content
stringlengths 15
7.12M
| publicationDate
stringlengths 10
10
| abstract
stringlengths 41
2.5k
| title
stringlengths 14
230
| id
stringlengths 11
12
|
---|---|---|---|---|
Squeezed driving induced entanglement and squeezing among cavity modes and magnon mode in a magnon-cavity QED system Ying Zhoua,b, Jingping Xua,, Shuangyuan Xiea,, Yaping Yanga aMOE Key Laboratory of Advanced Micro-Structured Materials, School of Physics Science and Engineering, Tongji University, Shanghai 200092, China bSchool of Electronics and Information Engineering, Taizhou University, Taizhou, Zhejiang, 318000, China Abstract We propose a scheme to generate entanglement between two cavity modes and squeeze magnon mode in a magnon- cavity QED system, where the two microwave cavity modes are coupled with a massive yttrium iron garnet (YIG) sphere through magnetic dipole interaction. The nonlinearity used in our system originates from a squeezed driving via parametric down-conversion process, which is the reason to cause entanglement and squeezing. By using the mean eld approximation and employing experimentally feasible parameters, we demonstrate that the system shows zero entanglement and squeezing without squeezed driving. Meanwhile, our QED system denotes that the entanglement between squeezed cavity mode and magnon mode can be transferred to the other cavity mode and magnon mode via magnon-cavity coupling interaction, and then the two cavity modes get entangled. A genuinely tripartite entangled state is formed. We also show that magnon mode can be prepared in a squeezed state via magnon-cavity beam-splitter interaction, which is as a result of the squeezed eld. Moreover, we show that it is a good way to enhance entanglement and squeezing by increasing the nonlinear gain coecient of squeezed driving. Our results denote that magnon-cavity QED system is a powerful platform for studying macroscopic quantum phenomena, which illustrates a new method to photon-photon entanglement and magnon squeezing. Keywords: Squeezed driving, Nonlinearity gain, Entanglement, Squeezed 1. Introduction In recent years, yttrium iron garnet (YIG) material, as an excellent ferrimagnetic material with high spin den- sity(about 4 :2210 27m 3) and low dissipation rate(about 1 MHz), has attracted considerable attention [1, 2, 3]. Moreover, YIG material is ferromagnetic at both cryogenic [4, 5, 6] and room temperature [7] because its Curie tem- perature is about 559K. The magnon mode, as a collective motion of a large number of spins with zero wavevector (Kittel mode [8]) via the Holstein-Primako transforma- tion [9] in YIG sphere, possesses unique properties. It can realize strong [4, 5, 6, 7, 10, 11] and ultrastrong [12, 13] coupling to microwave cavity photons at either cryo- genic or room temperature, and then lead to magnon- cavity polaritons. Thus, a lot of meaningful development about magnons is found, including the observation of cav- ity spintronics [10, 14], bistability [15], magnon gradient memory[16], magnetically controllable slow light [17], level attraction[18], magnon-induced transparency [19, 20], and the magnon squeezed state [21]. It is noted that magnon Corresponding author Email addresses: xx_ jj_ pp@tongji.edu.cn (Jingping Xu), xieshuangyuan@tongji.edu.cn (Shuangyuan Xie)squeezed state is an important macroscopic quantum state, which can be used to improve the measurement sensitiv- ity [22] and study decoherence theories at large scales[23]. Meanwhile, by virtue of strong coupling among magnons, other interesting phenomena, including coupling the magnon mode to a single superconducting qubit [24], to photons and to phonon mode [20, 25], have also been studied. This of- fers a possibility to enable coherent information transfer between dierent information carriers. Clearly, compared to atom, the size of YIG sphere is in mesoscopic or macro- scopic scale usually with a size of 250m, which holds the potentiality for implementing quantum states, espe- cially the entanglement in more massive object. Thus, it provides a promising and completely new platform for the study of macroscopic quantum phenomena [25], which is a key step to test decoherence theories at macroscopic scale [23, 26], and probe the boundary between the quan- tum and classical worlds [27, 28, 29]. In the microwave region, one important quantum state is entangled state, which is typically produced by exploiting the nonlinearity of magnetostrictive interaction in cavity magnomechani- cal system[30], by utilizing Kerr nonlinearity results from magnetocrystalline anisotropy[31], and by using the non- linearity of quantum noise in Josephson parametric am- pliers (JPA) [32]. Meanwhile, another important macro- Preprint submitted to Physics Letters A January 25, 2022arXiv:2201.09154v1 [quant-ph] 23 Jan 2022scopic quantum state is magnon squeezed state, which is usually generated by the quantum noise of JPA process[21]. Recent interest has focused on generating entangle- ment and squeezing in a hybrid cavity magnon QED sys- tem, especially in a hybrid cavity magnomechanics sys- tem including phonons. A genuine tripartite entangle- ment is shown by using the nonlinearity of magnon-phonon coupling in a cavity magnomechanical system consisting of magnons, microwave photons and phonons [33], where the magnons couple to microwave photons and phonons via magnetic dipole interaction and magnetostrictive in- teraction, respectively. When driving the above cavity (Ref. [33]) by a weak squeezed vacuum eld generated by a
ux-driven JPA process, the magnons and phonons are squeezed in succession, and larger squeezing could be real- ized by increasing the degree of squeezing of the drive eld and working at a lower temperature [21]. A hybrid cav- ity magnomechanical system includes two magnon modes in two macroscopic YIG spheres, which couple to the sin- gle microwave cavity mode via magnetic dipole interaction. By activating the nonlinear magnetostrictive interaction in one YIG sphere, realized by driving the magnon mode with a strong red-detuned microwave eld, the two magnon modes get entangled [34]. When two YIG spheres are placed inside two microwave cavities driven by a two-mode squeezed microwave eld. Each magnon mode couples to the cavity mode via magnetic dipole interaction. The quantum correlation of the two driving elds can eciently transferred to the two magnon modes and magnon-magnon entanglement can be achieved. The two cavity modes also can entangle to each other [35]. When considering the vibrational modes in the above cavity (Ref. [35]), each phonon mode couples to the magnon mode via magne- tostrictive interaction. By directly driving magnon mode with a strong red-detuned microwave eld to active the magnomechanical anti-stokes process, and further driving the two cavities by a two-mode squeezed vacuum eld as above scheme (Ref. [35]), the two phonon modes in two YIG spheres can also get entangled [36]. All the above solutions denote that magnetic dipole coupling interaction and nonlinearity are two main elements to produce entan- gled and squeezed states. The main nonlinearity used is the nonlinearity of magnetostrictive interaction[34, 30, 25, 33] and the nonlinearity generated by quantum noise of JPA process[35, 32] in cavity magnon system. In this letter, we propose a scheme to generate photon- photon entanglement and squeeze magnon in a magnon- cavity QED system. Magnon mode in a YIG sphere is coupled to two microwave elds via magnetic dipole in- teraction, respectively. Since YIG material that generates magnons is a massive object, it is considered to be a the- oretically innovation to realize the entanglement of two mesoscopic objects through a cavity mode. However, for the cavity mode or photon being a good carrier of infor- mation, we hope to entangle the two cavity modes with the magnons as a mesoscopic medium, and we think this is more important from the perspective of information.Squeezing is also a very important quantum resource, and we then emphasize the squeezing of magnons. It is found that squeezing can be transmitted to various objects in this QED system. Dierent from previous propose, the nonlinearity we used is generated by parametric down- conversion of JPA process. The intensity of that is
exible tunable, resulting in a squeezed cavity mode. Meanwhile, for the phonon mode can provide another non-linearity and make the system into a more complex one, we did not take it into consideration. In our QED system, entan- glement transferred from magnon-cavity a1subsystem to magnon-cavity a2subsystem, and then transferred to two cavity modes subsystem. A genuinely tripartite entangled state is formed. Meanwhile, the squeezed driving also pre- pares the magnon mode in a squeezing state. Further, we show that increasing the nonlinearity gain coecient of squeezed driving is a good way to enhance entanglement and squeezed. Moreover, we show that the optimal entan- glement and squeezing generated when the coupling rates between the two cavity modes and magnon mode are the same. 2. The model We consider a hybrid magnon-cavity QED system, which consists of two microwave cavity modes and a magnon mode, as depicted in Fig.1. A squeezed microwave cavity 1 (with frequency !1) is implemented by parametric down- conversion in JPA process. We assume that the nonlinear gain coecient of JPA is . The second microwave cav- ity (with frequency !2) is perpendicular to the microwave cavity 1 without any non-linearity driving, the resonance frequency!2is close to that of cavity mode a1. To achieve strong couplings between the YIG sphere and these two cavity modes, we place the YIG sample at the center of both cavities. Meanwhile, the magnons are quasiparticles, a collective motion of a large number of spins spatially uniform mode (Kittel mode [8]) in a massive YIG sphere. The magnetic eld of cavity mode a1anda2are alongx andydirection, respectively. The bias magnetic eld H is alongz-axis for producing the Kittel mode. Strongly coupled is implemented via magnetic dipole interaction. Moreover, a microwave eld with angular frequency !0 and Rabi frequency "pis applied along the xdirection to drivinga1. We assume the size of the YIG sample to be much smaller than the microwave wavelengths in our QED system, so the radiation pressure on YIG sample induced by microwave elds can be neglected. The Hamiltonian of the system reads H=h=X j=1;2!jay jaj+!mmym+X j=1;2gj(ay j+aj)(m+my) +"p(a1ei!0t+ay 1e i!0t) + (a2 1e2i!0t+ay 12e 2i!0t)(1) whereajanday jare, respectively, the annihilation and creation operators of cavity mode j. m(my) is annihi- 2lation (creation) operator of magnon mode [37], which represent the collective motion of spins via the Holstein- Primako transformation [9] in terms of Bosons, satisfy- ing [O;Oy] = 1 (O=a1;a2;m).!j(j= 1;2) and!m present the resonance frequency of cavity modes ajand magnon mode, respectively. The frequency of magnon mode can be adjusted by the external bias magnetic eld Hvia!m=
H, where
=2= 28GHz/T is the gyromag- netic ratio. gjdenotes the linear coupling rate between magnon mode and cavity mode aj, which currently can be (much) larger than the dissipation rates jandm of cavity mode ajand magnon mode, i.e. gj> j,m (j= 1;2). It denotes the magnon-cavity QED system is in the strong coupling regime, but not in the ultrastrong coupling regime, and the rotating-wave approximation can be applied for the magnon-cavity interaction terms in our magnon-cavity QED system. Fig. 1. Schematic of magnon-cavity QED system. The rst cav- ity is driven by a microwave eld with "pthe Rabi frequency and a squeezed eld with the gain coecient of parametric down- conversion, the resonance frequency of which is !1. The second cavity (with frequency !2) is perpendicular to the rst one with a close angular frequency. The magnetic eld of cavity mode a1anda2 are alongxandydirection, respectively. A YIG sphere is mounted at the center of the both microwave cavities. Simultaneously, it is also in a bias magnetic eld Halongz-axis for producing the Kittel mode, resulting in the resonance frequency !m. Here,1,2and mare the dissipation rates of cavity mode a1, cavity mode a2and magnon mode, respectively. Under the rotating-wave approximation, the magnon- photon interaction term gj(aj+ay j)(m+my) becomesgj(ajmy+ ay jm). We then switch to the rotating frame with respect to the driving frequency !0, the Hamiltonian of the system can be written as: H=h=X j=1;2jay jaj+ mmym+X j=1;2gj(ay jm+ajmy) +"p(a1+ay 1) + (a2 1+ay 12) (2) Where j=!j !0, and m=!m !0are the detunings of cavity mode j and magnon mode, respectively. By including input noises and dissipations of the system, the quantum Langevin equations describing the system areas follows, _a1= (i1+1)a1 ig1m i"p 2i ay 1+p 21ain 1(3) _a2= (i2+2)a2 ig2m+p 22ain 2 (4) _m= (im+m)m ig1a1 ig2a2+p 2mmin(5) Whereain jandminare input noise operators for the cavity mode ajand magnon mode m, respectively, which are zero mean value acting on the cavity and magnon modes. The Gaussian nature of quantum noises can be characterized by the following correlation function [38]: hain j(t)ainy j(t0)i= [Nj(!j) + 1](t t0),hainy j(t)ain j(t0)i= Nj(!j)(t t0)(j= 1;2), andhmin(t)miny(t0)i= [Nm(!m)+ 1](t t0),hminy(t)min(t0)i=Nm(!m)(t t0) where Nl(!l) = [exp(h!l=kBT) 1] 1(l= 1;2;m) are the equilib- rium mean thermal photon numbers and magnon number, respectively, with kBthe Boltzmann constant and Tthe environmental temperature. Since the rst cavity is under strong driving by the mi- crowave eld "pand squeezed eld , which results in a large amplitudejha1ij1 at the steady state. Meanwhile, due to the beam-splitter-like coupling interaction between cavity modes and magnon mode, magnon mode and cav- ity modea2are also of large amplitudes in steady state. This allows us to linearize the system dynamics around the semiclassical averages and write any mode operator asO=hOi+O(O=a1;a2;m), neglecting small second- order
uctuation terms. Here, hOiis the mean value of the operatorO, andOis the zero-mean quantum
uctuation. We then obtain two sets of equations for semiclassical av- erages and for quantum
uctuations. The former set of equations are given by: (i1+1)ha1i ig1hmi i"p 2i hay 1i= 0 (6) (i2+2)ha2i ig2hmi= 0 (7) (im+m)hmi ig1ha1i ig2ha2i= 0 (8) By solving Eqs.(6)-(8), we obtain the steady-state solution for the average values ha1i=2 "p P "p 1 i1 4 2 P g2 1(2 i2) (m im)( 2 i2) g2 2(9) ha2i=g1g2ha1i (m im)(2 i2) g2 2(10) hmi= g1(2 i2)ha1i (m im)(2 i2) g2 2(11) whereP= 1+i1 g2 1(2+i2) (m+im)( 2+i2) g2 2. Thus, we can obtain the mean photon numbers and mean magnon number from Eqs.(9)-(11). On the other hand, quantum
uctuations is related to entanglement and squeezing. To study the quantum char- acteristics of the two cavity modes and magnon mode, the quadratures of quantum
uctuations about cavity modes and magnon mode are as X1= (a1+ay 1)=p 2,Y1= i(ay 1 a1)=p 2,X2= (a2+ay 2)=p 2,Y2=i(ay 2 3a2)=p 2,x= (m+my)=p 2, andy=i(my m)=p 2, and similarly for the input noise operators. The quantum Langevin equations describing quadrature
uctuations ( X1; Y1; X 2; Y2; x; y ) can be written as _f(t) =Af(t) + (12) wheref(t) = [X1(t),Y1(t),X2(t),Y2(t),x(t),y(t)]Tand (t) = [p21Xin 1(t),p21Yin 1(t),p22Xin 2(t),p22Yin 2(t),p2mxin(t),p2myin(t)]Tare the vectors for quantum
uctuations operator and noises operator, respectively. The drift matrix A is given by A=0 BBBBBB@ 11 2 0 0 0 g1 1 2 1 0 0 g10 0 0 220g2 0 0 2 2 g20 0g1 0g2 mm g1 0 g20 m m1 CCCCCCA Due to the linearized dynamics and the Gaussian na- ture of the quantum noises in our system, the steady state of quantum
uctuations is a continuous variable three mode Gaussian state, which is completely characterized by a 66 covariance matrix Vdened asVij=hfi(t)fj(t0) + fj(t0)fi(t)i=2 (i;j= 1;2;:::;6). In generally, the steady- state covariance matrix Vcan be obtained straightfor- wardly by solving the Lyapunov equation [39, 40] AV+VAT= D (13) whereD=diag[1(2N1+ 1),1(2N1+ 1),2(2N2+ 1), 2(2N2+ 1),m(2Nm+ 1),m(2Nm+ 1)] is the diu- sion matrix, which is dened as Dij(t t0) =hi(t)j(t0)+ j(t0)i(t)i=2. With the covariance matrix in hand, we can get the quantities related to entanglement and squeezing. To quantify entanglement between the two cavity modes and magnon mode, we adopt quantitative measures of the logarithmic negativity [41, 42] ENfor the bipartite entan- glement, which is dened as ENmax[0; ln2~ ] (14) where ~ =min[eigji 2~V4j] is the minimum symplectic eigen- value of the ~V4=P1j2V4P1j2.V4is the 44 covariance matrix, which can be obtained by directly removing in V the rows and columns of uninteresting mode. Meanwhile, to realize partial transposition at the level of covariance matrix, we set P1j2=diag(1; 1;1;1). 2is symplectic matrix with 2=2 j=1iyandyis they-Pauli matrix. A nonzero logarithmic negativity EN>0 denotes the pres- ence of bipartite entanglement in our QED system. Meanwhile, a quantication of continuous variable tri- partite entanglement is given by the minimum residual contangle [43, 44], dened as Rmin min[Rajm1m2 ;Rm1jam2 ;Rm2jam1 ] (15) whereRijjk Cijjk Cijj Cijk0 (i;j;k =a;m 1;m2) is the residual contangle, with Cujvthe contangle of sub- systems of uandv(vcontains one or two modes), whichis a proper entanglement monotone dened as the squared logarithmic negativity. When vcontains two modes, loga- rithmic negativity Eijjkcan be calculated by the denition of Eq.(14). We only need to use 3=3 j=1iyinstead of 2=2 j=1iyand ~V6=PijjkVPijjkinstead of ~V4= P1j2V4P1j2, whereP1j23=diag(1; 1;1;1;1;1),P2j13= diag(1;1;1; 1;1;1) andP3j12=diag(1;1;1;1;1; 1) are partial transposition matrices. Rmin 0 denotes the pres- ence of genuine tripartite entanglement in three modes Gaussian system. Meanwhile, squeezing can be calculated by the covari- ance matrix of quantum
uctuations. The variances of squeezed magnon quadratures are amplitude quadrature hx(t)2i, phase quadrature hy(t)2i, and amplitude quadra- turehY2(t)2iis quadrature of cavity mode a2,x= (my+ m)=p 2,y=i(my m)=p 2, andY2=i(ay 2 a2)=p 2. In our denition, hQ(t)2ivac= 1=2 (Q is a mode quadra- ture) denotes vacuum
uctuations. The degree of squeez- ing can be expressed in the dB unit, which can be evalu- ated by 10log10[hQ(t)2i=hQ(t)2ivac], wherehQ(t)2ivac= 1=2. 3. Results and discussion To show whether the squeezed driving can induce en- tanglement, we consider a simpler magnon-cavity QED system at rst, where no coupling interaction exists be- tween the magnon mode and cavity mode a2, i.e.,g2= 0. Fig.2(a) shows the bipartite entanglement between cav- ity modea1and magnon mode versus detunings 1and min steady state. We employed experimentally fea- sible parameter [5] at low temperature T= 10mK, as !1=2= 10GHz, m=2= 1MHz,1=2=2=2= 5MHz,g1=2= 20MHz. Moreover, Rabi frequency of microwave eld we employed is "p= 10m. Squeezed eld used in our system is to generate nonlinear term by the JPA process with gain coecient = 2 :5m. This is the nonlinearity that causes entanglement in our QED system. Fig.2(a) shows that the photon-magnon entan- glement described by logarithmic negativity can achieve to 0.3. Meanwhile, due to the state-swap interaction be- tween the cavity mode a1and magnon mode, the squeez- ing can be transferred from squeezed cavity mode a1to the magnon mode, as shown in Fig.2(b). Note that the above results are valid only when the assumption of low-lying excitations, i.e. magnon excita- tion numberhmymi 2Ns, wheres= 5=2 is the spin number of ground-state Fe3+ion in YIG sphere. The to- tal number of spins N=Vwith= 4:221027m 3 the spin density of YIG and Vthe volume of sphere. For a 250-m-diameter YIG sphere, the number of spins N'3:51016. We then calculate the mean photon numbers of cavity mode a1N1=hay 1a1i, cavity mode a2 N2=hay 2a2i, and mean magnon number Nm=hmymivia Eqs.(9)-(11), which are closely related to the input inten- sity of microwave eld and squeezed eld. Fig.2(c) and 4(d) show the mean photon number N1and mean magnon numberNmversus detunings 1and min steady state wheng2= 0. They are drawn with logarithmic log10. We show that both the maximum number of photons and magnons are above 10, but less than 103in Fig.2(c) and (d). Meanwhile, we also get N2= 0. so the assumption of low-lying excitations is well satised. Fig. 2. (a)Density plot of photon-magnon bipartite entanglement Ea1m, (b)variance of the magnon amplitude quadrature hx(t)2i, (c)logarithm of mean photon number of cavity mode a1N1, and (d)logarithm of mean magnon number Nmversus detunings 1and m. We choose = 2 :5m,"p= 10m. The blank area denotes hQ(t)2ivac>1=2, i.e., above vacuum
uctuations. We take g2= 0 for all the plots. See text for the detail of other parameters. We then take g2into consideration. To be more gen- eral, we assume that coupling rate g2is the same as that between the cavity mode a1and magnon mode, i.e., g2= g1. In Fig.3(a)-3(c), mean photon numbers and mean magnon number, N1,N2, andNm, are plotted as func- tions of detunings 2and m, respectively. They are also drawn with logarithmic log10. It is noted that P= 0 is the extreme value of Eqs.(9)-(11). Ignoring dissipa- tive terms and analyzing the extreme value, we can ob- tain a simple form m= ( 1g2 2+ 2g2 1)=(12). The black dashed curves in Fig.3(a)-(c) denote m= ( 1g2 2+ 2g2 1)=(12), and from which we can see that the max- imum numbers of photons and magnons are located at about this region. -1.01.0 -2.0 (a) (b) (c) 0 Fig. 3. (a)Mean photon number of cavity mode a1(squeezed cav- ity mode)N1, (b)mean photon number of cavity mode a2N2, and (c)mean magnon number Nmversus detunings 2and m. Black dash curves indicate m= ( 1g2 2+ 2g2 1)=(12). All Figures are drawn with logarithmic log10. We take Rabi frequency of microwave eld"p= 10mand the nonlinear gain coecient of squeezed eld = 2:5m. We assume the coupling rate between the two cavity modes and magnon mode are the same, i.e., g2=g1. The detuning of cavity mode a11= 20m. The other parameters are as in Fig.2.After coupling cavity mode a2to magnon mode ( g2> 0), the magnon-cavity a1entanglement Ea1mdecreased while cavity mode a2and magnon mode get entangled, as shown in Fig.4(a) and (c) with assuming g2=g1. It denotes that the quantum correlations can be trans- ferred from magnon mode and cavity mode a1to magnon mode and cavity mode a2. All results are in the steady state guaranteed by the negative eigenvalues (real parts) of the drift matrix A. We also choose the Rabi frequency of microwave eld "p= 10mand the gain coecient = 2:5m. The Black dashed curves in Fig.4(a) and (c) denote m= ( 1g2 2+ 2g2 1)=(12). It clearly shows that the optimal photon-magnon entanglement is achieved near the maximum of mean particle numbers. We then calculated the squeezing by the covariance ma- trix of quantum
uctuations applying mean eld approxi- mation, and found that the cavity modes and the magnon mode can be squeezed. Compared with photons, it is more meaningful to study squeezed magnons, a mesoscopic ob- ject. Two quadratures of magnon mode are amplitude quadraturehx(t)2iand phase quadrature hy(t)2i, these two quadratures also obey the uncertainty relationship, i.e., when the phase (amplitude) quadrature is squeezed, the amplitude (phase) quadrature will not be squeezed. That is, the squeezed of one quadrature is at the expense of increasing the other one. Variance of magnon ampli- tude quadrature hx(t)2iand phase quadrature hy(t)2i versus detunings 2and mare shown in Fig.4(b) and (d), respectively. The blank area denotes above vacuum
uctuations, i.e., hQ(t)2ivac>1=2, (Q=x;y). Fig. 4. (a)Density plot of bipartite entanglement Ea1m, (b)variance of the magnon amplitude quadrature hx(t)2i, (c)density plot of bi- partite entanglement Ea2m, and (d)variance of the magnon phase quadraturehy(t)2iversus detunings 2and m. We choose = 2:5m,"p= 10m. The detuning of cavity mode a1 1= 20m. Black dash curves in Fig.4(a) and (c) indicate m= ( 1g2 2+ 2g2 1)=(12). The blank area in Fig.4(b) and (d) denoteshQ(t)2ivac>1=2, i.e., above vacuum
uctuation. We takeg2=g1for all the plots. See text for the other parameters. Further, Fig.5(a) shows that the two cavity modes get entangled, which denotes that the photon-photon entan- glementEa1a2is transferred from magnon-cavity entangle- mentEa1mandEa2mdue to the state-swap interaction be- tween the two cavity modes and magnon mode. The cou- 5pling rateg2also induces the squeezing transferred from cavity mode a1to cavity mode a2via magnon mode, as shown in Fig.5(b). The blank area denotes above vac- uum
uctuations, i.e., hQ(t)2ivac>1=2. Comparing to Fig.2(b), the maximum of variance of the magnon am- plitude quadratures hx(t)2iandhy(t)2idecreases, and cavity mode a2get squeezed. It denotes that the two cav- ity modes and magnon mode are all prepared in squeezed states due to the state-swap interaction between the two cavity modes and magnon mode, meaning that the mag- netic dipole interaction is an essential element to generate squeezed states. Logarithmic negativity Ea1a2as a func- tion of bath temperature is shown in Fig.5(c). It denotes that photon-photon entanglement Ea1a2is robust again bath temperature and survives up to about 200 mK. Tri- partite entanglement in terms of the minimum residual contangleRmin detunings 2and mis shown in Fig5(d). It shows that the tripartite entanglement does exist in our QED system. The black dashed curves in Fig.5(d) denote m= ( 1g2 2+ 2g2 1)=(12), and from which we can see that the maximum of tripartite entanglement located at about this region. Fig. 5. (a)Density plot of photon-photon bipartite entanglement Ea1a2, and (b)variance of cavity mode a2amplitude quadrature hY2(t)2iversus detunings 2and m. (c)Logarithmic negativ- ityEa1a2vs bath temperature T. (d) Tripartite entanglement in terms of the minimum residual contangle Rmin detunings versus 2 and m. The blank area in Fig.5(b) denotes hQ(t)2ivac>1=2, and black dash curves indicate m= ( 1g2 2+ 2g2 1)=(12) in Fig.5(d). We take 2= 35m, m= 45mfor (c), = 2 :5m, "p= 10m, 1= 20mandg2=g1for all the plots. See text for the other parameters. Squeezing does not increase linearly with increasing the gain coecient. We choose 2= 0 , and nd that squeez- ing rst increases and then decreases with the increase of the gain coecient, as shown in Fig.6(a) and (b), respec- tively.The blank area denotes hQ(t)2ivac>1=2. Squeez- ing reaches the maximum near = 8 mfor the amplitude quadrature and near = 2 mfor phase quadrature. To obtain the optimal entanglement between the two cavity modes, we show photon-photon entanglement Ea1a2 versus gain coecient and the rate of magnon-cavity coupling strength g2=g1in Fig.7(a). All results are cal- culated in the steady state, and the blank area denotes Fig. 6. (a)Variance of the magnon amplitude quadrature hx(t)2i, (b)Variance of the magnon phase quadrature hy(t)2iversus gain co- ecient and detunings m. The blank area denotes hQ(t)2ivac> 1=2. We take 1= 20m, 2= 0,"p= 10mfor all the plots. See text for the other parameters. Non equilibrium state. As shown in Fig.7(a), the two cavity modes show zero entanglement in the absence of gain coecient, i.e., = 0, meaning that it is squeezed driving that induced entanglement in our QED system. It demonstrates that the nonlinearity produced by para- metric down-conversion is the reason to generate entan- glement. Bipartite entanglement Ea1mincreases with the increase of gain coecient , and then the entanglement transferred from Ea1mtoEa2mandEa1a2. But, to keep the system in steady state, the gain coecient can not be too large. Fig.7(a) denotes that increasing gain coecient is a good way to improve entanglement in our QED system. Due to the
exible tunability of gain coecient, which makes large entanglement possible. Further, we show that the optimal entanglement can be generated when the rate of photon-magnon coupling strength are almost the same, i.e.,g2=g11. Meanwhile, we show variance of magnon amplitude quadraturehx(t)2iversus gain coecient and the rate of magnon-cavity coupling strength g2=g1in Fig.7(b). The blank area represents above vacuum
uctuation, i.e., hQ(t)2ivac>1=2. It shows that the magnon mode can not be squeezed in the absence of squeezed eld, i.e., = 0, and the strength of squeezed magnon mode transferred from squeezed cavity mode a1can increase a lot as gain coecient increasing. It provides a good scheme to im- prove macroscopic quantum state. Further, we also show that the optimal squeezing is also located at about the regiong2=g1= 1. Fig. 7. (a)Density plot of photon-photon bipartite entanglement Ea1a2, (b)variance of the magnon amplitude quadrature hx(t)2i versus nonlinear gain coecient and the rate of magnon-cavity coupling strength g2=g1. The blank area in Fig.7(a) presents non equilibrium states and hQ(t)2ivac>1=2 in Fig.7(b), i.e., above vacuum
uctuations. We take 2= 35m, m= 45mfor (a), 2= 45m, m= 15mfor (b), and 1= 20m,"p= 10m for all the plots. See text for the other parameters. 64. Conclusion In summary, we have presented a scheme to gener- ate bipartite entanglement between two cavity modes and squeeze magnon mode in a magnon-cavity QED system by using a squeezed driving. With experimentally reach- able parameters, we show that without the nonlinearity induced by parametric down-conversion process, our QED system denotes zero entanglement and above vacuum
