from large sheet\nin PM phase, showing a 2D character with neglectable\ndispersion along Γ-Z direction. For the electron pock-\nets atΓpoint, the heavier one is mainly contributed\nbydz2orbital of Nd and Ni, so it comes from the HEP\natΓpoint of PM phase. While for the lighter one, it\ncomes from the LEP at A point of PM phase which\nis folded into the Γpoint of (π,π,π)AFM phase [see\nFig. 1(a)]. The orbital composition can also be used\nto check this folded band, which is contributed by dxy\norbital of Nd, dxz/yz (dz2) orbital of spin-down (-up)\nNi.\nThese pockets have a different evolution with the\nincreasing value of Hubbard U. For electron pockets,\nthe heavier one is very sensitive to Hubbard U and\ndisappears at U = 1 eV. Meanwhile the lighter one\ndoesn’t appear until U = 6 eV. In addition, the orbital\ncomponents of lighter electron pockets are purified by\nelectron-electron interaction and it mainly contributed\nbydxyof Nd in the large U limit as shown in Fig.\n4. The case for hole pockets is rather complicated.\nFirstly, the bands of hole pockets become flat with the\nincreasing value of Hubbard U. Secondly, the original\nhole pockets formed by dx2−y2orbital of Ni gradually\ndisappear, meanwhile, a new hole pocket formed by\ndz2orbital of Ni appears along Γ-M as shown in Fig.\n3(d). At U = 6 eV, NdNiO 2is a compensated metal\nwith a small electron pocket at Γpoint and four hole\npockets along Γ-M as displayed in Fig. 3(e). Further\nincreasing the value of Hubbard U, the system enters\nan AFM insulator, just like cuprates. Therefore, the\nmetal-to-insulator phase transition point is near U ∼6\neV. If Hubbard U is less than 6 eV, NdNiO 2is an AFM4\nFigure 3: (a) Band structures of (π,π,π)AFM phase with different values of Hubbard U. (b)-(e) Perspec tive view of the\nFermi surfaces of (π,π,π)AFM with U = 0, 2, 4 6 eV. The two hole pockets in (b)-(d) are dege nerated and can not be\ndistinguished from this picture. The high symmetry k-point s in (e) are labelled to guide the eye. (f) Schematic diagram\nof major self-doping channel at U = 0 and U = 6 eV. The red/blue c olor represents dorbital of Nd/Ni respectively.\nmetal with relatively small amount of holes that are\nself-doped27,32–35intodorbitals of Ni. Interestingly,\nthere is an orbital shift from 3dx2−y2orbital of Ni at\nU = 0 eV to 3dz2orbital of Ni at U = 6 eV in the\nNiO2plane, as depicted in Fig. 3(f). We speculate\nthat this orbital shift may change the paradigm after\ndoping8,36. Moreover, without the Hubbard U, the 2p\norbital of O is far away from the Fermi level, just like\nthe case of PM phase. However, with the increasing\nvalue of Hubbard U, the gap between 3dorbital of\nNi and2porbital of O gradually decreases (Fig. 4),\ndemonstrating an evolution from Mott-Hubbard metal\nto charge-transfer insulator.\nIn order to quantitatively describe such a phe-\nnomenon, the phase transition temperature is further\ncalculated. For (π,π,π)AFM phase, the magnetic mo-\nmentum of Niincreases from 0.58 µB(U = 0 eV) to\n1.04µB(U = 8 eV) and becomes gradually saturated,\nas shown in Fig. 5(a). This is also consistent with\nthe fact that dx2−y2orbital of Ni is closer to single\noccupation with the increasing value of Hubbard U.\nTherefore, Niis spin one half ( S= 1/2) in infinite-layer\nNdNiO 2, just like the case in cuprates. To extract the\nexchange coupling parameters of J1,J2,J3andJ4(as\nlabelled in Fig. 2(a)), the total energy of five AFM\nconfigurations obtained from DFT + U calculations\nare mapped onto the Heisenberg spin Hamiltonian. In\nthe2×2×2supercell, there are 8 Ni atoms and the\ntotal energy of different AFM configurations are:EAFM1=E0−16J1S2−8J2S2+16J3S2+32J4S2\nEAFM2=E0−16J1S2+8J2S2+16J3S2−32J4S2\nEAFM3=E0+16J1S2−8J2S2+16J3S2−32J4S2\nEAFM4=E0+8J2S2−16J3S2+32J4S2\nEAFM5=E0−8J2S2−16J3S2−32J4S2(1)\nwhereE0is the reference energy without magnetic or-\nder. The calculated exchange coupling parameters as\na function of Hubbard U are shown in Fig. 5(b). We\nwould like to make several remarks here: 1) the NN\nintralayer exchange coupling ( J1), mediated by dx2−y2\norbital of Ni, demonstrates a 1/Ulaw; 2) the NN in-\nterlayer exchange coupling ( J2) is∼10 meV with lit-\ntle variation. The positive value of J2indicates an\nAFM coupling between NiO2planes; 3) the next NN\nintralayer exchange coupling ( J3), mediated by dxy\norbital of Nd, is comparable to J1at large value of\nHubbard U, which is dramatically different to that in\ninfinite-layer SrFeO 237–39. This large value could be\nattributed to the relative robustness of lighter electron\npocket and orbital purification; 4) the next NN inter-\nlayer exchange coupling ( J4) is∼0 meV, indicating\nthe validity of our Hamiltonian up to the third NN;\n5)J1,J2andJ3have the same strength at large U,\nsuggesting NdNiO 2is a 3D magnet rather than 2D\nmagnet.\nBased on the above exchange coupling parameters,\nthe phase transition temperature ( TN) is calculated by5\nFigure 4: Orbital resolved band structures of (π,π,π)AFM phase with different values of Hubbard-U. The first, secon d,\nthird and forth row represents Nd, spin-up Ni, spin-down Ni a nd O, respectively. The filled circles with different colors\nhave the same meaning as those in Fig. 1. The name of dorbitals in the AFM supercell has been aligned to that of unit\ncell.\nclassical Monte Carlo method in a 12×12×12super-\ncell based on the classical spin Hamiltonian:\nH=/summationdisplay\n/angbracketlefti,j/angbracketrightJij/vectorSi·/vectorSj (2)\nwhere the spin exchange parameters Jijhave been de-\nfined above. First, we calculate the specific heat ( C)\nafter the system reaches equilibrium at a each given\ntemperature ( T), as shown in Fig. 5(c). Then, TNis\nextracted from the peak position in the curve of C(T),\nas shown in Fig. 5(c). For U = 1 eV, TNis as high as220 K, which can be ascribed to the large value of J1.\nWith the increasing value of Hubbard U, TNgradually\ndecreases and becomes ∼70 K at U = 6 eV. To fur-\nther check the effect of interlayer exchange coupling on\n3D magnet, an additional Monte Carlo calculation is\nperformed without J2andJ4. As shown in Fig. 5(e),\ntheC(T)vsTplot shows a broaden peak at a lower\ntemperature. Since Mermin-Wagner theorem prohibit\nmagnetic order in 2D isotropic Heisenberg model at\nany nonzero temperatures40, the broad peak in C(T)\nvsTplot implies the presence of short-range order.6\nFigure 5: The Hubbard U dependence of (a) the magnetic moment um ofNiin(π,π,π)AFM phase, (b) the exchange\ncoupling parameters, (c) specific heat ( C) vsTwith four J’s, (d) the estimated TNwith four J’s, (e) specific heat ( C) vs\nTwithJ1,J3only and (f) the estimated TNwithJ1,J3only. The shaded region in (d) highlights the possible TNof 70\n∼90 K with a reasonable U = 5 ∼6 eV.\nRegarding the small drop of TN(about 30 K in Fig.\n5(f)), our MC simulations indicate that the weak in-\nteractions between NiO2planes.\nAlthough the exact value of Hubbard U cannot be\ndirectly extracted from the first principles calculations,\nits value range can still be estimated based on similar\ncompounds. The infinite-layer nickelates are undoubt-\nedly worse metals compared to elemental nickel with U\n∼3 eV41, which can be considered as a lower bound of\nHubbard U. The Coulomb interaction in infinite-layer\nnickelates should be smaller than that in the charge-\ntransfer insulator NiO with U ∼8 eV17, which can be\nconsidered as a upper bound of Hubbard U. Therefore,\na reasonable value of Hubbard U in NdNiO 2will be-\ntween 3 eV and 8 eV. In the following discussions, we\nuse U = 5 ∼6 eV27,42to draw our conclusions: 1) with\nthe decreasing of temperature, there is a phase tran-\nsition from PM phase to (π,π,π)AFM phase near 70\n∼90 K; 2) the exchange-coupling parameters are ∼10\nmeV, which is one order smaller than cuprates43–47and\nresults in a low TNcompared with cuprates; 3) the self-\ndoing effect from 5dorbital of Nd and 3dorbital of Ni\nmay screen the local magnetic momentum in dx2−y2\norbital of Ni, which gives a small magnetic momen-\ntum less than 1 µBand makes the long-range AFM\norder unstable6,21,22,48; 4) the Fermi surface of PM\nphase is quite large with two 3D-liked electron pock-\nets, while the Fermi surface of (π,π,π)AFM phase\nis quite small with one 3D-liked electron pocket and\nfour 2D-liked hole pocket. Therefore, there could ex-\nist a crossover from normal metal to bad AFM metal\naroundTN∼70−90K, which provides a plausible\nunderstanding of minimum of resistivity and Hall co-efficient drop in infinite-layer NdNiO 26. We envision\nthat our calculations will intrigue intensive interests\nfor studying the magnetic properties of high quality\ninfinite-layer NdNiO 2samples.\nLastly, we would like to make some remarks on the\nexisting experiments. Some recent experiments fail to\nfind bulk superconductivity in NdNiO 2systems, and\nthe parent samples show strong insulating behaviors49.\nThe insulating behavior could be attributed to strong\ninhomogenious disorder or improper introduction of\nH during the reaction with CaH 250. Especially, it is\nworth noting, the experiments cannot rule out the pos-\nsibility of weak AFM ordering, due to the presence of\nNi impurities in their samples. (Actually this prob-\nlem has been pointed out before21,22.) The strong fer-\nromagnetic order from elemental Ni would dominate\nover and wash out the weak signal of AFM ordering\nfromNdNiO 2as we suggested in this work. In this re-\ngarding, the upcoming inelastic neutron scattering on\nhigh-quality samples is highly desired.\nW.Z. thanks Chao Cao for sharing their unpub-\nlished DMFT results, and thanks Filip Ronning, H. H.\nWen, G. M. Zhang for helpful discussion. This work\nwas supported by NSFC (No. 11774325, 21603210,\n21603205, 21688102), National Key Research and De-\nvelopment Program of China (No. 2017YFA0204904,\n2016YFA0200604), Anhui Initiative in Quantum Infor-\nmation Technologies (No. AHY090400), Fundamen-\ntal Research Funds for the Central Universities and\nthe Start-up Funding from Westlake University. We\nthank Supercomputing Center at USTC for providing\nthe computing resources.\n∗E-mail: jlyang@ustc.edu.cn1J. G. Bednorz and K. A. Müller, Z. Phys. B 64, 189-1937\n(1986).\n2J. Nagamatsu, N. Nakagawa, T. Muranaka, Y. Zenitani\nand J. Akimitsu, Nature 410, 63-64 (2001).\n3A. P. Drozdov, M. I. Eremets, I. A. Troyan, V. Kseno-\nfontov and S. I. Shylin, Nature 525, 73-76 (2015).\n4Y. Kamihara, H. Hiramatsu, M. Hirano, R. Kawamura,\nH. Yanagi, T. Kamiya and H. Hosono, J. Am. Chem.\nSoc.128, 10012-10013 (2006).\n5Q.-Y. Wang et al., Chin. Phys. Lett. 29, 037402 (2012).\n6D. Liet al., Nature 572, 624-627 (2019).\n7V. I. Anisimov, D. Bukhvalov and T. M. Rice, Phys.\nRev. B 59, 7901-7906 (1999).\n8K.-W. Lee and W. E. Pickett, Phys. Rev. B 70, 165109\n(2004).\n9A. L. Patrick, N. Nagaosa and X.-G. Wen, Rev. Mod.\nPhys.78, 17-85 (2006).\n10B. Keimer, S. A. Kivelson, M. R. Norman, S. Uchida\nand J. Zaanen, Nature 518, 179-186 (2015).\n11P. Dai, J. Hu and E. Dagotto, Nat. Phys. 8, 709-718\n(2012).\n12P. Dai, Rev. Mod. Phys. 87, 855-896 (2015).\n13M. P. Teter, M. C. Payne and D. C. Allan, Phys. Rev.\nB40, 12255 (1989).\n14G. Kresse and J. Furthmüller, Phys. Rev. B 54, 11169\n(1996).\n15P. E. Blöchl, Phys. Rev. B 50, 17953 (1994).\n16J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev.\nLett.77, 3865-3868 (1996).\n17V. I. Anisimov, J. Zaanen and O. K. Anderson, Phys.\nRev. B 44, 943-954 (1991).\n18S. L. Dudarev, G. A. Botton, S. Y. Savrasov, C. J.\nHumphreys and A. P. Sutton, Phys. Rev. B 57, 1505\n(1998).\n19A. Rohrbach, J. Hafner and G. Kresse, J. Phys.: Con-\ndens. Matter 15, 979-996 (2003).\n20H. Casalta, P. Bourges, M. d’Astuto, D. Petitgrand and\nA. Ivanov, Phys. Rev. B 57, 471-475 (1998).\n21M. A. Hayward, M. A. Green, M. J. Rosseinsky and J.\nSloan, J. Am. Chem. Soc. 121, 8843-8854 (1999).\n22M. A. Hayward and M. J. Rosseinsky, Solid State Sci.\n5, 839-850 (2003).\n23A. A. Mostofi, J. R. Yats, Y.-S. Lee, I. Souza, D. Van-\nderbilt and N. Marzari, Comput. Phys. Commun. 178,\n685 (2008).\n24N. Marzari, A. A. Mostofi, J. R. Yats, I. Souza and D.\nVanderbilt, Rev. Mod. Phys. 84, 1419-1475 (2012).\n25A. S. Botana and M. R. Norman, arXiv: 1908.10946\n(2019).26J. Zaanen, G. A. Sawatzky and J. W. Allen, Phys. Rev.\nLett.55, 418-421 (1985).\n27M. Hepting et al., arXiv: 1909.02678 (2019).\n28Y. Gu, S. Zhu, X. Wang, J. Hu and H. Chen, arXiv:\n1911.00814 (2019).\n29J. B. Goodenough, Phys. Rev. 100, 564 (1955).\n30P. W. Anderson, Phys. Rev. 115, 2 (1959).\n31J. Kanamori, J. Phys. Chem. Solids 10, 87-89 (1959).\n32X. Wu, D. Di Sante, T. Schwemmer, W. Hanke, H. Y.\nHwang, S. Raghu and R. Thomale, arXiv: 1909.03015\n(2019).\n33Y. Nomura, M. Hirayama, T. Tadano, Y. Yoshimoto, K.\nNakamura and R. Arita, arXiv: 1909.03942 (2019).\n34H. Sakakibara, H. Usui, K. Suzuki, T. Kotani, H. Aoki,\nand K. Kuroki, arXiv: 1909.00060 (2019).\n35G.-M. Zhang, Y.-F. Yang and F.-C. Zhang, arXiv:\n1909.11845 (2019).\n36F. Lechermann, arXiv: 1911.11521 (2019).\n37Y. Tsujimoto et al., Nature 450, 1062 (2007).\n38M. A. Hayward and M. J. Rosseinsky, Nature 450, 960\n(2007).\n39H. J. Xiang, S.-H. Wei and M.-H. Whangbo, Phys. Rev.\nLett.100, 167207 (2008).\n40N. D. Mermin and H. Wagner, Phys. Rev. Lett. 100,\n1133-1136 (1966).\n41F. Aryasetiawan, K. Karlsson, O. Jepsen and U. Schön-\nberger, Phys. Rev. B 74, 125106 (2006).\n42M. Cococcioni and S. de Gironcoli, Phys. Rev. B 17,\n1133 (1966).\n43P. Bourges, H. Casalta, A. S. Ivanov and D. Petitgrand,\nPhys. Rev. Lett. 79, 4906 (1997).\n44R. Coldea, S. M. Hayden, G. Aeppli, T. G. Perring, C.\nD. Frost, T. E. Mason, S.-W. Cheong and Z. Fisk, Phys.\nRev. Lett. 86, 5377 (2001).\n45L. Braicovich et al., Phys. Rev. Lett. 102, 167401\n(2009).\n46G. Blumberg, R. Liu, M. V. Klein, W. C. Lee, D. M.\nGinsberg, C. Gu, B. W. Veal and B. Dabrowski, Phys.\nRev. B 49, 13295 (1994).\n47G. Blumberg, P. Abbamonte, M. V. Klein, W. C. Lee,\nD. M. Ginsberg, L. L. Miller and A. Zibold, Phys. Rev.\nB53, R11930 (1994).\n48Y. Fu et al., arXiv: 1911.03177 (2019).\n49Q. Li, C. He, S. Jin, X. Zhu, Y. Zhang and H.-H. Wen,\narXiv: 1911.02420 (2019).\n50L. Si, W. Xiao, J. Kaufmann, J. M. Tomczak, Y. Lu, Z.\nZhong and K. Held, arXiv: 1911.06917 (2019)." }, { "title": "1611.10062v2.Optical_properties_of_the_optimally_doped_Ca___8_5__La___1_5___Pt__3_As__8___Fe___2__As___2_____5__single_crystal.pdf", "content": "Optical properties of the optimally doped Ca 8:5La1:5(Pt3As8)(Fe 2As2)5single crystal\nYu-il Seo1, Woo-Jae Choi1, Shin-ichi Kimura2, Yunkyu Bang3,\u0003and Yong Seung Kwon1y\n1Department of Emerging Materials Science, Daegu Gyeongbuk Institute\nof Science and Technology (DGIST), Daegu 711-873, Republic of Korea\n2FBS and Department of Physics, Osaka University, Suita 565-0871, Japan\n3Department of Physics, Chonnam National University, Kwangju 500-757, Republic of Korea\n(Dated: August 6, 2021)\nWe have measured the reflectivity of the optimally doped Ca 8:5La1:5(Pt3As8)(Fe 10As10) single crystal ( Tc\n= 32.8K) over the broad frequency range from 40 cm\u00001to 12000cm\u00001and for temperatures from 8K to 300\nK. The optical conductivity spectra of the low frequency region ( <1;000cm\u00001) in the normal state (80 K\n< T\u0014300 K) is well fitted with two Drude forms, which indicates the presence of multiple bands at the\nFermi level. Decreasing temperature below 80 K, this low frequency Drude spectra develops pseudogap (PG)\nhump structure at around \u0019100cm\u00001and continuously evolves into the fully opened superconducting (SC)\ngap structure below Tc. Theoretical calculations of the optical conductivity with the preformed Cooper pair\nmodel provide an excellent description of the temperature evolution of the PG structure above Tcinto the SC\ngap structure below Tc. The extracted two SC gap sizes are \u0001S= 4.9meV and\u0001L= 14.2meV , suggesting\nCa8:5La1:5(Pt3As8)(Fe 10As10) as a multiple gap superconductor with a mixed character of the weak coupling\nand strong coupling superconductivity.\nPACS numbers: 74.25.nd,74.20.Rp,74.70.Xa\nI. INTRODUCTION\nRecently discovered Fe-pnictide compounds,\nCa10(Pt3As8)(Fe 2As2)5(so-called 10-3-8 compound)\nand Ca 10(Pt4As8)(Fe 2As2)5(so-called 10-4-8 compound)1–4,\nadded another new class to the family of the Fe-based\nsuperconductors (IBSs). They are showing the prototype\nbehavior of the subtle balance and competition between\nmagnetism and superconductivity commonly observed in\nother IBSs. Both compounds share similar overall crystal\nstructure, consisting of tetrahedral FeAs planes sandwiched\nbetween the planar Pt nAs8(n= 3;4) intermediary layers.\nHowever, in fine details, there exist also distinct differences\nbetween them.\nThe parent 10-4-8 compound has a tetragonal crystal struc-\nture and is metallic. It becomes superconducting (SC) with-\nout doping with the maximum Tcof\u001838K1,5, and doping\nof electrons only suppresses the SC transition temperature4,6.\nOn the other hand, the parent 10-3-8 compound has a triclinic\ncrystal structure. Sturzer et al.6reported that this compound\nis semiconducting and becomes antiferromagnetically ordered\nbelowTN\u0019120K without any further reduction in the crys-\ntal symmetry. Superconductivity can be induced by doping\nof electrons: for instance, La substitution at the Ca site, or\nPt substitution at the Fe site. The maximum Tcfor each sub-\nstitution can be\u001832K and\u001815K, respectively6. How-\never, Neupane et al.7reported that the undoped 10-3-8 com-\npound has the multiple Fermi surfaces like other typical FeAs-\nsuperconducting compounds and becomes superconducting at\nTc\u00188Kwithout doping.\nThe band structure calculations8suggested that these dif-\nferences arise from the increased metallicity of the PtAs lay-\ners. And recent angle resolved photoemission spectroscopy\n(ARPES) experiment further suggested that this difference be-\ntween two compounds could be developed by the number of\nthe band-edge singularities9. In order to study this subtle re-lation between the electronic structure and superconductivity\nin these materials, various experiments have already been per-\nformed: such as transport10, pressure effect11, ARPES7, mag-\nnetic force microscope (MFM)12, upper critical field ( Hc2)13,\nnuclear magnetic resonance (NMR)14, and penetration depth\n\u0015(T)15measurements, etc. However, the IR spectroscopy ex-\nperiment has not yet been carried out with these compounds.\nThis technique is particularly a useful tool to directly study\nthe low energy dynamics of the correlated materials with tem-\nperature variation, and, therefore, it is able to investigate the\nsystematic development of superconductivity with tempera-\nture and any non-trivial evolution of the correlation effects, if\nexists.\nIn this paper, we have measured the IR reflectivity and an-\nalyzed the optical properties of the optimally La-doped 10-\n3-8 compound, Ca 8:5La1:5(Pt3As8)(Fe 2As2)5single crystal\n(Tc\u001932.8 K). In view of the phase diagram with electron dop-\ning of the previous study10, our compound is located far from\nthe antiferromagnetic (AFM) quantum critical point (QCP)\nwhereTNgoes to zero in the phase diagram. However, we\nspeculate that our compound is located near another non-\nmagnetic QCP in the universal phase diagram, from the ob-\nservations of (1) the maximum Tcwith doping variation, and\n(2) theT-linear resistivity data above Tcup to 300 K. It is\ninteresting to note that these QCP behaviors far away from\nthe AFM QCP are similar to the case of the cuprate super-\nconductors. Then, most interestingly, we have also found the\npseudogap (PG) behavior in the optical conductivity of our\nsample up to the temperature about three times higher ( \u001880\nK) thanTc(\u001933 K). With the observation that our compound\nis far away from the AFM QCP, it is logical to conclude that\nthe PG behavior of the optimal doped 10-3-8 compound is\nnot related to the AFM correlation. We have employed the\npreformed Cooper pair model to calculate the real part of the\noptical conductivity \u001b(!;T)and successfully reproduced the\nPG feature far above Tcand its consistent evolution to the SCarXiv:1611.10062v2 [cond-mat.supr-con] 28 Feb 20172\n05 01 001 502 002 503 00050100150C\na8.5La1.5(Pt3As8)(Fe2As2)5 \n ρ(µΩ⋅cm)T\n(K)Tc=35.2 KT\nc,zero=32.8K\nFIG. 1: (Color online) Temperature dependence of the DC electrical\nresistivity\u001a(T)of the Ca 8:5La1:5(Pt3As8)(Fe 2As2)5single crystal.\ngap structure below Tc.\nCombining the previous reports of the PG observations in\nseveral IBS compounds and cuprate superconductors, our ob-\nservation establishes that the PG phenomena induced by a SC\ncorrelation is a generic phenomena of the strongly correlated\nunconventional superconductors.\nII. EXPERIMENTAL DETAILS\nHigh quality single crystals of\nCa8:5La1:5(Pt3As8)(Fe 2As2)5were grown by a Bridge-\nman method with a sealed molybdenum crucible at 1250\u000e\nC. Before performing the Bridgeman method, we made\nthe precursors of CaAs, LaAs and FeAs in advance. Fig-\nure 1 shows the in-plane electrical resistivity \u001a(T)of the\nCa8:5La1:5(Pt3As8)(Fe 2As2)5single crystal. We do not see\nany noticeable anomaly in \u001a(T)aroundT0\nN\u0019120 K (the\nNeel temperature of the undoped parent 10-3-8 compound)\nand below due to the spin density wave (SDW) transition.\nThe SC transition begins at Tc=35.2 K and zero resistivity\noccurs atTc;zero = 32:8K, respectively. These transition\ntemperatures are about 7-9 K higher than the one with similar\nnominal doping of La ( x= 1:45) in the previously reported\nwork10, indicating that our sample is a substantially higher\nquality single crystal.\nAnother noticeable feature is that \u001a(T)shows theT-linear\ntemperature dependence above Tcup to 300 K. This linear-\nin-Tbehavior in \u001a(T)is commonly observed in other opti-\nmally doped IBSs16–19and is a signature that the sample is\nnear the QCP. This is consistent with the phase diagram of\nRef.10, where the maximum SC Tcwas also located with the\nLa doping ” xsc\nmax”\u00191:5. Moreover the phase diagram of\nRef.10shows that the magnetic critical doping ” xmag\nc”, where\nTN!0, is near\u00190:3, which is far apart from the optimaldopingxsc\nmax for the maximum SC Tc. Therefore, in view\nof the phase diagram of Ref.10and our result, we speculate\nthat there exist two QCPs in the La-doped 10-8-3 compound:\none is the QCP due to magnetic fluctuations where TN!0,\nand the other is the QCP due to unknown quantum fluctuations\nwhere the SC Tcbecomes maximum. The former type of QCP\nis more common with many IBSs as well as most of heavy\nfermion superconductors, while the latter type of QCP is well\nrepresented in the high- Tccuprate superconductors. There-\nfore, clarifying the nature of the QCP in the 10-3-8 compound\nwill be particularly interesting in connection with the mystery\nof the high-Tccuprate superconductors.\nIn order to investigate the electronic structure as well as the\nSC properties of Ca 8:5La1:5(Pt3As8)(Fe 2As2)5, we have mea-\nsured the optical reflectivity R(!;T)of this single crystal in\na broad frequency range from 40 to 12000 cm\u00001and for vari-\nous temperatures from 8 to 300 K. We used a Michelson-type\nrapid-scan Fourier spectrometer (Jasco FTIR610). In particu-\nlar, we used a specially designed feedback positioning system\nto drastically reduce the overall uncertainty level. The uncer-\ntainty of our data was maintained less than 0.5 %. For more\ndetails of experimental methods of reference setting and con-\ntrol of data uncertainty, we refer to our previous paper20.\nIII. RESULTS AND DISCUSSIONS\nA. Reflectivity and Optical Conductivity\nFigure 2 shows the reflectivity spectra R(!;T)of\nCa8:5La1:5(Pt3As8)(Fe 2As2)5single crystal at several differ-\nent temperatures from 8 K to 300 K. The main panel shows\nthe full measured data for frequencies from 50 cm\u00001to 12000\ncm\u00001in log scale, and the inset shows the close-up view of the\ndata from 0 cm\u00001to 300cm\u00001in linear scale; here the data\nfrom 0cm\u00001to 60cm\u00001(dotted lines) are not the measured\nones but extrapolated ones as explained below (although we\nhave data from 40 cm\u00001, we didn’t use the data of 40 cm\u00001-\n60cm\u00001because the data in this part are not uniformly clean).