uc- tuations. We also show the photon-magnon entanglement can transfer to photon-photon entanglement by state-swap interaction between cavity and magnon modes in the steady state. A genuinely tripartite entangled state is formed. Meanwhile, magnon squeezed state also can be realized due to the squeezing from squeezed driving cavity mode. Moreover, our QED system shows that increasing the non- linear gain coecient is a good way to enhance entangle- ment and squeezing. Further, the optimal entanglement and squeezing is located at about the region where the cou- pling rates between two cavity modes and magnon mode are almost the same. Our results denote that magnon- cavity QED system is a powerful platform for studying macroscopic quantum phenomena, and squeezed drive pro- vides an new method for generating macroscopic quantum state. Acknowledgements This work has been supported by the National Natural Science Foundation of China (Grant No. 12174288, Grant No. 11874287, Grant No. 11774262, Grant No. 61975154), and the Shanghai Science and Technology Committee (Grant No. 18JC1410900). References [1] Dany Lachance-Quirion, Yutaka Tabuchi, Arnaud Gloppe, Koji Usami, and Yasunobu Nakamura. Hybrid quantum systems based on magnonics. Appl. Phys. Exp. , 12(7):070101, Jun 2019. [2] Graeme Flower, Jeremy Bourhill, Maxim Goryachev, and Michael E Tobar. Broadening frequency range of a ferromag- netic axion haloscope with strongly coupled cavity{magnon po- laritons. Phys. Dark Universe , 25:100306, 2019. [3] Na Zhu, Xufeng Zhang, Xu Han, Chang-Ling Zou, Changchun Zhong, Chiao-Hsuan Wang, Liang Jiang, and Hong X Tang. Waveguide cavity optomagnonics for microwave-to-optics con- version. Optica , 7(10):1291, 2020. [4] Hans Huebl, Christoph W. Zollitsch, Johannes Lotze, Fredrik Hocke, Moritz Greifenstein, Achim Marx, Rudolf Gross, and Sebastian T. B. Goennenwein. High cooperativity in coupled microwave resonator ferrimagnetic insulator hybrids. Phys. Rev. Lett., 111(12):127003, 2013. [5] Yutaka Tabuchi, Seiichiro Ishino, Toyofumi Ishikawa, Rekishu Yamazaki, Koji Usami, and Yasunobu Nakamura. Hybridizing ferromagnetic magnons and microwave photons in the quantum limit. Phys. Rev. Lett. , 113(8):083603, 2014. [6] Maxim Goryachev, Warrick G. Farr, Daniel L. Creedon, Yao- hui Fan, Mikhail Kostylev, and Michael E. Tobar. High- cooperativity cavity qed with magnons at microwave frequen- cies. Phys. Rev. Appl. , 2(5):054002, 2014. [7] Xufeng Zhang, Changling Zou, Liang Jiang, and Hong X. Tang. Strongly coupled magnons and cavity microwave photons. Phys. Rev. Lett. , 113(15):156401, 2014.[8] Charles Kittel. On the theory of ferromagnetic resonance ab- sorption. Phys. Rev. , 73(2):155, 1948. [9] H. Primako and T. Holstein. Field dependence of the in- trinsic domain magnetization of a ferromagnet. Phys. Rev. , 58(12):1098, 1940. [10] Lihui Bai, M. Harder, Y. P. Chen, X. Fan, J. Q. Xiao, and C. M. Hu. Spin pumping in electrodynamically coupled magnon- photon systems. Phys. Rev. Lett. , 114(22):227201, 2015. [11] Dengke Zhang, Xinming Wang, Tiefu Li, Xiaoqing Luo, Wei- dong Wu, Franco Nori, and J. Q. You. Cavity quantum electro- dynamics with ferromagnetic magnons in a small yttrium-iron- garnet sphere. NPJ Quan. Inf. , 1(1):15014, 2015. [12] J. Bourhill, N. Kostylev, M. Goryachev, D. L. Creedon, and M. E. Tobar. Ultrahigh cooperativity interactions between magnons and resonant photons in a yig sphere. Phys. Rev. B, 93(14):144420, 2016. [13] Nikita Kostylev, Maxim Goryachev, and Michael E. Tobar. Su- perstrong coupling of a microwave cavity to yttrium iron garnet magnons. Appl. Phys. Lett. , 108(6):3534, 2016. [14] Lihui Bai, Michael Harder, Paul Hyde, Zhaohui Zhang, Can- ming Hu, Y. P. Chen, and John Q. Xiao. Cavity mediated manipulation of distant spin currents using a cavity-magnon- polariton. Phys. Rev. Lett. , 118(21):217201, 2017. [15] Yipu Wang, Guoqiang Zhang, Dengke Zhang, Tiefu Li, C. M. Hu, and J. Q. You. Bistability of cavity magnon polaritons. Phys. Rev. Lett. , 120(5):057202, 2018. [16] Xufeng Zhang, Changling Zou, Na Zhu, Florian Marquardt, Liang Jiang, and Hong X. Tang. Magnon dark modes and gra- dient memory. Nat. Comm. , 6(1):8914, 2015. [17] Cui Kong, Bao Wang, Zengxing Liu, Hao Xiong, and Ying Wu. Magnetically controllable slow light based on magnetostrictive forces. Opt. Exp. , 27(4):5544, 2019. [18] M. Harder, Y. Yang, B. M. Yao, C. H. Yu, J. W. Rao, Y. S. Gui, R. L. Stamps, and C. M. Hu. Level attraction due to dissipa- tive magnon-photon coupling. Phys. Rev. Lett. , 121(13):137203, 2018. [19] Bao Wang, Zengxing Liu, Cul Kong, Hao Xiong, and Ying Wu. Magnon-induced transparency and amplication in pt- symmetric cavity-magnon system. Opt. Exp. , 26(16):20248, 2018. [20] Xufeng Zhang, Changling Zou, Liang Jiang, and Hong X. Tang. Cavity magnomechanics. Sci. Adv. , 2(3):e1501286, 2016. [21] Jie Li, Shiyao Zhu, and G. S. Agarwal. Squeezed states of magnons and phonons in cavity magnomechanics. Phys. Rev. A, 99(2):021801, 2019. [22] MS Ebrahimi, Ali Motazedifard, and M Bagheri Harouni. Single-quadrature quantum magnetometry in cavity electro- magnonics. Phys. Rev. A , 103(6):062605, 2021. [23] Angelo Bassi, Kinjalk Lochan, Seema Satin, Tejinder P. Singh, and Hendrik Ulbricht. Models of wave-function collapse, un- derlying theories, and experimental tests. Rev. Mod. Phys. , 85(2):471, 2013. [24] Yutaka Tabuchi, Seiichiro Ishino, Atsushi Noguchi, Toyofumi Ishikawa, Rekishu Yamazaki, Koji Usami, and Yasunobu Naka- mura. Coherent coupling between a ferromagnetic magnon and a superconducting qubit. Science , 349(6246):405, 2015. [25] Wei Zhang, Dongyang Wang, Chenghua Bai, Tie Wang, Shou Zhang, and Hongfu Wang. Generation and transfer of squeezed states in a cavity magnomechanical system by two-tone mi- crowave elds. Opt. Exp. , 29(8):11773, 2021. [26] M. J. Weaver, D. Newsom, F. Luna, W. Loer, and D. Bouwmeester. Phonon interferometry for measuring quan- tum decoherence. Phys. Rev. A , 97(6):063832, 2018. [27] Yanbei Chen. Macroscopic quantum mechanics: theory and experimental concepts of optomechanics. J. Phys. B: At. Mol. Opt. Phys. , 46(10):104001, 2013. [28] A. J. Leggett. Testing the limits of quantum mechanics: mo- tivation, state of play, prospects. J. Phys.: Cond. Matt. , 14(15):R415, 2002. [29] Florian Froewis, Pavel Sekatski, Wolfgang D Ur, Nicolas Gisin, and Nicolas Sangouard. Macroscopic quantum states: 7Measures, fragility, and implementations. Rev. Mod. Phys. , 90:025004, 2018. [30] Mei Yu, Heng Shen, and Jie Li. Magnetostrictively induced stationary entanglement between two microwave elds. Phys. Rev. Lett. , 124(21):213604, 2020. [31] Zhedong Zhang, Marlan O. Scully, and Girish S. Agarwal. Quantum entanglement between two magnon modes via kerr nonlinearity driven far from equilibrium. Phys. Rev. Res. , 1(2):023021, 2019. [32] Jayakrishnan M. P. Nair and G. S. Agarwal. Deterministic quantum entanglement between macroscopic ferrite samples. Appl. Phys. Lett. , 117(8):084001, 2020. [33] Jie Li, Shiyao Zhu, and G. S. Agarwal. Magnon-photon-phonon entanglement in cavity magnomechanics. Phys. Rev. Lett. , 121(20):203601, 2018. [34] Jie Li and Shiyao Zhu. Entangling two magnon modes via mag- netostrictive interaction. N. J. Phys. , 21(8):85001, 2019. [35] Mei Yu, Shi Yao Zhu, and Jie Li. Macroscopic entanglement of two magnon modes via quantum correlated microwave elds. J. Phys. B: At. Mol. Opt. Phys. , 53(6):065402, 2020. [36] Jie Li and Simon Groeblacher. Entangling the vibrational modes of two massive ferromagnetic spheres using cavity mag- nomechanics. Quan. Sci. Tech. , 6(2):024005, 2021. [37] Hoa T. Nguyen and A. G. Cottam. Dipole-exchange theory of spin waves in nanowires: Application to arrays of interacting ferromagnetic stripes. In AIP Conference Proceedings , volume 1147, page 117. Springer, 2009. [38] C. W. Gardiner and P. Zoller. Quantum Noise . Springer Berlin,Germany, 2004. [39] D. Vitali, S. Gigan, A. Ferreira, H. R. B Ohm, P. Tombesi, A. Guerreiro, V. Vedral, A. Zeilinger, and M. Aspelmeyer. Op- tomechanical entanglement between a movable mirror and a cavity eld. Phys. Rev. Lett. , 98:030405, 2007. [40] P. C. Parks and V. Hahn. Stability theory . Springer New York, 1993. [41] G. Vidal and R. F. Werner. Computable measure of entangle- ment. Phys. Rev. A , 65:032314, 2002. [42] M. B. Plenio. Logarithmic negativity: a full entanglement monotone that is not convex. Phys. Rev. Lett. , 95(9):090503, 2005. [43] G. Adesso and F. Illuminati. Continuous variable tangle, monogamy inequality, and entanglement sharing in gaussian states of continuous variable systems. N. J. Phys. , 8:15, 2006. [44] Gerardo Adesso and Fabrizio Illuminati. Entanglement in continuous-variable systems: recent advances and current per- spectives. J. Phys. A: Math. Theo. , 40(28):7821, 2007. 8 | 2022-01-23 | We propose a scheme to generate entanglement between two cavity modes and
squeeze magnon mode in a magnon-cavity QED system, where the two microwave
cavity modes are coupled with a massive yttrium iron garnet (YIG) sphere
through magnetic dipole interaction. The nonlinearity used in our system
originates from a squeezed driving via parametric down-conversion process,
which is the reason to cause entanglement and squeezing. By using the mean
field approximation and employing experimentally feasible parameters, we
demonstrate that the system shows zero entanglement and squeezing without
squeezed driving. Meanwhile, our QED system denotes that the entanglement
between squeezed cavity mode and magnon mode can be transferred to the other
cavity mode and magnon mode via magnon-cavity coupling interaction, and then
the two cavity modes get entangled. A genuinely tripartite entangled state is
formed. We also show that magnon mode can be prepared in a squeezed state via
magnon-cavity beam-splitter interaction, which is as a result of the squeezed
field. Moreover, we show that it is a good way to enhance entanglement and
squeezing by increasing the nonlinear gain coefficient of squeezed driving. Our
results denote that magnon-cavity QED system is a powerful platform for
studying macroscopic quantum phenomena, which illustrates a new method to
photon-photon entanglement and magnon squeezing. | Squeezed driving induced entanglement and squeezing among cavity modes and magnon mode in a magnon-cavity QED system | 2201.09154v1 |
Detection of entanglement by harnessing extracted work in an opto-magno-mechanics M’bark Amghar1and Mohamed Amazioug1,∗ 1Department of Physics, Ibnou Zohr University, Agadir 80000, Morocco (Dated: May 30, 2024) The connections between thermodynamics and quantum information processing are of paramount impor- tance. Here, we address a bipartite entanglement via extracted work in a cavity magnomechanical system contained inside an yttrium iron garnet (YIG) sphere. The photons and magnons interact through an interaction between magnetic dipoles. A magnetostrictive interaction, analogous to radiation pressure, couple’s phonons and magnons. The extracted work was obtained through a device similar to the Szil ´ard engine. This engine operates by manipulating the photon-magnon as a bipartite quantum state. We employ logarithmic negativity to measure the amount of entanglement between photon and magnon modes in steady and dynamical states. We explore the extracted work, separable work, and maximum work for squeezed thermal states. We investigate the amount of work extracted from a bipartite quantum state, which can potentially determine the degree of entanglement present in that state. Numerical studies show that entanglement, as detected by the extracted work and quantified by logarithmic negativity, is in good agreement. We show the reduction of extracted work by a second measurement compared to a single measurement. Also, the e fficiency of the Szilard engine in steady and dynamical states is investigated. We hope this work is of paramount importance in quantum information processing. I. INTRODUCTION Entanglement, a cornerstone of quantum mechanics pioneered in works like [1–3], holds immense potential across various fields. It has exciting applications in areas like precision measurement, quantum key distribution [4], teleporting quantum information [5], and building powerful quantum computers as explored by [6]. Entanglement creates a spooky connection between particles. Measuring one instantly a ffects the other, defying classical ideas of locality. Cavity optomechanics is a field of physics that studies the interaction between light and mechanical objects via radiation pressure at very small scales [7]. Cavity optomechanics has been used to develop new methods for generating and manipulating squeezed states of light, which are a type of quantum state. Recently this cavity has paramount importance applications in quantum information processing such as quantum entangled states [8–19], cooling the mechanical mode to their quantum ground states [20–24], photon blockade [25], enhancing precision measurements [27, 28], superconducting elements [29], and also between two massive mechanical oscillators [30, 31] have been observed. Cavity quantum electrodynamics (CQED) paved the way for the emergence of a new field called cavity optomechanics. Cavity quantum electrodynamics (CQED) o ffers control over how light (photons) interacts with atoms at the quantum level. Actually, single quanta can significantly impact the atom-cavity dynamics in the strong coupling regime, which is made possible by strong confinement. Recently, we have successfully achieved strong coupling in numerous experiments, leading to the demonstration of fascinating quantum phenomena such as quantum phase gates [32], the Fock state generation [33], and quantum nondemolition detection of a single cavity photon [34]. Building on the success of cavity QED, exploring how magnon systems interact within cavity optomechanics o ffers a promising avenue for unlocking their unique quantum properties. The first experimental demonstration of interaction between magnons, photons, and phonons has been achieved [35]. This system combines magnon-photon coupling, similar to what’s found in magnon QED, with an additional coupling between magnons and phonons. While the cavity output reflects the impact of magnon-phonon coupling, a more comprehensive understanding would require a full quantum treatment that incorporates these fluctuations. From the standpoint of cavity quantum electrodynamics (QED), ferrimagnetic systems. Particularly, the yttrium iron garnet (YIG) sphere has garnered a lot of attention. Studies have shown that the YIG sphere’s Kittel mode [36] can achieve strong coupling with microwave photons trapped within a high-quality cavity. This strong coupling leads to the formation of cavity polaritons [37] and a phenomenon known as vacuum Rabi splitting. The success of cavity QED has opened doors to applying many of its concepts to the emerging field of magnon cavity QED [38]. This new field has already seen exciting advancements, including the observation of bistability [39] and the groundbreaking coupling of a single superconducting qubit to the Kittel mode [40]. Recently, magnons (as spin waves) have been studied extensively in the field of quantum information processing [41–44]. Studies suggest that physical system information can be utilized to extract work with suitable operations [45]. This, described as information thermodynamics, explores the relationship between information theory and thermodynamics. By studying how information can be manipulated to perform physical tasks, researchers hope to uncover new ways to improve e fficiency in ∗amazioug@gmail.comarXiv:2405.19205v1 [quant-ph] 29 May 20242 various processes. Szilard’s engine [46] is a classic example of this concept. It demonstrates how information processing is capable of extracting work from a physical system. Interestingly, research has shown that the specific way information is encoded, particularly in entangled or correlated states, plays a crucial role in how much work can be extracted [47–49]. In this letter, we investigate the potential of exploiting both extracted work and e fficiency in optomagnomechanical systems. This study reveals the presence of entanglement between magnons and cavity photons in an optomechanical system. We achieve this detection by examining the extractable work. Our work breaks new ground by employing extractable work as a tool to detect entanglement between magnons and cavity photons in optomechanical systems. Our study, utilizing realistic experimental parameters, reveals excellent agreement between the entanglement region detected via extractable work and the results of Jie Li et al [50]. This highlights the validity of our method for entanglement detection in magnomechanical systems. This applies not only to steady-state conditions but also to dynamical states under thermal influence. Our study further explores the influence of various parameters, including detuning and magnon-phonon coupling, on the entanglement properties of the system. In addition, we analyze the information-work e fficiency under thermal noise, considering both steady-state and dynamical regimes. The article is outlined as follows: In Section II, we introduce the model for the optomagnomechanical system, its Hamilto- nian, and the quantum nonlinear Langevin equations for the interacting photon-magnon-phonon system. Section III tackles the linearization of the QLEs, and we then assess the covariance matrix for steady and dynamical states. Section IV delves into the connection between quantum thermodynamics and quantum entanglement in cavity magnomechanical systems. We employ logarithmic negativity to quantify the entanglement between the photon and magnon modes and investigate how it relates to the amount of work that can be extracted from the system. Besides, we have also investigated the e fficiency of a Szilard engine. The results obtained are discussed. Concluding remarks close this paper. II. MODEL In this section, we consider on a system that combines a microwave cavity with magnetic excitations (magnons) and mechan- ical vibrations (phonons). This cavity magnomechanical system is illustrated in Fig. 1. Magnons are a collective motion of numerous spins in a ferrimagnet, such as an YIG sphere (250- µm-diameter sphere used in Ref. [35]). A sphere made of YIG (Yttrium Iron Garnet) is positioned within a microwave cavity at a location with the strongest magnetic field. Additionally, a uniform bias magnetic field is applied throughout the entire system. These combined fields allow the microwave photons in the cavity to interact with the YIG sphere’s magnons through the magnetic dipole interaction. To improve the coupling between magnons and phonons, the experiment utilizes a microwave source (not shown) to directly driven the magnon mode magne- tostrictive interaction. Due to the YIG sphere’s small size compared to the microwave wavelength, we can ignore the interaction between microwave photons and phonons. We consider the three magnetic fields: a bias field pointing in the z-axis, a drive field in the y-axis, and the magnetic field of the cavity mode oriented along the x-axis, as depicted in Fig. 1. These three fields are mu- tually perpendicular at the position of the YIG sphere. The YIG sphere experiences a deformation of its geometry structure due to the creation of vibrational modes, or phonons, which influence the magnon excitations within the sphere, and vice versa [51]. The Hamiltonian writes as [50] ˆH=ℏΩcˆc†ˆc+ℏΩnˆn†ˆn+Ωd 2( ˆx2+ˆy2) +ℏgndˆn†ˆnˆx+ℏGnc(ˆc+ˆc†)(ˆn+ˆn†) +iℏΩ(ˆn†e−iΩ0t−ˆneiΩ0t),(1) where the creation and annihilation operators for the cavity (ˆ c,ˆc†) and magnon (ˆ n,ˆn†) modes satisfy the canonical commutation relation [ ˆO,ˆO†]=1 (where ˆOcan be ˆ cor ˆn). Additionally, dimensionless position ˆ xand momentum ˆ yoperators for the mechan- ical mode are included, with the commutation relation [ ˆ x,ˆy]=i. The Hamiltonian also incorporates the resonance frequencies (Ωc,Ωn, and Ωd) of the cavity, magnon, and mechanical modes, respectively. The magnon frequency Ωnis dictated by the exter- nal bias magnetic field Hand the gyromagnetic ratio γfollowing the relation: Ωn=γH. Interestingly, the magnon-microwave coupling rateGncsurpasses the dissipation rates of both the cavity λcand magnon modes λn, satisfying the condition for strong coupling:Gnc>λ c,λn[37]. The inherent coupling rate between a single magnon and the mechanical vibrations, denoted by gnd, is typically low. This limitation can be overcome by strategically applying a strong microwave field directly to the YIG sphere. This approach, employed in earlier works [39, 52], e ffectively enhances the magnomechanical interaction. The Rabi frequency Ω, derived under the assumption of low-lying excitations ( ⟨ˆn†ˆn⟩≪2Ns, where s=5/2 is the spin of the Fe3+ground state ion), characterizes the coupling strength between the driving magnetic field (amplitude B0and frequency ω0) and the magnon mode. It is expressed as Ω =√ 5 4γg√ NB0, whereγg/2π=28 GHz /T is the gyromagnetic ratio of the material and N=ϱVrepresents the total number of spins in the YIG sphere. Here, ϱ=4.22×1027m−3is the spin density and Vis the sphere’s volume. Using a rotating frame at the driving frequency Ω0and the rotating-wave approximation Gnc(ˆc+ˆc†)(ˆn+ˆn†)→G nc(ˆcˆn†+ˆc†ˆn) valid3 when Ωc,Ωn≫G nc,κc,κn[35], the system’s dynamics are described by quantum Langevin equations (QLEs). δ˙ˆc=−(iδc+λc)ˆc−iGncˆn+p 2λcˆcin, δ˙ˆn=−(iδn+λn)ˆn−iGncˆc−igndˆnˆx+ Ω +p 2λnˆnin, δ˙ˆx=ωˆy, δ ˙ˆy=−ωˆx−γdˆy−gndˆn†ˆn+η,(2) withδc= Ω c−Ω0,δn= Ω n−Ω0andγdis the mechanical damping rate. The input noise operators for the cavity and magnon modes are respectively, ˆ cinand ˆnin, with zero mean, i.e., ⟨ˆcin⟩=⟨ˆnin⟩=0, and described by the following correlation functions [53]: ⟨ˆcin(t) ˆcin†(t′)⟩=[Nc(Ωc)+1]δ(t−t′),⟨ˆcin†(t) ˆcin(t′)⟩=Nc(Ωc)δ(t−t′) ⟨ˆnin(t) ˆnin†(t′)⟩=[Nn(Ωn)+1]δ(t−t′),⟨ˆnin†(t) ˆnin(t′)⟩=Nn(Ωn)δ(t−t′),(3) we assume the mechanical mode follows a Markovian process. This means a large mechanical quality factor Q≫1, i.e., Ωd≫γd[54]. Furthermore, the noise operators for this mode possess non-zero correlation properties (with ⟨η(t)⟩=0), writes as ⟨η(t)η(t′)+η(t′)η(t)⟩/2≃γd[2Nd(Ωd)+1]δ(t−t′), (4) where Nc,NnandNdcorrespond to the equilibrium mean thermal occupation numbers for the cavity photons, magnons, and phonons, respectively. Thus Nj(Ωj)=exp ℏΩj kBT−1−1(j=ˆc,ˆn,ˆd), where kBis the Boltzmann constant. III. COV ARIANCE MATRIX We consider the case where the magnon mode is highly driven. We linearize the non-linear quantum Langevin equation by assuming small fluctuations around a steady state amplitude, i.e., ˆO=ˆOss+δˆO(ˆO=ˆc,ˆn,ˆx,ˆy), where ˆ nsswrites as ˆnss=Ω(iδc+λc) G2nc+(i˜δn+λn)(iδc+λc), (5) where ¯δn=δn+gndˆxssis the e ffective magnon-drive detuning taking into account a frequency shift resulting from the interaction between magnons and phonons. This interaction is known as magnomechanical interaction. Under the condition |¯δn|,|δc|≫ λc,λn; ˆnssis given by ˆnss=iΩδc G2nc−˜δnδc. (6) and ˆxss=−gnd Ωd|ˆnss|2. The system is described by linearized quantum Langevin equations (LQLEs) δ˙ˆc=−(iδc+λc)ˆc−iGncˆn+p 2λcˆcin, δ˙ˆn=−(i¯δn+λn)ˆn−iGncˆc−igndˆnssˆx+ω+p 2λnˆnin, δ˙ˆx=ωˆy, δ ˙ˆy=−ωˆx−γdˆy−gnd(ˆnssˆn†+ˆn∗ ssˆn)+η,(7) where ˆ nss=−iGnd√ 2gndis the magnon-phonon coupling. The quadrature fluctuations ( δXc,δYc,δXn,δYn,δx,δy) are described as δXc=(δˆc+δˆc†)/√ 2, δYc=i(δˆc†−δˆc)/√ 2 δXn=(δˆn+δˆn†)/√ 2, δYn=i(δˆn†−δˆn)/√ 2 We can rewrite equation (7) as ˙v(t)=Fv(t)+χ(t), (8)4 where v(t) =δXc(t),δYc(t),δXn(t),δYn(t),δx(t),δy(t)Tis the quadrature vector, χ(t) =√2λcXin c(t),√2λcYin c(t),√2λnXin n(t),√2λnYin n(t),0,η(t)Tis the noise vector and the drift matrix Fis written as F=−λcδc 0Gnc 0 0 −δc−λc−Gnc0 0 0 0Gnc−λn˜δn−Gnd 0 −Gnc0−˜δn−λn0 0 0 0 0 0 0 ωd 0 0 0Gnd−ωd−λd, (9) The system under consideration is considered stable if all the eigenvalues of a drift matrix, have real parts that are negative [55]. By using the Lyapunov equation, the system’s state is expressed as [56, 57] FC+CFT=−L, (10) whereCi j=1 2⟨vi(t)vj(t′)+vj(t′)vi(t)⟩(i,j=1,2,...,6) is the covariance matrix and L=diagλc(2Nc+1),λc(2Nc+1),λn(2Nn+ 1),λn(2Nn+1),0,λd(2Nd+1)is the di ffusion matrix achieved through Li jδ(t−t′)=1 2⟨χi(t)χj(t′)+χj(t′)χi(t)⟩. IV . SZIL ´ARD ENGINE L´eo Szil ´ard introduced Szil ´ard’s engine as a thought experiment in 1929. This experiment simplified the famous Maxwell’s demon paradox by using just one molecule of gas and replacing the demon with a mechanical device. Szil ´ard’s engine operates in four key steps: (i) The experiment begins with a single gas molecule bouncing around freely in a container with a volume ofV. (ii) A separator is placed inside a container, dividing it into two equal chambers with a volume of V/2, ensuring no heat exchange during the process. (iii) The engine’s function relies on determining the molecule’s location in the left or right chamber. According to the measurement results, a tiny weight has been attached to the same side of the partition using a pulley system. (iv) The final stage involves connecting the entire setup to a constant temperature heat source, allowing a gas molecule to expand and fill the container, crucial for the engine’s theoretical work. Szil ´ard’s engine, a concept that challenges our understanding of thermodynamics, involves a single molecule expanding to fill a container, absorbing heat from a constant temperature bath, and converting this heat into usable work by lifting the weight attached to the partition. The amount of work extracted, can be calculated using the formula W=kBTln 2, where kBis Boltzmann’s constant and ln(2) represents the information gain from measuring the molecule’s location. This process relies on connecting a weight to the partition and allowing the single gas molecule to expand in a controlled, constant temperature, i.e., isothermal manner. Szil ´ard’s engine can theoretically extract a specific amount of work per cycle, as described in [59] W=kBTln 2[1−H(X)] (11) The uncertainty about where the molecule is situated (left or right) can be quantified using a concept called Shannon entropy. This entropy is denoted by H(X)=−P xpxlnpx, where pxis the probability of capturing the molecule in each location ( x=R orx=L). Thus, the more uncertain we are about the molecule’s location, the higher the Shannon entropy will be. Equation (11) presents a potential challenge to the second law of thermodynamics. It suggests that under specific conditions, perfect knowledge about a the microscopic information of the system state might allow for work extraction using only a single heat bath. A significant link between information processing and the physical world was suggested by physicist Rolf Landauer in 1961. He theorized that whenever a single bit of information is erased in a computer system, it leads to an increase in energy dissipation as heat. This principle suggests a fundamental link between logical operations within a computer and the laws of thermodynamics that govern physical processes [60]. Recent experiments explore innovative techniques inspired by Maxwell’s demon and Landauer’s principle, which link information processing and energy dissipation, despite the potential to defy thermodynamic laws [61, 62]. By separating two entangled particles into di fferent containers, we essentially create two Szil´ard engines, AandB. These engines are unique because their functionality is intrinsically linked. Thus, the amount of work extractable from engine Ais dependent on the specific state of its entangled partner in engine B W(A|B)=kBTlog[1−H(A|B)] (12) Due to the entanglement, any event a ffecting engine Bhas an immediate impact on our understanding of engine A. Mutual information I(A:B)=H(A)−H(A|B)≥0 quantifies the link between AandB, with a non-negative value indicating that knowing the state of Breduces uncertainty about A. The reduced uncertainty leads to a significant increase in work extraction from engine Acompared to a scenario without entangled two Szil ´ard engines, as indicated by W(A|B) being greater than or equal toW(A), i.e., W(A|B)≥W(A).5 V . NEGATIVITY LOGARITHMIC, WORK EXTRACTION AND EFFICIENCY In this section, we will quantify and harness negativity logarithmic compared to the extracted work in a two-mode Gaussian state, shared by Alice ( A: photon) and Bob ( B: magnon). The e fficiency of a Szilard engine will be adopted. A. Negativity logarithmic The covariance matrix corresponding to the photon and magnon modes in the ( δXc(t),δYc(t),δXn(t),δYn(t)) basis can be expressed as CAB=CcCcn CT cnCn, (13) CcandCndepict the covariance matrix 2 ×2, respectively, representing the photon mode and magnon mode. The correlations between photon and magnon modes in standard form are denoted by Ccn Cc=diag(α,α),Cn=diag(β,β),Ccn=diag(∆,−∆). (14) For measuring bipartite entanglement, we employ the logarithmic negativity EN[59, 63, 64], that is given by Eom=max[0,−log(2ν−)], (15) whereν−=p Y− (Y2−4 detCAB)1/2/√ 2 is the minimum symplectic eigenvalue of the CAB, whereY=detCc+detCn− 2 detCcn. B. Magnon only performs Gaussian measurement The medium under consideration is a two-mode Gaussian state, i.e., photon and magnon modes. When Bob executes a Gaussian measurement on his assigned part of the system, the measurement has an impact on Alice’s state. This measurement can be described by ˜Nn(X)=π−1˜Dn(X)˜ρNn˜D† n(X), (16) where ˜Dn(X)=eXδˆn†−X∗δˆnis the displacement operator, ˜ ρNnis a pure Gaussian state without first moment and the its covariance matrix is given by ΓNn=1 2R(ξ)diag(λ,λ−1)R(ξ)T, (17) whereλis a positive real number, R(ξ)=[cosξ,−sinξ; sinξ,cosξ] is a rotation matrix and the detection of homodyne (hetero- dyne) is suggested by λ=0 (λ=1), individually. The outcome XBob gets from his measurement, it doesn’t a ffect the state of Alice’s mode δˆc, i.e.,CNn c|X=CNnc. The constrained state of mode A’s covariance matrix can be explicitly expressed as CNnc=Cc−C cn(Cn+ ΓNn)−1CT cn. (18) Bob measurement does push the state of mode Aout of equilibrium. However, by interacting with a heat bath for long time, mode Aeventually returns to an equilibrium state Ceq c. Its average entropy is solelyR dXpXS(CNn c|X)=S(CNnc) because her state is una ffected by the result. Work can be extracted by Alice from a surrounding heat bath [65] W=kBTh S(Ceq c)−S(CNnc)i . (19) We adopt the case of the covariance matrix in a squeezed thermal state, as depicted in equation (14) and Ceq c=Cc. The entropy of the covariance matrix described by equation (18) is quantified by considering the second-order R ´enyi entropy S2(ϱ)=−lnTrϱ2 [66]. In the case of two modes, Gaussian states (see equation (14)) are written as S2(CAB)=1 2ln(detCAB). (20)6 The extracted work, Eq. (19), became W(λ)=kBT 2ln detCc detCNnc! . (21) The extractable work for both homodyne ( λ=0) and heterodyne ( λ=1) detection in the case of STSs, writes as W(0) om=kBT 2ln αβ αβ−∆2! ,W(0) omSep=kBT 2ln 4αβ 2α+2β−1! ,W(0) omMax=kBT 2ln"4αβ 1+2|α−β|# . (22) W(1) om=kBTln"2αβ+α 2αβ+α−2∆2# ,W(1) omSep=kBTln"2α(2β+1) 4α+2β−1# ,W(0) omMax=kBTln 2α ifα≤β kBTlnh2α(1+2β) 1+4α−2βi otherwise(23) The works remain independent of the measurement angle. C. Both magnon and photon perform Gaussian measurement This subsection explores the case where Alice and Bob, each make Gaussian measurements on their state. Alice now performs a second Gaussian measurement on her reduced state of the system, it can be described by ˜Nc(X)=π−1˜Dc(Y)˜ρNc˜D† c(Y), (24) where ˜Dc(Y)=eYδˆc†−Y∗δˆcis the displacement operator, ˜ ρNccorresponds to a pure Gaussian state without first moment and the its covariance matrix is given by ΓNn=1 2R(χ)diag( Λ,Λ−1)R(χ)T, (25) where R(χ) represents a rotation matrix and Λ∈[0,∞]. The probability distribution describing a Gaussian measurement on Alice mode δˆcis influenced by the measurement ˜Nn(X) performed on Bob mode δˆn. However, interestingly, the uncertainty in Alice mode δˆcremains una ffected by the outcome Xthat Bob obtains from his Gaussian measurement, i.e., CNn,Nccn=CNnc+ ΓNc, WhileCNncis provided by Eq. (18). The extracted work by Alice (photon), can be measured via the Shannon entropy of the appropriate probability distribution H(Pr(X,Y)) is similar to the entropy of the Gaussian distribution H(CNn,Nccn). Its expression writes as W(λ,Λ)(ξ,χ)=kBT 2ln detCNnc detCNn,Nccn! . (26) In the case of STSs the extractable work for both homodyne ( λ=0) and heterodyne ( λ=1), writes as W(0,0) om(ξ,χ)=kBTlns 4αβ 4αβ−2∆2[1+cos(2ξ+2χ)],W(1,1) om=kBTln"(1+2α)(1+2β) 1+2β+α(2+4β)−4∆2# . (27) D. Efficiency of the work extraction According to Zhuang et al. (2014), the information-work e fficiency of a Szilard engine can be expressed as the ratio of extracted work to erasure work [67] µ=W Weras, (28) In this case, the information contained in the system is proportionate to Weras Weras=kBT H(P) ln 2, (29) where the Shannon entropy connected to the probability Pjdistribution is expressed as ?? We exploit the density operators ρand the von Neumann entropy to serve as the quantum mechanical equivalents of probability distributions [68] S(ρ)=Tr(ρlog(ρ)), (30)7 For two modes Gaussian state ρGthe von Neumann entropy sVcan be written as S(ρG)=2X l=1sV(Φl), (31) withΦl,l=1,2, represent the symplectic eigenvalues of the matrix CAB(see equation (13)) writes as Φ±=s κ±p κ2−4 detCAB 2, (32) andsVcan be expressed sV(w)=2w+1 2 log2w+1 2 −2w−1 2 log2w−1 2 , (33) whereκ=detCc+detCn+2 detCcn. E. Results and discussions In this section, we will explore how light (photons) and magnetic excitation’s (magnons) interact and share quantum corre- lations and e fficiency in a steady and dynamical state, considering various factors. We’ve selected parameters that are suitable for experimentation [35]: Ωc/2π=10×106Hz,Ωd/2π=10×106Hz,λd/2π=102Hz,λc/2π=λn/2π=1×106Hz, gnc/2π=Gnd/2π=3.2×106Hz, and at low temperature T=10×10−3K. Under these conditions, the coupling between the magnon mode and cavity mode gncis significantly weaker than the product of the detuning between the magnon and cavity modes and the mechanical resonance frequency, i.e., g2 nc≪|˜δnδc|≃Ω2 d. In this case, we adopt the approximate of the e ffective magnomechanical coupling as Gnd≃√ 2gndΩ Ωdsee Eq. (6), whereGnd/2π=3.2×106Hz leading to|⟨n⟩|≃ 1.1×107for a 250-µm-diameter YIG sphere, is regarding the drive magnetic field B0≃3.9×10−5T for gnd/2π≃0.2 Hz and the drive power P=8.9×10−3W. In this order, one can make the Kerr e ffect negligible because of the realization of the K|⟨n⟩|3≪ω. Eom Wom(0) WomSep(0) WomMax(0) 0.00 0.05 0.10 0.15 0.200.000.050.100.150.200.250.30 T(K)(a) Eom Wom(1) WomSep(1) WomMax(1) 0.00 0.05 0.10 0.15 0.200.000.050.100.150.200.25 T(K)(b) FIG. 1: Plot of logarithmic negativity Eom, extracted work W(λ) om(in units of kBT), maximum of extractable work W(λ) omMaxand extracted work at separable state W(λ) omS epbetween photon and magnon against temperature Tfor various Gaussian measurements. (a) λ=0 (homodyne); (b) λ=1 (heterodyne). In Fig. (1), we plot the logarithmic negativity Eom, extractable work W(λ) om(in units of kBT), separable work W(λ) omsepand maximum work W(λ) ommaxbetween optical mode and magnon mode versus the temperature Tfor di fferent measurements. The extractable work W(λ) omand separable work W(λ) omsepare always bound by maximum work W(λ) ommax, as depicted in Fig. (1). We remark that photon and magnon modes are entangled in the region where W(λ) om>W(λ) omsep. This agrees with entanglement quantified by logarithmic negativity Eom[69]. This figure exhibits that W(λ) om,W(λ) omsepandW(λ) ommaxall increase with increasing temperature. Conversely, logarithmic negativity diminishes to zero around 0.17 K., i.e., the two modes photon and magnon are in separable state and W(λ) om≤W(λ) omsep, as depicted in Fig. (1)(a-b). We note that for a large value of the temperature Tthe mode corresponds to8 the optimal performance of a Szilard engine. This is for homodyne and heterodyne detection ( λ=0,1). Besides, the maximum work is larger at high temperatures. Furthermore, in homodyne detection, the maximum work W(λ) ommax(in units of kBT) achieves 0.30 at T=0.2 K (a), while in heterodyne detection it achieves 0.27, as depicted in figure (1). Thus, one can say that extractable work provides a su fficient condition to witness entanglement in generic two-mode states, which is also necessary for squeezed thermal states. Eom Wom(0) WomSep(0) WomMax(0)0.0 0.2 0.4 0.6 0.8 1.00.00.10.20.30.40.50.60.7 t(μs)(a) Eom Wom(1) WomSep(1) WomMax(1)0.0 0.2 0.4 0.6 0.8 1.00.00.10.20.30.40.50.60.7 t(μs)(b) FIG. 2: Time evolution of logarithmic negativity Eom, extracted work W(λ) om(in units of kBT), maximum of extractable work W(λ) omMaxand extracted work at separable state W(λ) omS epbetween photon and magnon for various Gaussian measurements. (a) λ=0 (homodyne) (b) λ=1 (heterodyne). In Fig. 2, we plot the time-evolution of the bipartite entanglement Eom, extractable work W(λ) om(in units of kBT), separable work W(λ) omsepand maximum work W(λ) ommaxbetween optical mode and magnon mode in homodyne measurement (a) and in hetrodyne measurement (b). This figure shows three entanglement regimes: The first regime is dedicated to classically correlated states (Eom=0), i.e., W(λ) om<W(λ) omsep. This means that the two modes (photon and magnon) are separated. Nevertheless, the extracted work increases in time while the separable work and maximum work decrease due to decoherence from the thermal bath. The second regime, W(λ) om>W(λ) omsepandEom>0, indicating entanglement sudden death between the two modes. Here, we observe the generation of oscillations, which can be explained by the Sorensen-Molmer entanglement dynamics discussed in Ref. [43, 70]. The last regime corresponds to a steady state, i.e., W(λ) omsepremains bounded by W(λ) omandEomis constant. The extracted work W(λ) om and separable work W(λ) omsepare bounded by the maximum work W(λ) ommax. Thus, the engine has the best performance for strongly squeezed vacuum states and small times of evolution. Eom Wom(0) WomSep(0) WomMax(0)-2.0 -1.5 -1.0 -0.5 0.00.000.050.100.150.200.25 δc/Ωd(a) Eom Wom(1) WomSep(1) WomMax(1)-2.0 -1.5 -1.0 -0.5 0.00.000.050.100.150.20 δc/Ωd(b) FIG. 3: Plot of logarithmic negativity Eom, extracted work W(λ) om(in units of kBT), maximum of extractable work W(λ) omMaxand extracted work at separable state W(λ) omS epbetween photon and magnon versus the normalized photon detuning for various Gaussian measurements. (a) λ=0 (homodyne) (b) λ=1 (heterodyne).9 Figure 4 presents the influence of normalized photon detuning δc/ωdon logarithmic negativity Eom, extractable work W(λ) om (in units of kBT), separable work W(λ) omsep, and maximum work W(λ) ommaxas functions of normalized magnon detuning δa/ωb. En- tanglement W(λ) om>W(λ) omsep, as expected from the logarithmic negativity Eom[69], is observed in Fig. 4, while the separable state Eom=0 and W(λ) om≤W(λ) omsepis depicted in Fig. 4(a-b). We remark that the peak in the logarithmic negativity corresponds to the dip in W(λ) omandW(λ) omsepfor both homodyne and heterodyne measurements. Eom Wom(0) WomSep(0) WomMax(0)0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.40.000.050.100.150.200.250.30 gnc/Ωd(a) Eom Wom(1) WomSep(1) WomMax(1)0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.40.000.050.100.150.200.250.30 gnc/Ωd(b) FIG. 4: Plot of logarithmic negativity Eom, extracted work W(λ) om(in units of kBT), maximum of extractable work W(λ) omMaxand extracted work at separable state W(λ) omS epbetween photon and magnon as a function of the magnon-photon coupling gcn/ωdfor various Gaussian measurements. (a)λ=0 (homodyne) (b) λ=1 (heterodyne). Figure ??explores the logarithmic negativity Eom, extractable work W(λ) om(in units of kBT), separable work W(λ) omsep, and maxi- mum work W(λ) ommaxbetween the optical mode and magnon mode versus temperature for both homodyne ( λ=0) and heterodyne (λ=1) measurements, as a function of the magnon-photon coupling gcn/ωd. As expected, entanglement W(λ) om>W(λ) omsepcoincides with non-zero logarithmic negativity Eom[69]. Conversely, the separable state W(λ) om≤W(λ) omsepandEom=0 indicates separabil- ity, as shown in Fig. ??(a-b). Interestingly, Fig. ??also shows that Eom,W(λ) om,W(λ) omsep, and W(λ) ommaxall increase with increasing magnon-photon coupling ( gcn/ωd) before gradually decreasing after reaching a maximum value. Additionally, we observe that in homodyne detection, W(λ) ommax>0 even for gcn/ωd=0, whereas in heterodyne detection, W(λ) ommax=0 for gcn/ωd=0. Wom(0) Wom(0,0) Wom(1) Wom(1,1)0.00.10.20.30.40.50.025330.025340.025350.025360.025370.025380.025390.02540 0.0 0.1 0.2 0.3 0.4 0.50.0000.0050.0100.0150.0200.0250.0300.035 T(K) FIG. 5: Plot of the extractable work W(in units of kBT) as a function of the temperature Tfor both measurement W(0,0) omandW(1,1) omand single homodyne measurement W(0) omand heterodyne measurement W(1) om.10 In Fig. 5, we represent the comparison between the extracted work from both measurement W(0,0) om(0,0) and W(1,1) omto a single homodyne measurement W(0) omand heterodyne measurement W(1) om. The decrease in extractable work observed in Fig. 5 is attributed to the second measurement introducing entropy into the system, which can be mathematically represented as a smearing of the distribution imparted by the single measurement, i.e., W(1,1) om<W(1) om, just for homodyne detection appears like identical W(0,0) om<W(0) om. λ=0 λ=1 0.00 0.05 0.10 0.15 0.200.040.050.060.070.080.09 T(K)μ(a) λ=0 λ=1 0.0 0.5 1.0 1.5 2.00.000.020.040.060.08 t(μs)μ(b) FIG. 6: Plot of the e fficiency of the work extraction as a function of (a) temperature Tand (b) time versus t(µs) for various Gaussian measurements with λ=0 (homodyne) and λ=1 (heterodyne). Fig. 6(a) shows information-work e fficiency monotonically decreasing towards zero with increasing temperature for both homodyne and heterodyne detection ( λ=0,1). The engine performs best at low temperatures. Notably, homodyne and hetero- dyne measurements achieve similar e fficiency. Fig. 6(b) explores the time-dependence of e fficiency. Here, we see e fficiency monotonically increasing to reach a steady-state value, indicating better e fficiency for longer times. Besides, the e fficiency for homodyne measurement is bound by homodyne measurement, as depicted in Fig. 6. VI. CONCLUSION In summary, we have discussed a possible strategy to measure the entanglement and separability of the two-mode Gaussian state in a steady and dynamical state by harnessing the extracted work (out of a thermal bath) by means of a correlated quantum system subjected to measurements. Our investigation focuses on a cavity magnomechanical system. Here, a microwave cavity mode is coupled with a magnon mode in a Yttrium Iron Garnet (YIG) sphere. This magnon mode further couples to a mechanical mode through the magnetostrictive interaction. We have used logarithmic negativity to quantify the entanglement between photons and magnons with experimentally reachable parameters. We have shown that the detecting of the entanglement between optic and magon modes is realized when W(λ) om>W(λ) omsep(λ=1,2), and it is in agreement with logarithmic negativity. We show that when W(λ) om≤W(λ) omsep(λ=1,2). Besides, extracted and separable work are always bounded by the maximum work in homodyne and hetrodyne measurement. We have found that all works are influenced by di fferent parameters like temperature T, normalized magnon detuning δc/ωd, magnon-photon coupling gnc. Also, we have shown that the information-work e fficiency of Szilard engine is better for small temperature and for large time. Besides, the e fficiency enhances in heterodyne measurement with request to the homodyne measurement. [1] A. Einstein, B. Podolsky and N. Rosen. Phys. Rev. 47, 777 (1935). [2] E. Schr ¨odinger. Mathematical Proceedings of the Cambridge Philosophical Society 31, 555 (1935). [3] J. S. Bell. On the Einstein Podolsky Rosen paradox. Physics Physique Fizika 1, 195 (1964). [4] A.K. Ekert. Quantum cryptography based on Bell’s theorem. Phys. Rev. Lett. 67, 661 (1991). [5] C.H. Bennett, G. Brassard, C. Crepeau, R. Jozsa, A. Peres and W.K. Wootters. Phys. Rev. Lett. 70, 1895 (1993). [6] C. H. Bennett and D. P. DiVincenzo. Nature 404, 247 (2000). [7] M. Aspelmeyer, T. J. Kippenberg and F. Marquardt. Rev. Mod. Phys. 86, 1391 (2014). [8] D. Vitali, et al., Phys. Rev. Lett. 98, 030405 (2007).11 [9] M. Abdi, S. Pirandola, P. Tombesi and D. Vitali. Physical Review Letters. 109, 143601 (2012). [10] E. A. Sete, H. Eleuch and C. R. Ooi. JOSA B. 31(11), 2821-2828 (2014). [11] M. Abdi, S. Pirandola, P. Tombesi and D. Vitali. Physical Review A. 89, 022331 (2014). [12] E. A. Sete and H. Eleuch. Physical Review A. 89(1), 013841 (2014). [13] J. Manninen, M. Asjad, E. Selenius, R. Ojajarvi, P. Kuusela and F. Massel. Physical Review A 98, 043831 (2018). [14] M. Amazioug, M. Nassik and N. Habiballah. Eur. Phys. J. D 72, 171 (2018). [15] M. Asjad, M. A. Shahzad and F. Saif. The European Physical Journal D 67, 1 (2013). [16] M. Amazioug, B. Maroufi and M. Daoud. Quantum Inf. Process. 19, 16 (2020). [17] B. Teklu, T. Byrnes and F. S. Khan. Phys. Rev. A 97, 023829 (2018). [18] T. G. Tesfahannes. Josa B. 37, A245-A252 (2020). [19] H. D. Mekonnen, T. G. Tesfahannes, T. Y . Darge and A. G. Kumela. Scientific Reports. 13, 13800 (2023). [20] J. Teufel, T. Donner, D. Li, J. Harlow, M. Allman, K. Cicak, A. Sirois, J. Whittaker, K. Lehnert and R. Simmonds. Nature 475359 (2011). [21] S. Machnes, J. Cerrillo, M. Aspelmeyer, W. Wieczorek, M.B. Plenio and A. Retzker. Phys. Rev. Lett. 108, 153601 (2012). [22] M. Asjad, N. E. Abari, S. Zippilli and D. Vitali. Optics Express 27, 32427 (2019). [23] J. Chan, T. P. M. Alegre, A.H. Safavi Naeini, J.T. Hill, A. Krause, S. Groblacher, M. Aspelmeyer and O. Painter. Nature 478(7367) 89 (2011). [24] M. Bhattacharya and P. Meystre. Phys. Rev. Lett. 99073601 (2007). [25] M. Amazioug, M. Daoud, S. K. Singh and M. Asjad. Quantum Information Processing, 22(8), 301 (2023). [26] J.Q. Liao and L. Tian. Phys. Rev. Lett. 116, 163602 (2016). [27] H. Xiong, Z.X. Liu and Y . Wu. Opt. Lett. 42, 3630 (2017). [28] A. Candeloro, S. Razavian, M. Piccolini, B. Teklu, S. Olivares and M. G. A. Paris. Entropy 23(10) (2021). [29] J. D. Teufel et al. , Nature (London) 475, 359 (2011). [30] R. Riedinger et al. Nature (London) 556, 473 (2018). [31] C. F. Ockeloen-Korppi et al. Nature (London) 556, 478 (2018). [32] Q. A. Turchette, C. J. Hood, W. Lange, H. Mabuchi and H. J. Kimble. Phys. Rev. Lett. 75, 4710 (1995). [33] B. T. H. Varcoe, S. Brattke, M. Weidinger and H. Walther. Nature 403, 743 (2000). [34] G. Nogues, A. Rauschenbeutel, S. Osnaghi, M. Brune, J. M. Raimond and S. Haroche. Nature 400, 239 (1999). [35] X. Zhang, C.-L. Zou, L. Jiang and H. X. Tang. Sci. Adv. 2, e1501286 (2016). [36] C. Kittel. Phys. Rev. 73, 155 (1948). [37] H. Huebl et al., Phys. Rev. Lett. 111, 127003 (2013). [38] D. Zhang et al. npj Quantum Information 1, 15014 (2015). [39] Y .-P. Wang et al. , Phys. Rev. Lett. 120, 057202 (2018). [40] Y . Tabuchi et al. , Science 349, 405 (2015). [41] B. Hussain, S. Qamar and M. Irfan. Physical Review A. 105, 063704 (2022). [42] M. Asjad, J. Li, S.-Y . Zhu and J. Q. You. Fundam. Res. 3, 3 (2023). [43] M. Amazioug, B. Teklu and M. Asjad. Scientific Reports. 13, 3833 (2023). [44] M. Amazioug, S. Singh, B. Teklu and M. Asjad. Entropy. 25, 1462 (2023). [45] K. Maruyama, F. Nori and V . Vedral. Rev. Mod. Phys. 81, 1 (2009) [46] J.M.R. Parrondo, J.M. Horowitz and T. Sagawa. Nat. Phys. 11, 131 (2015) [47] J. Oppenheim, M. Horodecki, P. Horodecki and R. Horodecki. Phys. Rev. Lett. 89, 180402 (2002) [48] K. Maruyama, F. Morikoshi and V . Vedral. Phys. Rev. A 71, 012108 (2005) [49] K. Funo, Y . Watanabe and M. Ueda. Phys. Rev. A 88, 052319 (2013) [50] J. Li, S.-Y . Zhu and G. Agarwal. Phys. Rev. Lett. 121, 203601 (2018). [51] C. Kittel. Phys. Rev. 110, 836 (1958). [52] Y .-P. Wang et al., Phys. Rev. B 94, 224410 (2016). [53] C. W. Gardiner and P. Zoller, Quantum Noise (Springer, Berlin, Germany, 2000). [54] V . Giovannetti and D. Vitali, Phys. Rev. A 63, 023812 (2001); R. Benguria and M. Kac, Phys. Rev. Lett. 46, 1 (1981). [55] E.X. DeJesus and C. Kaufman, Phys. Rev. A 35, 5288 (1987) [56] D. Vitali et al., Phys. Rev. Lett. 98, 030405 (2007). [57] P. C. Parks and V . Hahn, Stability Theory . Prentice Hall, (1993). [58] G. Vidal and R. F. Werner, Phys. Rev. A 65, 032314 (2002); M. B. Plenio, Phys. Rev. Lett. 95, 090503 (2005). [59] M. B. Plenio. Physical Review Letters. 95, 090503 (2005). [60] R. Landauer. IBM J. Res. Dev. 5(3), 183–191 (1961). [61] J. V . Koski, A. Kutvonen, I. M. Khaymovich, T. Ala-Nissila and J. P. Pekola, Phys. Rev. Lett. 115, 260602 (2015). [62] M. D. Vidrighin, O. Dahlsten, M. Barbieri, M. S. Kim, V . Vedral and I. A. Walmsley. Phys. Rev. Lett. 116, 050401 (2016). [63] G. Vidal and R. F. Werner. Physical Review A. 65, 032314 (2002). [64] G. Adesso, A. Serafini and F. Illuminati. Physical Review A. 70, 022318 (2004). [65] M. Brunelli, M. G. Genoni, M. Barbieri and M. Paternostro. Physical Review A. 96(6), 062311 (2017) [66] G. Adesso, D. Girolami and A. Serafini, Phys. Rev. Lett. 109, 190502 (2012). [67] Z. Zhuang and S.-D. Liang. Phys. Rev. E 90, 052117 (2014). [68] M. Cuzminschi, A. Zubarev, S. M. Iordache and A. Isar. iScience, 26(12) (2023). [69] J. Li, S.-Y . Zhu and G. S. Agarwal. Phys. Rev. Lett. 121, 203601 (2018). [70] J. Li, G. Li, S. Zippilli, D. Vitali and T. Zhang. Phys. Rev. A 95, 043819 (2017). [71] T. Holstein and H. Primako ff, Phys. Rev. 58, 1098 (1940). | 2024-05-29 | The connections between thermodynamics and quantum information processing are
of paramount importance. Here, we address a bipartite entanglement via
extracted work in a cavity magnomechanical system contained inside an yttrium
iron garnet (YIG) sphere. The photons and magnons interact through an
interaction between magnetic dipoles. A magnetostrictive interaction, analogous
to radiation pressure, couple's phonons and magnons. The extracted work was
obtained through a device similar to the Szil\'ard engine. This engine operates
by manipulating the photon-magnon as a bipartite quantum state. We employ
logarithmic negativity to measure the amount of entanglement between photon and
magnon modes in steady and dynamical states. We explore the extracted work,
separable work, and maximum work for squeezed thermal states. We investigate
the amount of work extracted from a bipartite quantum state, which can
potentially determine the degree of entanglement present in that state.
Numerical studies show that entanglement, as detected by the extracted work and
quantified by logarithmic negativity, is in good agreement. We show the
reduction of extracted work by a second measurement compared to a single
measurement. Also, the efficiency of the Szilard engine in steady and dynamical
states is investigated. We hope this work is of paramount importance in quantum
information processing. | Detection of entanglement by harnessing extracted work in an opto-magno-mechanics | 2405.19205v1 |
arXiv:1807.08614v2 [physics.gen-ph] 6 Aug 2018Magnon-phonon conversion experiment and phonon spin S. C. Tiwari Department of Physics, Institute of Science, Banaras Hindu Unive rsity, Varanasi 221005, and Institute of Natural Philosophy Varanasi India Recent experiment demonstrates magnon to phonon conversio n in a YIG film under the appli- cation of a non-uniform magnetic field. Light scattered from phonons is observed to change its polarization state interpreted by the authors signifying p honon spin. In this note we argue that the experimental data merely shows the exchange of angular mome ntum±¯hper photon. We suggest that it has physical origin in the orbital angular momentum o f phonons. The distinction between spin and orbital parts of the total angular momentum, and bet ween phonons and photons with added emphasis on their polarizations is explained. The mai n conclusion of the present note is that phonon spin hypothesis is unphysical. PACS numbers: 63.20.-e, 63.20.kk I. INTRODUCTION Magneto-elastic waves or magnon-phonon excitations have been of interest for various reasons, one of them be- ing the field of spintronics. A recent experimental study on the magnon-phonon conversion in the ferrimagnetic insulator YIG addresses a question of fundamental im- portance whether phonons carry spin [1]. We recall that in 1988 McLellan [2] showed that sharp angular momen- tum could be attributed to circular or elliptical phonon polarizations. Note that this angular momentum can- not be identified with the spin of phonon. In this note we discuss the recent experiment [1] and argue that the measurements show that phonons exchange angular mo- mentum with light but it is not spin of the phonons. We emphasize that this distinction is not just semantic [3] but of fundamental nature [4]. The experiment [1] first shows by time-resolved measurements that under the application of a non- uniform magnetic field on a YIG film, spin wavepackets launched by pulsed microwave signals, convert into elas- tic wavepackets, i.e. magnon-phonon conversion. Next using wavevector resolved Brillouin light scattering ex- periment the measurementsshow i) magnon-phononcon- version with constant energy and linearly varying mo- mentum, and ii) the light scattered by the phonons is circularly polarized. The meticulous data presented in Figures (4) and (5) of their paper by the authors could hardly be doubted. The question that concerns me is regarding their claim, ’that phonons created by the con- version of magnons do carry spin’. It is true that the change in the polarization state of light involves exchange of angular momentum, for ex- ample, to transform linearly polarized light to circularly polarized light an angular momentum of ±¯hper photon is required. However this angular momentum need not be associated with the spin of the medium or the light- scattering object. At the macroscopic level, Beth experi- ment [5] detected a direct mechanical effect in terms of a torque exerted by a circularly polarized beam of light ona doubly refracting medium which changes the polariza- tion state of light. The photon spin angular momentum is transferred to the body of the medium imparting or- bital rotation; the aim of the Beth experiment was, of course, to demonstrate that photons had spin. Holanda et al experiment [1], on the other hand, assumes photon spin, andinfersthatphononscarryspin frommicroscopic scattering data with photons. The crucial point is that the experiment only proves that angular momentum in the unit of ±¯his exchanged. It cannot be attributed to the phonon spin: non-zero spin of phonon does not make physical sense. We argue that the orbital angular momentum of elastic waves or phonons is responsible for angular momentum transfer. In this note we address the question: Why not spin? First photon physics is briefly reviewed in the next sec- tion to highlight the intricate relationship between polar- ization and spin. In section III elementary discussion on phononsshowsthatphoton-phononanalogyisuntenable, and spin cannot be associated with polarized phonons. FurtherelaborationconstitutessectionIV.Thenoteends with a short conclusion. II. PHOTON AND LIGHT POLARIZATION A brief review on elementary considerations on the meaning of angular momentum and its decomposition into orbital and spin parts seems necessary. For a sys- tem with rotational symmetry the angular momentum is a constant of motion; if linear momentum is pone de- fines angular momentum simply as r×p. In field theory one may construct the expression for angular momentum from the momentum density of the field or directly cal- culate the angular momentum density tensor as Noether current from the rotational invariance of the action. In general, it is useful to separate total angular momentum Jinto orbital and spin components J=L+S (1)2 For a scalar particle the spin part is zero. For a vector particle, following a textbook discussion [6], see pp197- 198, a simplified picture is obtained in terms of Lthat depends on the space or position and spin part that de- pends on the three components of the vector wavefunc- tionVwhere 3 ×3 spin matrices Sx,Sy,Szact onV, see (27.11) in [6]. Note that similar arguments hold for the second quantized field theory. Field Vbecomes an operator Vop=aλVλ+a† λV∗ λ (2) The annihilation and creation operators aλanda† λfor a modeλsatisfy the commutation rules [aλ,a† λ′] =δλλ′ (3) ThefieldoperatorsactinFockspacespannedbytheFock state vectors. The orbital angular momentum operator L=−i(r×∇) (4) acts on the space-dependence of the mode functions Vλ, whereas the spin operator S=−iǫijk (5) acts on the components of Vλ. Photonisavectorparticlewithrestmasszeroandspin one having only two projections - better understood in terms of helicity. Photon as a quantized electromagnetic radiation field continues to have fundamental questions: gauge-invariance, transversality and Lorentz covariance arecontroversialand unsettled issues, seereferencescited in [4] and also [7, 8]. Let us try to explain the problem. Classical fields E,B,Aµsatisfy the wave equation, and one assumes a plane wave representation. Introducing canonically con- jugate field variables canonical quantization is carried out. In the normal mode expansion the annihilation and creation operators can be defined, and polarization 4- vectorǫµcomprising of four mutually orthogonal unit vectors takes care of the vector nature of the field. In QED a manifest Lorentz covariant quantization results into longitudinal and time-like photons besides the phys- ical photons. HoweverincontrasttoQEDwheretheelectromagnetic potentials Aµarefundamentalfieldvariables,inquantum optics literature the utility of the electric and magnetic field operators is well known. In a simpler field quantiza- tion for the radiation field polarization index s= 1,2 for transverse fields is sufficient. The normalized eigenstate of the number operator nks=a† ksaksgives the number of photons in the mode ( k,s) as nks|nks>=nks|nks> (6) The Fock state is a direct product of number states over all possible modes |{n}>=/productdisplay ks|nks> (7)The assumption of transverse mode functions, for exam- ple,A⊥eliminates longitudinal and time-like photons in quantum optics. Physical quantities like energy, momentum and angu- lar momentum are obtained using their classical expres- sions and transforming them to the quantized field oper- ators. In the classical radiation field theory the Poynting vectorE×Brepresents the momentum density and the total angular momentum density becomes J=r×(E×B) (8) Separation of (8) into orbital and spin parts can be made similar to (1). The spin angular momentum density is identified with the expression S=E×A (9) Regarding spin angular momentum a remarkable result pointed out by van Enk and Nienhuis [9] is worth men- tioning. For a circularly polarized plane wave it is found that the spin operator corresponding to (9) has a simple form Sr=/summationdisplay kk |k|(nk+−nk−) (10) Heres=±for right and left circular polarization. The components of Srcommute with each other [Sr i,Sr j] = 0 (11) Authors [9] argue that the spin operator(10) cannot gen- erate polarization rotation of the field, and cannot be in- terpreted as spin angular momentum of photon. Note that Jauch and Rohrlich [10] define Stokes operators sat- isfying the angular momentum commutation rules which provide interpretation of the photon spin [11]. To conclude this section, in both QED and quantum optics photon spin and the role of polarization state in- volve intricate issues. One thing is, however unambigu- ous, namely that spin angular momentum is an intrinsic property associated purely with the nature of the fields. In fact, spin for electron also depends only on the Dirac field Σ= Ψ†γγ5Ψ (12) III. PHONON SPIN In the abstract of [1] the authors state that, ’while it is well established that photons in circularly polarized light carry a spin, the spin of phonons has had little attention in the literature’. Now keeping in mind the conceptual problems associated with photon physics highlighted in the preceding section the photon spin has to be inter- preted with great care. The second part of the statement is, however not correct. The condensed matter literature tacitly accepts phonon to be a zero spin boson, in spite3 of the transverse modes and the known polarization of acoustic and optical phonons. Polarization of phonon modes is not related with spin but orbital angular mo- mentum [2]. A brief discussion seems useful for the sake of clarity. Phonons are quantized lattice vibrations; phonon modes are described by wavevector k, a branch number j and the orientation of the coordinate axes [2]. The branch number has two values for crystals with two sub- lattices and there are two triplets of phonons for acoustic and optical branches. McLellan defines phonon angular momentum in terms of phonon annihilation and creation operators to be Lph=/summationdisplay kjakj×a† kj(13) This expression is, as pointed by the author [2], in agree- ment with that defined using the displacement vector ulκ L=/summationdisplay lκulκ×plκ (14) Here the index lcorresponds to the unit cell and κfor the atom on a sub-lattice. Expression (52) in [2] for the total angular momentum of the lattice includes that of the rigid body rotation of the crystal. What are the implications of above discussion? It throws light on the issue of phonon polarization and spin as follows. [1]Phononisaquasi-particlehavingnodynamicalfield equations like Maxwell field equations for photon. The most crucial point that seems to have gone unnoticed in the discussions on phonon spin and phonon-photon anal- ogy is that the displacement vector representing lattice vibrations is a real space coordinate. Canonical quanti- zation and the field operators for phonons are based on the coordinate and momentum, for example those ap- pearing in Eq.(14). On the other hand, for photon the field variable Aµis treated as a coordinate variable, and ∂L ∂˙Aµis the canonically conjugate “momentum” variable for the quantization. Here Lis the Lagrangian density of the Maxwell field. [2] Phonon polarization is physically entirely differ- ent than light or photon polarization. McLellan’s analysis clearly establishes the physical significance of phonon polarization in terms of orbital angular momen- tum. Isotropic 2D quantum oscillator best illustrates the meaning of polarization of elastic waves or phonons. In cartesian coordinates the raising and lowering operators separate into 1D oscillators; it is akin to linear polar- ization. A circular basis ( a† x±ia† y) formally resembles circular polarization. In the circular basis one gets well- defined orbital angular momentum of the oscillator. Transverse modes in paraxial optics also represent physical realizationof this example. First order Hermite- Gaussian modes HG10andHG01are not eigenstates of angular momentum operator (4). However, Laguerre-Gaussian modes LG±1 0=1√ 2(HG10±iHG01) (15) possess sharp angular momentum. Thus phonon polar- ization is related with orbital angular momentum not spin. IV. DISCUSSION Let us try to elucidate further why photon-phonon analogy is misleading. Photon as a quantized vector field has intrinsic spin one. Wigner’s group theoretical argu- ments establishthat foranymasslessorlight-likeparticle with non-zero spin there exist only two helicity states. In the classical picture the intrinsic spin is identified with the vector product of the electric field and the vector po- tential (9). Assumed transverse vector potential leads to the electric field E⊥=−∂A⊥ ∂t(16) In the field quantization assuming monochromatic light the electric and magnetic fields are obtained using (16) and∇×A⊥respectively, e. g. the expression (6) in [9]. The oscillations or vibrations around equilibrium posi- tion of ions collectively lead to the elastic waves and are analyzed in the harmonic approximation in terms of the normal modes. The mode expansion includes wave vec- tor and polarization specifications [12, 13]. Phonon field is understood in terms of the displacement of a point in the material medium u(r,t) and the corresponding mo- mentum p=/integraldisplay ρ˙u(r,t)dV (17) whereρis the mass density. Standard coordinate and momentum quantization rule, and plane wave represen- tation yield quantized phonon field. It is easy to see that expression (14) is just the orbital angular momentum. A deceptive formal analogy with the photon spin expres- sion (9) is obvious considering expression (16) and using (17) for phonon. Physical interpretation depends on the fundamentaldistinctionbetweenthe vectorpotentialand the displacement vector since the later is a real space co- ordinate variable. Thus the suggested interpretation for the phonon angular momentum corresponding to the cir- cularly polarizedmodes in [2] seems justified. We remark that in spite of the usage of phonon polarization in the literature [13], and transverse polarization in the des- ignation of creation and annihilation operators phonon spin and vector nature of the phonon field is nowhere mentioned. To avoid confusion, it has to be understood that scalar field could possess well defined orbital angu- lar momentum and laser light beams with sharp orbital angular momentum have been extensively studied in the4 literature, see references in [4]. Longitudinal modes have no spin or orbital angular momentum, however linearly polarized light could possess orbital angular momentum but not spin. Thus the conventional phonon theory has no analogy with the photon theory, and non-zero phonon spin does not make physical sense. The origin of the an- gular momentum transfer from phonon to photon in the reported experiment [1] may be logically attributed to the orbital angular momentum of phonons. In a hypothetical scenario assuming phonon has spin one it would be of interest to find its physical conse- quences. I think electron-phonon interaction and Cooper pair formation via phonon mediated electron-electron in- teraction may be re-examined: phonon creation and an- nihilation operators [13] could be generalized for the cir- cularly polarized modes in the interaction Hamiltonian and treated as spin one particles. There is another prob- leminsuperconductivityhighlightedbyPost[14], namely the angular momentum conservation in a superconduct- ing ring. Though Post sets the problem in the form of Onsager-Feynman controversyhe offers insightful discus- siononthe mechanismoftheangularmomentumbalance when supercurrent in a ring vanishes as the temperature is raised above the transition temperature. Note that Post rules out any role of lattice, therefore it may be of interest to examine the role of phonon spin in this prob- lem. We could, of course, explore new physics or unconven- tional ideas [4]. Departing from the phonon picture new kind of field excitations in the spirit of Cosserat medium was suggested in [4]. Analogy of displacement vectorwith the vector potential is not justified, however the velocity field in a rotating fluid may be treated as a vec- tor potential: postulating rotating space-time fluid with nontrivial topology of vortices we have re-interpreted the electromagnetic field tensor as the angular momentum (density or more appropriately flux) of photon fluid [8], and proposed a topological photon [15]. Note that the netangularmomentumofthemicroscopicparticlesinthe rotating fluid implies antisymmetric stress tensor. Such speculations relate spin with topological invariants. V. CONCLUSION It has been pointed out [4] that phonon angular mo- mentum discussed in [16] is ambiguous as compared to that discussed in [2]. We have shown that non-zero phonon spin hypothesis and phonon-photon analogy [17] are conceptually flawed, giving further support to the ar- gumentspresentedin [4]. Phononspin hasno experimen- tal evidence. The correct physical interpretation of the reported experiment [1] is that orbital angular momen- tum of phonons is exchanged with light beam resulting into the change in the polarization of the light. Acknowledgments I thank S. Streub for raising specific questions on the photon-phonon analogy. I also acknowledge correspon- dence with S. M. Rezende, M. Wakamatsu, and A. Hoff- mann, and conversation with D. Sa and V. S. Subrah- manyam. [1] J. Holanda et al, Nature Physics 14, 500 (2018) [2] A. G. McLellan, J. Phys. C Solid State Phys. 21, 1177 (1988) [3] S. Streub et al, Phys. Rev. Lett. 121, 027202 (2018); arXiv: 1804.07080v1 [cond-mat.mes-hall] [4] S. C. Tiwari, arXiv: 1708.07407v3 [physics.gen-ph] [5] R. A. Beth, Phys. Rev. 48, 471 (1935) [6] L. I. Schiff, Quantum Mechanics (McGraw-Hill, 1968) Third Edition [7] S. C. Tiwari, arXiv:08070.0699v1 [physics-gen.ph] [8] S. C. Tiwari, J. Mod. Optics, 46, 1721 (1999) [9] S. J. van Enk and G. Nienhuis, Europhys. Lett. 25, 497(1994) [10] J. M. Jauch and F. Rohrlich, The Theory of Photons and Electrons (Reading, Addison-Wesley, 1955) [11] S. C. Tiwari, J. Mod. Optics, 39, 1097 (1992) [12] L. D. Landau and E. M. Lifshitz, Theory of Elasticity (Pergamon, 1970) [13] D. J. Scalapino, Chapter 10in SuperconductivityVolum e 1, Edited by R. D. Parks (M. Dekker, 1969) [14] E. J. Post, Quantum Reprogramming (Kluwer, 1995) [15] S. C. Tiwari, J. Math. Phys. 49, 032303 (2008) [16] L. ZhangandQ.Niu, Phys.Rev.Lett.112, 085503 (2014) [17] D. A. Garanin and E. M. Chudnovsky, Phys. Rev. B 92, 0244421 (2015) | 2018-07-17 | Recent experiment demonstrates magnon to phonon conversion in a YIG film
under the application of a non-uniform magnetic field. Light scattered from
phonons is observed to change its polarization state interpreted by the authors
signifying phonon spin. In this note we argue that the experimental data merely
shows the exchange of angular momentum $\pm \hbar$ per photon. We suggest that
it has physical origin in the orbital angular momentum of phonons. The
distinction between spin and orbital parts of the total angular momentum, and
between phonons and photons with added emphasis on their polarizations is
explained. The main conclusion of the present note is that phonon spin
hypothesis is unphysical. | Magnon-phonon conversion experiment and phonon spin | 1807.08614v2 |