\nIn the normal state (the data sets for T\u001540K),R(!)\napproaches to unity at zero frequency and increases as de-\ncreasing temperature in far-infrared region, showing a typi-\ncal metallic behavior of Ca 8:5La1:5(Pt3As8)(Fe 2As2)5. Upon\nentering the SC phase (the data sets below 40 K), the low fre-\nquency reflectivity turns up quickly, and reaches the flat unity\nresponse below\u001880cm\u00001reflecting the SC gap opening. As\nthe temperature increases toward Tcthe flat unity response\nshrinks to the lower frequency region as the SC gap size is\ndecreasing. The shape of the flat part in R(!)at low frequen-\ncies and its temperature dependence are clear signatures of the\nfully opened SC gap. However, the most interesting develop-\nment of the reflectivity spectra in Fig.2 occurs in the data sets\nfor 80 K (cyan color) and 40 K (purple color), which are still\nin the normal state. Both data sets show a curious suppression\nof spectra in the low frequency region around 100 cm\u00001(see\nthe inset of Fig.2).\nFor more convenient analysis, we converted our reflectivity\ndataR(!)into the real part of the optical conductivity \u001b1(!)3\nby the Kramers-Kronig (KK) transformation. For extrapola-\ntion of the data for the KK-transformation, at low frequencies,\nwe used the Hagen-Rubens extrapolation formula for the nor-\nmal state and the form ( 1\u0000A!4) below the gap in the SC\nstate. For high frequencies above 12000 cm\u00001, the standard\n1=!4form plus a constant reflectivity are used up to 40 eV.\n1001 0001 00000.40.50.60.70.80.91.00\n5 00.920.940.960.981.001\n001 502 002 503 00 \n ReflectivityF\nrequency (cm-1) \n Ca8.5La1.5(Pt3As8)(Fe2As2)5F\nrequency (cm-1) \n Reflectivity 8 K \n17 K \n28 K \n40 K \n80 K \n130 K \n240 K \n300 K\nFIG. 2: (Color online) Reflectivity spectra R(!)of the\nCa8:5La1:5(Pt3As8)(Fe 2As2)5single crystal at several temperatures.\nThe inset shows the enlarged view of spectra below 300 cm\u00001; dot-\nted lines are the part from extrapolation as explained in the main text.\n02 004 006 008 001 00002000400060008000F\nrequency (cm-1)\n02 0004 0006 0008 0001 0000100015002000250030003500F\nrequency (cm-1)σ1 (Ω-1cm-1) \n Ca8.5La1.5(Pt3As8)(Fe2As2)5 \n \n8K \n17K \n28K \n40K \n80K \n130K \n240K \n300Kσ1(Ω−1cm-1)\nFIG. 3: (Color online) Real part of optical conductivity \u001b1(!)of\nCa8:5La1:5(Pt3As8)(Fe 2As2)5single crystal for 8, 17, 28, 40, 80,\n130, 240, and 300 K, respectively. The pseudogap feature of the 40K\nand 80 K data is indicated inside the dotted circle. The inset shows\nthe same data \u001b1(!)in wider view up to 10,000 cm\u00001.\nFigure 3 shows these results of \u001b1(!), obtained from the\ndata of Fig.2, in the frequency range from 0 to 1000 cm\u00001\nfor various temperatures. The optical conductivity spectra at\n130, 240 and 300 K – which are in normal state – show a verybroad Drude peak centered at != 0 and then monotonically\ndecrease until the interband transition starts appearing around\n1000cm\u00001(see the inset of Fig.3). The interband transition\nspectra continues to form a broad hump around 6000 cm\u00001;\nwhile the peak position of the hump is almost temperature in-\ndependent, there exists an interesting spectral weight transfer\nwith temperature variation which will be analyzed in section\nIII.B.\nAs the temperature decreases down to 80 K (cyan color) and\n40 K (blue color) – which are still in normal state by trans-\nport/thermodynamic measurements – (1) the Drude part of\nspectra rapidly sharpens, and then (2) a peculiar hump struc-\nture around 80-150 cm\u00001appears on top of the smooth Drude\nspectra. The former is the expected evolution due to the for-\nmation of coherent quasiparticles as temperature decreases.\nThe striking feature is the latter, namely, the appearance of the\nshoulder-like hump structure far above Tcon top of the Drude\nresponse in the frequency range of 80-150 cm\u00001, indicating\nan incomplete gap – hence called PG – formation over the\nFermi surfaces or in a part of the Fermi surfaces. Then below\nTc, i.e., at 28 K, 17 K and 8 K, \u001b1(!)shows a clean open-\ning of the SC energy gap with the absorption edge at about 80\ncm\u00001and a peak position at around 150 cm\u00001. Comparing\nthis SC gap structure at around 80-150 cm\u00001belowTc, the\nhump structure around 80-150 cm\u00001, aboveTcat 40 K and\n80 K, appears to be a continuous evolution of the SC correla-\ntion belowTc.\nThe observation of the PG features in the IBSs with the\noptical spectroscopy measurements is now quite common.\nFor example, the underdoped ( x\u00190:12;0:2)21and the\nslightly underdoped ( x\u00190:3)20Ba1\u0000xKxFe2As2have re-\nported the similar behavior in their optical measurements:\nnamely, the hump structures in the optical conductivity \u001b1(!)\naboveTc(about three times of Tcfor all cases) in the same\nenergy scale as the SC gap energy, and their continuous evo-\nlution into the SC gap. These authors concluded that this\nhump structure is not related with a magnetic correlation but\nrather connected to the SC gap, hence possibly a precur-\nsor of the preformed Cooper pairs. The optimally Co-doped\nBa(Fe 0:92Co0:08)As 222also has shown a similar hump struc-\nture at 30 K ( Tc=22.5 K) and its evolution into the SC gap be-\nlowTc; however, the authors of Ref.22advocated, as its origin,\nthe impurity bound state or a low-energy interband transition,\nwith which we do not agree.\nMore recently, Surmach et al.23have performed a com-\nprehensive measurements of muon-spin relaxation ( \u0016SR),\ninelastic neutron scattering (INS), and NMR on Pt-doped\n(CaFe 1\u0000xPtxAs) 10Pt3As8(Tc= 13 K) – the same 10-3-8 com-\npound we studied in this paper but with Pt doping on Fe sites.\nThey found a PG behavior in the 1=T1data of75As NMR be-\nlowT\u0003\u001945K (about three times higher than Tc). These\nauthors concluded from a combination of measurements of\n\u0016SR, INS, together with NMR that this PG behavior is likely\nto be associated with the preformed Cooper pairs. There-\nfore, our observation of a similar PG behavior in the optimally\ndoped Ca 8:5La1:5(Pt3As8)(Fe 2As2)5compound has strength-\nened the universal nature of the PG phenomena originating\nfrom the preformed Cooper pairs or a precursor effect from4\n05 01 001 502 002 503 000.00.51.01.505 01 001 502 002 503 000.60.81.0(\na)c\nutoff freq. (cm-1) \n SW(T)/SW(300K) \n1000 \n500 \n200(b) \n SW(T)/SW(300K) \n1000 \n2000 \n4000 \n5000 \n7000 \n10000cutoff freq. (cm-1)\n05 01 001 502 002 503 000.00.51.01.52.0c\nutoff freq. (cm-1) \n SW(T)/SW(300K)T\n(K) 200 \n100 \n 80 \n 60 \n 40(c)\nFIG. 4: (Color online) Temperature dependence of the normalized\n”partial” spectral weights SW(T;!c)=SW (300K;!c)of the opti-\ncal conductivity data in Fig.3 with different cutoff frequency !c. (A)\ndata with!c\u00151000cm\u00001; (B) data with !c\u00141000cm\u00001; and (C)\ndata with finer variation of cutoff frequencies for !c\u0014200cm\u00001:\n200, 100, 80, 60, and 40 cm\u00001, respectively.\nthe SC correlation in the 10-3-8 compound.\nB. Temperature Dependence of Spectral Weight Transfer\nTo further understand the temperature evolution of the opti-\ncal conductivity shown in Fig.3, we have analyzed the temper-\nature dependence of the ”partial” spectral weight SW(T;!c)\nwith different cutoffs defined as\nSW(T;!c) =Z!c\n0+\u001b1(!;T)d!; (1)\n0 2000 4000 6000 8000 10000100015002000250030003500\nFrequency (cm-1)1 (-1cm-1)\n FIG. 5: (Color online) A schematic illustration of the spectral weight\ntransfer with temperature variation. The spectral density in mid-\nfrequency range ( \u0018500cm\u00001< ! < \u00185;000cm\u00001) is contin-\nuously depleted with decreasing temperature. This depleted spectral\nweight is roughly divided by !\u0003\u00181;000cm\u00001(black vertical line),\nand the spectral weight below !\u0003and the spectral weight above !\u0003\nare separately conserved (of course, this separate conservation rule\nholds only approximately).\nwhere!cis a cutoff frequency. Figure 4 shows the results of\nSW(T;!c), normalized by SW(T= 300K;!c), that reveal\nnon-trivial information of the correlated electron system.\nFig.4(A) shows that the total spectral weight up to 10,000\ncm\u00001is constant on changing temperature from Tcto 300 K\nconfirming that the sum rule is satisfied24. However, with low-\nering cutoff frequency !c, the sum rule is being deviated, as\nit should be, but in a non-trivial way. First, the normalized\n”partial” spectral weight SW(T;!c)=SW (300K;!c)with\n1;000cm\u00001< !c<10;000cm\u00001is monotonically de-\ncreasing with lowering temperature down to Tc. This means\nthat the spectral weight below the cutoff frequency !cis trans-\nferred to the higher frequency region above the cutoff fre-\nquency!cwhen temperature decreases, which is an opposite\nbehavior from a standard Drude type metallic state. Second,\nthe rate of this spectral weight transfer is not monotonously\nincreasing with lowering the cutoff frequency !c; the de-\nceasing rate increases to the maximum when the cutoff fre-\nquency is lowered to !c= 4;000cm\u00001(blue triangles) and\nthen it becomes weaker with lowering cutoff frequencies to\n!c= 2;000cm\u00001;and1;000cm\u00001. Third, in particular,\nwhen!c= 1;000cm\u00001, the partial sum rule is ”almost” re-\ncovered again: namely, the spectral weight below !\u0003\nc= 1;000\ncm\u00001(red circles) is separately conserved with respect to the\ntemperature variation up to Tc.\nFig.4(B), on the other hand, shows the similar plots of\nSW(T;!c)=SW (300K;!c)with!c\u0014!\u0003\nc(= 1;000cm\u00001).\nIt clearly shows that the partial spectral weight is ”increasing”5\n– not decreasing –with decreasing temperature. This is an op-\nposite behavior to the ones in Fig.4(A).\nThis complicated spectral weight transfer of SW(T;!c)\nwith temperature variation is summarized in Fig.5. It shows\nthat the spectral weight transfer of \u001b1(!)with temperature\nvariation is roughly divided into two parts: one part below\n!\u0003\nc= 1;000cm\u00001and the other part above !\u0003\nc= 1;000\ncm\u00001, and each part separately conserves the spectral weight.\nThe physical meaning of it is that the correlated electrons are\ndivided into the low energy itinerant part (Drude spectra) and\nthe high energy localized part (Lorentzian spectra). At high\ntemperatures, a large portion of spectra in the intermediate\nenergy range – in this particular compound, in between 500\ncm\u00001and 5000cm\u00001– is undetermined, hence remains in-\ncoherent. Lowering temperature, these incoherent spectra are\ncontinuously depleted and transferred either into the low fre-\nquency Drude part or into the high frequency Lorentzian part:\nthe final fate of the incoherent spectra was roughly predeter-\nmined by the frequency !\u0003\nc= 1;000cm\u00001– which is the\ndeepest valley point of the \u001b1(!)spectra in Fig.5 – in this\nparticular 10-3-8 compound. The sharpening of the Drude\nspectra with decreasing temperature is well understood as a\ndevelopment of coherent quasiparticles. The spectra transfer\nof the high frequency localized part should be associated with\nthe strong correlation effects of the local interactions such as\nHubbard interactions ( U;U0) and Hund’s couplings ( J)25–27.\nFinally, in order to scrutinize the development of hump\nstructure around 80-150 cm\u00001with temperature variation, we\nanalyzed the low frequency spectra transfer in Fig.4(C), with\nfiner variation of cutoff frequencies for !c\u0014200cm\u00001: 200,\n100, 80, 60, and 40 cm\u00001, respectively. The main observa-\ntion is that there is a sudden change of the slopes of the spec-\ntral weight transfer with the cutoff frequency !cbelow and\nabove!c= 100cm\u00001; the slopes suddenly increase when\n!c\u0014100cm\u00001. This means that decreasing temperature be-\nlow\u0018130K, there appears another drain of spectral weight\ntransfer to the region in between 100cm\u00001\u0000200cm\u00001be-\nsides the narrowing of Drude spectra. This another drain of\nspectral weight transfer is the formation of ”hump” structure\naround 80cm\u00001\u0000150cm\u00001as seen with the 40K and 80K\ndata in Fig.3.