arXiv:1302.1352v1 [cond-mat.mes-hall] 6 Feb 2013Theory of spin Hall magnetoresistance Yan-Ting Chen1, Saburo Takahashi2, Hiroyasu Nakayama2, Matthias Althammer3,4, Sebastian T. B. Goennenwein3, Eiji Saitoh2,5,6,7, and Gerrit E. W. Bauer2,5,1 1Kavli Institute of NanoScience, Delft University of Techno logy, Lorentzweg 1, 2628 CJ Delft, The Netherlands 2Institute for Materials Research, Tohoku University, Send ai, Miyagi 980-8577, Japan 3Walther-Meißner-Institut, Bayerische Akademie der Wisse nschaften, 85748 Garching, Germany 4University of Alabama, Center for Materials for Informatio n Technology MINT, Dept Chem, Tuscaloosa, AL 35487, USA 5WPI Advanced Institute for Materials Research, Tohoku University, Sendai 980-8577, Japan 6CREST, Japan Science and Technology Agency, Sanbancho, Tok yo 102-0075, Japan and 7The Advanced Science Research Center, Japan Atomic Energy Agency, Tokai 319-1195, Japan (Dated: February 7, 2013) We present a theory of the spin Hall magnetoresistance (SMR) in mu ltilayers made from an insulating ferromagnet F, such as yttrium iron garnet (YIG), an d a normal metal N with spin-orbit interactions, such as platinum (Pt). The SMR is induced by the simultaneous action of spin Hall and inverse spin Hall effects and therefore a non- equilibrium proximity phenomenon. We compute the SMR in F |N and F|N|F layered systems, treating N by spin- diffusion theory with quantum mechanical boundary conditions at th e interfaces in terms of the spin-mixing conductance. Our results explain the experiment ally observed spin Hall magnetoresistance in N |F bilayers. For F |N|F spin valves we predict an enhanced SMR amplitude when magnetizations are collinear. The SMR and the spin-tr ansfer torques in these trilayers can be controlled by the magnetic configuration. PACS numbers: 85.75.-d, 73.43.Qt, 72.15.Gd, 72.25.Mk I. INTRODUCTION Spin currents are a central theme in spintronics since they a re intimately associated with the manipulation and transport of spins in small structures and devices.1,2Spin currents can be gener- ated by means of the spin Hall effect (SHE) and detected by the in verse spin Hall effect (ISHE).3 Of special interest are multilayers made of normal metals (N ) and ferromagnets (F). When an electric current flows through N, an SHE spin current flows tow ards the interfaces, where it can be absorbed as a spin-transfer torque (STT) on the ferromagnet . This STT affects the magnetization damping4or even switches the magnetization.5,6The ISHE can be used to detect spin currents pumped by the magnetization dynamics excited by microwaves7–10or temperature gradients (spin Seebeck effect).11,12 Recently, magnetic insulators have attracted the attentio n of the spintronics community. Yt- trium iron garnets (YIG), a class of ferrimagnetic insulato rs with a large band gap, are interesting because of their very low magnetization damping. Their magn etization can be activated thermally to generate the spin Seebeck effect in YIG |Pt bilayers.13,14By means of the SHE, spin waves can be electrically excited in YIG via a Pt contact, and, via the ISH E, subsequently detected electrically2 in another Pt contact.15Spin transport at an N |F interface is governed by the complex spin-mixing conductance G↑↓.16The prediction of a large real part of G↑↓for interfaces of YIG with simple metals by first principles calculations17has been confirmed by experiments.18 Magnetoresistance (MR) is the property of a material to chan ge the value of its electrical resistance under an external magnetic field. In normal metal s its origin is the Lorentz force.19 The dependence of the resistance on the angle between curren t and magnetization in metallic ferromagnets is called anisotropic magnetoresistance (AM R). The transverse component of the AMR is also called the planar Hall effect (PHE), i.e.the transverse (Hall) voltage found in ferromagnets when the magnetization is rotated in the plane of the film.20,21Both effects are symmetric with respect to magnetization reversal, which di stinguishes them from the anomalous Hall effect (AHE) for magnetizations normal to the film, which c hanges sign under magnetization reversal.22The physical origin of AMR, PHE, and AHE is the spin-orbit int eraction, in contrast to the giant magnetoresistance (GMR), which reflects the cha nge in resistance that accompanies the magnetic field-induced magnetic configuration in magnet ic multilayers.23 Here we propose a theory for a recently discovered magnetore sistance effect in Pt |YIG bilayer systems.14,24,25This MR is remarkable since YIG is a very good electric insula tor such that a charge current can only flow in Pt. We explain this unusual mag netoresistance not in terms of an equilibriumstatic magnetic proximity polarization in Pt,24but rather in terms of anon-equilibrium proximity effect caused by the simultaneous action of the SHE a nd ISHE and therefore call it spin Hall magnetoresistance (SMR). This effect scales like the squ are of the spin Hall angle and is modulated by the magnetization direction in YIG via the spin -transfer at the N |F interface. Our explanationissimilartotheHanleeffect-inducedmagnetore sistanceinthetwo-dimensional electron gas proposed by Dyakonov.26Here we present the details of our theory, which is based on th e spin- diffusion approximation in the N layer in the presence of spin- orbit interactions27and quantum mechanical boundary conditions at the interface in terms of the spin-mixing conductance.16,17We also address F|N|F spin valves with electric currents applied parallel to the interface(s) with the additional degree of freedom of the relative angle between t he two magnetizations directions. This paper is organized as follows. We present the model, i.e.spin-diffusion with proper boundary conditions in Sec. II. In Sec. III, we consider an N |F bilayer as shown in Fig. 1 (a). We obtain spinaccumulation, spincurrentsandfinallythemeas uredchargecurrentsthat arecompared with the experimental SMR. We also find and discuss that the im aginary part of the spin-mixing conductance generates an AHE. F |N|F (Fig. 1 (b)) spin valves are investigated in Sec. IV, which show an enhanced SMR for spacers thinner than the spin-flip di ffusion length. We summarize the results and give conclusions in Sec. V. II. TRANSPORT THEORY IN METALS IN CONTACT WITH A MAGNETIC INSULATOR The spin current density in the non-relativistic limit ← →js=en/an}b∇acketle{t/vector v⊗/vector σ+/vector σ⊗/vector v/an}b∇acket∇i}ht/2 =/parenleftig /vectorjsx,/vectorjsy,/vectorjsz/parenrightigT =/parenleftig /vectorjx s,/vectorjy s,/vectorjz s/parenrightig (1) is a second-order tensor (in units of the charge current dens ity/vectorjc=en/an}b∇acketle{t/vector v/an}b∇acket∇i}ht), where e=|e|is the electron charge, nis the density of the electrons, /vector vis the velocity operator, /vector σis the vector of Pauli spin matrices, and /an}b∇acketle{t···/an}b∇acket∇i}htdenotes an expectation value. The row vectors /vectorjsi=en/an}b∇acketle{t/vector vσi+σi/vector v/an}b∇acket∇i}ht/2 in Eq. (1) are the spin current densities polarized in the ˆ ı-direction, while the column vectors /vectorjj s=en/an}b∇acketle{tvj/vector σ+/vector σvj/an}b∇acket∇i}ht/2 denote the spin current densities with polarization /vector σflowing in the ˆ - direction. Ohm’s law for metals with spin-orbit interactio ns can be summarized by the relation3 FIG. 1: (a) N|F bilayer and (b) F |N|F trilayer systems considered here, where F is a ferromagnetic insu lator and N a normal metal. between thermodynamic driving forces and currents that refl ects Onsager’s reciprocity by the symmetry of the response matrix:27 /vectorjc /vectorjsx /vectorjsy /vectorjsz =σ 1θSHˆx×θSHˆy×θSHˆz× θSHˆx×1 0 0 θSHˆy×0 1 0 θSHˆz×0 0 1 −/vector∇µ0/e −/vector∇µsx/(2e) −/vector∇µsy/(2e) −/vector∇µsz/(2e) , (2) where/vector µs= (µsx,µsy,µsz)T−µ0ˆ1 is the spin accumulation, i.e.the spin-dependent chemical potential relative to the charge chemical potential µ0,σis the electric conductivity, θSHis the spin Hall angle, and “ ×” denotes the vector cross product operating on the gradient s of the spin- dependent chemical potentials. The spin Hall effect is repres ented by the lower non-diagonal elements that generate the spin currents in the presence of a n applied electric field, in the following chosen to be in the ˆ x-direction /vectorE=Exˆx=−ˆx∂xµ0/e. The inverse spin Hall effect is governed by elements above the diagonal that connect the gradients of th e spin accumulations to the charge current density. The spin accumulation /vector µsis obtained from the spin-diffusion equation in the normal met al ∇2/vector µs=/vector µs λ2, (3) where the spin-diffusion length λ=√Dτsfis expressed in terms of the charge diffusion constant Dand spin-flip relaxation time τsf.28For films with thickness dNin the ˆz-direction /vector µs(z) =/vectorAe−z/λ+/vectorBez/λ, (4) where the constant column vectors /vectorAand/vectorBare determined by the boundary conditions at the interfaces. AccordingtoEq. (2), thespincurrentinNconsists of diffusio nandspinHall driftcontributions. Since we are considering a system homogeneous in the x-yplane, we focus on the spin current density flowing in the ˆ z-direction /vectorjz s(z) =−σ 2e∂z/vector µs−jSH s0ˆy, (5)4 wherejSH s0=θSHσExis the bare spin Hall current, i.e., the spin current generated directly by the SHE. The boundary conditions require that /vectorjz s(z) is continuous at the interfaces z=dNandz= 0. The spin current at a vacuum (V) interface vanishes, /vectorj(V) s= 0. The spin current density /vectorj(F) sat a magnetic interface is governed by the spin accumulation and spin-mixing conductance:16 e/vectorj(F) s(ˆm) =Grˆm×(ˆm×/vector µs)+Gi(ˆm×/vector µs), (6) where ˆm= (mx,my,mz)Tis a unit vector along themagnetization and G↑↓=Gr+iGithe complex spin-mixing interface conductance per unit area. The imagi nary part Gican be interpreted as an effectiveexchangefieldactingonthespinaccumulation. Apos itivecurrentinEq.(6)correspondsto up-spins flowing from F towards N. Since F is an insulator, thi s spin current density is proportional to the spin-transfer acting on the ferromagnet /vector τstt=−/planckover2pi1 2eˆm×/parenleftig ˆm×/vectorj(F) s/parenrightig =/planckover2pi1 2e/vectorj(F) s (7) With these boundary conditions we determine the coefficients /vectorAand/vectorB, which leads to the spin accumulation /vector µs=2eλ σ/bracketleftbigg −/parenleftig jSH s0ˆy+/vectorjz s(dN)/parenrightig coshz λ+/parenleftig jSH s0ˆy+/vectorj(F) s(ˆm)/parenrightig coshz−dN λ/bracketrightbigg /sinhdN λ,(8) where/vectorjz s(dN) = 0 for F(ˆ m)|N|V bilayers and /vectorjz s(dN) =−/vectorj(F) s(ˆm′) for F(ˆm)|N|F(ˆm′) spin valves. III. N|F BILAYERS In the bilayer the spin accumulation (8) is /vector µs(z) =−ˆyµ0 ssinh2z−dN 2λ sinhdN 2λ+/vectorj(F) s(ˆm)2eλ σcoshz−dN λ sinhdN λ, (9) whereµ0 s≡|/vector µs(0)|= (2eλ/σ)jSH s0tanh[dN/(2λ)] is the spin accumulation at the interface in the absence of spin-transfer, i.e., whenG↑↓= 0. Using Eq. (6), the spin accumulation at z= 0 becomes /vector µs(0) = ˆyµ0 s+2λ σ{Gr[ˆm(ˆm·/vector µs(0))−/vector µs(0)]+Giˆm×/vector µs(0)}cothdN λ. (10) With ˆm·/vector µs(0) =myµ0 s, (11) ˆm×/vector µs(0) =µ0 sσˆm׈y+ ˆmmy2λGicothdN λ σ+2λGrcothdN λ−/vector µs(0)2λGicothdN λ σ+2λGrcothdN λ, (12) /vector µs(0) = ˆyµ0 s1+2λ σGrcothdN λ/parenleftig 1+2λ σGrcothdN λ/parenrightig2 +/parenleftig 2λ σGicothdN λ/parenrightig2 + ˆmmyµ0 s2λ σGrcothdN λ/parenleftig 1+2λ σGrcothdN λ/parenrightig +/parenleftig 2λ σGicothdN λ/parenrightig2 /parenleftig 1+2λ σGrcothdN λ/parenrightig2 +/parenleftig 2λ σGicothdN λ/parenrightig2 +(ˆm׈y)µ0 s2λ σGicothdN λ/parenleftig 1+2λ σGrcothdN λ/parenrightig2 +/parenleftig 2λ σGicothdN λ/parenrightig2, (13)5 the spin current through the F |N interface then reads /vectorj(F) s=µ0 s eˆm×(ˆm׈y)σReG↑↓ σ+2λG↑↓cothdN λ+µ0 s e(ˆm׈y)σImG↑↓ σ+2λG↑↓cothdN λ.(14) The spin accumulation /vector µs(z) µ0s=−ˆysinh2z−dN 2λ sinhdN 2λ+[ˆm×(ˆm׈y)Re+(ˆm׈y)Im]2λG↑↓ σ+2λG↑↓cothdN λcoshz−dN λ sinhdN λ,(15) then leads to the distributed spin current in N /vectorjz s(z) jSH s0= ˆycosh2z−dN 2λ−coshdN 2λ coshdN 2λ−[ˆm×(ˆm׈y)Re+(ˆm׈y)Im]2λG↑↓tanhdN 2λ σ+2λG↑↓cothdN λsinhz−dN λ sinhdN λ. (16) The ISHE drives a charge current in the x-yplane by the diffusion spin current component flowing along the ˆ z-direction. The total longitudinal (along ˆ x) and transverse or Hall (along ˆ y) charge currents become jc,long(z) j0c= 1+θ2 SH/bracketleftigg cosh2z−dN 2λ coshdN 2λ+/parenleftbig 1−m2 y/parenrightbig Re2λG↑↓tanhdN 2λ σ+2λG↑↓cothdN λsinhz−dN λ sinhdN λ/bracketrightigg ,(17) jc,trans(z) j0c=θ2 SH(mxmyRe−mzIm)2λG↑↓tanhdN 2λ σ+2λG↑↓cothdN λsinhz−dN λ sinhdN λ, (18) wherej0 c=σExis the charge current driven by the external electric field. The charge current vector is the observable in the experimen t that is usually expressed in terms of the longitudinal and transverse (Hall) resistivities. A veraging the electric currents over the film thickness zand expanding the longitudinal resistivity governed by the current in the ( x-)direction of the applied field to leading order in θ2 SH, we obtain ρlong=σ−1 long=/parenleftbiggjc,long Ex/parenrightbigg−1 ≈ρ+∆ρ0+∆ρ1/parenleftbig 1−m2 y/parenrightbig , (19) ρtrans=−σtrans σ2 long≈−jc,trans/Ex σ2= ∆ρ1mxmy+∆ρ2mz, (20) where ∆ρ0 ρ=−θ2 SH2λ dNtanhdN 2λ, (21) ∆ρ1 ρ=θ2 SHλ dNRe2λG↑↓tanh2dN 2λ σ+2λG↑↓cothdN λ, (22) ∆ρ2 ρ=−θ2 SHλ dNIm2λG↑↓tanh2dN 2λ σ+2λG↑↓cothdN λ, (23) whereρ=σ−1is the intrinsic electric resistivity of the bulk normal met al. ∆ρ0<0 seems to imply that the resistivity is reduced by the spin-orbit interacti on. However, this is an effect of the order ofθ2 SHthat becomes relevant only when dNis sufficiently small. The spin-orbit interaction also generatesspin-flipscatteringthatincreasestheresistan cetoleadingorderaccordingtoMatthiesen’s rule. We see that ∆ ρ1(caused mainly by Gr) contributes to the SMR, while ∆ ρ2(caused mainly byGi) contributes only when there is a magnetization component n ormal to the plane (AHE), as discussed below.6 /s48 /s53 /s49/s48/s45/s49/s48/s49 /s115/s120/s61 /s106 /s115/s120/s61/s48/s40/s97/s41/s118/s97/s99/s117/s117/s109 /s70/s109 /s124/s124 /s121/s32 /s40/s114/s101/s102/s108/s101/s99/s116/s105/s110/s103/s41 /s32/s32 /s122/s32/s40/s110/s109/s41/s115/s121/s47/s48 /s115 /s106 /s115/s121/s47/s106/s83/s72 /s115 /s48 /s48 /s53 /s49/s48/s45/s49/s48/s49 /s40/s98/s41 /s109 /s124/s124 /s40/s120/s43/s121 /s41/s32/s118/s97/s99/s117/s117/s109 /s70 /s32/s32 /s122/s32/s40/s110/s109/s41/s115/s121/s47/s48 /s115 /s106 /s115/s121/s47/s106/s83/s72 /s115 /s48/s106 /s115/s120/s47/s106/s83/s72 /s115 /s48/s115/s120/s47/s48 /s115 /s48 /s53 /s49/s48/s45/s49/s48/s49 /s40/s99/s41/s118/s97/s99/s117/s117/s109 /s70/s109 /s124/s124 /s120/s32 /s40/s97/s98/s115/s111/s114/s98/s105/s110/s103/s41 /s32/s32 /s122/s32/s40/s110/s109/s41/s115/s121/s47/s48 /s115 /s106 /s115/s121/s47/s106/s83/s72 /s115 /s48/s115/s120/s61 /s106 /s115/s120/s61/s48 FIG. 2: (Color online). Normalized µsx,µsy,jsx, andjsyas functions of zfor magnetizations (a) ˆ m= ˆy, (b) ˆm= (ˆx+ ˆy)/√ 2, and (c) ˆ m= ˆxfor a sample with dN= 12 nm. We adopt the transport parameters ρ= 8.6×10−7Ωm,λ= 1.5 nm, and Gr= 5×1014Ω−1m−2. For magnetizations ˆ m= ˆyand ˆm= ˆx, both µsxandjsxare 0. A. Limit of Gi= ImG↑↓≪ReG↑↓=Gr According to first principles calculations,17|Gi|is at least one order of magnitude smaller than Grfor YIG, so Gi= 0 appears to be a good first approximation. In this limit, we p lot normalized components of spin accumulation ( µsxandµsy) and spin current ( jsx=/vectorjz s·ˆxandjsy=/vectorjz s·ˆy) as functions of zfor different magnetizations in Fig 2. When the magnetization of F is along ˆy, the spin current at the N |F interface ( z= 0) vanishes just as for the vacuum interface. By rotating the magnetization from ˆ yto ˆx, the spin current at the N |F interface and the torque on the magnetization is activated, while the spin accumulatio n is dissipated correspondingly. We note that the x-components of both spin accumulation and spin current vani sh when the magnetization is along ˆxand ˆy, and reach a maximum value at (ˆ x+ ˆy)/√ 2. ForGi= 0 the observable transport properties reduce to ρlong≈ρ+∆ρ0+∆ρ1/parenleftbig 1−m2 y/parenrightbig , (24) ρtrans≈∆ρ1mxmy, (25)7 where ∆ρ0 ρ=−θ2 SH2λ dNtanhdN 2λ, (26) ∆ρ1 ρ=θ2 SHλ dN2λGrtanh2dN 2λ σ+2λGrcothdN λ. (27) Equations (24-25) fully explain the magnetization depende nce of SMR in Ref. 25, while Eq. (27) shows that an SMR exists only when the spin-mixing conductan ce does not vanish. Since results do not depend on the z-component of magnetization, the AHE vanishes in our model w henGi= 0. B.Gr≫σ/(2λ) Here we discuss the limit in which the spin current transvers e to ˆmis completely absorbed as an STT without reflection. This ideal situation is actually n ot so far from reality for the recently found large Grbetween YIG and noble metals.17,18The spin current at the interface is then /vectorj(F) s jSH s0Gr≫σ/(2λ)= ˆm×(ˆm׈y)tanhdN λtanhdN 2λ, (28) and the maximum magnetoresistance for the bilayer is ∆ρ1 ρ=θ2 SHλ dNtanhdN λtanh2dN 2λ. (29) In Sec. IIIE we test this limit with available parameters fro m experiments. C.λ/dN≫1 When the spin-diffusion length is much larger than the thickne ss of N /vector µs(z) µ0sλ/dN≫1= ˆm×(ˆm׈y)−ˆy2z−dN dN, whilespincurrentandmagnetoresistance vanish. We can int erpretthis as multiplescattering of the spincurrent at theinterfaces; the ISHEhas both positive an dnegative charge current contributions that cancel each other. D. Spin Hall AHE Recent measurements in YIG |Pt display a small AHE-like signal on top of the ordinary Hall effect,i.e. a transverse voltage when the magnetization is normal to th e film.30As mentioned above, an imaginary part of the spin-mixing conductance Gican cause a spin Hall AHE (SHAHE). The component of the spin accumulation µsx µsx(z) µ0s=2λ σcoshz−dN λ sinhdN λ[mxmyRe−mzIm]σG↑↓ σ+2λG↑↓cothdN λ(30)8 /s50 /s52 /s54 /s56 /s49/s48/s48/s50/s52/s54/s56/s49/s48 /s32/s32/s40/s49/s48/s45/s32/s52 /s41 /s40/s110/s109 /s41/s32 /s83/s72/s61/s48/s46/s48/s50 /s32 /s83/s72/s61/s48/s46/s48/s52 /s32 /s83/s72/s61/s48/s46/s48/s54 /s32 /s83/s72/s61/s48/s46/s48/s56 /s32/s69/s120/s112/s46/s83/s97/s109/s112/s108/s101/s32/s49 /s83/s97/s109/s112/s108/s101/s32/s50/s40/s97/s41 /s50 /s52 /s54 /s56 /s49/s48/s48/s50/s52/s54/s56/s49/s48/s40/s49/s48/s45/s32/s52 /s41 /s40/s110/s109 /s41 /s32/s32 /s40/s98/s41 /s50 /s52 /s54 /s56 /s49/s48/s48/s50/s52/s54/s56/s49/s48/s40/s49/s48/s45/s32/s52 /s41 /s40/s110/s109 /s41 /s32/s32 /s40/s99/s41 /s50 /s52 /s54 /s56 /s49/s48/s48/s50/s52/s54/s56/s49/s48 /s40/s100/s41 /s40/s49/s48/s45/s32/s52 /s41 /s40/s110/s109 /s41 /s32/s32 FIG.3: (Coloronline)Calculated∆ ρ1/ρasafunctionof λfordifferentspinHallangles θSHwith(a)Gr= 1× 1014Ω−1m−2, (b)Gr= 5×1014Ω−1m−2, (c)Gr= 10×1014Ω−1m−2, and (d) the ideal limit Gr≫σ/(2λ). The Pt layers are 12-nm-thick with resistivity 8 .6×10−7Ωm (Sample 1, solid curve) and 7-nm-thick with resistivity 4 .1×10−7Ωm (Sample 2, dashed curve). Experimental results are shown as h orizontal lines for comparison.25 contains a contribution that scales with mzand contributes a charge current in the transverse (ˆ y-) direction j(SHAHE) c,trans(z) j0c=−2λθ2 SHmzsinhz−dN λ sinhdN λImG↑↓tanhdN 2λ σ+2λG↑↓cothdN λ. (31) The transverse resistivity due to this current is ρ(SHAHE) trans≈−j(SHAHE) c,trans/Ex σ2=−∆ρ2mz, (32) where ∆ρ2 ρ≈2λ2θ2 SH dNσGitanh2dN 2λ/parenleftig σ+2λGrcothdN λ/parenrightig2 +/parenleftig 2λGicothdN λ/parenrightig2≈2λ2θ2 SH dNσGitanh2dN 2λ/parenleftig σ+2λGrcothdN λ/parenrightig2. E. Comparison with experiments There are controversies about the values of the material par ameters relevant for our theory, i.e. the spin-mixing conductance G↑↓of the N|F interface, as well as spin-flip diffusion length λand spin Hall angle θSHin the normal metal.9 Experimentally, Burrows et al.18found for an Au|YIG interface with G0=e2/h. Gexp r G0= 5.2×1018m−2;Gexp r= 2×1014Ω−1m−2. (33) On the theory side, the spin-mixing conductance from scatte ring theory for an insulator reads16 G↑↓ G0=NSh−/summationdisplay nr∗ n↑rn↓=NSh−/summationdisplay nei(δn↓−δn↑), (34) wherern↑(↓)=eiδn↑(↓)is the reflection coefficient of an electron in the quantum chan nelnon a unit area at the N |F interface with unit modulus and phase δn↑(↓)for the majority (minority) spin, andNShis the number of transport channels (per unit area) at the Fer mi energy, i.e.NShis the Sharvin conductance (for one spin). Therefore Gr G0≤2NSh;|Gi| G0≤NSh, (35) Jiaet al.17computed Eq. (34) for a Ag |YIG interface by first principles. Theaverage of different crystal interfaces G(0) r= 2.3×1014Ω−1m−2, (36) is quite close to the Sharvin conductance of silver ( NShG0≈4.5×1014Ω−1m−2). For comparison with experiment we have to include the Schep d rift correction:31 1 ˜Gr/G0=1 G(0) r/G0−1 2NSh, (37) which leads to ˜Gr≈3.1×1014Ω−1m−2. (38) One should note that the mixing conductance of the Pt |YIG interface can then be estimated to be ˜Gr≈1015Ω−1m−2since the Pt conduction electron density and Sharvin conduc tance are higher than those of noble metals. Using parameters ρ=σ−1= 8.6×10−7Ωm,dN= 12 nm, and λ= 1.5 nm,29we see that the absorbed transverse spin currents with Gr=˜GrandGr=Gmax robtained from above for a Ag |YIG interface are 44% and 70% of the value for a perfect spin sink Gr→∞, respectively. For a Pt |YIG interface this value should be even larger. In order to compare our results with the observed SMR, we have to fill in or fit the parameters. The values of the spin-diffusion length and the spin Hall angle differ widely.29In Fig. 3 we plot the SMR for three fixed values of Gr. We observe that the experiments can be explained by a sensib le set of transport parameters ( Gr,λ,θSH) that somewhat differ for the two representative samples reported in Ref. 25. Generally, the SMR increases with a larg er value of Grbut decreases when λ is getting longer. These features are in agreement with the d iscussion of the simple limits above. Sample 1 in Ref. 25 has a larger resistivity but a smaller SMR ( ratio), implying a smaller spin Hall angle and/or smaller spin-diffusion length. When we fix th e spin Hall angle θSH= 0.06 and the spin-mixing conductance Gr= 5×1014Ω−1m−2, the corresponding estimated spin-diffusion lengths of Samples 1 and 2 are λ1≈1.5nm and λ2≈3.5nm, respectively. Finally we discuss the AHE equivalent or SHAHE. From experim ents ∆ρ2/ρ≈1.5×10−5for ρ= 4.1×10−7Ωm and dN= 7 nm.30Choosing θSH= 0.05,λ= 1.5nm, and Gr= 5×1014Ω−1m−2, we would need a Gi= 6.2×1013Ω−1m−2to explain experiments, a number that is supported by first principle calculations.1710 IV. SPIN VALVES In this section we discuss F(ˆ m)|N|F(ˆm′) spin valves fabricated from magnetic insulators with magnetization directions ˆ mand ˆm′. The general angle dependence for independent rotations of ˆm and ˆm′is straightforward buttedious. We discussinthe following two representative configurations in which the two magnetizations are parallel and perpendicu lar to each other. We disregard in the following the effective field due to Gisuch that the parallel and antiparallel configurations ˆm=±ˆm′are equivalent. Moreover, we limit the discussion to the sim ple case of two identical F |N and N|F interfaces, i.e., the spin-mixing conductances at both interfaces are the sa me. A. Parallel Configuration ( ˆm·ˆm′=±1) When the magnetizations are aligned in parallel or antipara llel configuration, the boundary condition /vectorj(z) s(dN) =−/vectorj(F) sapplies. We proceed as in Sec. III to obtain the spin accumula tion /vector µs µ0s=−/bracketleftigg ˆy+ ˆm×(ˆm׈y)2λGrtanhdN 2λ σ+2λGrtanhdN 2λ/bracketrightigg sinh2z−dN 2λ sinhdN 2λ, (39) and the spin current /vectorjz s jSH s0= ˆy/parenleftigg cosh2z−dN 2λ coshdN 2λ−1/parenrightigg + ˆm×(ˆm׈y)2λGrtanhdN 2λ σ+2λGrtanhdN 2λcosh2z−dN 2λ coshdN 2λ. The spin currents at the bottom and top of N are absorbed as STT s and read /vectorjz s(0) jSH s0=/vectorjz s(dN) jSH s0= ˆm×(ˆm׈y)2λGrtanhdN 2λ σ+2λGrtanhdN 2λ, (40) leading to opposite STTs at the bottom ( /vector τ(B) stt) and top ( /vector τ(T) stt) ferromagnets /vector τ(B) stt=/planckover2pi1 2e/vectorj(z) s(0) =−/vector τ(T) stt (41) since/vectorj(F) s(ˆm) =/vectorjz s(0) =/vectorjz s(dN) =−/vectorj(F) s(ˆm′). The longitudinal and transverse (Hall) charge currents are jc,long j0c= 1+θ2 SH/bracketleftigg 1−/parenleftbig 1−m2 y/parenrightbig2λGrtanhdN 2λ σ+2λGrtanhdN 2λ/bracketrightigg cosh2z−dN 2λ coshdN 2λ, (42) jc,trans j0c=−θ2 SHmxmy2λGrtanhdN 2λ σ+2λGrtanhdN 2λcosh2z−dN 2λ coshdN 2λ. (43) and the longitudinal and transverse resistivities read ρlong=ρ+∆ρ0+∆ρ1/parenleftbig 1−m2 y/parenrightbig , (44) ρtrans= ∆ρ1mxmy, (45) where ∆ρ0 ρ=−θ2 SH2λ dNtanhdN 2λ, (46) ∆ρ1 ρ=θ2 SH dN4λ2Grtanh2dN 2λ σ+2λGrtanhdN 2λ. (47)11 /s50 /s52 /s54 /s56 /s49/s48/s48/s46/s48/s48/s46/s50/s48/s46/s52/s48/s46/s54/s48/s46/s56 /s78/s47/s70 /s32/s32/s32/s32/s50 /s83/s72 /s110/s109/s70/s47/s78/s47/s70 FIG. 4: (Color online) Calculated ∆ ρ1//parenleftbig ρθ2 SH/parenrightbig in an F|N|F spin valve as a function of spin-diffusion length λwithdN= 12 nm, Gr= 5×1014Ω−1m−2, andρ= 8.6×10−7Ωm chosen from Sample 1 in Ref. 25. ∆ρ1//parenleftbig ρθ2 SH/parenrightbig in an N|F bilayer is plotted as a dotted line for comparison. Figure 4 shows ∆ ρ1//parenleftbig ρθ2 SH/parenrightbig with respect to the spin-diffusion length in an F |N|F spin valve with parallel magnetization configuration. Compared to N |F bilayers, the SMR in spin valves is larger and does not vanish in the limit of long spin-diffusion lengths . B. Limit λ/dN≫1 The spin accumulation for weak spin-flip reads /vector µs µ0sλ/dN≫1=−/bracketleftbigg ˆy+dNGr σ+dNGrˆm×(ˆm׈y)/bracketrightbigg2z−dN dN, (48) leading to the spin current /vectorjz s jSH s0λ/dN≫1=dNGr σ+dNGrˆm×(ˆm׈y). (49) In contrast to the bilayer, we find a finite SMR in this limit for spin valves: jc,long j0cλ/dN≫1= 1+ θ2 SH/bracketleftbigg 1−dNGr σ+dNGr/parenleftbig 1−m2 y/parenrightbig/bracketrightbigg Gr≫σ/dN= 1+ θ2 SHm2 y, (50) jc,trans j0cλ/dN≫1=−θ2 SHdNGr σ+dNGrmxmyGr≫σ/dN=−θ2 SHmxmy (51) or ∆ρ0 ρ=−θ2 SH, (52) ∆ρ1 ρ=θ2 SHdNGr σ+dNGrGr≫σ/dN=θ2 SH. (53) Here we find the maximum achievable SMR effects in metals with sp in Hall angle θSHby taking the limit of perfect spin current absorption. Clearly this r equires spin valves with sufficiently thin spacer layers. We interpret these results in terms of spin an gular momentum conservation: The finite SMR is achieved by using the ferromagnet as a spin sink t hat suppresses the back flow of spins and the ISHE. This process requires a source of angular momentum, which in bilayers can only be the lattice of the normal metal. Consequently, the SM R is suppressed in the F |N system when spin-flip is not allowed. In spin valves, however, the se cond ferromagnet layer can act as a spin current source, thereby allowing a finite SMR even in the absence of spin-flip scattering.12 C. Perpendicular Configuration ( ˆm·ˆm′= 0) We may consider two in-plane magnetizations ˆ m= (cosα,sinα,0) and ˆm′= (−sinα,cosα,0), which are perpendicular to each other. When α= 0, the first layer maximally absorbs the SHE spin current, while ˆ m′is completely reflecting, just as the vacuum interface in the bilayer. For generalα: µsx(z) µ0s=2λGr σ+2λGrcothdN λ/parenleftigg coshz−dN λ sinhdN λ+coshz λ sinhdN λ/parenrightigg cosαsinα, (54) µsy(z) µ0s=−sinh2z−dN 2λ sinhdN 2λ−2λGr σ+2λGrcothdN λ/parenleftigg coshz−dN λ sinhdN λcos2α−coshz λ sinhdN λsin2α/parenrightigg ,(55) µsz(z) = 0, (56) which leads to the components of spin current normal to the in terfaces jsx(z) jSH s0=−2λGrtanhdN 2λ σ+2λGrcothdN λ/parenleftigg sinhz−dN λ sinhdN λ+sinhz λ sinhdN λ/parenrightigg cosαsinα, (57) jsy(z) jSH s0=cosh2z−dN 2λ−coshdN 2λ coshdN 2λ+2λGrtanhdN 2λ σ+2λGrcothdN λ/parenleftigg sinhz−dN λ sinhdN λcos2α−sinhz λ sinhdN λsin2α/parenrightigg .(58) The total current is the sum of those from the two ferromagnet s at the top and bottom; in contrast to the parallel ˆ m=±ˆm′configuration, they do not feel each other. We can extend the d iscussion from the previous subsection: the second F can be a spin curre nt source, and we can switch this source on by rotating the magnetization from perpendicular to (anti)parallel configuration. The longitudinal and transverse electric currents read jc,long(z) j0c= 1+θ2 SHcosh2z−dN 2λ coshdN 2λ+θ2 SH2λGrtanhdN 2λ σ+2λGrcothdN λ/parenleftigg sinhz−dN λ sinhdN λcos2α−sinhz λ sinhdN λsin2α/parenrightigg , (59) jc,trans(z) j0c=θ2 SH2λGrtanhdN 2λ σ+2λGrcothdN λ/parenleftigg sinhz−dN λ sinhdN λ+sinhz λ sinhdN λ/parenrightigg cosαsinα. (60) Since the angle-dependent contributions vanish upon integ ration over z, there is no magnetoresis- tance in the perpendicular configuration. D. Controlling the spin-transfer torque Like the SMR, the STT at the N |F interface depends on the relative orientation between ˆ m and ˆm′, too. We may pin ˆ m= ˆxand observe how the STT at the bottom magnet, /vector τ(B) stt(ˆm,ˆm′), changes with rotating ˆ m′= ˆxcosα+ ˆysinα. Figure 5 displays the ratio βdefined as β(α)≡/vextendsingle/vextendsingle/vextendsingle/vector τ(B) stt(ˆx,ˆx)−/vector τ(B) stt(ˆx,ˆxcosα+ ˆysinα)/vextendsingle/vextendsingle/vextendsingle /vextendsingle/vextendsingle/vextendsingle/vector τ(B) stt(ˆx,ˆx)/vextendsingle/vextendsingle/vextendsingle, (61) as a function of αfor some spin-diffusion lengths. Only when λ≪dN,βremains constant under rotation of ˆ m′. A larger spin-mixing conductance and smaller dNenhances the SMR as well as angle dependence of β. This modification of the STT should lead to complex dynamics of the spin valve in the presence of an applied current and will be the sub ject of a subsequent study.13 /s48/s46/s48 /s48/s46/s53/s48/s49/s50 /s32/s32 /s32 /s61/s50 /s110/s109 /s32 /s61/s52 /s110/s109 /s32 /s61/s54 /s110/s109 /s32 /s61/s56 /s110/s109 /s32 /s61/s49/s48 /s110/s109 FIG. 5: (Color online) The ratio β(α) which characterize how /vector τ(B) sttchanges with respect to the relative orientation between ˆ mand ˆm′. We adopt the transport parameters dN= 12 nm, ρ= 8.6×10−7Ωm, and Gr= 5×1014Ω−1m−2. V. SUMMARY We developed a theory for the SMR in N |F and F|N|F systems that takes into account the spin-orbit coupling in N as well as the spin-transfer at the N |F interface(s). In a N |F bilayer system, the SMR requires spin-flip in N and spin-transfer at t he N|F interface. Our results explain the SMR measured in Ref. 25 both qualitatively and quantitat ively with transport parameters that are consistent with other experiments. The degrees of s pin accumulation in N that can be controlled by the magnetization direction is found to be v ery significant. In the presence of an imaginary part of the spin-mixing conductance Giwe predicted a AHE-like signal (SHAHE). Such a signal was observed in Ref. 30 and can be explained with values of Githat agree with first principles calculations.17We furthermoreanalyzed F |N|F spin valves for parallel and perpendicular magnetization configurations. A maximal SMR ∼θ2 SHis found for a collinear magnetization configuration in the limit that the spin-diffusion length is mu ch larger than the thickness of the normal spacer. TheSMR vanishes when rotating the two magnet izations into a fixed perpendicular constellation. The SMR torques under applied currents in N a re expected to lead to magnetization dynamics of N|F and F|N|F structures. Acknowledgments This work was supported by FOM (Stichting voor Fundamenteel Onderzoek der Materie), EU- ICT-7 “MACALO,” the ICC-IMR, DFG Priority Programme 1538 “S pin-Caloric Transport” (GO 944/4), and KAKENHI (Grant-in-Aid for Scientific Research) C (22540346). 1S. D. Bader and S. S. P. Parkin, Ann. Rev. Cond. Matt. Phys. 1, 71 (2010). 2J. Sinova and I ˇZuti´ c, Nature Mater. 11, 368 (2012). 3For a review see: T. Jungwirth, J. Wunderlich, and K. Olejn´ ık, Natu re Mater. 11, 382 (2012). 4K. Ando, S. Takahashi, K. Harii, K. Sasage, J. Ieda, S. Maekawa, a nd E. Saitoh, Phys. Rev. Lett. 101, 036601 (2008). 5I. M. Miron, K. Garello, G. Gaudin, P.-J. Zermatten, M. V. Costache , S. Auffret, S. Bandiera, B. Rodmacq, A. Schuhl, and P. Gambardella, Nature 476, 189 (2011). 6L. Liu, C. F. Pai, Y. Li, H. W. Tseng, D. C. Ralph, and R. A. Buhrman, S cience336, 555 (2012).14 7E. Saitoh, M. Ueda, H. Miyajima, and G. Tatara, Appl. Phys. Lett. 88, 182509 (2006). 8O. Mosendz, J. E. Pearson, F. Y. Fradin, G. E. W. Bauer, S. D. Bad er, and A. Hoffmann, Phys. Rev. Lett.104, 046601 (2010). 9O. Mosendz, V. Vlaminck, J. E. Pearson, F. Y. Fradin, G. E. W. Baue r, S. D. Bader, and A. Hoffmann, Phys. Rev. B 82, 214403 (2010). 10F. D. Czeschka, L. Dreher, M. S. Brandt, M. Weiler, M. Althammer, I.-M. Imort, G. Reiss, A. Thomas, W. Schoch, W. Limmer, H. Huebl, R. Gross, and S. T. B. Goennenwein , Phys. Lett. Rev. 107, 046601 (2011). 11K. Uchida, S. Takahashi, K. Harii, J. Ieda, W. Koshibae, K. Ando, S. Maekawa, and E. Saitoh, Nature 455, 778 (2008). 12C. M. Jaworski, J. Yang, S. Mack, D. D. Awschalom, J. P. Heremans , and R. C. Myers, Nature Mater. 9, 898 (2010). 13K. Uchida, J. Xiao, H. Adachi, J. Ohe, S. Takahashi, J. Ieda, T. Ota , Y. Kajiwara, H. Umezawa, H. Kawai, G. E.W. Bauer, S. Maekawa, and E. Saitoh, Nature Mater. 9, 894 (2010). 14M. Weiler, M. Althammer, F. D. Czeschka, H. Huebl, M. S. Wagner, M. Opel, I.-M. Imort, G. Reiss, A. Thomas, R. Gross, and S. T. B. Goennenwein, Phys. Rev. Lett. 108, 106602 (2012). 15Y. Kajiwara,K. Harii, S. Takahashi, J. Ohe, K. Uchida, M. Mizuguchi, H. Umezawa, H. Kawai, K. Ando, K. Takanashi, S. Maekawa, and E. Saitoh, Nature 464, 262 (2010). 16A. Brataas, Yu. V. Nazarov, and G. E. W. Bauer, Phys. Rev. Lett .84, 2481 (2000); Eur. Phys. J. B 22, 99 (2001). 17X. Jia, K. Liu, K. Xia, and G. E. W. Bauer, Eurphys. Lett. 96, 17005 (2011). 18C. Burrowes, B. Heinrich, B. Kardasz, E. A. Montoya, E. Girt, Y. S un, Y. Y. Song, and M. Wu, Appl. Phys. Lett. 100, 092403 (2012). 19N. W. Ashcroft and N. D. Mermin, Solid State Physics (Saunders, Ph iladelphia, 1976). 20T. R. McGuire and R. I. Potter, IEEE Trans. Magn. MAG-11 , 1018 (1975). 21D. A. Thompson, L. T. Romankiw, and A. F. Mayadas, IEEE Trans. M agn.MAG-11 , 1039 (1975). 22N. Nagaosa, J. Sinova, S. Onoda, A. H. MacDonald, and N. P. Ong, R ev. Mod. Phys. 82, 1539 (2010). 23A. Fert, Rev. Mod. Phys. 80, 1517 (2008). 24S. Y. Huang, X. Fan, D. Qu, Y. P. Chen, W. G. Wang, J. Wu, T. Y. Che n, J. Q. Xiao, and C. L. Chien, Phys. Rev. Lett. 109, 107204 (2012). 25H. Nakayama, M. Althammer, Y.-T. Chen, K. Uchida, Y. Kajiwara, D. Kikuchi, T. Ohtani, S. Gepr¨ ags, M. Opel, S. Takahashi, R. Gross, G. E. W. Bauer, S. T. B. Goennenw ein, and E. Saitoh, arXiv:1211.0098 (cond-mat.mes-hall). 26M. I. Dyakonov, Phys. Rev. Lett. 99, 126601 (2007). 27S. Takahashi, H. Imamura, and S. Maekawa, in Concepts in Spin Electronics , edited by S. Maekawa (Oxford University Press, U.K., 2006), pp. 343-370. 28T. Valet and A. Fert, Phys. Rev. B 48, 7099 (1993). 29L. Liu, R. A. Buhrman, and D. C. Ralph, arXiv:1111.3702 (cond-mat.m es-hall). 30M. Althammer et al., unpublished. 31K. M. Schep, J. B. A. N. van Hoof, P. J. Kelly, G. E. W. Bauer, and J. E. Inglesfield, Phys. Rev. B 56, 10805 (1997). | 2013-02-06 | We present a theory of the spin Hall magnetoresistance (SMR) in multilayers
made from an insulating ferromagnet F, such as yttrium iron garnet (YIG), and a
normal metal N with spin-orbit interactions, such as platinum (Pt). The SMR is
induced by the simultaneous action of spin Hall and inverse spin Hall effects
and therefore a non-equilibrium proximity phenomenon. We compute the SMR in
F$|$N and F$|$N$|$F layered systems, treating N by spin-diffusion theory with
quantum mechanical boundary conditions at the interfaces in terms of the
spin-mixing conductance. Our results explain the experimentally observed spin
Hall magnetoresistance in N$|$F bilayers. For F$|$N$|$F spin valves we predict
an enhanced SMR amplitude when magnetizations are collinear. The SMR and the
spin-transfer torques in these trilayers can be controlled by the magnetic
configuration. | Theory of spin Hall magnetoresistance | 1302.1352v1 |
1 Submitted Oct 11, 2012 Intrinsic Spin Seebeck Effect in Au/YIG D. Qu1, S. Y. Huang1,2, Jun Hu3, Ruqian Wu3, and C. L. Chien1* Affiliations: 1Department of Physics and Astronomy, Johns Hopkins University, Baltimore Maryland 21218, USA 2Department of Physics, National Tsing Hua University, Hsinchu 300, Taiwan 3 Department of Physics and Astronomy, University of California, Irvine , California 92697, USA *clc@pha.jhu.edu Abstract: The acute magnetic proximity effects in Pt/YIG compromise the suitability of Pt as a spin current detector. We show that Au/YIG, with no anomalous Hall effect and a negligible magnetoresistance, allows the measurements of the intrinsic spin Seebeck effect with a magnitude much smaller than that in Pt/Y IG. The experiment results are consistent with the spin -polarized density -functional calculations for Pt with a sizable and Au with a negligible magnetic moment near the interface with YIG. PACS numbers : 72.15.Jf, 72.20.Pa, 85.80. -b, 85.75. -d 2 The expl oration of spintronic phenomena has been advanced towards the manipulation of a pure spin current without a charge current. A pure spin current can be realized by compelling electrons of opposite spins to move in opposite directions, or be carried by spin waves (magnons). Pure spin current is beneficial for spintronic operations with the attributes of maximal angular momentum and minimal charge current thus with much reduced Joule heating, circuit capacitance and electromigration. In the spin Hall effect (S HE), a charge current driven by a voltage gradient can generate a transverse spin current [ 1]. Using the spin Seebeck effect (SSE), a temperature gradient can also generate a spin current. Consequently, the SSE, within the emerging field of “spin caloritronics”, where one exploits the interplay of spin, charge, and heat, has attracted much attention. SSE has been reported in a variety of ferromagnetic (FM) materials (metal [2], semiconductor [3], or insulator [4]), where the pure spin current is detected in the Pt strip patterned onto the FM material by the inverse spin Hall effect (ISHE) with an electric field of ESHE = DISHE jS, where DISHE is the ISHE efficiency, jS is the pure spin current density diffusing into the Pt strip and is the spin direction. Consider a FM layer sample in the xy-plane, there are two ways to observe SSE using either the transverse or the longitudinal SSE configuration with a temperature gradient applied either in the sample plane (∇xT) or out -of-plane (∇zT) respectively. Various potential applications of SSE have already been proposed [ 5]. However, the SSE is not without controversies and complications. One fundamental mystery is that SSE has been reported in macroscopic structures on the mm scale whereas the spin diffusion length within which the spin coh erence is preserved is only on the nm scale [2-4]. Furthermore, the previous reports of SSE have other 3 unforeseen complications. In the transverse geometry of SSE with an intended in- plane∇xT, due to the overwhelming heat conduction through the substrate, there exists also an out -of-plane ∇zT, which gives rise to the anomalous Nernst effect (ANE) with an electric field of EANE zT m, where m is the magnetization direction [ 6]. The ANE is very sensitive in detecting ∇zT in a manner similar to the high sensitivity of the anomalous Hall effect to perpendicular magnetization in FM layer s less than 1 nm in thickness. As a result, ESHE jS due t o SSE with jS in the z -direction and EANE zT m due to ANE are both along the y -direction and asymmetric in magnetic field. The voltages of SSE due to ∇xT and ANE due to ∇zT are additive, entangled, and inseparable [6]. In the longitudinal SSE using Pt on a FM insulator (e.g., YIG), while the temperature gradient ∇zT is unequivocally out -of-plane, one encounters a different issue of magnetic proximity effects (MPE) in Pt in contact with a FM material. As a result , in the longitudinal configuration there is also entanglement of SSE and ANE [7]. These complications, when present, prevent the unequivocal establishment of SSE in either the transverse or the longitud inal configuration. The characteristics of the intrinsic SSE including its magnitude, remain outstanding and unresolved issues. In this work, we report the measurements of intrinsic SSE in gold (Au) using the longitudinal configuration with an unambiguo us out -of-plane (zT) gradient near room temperature . It is crucial to identify metals other than Pt that can unequivocally detect the pure spin current without MPE . Gold offers good prospects since it has been successfully used as a substrate or underlaye r for ultrathin magnetic films. We use polished 4 polycrystalline yttrium iron garnet (YIG=Y 3Fe5O12) a well -known FM insulator with low loss magnons as the substrate . A large spin -mixing conductance at Au/YIG interface has been reported using ferromagnetic r esonance [ 8]. Our results show that Au( t)/YIG does not have the large anomalous Hall effect and large MR that plagued Pt( t)/YIG but exhibits an unusual thickness dependence in the thermal transport. These resul ts allow us to place an upper limit for the intrinsic SSE of about 0.1 µ V/K much smaller than the thermal effect in Pt/YIG [ 7]. Thin Au films have been made by magnetron sputtering on YIG and pattern ed into parallel wires and Hall bars. As shown in the inset of Fig. 1(a), the xyz axes are parallel to the edges of the YIG substrates. The parallel wires with ascending order of thickness from 4 nm to 12 nm are in the xy-plane and oriented in the y-direction, where each wire is 4 mm long, 0.1 mm wide, and 2 mm apart. The Hall bar samples (inset of Fig. 1b) consist of one long segment along the x-direction and several short segments along the y-direction. For the MR measurements current is along the x-axis, for the thermal transport measurement ∇zT is along the z-axis, and the magnetic field is in the xy-plane in both cases. We use 4 -probe and 2 -probe measurements for MR and thermal voltage respectively. The multiple wires facilitate a systematic study of the thickness dependence of electric transport and thermal measurement under the same uniform thermal gradient with a temperature difference of Tz ≈ 10 K. The sample was sandwiched between, and in thermal contact with, two large Cu blocks kept at constant temperatures differing by 10 K. We first describe the thickness dependence of electrical resistivity ( ρ) of the Au wires. As expected ρ increases with decreasing film thickness as shown in Fig. 1(a). The 5 results can be well described by a semi classical theoretical model in the frame of Fuchs - Sondheimer (FS) theory [9], which includes the contributions from thickness ( t) as well as surface scattering ( p) and grain boundary scattering ( ), ρ = ρ∞{1-(1/2+3λ/4t)[1 -pexp(- tξ/λ)]exp( -t/λ)}-1 for t/ > 0.1 . Using bulk resistiv ity (ρ∞ = 2.2 cm) and the mean free path (λ=37 nm) [ 10], we find the data can be well described by p = 0.89 and = 0.37 as shown by the solid line in Fig. 1(a) . The anomalous Hall effect (AHE) is an essential measurement for assessing MPE . Hall measurements of the Au/YIG Hall bar samples have been made from 2K to 300K as shown in Fig. 1(b). The Hall resistance of Au/YIG is linea r in magnetic field at all temperatures (2 -300 K) showing only the ordinary Hall effect (OHE) with no observable AHE. In contrast, strong AHE has been observed in Pt/YIG due to the acute MPE [ 7]. The Hall const ant ( RH = 1/ne) of Au/YIG indicates the carrier concentration n ≈ 6×1022 cm-3 as shown in Fig. 1(c), essentially constant from 2 K to 300 K, is in good agreement with the bulk carrier concentration of n =5.9× 1022 cm-3 [11]. The spin polarized moment induced in Au, is very small, if any, i.e., Au is not appreciably affected by MPE and will be further discussed below. We employ the longitudinal configuration with spin current along the out -of-plane temperature gr adient zT to determine the thickness dependence of the thermal transport of Au/YIG, and to compare the results with those of the Pt/YIG. As shown in Fig. 2(a), the transverse thermal voltage (in the y-direction) across the Au strip is asymmetrical when the magnetic field is along the x-axis with the same sense as that for the Pt strip. The same sign of the thermal spin -Hall voltage between Pt and Au is consistent with the theoretical calculation of positive values of spin Hall conductivity in Pt and Au [ 12]. 6 However, there are several distinct differences between the thermal results of Au/YIG and Pt/YIG. We take Vth as the magnitude of spin -dependent thermal voltage between the positive and the negative switchin g fields. As shown in Fig. 2(b), the value of Vth of the Pt( t)/YIG is far larger, increasing sharply and unabatedly with decreasing t to a value of 64 µ V at t = 2.2 nm , due to the strong MPE at the interface between Pt and YIG. In contrast, the thermal v oltage Vth of the Au/YIG samples is much smaller than that of Pt/YIG and it varies with thickness ( t) in a non -monotonic manner as shown in Fig. 2(c). The value of Vth is negligible (less than 0.2 µ V) for t ≤ 7 nm , increasing to a maximum of 1.3 µ V at t = 8 nm before decreasing at larger thicknesses. This contrasting behavior shows that there is much smaller, perhaps negligible, MPE in Au/YIG. Consequently, the measured thermal voltage may be attributed enti rely to intrinsic SSE. With a maximal (Vth)max ≈1.3 µV at t = 8 nm at T of 10 K, the strength of the intrinsic SSE in Au/YIG is about 0.1 µ V/K, far smaller than the values in Pt( t)/YIG of 6 µ V/K at t = 2.2 nm, and 1 µ V/K at t = 10 nm, by one to two order s of magnitude. This suggests most of the spin - dependent thermal voltage in Pt/YIG is due to ANE and not SSE. From the value of Sxy ≈ 610-3 μV/K (Sxy=E xy/∇T=(Vth/l)/(T/d), where Vth is the thermal voltage, l is the distance between the voltage leads, T is the temperature difference and d is the thickness of Au/YIG sample) we measured and using the Seebeck coefficient Sxx ≈1.9 μ V/K of Au at 300 K [ 13], we obtain a spin Nernst angle of N = Sxy/Sxx ≈ 0.003, wh ich is very close to the spin Hall angle H = 0.0016, defined as the ratio of spin Hall and charge conductivities, from spin pumping measurement [ 14]. 7 However, we have observed MR, albeit with very small but c lear signals, in Au(t)/YIG. The MR result of Au(7 nm)/YIG Hall bar sample is shown in Fig. 3 (a). It is of a very small magnitude of ≈ - 4 x 10-6, where = || - T, about two orders of magnitude smaller than those of Pt( t)/YIG as shown in Fig. 3(b). Nevertheless all the Au(t)/YIG with 4 nm ≤ t ≤ 11 nm show similarly small but measurable More unexpectedly, the MR of Au(t)/YIG has the opposite angular dependence as that of the usual anisotropic MR (AMR). In the AMR of most 3d ferromagnetic met als of Fe, Co, Ni, and their alloys, the common behavior is positive , that is || > T, the resistivity with current parallel to, is higher than that with the current perpendicular, to the magnetization aligned by a magnetic field. The MR observed in P t(t)/YIG also has the same behavior of > 0. In contrast, the small MR in Au( t)/YIG is opposite with T > ||, or inverse AMR, The mechanism of this up behavior in Au( t)/YIG is not yet fully understood, but probably due to spin -dependent scattering at i nterface between Au and YIG, supported by the fact that | | increases with decreasing Au films thickness. One notes that inverse AMR has occasionally been reported in thin Co films. The s -d scattering influenced by spin -orbital and electron -electron interactions may be enhanced by the disorder in thin films [ 15]. To assess the magnetic moments of Pt and Au near the interface with YIG, spin - polarized density functional calculations have been carried out with th e Vienna ab initio simulation package (VASP), [16,17] at the level of the generalized gradient approximation (GGA) [18] with a Hubbard U correction for Fe -3d orbitals in YIG. We use the projector augmented wave (PAW) method for the description of the core -valence interaction [19,20]. The YIG structure has two Fe sites: tetrahedral Fe t and octahedral Fe o. 8 To model the Pt/YIG and Au/YIG interfaces, we construct a superlattice structure with a slab of YIG(111) of about 6 Å thick along with a 4 -layer Pt or Au film of about 7 Å thick. In the initial configuration, the Fe o atoms match the hcp sites of Pt(111) or Au(111) slab. During the relaxation process, the in -plane lattice constant has been fixed at the experimental value of the bulk YIG, with a dimension of 17.5× 17.5 Å2, and thickness of superlattice [notated as c in Fig. 4 (a)] is allowed to change. All atoms are fully relaxed until the calculated force on each atom is smaller than 0.02 eV/Å. For this lar ge unit cell with 274 atoms, we find that a single Γ point is enough to sample the Bril louin zone. The optimized atom ic structure of Pt/YIG in Fig. 4(a) shows significant reconstructions in both Pt and YIG layers. The average bond lengths are: d Pt-O ~ 2.2 Å, d Pt-Fe ~ 2.6 Å. Au/YIG has a similar structure. It is important to note that all fo ur Pt layers are significantly spin polarized as shown in Fig. 4(b) . The Pt layers adjacent to the interfaces [labeled by 1 and 4 in Fig. 4(b)] tend to cou ple ferromagnetically to their neighboring Fe atoms in YIG, as found in most studies for Pt on magnet ic substrates. The local spin moments of Pt atoms in the Pt 2 and Pt 3 layers can still be as large as 0.1 μB. By integrating spin density in the lateral planes, we can obtain the z -dependent spin density as shown in Fig. 4(c). Clearly, the spin polarization in all Pt layers is significant for the measurement of SSE. In particular, the total spin moments of th e Pt 2 and Pt 3 layers (each has 36 Pt atoms) are about 0.8 μ B and 1.1 μ B, respectively, even after the mutual cancelation with the intra -layer antiferromagnetic ordering. In contrast, spin polarizations induced in the Au layers are much weaker , with the max imum local spin moment smaller than 0.05 μB and the integrated spin moment in the entire Au layers smaller than 0. 1 μ B. Therefore, one can 9 view Au as nearly “nonmagnetic” in contact with YIG, in contrast to Pt. The sizable magnetic moments of Pt near the interface from the theoretical calculations is consisting with the strong MPE shown in Pt( t)/YIG by the electric transport . Therefore, the ANE and SSE are not only entangled but with ANE dominating in Pt/YIG. In contrast, the negligible Au moments from the oretical calculations is also consistent with no apparent AHE in Au( t)/YIG . The only noticeable magnetic characteristic is the inverse AMR of Au( t)/YIG but with a magnitude two orders smaller than that of Pt/YIG. This indicates that most, if not all, of the thermal voltage measured in Au/YIG is due to the intrinsic SSE as a result of the pure spin current injected from YIG. As shown in Fig. 2(c), the measurement of the thickness dependence is essential in revealing the non -monotonic dependence of intrin sic SSE voltage in Au/YIG due to the spin diffusion length SF. For very thin Au layer with t < 6 nm, SF is short due to the large resistivity from interface and boundary scattering, thus no appreciable spin current could survive intact , and this results in negligible Vth. As the Au film thickness increases, t he value of Vth exhibits a rapid rise reaching a maximum of 1.3V at t~8 nm and then decreases owing to the spin flip relaxation mechanism. Using the expression xx sf F sf / h/e/ k =l ) ( )23()2/(2 2 including the Fermi wave vector kF, the conductivity xx, the mean time between collisions and the mean time between spin - flipping collision sf, we estimate SF ≈ 40 nm. [ 21] The critical thickness of 8 nm is close to spin diffusion length 10.5 nm evaluated from weak localization [ 10]. Given the weak inverse AMR and the nonexistent AHE, the thermal signal of 0.1 µ V/K measured in Au/YIG at an optimal thickness of 8 nm should be considered as an upper limit of the 10 intrinsic S SE effect. The spin Hall angle between Au and YIG might be further enhanced by chemical modification on the YIG surface at high temperature . But a careful surface treatment is very important to avoid the metallic state of Fe developed, which could result in a reduction of spin mixing conductance and contamination in SSE [ 22]. In summary, we use Au(t)/YIG with no anomalous Hall signals and a very weak inverse MR results with non -monotonic dependence of spin -thermal voltage to show that the acute magnetic proximity effects that plagued Pt/YIG do not affect Au/YIG . The thermal voltage in Au/YIG is thus due to primarily intrinsic spin Seebeck effect with an upper limit of 0.1 µ V/K. Although the spin Hall angle of Au is smaller th an that of Pt, Au is a good spin current detector, far better than Pt. Acknowledgments: The work is supported at Johns Hopkins University by US NSF (DMR 05 -20491) and Taiwan NSC (99 -2911 -I-007- 510), and at University of California by DOE -BES (Grant No: D E-FG02 -05ER46237) and by NERSC for computing time. References : 1. J. E. Hirsch, Phys. Rev. Lett. 83, 1834. (1999). 2. K. Uchida, S. Takahashi, K. Harii, J. Ieda, W. Koshibae, K. Ando, S. Maekawa, and E. Saitoh, Nature 455, 778 (2008). 3. C. M. Jaworski, J. Yang, S. Mack, D. D. Awschalom, J. P. Heremans, and R. C. Myers, Nat. Mater. 9, 898 (2010). 4. K. Uchida, J. Xiao, H. Adachi, J. Ohe, S. Takahashi, J. Ieda, T. Ota, Y. Kajiwara, H. Umezawa, H. Kawai, G. E. W. Bauer, S. Maekawa, and E. Saitoh, Nat. Mater. 9, 894 ( 2010). 5. A. Kirihara, K. Uchida, Y. Kajiwara, M. Ishida, Y. Nakamura 1, T. Manako 1, E. Saitoh and S. Yorozu, Nature Mater . 11, 686 (2012). 6. S. Y. Huang, W. G. Wang, S. F. Lee, J. Kwo, and C. L. Chien, Phys. Rev. Lett. 107, 216604 (2011). 7. S. Y. Huang, X. F an, D. Qu, Y. P. Chen, W. G. Wang, J. Wu, T. Y. Chen, J. Q. Xiao and C. L. Chien, Phys. Rev. Lett. 109, 107204 (2012). 11 8. B. Heinrich, C. Burrowes, E. Montoya, B. Kardasz, E. Girt, Young -Yeal Song, Yiyan Sun, and Mingzhong Wu, Phys. Rev. Lett. 107, 066604 (2011). 9. P. Fan, K. Ti. J. D. Shao, and Z. X. Fau, J. Appl. Phys. 95, 2527 (2004). 10. J. Bass and W. P. Pratt Jr., J. Phys.: Condes. Matter 19, 183201 (2007). 11. C. L. Chien and C. R. Westgate, The Hall effect and its applications. (Plenum Press, New York, 1980). 12. T. Tanaka, H. Kontani, M. Naito, T. Naito, D. S. Hirashima, K. Yamada, and J. Ionoue, Phys. Rev. B 77, 165117 (2008). 13. D. M. Rowe , CRC handbook of thermoelectrics (CRC Press, New York , 1995) 14. O. Mosendz , J. E. Pearson, F. Y. Fradin , G. E. W. Bauer, S. D. Bade r, and A. Hoffmann, Phys. Rev. Lett. 104, 046601 (2010). 15. T. Y. Chung and S. Y. Hsu, J Phys.: Conf. Ser. 150, 042063 (2009). 16. G. Kresse and J. Furthmuller, Comput. Mater. Sci. 6, 15 (1996). 17. G. Kresse and J. Furthmuller, Phys. Rev. B 54, 11169 (1996). 18. J. P. Perdew, K. Burke, and M. Ernzerhof, P hys. Rev. Lett. 77, 3865 (1996). 19. P. E. Blochl, Phys. Rev. B 50, 17953 (1994). 20. G. Kresse and D. Joubert, Phys. Rev. B 59, 1758 (1999). 21. M. Gradhand, D. V. Fedorov, P. Zahn, and I. Mertig, Phys. Rev. B. 81, 245109 (2010). 22. C. Burrowes, B. Heinrich, B. Kardasz, E. A. Montoya, E. Girt, Yiyan Sun, Young -Yeal Song, and Mingzhong Wu, Appl. Phys. Lett. 100, 092403 (2012). 12 Fig. 1 (color online). (a) Resistivity as a function of Au thickness t for Au /YIG. The solid line represent s semiclassical theoretic fittings . Inset is schematic diagram of multiple patterned strips with ascending thickness. (b) Field dependence of Hall resistance RH at different temperatures for Au(7 nm)/YIG. Inset is schematic diagr am of patterned Hall bar. (c) Carrier concentration as a function of temperature for Au(7 nm)/YIG. 13 Fig. 2 (color online). (a) Field -dependent thermal voltage for Pt(5.1nm)/YIG and Au(10nm)/YIG ). Thermal voltage (left scale) and (right scale) for multiple strips as a function of Pt thicknesses (b) and Au thicknesses (c) on YIG. All thermal results are under a temperature difference of T ≈ 10 K. 14 Fig. 3 (color online). (a) Magnetor esistance (MR) result of Au(7nm)/YIG as a function of magnetic fi eld H at ||) and 90 (T). (b) AMR ratio as a function of metal layer thickness t for Pt/YIG (open triangles) and Au/Pt (open squares). 15 Fig. 4 (color online). (a) The optimized structural model of Pt/YIG. The teal, coral, purple, cyan and red spheres represent for Pt, Fe o (center of octahedron), Fe t (center of tetrahedron), Y and O atoms, respectively. The thickness of the superlattice, denoted as c, is 15.6 Å after relaxation . The numerals in the left side label the Pt layers for the convenien ce of discussions. (b) Isosurfaces of spin density (at 0.03 e/ Å3) of Pt/YIG. The blue and yellow isosurfaces are positive and negative spin polarizations. (c) Planar averaged spin density along c axis. The vertical dashed lines indicate the average z - coordinates of Pt and Au layers. Arrows ↑ and ↓ stand for the majority spin and minority spin contributions, respectively. | 2013-01-25 | The acute magnetic proximity effects in Pt/YIG compromise the suitability of
Pt as a spin current detector. We show that Au/YIG, with no anomalous Hall
effect and a negligible magnetoresistance, allows the measurements of the
intrinsic spin Seebeck effect with a magnitude much smaller than that in
Pt/YIG. The experiment results are consistent with the spin-polarized
density-functional calculations for Pt with a sizable and Au with a negligible
magnetic moment near the interface with YIG. | Intrinsic Spin Seebeck Effect in Au/YIG | 1301.6164v1 |
Giant Enhancement of Vacuum Friction in Spinning YIG Nanospheres Farhad Khosravi,1,2Wenbo Sun,2Chinmay Khandekar,2Tongcang Li,3,2and Zubin Jacob2,∗ 1Department of Electrical and Computer Engineering, University of Alberta, Edmonton, Alberta T6G 1H9, Canada 2Elmore Family School of Electrical and Computer Engineering, Birck Nanotechnology Center, Purdue University, West Lafayette, Indiana 47907, USA 3Department of Physics and Astronomy, Purdue Quantum Science and Engineering Institute, Purdue University, West Lafayette, Indiana 47907, USA (Dated: January 19, 2024) Experimental observations of vacuum radiation and vacuum frictional torque are challenging due to their vanishingly small effects in practical systems. For example, a rotating nanosphere in free space slows down due to friction from vacuum fluctuations with a stopping time around the age of the universe. Here, we show that a spinning yttrium iron garnet (YIG) nanosphere near aluminum or YIG slabs exhibits vacuum radiation eight orders of magnitude larger than other metallic or dielectric spinning nanospheres. We achieve this giant enhancement by exploiting the large near- field magnetic local density of states in YIG systems, which occurs in the low-frequency GHz regime comparable to the rotation frequency. Furthermore, we propose a realistic experimental setup for observing the effects of this large vacuum radiation and frictional torque under experimentally accessible conditions. I. INTRODUCTION The physics of rotating nanoparticles is gaining more attention as recent technological advancements provide experimental platforms for rotating levitated nanoparti- cles at GHz speeds [1–8]. Besides having implications in the fields of quantum gravity [9], dark energy de- tection [10], and superradiance [11], rotating nanopar- ticles are crucial for studying the effects of quantum vac- uum fluctuations [12–17]. Rotating nanoparticles can emit real photons and experience frictional torques from the fluctuating quantum vacuum even at zero temper- ature [18, 19]. Although Casimir forces between static objects have been measured extensively [20–22], the ex- perimental sensitivity is only starting to reach the limit needed to measure the frictional torque exerted on ro- tating nanoparticles from the vacuum [23]. Meanwhile, direct observation of vacuum radiation from rotating nanoparticles remains challenging due to the extremely low number of radiated photons. In the specific case of moving media or rotating par- ticles, a unique regime of light-matter interaction occurs when the material resonance frequency becomes compa- rable to the mechanical motion frequency [24–26]. In particular, a giant enhancement or even a singularity is possible in vacuum fluctuation effects [24–26]. Re- cently, world record rotation frequencies were achieved for optically levitated nanospheres [2, 3, 6]. This immedi- ately opens the question of whether unique material res- onances comparable to this rotation frequency can help enter a new regime of light-matter interaction. Here, we show that gyromagnetic yttrium iron garnet (YIG) ex- hibits the magnon polariton resonance at GHz frequen- ∗zjacob@purdue.educies[27,28]comparabletothelevitatednanoparticle’sro- tation frequency, providing a unique opportunity for en- hancing vacuum fluctuation effects on rotating nanopar- ticles. Inthisarticle,weputforthanapproachtoenhanceand observe the vacuum radiation and frictional torques by leveraging a YIG nanosphere spinning at Ω = 1 GHz in the vicinity of a metallic or YIG interface. Our proposal exploits an asymmetry between the electric and magnetic local density of states (LDOS) which was previously re- ported in Ref. [29]. In particular, near conventional met- als, the electric LDOS is enhanced at optical frequen- cies, whereas the magnetic LDOS becomes dominant at GHz frequencies. Therefore, our proposal exploits mag- netic materials with magnon polaritons to enhance the magnetic local density of states beyond those of con- ventional plasmonic metals. Due to the large magnetic LDOS and YIG magnetic resonance at GHz frequencies, the fluctuating magnetic dipoles of the YIG nanosphere canstrongly coupleto alargedensityofevanescentwaves inthe near-fieldofmetallic andmagneticinterfaces, lead- ing to colossal vacuum radiation. We demonstrate that a spinning YIG nanosphere gen- erates vacuum radiation eight orders of magnitude larger than other metallic or dielectric nanospheres in the vicin- ity of a metallic or magnetic slab. We show that, near magnetic materials, most of this radiated energy can be transferred to surface magnon polaritons. Furthermore, we reveal that the large vacuum radiation and vacuum frictionhaveexperimentallyobservableeffectsonthebal- ance rotation speed, stopping time, and balance temper- ature of the spinning YIG nanospheres under experimen- tally accessible rotation speeds, particle sizes, tempera- tures, and vacuum pressures. Therefore, the setup pro- posed in this article based on spinning YIG nanospheres represents a unique tool for detecting and analyzing vac- uum radiation and frictional torques.arXiv:2401.09563v1 [quant-ph] 17 Jan 20242 II. GIANT VACUUM RADIATION FROM SPINNING YIG NANOSPHERES We first consider the vacuum radiation from a spinning YIG nanosphere with a radius of 200 nm, as illustrated in Fig. 1(a, b). A stationary nanosphere at the equilibrium temperature exhibits zero net radiation since the num- ber of photons emitted by the fluctuating dipoles of the nanosphere is equal to the number of photons absorbed by the nanosphere from the fluctuating electromagnetic fields in the vacuum. However, for rotating nanospheres, the balance between the emitted and absorbed photons is broken. A net radiated power from the nanosphere arises even at zero temperature due to the extra boost of mechanical rotational energy [30]. The source of this vacuum radiation energy is the non-inertial motion of the nanosphere, which is transferred to generate real photons from vacuum fluctuations [19]. Based on fluctuational electrodynamics (see derivations in Appendix A), we find thetotalradiatedpowerfromaspinningYIGnanosphere Prad=R∞ 0dωℏω ΓH(ω)−ΓH(−ω) can be determined from ΓH(ω), which is the spectral density of the radiation power arising from magnetic dipole fluctuations. In the absence of any interface, vacuum radiation from a spinning YIG nanosphere does not exhibit any sub- stantial enhancement. However, metallic or magnetic in- terfaces can drastically change this observation. Metal- lic nanospheres are known to possess higher radiation rates compared to dielectric nanospheres near material interfaces [30, 31]. Here, we observe that magnetic nanospheres exhibit even larger radiation rates, which are about eight orders of magnitude compared to metal- lic nanospheres near metallic or magnetic interfaces, as shown in Fig. 1(c, d). We demonstrate that radiated photons per second per frequency expressed through ΓH(ω)−ΓH(−ω)from spinning YIG nanospheres (blue curves) are much more than those from the aluminum nanospheres(orangecurves)nearAlinterfaces(Fig.1(c)) andYIGinterfaces(Fig.1(d)). Furthermore,wefindthat a spinning YIG nanosphere radiates about 6femtowatts of power, in stark contrast to the Al sphere, which ra- diates about 6×10−7femtowatts near Al interfaces (Fig. 1(c)). In the vicinity of YIG interfaces (Fig. 1(d)), we find about 61.3femtowatts and 4.63×10−7fem- towatts of radiated power from YIG and Al nanospheres, respectively. The radiated energy mostly goes into the lossy surface waves in both metallic and magnetic mate- rials [32]. However, if the magnetic material is properly biased, as is the case studied here with a bias magnetic field of 812Oe for the YIG slab, the magnetic resonance in the magnetic slab can become resonant with the mag- netic resonance in the magnetic sphere. As a result, most of the radiated energy is transferred to surface magnon polaritons. These results clearly show the advantage of YIG over Al nanospheres for probing vacuum radiation. The above results are explained by the YIG magnon polariton resonance at GHz frequencies and differences in the low-frequency electric and magnetic LDOS near FIG. 1. (a) A YIG sphere trapped in the laser beam and spinning at 1 GHz rotation frequency in the vacuum. The stopping time for the sphere is on the order of the age of the universe. (b) YIG sphere spinning in the vicinity of an Aluminum or YIG interface exhibits colossal vacuum radi- ation. The stopping time, due to the presence of the in- terface, is reduced to about 1 day. (c, d) Number of pho- tons emitted per second per radiation frequency, defined as 1 ℏωdP/dω = Γ(ω)−Γ(−ω), for a YIG (blue solid curve) or Aluminum (dashed orange curve) nanosphere of radius 200 nm at distance d= 0.5µmfrom (c) an aluminum slab or (d) a YIG slab at room temperatures. For the Al slab, a non-local model has been used. The YIG slab in panel (d) is biased along the y direction (panel (a)) with a magnetic field ofH0= 812Oe. metallic and magnetic interfaces. Vacuum fluctuation effects on rotating nanoparticles can be significantly en- hanced when the rotation frequency is comparable to res- onance frequencies. In addition, as shown by Joulain et al.[29], LDOS near metals is dominated by the magnetic LDOS at wavelengths above a few microns. Here, we ex- tend this observation to magnetic materials and take into account the effects of non-local electromagnetic response in Al [32] (also see Appendix F). Higher magnetic LDOS thanelectricLDOSatlowfrequenciesoriginatesfromdif- ferences in the reflection of the s- and p-polarized evanes- cent waves. The near-field electric LDOS is mainly in- fluenced by p-polarized evanescent waves since their con- tributions to the electric LDOS are strongly momentum- dependent and dominate the high momentum contribu- tions crucial for near-field LDOS. In contrast, the oppo- siteistrueforthenear-fieldmagneticLDOS,andthecon- tributions from the s-polarized evanescent waves dom- inate. At GHz frequencies, the imaginary part of the reflection coefficient for evanescent s-polarized waves is much larger than that for evanescent p-polarized waves. Thus, the spolarization contributes more to the LDOS3 than the ppolarization, leading toa more dominant mag- netic LDOS near metallic and magnetic interfaces. These near-fieldLDOScanbefurtherenhancedbymaterialres- onances [24–26, 33, 34]. To this end, we discuss the spectral density ΓH(ω)that determines the vacuum radiation. Through a sim- ilar approach as the methods used by Abajo and Man- javacas [18], our result for the radiation spectral den- sityΓH(ω)of a spinning gyromagnetic nanosphere due to magnetic dipole fluctuations is (see derivations in Ap- pendix A): ΓH(ω) = (ωρ0/8)(h gH ⊥,2(ω) + 2gH ∥(ω) + 2gH g,2(ω)i Im αm,⊥(ω−) −Re αm,g(ω−) n1(ω−)−n0(ω) +gH ⊥,1(ω)Im αm,∥(ω) [n1(ω)−n0(ω)]) ,(1) where ρ0=ω2/c2π3is the vacuum density of states, gH ⊥,1 ,gH ⊥,2are the two components of the magnetic Green’s function in the plane of the interface (the xxandzz components for the setup shown in Fig. 1(b)), gH ∥is the component normal to the interface (the yycomponent here), and gH g,2is the off-diagonal component between the in-plane and normal directions (the xycomponent here), all normalized by πωρ 0/8.