\nC. Drude-Lorentz model fitting for \u001bexp\n1(!;T >T c)in the\nnormal state\nWe have shown that the total spectra of the normal state op-\ntical conductivity \u001b1(!)consists of two spectral parts – Drude\npart and Lorentz part – and each part conserves their spectral\nweights: this separate conservation of sum rule will be con-\nfirmed once more in this section. Therefore we tried to fit the\ndata with the standard Drude-Lorentz model\n\u001b1(!) =1\n4\u0019RehX\nj!2\np;j\n1\n\u001cD;j\u0000i!+X\nkSk!\n!\n\u001cL;k+i(!2\nL;k\u0000!2)i\n(2)\nwhere!p;jand1\n\u001cD;jare the plasma frequency and the scatter-\ning rate for the j-th free carrier Drude band, respectively, andSk,!L;kand1\n\u001cL;kare the spectral weight, the Lorentz oscil-\nlator frequency, and the scattering rate of the k-th oscillator,\nrespectively.\n1001 0001 000002000400060008000100001\n001 0001 00000200040006000800010000Lorentz 3Drude 2 \n \ndata (T=130 K) \nTotal \nSum of Drudes \nSum of LorentzesCa8.5La1.5(Pt3As8)(Fe2As2)5σ1 (Ω-1cm-1)F\nrequency (cm-1)Drude 1L\norentz 1L\norentz 2(a)(\nb)D\nrude 2Ca8.5La1.5(Pt3As8)(Fe2As2)5F\nrequency (cm-1)σ1 (Ω-1cm-1) \n \n data (T=40 K) \nTotal \nSum of Drudes \nSum of LorentzesD\nrude 1\nFIG. 6: (Color online) Typical Drude-Lorentz model fittings of the\noptical conductivity \u001b1(!)in the normal state: (A) 130 K data; (B)\n40 K data. Both data were fitted with two Drude forms and three\nLorentz oscillators. The Lorentz oscillator parts and Drude-2 form do\nnot change much with temperatures, but the Drude-1 form becomes\nsubstantially sharpened with decreasing temperature.\nFigure 6(A) shows that the 130 K data of \u001b1(!)and its\nDrude-Lorentz model fitting. It is well fitted with two Drude\nterms (one narrow and the other broad) and three Lorentz os-\ncillator terms. Here two Drude terms show the presence of\nmultiple bands in Ca 8:5La1:5(Pt3As8)(Fe 2As2)5, as in many\nIBS compounds, e.g., LiFeAs compound28. The data of \u001b1(!)\nof 300 K and 240 K are also well fitted with the Drude-Lorentz\nmodel with the similar parameters: the spectral weights,\n!2\np;j=1;2andSk=1;2;3, and the Lorentz oscillator frequencies\n!L;k=1;2;3are the same, but only the scattering rates,1\n\u001cD;j=1;2\nand1\n\u001cL;k=1;2;3, need to be adjusted (see Fig.7).\nHowever, the 40 K and 80 K data are different. While the\nLorentzian part can be fitted as above only with the scattering\nrates1\n\u001cL;k=1;2;3adjusted, the low frequency Drude part has an\nextra hump structure around \u0018100cm\u00001, as seen in Fig.3,6\n600080001000012000 \nDrude band 2 \nDrude band 1 \n ωp,j=1,2 (cm-1)(\na)0\n5 01 001 502 002 503 003006009001200T\n(K)(b) \n1/τD,j=1,2 (cm-1)x\n 5\nFIG. 7: (Color online) (a) The plasma frequency of the Drude bands,\n!p1(black solid square), and !p2(red solid circle). (b) The cor-\nresponding scattering rates, 1=\u001cD1(black solid square), and 1=\u001cD2\n(red solid circle).\nhence can not be fitted by smooth Drude spectra. Neverthe-\nless, it was shown in the previous section that the low fre-\nquency spectra of \u001b1(!;T)for! < 1;000cm\u00001separately\nconserve the spectral weight with temperature variation (see\nthe pink inverse triangle data in Fig.4). Therefore, we tried to\ncontinue to use the same Drude-Lorentz model to fit the 40 K\nand 80 K data.\nFigure 6(B) shows the fitting result of the 40 K data. The\nLorentzian part is fitted well as in the case of T= 130 K,\nconfirming the separate sum rule, but the Drude part show a\nclear deviation. Comparing the experimental data (black solid\nline) and the model fitting (red solid line), it shows that the\nhump structure (extra spectral peak) around !\u0018100cm\u00001\nis formed by draining a spectral density from the lower fre-\nquency part ( ! < 50cm\u00001) of Drude spectra. Combining\nwith the separate sum rule of the lower frequency \u001b1(!;T)\nfor! < 1;000cm\u00001, this hump spectra should be due to a\npartial gapping in the Drude spectra. If it were from an addi-\ntional inter-band transition29, the sum rule should have been\nviolated. Impurity bound state scenario is also unrealistic be-\ncause any potential scattering cannot form a bound state inside\na Drude spectra without first forming a deep gap.\nTo summarize Fig.6, (1) the spectral weight of the hump\nstructure around !\u0018100cm\u00001is drained from the Drude\nspectra with a partial gapping in them; (2) this hump struc-\nture at 80 K and 40 K continuously evolve into the SC gap\nstructure below Tc. We then conclude that the most plausible\nscenario for the hump structure around !\u0018100cm\u00001is due\nto the preformed Cooper pairs.\nIn Fig.7, we show the temperature dependence of the fit-\nting parameters of the Drude spectra of the normal state\noptical conductivity \u001b1(!), i.e., plasma frequencies of the\nDrude bands, !p;i=1;2, and the corresponding scattering\nrates, 1=\u001cD;i=1;2, for all measured temperatures. It is in-TABLE I: Fitting parameters of Lorentz Oscillators\n!L;k=1;2;3TSk 1=\u001cL;k\n(cm\u00001)(K) (cm\u00001) (cm\u00001)\n40K 14224 5700Lorentz #1 2734130K 14224 5744\n40K 16811 8371Lorentz #2 5000130K 16812 10000\n40K 42977 15000Lorentz #3 8500130K 43319 15854\nteresting to note the T-linear scattering rates 1=\u001cD;1(T)\nof the narrow Drude band in Fig.7(b) and the T-linear\nresistivity data \u001aDC(T)in Fig.1, consistently indicating\nthat the optimal doped Ca 8:5La1:5(Pt3As8)(Fe 2As2)5is lo-\ncated near the QCP, as discussed in the Introduction sec-\ntion. A similar behavior was also observed with the op-\ntimally doped Ba 0:6K0:4Fe2As2by optical spectroscopy30,\nindicating the proximity to the QCP. However, the case of\nBa0:6K0:4Fe2As2is an antiferromagnetic QCP30, while our\ncase of Ca 8:5La1:5(Pt3As8)(Fe 2As2)5is a non-magnetic QCP,\nlocated far away from the AFM QCP10. Finally, Table I shows\nthe fitting parameters for the Lorentz oscillators of Fig.6: the\n40K and 130K data.\nD.\u001bexp\n1(!;T Tc)in the normal state, as shown in the\nprevious section, it is natural to assume two s-wave gaps in\nthe SC state. Indeed, the line shape of the 8 K data (black\nsolid line) of \u001bexp\n1(!)appears to have two gaps with different\nsizes, \u00011and\u00012. Therefore, we use the generalized Mattis-\nBardeen formula for two band superconductor to fit the 8 K\ndata of\u001b1(!). The spectral weight of two band and its scat-\ntering rate were taken as fitting parameters for optimal fitting.\nThe results are shown in Fig.8.\nThe black solid line in Fig. 8 shows the SC optical con-\nductivity after subtracting the Lorentz oscillator contribution\nat high frequencies, which is almost independent of tempera-\nture, from the optical conductivity of 8 K. The data shows the\nabsorption edge at about 80 cm\u00001, proportional to the first\nSC gap (\u00182\u00011), and a weak kink at about 220 cm\u00001indicat-\ning a second SC gap ( \u00182\u00012). With the generalized Mattis-\nBardeen model31, the SC gap sizes (the scattering rates) were\ndetermined to be \u00011= 4.9meV (1=\u001c1= 13:1meV ) and\n\u00012= 14.2meV (1=\u001c2= 37:1meV ), respectively. The ra-\ntios of gap magnitude to Tc,R= 2\u0001=kBTc, were evalu-\nated as 3.6 and 10.2, respectively. These mixed values of\nR, compared to the value of RBCS = 3:5, suggest that the\nsuperconductivity in Ca 8:5La1:5(Pt3As8)(Fe 2As2)5could be7\n05 01 001 502 002 503 003 504 004 50050010001500200025003000 \nσ8K - σinterband \nσS (Δ1= 4.9meV, 1/τ1=13.1meV) \nσL (Δ2= 14.2meV, 1/τ2=100meV) \nσS + σL\n51 01 52 02 53 0540056005800600062006400 Ωp,s (cm-1)T\n (K)250260270280290λ (nm) \n σ1 (Ω-1cm-1)F\nrequency (cm-1)\nFIG. 8: (Color online) Mattis-Bardeen model fitting for the 8 K data\n(black solid line) of \u001bexp\n1(!)after the Lorentz oscillator contribution\nsubtracted. The data are decomposed into two Mattis-Bardeen terms\n(blue dashed line, and purple dash dotted line). The sum of the two\nMattis-Bardeen terms is plotted as a red solid line. The inset repre-\nsents the SC plasma frequency \np;S(T)(black solid square) and the\ncorresponding penetration depth \u0015(T)(red open square).\na mixture of weak coupling and strong coupling SC states.\nTo our knowledge, there is not yet an experimental report of\nthe gap sizes of 10-3-8 compound. However, our gap values\nare consistent with the gap values of a typical FeAs-SC com-\npounds with similar Tc; e.g., Ba 0:6K0:4Fe2As2(Tc= 37K;\n\u00011\u00186meV; \u00012\u001812meV )32.\nThe relative sizes of the plasma frequency of each band\nare 1 (small gap band) to 2 (large gap band), consistent with\nthe normal state Drude weights of two band. The scattering\nrate of the small gap band 1=\u001c1= 13:1meV\u0018104cm\u00001\nis slightly larger than the value of 1=\u001cD1\u001875cm\u00001at the\nnormal state, while the scattering rate of the larger gap band\n1=\u001c2= 100meV\u0018800cm\u00001is smaller than the value\nof1=\u001cD2\u00181100cm\u00001; the former is physically not very\nconsistent while the latter is more reasonable. This inconsis-\ntency comes from the difference of two conductivity formula –\nMattis-Bardeen model31and the Drude model – which are not\ncontinuously connected. Considering this, the small inconsis-\ntency of the estimated scattering rates of the narrow band is in\nan understandable range.\nIn the inset, we also calculated the condensation strength\n(the SC plasma frequency \np;S(T)) from the missing spec-\ntral weight and the corresponding penetration depth \u0015(T)us-\ning the London formula \n2\np;S(T) =c2=\u00152(T). Having only\nthree data points below Tc(8K, 17K, and 28K, respectively),\nwe cannot extract much about the temperature dependence of\nthese quantities. However, the overall trend of temperature\nvariation – a flatter behavior at lower temperatures (i.e., 0<\nT T c)\nby the incoherent preformed Cooper pair model with T= 40 K,\noverlayed on the experimental data of \u001bexp\n1(!;T)forTbelow and\naboveTc= 32:8K.\nwave SC gap(s). Also the absolute magnitude of the low tem-\nperature penetration depth \u0015(T= 8K)\u0018250nm\u00180:25\u0016m\nbelongs to the standard range ( 0:2\u0016m<\u0015ab(0)<0:4\u0016m) of\ntypical FeAs-superconductors33.\nE.\u001bexp\n1(!;TcTc) to the 28K,\n17K, and 8K data ( < Tc) of the SC gap structure. For this\nhump structure to be from a magnetic gap, we need a unusual\naccidental coincidence to have the almost same gap energy\nscale both for the magnetic gap and the SC gap.\nIV . CONCLUSIONS\nWe have measured the optical reflectivity R(!;T)of the\noptimally doped Ca 8:5La1:5(Pt3As8)(Fe 2As2)5single crystal\n(Tc= 32.8 K) for frequencies from 40 cm\u00001to 12,000 cm\u00001\nand for temperatures from 8 K to 300 K. In the normal state\nforT=130 K, 240 K, and 300 K, the optical conductivity\ndata\u001b1(!;T)are well fitted by the Drude-Lorentz model with\ntwo Drude forms and three Lorentz oscillators. We have also\nfound that (1) despite a huge variation of the spectral weight\nredistribution with temperature variation, the f-sum rule is\nsatisfied; (2) also, divided by !\u0003\u00181;000cm\u00001, the spectral\nweights above and below !\u0003are separately conserved, whichsuggests that the original bare conduction electrons are split\ninto the low energy itinerant carriers (Drude spectra) and the\nhigh energy localized carriers (Lorentzian oscillators) due to\nthe strong correlation effects.\nIn the SC state for T=28 K, 17 K, and 8 K, the \u001b1(!;T <\nTc)data show a clean gap opening at \u001880cm\u00001and the\nsecond gap structure at \u0018220cm\u00001. The\u001b1(!;T = 8\nK)data is well fitted with two SC gaps, \u00011= 4:9meV\nand\u00012= 14:2meV , consistent with the presence of two\nDrude bands observed in the normal state data. From the es-\ntimated fraction of the condensate spectral weight below Tc,\nwe have estimated that only about \u001829%37of the conduction\nband carriers participates in the SC condensation indicating\nthat Ca 8:5La1:5(Pt3As8)(Fe 2As2)5is in dirty limit.\nMost interestingly, the \u001b1(!;T = 40;80K)data, in the\nintermediate temperature region above Tcbut below\u001880\nK, showed a PG-like hump structure at the exactly same\nenergy scale as the SC gap energy on top of the smooth\nDrude-like spectra. We have demonstrated that this PG-like\nhump structure can be consistently fitted with the preformed\nCooper pair model using the same SC gap values of \u00011;2\nof the SC state. Our work showed that the optimally doped\nCa8:5La1:5(Pt3As8)(Fe 2As2)5is a multigap superconductor\nhaving a clear PG formed by incoherent preformed Cooper\npairs up to 80 K (about three times of Tc), therefore added\none more case to the list of the IBSs with the PG due to the\nSC correlation20,21.\nAcknowledgments\nYSK was supported by the NRF grant\n2015M2B2A9028507 and 2016R1A2B4012672. YKB\nwas supported by NRF Grant 2016-R1A2B4-008758.\nAppendix A: DC limit of conductivity \u001b1(T;!)\nIn Fig.A.1, we show the DC limits of \u001b1(T;!)in the nor-\nmal state for 40, 80, 130, 240, and 300 K, respectively, over-\nlayed with the DC conductivities directly obtained from the\n\u001aDC(T)data of Fig.1 as \u001bDC(T) = 1=\u001aDC(T). The agree-\nment at the DC limit is excellent for all temperatures which\ndemonstrates the quality of our reflectivity data R(T;!)\n(Fig.2) and the faithfulness of the KK-transformation.\nFor the low frequency region below 60cm\u00001, we use the\nHagen-Rubens extrapolation formula RHR(T;!) = 1\u0000\n2q\n2\u000f0!\n\u001bDC, where the only free parameter \u001bDC(T)can be sub-\nstituted by the DC resistivity data of Fig.1 as \u001bDC(T) =\n1=\u001aDC(T). However, to make the best smooth connection\nbetweenRHR(T;!)and our measured data R(T;!), in the\nregion of 40\u000060cm\u00001(our data exist from 40cm\u00001), we al-\nlow some adjustment of the value \u001bDCin the Hagen-Rubens\nformula. The necessary adjustments were less than 10% for\nall temperatures as demonstrated in Fig.A.1, which shows the\nexcellent agreement between \u001b1(T;!!0)and\u001bDC(T) =\n1=\u001aDC(T).9\n02 004 006 008 001 00005000100001500020000\n-100 1 02 03 04 05 05000100001500020000F\nrequency (cm-1)σ1 (Ω-1cm-1) \n C\na8.5La1.5(Pt3As8)(Fe2As2)5σ1 (Ω-1cm-1) \n F\nrequency (cm-1) 40K \n80K \n130K \n240K \n300K \n40K \n80K \n130K \n240K \n300K\nFIG. A.1: (Color online) The same plot of Fig.3, \u001b1(!), but for wider\ny-axis range, to show the DC limit of \u001b1(!). The data are shown only\nfor the normal state for 40, 80, 130, 240, and 300 K, respectively, and\noverlayed with the DC conductivities (solid square symbols) directly\nobtained from the data of Fig.1 as \u001bDC(T) = 1=\u001aDC(T).\n\u0003ykbang@chonnam.ac.kr\nyyskwon@dgist.ac.kr\n1N. Ni, J. M. Allred, B. C. Chan, and R. J. Cava, Proc. Natl. Acad.\nSci.108, 45 (2011)\n2S. Kakiya, K. Kudo, Y . Nishikubo, K. Oku, E. Nishibori, H. Sawa,\nT. Yamamoto, T. Nozaka, and M. Nohara, J. Phys. Soc. Jpn. 80,\n093704 (2011)\n3T. Sturzer, G. Derondeau, and D. Johrendt, Phys. Rev. B 86,\n060516 (2012)\n4A. Sapkota, G. S. Tucker, M. Ramazanoglu, W. Tian, N. Ni, R. J.\nCava, R. J. McQueeney, A. I. Goldman, and A. Kreyssig, Phys.\nRev. B 90, 100504 (2014)\n5C. Loehnert, T. Sturzer, M. Tegel, R. Frankovsky, G. Friederichs,\nand D. Johrendt, Angew. Chem. Int. Ed. 50, 9195 (2011).\n6T. Sturzer, G. M. Friederichs, H. Luetkens, A. Amato, H. H.\nKlauss, F. Winter, R. Pottgen, and D. Johrendt, J. Phys.: Condens.\nMatter 25, 122203 (2013).\n7M. Neupane, C. Liu, S. Y . Xu, Y . J. Wang, N. Ni, J. M. Allred, L.\nA. Wray, N. Alidoust, H. Lin, R. S. Markiewicz, A. Bansil, R. J.\nCava, and M. Z. Hasan, Phys. Rev. B 85, 094510 (2012)\n8I. R. Shein and A. L. Ivanovskii, arXiv:1108.0495 (2011)\n9S. Thirupathaiah, T. Sturzer, V . B. Zabolotnyy, D. Johrendt, B.\nBuchner, and S. V . Borisenko, Phys. Rev. B 88, 140505 (2013)\n10N. Ni, W. E. Straszheim, D. J. Williams, M. A. Tanatar, R. Pro-\nzorov, E. D. Bauer, F. Ronning, J. D. Thompson, and R. J. Cava,\nPhys. Rev. B 87, 060507 (2013)\n11P. Gao , L. Sun , N. Ni , J. Guo , Q. Wu , C. Zhang , D. Gu , K.\nYang , A. Li , S. Jiang , R. J. Cava , and Z. Zhao, Adv. Mater. 26,\n2346 (2014)\n12J. Kim, F. Ronning, N. Haberkorn, L. Civale, E. Nazaretski, N.\nNi, R. J. Cava, J. D. Thompson, and R. Movshovich, Phys. Rev. B\n85, 180504 (2012)\n13E. Mun, N. Ni, J. M. Allred, R. J. Cava, O. Ayala, R. D. McDon-\nald, N. Harrison, and V . S. Zapf, Phys. Rev. B 85, 100502 (2012)\n14T. Zhou, G. Koutroulakis, J. Lodico, N. Ni, J. D. Thompson, R.J. Cava and S. E. Brown, J. Phys.: Condens. Matter 25, 122201\n(2013)\n15K. Cho, M. A. Tanatar, N. Ni and R. Prozorov, Supercond. Sci.\nTechnol. 27, 104006 (2014)\n16Y . Kamihara, T. Watanabe, M. Hirano, and H. Hosono, J. Am.\nChem. Soc. 130, 3296 (2008)\n17X. H. Chen, T. Wu, G. Wu, R. H. Liu, H. Chen, and D. F. Fang,\nNature. 453, 761 (2008)\n18Z. A. Ren, G. C. Che, X. L. Dong, J. Yang, W. Lu, W. Yi, X.\nL. Shen, Z. C. Li, L. L. Sun, F. Zhou, Europhys. Lett. 83, 17002\n(2008)\n19F. C. Hsu, J. Y . Luo, K. W. Yeh, T. K. Chen, T. W. Huang, P. M.\nWu, Y . C. Lee, Y . L. Huang, Y . Y . Chu, D. C. Yan, and M. K. Wu,\nProc. Natl. Acad. Sci. 105, 38 (2008)\n20Y . S. Kwon, J. B. Hong, Y . R. Jang, H. J. Oh, Y . Y . Song, B. H.\nMin, T. Iizuka, S. Kimura, A. V . Balatsky, and Y . Bang, New J.\nPhys. 14, 063009 (2012)\n21Y . M. Dai, B. Xu, B. Shen, H. H. Wen, J. P. Hu, X. G. Qiu, and R.\nP. S. M. Lobo, Phys. Rev. B 86, 100501 (2012)\n22R. P. S. M. Lobo, Y . M. Dai, U. Nagel, T. Room, J. P. Carbotte,\nT. Timusk, A. Forget, and D. Colson, Phys. Rev. B 82, 100506\n(2010)\n23M. A. Surmach, F. Bruckner, S. Kamusella, R. Sarkar, P. Y .\nPortnichenko, J. T. Park, G. Ghambashidze, H. Luetkens, P. K.\nBiswas, W. J. Choi, Y . I. Seo, Y . S. Kwon, H.-H. Klauss, and D.\nS. Inosov, Phys. Rev. B 91, 104515 (2015)\n24M. Tinkham, and R. A. Ferrell, Phys. Rev. Lett. 2, 331 (1959).\n25N. L. Wang, W. Z Hu, Z. G Chen, R. H. Yuan, G. Li, G. F. Chen\nand T. Xiang, Jour. of Phys: Conden. Matt. 24, 294202 (2012).\n26A. A. Schafgans, S. J. Moon, B. C. Pursley, A. D. LaForge, M. M.\nQazilbash, A. S. Sefat, D. Mandrus, K. Haule, G. Kotliar, and D.\nN. Basov, Phys. Rev. Lett., 108, 147002 (2012).\n27B. Xu et al. Chin. Phys. B 23, 087401 (2014).\n28B. H. Min, J. B. Hong, J. H. Yun, T. Iizuka, S. Kimura, Y . Bang,\nand Y . S. Kwon, New J. Phys. 15, 073029 (2013)10\n29E. van Heumen, Y . Huang, S. de Jon, A.B. Kuzmenko, M.S.\nGolden, and D. van der Marel, Europhys. 90, 37005 (2010).\n30Y . M. Dai, B. Xu, B. Shen, H. Xiao, H. H. Wen, X. G. Qiu, C. C.\nHomes, and R. P. S. M. Lobo Phys. Rev. Lett. 111, 117001 (2013).\n31W. Zimmermann, E. H. Brandt, M. Bauer, E. Seider, and L. Gen-\nzel, Physica C 183, 99 (1991).\n32K. Nakayama et al., Europhysics Letters 85, 67002 (2009).\n33R. Prozorov, V . G. Kogan, Rep. Prog. Phys. 74, 124505 (2011).\n34V . J. Emery and S. A. Kivelson, Nature 374, 434 (1995).\n35M. Franz and A. J. Millis, Phys. Rev. B 5814572 (1998).36W. Z. Hu, J. Dong, G. Li, Z. Li, P. Zheng, G. F. Chen, J. L. Luo,\nand N. L. Wang, Phys. Rev. Lett. 101, 257005 (2008).\n37This condensate spectral weight is calculated from the total Drude\nspectral weight of the 40K data (Fig.6(B)) and the missing spectral\nweight of the 8K data (Fig.8). This number 29% is smaller than\nthe number \u001840% which can be read with the data (red circles)\nwith the cutoff frequency 1000cm\u00001in Fig.4(b). This is because\nthe Drude spectra have long tail beyond !>1000cm\u00001." }, { "title": "1505.04770v2.Van_Hove_Singularity_as_the_Driver_of_Pseudogap_Physics_in_Cuprate_High__T__Superconductors.pdf", "content": "Condensation Bottleneck Driven by a Hidden Van Hove Singularity as the Origin of\nPseudogap Physics in Cuprate High- TcSuperconductors\nR.S. Markiewicz, I.G. Buda, P. Mistark, and A. Bansil1\n1Physics Department, Northeastern University, Boston MA 02115, USA\nWe propose a new approach to understand the origin of the pseudogap in the cuprates. The near-\nsimultaneous softening of a large number of different q-bosons yields an ordering bottleneck, wherein\nthe growth of magnetic correlations with decreasing temperature is anomalously slow, leading to\nextended ranges of short-range order. This effect is not tied to the Fermi level, but driven by a\nVan Hove singularity (VHS) nesting that is strongest near a temperature TVHS that scales as the\nenergy separation between the Fermi energy and the energy of the VHS peak. By identifying TVHS\nas the pseudogap onset temperature T∗, we explain many characteristic features of the pseudogap,\nincluding the observed transport anomalies and the termination of the pseudogap phase when EVHS\ncrosses the Fermi level. The condensation bottleneck (CB) provides a new pathway for understanding\nstrong correlation effects in the cuprates. We find that LSCO lies close to an anomalous disorder-free\nspin glass quantum critical point, where the frustration is due to the CB. To study this point we\ndevelop an approach to interpolate between different cuprates.\nPACS numbers:\nI. INTRODUCTION\nEvidence is growing that the ‘pseudogap phase’ found in cuprates is actually home to one or more competing phases,\nincluding a variety of stripe, spin-, or charge-density wave (S/CDW) phases.1-12The CDW phase, in particular,\nhas stimulated considerable interest13-22. Many of these phases, including superconductivity, seem to appear at\ntemperatures well below the pseudogap temperature T∗, so the exact relation between the pseudogap and these other\nphases remains elusive. Indeed, the real puzzle is understanding why the pseudogap phase bears so little resemblance\nto a conventional phase transition. Here we demonstrate that the pseudogap phenomenon arises from a condensation\nbottleneck (CB), where a large number of density waves (DWs) with similar q-vectors attempt to soften and condense\nat the same time, leading to anomalously low transition temperatures and extended ranges of short-range order,\ncharacteristic of the pseudogap phase. We note that many treatments of correlation effects – e.g., path-integral\napproaches – begin by finding a mean-field order parameter at some particular q-vector – a form of random-phase\napproximation (RPA) – and then include the effect of fluctuations. Then if several instabilities compete, the one\nwith lowest free energy wins out. In this case, the CB represents a particularly strong correlation effect, insofar as\none cannot solve the problem one- qat a time, but must account for the q-competition from the start, to account\nfor the entropy associated with the q-manifold. We note that RPA-type approximations extend well beyond simple\nmean-field theories, arising in spin-fermion models and in Hertz-Millis23,24type of quantum critical theories when the\nStoner denominator is approximated by a T-independent Ornstein-Zernicke25form, and in all cases they miss the\nessential and strongly T-dependent mode-coupling physics.\nSimilar effects have arisen in the past. In Overhauser’s theory of CDWs in alkali metals,26theq-vectors for\nall points on a spherical Fermi surface (FS) become unstable simultaneously and cannot be handled one at a time.\nThese effects are often referred to as phonon entropy27, or more generally as boson entropy, and can be analyzed via\nvertex corrections which take proper account of mode coupling28,29. In excitonic theories, the various DW instabilities\nrepresent condensation of some electronic boson at a single q-vector, with triplet excitons corresponding to SDWs and\nsinglet excitons to CDWs30. Summing only bubble and ladder diagrams reproduces the RPA results with BCS-like\nratios of gap to critical temperature, 2∆ /kBTc. Going beyond ladders (e.g., Bethe-Salpeter equation) incorporates\nmode coupling with enhanced ∆ /Tcratios (compare Refs. [31] and [32]; for a review of excitonic insulators, see\nRef. [30]). A particularly interesting analogy is provided by the Bose condensation of excitons. The mean-field\ntransition is found to correspond to the temperature at which excitons are formed. However, since the excitons are\nlocalized in real space, they are greatly spread in q. When fluctuations are included, the real transition lies at a much\nlowerT, when all excitons condense into the lowest qstate. Here we develop a similar theory for the cuprates via a\nself-consistent renormalization calculation of the vertex corrections. We find that the CB not only drives pseudogap\nphysics, but can potentially lead to a novel disorder-free spin-glass phase. Similar effects are likely to be present in\nmany other families of correlated materials.