αm,⊥(ω),αm,g(ω), and αm,∥(ω)are the xx(oryy),xy, and zzcomponents of the YIG nanosphere magnetic polarizability tensor in the ro- tating sphere frame (see Appendix D for derivations). Ω is rotating frequency of the nanosphere and ω−=ω−Ω. n1(ω)andn0(ω)are the Bose-Einstein distribution func- tions pertinent to the sphere temperature T1and the en- vironment temperature T0, respectively. Detailed deriva- tions for all these quantities and discussions of various YIG interface orientations and bias magnetic field direc- tions are provided in Appendix B. When the sphere is stationary ω−=ω, and the sphere temperature is equal tothetemperatureoftheenvironment T1=T0,theterms n1(ω−)−n0(ω)andn1(ω)−n0(ω)become zero; thus, the radiation becomes zero as expected. Here, we emphasize one important aspect of ΓH(ω)re- garding the rotation-induced magnetization of the YIG nanosphere, which can occur without any external mag- netic field. This is known as the Barnett effect and origi- natesfromtheconservationofangularmomentum, where the mechanical angularmomentumof the sphereistrans- ferred to the spin of the unpaired electrons in the mag- netic material [35]. Assuming the magnetic field is paral- lel to the rotation axis, the Larmor precession frequency ω0of the electrons inside the sphere is [36] (also see Ap- pendix E): ω0= Ω + µ0γH0, (2) for the electron gyromagnetic ratio γ, vacuum permeabil- ityµ0, and applied external magnetic field H0. We in- corporate this effect on ω0to find the magnetic response of the spinning YIG nanosphere.III. ENHANCEMENT OF VACUUM FRICTIONAL TORQUE We now discuss the vacuum frictional torque exerted on the rotating YIG nanosphere in the vicinity of YIG and Al interfaces. We use a similar approach to find the vacuum torque exerted on the spinning gyromagnetic YIG sphere due to magnetic dipole and magnetic field fluctuations (detailed derivations are provided in Ap- pendix G). The torque along the axis of rotation is given byMz=R∞ 0dωℏ ΓH M(ω) + ΓH M(−ω) , where the expres- sion for ΓH M(ω)is similar to the expression for ΓH(ω)in Eq. (1), with the difference being that the last term on thesecondlineisnotpresentin ΓH M(ω)(seeAppendixG). Additionally, wefindthatothercomponentsofthetorque (MxandMycomponents) are not necessarily zero in the vicinity of the YIG interface, in contrast to the Al slab. Due to the anisotropy of the YIG slab, MxandMydo not vanish for some directions of the bias magnetic field. We provide further discussions of these cases in the sup- plementary material. In Fig. 2, we compare vacuum torques exerted on spin- ning YIG nanospheres (Fig. 2(a, c)) and spinning Al nanospheres (Fig. 2(b, d)), on nanospheres spinning in the vicinity of YIG slabs (Fig. 2(a, b)) and Al slabs (Fig. 2(c, d)), as well as on nanospheres spinning in the vicinity of slabs (solid colored curves) and spinning in vacuum(dashedblackcurves). Wedemonstratethatvac- uum torques exhibit more than 10 orders of magnitude enhancement in the vicinity of YIG and Al slabs com- pared to the vacuum, and about 4 orders of magnitude enhancement due to employing YIG nanospheres instead ofAlnanospheres. Theseresultsunraveltheadvantageof utilizing YIG nanospheres for probing vacuum frictional torques at GHz frequencies. In Fig. 2, we consider non- local electromagnetic response [32] for Al interfaces and incorporate effects from the magnetic and electric dipole and field fluctuations on vacuum torques. We notice that the vacuum torque is dominated by magnetic rather than electric fluctuations in all cases (see Appendix G). In ad- dition, we have taken into account the effect of recoil4 FIG. 2. The negative vacuum frictional torque experienced by the YIG and aluminum nanosphere with a radius of 200 nm at room temperature. (a) Torque experienced by a YIG sphere in the vicinity of the YIG slab (solid blue curve) and in vacuum (dashed black curve). (b) Torque exerted on an Al sphere in the vicinity of the YIG slab (solid orange curve) and in vacuum (dashed black curve). (c), (d) the same as (a) and (b) with the YIG slab replaced by an Al slab. The YIG slab is biased along the ydirection with H0= 812Oe (see Fig. 1(a)). A non-local model is used for the Al slabs. The distance between the spinning spheres and slabs is d= 0.5µm for all cases. Placing the YIG or Al interface in the vicinity of spinning nanospheres results in about 12 orders of magnitude increase in the exerted vacuum torque. torque [37] – the torque exerted on the sphere due to the scatteringofvacuumfieldfluctuationsofftheparticle. As discussed in Appendix G, we find that effects from this second-order torque are negligible compared with the ef- fects of magnetic fluctuations in the studied cases. IV. OBSERVABLE OUTCOMES OF GIANT VACUUM FRICTION IN SPINNING YIG NANOSPHERES The observable effects of the colossal vacuum radia- tion and frictional torques come down to changes in ex- perimentally measurable parameters when the spinning nanosphereisbroughtclosertothevicinityofAl/YIGin- terfaces. In Fig. 3(a), we show the proposed experimen- tal setup for this observation where a YIG nanosphere is trapped inside an Al or YIG ring. We note that the size of the ring is much larger than that of the nanosphere, and it does not lead to any resonant behavior. However, for smaller ring sizes, LDOS can be further enhanced compared to the slab interface case due to the presence FIG. 3. Experimental considerations of the setup. (a) Pro- posed experimental setup with nanosphere trapped inside a ring. (b) Balance rotation speed Ωbfor Al sphere (red curve) and YIG sphere in the presence of Al (blue curve) and YIG (pink curve) interfaces, as a function of distance dfrom the interfacefora 200nmradiussphereat 10−4Torrvacuumpres- sure. The values are normalized by the vacuum balance rota- tion speed Ω0. (c) Characteristic stopping time as a function of distance from the interface at 10−6Torr vacuum pressure. (d) Balance temperature of the YIG sphere Tsatd= 500 nm distance from Al (blue curve) and YIG (pink curve) in- terfaces as a function of lab temperature T0, at10−4Torr vacuum pressure. For Al spheres, there is no final tempera- ture as the temperature keeps rising with time. of interfaces on all sides. We evaluated some observable experimental outcomes due to large vacuum radiation and friction. This analy- sis is based on the experimentally accessible parameters from Refs. [3, 38, 39]. In Fig. 3(b), we show the balanced rotation speed Ωbof the spinning nanosphere normalized bytherotationspeed Ω0intheabsenceofanyinterfaceas a function of distance dfrom the interface. The balance rotation speed is defined as the sphere’s stable, perpetual rotation speed and occurs when the driving force due to the laser is equal to the drag force due to the vacuum chamber. In the absence of any interface, due to the neg- ligible value of vacuum radiation, the balance rotation speed Ω0is obtained when the torque from the trapping laser balances the frictional torque from air molecules in the imperfect vacuum [3] (also see Apeendix H). We as- sume the laser driving torque is constant and the drag force from air molecules has a linear dependence on rota- tional speeds [3]. In Fig. 3(b), we show that the balance rotation speed of the YIG nanosphere is reduced when it is closer to Al (blue curve) or YIG (pink curve) in- terfaces, as a result of the large frictional torques from vacuum fluctuations. Remarkably, we notice that there5 is no observable change in the balance speed for spinning Al nanospheres in the vicinity of Al or YIG interfaces (red curve). In Fig. 3(c, d), we further demonstrate outcomes of the large vacuum radiation in other experimental observ- ables, such as the stopping time as a function of distance (Fig. 3(c)) and the balance temperature as a function of the vacuum temperature T0(Fig. 3(d)). Stopping time is the time constant of the exponential decrease of the nanosphere rotation velocity after the driving torque is turned off. The torque can be switched off by chang- ing the polarization of the trapping laser from circular to linear without having to switch off the trapping laser. The balance temperature refers to the nanosphere tem- perature Ts, at which the loss of mechanical rotational energy due to vacuum frictional torque stops heating the nanospheres. As shown in Fig. 3(c, d), YIG nanospheres exhibit distinct behaviors in the stopping time and bal- ance temperature compared to Al nanospheres near YIG and Al interfaces. The results of Fig. 3 show that the vacuum radiation and frictional torque can be experimentally measured through the balance speed, balance temperature, and stopping time of the YIG nanosphere. In stark contrast, the Al nanosphere (or any other metallic nanospheres) may not experience enough vacuum friction to exhibit observable outcomes unless it is in a sensitive setup with very low vacuum pressure [3, 23]. V. DISCUSSION AND CONCLUSION Our results show that due to YIG magnon polariton resonance and the dominance of magnetic LDOS over electric LDOS in the vicinity of metallic or magnetic ma- terials at GHz frequencies, spinning YIG nanospheres can exhibit orders of magnitude larger vacuum radia- tion and frictional torque compared to any metallic or dielectric nanosphere. By investigating the case of a YIG nanosphere spinning at 1 GHz speed, we have shown that the effect of colossal vacuum fluctuations can be observed in an experimentally accessible setup. Our re- sults set a new perspective for observing and understand- ing radiation and frictional torques from vacuum fluctu- ations. Furthermore, our discussions of magnetic LDOS near YIG interfaces under various bias fields pave the way for magnetometry [40] and spin measurement [41] applications. ACKNOWLEDGEMENTS This research was supported by the Army Research Office under grant number W911NF-21-1-0287 and the Office of Naval Research under award number N000142312707.Appendix A: Radiation Power due to Magnetic Fluctuations In this appendix, we provide detailed derivations of the radiation power Pradfrom a spinning YIG nanosphere and its spectral density ΓH(ω)due to magnetic fluctua- tions. Using an approach similar to that taken by Abajo et. al[18, 30], we can write the radiated power due to the magnetic fluctuations of dipoles and fields as, Pmag=−⟨Hind·∂mfl/∂t+Hfl·∂mind/∂t⟩,(A1) where Hindis the induced magnetic field due to the mag- netic dipole fluctuations mflof the particle and mindis the induced magnetic dipole in the particle due to the fluctuations of the vacuum magnetic field Hfl. Note that all of these quantities are written in the lab frame. For the sphere spinning at the rotation frequency Ω, we can write, mfl x=m′fl xcos Ω t−m′fl ysin Ωt, mfl y=m′fl xsin Ωt+m′fl ycos Ω t, mfl z=m′fl z,(A2) where the primed quantities are written in the rotat- ing frame. Performing a Fourier transform as m′fl(t) =Rdω 2πe−iωtm′fl(ω), we can write in the frequency domain mfl x(ω) =1 2h m′fl x(ω−) +m′fl x(ω+) +im′fl y(ω+)−im′fl y(ω−)i , mfl y(ω) =1 2h im′fl x(ω−)−im′fl x(ω+) +m′fl y(ω+) +m′fl y(ω−)i , (A3) where ω±=ω±Ω. We can similarly write for the mag- netic fields H′fl x(ω) =1 2 Hfl x(ω+) +Hfl x(ω−)−iHfl y(ω+) +iHfl y(ω−) , H′fl y(ω) =1 2 iHfl x(ω+)−iHfl x(ω−) +Hfl y(ω+) +Hfl y(ω−) . (A4) Thus, using the fact that, m′ind(ω) =¯αm(ω)·H′fl(ω), (A5) with ¯αm(ω) = αm,⊥(ω)−αm,g(ω) 0 αm,g(ω)αm,⊥(ω) 0 0 0 αm,∥(ω) (A6) being the magnetic polarizability tensor of the YIG sphere biased along the zaxis, we find in the lab frame mind(ω) =¯αeff m(ω)·Hfl(ω), (A7)6 where ¯αeff m= αeff m,⊥(ω)−αeff m,g(ω) 0 αeff m,g(ω)αeff m,⊥(ω) 0 0 0 αeff m,∥(ω) , (A8a) αeff m,⊥(ω) =1 2 αm,⊥(ω+) +αm,⊥(ω−) +iαm,g(ω+)−iαm,g(ω−) , (A8b) αeff m,g(ω) =−i 2 αm,⊥(ω+)−αm,⊥(ω−) +iαm,g(ω+) +iαm,g(ω−) , (A8c) αeff m,∥=αm,∥(ω). (A8d) Note that we have used an expression similar to Eq. (A3) but written for the induced magnetic dipole moments. Expression for αm,⊥(ω)andαm,g(ω)are given in Appendix D. Using the fluctuation-dissipation theorem (FDT) [42], ⟨Hfl i(ω)Hfl j(ω′)⟩= 4πℏ[n0(ω) + 1]( GH ij(ω)−GH∗ ji(ω) 2i) δ(ω+ω′), (A9) with GH ij(ω) =GH ij(r,r′=r, ω)defined as the equal-frequency magnetic Green’s function of the environment defined through the equation, Hi(r,r′, ω) =GH ij(r,r′, ω)mj(r′, ω), (A10) we find the second term in Eq. (A1) employing Eqs. (A4) and (A5): ⟨Hfl i(ω)∂mind i(ω′)/∂t⟩=−iω′2πℏh n0(ω) + 1i δ(ω+ω′)× ( Im GH xx(ω) +Im GH yy(ω) h αm,⊥(ω′+) +αm,⊥(ω′−) +iαm,g(ω′+)−iαm,g(ω′−)i + Re GH xy(ω) −Re G∗H yx(ω) h αm,⊥(ω′+)−αm,⊥(ω′−) +iαm,g(ω′+) +iαm,g(ω′−)i +2Im GH zz(ω) αm,∥(ω′)) , (A11) where n0(ω) = 1 /(eℏω/kBT0−1)is the Planck distribution at the temperature of the lab T0. Writing FDT for the fluctuating dipoles, ⟨m′fl i(ω)m′fl j(ω′)⟩= 4πℏ[n1(ω) + 1]αm,ij(ω)−α∗ m,ji(ω) 2i δ(ω+ω′), (A12)7 we find the first term in Eq. (A1) employing Eq. (A3) and Hind i(ω) =GH ij(ω)mfl j(ω): ⟨Hind i(ω)∂mfl i(ω′)/∂t⟩=−2πℏiω′ n1(ω−) + 1( δ(ω+ω′)h GH xx(ω)Im αm,⊥(ω−) −GH xx(ω)Re αm,g(ω−) −GH yy(ω)Re αm,g(ω−) +GH yy(ω)Im αm,⊥(ω−) +iGH xy(ω)Im αm,⊥(ω−) −iGH xy(ω)Re αm,g(ω−) +iGH yx(ω)Re αm,g(ω−) −iGH yx(ω)Im αm,⊥(ω−) i) −2πℏiω′ n1(ω+) + 1( δ(ω+ω′)h GH xx(ω)Im αm,⊥(ω+) +GH xx(ω)Re αm,g(ω+) +GH yy(ω)Re αm,g(ω+) +GH yy(ω)Im αm,⊥(ω+) −iGH xy(ω)Im αm,⊥(ω+) −iGH xy(ω)Re αm,g(ω+) +iGH yx(ω)Re αm,g(ω+) +iGH yx(ω)Im αm,⊥(ω+) i) −4πℏiω′[n1(ω) + 1]( δ(ω+ω′)GH zz(ω)Im αm,∥(ω) ) ,(A13) where n1(ω)is the Planck distribution at the sphere temperature T1. Taking the inverse Fourier transform, adding Eqs. (A11) and (A13), taking the real part of the radiated power, and changing integral variables, we find Pmag=ℏ πZ+∞ −∞ωdω( n1(ω−)−n0(ω) Im GH xx(ω) +Im GH yy(ω) +Re GH xy(ω) −Re GH yx(ω) × Im αm,⊥(ω−) −Re αm,g(ω−) + [n1(ω)−n0(ω)]Im GH zz(ω) Imn αH m,∥(ω)o) .(A14) In this derivation, we have used the property αm(−ω) =α∗ m(ω). The expressions for Green’s functions in different YIG and aluminum interface arrangements are given in Appendix B. Plugging these expressions into Eq. (A14), we obtain Eq. (1) in the main text. Appendix B: Green’s Function Near an Anisotropic Magnetic Material In this appendix, we provide the Green’s function near a half-space of magnetic material, which would change due to the anisotropy of the material. We study two cases when the interface is the x−yplane and x−zplane, as shown in Fig. 4. We can write the electric and magnetic fields in the vacuum as E=Ei+Er,H=Hi+Hr, (B1a) Ei= (E0sˆs−+E0pˆp−)eik−·r, (B1b) Er= (E0srssˆs++E0prppˆp++E0srpsˆp++E0prspˆs+)eik+·r, (B1c) Hi=1 η0(−E0sˆp−+E0pˆs−)eik−·r, (B1d) Hr=1 η0(−E0srssˆp++E0prppˆs++E0srpsˆs+−E0prspˆp+)eik+·r, (B1e) where ˆs±,ˆp±, and ˆk±/k0form a triplet with ˆk±=k0(κcosϕˆx+κsinϕˆy±kzˆz),ˆs±= sin ϕˆx−cosϕˆy,ˆp±=−(±kzcosϕˆx±kzsinϕ−κˆz),(B2)8 FIG. 4. Schematic of the problem for the two cases of when the interface is in (a) x−yplane and (b) x−zplane. andη0=p µ0/ϵ0,k0=ω/c,κ2+k2 z= 1, and k0kzis the zcomponent of the wavevector. Similarly, we can write the electric and magnetic fields inside the magnetic material as E′=Et,H′=Ht, (B3a) Et= E0stssˆs′ −+E0ptppˆp′ −+E0stpsˆp′ −+E0ptspˆs′ − eik′ −·r, (B3b) ¯¯µHt=p κ2+k′2z η0 −E0stssˆp′ −+E0ptppˆs′ −+E0stpsˆs′ −−E0ptspˆp′ − eik′ −·r, (B3c) where k′ ±=k0ˆk′ ±=k0(κcosϕˆx+κsinϕˆy±k′ zˆz),ˆs′ ±= sin ϕˆx−cosϕˆy,ˆp′ ±=−±k′ zcosϕˆx±k′ zsinϕˆy−κˆzp κ2+k′2z.(B4) Note that κis the same in the two media due to the boundary conditions. Also ˆk′ ±×ˆp′ ±= ˆs′ ±. We can write Maxwell’s equations in the magnetic material in matrix form as [43] (M+Mk)ψ=¯¯ϵ0 0¯¯µ + 0¯¯κ −¯¯κ0 Et η0Ht = 0, (B5) where ¯¯κ= 0 −k′ zκsinϕ k′ z 0−κcosϕ −κsinϕ κcosϕ 0 . (B6) Setting the det (M+Mk) = 0we get the solutions for k′ zin terms of κandϕ[43]. From these solutions and applying the boundary conditions, we can find the values of rss,rsp,rps,rppfor a given κandϕ. Note that different bias directions for the magnetic field of the YIG slab change the ¯¯µtensor and thus change the reflection coefficients rss, rsp,rps,rpp. In the following, we first provide the expression for the magnetic dyadic Green’s function ¯GHfor a source at z′=d when the interface is in the x−yplane (Fig. 4(a)). Here, we take the spinning sphere to be at the origin to simplify the derivations and move z= 0toz′=d. This would not change the Fresnel reflection coefficients. The incident magnetic Green’s function at the location of the source is thus, ¯GH i(z=z′, ω) =ik2 0 8π2ϵmZdkxdky kz(ˆsˆs+ ˆp−ˆp−)eikx(x−x′)+iky(y−y′). (B7) The reflected magnetic Green’s function at the location of the source is ¯GH r(z=z′, ω) =ik2 0 8πϵmZdkxdky kz(ˆsrppˆs+ ˆp+rspˆs+ ˆsrpsˆp−+ ˆp+rssˆp−)e2ikzd, (B8) where kx=κcosϕandky=κsinϕ. NotethatheretheFresnelreflectioncoefficientsgenerallydependontheincidence angle ϕ. For the special case of magnetization along the zaxis, they become independent of ϕ. Using Eq. (B4) and9 dropping the terms that vanish after integration over ϕ, we can write the total magnetic Green’s function at the location of source as, ¯GH(r,r, ω) =ik3 0 8π2Z2π 0dϕZ+∞ 0κdκ p( sin2ϕˆxˆx+ cos2ϕˆyˆy−sinϕcosϕ(ˆxˆy+ ˆyˆx) 1 +rppe2ikzd +p2cos2ϕˆxˆx+p2sin2ϕˆyˆy+κ2ˆzˆz +e2ik0pdrss −p2cos2ϕˆxˆx−p2sin2ϕˆyˆy+κ2ˆzˆz−p2cosϕsinϕ(ˆxˆy+ ˆyˆx) −pκcosϕ(ˆxˆz−ˆzˆx)−pκsinϕ(ˆyˆz−ˆzˆy) +e2ik0pdrps psinϕcosϕ(ˆxˆx−ˆyˆy) +psin2ϕˆxˆy−pcos2ϕˆyˆx+κsinϕˆxˆz−κcosϕˆyˆz +e2ik0pdrsp −pcosϕsinϕ(ˆxˆx−ˆyˆy) +pcos2ϕˆxˆy−psin2ϕˆyˆx+κsinϕˆzˆx−κcosϕˆzˆy) .(B9) Note that the electric Green’s function can be obtained by changing rsstorpp,rpptorss,rpstorspandrsptorps and dividing by ϵ0. In general, the non-diagonal parts of the Green’s function are not zero. Using this equation, we find, Im GH xx(ω) =πωρ 0 8gH ⊥,1(ω), (B10a) Im GH yy(ω) =πωρ 0 8gH ⊥,2(ω), (B10b) Re GH xy(ω) −Re GH yx(ω) =πωρ 0 4gH g,1(ω), (B10c) Im GH zz(ω) =πωρ 0 4, gH ∥(ω) (B10d) where ρ0=ω2/π2c3is the vacuum density of states and, gH ⊥,1(ω) =1 πZ2π 0dϕ(Z1 0κdκ ph 1 + sin2ϕRe rppe2ik0pd −κ2cos2ϕ+ cos2ϕ κ2−1 Re rsse2ik0pd +psinϕcosϕRe e2ik0pd(rps−rsp) i +Z∞ 1κdκ |p| sin2ϕIm{rpp}+ cos2ϕ κ2−1 Im{rss}+|p|sinϕcosϕRe{rps−rsp} e−2k0|p|d) , (B11a) gH ⊥,2(ω) =1 πZ2π 0dϕ(Z1 0κdκ ph 1 + cos2ϕRe rppe2ik0pd −κ2sin2ϕ+ sin2ϕ κ2−1 Re rsse2ik0pd −psinϕcosϕRe e2ik0pd(rps−rsp) i +Z∞ 1κdκ |p| cos2ϕIm{rpp}+ sin2ϕ κ2−1 Im{rss} − |p|sinϕcosϕRe{rps−rsp} e−2k0|p|d) , (B11b) gH g,1(ω) =−1 πZ2π 0( dϕZ1 0κdκ sin2ϕIm rpse2ik0pd + cos2ϕIm rspe2ik0pd +Z∞ 1κdκ sin2ϕIm{rps}+ cos2ϕIm{rsp} e−2k0|p|d) ,(B11c)10 gH ∥(ω) =1 2πZ2π 0dϕ(Z1 0κ3dκ p 1 +Re rsse2ik0pd +Z∞ 1κ3dκ |p|e−2k0|p|dIm{rss}) . (B11d) Plugging Eq. (B10) into Eq. (A14), we find, Pmag=Z∞ −∞dωℏωΓH(ω), (B12) with, ΓH(ω) = (ωρ0/8)( gH ⊥,1(ω) +gH ⊥,2(ω) + 2gH g,1(ω) Im αm,⊥(ω−) −Re αm,g(ω−) n1(ω−)−n0(ω) +2gH ∥(ω)Im αm,∥(ω) [n1(ω)−n0(ω)]) .(B13) For the case when the YIG interface is the x−zplane (Fig. 4(b)), we find the radiated power by exchanging the axes ˆx→ˆz,ˆy→ˆx, and ˆz→ˆyin Eq. (B9). In this case, we have Im GH xx(ω) =πωρ 0 8gH ⊥,2(ω), (B14a) Im GH yy(ω) =πωρ 0 4gH ∥(ω), (B14b) Im GH zz(ω) =πωρ 0 8gH ⊥,1, (B14c) where gH ⊥,1,gH ⊥,2, and gH ∥given by Eq. (B11). For the xyandyxcomponent of the Green’s function, however, we get Re GH xy(ω) −Re GH yx(ω) =πωρ 0 4gH g,2(ω), (B15) with gH g,2(ω) =1 πZ2π 0dϕ(Z1 0κ2dκ p psinϕIm rsse2ik0pd +cosϕ 2Im (rps−rsp)e2ik0pd +Z∞ 1κ2κ |p| |p|sinϕIm{rss} −cosϕ 2Re{rsp−rps} e−2k0|p|d) ,(B16) and thus we have for the case when the YIG interface is the x−zplane, ΓH(ω) = (ωρ0/8)(h gH ⊥,2(ω) + 2gH ∥(ω) + 2gH g,2(ω)i Im αm,⊥(ω−) −Re αm,g(ω−) n1(ω−)−n0(ω) +gH ⊥,1(ω)Im αm,∥(ω) [n1(ω)−n0(ω)]) ,(B17) with gH ⊥,1,gH ⊥,2, and gH ∥given by Eq. (B11) and gH g,2by Eq. (B16). This is the same as Eq. (1) in the main manuscript. Appendix C: Dominance of Magnetic Local Density of States Although the expressions found in the previous sections for the radiated power Pradare not, in general, exactly proportional to the local density of states (LDOS), they are proportional to terms of the same order as the LDOS. The expression for LDOS is given by [29], ρ(r, ω) =1 πωTr ϵ0Im GE(r,r, ω) +Im GH(r,r, ω) , (C1)11 where the Tr represents the trace operator. Using the expressions of the previous section, it is easy to see that the LDOS at the location of the nanosphere is given by, ρ(ω) = (ρ0/8)h ϵ0(gE ⊥,1+gE ⊥,2+ 2gE ∥) +gH ⊥,1+gH ⊥,2+ 2gH ∥i , (C2) wheretheexpressionsfor gH ⊥,1,gH ⊥,2, and gH ∥aregivenbyEq.(B11)andtheexpressionfortheelectricGreen’sfunctions are found from the magnetic ones by replacing s→pandp→sand dividing by ϵ0. As discussed before, the magnetic Green’s functions are about eight orders of magnitude larger than the electric ones at GHz frequencies, and thus, the LDOS is dominated by the magnetic LDOS. This shows that the magnetic field fluctuations dominate the vacuum radiation, vacuum torque, and LDOS simultaneously. Appendix D: Magnetic Polarizability Tensor of YIG In the appendix, we provide derivations of the YIG polarizability tensor. We consider the Landau-Lifshitz-Gilbert formula to describe the YIG permeability tensor [36], ¯¯µ= µ⊥−µg0 µgµ⊥0 0 0 µ∥ , (D1) where µ⊥(ω) =µ0(1 +χ⊥) =µ0( 1 +ω0ωm(ω2 0−ω2) +ω0ωmω2α2+i αωω m ω2 0+ω2(1 +α2) [ω2 0−ω2(1 +α2)]2+ 4ω2 0ω2α2) ,(D2a) µg(ω) =µ0χg=µ0−2ω0ωmω2α+iωωm ω2 0−ω2(1 +α2) [ω2 0−ω2(1 +α2)]2+ 4ω2 0ω2α2, (D2b) µ∥=µ0, (D2c) andω0=µ0γH0is the Larmor precession frequency with γbeing the gyromagnetic ratio and H0the bias magnetic field (assumed to be along ˆzdirection), ωm=µ0γMswith Msbeing the saturation magnetization of the material, andαis the YIG damping factor related to the width of the magnetic resonance through ∆H= 2αω/µ 0γ. In the main text, we considered Ms= 1780 Oe and∆H= 45 Oe [36] in our calculations. When the magnetic field is reversed (along −ˆzdirection), we can use the same results by doing the substitutions ω0→ −ω0, ω m→ −ωm, α→ −α, (D3) which gives µ⊥→µ⊥, µ g→ −µg. (D4) Using the method in Ref. [44] for the polarizability tensor of a sphere with arbitrary anisotropy, we find the polarizability tensor of YIG with the permeability tensor described by Eq. (D1), ¯¯αm= 4πa3 (µ⊥−µ0)(µ⊥+2µ0)+µ2 g (µ⊥+2µ0)(µ⊥+2µ0)+µ2g−3µ0µg (µ⊥+2µ0)(µ⊥+2µ0)+µ2g0 3µ0µg (µ⊥+2µ0)(µ⊥+2µ0)+µ2g(µ⊥−µ0)(µ⊥+2µ0)+µ2 g (µ⊥+2µ0)(µ⊥+2µ0)+µ2g0 0 0µ∥−µ0 µz+2µ0 . (D5) Therefore the magnetic polarizability terms in Eqs. (B13) and (B17) are given by, αm,⊥(ω) = 4 πa3(µ⊥−µ0)(µ⊥+ 2µ0) +µ2 g (µ⊥+ 2µ0)(µ⊥+ 2µ0) +µ2g, (D6a)12 αm,g(ω) = 4 πa3 3µ0µg (µ⊥+ 2µ0)(µ⊥+ 2µ0) +µ2g, (D6b) where µ⊥andµgare frequency dependent terms give by Eq. (D2). It is important to note that magnetostatic approximation has been assumed in the derivation of the magnetic polarizability. This is similar to the electrostatic approximation used for the derivation of the electric polarizability [45], where, using the duality of electromagnetic theory, the electric fields and electric dipoles have been replaced by the magnetic fields and magnetic dipoles. In this approximation, the fields inside the sphere are assumed to be constant. One can apply the Mie theory to find the magnetic polarizability to the first order in the Mie scattering components. This, however, is mathematically challenging due to the anisotropy of the magnetic material. For the purpose of our study, the magnetostatic assumption is enough to find the polarizability properties of YIG since the size of the sphere is much smaller compared to the wavelength, and the polarizability is dominated by the magneto-static term. For metals, however, higher order terms are important for finding the magnetic polarizability since the magneto- static terms are zero and only higher order terms due to electric dipole fluctuations give rise to the magnetic polar- izability of metals [30]. We provide derivations based on Mie theory for the polarizability constant of an aluminum particle in Section S1 in the supplementary material. Appendix E: Barnett Effect In the simplest models of magnetic materials, electrons are assumed to be magnetic dipoles with the moments µBspinning about the magnetization axis determined by the applied magnetic field H0with the Larmor precession frequency ω0=µ0γH0, where γis the gyromagnetic ratio of the material [36]. Barnett showed that the spontaneous magnetization of a material with the magnetic susceptibility of χis given by [35] Mrot=χΩ/γ, (E1) where Ωis the rotation frequency of the magnetic material. This magnetization can be assumed to be caused by an applied magnetic field Hrotwhich is Hrot=Mrot/χ=Ω γµ0. We thus get the Larmor frequency due to rotation, ω0,rot= Ω. (E2) Therefore, the Larmor frequency of a spinning magnetic material is the same as the rotation frequency. We thus can write the total Larmor frequency of spinning YIG as ω0= Ω + µ0γH0. (E3) We use this expression to find the permeability tensor of a spinning YIG nanosphere discussed in Appendix D. Appendix F: Non-local Model for Aluminum Since the sphere is spinning in close proximity to material interfaces, the non-local effects in aluminum electromag- netic response can become important. Here, we employ the non-local Fresnel reflection coefficients from Ref. [46]. rss=Zs−4π cp Zs+4π cp, r pp=4πp/c−Zp 4πp/c +Zp, (F1) where p=√ 1−κ2, and Zs=8i cZ∞ 0dq1 ϵt(k, ω)−(q2+κ2), (F2a) Zp=8i cZ∞ 0dq1 q2+κ2q2 ϵt(k, ω)−(q2+κ2)+κ2 ϵl(k, ω) , (F2b) with the longitudinal and transverse dielectric permittivities given by ϵl(k, ω) = 1 +3ω2 p k2v2 F(ω+iΓ)fl(u) ω+iΓfl(u), (F3a)13 ϵt(k, ω) = 1−ω2 p ω(ω+iΓ)ft(u), (F3b) with k2= (ω/c)2 q2+κ2 ,u= (ω+iΓ)/(kvF), and fl(u) = 1−1 2ulnu+ 1 u−1, f t(u) =3 2u2−3 2u(u2−1) lnu+ 1 u−1. (F4) These expressions give the non-local reflection coefficients at a metallic interface for the semi-classical infinite barrier (SCIB) model. The SCIB model is accurate as long as z=k 2kF∼0, where kF=mvF/ℏwith mbeing the free-electron mass. For example, for aluminum with vF≃2.03×106m/s, we have kF≃1.754×1010andk=ω/c≃20, which shows that for our case the SCIB model is valid. Appendix G: Vacuum Frictional Torque In this section, we provide the derivations of the vacuum frictional torque exerted on the spinning YIG nanosphere due to vacuum fluctuations. The torque on a magnetic dipole is given by M=m×H. (G1) Since we are interested in the torque along the rotation axis ( zdirection), we can write the torque as Mz=ˆz· ⟨mfl×Hind+mind×Hfl⟩ =⟨mfl xHind y−mfl yHind x+mind xHfl y−mind yHfl x⟩,(G2) using the Fourier transform, we get Mz=Zdωdω′ (2π)2e−i(ω+ω′)th ⟨mfl x(ω)Hind y(ω′)⟩ − ⟨mfl y(ω)Hind x(ω′)⟩+⟨mind x(ω)Hfl y(ω′)⟩ − ⟨mind y(ω)Hfl x(ω′)⟩i .(G3) Through a similar approach as that used in Appendix A, after some algebra, we find Mz=ℏ 2πZ∞ −∞dω( Im GH yy(ω) +Im GH xx(ω) +Re GH yx(ω) −Re GH xy(ω) × Im αm,⊥(ω+) +Re αm,g(ω+) n1(ω+)−n0(ω) − Im GH yy(ω) +Im GH xx(ω) −Re GH yx(ω) +Re GH xy(ω) × Im αm,⊥(ω−) −Re αm,g(ω−) n1(ω−)−n0(ω)) ,(G4) which can be written as Mz=−Z+∞ −∞dωℏΓH M(ω). (G5) For an interface in the x−yplane ΓH Mis given by ΓH M(ω) = (ωρ0/8) gH ⊥,1(ω) +gH ⊥,2(ω) + 2gH g,1(ω) Im αm,⊥(ω−) −Re αm,g(ω−) n1(ω−)−n0(ω) ,(G6) which is the same expression for the radiated power minus the term related to the axis of rotation z. For an interface in the x−zplane, on the other hand, ΓH Mwe have ΓH M(ω) = (ωρ0/8)h gH ⊥,2(ω) + 2gH ∥(ω) + 2gH g,2(ω)i Im αm,⊥(ω−) −Re αm,g(ω−) n1(ω−)−n0(ω) .(G7) This expression is the same as Eq. (1) in the main manuscript, with the difference that it does not have the last term involving the term n1(ω)−n0(ω). Compared to the vacuum radiation expression, vacuum torque has an extra minus sign in Eq. (G5), indicating that this torque acts as friction rather than a driving force, as expected.14 1. Other components of torque In the previous section, we only derived the zcomponents of the torque exerted on the nanosphere. The xandy components can be written as Mx=⟨mfl yHind z−mfl zHind y+mind yHfl z−mind zHfl y⟩, (G8a) My=⟨mfl zHind x−mfl xHind z+mind zHfl x−mind xHfl z⟩. (G8b) UsingasimilarapproachasthatusedintheprevioussectionandsectionA,incorporatingthetorqueduetotheelectric field fluctuations of vacuum and the magnetic dipole fluctuations of the YIG sphere, we find for the xcomponent of torque, Mx=ℏ 4πZ∞ −∞dω( 2n1(ω−) + 1 Im αm,⊥(ω−) −Re αm,g(ω−) 2Im GH zx(ω) + 2Re GH zy(ω) −4 [n1(ω) + 1]Im αm,∥(ω) Re GH yz(ω) + [2n0(ω) + 1]( Re{αm,⊥(ω−)}+Im{αm,g(ω−)} Re{GH xz(ω)} −Re{GH zx(ω)}+Im{GH yz(ω)}+Im{GH zy(ω)} + Im{αm,⊥(ω−)} −Re{αm,g(ω−)} −Im{GH xz(ω)} −Im{GH zx(ω)}+Re{GH yz(ω)} −Re{GH zy(ω)}) + [n0(ω) + 1]( −2Re{αm,∥(ω)} Im{GH zy(ω)}+Im{GH yz(ω)} + 2Im{αm,∥(ω)} −Re{GH zy(ω)}+Re{GH yz(ω)}) , (G9) and for the ycomponent, My=ℏ 4πZ∞ −∞dω( 2n1(ω−) + 1 Im αm,⊥(ω−) −Re αm,g(ω−) −2Re{GH zx(ω)}+ 2Im{GH zy(ω)} +4 [n1(ω) + 1]Im αm,∥(ω) Re{GH xz(ω)}) −[2n0(ω) + 1]( Re{αm,⊥(ω−)}+Im{αm,g(ω−)} Im{GH xz(ω)}+Im{GH zx(ω)} −Re{GH yz(ω)}+Re{GH zy(ω)} + Im{αm,⊥(ω−)} −Re{αm,g(ω−)} Re{GH xz(ω)} −Re{GH zx(ω)}+Im{GH yz(ω)}+Im{GH zy(ω)}) −[n0(ω) + 1]( −2Reαm,∥(ω) Im{GH zx(ω)}+Im{GH xz(ω)} + 2Imαm,∥(ω) −Re{GH zx(ω)}+Re{GH xz(ω)}) . (G10) We can find the xandycomponents of frictional torque by plugging magnetic Green’s function expressions into Eqs. (G9) and (G10). Remarkably, we find that the spinning YIG nanosphere can experience a large torque along the x or y direction when the YIG interface is biased by external magnetic fields in the x or y direction. This means that in these cases, the sphere can rotate out of the rotation axis and start to precess. This will change the validity of the equations found for the vacuum radiation and frictional torque along the zaxis since it has been assumed that the sphere is always rotating around the zaxis and is also magnetized along that axis. However, this torque is still small enough compared to the driving torque of the trapping laser and it will still give enough time to make the observations of vacuum fluctuation effects. In Section S2 in the supplementary material, we present the plots of these torques when the interface is the x−yorx−zplane and provide more detailed discussions.15 2. Recoil torque Another contribution to the torque comes from the case when the induced dipole moments on the YIG sphere re-radiate due to the vacuum electric field fluctuations. This causes a recoil torque on the sphere and can be written as Mrec=⟨mind×Hsc⟩, (G11) where Hscis the scattered fields from the dipole and are given by, Hsc(r, ω) =¯GH(r,r′, ω)·mind(r′, ω), (G12) which shows that this term is of higher order contribution. We find that this recoil torque is much smaller than the torque derived in Eq. (G5) for YIG spheres spinning near YIG or Al interfaces and can thus be ignored in all studied cases. We provide detailed derivations of Mrecand quantitative comparisons in Section S2 in the supplementary material. Appendix H: Experimental Analysis In this section, we present the analytical steps for finding the experimental prediction plots provided in the last section of the main text. 1. Effects of drag torque due to imperfect vacuum In the real system of a spinning sphere, the environment is not a pure vacuum. This causes an extra torque on the spinning sphere from air molecules in the imperfect vacuum. The steady-state spin of the sphere happens when the driving torque of the trapping laser is equal to the drag and vacuum friction torques. In the case when there is no interface present, the only important counteracting torque is the drag torque given by [47] Mdrag=2πµa4 1.497λΩ, (H1) where ais the sphere radius, µis the viscosity of the gas the sphere is spinning in, λis the mean free path of the air molecules, and Ωis the rotation frequency. We further have for gases [48], λ=µ pgasr πKBT 2m, (H2) where pgasandmare the pressure and the molecular mass of the gas, respectively. Thus, we get the drag torque, Mdrag=2a4pgas 1.479r 2πm kBTΩ. (H3) For 1 GHz rotation of a sphere, the balance between the drag torque and the optical torque Mopthappens at about pgas= 10−4torr. Therefore we get, at room temperature and for a molecular mass of 28.966gram /mol, r 2πm KBT= 8.542×10−3, (H4) and thus [3], Mopt= 1.568×10−21N·m, (H5) This is important for studying the effects of vacuum torque on the rotation speed of the sphere. As shown in the main text, we find that for vacuum pressures of about 10−4torr, changes in the balance speed of the YIG nanoparticle when it is closer to material interfaces are detectible in the power spectral density (PSD) of the nanosphere [3].16 2. Effects of negative torque and shot noise heating due to the trapping laser When the trapping laser is linearly polarized, it can exert a negative torque on the spinning particle. The torque on the sphere due to the laser is given by Mopt=1 2Re{p∗×E}[3], where pis the dipole moment of the sphere, given by p=¯αeff·E, with ¯αeffbeing the effective polarizability of the sphere as seen in the frame of the lab, and E is the electric field from the laser. As derived in Section S3 in the supplementary material, in the case when the laser is linearly polarized, the negative torque from the laser is proportional to Im {α(ω0+ Ω)} −Im{α(ω0−Ω)}, where ω0= 1.21×1016is the frequency of the laser, and Ω = 6 .28×109is the rotation frequency. Since Ω≪ω0, we get α(ω+)≃α(ω−)and thus the second term is negligible. We can thus ignore the negative torque coming from the laser when the laser is linearly polarized. Another effect from the trapping laser is the heating of nanoparticles due to the shot noise. The rate of temperature change due to shot noise heating can be determined by the laser frequency, the power of the laser per unit area, the mass of the particle, and the scattering cross section for the nanoparticles [39]. For YIG nanospheres of density 5110kg/m3and radius 200nm, and trapping laser of 1550nm wavelength and of 500mW power focused on an area of radius 0.7566µm, we find that the temperature change due to shot noise is small compared to the time scale of the rotation, which is 1ns. Therefore, the thermodynamic equilibrium condition for the FDT is valid. We provide further details for the derivations and calculations of negative torque and shot noise heating due to the trapping laser in Section S3 in the supplementary material. [1] A.Kani, F.Quijandría,andJ.Twamley,PhysicalReview Letters129, 257201 (2022). [2] R. Reimann, M. Doderer, E. Hebestreit, R. Diehl, M. Frimmer, D. Windey, F. Tebbenjohanns, and L. Novotny, Phys. Rev. Lett. 121, 033602 (2018). [3] J. Ahn, Z. Xu, J. Bang, Y.