\nWe find that the phase diagram of La 2−xSrxCuO 4(LSCO) is different from that of other cuprates. To explore\nthe transition between them, we simplify the first principles dispersions to equivalent ‘reference family’ dispersions,\ndepending only on the three nearest neighbor hopping parameters, t,t/prime, andt/prime/prime(Supplementary Material Section I).arXiv:1505.04770v2 [cond-mat.supr-con] 28 Mar 20162\nWe find that all cuprates fall along a single cut in t/prime/t−t/prime/prime/t-space, namely t/prime/prime=−t/prime/2, with LSCO characterized by\na smaller value of t/prime.\nII. RESULTS\nA. Bosonic VHS and Origin of the Susceptibility Plateau\nA proper understanding of the CB involves two different issues: firstly, since Fermi surface (FS) nesting typically\nsingles out only a few discrete q-vectors, what causes many q–vectors to soften together? Secondly, a formalism needs\nto be developed to handle the effects of bottleneck. Both of these issues are addressed here. Remarkably, in cuprates\nwe find that the underlying source of strong mode coupling can be traced back to the Lindhard susceptibility χ0.\nIn a typical calculation of classical or quantum phase transitions23,24, material parameters are introduced via χ0,\nwhich is used to define an interacting susceptibility χ. In our calculation, χ0is calculated from density functional\ntheory (DFT) bands corrected by a GW self-energy [Methods Section], and χis the resulting RPA susceptibility. The\nphase transition then corresponds to the vanishing of the denominator of χat frequency ω= 0 for some momentum q.\nTo properly incorporate fluctuation effects, this ‘Stoner denominator’ is then typically reduced to Ornstein-Zernicke\n(OZ)25form,\nχ(q,ω)∼1\nq2+ξ−2+i(ω/ωc)z, (1)\nin terms of various deviations from the critical point (in q,ω, and a ‘tuning parameter’ which is proportional to ξ−2,\nwhereξis the correlation length). Here zis a dynamic exponent and qis measured from the Qwhere the susceptibility\nhas a peak. An important result of our calculation is that in cuprates, this OZ form must be treated carefully, since\nthe coefficients of the deviation parameters can be strong functions of temperature and doping. In particular, we\nfind that the susceptibility inverse curvature (coefficient of q2) can diverge due to a competition between conventional\nFermi-surface nesting and Van Hove singularity (VHS) nesting.\nWhereas many properties of a Landau Fermi liquid are determined by their values at the Fermi level, the suscep-\ntibility is an exception, having both bulk and Fermi surface contributions. While the FS part contributes a ridge\ntoχ0that is a map of the FS at q= 2kF34(wherekFis the Fermi wave vector), in the cuprates there is also an\nimportant bulk contribution, which provides a smoothly varying background, peaking at ( π,π) and giving rise to the\nnear-(π,π)-plateau in the susceptibility. This peak can shift the balance of the FS nesting to q-vectors closer to the\npeaks, and in special cases can lead to commensurate nesting away from the FS nesting vector, at exactly ( π,π).\nMoreover, as Tincreases, coherent FS features are washed out, leaving behind only the commensurate bulk contribu-\ntion. This peak is a bosonic VHS (b-VHS), the finite- qanalog of the Van Hove excitons found in optical spectra35,\nbut present in the intraband susceptibility. However, it is a “hidden” b-VHS. Despite the fact that it is pinned to\nzero energy independent of doping or hopping parameters , it is hidden in the sense that the effective density-of-states\n(DOS) exactly at the b-VHS peak almost always vanishes. Some consequences of this are discussed in Supplementary\nMaterials Section II.\nThe imaginary part of the susceptibility can be thought of as the DOS of electronic bosons, electron-hole (e-h)\npairs, which may become excitons when a Coulomb interaction is turned on. If the renormalized dispersion of a single\nelectron is/epsilon1kwith wave vector k, then an e-h pair at wave vector qhas a dispersion ωq(k) =/epsilon1k+q−/epsilon1k=−2/epsilon1q−(k),\nwhere/epsilon1q±(k) = (/epsilon1k±/epsilon1k+q)/2, and a Pauli blocking factor ∆ fk,q=f(/epsilon1k+q)−f(/epsilon1k). Then the corresponding pair DOS\nisDq(ω) =/summationtext\nk∆fk,qδ(ω−ωq(k)) =χ/prime/prime\n0(q,ω). For LSCO, the dominant pairs are those at q=Q≡(π,π), the AF\nnesting vector. The associated dispersion ωQ(π,π), plotted in Fig. 1(a), resembles the electronic dispersion /epsilon1k, but\nwith an important distinction: it depends only on /epsilon1Q−(k), whereas all of the hopping terms that shift the electronic\nVHS away from half filling ( t/prime,t/prime/prime) are contained in /epsilon1Q+(k), i.e. the b-VHS is pinned at ω= 0. Since χ/prime/prime\n0is an odd\nfunction of ω,χ/prime/prime\n0(Q,ω = 0) = 0. However, excitonic instabilities depend on a Stoner criterion, and hence on\nχ/prime\n0(q,ω= 0) = 2/integraldisplay∞\n0dω/prime\nπχ/prime/prime\n0(q,ω/prime)\nω/prime. (2)\nThe FS contribution thus arises from ω/prime∼0, while bulk contributions arise from peaks in the bosonic DOS, such\nas the b-VHS peak in χ/prime/prime\n0near (π,π). However, while the b-VHS is pinned at ω= 0, Fig. 1(a), its weight vanishes\natT= 0, due to the Pauli blocking factor, ∆ fk,Q= 0 neark= (π,0) atT= 0, Fig. 1(b). Finite Trestores weight,\noptimally near 1000K, although ∆ falways vanishes exactly at ( π,0). In the Supplementary Materials, Section II, we\ndeconvolve the near-( π,π) susceptibility to show that it is a superposition of bulk and Fermi surface features. The3\n1.0\n0.5\n0.0 | ∆f |\n1.0ka(b)\nΓ (π,0) (π,π)1.0\n0.0\n-1.0ωQ (eV)\n3 2 1 0ka(a)\nΓ Γ (π,0) (π,π)\n2000\n1000\n0T (K)\n0.20 0.15 0.10 0.05 0.00\nx0TVHS(π,π−δ)(π,π)\n 5 1(c)\nFIG. 1: Coherent-incoherent crossover at the VHS. (a) Pair dispersion, ωQ, as a function of k, and (b) the corresponding\nweight ∆f, atT= 0K (red), 500K (blue), and 2000K (violet) for LSCO ( x= 0). (c) Phase diagram for LSCO, showing\nVHS-related crossovers and regions where the susceptibility peaks at different q-vectors. Here x0is the doping at T= 0\n(plots are at constant EF), but for qualitative purposes we can assume x/similarequalx0. Dominant fluctuations are at commensurate\n(π,π) (white shaded region) or incommensurate ( π,π−δ) (green shaded region); crossovers are TVHS (red short-dashed line),\nDOS peak (green long-dashed line), coherent-incoherent crossover (pink shaded region), Tγ(violet dot-dashed curve), and the\nposition of the ( π,π) peak vs doping (light blue dot-dashed line) – the approximate electron-hole symmetry point.\nbulk feature is dominated by the b-VHS, while the FS feature smears out with increasing Tin a coherent-incoherent\ncrossover, pink shaded region in Fig. 1(c).\nTo demonstrate the close similarity to VHS effects, we also plot two characteristic features of the VHS. The crossover\nscales with both kBTVHS =EF−EVHS, dark red dashed curve in Fig. 1(c), and with Tγ(violet dot-dashed line), the\ntemperature at which the Sommerfeld constant γ=dS/dT , has a peak. Since the Fermi function evolves smoothly\nwithT, we wxpect sharp Fermi surface features to wash out as Tis increased, but why should this coherent-incoherent\ncrossover move to lower Tas doping is increased? Because a source of entropy, the b-VHS, is moving closer to EF. To\ndemonstrate the connection with entropy, we look at the Sommerfeld constant γ. We calculate γfrom the electronic\ndispersion, assumng a paramagnetic phase, to avoid compications arising from phase transitions. At T= 0,γis\nproportional to the DOS, and hence diverges when the VHS crosses EF. At finite Tthe peak in γrepresents the\nexcess entropy associated with mode coupling. The other curves in Fig. 1(c) will be considered in the Discussion\nSection below.\nB. Strong Mode Coupling Leads to an Extended Range of Short-range Order\nIn conventional mode-coupling calculations33,36-38, the OZ parameters are assumed to be T- and doping-\nindependent, and the resulting physics becomes quite simple. For 2D materials, the Mermin-Wagner (MW) theorem39\nis satisfied, and the mean-field transition at Tmfturns into a pseudogap onset at T∗∼Tmf, with a crossover to long-\nrange order when interlayer coupling is strong enough – in short, not much changes from the mean-field results.\nHowever, when we incorporate realistic susceptibilities we find that this OZ form fails to properly account for the\nstrong mode-coupling effects, leading to dramatically different results. In particular, the entropic effects are encoded\nin a strong T-dependence of the model parameters, which greatly slows down the growth of correlations. This in\nturn leads to an extended T-range of short-range fluctuations, which is typical of pseudogap physics. We find that,\nparticularly at high T, the doping dependence reflects the evolution of the anomalous VH scattering. We note that\nwhile the OZ form is used in quantum critical theory23,24, it is explicitly stated that it is to be used only in a limited\nT-range, and only in the absence of FS nesting24, both of which are violated here.\nMode coupling modifies the bare Coulomb interaction U, producing a vertex-renormalized effective U,Usp= ΓU\nwith Γ = 1/(1 +λ), whileλis found self-consistently from [see Methods section below]\nλ= ΓTA 0/integraldisplay∞\nUcdX−N−(X−)\nX−−Usp. (3)\nwhereA0= 12u/χ 0(Q0,0),χ0(Q0,0) =maxq[χ/prime\n0(q,0)],uis the mode-coupling parameter, and Uc= 1/χ0(Q0,0).\nHere,χ−1\n0, denoted by X−, is the variable of integration and we introduce a corresponding susceptibility density of\nstates (SDOS) N−. In Section III of Supplementary Materials, we show that Eq. 3 is closely related to excitonic Bose\ncondensation, with λproportional to the effective number of bosons.4\n10-1100101102103104105106ξ/a\n2345678\n1002345678\n10002\nT (K)(b)\nIII\nI II4\n3\n2\n1\n0N-\n2.5 2.0 1.5 1.0X- (eV)T= 5K50K250K500K2000K(a)3\n2\n1\n00.94 0.93 0.92\n1 100 ξ/a\n5x10-34321\n1/T (1/K)X2 (c)\nFIG. 2: Structures in LSCO correlation length. (a) SDOS for undoped LSCO at several temperatures; light-blue dotted\nline gives OZ form of SDOS. (b) Corresponding temperature dependence of correlation length ξth. Forx= 0, the red solid line\nwith filled dots is ξthcalculated from the solid lines in (a), while the thin green dot-dashed line is based on the dashed lines in\n(a), showing that ξthis relatively insensitive to the structures away from threshold. In contrast, the blue dashed line is based\non the shape of the SDOS at T= 2000K, but shifted and renormalized to match the SDOS at lower T, illustrating sensitivity\nto the leading edge structure. The black dotted line illustrates the scaling ξth∝T−1/2. (c) Calculated ξthreplotted for x= 0\n(red solid line with filled circles) compared with ξ(blue filled circles) and with experiment (green dot-dot-dashed line)40; the\nblue dotted line is twice ξth.\nFigure 2 illustrates the profound effects that strong mode coupling has in LSCO, as well as the complete inability\nof the OZ approximation to capture this physics. The SDOS, Fig. 2(a), contains VHS-like features characteristic of\nconventional DOSs. However, the singular behavior of Eq. 3 involves only features near threshold, Uc=min(X−),\nwhich evolve strongly with T, see inset to Fig. 2(a). For T > 0, the threshold behavior is always a step at Uc,\nindicative of a parabolic peak in χ0with curvature inversely proportional to the step height. Thus for a qualitative\nunderstanding of the evolution of ξ(Eq. 1) with Twe can assume an OZ form of χ, but with a strongly T-dependent\nstep height A2(T). This allows a threshold correlation length ξth(red solid lines in Figs. 2(b) and (c)) to be defined\nfrom Eq. 1 as the inverse half-width in q, assuming that the curvature ( ∝A−1\n2) isq-independent. In reality, the\ncurvature increases with q, and the correlation length ξobtained from the renormalized susceptibility half-width is\ntypically a factor of 2 larger. Fig. 2(c) compares the x= 0 values of ξth(solid red line) and ξ(filled blue circles)\nvs 1/T; the values of ξlie on the blue dotted curve, which represents twice ξth, and are in good agreement with\nexperiment (green dot-dot-dashed line)40.\nFigure 2(b) illustrates how strong mode coupling slows down the correlation length divergence. Undoped LSCO\n(red solid line) shows two regions I and III of exponential growth of ξthwith decreasing T, separated by an anomalous\nregion II where ξthactually decreases with decreasing T. We will not discuss the MW-like divergence at low T(region\nIII). In the high- Tlimit (region I), the leading-edge parabolic curvature is quite small, and if it were T-independent, as\nin the OZ approximation, the growth in ξthwould follow the blue dashed line (Eq. 11 below), but thermal broadening\ncauses the curvature to decrease with increasing T, leading to the faster growth of the red solid line. The anomalous5\nFIG. 3: t’ dependence of correlation bottleneck. (a)ξthfor the minimal reference family at several values of t/prime/t[see\nlegend] and t/prime/prime= 0. (b-d): SDOS at three values of t/prime/t= -0.21 (b) [as in Fig. 2(b)], -0.27 (c), and -0.35 (d). All curves are\nshifted to line up the ( π,π) data at ∆ X−=X−−X−(π,π) = 0. (e)ξthfor the PA reference family at several values of t/prime/t\n[see legend] and t/prime/prime=−t/prime/2.\nbehavior in region II will be discussed next.\nC. Exploring parameter space\nTo understand the origin of the CB in region II, it is necessary to explore hopping parameter space away from the\nphysical cuprates. To do this, we introduce the notion of reference states: states with simplified hopping parameters\n(onlyt,t/prime, andt/prime/primenonzero) but which match the phase diagram of the real cuprates in a well-defined way [Supplemen-\ntary Material Section I]. This allows us to tune the system between LSCO and Bi2212, and explore phase space beyond\nthese limits. We study two important cuts in t/prime/t−t/prime/prime/tspace, a minimal cut ( t/prime/prime= 0) and the Pavarini-Andersen\n[PA] cut (t/prime/prime=−t/prime/2) – the latter seems to best capture the physics of the cuprates. By tuning t/primewe unveil the origin\nof the CB as a localization-delocalization crossover tied to the crossover from ( π,π)- to FS-nesting. As a byproduct,\nwe gain insight into why LSCO is so different from other cuprates, and how cuprates evolve from the pure Hubbard\nlimit (t/prime= 0).\nFigure 3(a) shows the T-evolution of ξthfor several values of t/primealong the minimal cut, including the data of Fig. 2(b).\nFort/prime/t >−0.17, the system is characterized by commensurate (C) ( π,π) order with a finite Neel temperature\nTN∼1000K [the correlation length grows so rapidly that interlayer correlations will drive a transition to full 3D\norder]. Similarly, for t/prime/t<−0.345 there is incommensurate (I) ( π,π−δ) order with TNabout a factor of 10 smaller.\nBut for intermediate t/prime/tthe C-I transition is highly anomalous, with correlation length orders of magnitude smaller\nthan expected. Figure 3(b) shows that a similar evolution follows along the PA cut in parameter space. The reason\nfor this anomalous behavior can be seen by looking at the leading edge SDOS in the crossover regime, shown for three\nvalues oft/primealong the minimal cut in Figs. 3(c)-(e). For ease in viewing, these curves have been shifted to line up\nthe SDOS at ( π,π) at allT. It is seen that the anomalous collapse of ξthis associated with a rapid growth of the\nstep height A2, culminating in a near-divergence at t/prime\nc=−0.27t, where the leading edge curvature goes to zero. This\ndivergence coincides with the C-I crossover of the leading edge SDOS [Supplementary Materials Section I]. Here many\ndifferentq-vectors compete simultaneously, frustrating the divergence of any particular mode. This is the electronic\nanalog of McMillan’s phonon entropy: if many phonons are simultaneously excited, the transition is suppressed to\nanomalously low temperatures. Note that ξthdrops by 9 orders of magnitude at T= 200Kwhent/prime/tchanges from\n-0.17 to -0.27, then grows by a similar amount at 100K when t/prime/tchanges from -0.27 to -0.345. [Over this same\nrange, the true ξwill be frozen at the value corresponding to the half-width of the ( π,π) plateau.] The green shaded\nregion in Fig. 3(c) shows that the range of the anomalous growth II of ξin Fig. 3(a) coincides with the range of rapid6\ngrowth of the SDOS leading edge; such behavior is absent if a T-independent OZ form (black dashed line) is assumed.\nThe divergence of the SDOS at T= 0 signals that the system at t/prime=−0.27tis highly anomalous, with an infinitely\ndegenerate ground state . This represents an anomalous form of spin glass arising in the absence of disorder, with the\nfrustration arising from strong mode coupling. Note also that when the C-I transition is at T > 300K,ξ(T) has a\nsharp downward cusp at the transition, while above the transition ξ T∗, the resistivity ρvaries linearly with T,9behavior expected\nnear a VHS46. For lower T, the resistivity is mixed, but ρ∼T2, as expected for a coherent Fermi liquid, is found\nbelow aTcoh< TVHS. This picture bears a resemblance to the Barzykin-Pines model of the cuprate pseudogap,47\nidentifying TVHS andTVHS/3 withT∗andT∗/3 in their model. The underlying physics of the incoherent-to-coherent\ncrossover in their model is related to Kondo lattice physics48, with the VHS peak standing in for the Kondo resonance\n(see Supplementary Materials Section V). This raises the question of whether a similar mode-coupling calculation in\nheavy-fermion compounds could lead to a similar anomalous entanglement at the f-electron C-I transition.7\nThe excitonic instability should be maximal when the susceptibility is approximately electron-hole (e-h) symmetrical\nin doping, and at T= 0 this happens when the VHS is at the Fermi level.34At finiteT, the e-h symmetrical point\ncontinues to coincide with a peak in the ( π,π) susceptibility at doping xp, butxpshows a remarkably rapid evolution\nwithTtowards half-filling (light-blue dot-dashed line in Fig. 1(c)), nearly coinciding with the coherent-incoherent\ncrossover. We interpret this temperature dependence as follows: The parameters t/primeandt/prime/primeare relevant perturbations\nshifting the VHS from the pure Hubbard model value at x= 0 where t/prime=t/prime/prime= 0. When kBT >|t/prime|, these\nperturbations become irrelevant, so that the effective VHS becomes electron-hole symmetric at x= 0. Lastly, the\nlower branch of the ( π,π)−(π,π−δ) commensurate-incommensurate transition (green shaded region in Fig. 1(c)) also\nscales, following TVHS very closely in this simple ( t−t/primeonly) model. Indeed, at T= 0, this transition corresponds\nto the Fermi energy falling off of the ( π,π) plateau, providing another indication that the effective VHS is shifting\ntowards half-filling as Tincreases. In Supplemental Material Section II we show that this is another consequence of\nPauli blocking.\nThe rapid thermal evolution of the ( π,π)-VHS peak should be contrasted with the much smaller change in conven-\ntional VHS effects found near Γ. In particular, the DOS peak is dominated by near-FS physics, and hence displays a\nmuch weaker doping dependence, green long-dashed line in Fig. 1(c).\nAn additional consequence is that in cuprates with small t/primehopping, such as LSCO, the anomalous VHS suscep-\ntibility can dominate even at T= 0, leading to strong deviations from FS-nesting. Thus, in most of Fig. 1(c), the\npeak susceptibility is associated with VHS nesting at q= (π,π), and only in an intermediate T-regime is FS nest-\ning is found at ( π,π−δ) (green shaded region). The resulting VHS-FS nesting competition plays a strong role in\nunderdoped LSCO, associated with retrograde correlation length change with T, region II of Fig. 2(b), which is a\nsignal of proximity to a novel disorder-free spin-glass QCP. Notably, in LSCO commensurate ( π,π) order disappears\nrapidly by∼2% doping, being replaced by incommensurate magnetic fluctuations and low- Tspin-glass effects, while\nneutron scattering has found that doped LSCO is close to a magnetic QCP.49The very different situation in most\nother cuprates is discussed in Sections I, IV of Supplementary Materials.\nB. Strong coupling physics\nWe recall that our self-energy formalism50is able to reproduce most spectral features of the insulating cuprates\nin terms of a ( π,π) ordered phase. We reproduce not only the photoemission dispersions, limited to the lower\nHubbard bands, but optical and x-ray spectra which depend sensitively on the Mott gap. In a related 3-band model,\nwe reproduced the Zhang-Rice result that the first doped holes are predominantly oxygen character, and our overall\ndispersions at half filling agree with [subsequent] DMFT results at least as well as DMFT results from different groups\nagree. [See further Supplementary Materials VI.] The problem with our earlier calculations is that they predict long-\nrange (π,π) AFM order at too high T. The new self-consistent renormalization calculations have only short range\norder, in which case the upper and lower Hubbard band dispersion is reproduced, with a broadening ∼1/ξ.33\nAn antiferromagnet is considered to be weakly coupled if the Neel temperature TNincreases with U, and strongly\ncoupled ifTNdecreases with increasing U[TN∼J= 4t2/U]. Since this is a finite Tcriterion, it is sensitive to the\nboson entropy effects we have been discussing. Indeed, for large Uthe system is nearly localized and the physics\nis again reminiscent of Bose condensation. The simultaneous softening of many magnetic modes can be readily\ndemonstrated from a simple Hartree-Fock model of an antiferromagnet. For arbitrary q, the resulting gap for large U\nis\n∆E=/radicalBig\nU2+/epsilon12\n−/similarequalU+J˜/epsilon12\n−, (4)\nwhere/epsilon1−= (/epsilon1k−/epsilon1k+q)/2 =−2t˜/epsilon1−. Thus, the difference in energy between any two spin configurations is a quantity\nof orderJ, so that when T∼J, the system gains entropy by mixing different q-states, and long-range AF order is\ndestroyed.33While this effect can be recognized in HF, only a theory that properly accounts for the mode competition\ncan resolve it. The present mode coupling model properly accounts for this effect via the vertex renormalization\nfactorλ, which is an integral over the RPA χ. This can be seen from Fig. 2(c), where the measured ξin LSCO\n(green dot-dot-dashed line)40is compared to our calculation (filled blue dots). The experimental data are shown only\nabove 300K, since at lower Tinterlayer coupling drives a transition to long-range order. In the Heisenberg model,\nξis a function only of T/J, and these data have been used to measure the exchange J. Hence, our mode coupling\ncalculation successfully reproduces this strong coupling result, as was found earlier for electron-doped cuprates33. The\nabove results suggest that the C-I, incoherent-coherent transition is also a localization-delocalization transition, with\nthe mode coupling associated with small effective hopping.\nFor more insight into the strong coupling limit, we note that in the two-particle self-consistent approach51,λis8\ndetermined by a sum rule involving double occupancy, which is fixed by assuming\nUsp/U=/(n/2)2,\nwhich leads to a saturation of UspasU→∞ (or∼1/U). In our calculation, this saturation arises naturally,\nsinceUspcan never exceed Uc.\nQuantum critical points (QCPs) in strongly correlated materials are often discussed in terms of deconfined QCPs\n(DQCPs), involving a non-Landau transition between two types of competing order, where a new form of excitation\nemerges exactly at the DQCP. This has been refined for cuprates into an underlying competition between Mott physics\nand Fermi liquid physics, masked by a low-energy order parameter of the FL52. This is an apt description of the\ncurrent results, with the Mott physics evolving into a low- TAFM and the FL to a spin-density wave (SDW), and a\nspin glass at the DQPT. This confirms the finding from dynamical cluster approximation calculations of the strong\nrole of the VHS in Mott physics.53\nC. Conclusions\nIn conclusion, we find that the pseudogap is driven by tendencies toward magnetic order at or near q= (π,π).