-H. Deng, T. M. Hoang, Q. Han, R.-M. Ma, and T. Li, Phys. Rev. Lett. 121, 033603 (2018). [4] J. Piotrowski, D. Windey, J. Vijayan, C. Gonzalez- Ballestero, A.delosRíosSommer, N.Meyer, R.Quidant, O. Romero-Isart, R. Reimann, and L. Novotny, Nature Physics19, 1009 (2023). [5] C. Gonzalez-Ballestero, M. Aspelmeyer, L. Novotny, R. Quidant, and O. Romero-Isart, Science 374, eabg3027 (2021). [6] Y. Jin, J. Yan, S. J. Rahman, J. Li, X. Yu, and J. Zhang, Photonics Research 9, 1344 (2021). [7] M. Fuwa, R. Sakagami, and T. Tamegai, Phys. Rev. A 108, 063511 (2023). [8] P. Ju, Y. Jin, K. Shen, Y. Duan, Z. Xu, X. Gao, X. Ni, and T. Li, Nano Letters 23, 10157 (2023). [9] C. Marletto and V. Vedral, Phys. Rev. Lett. 119, 240402 (2017). [10] A. D. Rider, D. C. Moore, C. P. Blakemore, M. Louis, M. Lu, and G. Gratta, Phys. Rev. Lett. 117, 101101 (2016). [11] J. D. Bekenstein and M. Schiffer, Phys. Rev. D 58, 064014 (1998). [12] J. R. Deop-Ruano and A. Manjavacas, Phys. Rev. Lett. 130, 133605 (2023). [13] Y. Guo and S. Fan, ACS Photonics 8, 1623 (2021). [14] D. Pan, H. Xu, and F. J. G. de Abajo, Physical Review Letters123, 066803 (2019). [15] Z. Xu, Z. Jacob, and T. Li, Nanophotonics 10, 537 (2020). [16] D. Hümmer, R. Lampert, K. Kustura, P. Maurer, C. Gonzalez-Ballestero, and O. Romero-Isart, Physical Review B 101, 205416 (2020). [17] D. Pan, H. Xu, and F. J. G. de Abajo, Physical ReviewA99, 062509 (2019). [18] A. Manjavacas and F. J. García de Abajo, Phys. Rev. Lett.105, 113601 (2010). [19] M. F. Maghrebi, R. L. Jaffe, and M. Kardar, Physical review letters 108, 230403 (2012). [20] R. S. Decca, D. López, E. Fischbach, and D. E. Krause, Phys. Rev. Lett. 91, 050402 (2003). [21] U. Mohideen and A. Roy, Phys. Rev. Lett. 81, 4549 (1998). [22] J. N. Munday, F. Capasso, and V. A. Parsegian, Nature 457, 170 (2009). [23] Z. Xu and T. Li, Phys. Rev. A 96, 033843 (2017). [24] S. Pendharker, Y. Guo, F. Khosravi, and Z. Jacob, Phys- ical Review A 95, 033817 (2017). [25] Y. Guo and Z. Jacob, Journal of Optics 16, 114023 (2014). [26] Y. Guo and Z. Jacob, Optics express 22, 26193 (2014). [27] R. Hisatomi, A. Osada, Y. Tabuchi, T. Ishikawa, A. Noguchi, R. Yamazaki, K. Usami, and Y. Nakamura, Physical Review B 93, 174427 (2016). [28] X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, Physical review letters 113, 156401 (2014). [29] K. Joulain, R. Carminati, J.-P. Mulet, and J.-J. Greffet, Phys. Rev. B 68, 245405 (2003). [30] A. Manjavacas and F. J. García de Abajo, Phys. Rev. A 82, 063827 (2010). [31] R. Zhao, A. Manjavacas, F. J. García de Abajo, and J. B. Pendry, Phys. Rev. Lett. 109, 123604 (2012). [32] G.FordandW.Weber,PhysicsReports 113,195 (1984). [33] A. Volokitin and B. N. Persson, Reviews of Modern Physics79, 1291 (2007). [34] J. Sloan, N. Rivera, J. D. Joannopoulos, and M. Soljačić, Physical Review Letters 127, 053603 (2021). [35] S. J. Barnett, Phys. Rev. 6, 239 (1915). [36] D. M. Pozar, Microwave Engineering 3e (Wiley, 2006). [37] M. Nieto-Vesperinas, Opt. Lett. 40, 3021 (2015). [38] T. Seberson, P. Ju, J. Ahn, J. Bang, T. Li, and F. Ro- bicheaux, J. Opt. Soc. Am. B 37, 3714 (2020). [39] T. Seberson and F. Robicheaux, Phys. Rev. A 102,17 033505 (2020). [40] J. Sloan, N. Rivera, J. D. Joannopoulos, I. Kaminer, and M. Soljačić, Phys. Rev. B 100, 235453 (2019). [41] H. J. Mamin, M. Kim, M. H. Sherwood, C. T. Rettner, K. Ohno, D. D. Awschalom, and D. Rugar, Science 339, 557 (2013). [42] L. D. Landau and E. M. Lifshitz, Statistical Physics: Vol- ume 5, Vol. 5 (Elsevier, 2013). [43] C. Khandekar and Z. Jacob, New Journal of Physics 21, 103030 (2019). [44] A. H. Sihvola, Opt. Lett. 19, 430 (1994). [45] C. F. Bohren and D. R. Huffman, Absorption and scat- tering of light by small particles (John Wiley & Sons, 2008). [46] G. W. Ford and W. H. Weber, Physics Reports 113, 195 (1984). [47] J. Corson, G. W. Mulholland, and M. R. Zachariah, Physical Review E 96, 013110 (2017). [48] C. H. Kruger and W. Vincenti, John Wiley & Sons , 414 (1965).SUPPLEMENTAL MATERIAL FOR ‘GIANT ENHANCEMENT OF VACUUM FRICTION IN SPINNING YIG N ANOSPHERES ’ Farhad Khosravi1,2, Wenbo Sun2, Chinmay Khandekar2, Tongcang Li2,3, and Zubin Jacob2,∗ 1Department of Electrical and Computer Engineering, University of Alberta, Edmonton, Alberta T6G 1H9, Canada 2Elmore Family School of Electrical and Computer Engineering, Birck Nanotechnology Center, Purdue University, West Lafayette, Indiana 47907, USA 3Department of Physics and Astronomy, Purdue Quantum Science and Engineering Institute, Purdue University, West Lafayette, Indiana 47907, USA ∗zjacob@purdue.edu Contents S1 Non-Electrostatic Limit and Magnetic Polarizability due to Electric Fluctuations 1 S2 Vacuum Frictional Torque 4 S2.1 Other components of torque . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 S2.2 Recoil torque . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6 S2.3 Plots of torque terms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 S3 Experimental Considerations 8 S3.1 Effect of torque due to the trapping laser . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8 S3.2 Effect of heating due to the shot noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 S1 Non-Electrostatic Limit and Magnetic Polarizability due to Electric Fluctuations In this section, we provide derivations for the magnetic polarizability of metallic nanoparticles due to the electric dipole terms based on Mie theory. If a sphere is placed in the direction of a plane wave polarized along ˆxdirection and propagating along zdirection Ei=E0eik0rcosθˆx, (S1) The scattered fields are given by [2], Es=−∞/summationdisplay n=1En/parenleftig ianN(1) e1n−bnM(1) o1n/parenrightig , (S2) Hs=−k0 ωµ∞/summationdisplay n=1En/parenleftig ibnN(1) o1m+anM(1) e1n/parenrightig , (S3) where Memn=−m sinθsinmϕPm n(cosθ)zn(kr)ˆθ−cosmϕdPm n(cosθ) dθzn(kr)ˆϕ, (S4a) Momn=m sinθcosmϕPm n(cosθ)zn(kr)ˆθ−sinmϕdPm n(cosθ) dθzn(kr)ˆϕ, (S4b) Nemn=zn(kr) krcosmϕn(n+ 1)Pm n(cosθ)ˆr+ cos mϕdPm n(cosθ) dθ1 krd d(kr)[krzn(kr)]ˆθ −msinmϕPm n(cosθ) sinθ1 krd d(kr)[krzn(kr)]ˆϕ,(S4c)Nomn=zn(kr) krsinmϕn(n+ 1)Pm n(cosθ)ˆr+ sin mϕdPm n(cosθ) dθ1 krd d(kr)[krzn(kr)]ˆθ +mcosmϕPm n(cosθ) sinθ1 krd d(kr)[krzn(kr)]ˆϕ,(S4d) the superscripts (1)forMandNindicate that the Bessel functions are the Hankel functions of the first kind h(1)(kr), En=inE0(2n+ 1)/n(n+ 1) , and anandbnare the Mie scattering coefficients. On the other hand, the radiated fields due to an electric dipole are given by Ed=k3 0 4πϵm/braceleftbigg (ˆr×p)׈reikr kr+ [3ˆr(ˆr·p)−p]/parenleftbigg1 (kr)3−i (kr)2/parenrightbigg eikr/bracerightbigg , (S5a) Hd=ck2 0 4π(ˆr×p)eikr r/parenleftbigg 1−1 ikr/parenrightbigg . (S5b) Using the facts that P1 1(cosθ) =−sinθ,dP1 1(cosθ) dθ=−cosθ, (S6) h(1) 1(kr) =−eikr/parenleftbiggi (kr)2+1 kr/parenrightbigg ,1 krd d(kr)/bracketleftig krh(1) 1(kr)/bracketrightig =−eikr/parenleftbigg −i (kr)3−1 (kr)2+i kr/parenrightbigg . (S7) The scattered fields to the first order of nbecome Es=3 2E0/braceleftigg ia1/bracketleftbigg eikr/parenleftbigg1 (kr)3−i (kr)2/parenrightbigg 2 cosϕsinθˆr−(cosϕcosθˆθ−sinϕˆϕ)eikr/parenleftbigg1 (kr)3−i (kr)2−1 kr/parenrightbigg/bracketrightbigg −b1/bracketleftbigg (cosϕˆθ−sinϕcosθˆϕ)eikr/parenleftbigg−1 (kr)2+i kr/parenrightbigg/bracketrightbigg/bracerightigg . (S8) Assuming that the dipole is along xdirection p=p0ˆx, the dipole fields become Ed=p0k3 0 4πϵm/braceleftbigg (cosθcosϕˆθ−sinϕˆϕ)eikr kr+ (2ˆrsinθcosϕ−ˆθcosθcosϕ+ˆϕsinϕ)/parenleftbigg1 (kr)3−i (kr)2/parenrightbigg eikr/bracerightbigg =p0k3 0 4πϵm/braceleftbigg 2ˆrsinθcosϕ/parenleftbigg1 (kr)3−i (kr)2/parenrightbigg eikr−(ˆθcosθcosϕ−ˆϕsinϕ)/parenleftbigg1 (kr)3−i (kr)2−1 kr/parenrightbigg eikr/bracerightbigg . (S9) In the low-frequency limit when kr=2πr λ≪1, the scattered fields are dominated by terms of the order (kr)−3. Thus, we can neglect the contribution from the Mterms or the b1terms in Eq. (S8). In this limit, the fields of the dipole and the scattered fields become equivalent, if we take p0=6πϵmia1 k3 0E0, (S10) or in other words, the sphere takes the polarizability αe=6πϵmc3 ω3ia1, (S11) where an=ϵ1jn(x1)[x0jn(x0)]′−ϵ0jn(x0)[x1jn(x1)]′ ϵ1jn(x1)[x0h(1) n(x0)]′−ϵ0h(1) n(x0)[x1jn(x1)]′, (S12) withx0=k0a,x1=k1a, and k1=ω√µ1ϵ1, and µ1andϵ1being properties of the sphere. Now, we look at the scattered magnetic fields. We have to the first order Hs=3 2k0 ωµ0E0/braceleftigg ib1/bracketleftbigg 2ˆrsinϕsinθ/parenleftbigg1 (kr)3−i (kr)2/parenrightbigg eikr−(ˆθsinϕcosθ+ˆϕcosϕ)/parenleftbigg1 (kr)3−i (kr)2−i kr/parenrightbigg eikr/bracketrightbigg −an/bracketleftbigg (ˆϕcosθcosϕ+ˆθsinϕ)/parenleftbiggi (kr)2+1 kr/parenrightbigg eikr/bracketrightbigg/bracerightigg . (S13) 2Again, we can ignore the second line or, in other words, anin this expression for low frequencies. Then, comparing this expression with the magnetic fields of a magnetic dipole polarized along ˆydirection m=m0ˆy, Hm=m0k3 0 4π/braceleftbigg 2ˆrsinθsinϕ/parenleftbigg1 (kr)3−i (kr)2/parenrightbigg eikr−(ˆθcosθsinϕ+ˆϕcosϕ)/parenleftbigg1 (kr)3−i (kr)2−1 kr/parenrightbigg eikr/bracerightbigg , (S14) Taking H0=k0 ωµ0E0, we find that the two are equivalent if we have m0=6πib1 k3 0H0, (S15) or if the sphere takes the magnetic polarizability αm=6πc3 ω3ib1, (S16) where bn=µ1jn(x1)[x0jn(x0)]′−µ0jn(x0)[x1jn(x1)]′ µ1jn(x1)[x0h(1) n(x0)]′−µ0h(1) n(x0)[x1jn(x1)]′. (S17) In the low-frequency limit, we have lim x→0jn(x) =2nn! (2n+ 1)!xn, (S18) and lim x→0yn(x) =−(2n)! 2nn!1 xn+1. (S19) Therefore, we have in this limit j1(x)≃x/3,y1(x)≃ −1/x2,[xj1(x)]′≃2x/3and[xy1(x)]′≃1/x2which gives a1≃ϵ1x1 32x0 3−ϵ0x0 32x1 3 ϵ1x1 3/parenleftig 2x0 3+i x2 0/parenrightig −ϵ02x1 3/parenleftig x0 3−i x2 0/parenrightig≃2k3 0a3 3iϵ1−ϵ0 ϵ1+ 2ϵ0, (S20a) b1≃2k3 0a3 3iµ1−µ0 µ1+ 2µ0. (S20b) We thus get for the polarizabilities αe≃4πϵ0a3ϵ1−ϵ0 ϵ1+ 2ϵ0, α m≃4πa3µ1−µ0 µ1+ 2µ0, (S21) which are exactly equal to the results derived using the electro-static and magneto-static approximations method. For a non-magnetic material, b1becomes b1≃x3 0 45ix2 0/parenleftbiggϵ1 ϵ0−1/parenrightbigg , (S22) which gives for the magnetic polarizability, αm≃2π 15k2 0a5/parenleftbiggϵ1 ϵ0−1/parenrightbigg =8π3 15a3/parenleftiga λ/parenrightig2/parenleftbiggϵ1 ϵ0−1/parenrightbigg . (S23) 3S2 Vacuum Frictional Torque S2.1 Other components of torque In this section, we provide further discussions of components of the torque other than the zcomponent exerted on a spinning nanosphere near YIG slabs under different bias fields. The xcomponent of torque, Mx=ℏ 4π/integraldisplay∞ −∞dω/braceleftigg /bracketleftbig 2n1(ω−) + 1/bracketrightbig/bracketleftbig Im/braceleftbig αm,⊥(ω−)/bracerightbig −Re/braceleftbig αm,g(ω−)/bracerightbig/bracketrightbig/bracketleftbig 2Im/braceleftbig GH zx(ω)/bracerightbig + 2Re/braceleftbig GH zy(ω)/bracerightbig/bracketrightbig −4 [n1(ω) + 1] Im/braceleftbig αm,∥(ω)/bracerightbig Re/braceleftbig GH yz(ω)/bracerightbig + [2n0(ω) + 1]/braceleftigg /bracketleftbig Re{αm,⊥(ω−)}+Im{αm,g(ω−)}/bracketrightbig/parenleftbig Re{GH xz(ω)} −Re{GH zx(ω)}+Im{GH yz(ω)}+Im{GH zy(ω)}/parenrightbig +/bracketleftbig Im{αm,⊥(ω−)} −Re{αm,g(ω−)}/bracketrightbig/parenleftbig −Im{GH xz(ω)} −Im{GH zx(ω)}+Re{GH yz(ω)} −Re{GH zy(ω)}/parenrightbig/bracerightigg + [n0(ω) + 1]/braceleftigg −2Re{αm,∥(ω)}/parenleftbig Im{GH zy(ω)}+Im{GH yz(ω)}/parenrightbig + 2Im{αm,∥(ω)}/parenleftbig −Re{GH zy(ω)}+Re{GH yz(ω)}/parenrightbig/bracerightigg , (S24) and for the ycomponent, My=ℏ 4π/integraldisplay∞ −∞dω/braceleftigg /bracketleftbig 2n1(ω−) + 1/bracketrightbig/bracketleftbig Im/braceleftbig αm,⊥(ω−)/bracerightbig −Re/braceleftbig αm,g(ω−)/bracerightbig/bracketrightbig/bracketleftbig −2Re{GH zx(ω)}+ 2Im{GH zy(ω)}/bracketrightbig +4 [n1(ω) + 1] Im/braceleftbig αm,∥(ω)/bracerightbig Re{GH xz(ω)}/bracerightigg −[2n0(ω) + 1]/braceleftigg /bracketleftbig Re{αm,⊥(ω−)}+Im{αm,g(ω−)}/bracketrightbig/parenleftbig Im{GH xz(ω)}+Im{GH zx(ω)} −Re{GH yz(ω)}+Re{GH zy(ω)}/parenrightbig +/bracketleftbig Im{αm,⊥(ω−)} −Re{αm,g(ω−)}/bracketrightbig/parenleftbig Re{GH xz(ω)} −Re{GH zx(ω)}+Im{GH yz(ω)}+Im{GH zy(ω)}/parenrightbig/bracerightigg −[n0(ω) + 1]/braceleftigg −2Reαm,∥(ω)/parenleftbig Im{GH zx(ω)}+Im{GH xz(ω)}/parenrightbig + 2Imαm,∥(ω)/parenleftbig −Re{GH zx(ω)}+Re{GH xz(ω)}/parenrightbig/bracerightigg . (S25) In the case when the interface is in the x−yplane, we have Re/braceleftbig GH xz(ω)/bracerightbig = (πωρ 0/8)1 π/integraldisplay2π 0dϕ/braceleftigg/integraldisplay1 0κ2dκ p/parenleftbig Im/braceleftbig rsse2ik0pd/bracerightbig pcosϕ−Im/braceleftbig rpse2ik0pd/bracerightbig sinϕ/parenrightbig +/integraldisplay∞ 1κ2dκ |p|e−2k0|p|d(Im{rss}|p|cosϕ+Re{rps}sinϕ)/bracerightigg ,(S26a) Re/braceleftbig GH zx(ω)/bracerightbig = (πωρ 0/8)1 π/integraldisplay2π 0dϕ/braceleftigg/integraldisplay1 0κ2dκ p/parenleftbig −Im/braceleftbig rsse2ik0pd/bracerightbig pcosϕ−Im/braceleftbig rspe2ik0pd/bracerightbig sinϕ/parenrightbig +/integraldisplay∞ 1κ2dκ |p|e−2k0|p|d(−Im{rss}|p|cosϕ+Re{rsp}sinϕ)/bracerightigg ,(S26b) Im/braceleftbig GH xz(ω)/bracerightbig = (πωρ 0/8)1 π/integraldisplay2π 0dϕ/braceleftigg/integraldisplay1 0κ2dκ p/parenleftbig −Re/braceleftbig rsse2ik0pd/bracerightbig pcosϕ+Re/braceleftbig rpse2ik0pd/bracerightbig sinϕ/parenrightbig +/integraldisplay∞ 1κ2dκ |p|e−2k0|p|d(−Re{rss}|p|cosϕ+Im{rps}sinϕ)/bracerightigg ,(S26c) 4Im/braceleftbig GH zx(ω)/bracerightbig = (πωρ 0/8)1 π/integraldisplay2π 0dϕ/braceleftigg/integraldisplay1 0κ2dκ p/parenleftbig Re/braceleftbig rsse2ik0pd/bracerightbig pcosϕ+Re/braceleftbig rspe2ik0pd/bracerightbig sinϕ/parenrightbig +/integraldisplay∞ 1κ2dκ |p|e−2k0|p|d(Re{rss}|p|cosϕ+Im{rsp}sinϕ)/bracerightigg ,(S26d) and Re/braceleftbig GH zy(ω)/bracerightbig = (πωρ 0/8)1 π/integraldisplay2π 0dϕ/braceleftigg/integraldisplay1 0κ2dκ p/parenleftbig −Im/braceleftbig rsse2ik0pd/bracerightbig psinϕ+Im/braceleftbig rspe2ik0pd/bracerightbig cosϕ/parenrightbig +/integraldisplay∞ 1κ2dκ |p|e−2k0|p|d(−Im{rss}|p|sinϕ−Re{rsp}cosϕ)/bracerightigg ,(S27a) Re/braceleftbig GH yz(ω)/bracerightbig = (πωρ 0/8)1 π/integraldisplay2π 0dϕ/braceleftigg/integraldisplay1 0κ2dκ p/parenleftbig Im/braceleftbig rsse2ik0pd/bracerightbig psinϕ+Im/braceleftbig rpse2ik0pd/bracerightbig cosϕ/parenrightbig +/integraldisplay∞ 1κ2dκ |p|e−2k0|p|d(Im{rss}|p|sinϕ−Re{rps}cosϕ)/bracerightigg ,(S27b) Im/braceleftbig GH zy(ω)/bracerightbig = (πωρ 0/8)1 π/integraldisplay2π 0dϕ/braceleftigg/integraldisplay1 0κ2dκ p/parenleftbig Re/braceleftbig rsse2ik0pd/bracerightbig psinϕ−Re/braceleftbig rspe2ik0pd/bracerightbig cosϕ/parenrightbig +/integraldisplay∞ 1κ2dκ |p|e−2k0|p|d(Re{rss}|p|sinϕ−Im{rsp}cosϕ)/bracerightigg ,(S27c) Im/braceleftbig GH yz(ω)/bracerightbig = (πωρ 0/8)1 π/integraldisplay2π 0dϕ/braceleftigg/integraldisplay1 0κ2dκ p/parenleftbig −Re/braceleftbig rsse2ik0pd/bracerightbig psinϕ−Re/braceleftbig rpse2ik0pd/bracerightbig cosϕ/parenrightbig +/integraldisplay∞ 1κ2dκ |p|e−2k0|p|d(−Re{rss}|p|sinϕ−Im{rps}cosϕ)/bracerightigg .(S27d) And for the case when it is in the x−zplane Re/braceleftbig GH xz(ω)/bracerightbig = (πωρ 0/8)1 π/integraldisplay2π 0dϕ/braceleftigg/integraldisplay1 0κdκ 2p/bracketleftig sin 2ϕIm/braceleftbig rppe2ik0pd/bracerightbig +p2sin 2ϕIm/braceleftbig rsse2ik0pd/bracerightbig +2pIm/braceleftbig rpse2ik0pd/bracerightbig cos2ϕ+ 2pIm/braceleftbig rspe2ik0pd/bracerightbig sin2ϕ/bracketrightig +/integraldisplay∞ 1κdκ 2|p|/bracketleftig −sin 2ϕ/parenleftig 1 +Re{rpp}e−2k0|p|d/parenrightig −p2sin 2ϕRe{rss}e−2k0|p|d +2Im{rps}|p|e−2k0|p|dcos2ϕ+ 2Im{rspe−2k0|p|d}sin2ϕ/bracketrightig/bracerightigg , (S28a) Re/braceleftbig GH zx(ω)/bracerightbig = (πωρ 0/8)1 π/integraldisplay2π 0dϕ/braceleftigg/integraldisplay1 0κdκ 2p/bracketleftig sin 2ϕIm/braceleftbig rppe2ik0pd/bracerightbig +p2sin 2ϕIm/braceleftbig rsse2ik0pd/bracerightbig −2pIm/braceleftbig rpse2ik0pd/bracerightbig sin2ϕ−2pIm/braceleftbig rspe2ik0pd/bracerightbig cos2ϕ/bracketrightig +/integraldisplay∞ 1κdκ 2|p|/bracketleftig −sin 2ϕ/parenleftig 1 +Re{rpp}e−2k0|p|d/parenrightig −p2sin 2ϕRe{rss}e−2k0|p|d −2Im{rps}|p|e−2k0|p|dsin2ϕ−2Im{rsp}|p|e−2k0|p|dcos2ϕ/bracketrightig/bracerightigg , (S28b) 5Im/braceleftbig GH xz(ω)/bracerightbig = (πωρ 0/8)1 π/integraldisplay2π 0dϕ/braceleftigg/integraldisplay1 0κdκ 2p/bracketleftig −sin 2ϕ/parenleftbig 1 +Re/braceleftbig rppe2ik0pd/bracerightbig/parenrightbig −p2sin 2ϕRe/braceleftbig rsse2ik0pd/bracerightbig −2pRe/braceleftbig rpse2ik0pd/bracerightbig cos2ϕ−2pRe/braceleftbig rspe2ik0pd/bracerightbig sin2ϕ/bracketrightig +/integraldisplay∞ 1κdκ 2|p|e−2k0|p|d/bracketleftig −sin 2ϕIm{rpp} −p2sin 2ϕIm{rss} −2Re{rps}|p|cos2ϕ−2Re{rsp}|p|sin2ϕ/bracketrightig/bracerightigg , (S28c) Im/braceleftbig GH zx(ω)/bracerightbig = (πωρ 0/8)1 π/integraldisplay2π 0dϕ/braceleftigg/integraldisplay1 0κdκ 2p/bracketleftig −sin 2ϕ/parenleftbig 1 +Re/braceleftbig rppe2ik0pd/bracerightbig/parenrightbig −p2sin 2ϕRe/braceleftbig rsse2ik0pd/bracerightbig +2pRe/braceleftbig rpse2ik0pd/bracerightbig sin2ϕ+ 2pRe/braceleftbig rspe2ik0pd/bracerightbig cos2ϕ/bracketrightig +/integraldisplay∞ 1κdκ 2|p|e−2k0|p|d/bracketleftig −sin 2ϕIm{rpp} −p2sin 2ϕIm{rss} +2Re{rps}|p|sin2ϕ+ 2Re{rsp}|p|cos2ϕ/bracketrightig/bracerightigg , (S28d) and the expressions for the real and imaginary parts of GH zyandGH yzare the same as the ones for GH xzandGH zx, respectively, for when the interface is in the x−yplane as given in Eq. (S26). We can find the xandycomponents of torque by plugging these expressions into Eqs. (S24) and (S25) for the two cases when the interface is the x−yor x−zplane. We present the plots of these torques at the end of this section. S2.2 Recoil torque There is also another contribution to the torque from the case when the induced dipole moments on the YIG sphere re-radiate due to the vacuum electric field fluctuations. This causes a recoil torque on the sphere and can be written as Mrec=⟨mind×Hsc⟩, (S29) where Hscis the scattered fields from the dipole and are given by, Hsc(r, ω) =¯GH(r,r′, ω)·mind(r′, ω), (S30) which shows that this term is of higher order contribution and is thus smaller than the torque discussed in the main text. Repeating a similar procedure used before and plugging in all of the induced terms and writing them in terms of the fluctuations, we find after some algebra, Mrec z=ℏ π/integraldisplay∞ −∞dω[n0(ω)+1] /braceleftigg Im{Gxx}/bracketleftbig Re{Gyx}αeff ⊥⊥−Re{Gxy}αeff gg+Re{α⊥g}Re{Gyy−Gxx}+Im{α⊥g}Im{Gyy+Gxx}/bracketrightbig +Im{Gyy}/bracketleftbig Re{Gyx}αeff gg−Re{Gxy}αeff ⊥⊥+Re{α⊥g}Re{Gxx−Gyy}+Im{α⊥g}Im{Gyy+Gxx}/bracketrightbig +Re{Gyx−Gxy}/bracketleftbigg Re{Gyx−Gxy}Im{α⊥g}+1 2Im{Gyy+Gxx}(α⊥⊥+αgg)/bracketrightbigg +Im{Gyx+Gxy}/bracketleftbigg −Re{Gyx+Gxy}Re{α⊥g}+1 2Re{Gxx−Gyy}(αgg−α⊥⊥)/bracketrightbigg +1 2Im/braceleftig αeff∗ m,∥/bracketleftbig (Gxz−G∗ zx)/parenleftbig G∗ yzαeff m,⊥−G∗ xzαeff m,g/parenrightbig −/parenleftbig Gyz−G∗ zy/parenrightbig/parenleftbig G∗ yzαeff m,g+G∗ xzαeff m,⊥/parenrightbig/bracketrightbig/bracerightig/bracerightigg ,(S31) 6where we have defined αeff m,⊥⊥(ω) =αeff m,⊥(ω)αeff m,⊥(−ω), αeff m,gg(ω) =αeff m,g(ω)αeff m,g(−ω), αeff m,⊥g(ω) =αeff m,⊥(ω)αeff m,g(−ω), αeff m,g⊥(ω) =αeff m,⊥(−ω)αeff m,g(ω),(S32) and have used the facts that αeff m,⊥⊥(ω)andαeff m,gg(ω)are real, and αeff m,⊥g(ω) =/bracketleftig αeff m,g⊥(ω)/bracketrightig∗ . Note that we have dropped the frequency dependence as well as the H superscript of the Green’s function in Eq. (S31) for simplicity. For the special case when the substrate material is isotropic, the non-diagonal elements of the Green’s function become zero, and we get Mrec z=ℏ π/integraldisplay∞ −∞dω[n0(ω) + 1]/braceleftigg Im{Gxx−Gyy}Re{Gyy−Gxx}Re{α⊥g}+[Im{Gxx+Gyy}]2Im{α⊥g}/bracerightigg .(S33) Note that the expressions for the real and imaginary parts of GxzandGyzare given by Eqs. (S26),(S27), and (S28) for the two possible interface directions while the imaginary parts of GxxandGyyare given by equations in Appendix B. Also note that Re/braceleftbig GH yx/bracerightbig for when the interface is the x−yplane is the same as Re/braceleftbig GH xz/bracerightbig for when the interface is in thex−zplane given by Eq. (S28). Also Re/braceleftbig GH yx/bracerightbig for when the interface is the x−zplane is the same as Re/braceleftbig GH zy/bracerightbig for when the interface is in the x−yplane given by Eq. (S27). Thus, the only new term is Re {Gyy−Gxx}which is given by Re/braceleftbig GH yy(ω)−GH xx(ω)/bracerightbig =(πωρ 0/8)1 π/integraldisplay2π 0dϕ/braceleftigg/integraldisplay1 0κdκ p/bracketleftig −cos 2ϕIm/braceleftbig rppe2ik0pd/bracerightbig −cos 2ϕIm/braceleftbig rsse2ik0pd/bracerightbig +2psinϕcosϕIm/braceleftbig (rps−rsp)e2ik0pd/bracerightbig/bracketrightig +/integraldisplay∞ 1κdκ |p|/bracketleftig cos 2ϕ/parenleftig κ2+Re{rpp}e−2k0|p|d/parenrightig +p2cos 2ϕRe{rss}e−2k0|p|d +2|p|sinϕcosϕIm{rps−rsp}e−2k0|p|d/bracketrightig/bracerightigg , (S34) when the interface is the x−yplane, and Re/braceleftbig GH yy(ω)−GH xx(ω)/bracerightbig =(πωρ 0/8)1 π/integraldisplay2π 0dϕ/braceleftigg/integraldisplay1 0κdκ p/bracketleftig cos2ϕIm/braceleftbig rppe2ik0pd/bracerightbig −/parenleftbig κ2+p2sin2ϕ/parenrightbig Im/braceleftbig rsse2ik0pd/bracerightbig −psinϕcosϕIm/braceleftbig (rps−rps)e2ik0pd/bracerightbig/bracketrightig +/integraldisplay∞ 1κdκ |p|/bracketleftig −cos2ϕ/parenleftig 1 +Re{rpp}e−2k0|p|d/parenrightig −p2sin2ϕ+κ2+/parenleftbig κ2+p2sin2ϕ/parenrightbig Re{rss}e−2k0|p|d −|p|sinϕcosϕIm{rps−rsp}e−2k0|p|d/bracketrightig , (S35) when the interface is the x−zplane. S2.3 Plots of torque terms In this section, we present the components of torque derived in previous sections for YIG slabs with various bias magnetic fields and for the two cases when the slab is the x−yandx−zplanes. Figure S2 shows the plots of Mx,My,Mz, and Mrecderived in the previous sections for the magnetic and electric fluctuations. The expressions for the torques due to the electric fields and dipoles fluctuations are found by changing s topandptosinrss, rpp, rsp, andrps, in the expressions for the Green’s functions. Moreover, magnetic polarizability is replaced by a simple isotropic electric polarizability, assuming a simple dielectric polarizability scalar for the YIG and Al interfaces. The results are for three directions of the bias magnetic field for the YIG interface labeled as x−, y−, and z−bias. The meaning of these bias directions is demonstrated in Fig. S1 when the YIG slab is the x−yand x−zplanes. 7YIG YIG yz x x-bias(a) YIG YIG yz x y-bias (b) YIG YIG yz x z-bias (c) YIG YIG yz x x-bias (d) YIG YIG yz x y-bias (e) YIG YIG yz x z-bias (f) Figure S1: Schematics of different bias directions for the YIG interface for the two cases of the interface being the x−y(top row) and x−zplanes (bottom row). The green arrow shows the direction of the bias magnetic field applied to the slab of YIG. It is interesting to note that in Figs. (S2a), (S2e), and (S2g), the sphere can experience a large value of torque along xorydirections for the x−ory−biases. This means that in these cases, the sphere can rotate out of the rotation axis and start to precess. This will, of course, change the validity of the equations found for the vacuum radiation and frictional torque along the zaxis since it has been assumed that the sphere is always rotating around the zaxis and is also magnetized along that axis. This torque is still small enough compared to the driving torque of the trapping laser and it will still give enough time to make the observations. A more careful investigation of these components of torque is out of the scope of this study and will be explored in the future. Figures S2i-S2p show the axial torque Mzas well as the recoil torque Mrecfor all orientations of the bias magnetic field and YIG slab. As expected, the recoil torque is much smaller than Mzsince it is a second-order term. Figure S3 shows the results for MzandMrecfor the case when the Al interface is placed in the vicinity of the spinning sphere. Because Al is an isotropic material, MxandMyvanish for both orientations of the interface and thus are not included in the plots of the torques. Note that similar to the YIG interface results, Mrecis much smaller than the Mz for all cases of the Al interface. These results show that the recoil torque Mreccan be ignored in all studied cases. S3 Experimental Considerations In this section, we present details of the experimental analysis regarding negative torque and shot noise heating due to the trapping laser discussed in Appendix H. S3.1 Effect of torque due to the trapping laser When the trapping laser is linearly polarized, it can exert a negative torque on the spinning particle. The torque on the sphere due to the laser is given by Mopt=1 2Re{p∗×E}[1], where pis the dipole moment of the sphere, given by p=¯αeff·E, with ¯αeffbeing the effective polarizability of the sphere as seen in the frame of the lab, and Eis the electric field from the laser. As shown in Appendix A, the polarizability tensor of the sphere when it is spinning in the x−yplane is given by ¯αeff(ω) = αeff ⊥(ω)−αeff g(ω) 0 αeff g(ω)αeff ⊥(ω) 0 0 0 αeff ∥(ω) , (S36) where αeff ⊥(ω) =1 2/bracketleftbig α(ω+) +α(ω−)/bracketrightbig , α g(ω) =−i 2/bracketleftbig α(ω+)−α(ω−)/bracketrightbig , α∥(ω) =α(ω), (S37) withα(ω)being the electric polarizability of YIG at the laser frequency. Note that here, we have assumed that the polarizability of the YIG is scalar in the range of frequencies around 1550 nm. Plugging these into the equation for 80 2 4 6 8 10 Frequency (GHz)-15-10-5051015YIG Magnetic Torque Mx - Z bias My - Z bias Mx - X bias My - X bias Mx - Y bias My - Y bias(a) 0 2 4 6 8 10 Frequency (GHz)-0.500.511.522.510-4YIG Electric Torque Mx - Z bias My - Z bias Mx - X bias My - X bias Mx - Y bias My - Y bias (b) 0 2 4 6 8 10 Frequency (GHz)-6-4-2024610-6Al Magnetic Torque Mx - Z bias My - Z bias Mx - X bias My - X bias Mx - Y bias My - Y bias (c) 0 2 4 6 8 10 Frequency (GHz)-505101510-12Al Electric Torque Mx - Z bias My - Z bias Mx - X bias My - X bias Mx - Y bias My - Y bias (d) 0 2 4 6 8 10 Frequency (GHz)-2024681012YIG Magnetic Torque Mx - Z bias My - Z bias Mx - X bias My - X bias Mx - Y bias My - Y bias (e) 0 2 4 6 8 10 Frequency (GHz)-2.5-2-1.5-1-0.500.510-4YIG Electric Torque Mx - Z bias My - Z bias Mx - X bias My - X bias Mx - Y bias My - Y bias (f) 0 2 4 6 8 10 Frequency (GHz)02468Al Magnetic Torque Mx - Z bias My - Z bias Mx - X bias My - X bias Mx - Y bias My - Y bias (g) 0 2 4 6 8 10 Frequency (GHz)-15-10-50510-12Al Electric Torque Mx - Z bias My - Z bias Mx - X bias My - X bias Mx - Y bias My - Y bias (h) 0 2 4 6 8 10 Frequency (GHz)-12-10-8-6-4-20YIG Magnetic Torque Mz - Z bias Mrec - Z bias Mz - X bias Mrec - X bias Mz - Y bias Mrec - Y bias (i) 0 2 4 6 8 10 Frequency (GHz)-20-15-10-5010-4YIG Electric Torque Mz - Z bias Mrec - Z bias Mz - X bias Mrec - X bias Mz - Y bias Mrec - Y bias (j) 0 2 4 6 8 10 Frequency (GHz)-0.2500.05Al Magnetic Torque Mz - Z bias Mrec - Z bias Mz - X bias Mrec - X bias Mz - Y bias Mrec - Y bias (k) 0 2 4 6 8 10 Frequency (GHz)-3-2-101210-7Al Eelctric Torque Mz - Z bias Mrec - Z bias Mz - X bias Mrec - X bias Mz - Y bias Mrec - Y bias (l) 0 2 4 6 8 10 Frequency (GHz)-15-10-50YIG Magnetic Torque Mz - Z bias Mrec - Z bias Mz - X bias Mrec - X bias Mz - Y bias Mrec - Y bias (m) 0 2 4 6 8 10 Frequency (GHz)-20-15-10-5010-3 YIG Electric Torque Mz - Z bias Mrec - Z bias Mz - X bias Mrec - X bias Mz - Y bias Mrec - Y bias (n) 0 2 4 6 8 10 Frequency (GHz)-0.30Al Magnetic Torque Mz - Z bias Mrec - Z bias Mz - X bias Mrec - X bias Mz - Y bias Mrec - Y bias (o) 0 2 4 6 8 10 Frequency (GHz)-4-2024681010-5 Al Electric Torque Mz - Z bias Mrec - Z bias Mz - X bias Mrec - X bias Mz - Y bias Mrec - Y bias (p) Figure S2: Plots of MxandMy(first two rows) and MzandMrec(second two rows) in the vicinity of the YIG slab when the slab is the x−yplane (first and third rows), and when the slab is x−zplane (second and fourth rows). The plots show the results for various magnetic field directions. The meanings of x−,y−, andz−bias are demonstrated in Fig. S1 for the two orientations of the interface. the exerted torque, we find the zcomponent of the torque Mopt=1 2Re/braceleftbig αeff∗ ⊥(ω)E∗ xEy−αeff∗ g(ω)E∗ yEy−αeff∗ g(ω)E∗ xEx−αeff∗ ⊥(ω)E∗ yEx/bracerightbig =1 2/bracketleftbig Im{αeff ⊥(ω)}Im{E∗×E} −Re{αeff g(ω)}/parenleftbig |Ex|2+|Ey|2/parenrightbig/bracketrightbig =1 2/bracketleftbig Im{α(ω+) +α(ω−)}Im{E∗×E} −Im{α(ω+)−α(ω−)}/parenleftbig |Ex|2+|Ey|2/parenrightbig/bracketrightbig .(S38) 90 2 4 6 8 10 Frequency (GHz)-3.5-3-2.5-2-1.5-1-0.50YIG Magnetic Torque Mz - local Mrec - local Mz - nonlocal Mrec - nonlocal(a) 0 2 4 6 8 10 Frequency (GHz)-1.4-1.2-1-0.8-0.6-0.4-0.2010-10 YIG Electric Torque Mz - local Mrec - local Mz - nonlocal Mrec - nonlocal (b) 0 2 4 6 8 10 Frequency (GHz)-0.150Al Magnetic Torque Mz - local Mrec - local Mz - nonlocal Mrec - nonlocal (c) 0 2 4 6 8 10 Frequency (GHz)-6-5-4-3-2-10110-16Al Electric Torque Mz - local Mrec - local Mz - nonlocal Mrec - nonlocal (d) 0 2 4 6 8 10 Frequency (GHz)-5-4-3-2-10YIG Magnetic Torque Mz - local Mrec - local Mz - nonlocal Mrec - nonlocal (e) 0 2 4 6 8 10 Frequency (GHz)-20-15-10-5010-11 YIG Electric Torque Mz - local Mrec - local Mz - nonlocal Mrec - nonlocal (f) 0 2 4 6 8 10 Frequency (GHz)-0.2500.05Al Magnetic Torque Mz - local Mrec - local Mz -nonlocal Mrec - nonlocal (g) 0 2 4 6 8 10 Frequency (GHz)-8-6-4-2010-16 Al Electric Torque Mz - local Mrec - local Mz - nonlocal Mrec - nonlocal (h) Figure S3: Plots of MzandMrecin the vicinity of the YIG slab when the slab is the x−yplane (first row) and when the slab is x−zplane (second row). Note that due to the isotropy of Al, the other components of torque, including MxandMy, vanish. The first term is proportional to the spin of the electromagnetic field and causes a positive torque on the particle. This is the term for the transferring of angular momentum from the laser to the particle. The second term is negative and thus causes a negative torque on the sphere. In the case when the laser is linearly polarized, this negative term is proportional to Im {α(ω0+ Ω)} −Im{α(ω0−Ω)}where ω0= 1.21×1016is the frequency of the laser, and Ω = 6 .28×109is the rotation frequency. Since Ω≪ω0, we get α(ω+)≃α(ω−)and thus the second term is negligible. We can thus ignore the negative torque coming from the laser when the laser is linearly polarized. S3.2 Effect of heating due to the shot noise The particle can heat up due to the shot noise of the trapping laser [4]. In this section, we calculate the rate of temper- ature change due to the shot noise and vacuum radiation, respectively. The rate of energy change in the nanosphere due to the shot noise is [4], ˙ETR=ℏk MIL cσ, (S39) where ωL=ckis the laser frequency, ILis the power of the laser per unit area, Mis the mass of the particle, andσis the cross section of scattering where, which is equal to σ=/parenleftbig8π 3/parenrightbig/parenleftig αk2 4πϵ/parenrightig2 for Rayleigh particles with the polarizability α= 4πϵ0a3/parenleftig ϵ−1 ϵ+2/parenrightig . For the range of wavelengths around visible and infrared, the Rayleigh limit is valid for particles of radii asmaller than 50nm. Since the radius of the particle in our case is 200nm, this expression may not be valid. Mie scattering parameters should be used to evaluate the scattering cross section. Assuming the trapping laser wavelength of λ= 1550 nm and using the Mie theory, the rate of energy change of YIG with refractive index n= 2.21[3] is close to that of the diamond with n = 2.39 in the Rayleigh limit [4]. Therefore, we get the energy change rate in the sphere ˙ETR=2ℏω0 ρc2AP0a3k4/parenleftbiggn2−1 n2+ 2/parenrightbigg2 , (S40) where A=πR2 Lis the area of the beam where the laser with the power P0is focused on, and ρis the mass density which for YIG is ρ= 5110 kg/m3. For a laser power of 500mW focused on an area of radius 0.7566µm, we find ˙TL= 15.45K/s. (S41) This is a very small temperature change compared to the time scale of the rotation, which is 1ns. Therefore, the thermodynamic equilibrium condition for the FDT is still valid. This temperature change gets damped by the radiated 10power of the sphere due to the rotation. For a YIG sphere spinning at about 0.5µm from the aluminum interface, the rate of change due to vacuum radiation at the equilibrium temperature T0= 300 K is, ˙TR=−362.973K/s, (S42) which is much larger than the temperature rise due to the shot noise of the laser, and this shows that the sphere will cool down. Note that this energy heats the aluminum instead. In this derivation, we have not included the heating due to the noise in the aluminum or YIG interface. The value found in Eq. (S42) is much smaller at lower temperatures. References [1] J. Ahn, Z. Xu, J. Bang, Y .-H. Deng, T. M. Hoang, Q. Han, R.-M. Ma, and T. Li. Optically levitated nanodumbbell torsion balance and ghz nanomechanical rotor. Physical review letters , 121(3):033603, 2018. [2] C. F. Bohren and D. R. Huffman. Absorption and scattering of light by small particles . John Wiley & Sons, 2008. [3] T. Seberson, P. Ju, J. Ahn, J. Bang, T. Li, and F. Robicheaux. Simulation of sympathetic cooling an optically levitated magnetic nanoparticle via coupling to a cold atomic gas. J. Opt. Soc. Am. B , 37(12):3714–3720, Dec 2020. [4] T. Seberson and F. Robicheaux. Distribution of laser shot-noise energy delivered to a levitated nanoparticle. Phys. Rev. A , 102:033505, Sep 2020. 11 | 2024-01-17 | Experimental observations of vacuum radiation and vacuum frictional torque
are challenging due to their vanishingly small effects in practical systems.