\nStrong mode coupling effects drive the characteristic anomalies of the pseudogap, in particular the existence of a broad\ndoping- and temperature-range of only short- to intermediate-range order. Finally, the underlying cause of the strong\nmode coupling is localization in particular the fact that the extended-to-localized transition is not a simple crossover\nbut a competition between FS-dominated and non-FS dominated (VHS-dominated) physics. This competition across\na manifold of q-states leads to a condensation bottlneck, which requires a new formalism for dealing with multi- q\nmode softening. The close similarity to heavy Fermion physics should be noted.\nThis can be restated slightly differently. The original ( t-only) Hubbard model is particularly difficult to solve,\nas three separate instabilities are simultaneously present at half filling: the Mott instability, FS nesting, and VHS\nnesting. A finite t/primeshifts the latter two to finite doping. For large |t/prime|the three instabilities become well separated,\nbut for small|t/prime|the two nesting instabilities overlap and compete, leading to frustration and pseudogap physics.\nSignificantly, we approach the problem from the intermediate-coupling side, suggesting that the full crossover from\nhalf-filling to large doping could be explored.\nIn turn, this sheds light on strong correlation effects in the cuprates, in particular Mott vs Slater physics. The\nlocalization associated with Mott physics has several manifestations. One is that coherent hopping is restricted to\nshorter range. That is consistent with stronger effects found at smaller t/prime. Particularly for Mott physics, a second\naspect is the breakdown of hybridization. The Cu-O hybridization in the cuprates spoils Mott physics [half filling does\nnot imply one electron on each Cu]. Zhang-Rice singlet formation can be considered as a breakdown of hybridization\n[pure oxygen states at the top of the lower magnetic band, pure Cu at the bottom of the upper magnetic band],\nand is found more generally for ( π,π) AFM fluctuations near half filling in a three-band model50. Hence the strong\nevolution with t/primefrom localized to extended physics, with only LSCO close to Mott physics.\nRice et al.54recently noted: “The need for theoretical methods to handle ... short range correlations ... is a key\nchallenge for the future.” Along this line, the present approach appears to explain a number of anomalous features\nassociated with the pseudogap, including a coherent crossover associated with a competition between two DW orders\n(see Supplementary Materials Section IV). In fact, we have incorporated an important ingredient for strong-coupling\ncalculations: a model of short-range AF order from which a t−Jmodel can be derived.\nIn Anderson’s RVB picture, he envisaged a regime of Mott physics where the FS played a neglible role; instead,\nmost experimental studies find clear evidence for a well-defined FS, with competing phases associated with FS nesting,\nand a QCP associated with FS reconstruction. Our results have a strong bearing on these competing scenarios: in\nparameter space ,t/primeis a relevant parameter tuning the system away from quasilocalized physics near t/prime= 0, where the\nFS has a negligible role, into a delocalized regime dominated by FS physics. The crossover is marked by a regime of\nstrong disorder, where the correlation length remains small down to very low T. Most families of cuprates lie on the\ndelocalized side of this crossover, and hence are most easily understood from a FL-type picture. LSCO appears to\nbe on the localized side of the crossover, but close to it, consistent with experiment.49We note that the disorder line\nmay be hard to directly access: since ordering transitions tend to lower the free energy, the line will be a high-energy\nstate with low energy states to either side of it, suggesting an instability to nanoscale phase separation.\nThe VHS has been predicted to play a significant role in many materials, particularly in lower dimensional systems\nwhereχdiverges, but clear evidence for this remains sparse. Thus, VH nesting was introduced as a possible cause\nof CDWs in dichalcogenides55, but this interpretation remains disputed. Even in VO 2, the striking metal-insulator\ntransition has been found to be driven by large phonon entropy.56Clearly, one problem is that the VHS has been\nassumed to play a role only when it is near the Fermi level, whereas Fig. 1(c) demonstrates that its influence extends9\nover a much wider doping range. For those who think condensed matter physics lies on the surface of the Fermi sea,\nthe VHS is the iceberg, lurking.\nIt will be interesting to examine how the present results are modified by effects of disorder and interlayer coupling.\nAn important issue is the extent to which the total number of QPs is conserved. More specifically, what is the relation\nbetween the number of incoherent electrons (1 −Z)Nand the effective number of excitons Neff? Proximity to a\nspin glass phase could explain the failure of recent attempts to create an ‘artificial cuprate from related nickelate\ncompounds, and may suggest a path to improved analog states.\nIV. METHODS\nOur calculation is a form of many-body perturbation theory (MBPT) based on Hedin’s scheme. The scheme\ninvolves four elements: electrons are described by Green’s functions Gwith DFT-based dispersions renormalized by a\nself-energy Σ; electronic bosons [electron-hole pairs] are described by a spectral weight [susceptibility] χrenormalized\nby vertex corrections Γ. Neglecting vertex corrections, the self energy can be calculated as a convolution of Gand\nW=U2χ, the GW approximation. This approach has been used to solve the energy gap problem in semiconductors,\nwhere the Γ correction leads to excitons via the solution of a Bethe-Salpeter equation, and in extending DMFT\ncalculations to incorporate more correlations [e.g., DMFT+GW, etc.]. Our approach here is to extend our previous\nGW calculations [quasiparticle-GW or QPGW50] to include vertex corrections.\nIn QPGW we introduce an auxiliary function GZ=Z/(ω−/epsilon1QP\nk), where the dressed, or QP dispersion is /epsilon1QP\nk=\nZ/epsilon1DFT\nk, and/epsilon1DFT\nk is the bare, or DFT dispersion. GZbehaves like the Green’s function of a Landau-type QP –\na free electron with renormalized parameters that describes the low-energy dressed electronic excitations. However,\nthis is a non-Fermi liquid type QP, since the frequency-integral of Im(GZ) isZand not unity. That is, the Z-QP\ndescribes only the coherent part of the electronic dispersion, and is not in a 1:1 correspondence with the original\nelectrons. The importance of such a correction can be readily demonstrated. Since a Z-QP has only the weight Z\nof a regular electron, the susceptibility [a convolution of two Gs] is weaker by a factor of Z2than an ordinary bare\nsusceptibility. To match this effect in the Stoner criterion requires introducing an effective Ueff=ZU. In contrast,\nMBPT calculations in semiconductors typically set the GW-corrected Green’s function to G−1\nGW=ω−/epsilon1QP\nk, where/epsilon1QP\nk\nis the average GW-renormalized dispersion57, thereby missing the reduced spectral weight of the low-energy, coherent\npart of the band.\nWe work in a purely magnetic sector, with a single Hubbard Ucontrolling all fluctuations; there is a competition\nin this model between near-nodal (NNN) and antinodal nesting (ANN) which mimics the SDW-CDW competition in\ncuprates, sharing the same nesting vectors22,34.\nThe self-consistent parameter λis found from a Matsubara sum of the susceptibility\nλ=A0T\nN/summationdisplay\nq,iωnχ(q,iωn), (5)\nwhereNis the number of q-points,\nχ(q,iωn) =χ0(q,iωn)\n1 +λ−Uχ0(q,iωn), (6)\nand the summation in Eq. 5 can be transformed:\nT\nN/summationdisplay\nq,iωnχ(q,iωn) =/integraldisplayd2qa2\n4π2/integraldisplay∞\n0dω\nπcoth(ω\n2T)χ/prime/prime(q,ω+iδ)\n/similarequalˆλ1+ˆλ2T, (7)\nwithathe in-plane lattice constant,\nˆλ1=/integraldisplayd2qa2\n4π2/integraldisplay∞\n0dω\nπχ/prime/prime(q,ω), (8)\nˆλ2=/integraldisplayd2qa2\n4π2χ/prime(q,0). (9)\nThe term ˆλ1introduces a small, nonsingular correction33to Eq. 5, which we neglect.10\nFIG. 4: Origin of SDOS features. (a) Inverse susceptibility χ−1\n0(q) for Bi2201 at doping x= 0. Note that Uc=min(χ/prime−1\n0)\n(circled region). Curves are at T= 10 (red) and 100K (blue). (b) Corresponding susceptibility density of states (SDOS)\nN−(X−), plotted horizontally, corresponding to the susceptibility of (a). (c,d) Blowups of circled regions in (a,b).\nIn order to explore the role of susceptibility plateaus for realistic band dispersions, the self-consistency equation is\nevaluated numerically. For this purpose, we note that since χ−1\n0has dimensions of energy, a plot of χ−1\n0(q,0) resembles\na dispersion map. Hence we can define a SDOS:\n/integraldisplayd2qa2\n4π2=/integraldisplay\nN−(X−)dX−. (10)\nFigure 4 illustrates how integrating over the inverse susceptibility, Fig. 4(a), leads to the SDOS, Fig. 4(b), at x=\n0 where the NNN plateau is dominant. Here we use hopping parameters appropriate to the DFT dispersion of\nBi2Sr2CuO 6+x(Bi2201),t= 419.5,t/prime=−108.2 andt/prime/prime= 54.1 eV (Supplementary Materials Section IV). This yields\na doping phase diagram that is qualitatively similar to that of most cuprates, except LSCO. By comparing Figs 4(a)\nand 4(b), one can see how features in χ−1\n0translate into features in N−. Thus, the intense, flat-topped peak in N−at\nsmall values of χ−1\n0represents the NNN plateau. Its broad leading edge (smaller χ−1\n0) is controlled by anisotropy of\nthe plateau edge between ( π,π−δ) and (π−δ,π−δ), while the sharp trailing edge corresponds to the local maximum\nofχ−1\n0at (π,π). For the ANN peak, its leading edge scarcely leaves any feature in N−, but a local maximum translates\ninto a large peak in N−. Finite temperature, T= 100K(blue line) rounds off the cusp in χ−1\n0, Fig. 4(c), leading to a\nstep inN−, Fig. 4(d), but otherwise has little effect.\nOnceN−has been calculated, Eq. 3 can be evaluated numerically. The singular part of the integral is treated\nanalytically, and the remainder numerically, with uandUapproximated as constants, u= 0.8eV−1, whileU= 2 eV\nunless otherwise noted. It is convenient to fix the SDOS at some temperature T/prime, and then solve Eq. 3 for T(ξth,T/prime),\nwith self-consistency requiring T(ξth,T/prime) =T/prime.\nWhen the OZ form of χis assumed, N−becomes a constant, which we denote Na, and Eq. 3 leads to long range\norder atT= 0 only, with correlation length ξth/a= 1/√4πNaδgiven by\nξthqc=eT2/T, (11)\nwhereδ=Uc−Usp,Uc= 1/χ0(Q0,0),T2=πA2λ/6Γua2,Aiis the coefficient of qi, andqcis a wave number cutoff.\nThe origin of the anomalous region II for undoped LSCO, Fig. 2(b), can be readily understood from Fig. 2(a),\nwhere forT < 1500Kthere is excess SDOS weight near Uc, leading to a strong peak (inset) as T→0. This feature\nrepresents the development of the ( π,π)-plateau. For the small- t/primematerials, the susceptibility on the plateau remains\nparabolic, but with a small curvature and a sharp cutoff. This leads to an additional contribution to the SDOS of the\nformN−=NpifX−≤Xp. Then the integral of Eq. 3 becomes\nI= (Na−Nbδ)ln(Xc\nδ+ 1) +NbXc+Npln(Xp\nδ+ 1). (12)\nwhereXcis the maximum of X−. The last term, which we denote Ip, has two distinct limits. If Xp<< δ , then\nIp→NpXp/δ. This is equivalent to assuming that the plateau has a flat top, in which case N−can be represented as11\naδ-function,N−=NpXpδ(X−−Uc), withNpXpthe excess height of Nat threshold, proportional to the area of the\nplateau atT= 0. AsTdecreases,δalso decreases, and the opposite limit for Ipbecomes appropriate, δ <