For example, a rotating nanosphere in free space slows down due to friction
from vacuum fluctuations with a stopping time around the age of the universe.
Here, we show that a spinning yttrium iron garnet (YIG) nanosphere near
aluminum or YIG slabs exhibits vacuum radiation eight orders of magnitude
larger than other metallic or dielectric spinning nanospheres. We achieve this
giant enhancement by exploiting the large near-field magnetic local density of
states in YIG systems, which occurs in the low-frequency GHz regime comparable
to the rotation frequency. Furthermore, we propose a realistic experimental
setup for observing the effects of this large vacuum radiation and frictional
torque under experimentally accessible conditions. | Giant Enhancement of Vacuum Friction in Spinning YIG Nanospheres | 2401.09563v1 |
arXiv:2009.04162v1 [cond-mat.mes-hall] 9 Sep 2020Sub-pico-liter magneto-optical cavities J. A. Haigh,1,∗R. A. Chakalov,2and A. J. Ramsay1 1Hitachi Cambridge Laboratory, Cambridge, CB3 0HE, United K ingdom 2Cavendish Laboratory, University of Cambridge, Cambridge , CB3 0HE, United Kingdom (Dated: September 10, 2020) Microwave-to-optical conversion via ferromagnetic magno ns has so-far been limited by the optical coupling rates achieved in mm-scale whispering gallery mode resonat ors. Towards overcoming this limitation, we pro- pose and demonstrate an open magneto-optical cavity contai ning a thin-film of yttrium iron garnet (YIG). We achieve a 0.1 pL (100 µm3) optical mode volume, ∼50 times smaller than previous devices. From this, we estimate the magnon single-photon coupling rate is G≈50Hz. This open cavity design offers the prospect of wavelength scale mode volumes, small polarization splitti ngs, and good magneto-optical mode overlap. With achievable further improvements and optimization, efficie nt microwave-optical conversion and magnon cooling devices become a realistic possibility. I. INTRODUCTION Magnetic-field tunable ferromagnetic modes can be easily strongly coupled to microwave resonators [ 1–3]. Further cou- pling to optical photons offers the prospect of useful trans duc- tion of microwave quantum signals to telecoms optical wave- lengths [ 4]. For this reason, the interaction of magnons and optical photons has been explored recently in the whisperin g gallery modes (WGM) of yttrium iron garnet (YIG) spheres [5–7]. However, despite the high Q-factor of the magnetic and optical modes [ 8], the optomagnonic coupling rates achieved in mm-scale YIG spheres have been limited to ∼1 Hz. If the coupling rate can be increased significantly, in turn raisin g the conversion efficiency, this would open a wide range of tech- nological opportunities [ 9], as well as the ability to coherently modify the magnetization dynamics, for example cooling or dynamical driving the magnon mode [ 10,11]. The low coupling rate for optical whispering gallery modes is due to the poor mode overlap and the large volume of mag- netic material involved. To overcome the poor overlap, it ma y be possible to exploit magnon whispering gallery modes in YIG spheres, with almost ideal overlap with the optical WGM [12]. A simpler strategy is to explore more compact struc- tures, as very recently shown in rib waveguide devices [ 13]. In that case, the mm-long structure confines both the magnons and photons, yielding excellent overlap inside the structu re, enabling a coupling rate of 17 Hz. The estimated maximum coupling rate for a YIG optical resonator is ≈0.1MHz [ 14], based on a mode volume of the order of the resonant wavelength cubed λ3. There are sev- eral candidate wavelength-scale optical resonators [ 15], which could get close to this maximum coupling rate. The choice of resonator, however, must take into account the significan t challenges of micro-patterning YIG [ 16]. We note that, while sub-wavelength mode confinement is possible with plasmonic devices [ 17–19], this typically comes with high optical losses in the metal components. A simple optical resonator design, with wavelength scale mode volumes combined with large Q-factors, is an open mi- ∗jh877@cam.ac.ukcrocavity [20]. These are typically hemispherical resonators where a reflection-coated microlens is positioned in close proximity to a mirror surface. Devices can be fiber-based [21,22], or fabricated on planar surfaces [ 23], and have previ- ously been used to obtain large coupling rates to single atom s [24], N-V centers [ 25], single organic dye molecules [ 26], and excitons in 2D materials [ 27]. The advantage of this structure is that any transferable material can be easily embedded [ 28], and the modes are tunable by the position of the lens. Optical mode volumes as small as 1 fL (1 µm3) have been achieved, withQ-factors in excess of 10,000 [ 29]. In this article, we demonstrate a viable route to low mode volume, high coupling rate magneto-optical cavities with mode volumes limited by the optical wavelength. We embed single crystal YIG layers in an open microcavity, and show a two orders of magnitude increase in the coupling rate over whispering gallery mode devices. With further reduction in mode volume, it is expected that the strong-coupling limite d can be reached. This work, therefore, shows a path towards efficient microwave-optical conversion, and optical magno n cooling. II. COUPLING RATE We first briefly review the enhanced scattering process in cavity optomagnonics. Magnetic Brillouin light scatterin g is an inelastic process where a photon is scattered from an input mode ˆaiinto an output mode ˆao, with absorption or emission of a magnon in mode ˆm. Brillouin light scatter- ing is most efficient between orthogonally polarized opti- cal modes, as this compensates the angular momentum lost or gained to the magnon mode, conserving angular momen- tum. To enhance the BLS significantly, we require two op- tical resonances, enhancing both the input and output optic al fields. These should be orthogonally polarized, and with fre - quency separation matching the magnon frequency. This is thetriple resonance condition , which has been observed pre- viously in whispering gallery mode resonators [ 7], and in a recent waveguide device [ 13]. The interaction Hamiltonian that governs the scattering is of the formHint=G(ˆa† iˆaoˆm+ˆaiˆa† oˆm), with interaction strength2 quantified by the coupling rate, G=−iθfc n/radicalbigg 4gµB Ms/radicalbiggηmagηopt Vopt. (1) The numerical constant µBis the Bohr magneton and cthe speed of light. The factors affecting this rate can be separa ted in two parts. Firstly, the materials parameters of the embed - ded magnetic material: the Faraday coefficient θf, refractive indexn, gyromagnetic ratio gand saturation magnetization Ms. These parameters can be optimized by materials devel- opment, finding new materials and improving the quality of those available. Secondly, the geometry of the optical cavi ty affects the coupling rate through the volumes of the optical modesVi≈Vo=Vopt=/integraltext |ui,o(r)|2, whereui,o(r)is the mode function with normalization max(|ui,o(r)|2) = 1 . The overlap is contained in the fill-factors ηmag=Vint/Vmagand ηopt=Vint/Vopt, which are the proportion of the magnetic Vmagand optical Voptmodes volumes that contribute to the coupling through the triple-mode overlap, Vint=/integraldisplay drum(r)·[u∗ i(r)×uo(r)]. (2) Here, the mode function umis also normalized such that max(|um(r)|2) = 1 . This expression includes the effect of the mode polarization. To maximize the geometric factors we would like a low optical mode volume resonator, with excel- lent overlap with the magnon mode, and orthogonal polariza- tion of all three modes. In this paper, we focus exclusively on the minimization of the optical mode volume. The design is such that lateral pat- terning of the continuous YIG layer to confine the magnon mode can incorporated at a later stage. III. DESIGN A schematic of our proposed device is shown in Fig. 1(a). The mirrors forming the open microcavities are purchased from Oxford HiQ [ 30]. They consist of one planar surface and one microlens array. Both surfaces have a high qual- ity reflective coating consisting of 22 layers of SiO 2and Ta2O5, deposited by sputtering. This distributed Bragg reflec- tor (DBR) has a reflectivity of 99.8 %at its design wavelength of 1300 nm. The microlens array contains 16 lenses with four different radii of curvature, from 100 µm down to 20 µm, fab- ricated by focused-ion-beam milling [ 31]. The lens array is situated on a raised pedestal to allow alignment of the two mirrors. To embed a high quality, single crystal YIG layer in the mi- crocavity, we avoid deposition techniques such as pulsed la ser deposition and sputtering, because the non-lattice matche d DBR substrate would lead to poly-crystalline YIG growth [32]. Furthermore, post growth annealing to improve the crys- tallinity of those layers requires temperatures above 700◦C, which has been found to be detrimental to the DBR [ 33]. In- stead, we use a lift-off technique [ 34] to remove a layer of Au antennalens array crack50ma) YIG/GGG layer Au antenna DBRGGG (7 m) YIG (2m) BCB (~1 m) DBR sapphire substrate lens array DBRsYIG/GGG c) substrate DBR magnet material b)i) ii) iii)~1m Au antenna Figure 1. (a) Schematic of the magneto-optical cavity desig n. The open microcavity consists of two parts: a concave lens mille d into a substrate and coated with a DBR, and a planar mirror with magn etic layer. Left: Cross section of design to show relative dimens ions of the lens and beam waist. Right: 3D representation of the devi ce structure. (b)Fabrication of magneto-optical microcavit y. (i) Flat side of open microcavity. A gold microwave antenna is patter ned on the DBR surface before the YIG/GGG layer is bonded. (ii) Cros s section of flat mirror, showing two-layer BCB bonding polyme r. (iii) Cross-section section of open microcavity, showing microl ens array. (c) Optical image through cavity structure, showing lens ar ray and microwave antenna. single-crystal YIG from a lattice matched GGG growth sub- strate. We later bond this layer to the mirror surface with a spin-on polymer. The advantage of the open microcavity design is that the polarization splitting is minimized due to the cylindrical sym- metry. In principle, a minor asymmetry can tune the splittin g to match the magnon frequency, given the precise control of lens profile that has been demonstrated [ 31]. For an optical resonator with an asymmetrical cross-section, the splitti ng is typically a fixed fraction of the free spectral range. There- fore, as the mode volume shrinks, the frequency separation can become too large. This can be seen in the whispering gallery mode resonators, where a 1 mm diameter YIG sphere was chosen to match the splitting to the magnon frequency [7]. It is not possible to decrease the size of the sphere, be- cause the splitting would become too large. This effect can also be seen in the rib waveguide geometry [ 13], where the length of the cavity must be long ( ≈4mm) to keep the polar- ization splitting small.3 IV . FABRICATION We start with a YIG film of thickness 2 µm grown by liq- uid phase epitaxy on a gadolinium gallium garnet (GGG) sub- strate [ 35]. The lift-off is achieved by inducing spontaneous delamination as follows. The sample is first subjected to a high dose ion implantation 5×1016cm−2with He ions at 3.5 MeV [ 36]. The penetration depth is approximately 9 µm, with straggle ≈1µm, creating a narrow layer in which the lat- tice is substantially damaged [ 37]. Annealing at 470◦C for 1 min leads to delamination of a bilayer consisting of the 2 µm of YIG and around 7 µm of GGG . This delamination occurs due to the slight lattice mismatch and different coefficient of expansion of YIG and GGG [ 38]. The lattice mismatch leads to a membrane with ∼10-mm radius of curvature at room tem- perature. The YIG wafer is diced into 1 mm square chips post im- plantation, but prior to delamination. Post delamination, the thin membrane is manipulated using a small piece of 25 µm thick Kapton film, where it is held in place by static. To bond the YIG/GGG membrane to the mirror surface, we use BCB cyclotene [ 39], a polymer used as a dielectric in mi- croelectronics, adhesive wafer bonding [ 40] and planarization applications [ 41]. It has excellent optical properties [ 42], al- lowing its use, for example, in bonding active III-V devices to silicon photonic wafers [ 43]. Prior to bonding, a strip-line antenna is patterned on to the surface of the mirror using photo-lithography and lift-off as shown in Fig. 1(b). A titanium adhesion layer of 7 nm is de- posited under 100 nm of gold, with a final 7 nm of titanium above. This final layer is required to avoid the poor adhesion of BCB cyclotene to gold [ 44]. We prepare the mirror surface with solvent cleaning in an ultrasonic bath. The device is then soaked in DI water, be- fore a 2 min plasma cleaning process in a reactive ion etcher. This is followed by a further 2 min soak in DI water. The sur- face is primed with an adhesion promoter AP3000 [ 44]. The BCB cyclotene is deposited and spun for 30s at 6000 RPM, followed by 1 min on a hot plate at 150◦C to remove the sol- vent. We use a double layer of BCB cyclotene [ 43]. The first layer is partially cured with a 2 min anneal at 250◦C on a strip annealer. This layer remains ‘tacky’ and bonds well to a second layer of BCB cyclotene, but is viscous enough to pre- vent pinch-through under the membrane during curing, where there can be significant re-flow of the polymer [ 43]. The sec- ond layer of BCB is spun under the same conditions. Separately, the YIG/GGG layer is prepared with a 2 min plasma ash to activate the surface, before a 12 hr evaporatio n of AP3000 is performed in a desiccator. The membrane is removed from the desiccator immediately prior to bonding. The bonding is performed in a simple spring-loaded clamp. The Kapton tape bearing the YIG/GGG membrane is placed on the mirror, with the YIG layer in contact with the BCB cyclotene. The clamp is closed to the point where the layer is held in place with minimal pressure and then heated on a hot plate to 150◦C. The pressure is then increased to the required load. The assembly is then transferred to an oven at 150◦C under nitrogen flow to prevent oxidation of the BCB cyclotene linear polarizerpol. beam splitterhalf-wave platephoto- diode electromagnetxyz tilt/roll obj. lenshalf-wave platephoto- diodes vector network analyserport 1port 2MW amps.tunable laser obj. lens H0 Figure 2. Experimental setup. The output polarization of a 1 270- 1370 nm tunable laser is controlled via a linear polarizer an d a half- wave plate, before being separated into a local oscillator a nd cavity drive. After passing through the cavity, the optical signal orthogonal to the input polarization is recombined with the local oscil lator and measured on a high frequency photodiode. The transmission t hrough the cavity is measured via the light with the same polarizati on as the input on a dc photodiode. A vector network analyzer drive s the magnetic modes and measures the microwave signal from the fa st photodiode. at elevated temperatures. The oven temperature is ramped to 250◦C at 1◦/min, for a 1 hr soak. After allowing the oven to cool to room temperature, the clamp is removed and the Kapton film peeled from the mirror surface. This leaves the YIG/GGG layer secured to the device. During the bonding process, there is some re-flow of the BCB cyclotene to the top surface of the YIG/GGG membrane. To remove this, a 3 min Ar/CF 4reactive ion etch descum is performed. V . EXPERIMENTAL SETUP For measurement, the planar mirror is glued over an aper- ture on a PCB patterned with input and output coplanar waveguides, which are connected to semi-rigid coaxial cabl es. The on-chip strip-line antenna is then wire-bonded to the PC B waveguides for microwave measurement and excitation of the magnon modes in the YIG. The PCB is mounted on a circu- lar stub which sits in an xyz-translation lens mount. The lens array is similarly mounted on a circular stub in a tilt-yaw le ns holder, for full control of the cavity geometry. The device is mounted in an electromagnet, with magnetic field applied orthogonal to the cavity length. Light is focus ed into and out-of the cavity using two aspheric lenses mounted onxyz stages. The cavity is selected by scanning the laser to the correct position. The input laser is an external cavit y diode laser with linewidth ≈1 MHz. The input polarization is set with a rotatable Glan-Thompson prism. On the output, a rotatable half-wave plate is used to select the measuremen t basis on a polarizing beam splitter. From the beam splitter, the transmitted signal with the same polarization as the inp ut light field is measured with a dc photodiode. The polariza- tion scattered light is focused into a single mode fiber, and combined with a local oscillator directly from the laser in a 50:50 fiber coupler. One output of this coupler is measured on a fast photodiode (12 GHz bandwidth) connected via a mi- crowave amplifier to a vector network analyzer (VNA). The VNA is also used to drive the magnetization dynamics via the4 1280 1300 1320 1340 13600.00.10.20.30.40.50.6transmitted intensity (arb. units) input laser wavelength (nm)transmitted intensity (arb.units) -100 -50 0 50 1000.00.10.20.30.40.5 laser detuning (GHz) a) b) c) -50 -25 0 25 50 laser detuning (GHz)0.00.10.20.30.4FSR Figure 3. Transmission spectroscopy of optical modes. (a) W ide wavelength scan, showing free spectral range. Insets show m ode profile imaged in transmission. (b) Measurement of polariza tion of modes. The linear polarization can be set so that only one mod e is excited. This device had a smaller polarization splitting ≈16 GHz. (c) Measurement of polarization splitting and optical line width of device used in BLS measurements. The linear polarization is set so that both modes are probed. This device is also measured in (a ). microwave antenna. VI. CHARACTERIZATION We first characterize the optical modes of the microcavi- ties. The transmitted intensity is measured as a function of input laser wavelength, and angle of input linear polarizat ion, as shown in Fig. 3. A measurement over a wide wavelength range (Fig. 3(a)) is used to determine the free spectral range ∆ωFSR/2π≈6.7THz. A number of spatial modes result- ing from the lateral confinement of the microlens are visible . These can be identified by imaging in transmission, see inset s of Fig. 3(a). The coupling to these higher order modes is min- imized by optimizing the transmitted intensity on resonanc e through the lowest order mode. By measuring the transmitted intensity as a function of the angle of linear polarization, we can find the axes of the ortho g- onal, linearly polarized modes, and the splitting between t he two. An example of this measurement is shown in Fig. 3(b), where the polarization splitting is 16 GHz. This splitting varies with different lens arrays, and is related to slight a sym- metries in the nominally-cylindrical fabricated lens. In t he device used for BLS measurements shown in this paper, the splitting is 32 GHz, as shown in Fig. 3(c). Because the ap- plied magnetic field from the electromagnet is limited to <1 T, we are unable to reach the triple resonance condition. We not e that the frequency splitting due to the magnetic linear bire frin- gence in YIG [ 45] is estimated as ∼900 MHz. This is not largeenough to explain the observed splittings. We note that by fa b- ricating arrays of lenses with varying ellipticity, it woul d be possible to obtain microcavities with a specific splitting. This would enable the triple resonance condition to be achieved. We extract the total dissipation of the optical mode from the linewidth of peak (Fig. 3(c))κ/2π= 11 GHz. This corre- sponds to a Q-factor of 20,000 and Finesse of 600. The expected external loss rate can be estimated from the DBR reflectivity R= 0.9986 asκext=−2∆ωFSRlogR [46], giving κex/2π≈3 GHz. Using these values, and κ=κext+κint, we can estimate the internal dissipation rate κint/2π=8 GHz. This is consistent with the transmitted in- tensity on resonance κ2 ext/κ2≈0.07. If this internal dissipa- tion were solely due to absorption in the YIG layer, we would expectκint=κabs= (αc/n YIG)(tYIG/L)≈1GHz. The dis- crepancy suggests that other dissipation mechanisms play a role. A likely source is the surface roughness on the GGG top surface, where the crack propagates during lift-off. This c ould be alleviated by post-bonding polishing. The choice of mirror reflectivity was conservative to en- sure good coupling to the cavity. If the scattering losses ca n be eliminated, then the mirror reflectivity could be increas ed, while keeping the system over-coupled. In this case, the min - imum possible dissipation rate would be κabs∼1GHz - as achieved in WGM cavities [ 6,8]. VII. BRILLOUIN LIGHT SCATTERING Next, we use homodyne detection to measure the magnon- scattered light, emitted from the microcavity with opposit e linear polarization to the input. The input laser wavelengt h is fixed and set to the lower wavelength optical mode, with frequency ωi. The VNA is used to drive the magnon modes via the microwave antenna, as well as detect the signal at the same frequency from the fast photodiode, where the scattere d light is combined with a local oscillator taken from the inpu t laser. A measurement using this method is shown in Fig. 4(a), as a function of microwave drive frequency and applied magnetic field. When the microwave drive is resonant with a magnon mode with frequency ωmwithin the microcavity, the magnons created scatter with the input optical photons to create opt ical photons at a frequency ωi±ωm. When combined with the local oscillator ωLO=ωi, and mixed on the photodiode, this results in a microwave signal at ±ωm, resulting in the bright lines in Fig. 4(a). The power plotted is the optical power at the photodiode, using the responsivity of the photodiode an d amplification of the amplifier chain to convert from the mea- sured microwave power at the VNA. To check this conversion, we measure the noise equivalent power of the photodiode in darkness, and compare to its specified value. To confirm that the modes result from the embedded mag- netic material, we compare the optical measurement to a stan - dard inductive ferromagnetic resonance (FMR). This is made via the reflected microwave power to the output port of the VNA, and is shown in Fig. 4(b). The change in microwave re- flection coefficient ∆|S11|with magnetic field is plotted over5 a) b) Figure 4. (a) Brillouin light scattering signal. The mixed p ower with the local oscillator, incident on the fast photodiode. A mag netic field independent background has been subtracted. (b) Microwave mea- surement of magnetic modes, via |S11|using the vector network an- alyzer. the same range as Fig. 4(a). We have confirmed that the res- onances in Fig. 4(b) results from the YIG layer in FMR mea- surements over a wider magnetic field. The fact that the slope with magnetic field is the same in both measurements con- firms that the optical signal results from Brillouin light sc at- tering in the YIG. The band of resonances also has the same upper limit in both measurements. The differences in the re- sponse – in particular, that the microwave reflection spectr a has more resonances than the optical BLS – can be explained by the fact that, in the inductive measurement, the entire st rip- line is probed, whereas the optical measurement is only sens i- tive to the region of the YIG film below the lens. The large number of resonances in the inductive measurement is due to strain inhomogeneity across the film from the film trans- fer process. The magnon modes observed in the optical measurement depend on overlap with both the microwave and optical fields [47]. The Kittel mode has the correct symmetry to fulfill these requirements, and we tentatively assign the strongest scat ter- ing to this uniform mode. There are two other modes at higher frequency visible in Fig. 4(a). The mode spacing of these is too large to be due to perpendicular standing spin waves, giv en the thickness of the YIG film [ 48]. A possible candidate for these modes would be magneto-static surface spin waves [ 49] with wavevector set by the width of the microstrip antenna [50]. However, the robust identification of these modes re- quires further measurement and will be the subject of future work. A fit to the Kittel mode in the BLS measurement gives a linewidth of Γ≈20MHz, a value larger than is typical for high quality YIG thin films [ 51]. This is expected, because the current device has imperfections in the YIG layer due to the ion-implantation process, and strain disorder from the bon d- ing process, such as the cracks visible in Fig. 1(c). These im- perfections can be improved by further fabrication process es. Firstly, the damage from ion-implantation can be alleviate d via annealing [ 52]. Secondly, the strain disorder can be re- duced by polishing the GGG from the back of the YIG/GGG bi-layer. It has also been possible to transfer a YIG layer crack-free.The peak measured optical power of the BLS signal is ≈1.2 nW. Given the local oscillator power PLO= 65µW and input microwave power 1 mW, the total conversion efficiency is calculated to be 8×10−16. This low value is to be ex- pected, since the microwave coupling and magnon mode over- lap in this devices have not been engineered. Therefore, to show the value of design we separate the coupling rate Eq. 1 into an optical part Goptand the magnetic fill-factor ηmag, G=Gopt√ηmag, and estimate the obtained rate for the fab- ricated cavity. The optical mode volume for a Gaussian beam can be esti- mated as [ 29] Vopt=πw2 1L/4, (3) whereLis the cavity optical path length and w1is the beam waist on the flat mirror surface. We estimate w2 1= (λ/π)√βL(1−L/β)in the parallax approximation, where βis the radius of curvature of the lens. With the parameters of the measured device, L≈12µm andβ= 70µm, this yields Vopt≈100µm3. This corresponding to Gopt≈50kHz. The magnon mode overlap in the device measured is poor. Taking the whole area of the cracked film part, we estimate η∼10−3, reducing the coupling rate to G≈50Hz. Compared to the whispering gallery mode Vopt≈ 5000µm3, and the waveguide device of Ref. 13Vopt≈ 105µm3, the optical mode volume achieved here is a signif- icant improvement. However, the microwave coupling and magnon confinement are lacking, severely limiting the con- version efficiency. The waveguide device [ 13] has optimized magnon modes ηmag≈1and optimized microwave coupling through an microwave resonator, and even WGM mode de- vices (ηmag≈10−5) benefit from impedance matched mi- crowave coupling to the Kittel mode [ 5]. VIII. CONCLUSIONS We propose and demonstrate an open magneto-optical cav- ity device with optical mode volume limited by the thickness of an embedded magnetic layer. This design is tunable, has the correct polarized modes, and a significantly reduced opt i- cal mode volume compared to previous devices [ 5–7,13]. We envisage that simple improvements in the demonstrated device design should enable the strong-coupling regime to b e reached. By removing the GGG from the device via polish- ing, the cavity length can be reduced to 3 um, and using the lowest radius of curvature lens β= 22µm, the resulting mode volume would be Vopt≈7µm3(from Eq. 3). Combining this with lateral patterning of the YIG layer to confine a magnon mode to a disk with diameter 5 µm, it should be possible to achieveG= 200 kHz using the open microcavity design. If we combine this with the discussed improvements in the magnon and optical linewidth to Γ = 1 MHz and κ= 1GHz, respectively, this would lead to a single photon cooperativ ity ofC= 4G2/Γκ= 10−4. We would then require an opti- cal pump power of ≈5mW to achieve the strong coupling regime√nG > κ, Γ. In order to achieve cooling of magnetic6 mode via optical damping, Γopt= 4nG2/κ[53,54] compa- rable to the magnetic damping would require only ≈1µW input power [ 10]. Finally, it will be necessary to couple microwaves effi- ciently into the resulting small volume of magnetic materia l. Elsewhere, we have demonstrated that this is possible using low impedance microwave resonators [ 55]. With careful mi- crowave circuit optimization it is possible to achieve coup ling to femtolitre magnetic volumes [ 56], to match that possible with open optical microcavities [ 29]. As well as demonstrating progress towards microwave- optical conversion [ 4], it is expected that the enhancement of the magnon-photon interaction demonstrated could have sig - nificant impact in magnonics [ 57], through the increased mea- surement sensitivity and in optical modification of the magn on dynamics. The versatility of the fabrication method meansthat antiferromagnetic materials could also be embedded in the microcavity in order to explore the interaction of optic al photons with THz magnon modes [ 18,58]. ACKNOWLEDGEMENTS We are grateful to Aurilien Trichet and Jason Smith (Ox- ford HiQ) for advice on open microcavities, Roger Webb (Surrey ion beam centre) for assistance with ion implantati on, and Miguel Levy, Dries Van Thourhout, Koji Usami, Andreas Nunnenkamp and Paul Walker for useful discussions. This work was supported by the European Union’s Horizon 2020 research and innovation programme under grant agreement No 732894 (FET Proactive HOT). The data plotted in the fig- ures can be accessed at the Zenodo repository [ 59]. [1] Y . Tabuchi, S. Ishino, T. Ishikawa, R. Yamazaki, K. Us- ami, and Y . Nakamura, “Hybridizing Ferromagnetic Magnons and Microwave Photons in the Quantum Limit,” Phys. Rev. Lett. 113, 083603 (2014) . [2] H. Huebl, C. W. Zollitsch, J. Lotze, F. Hocke, M. Greifens tein, A. Marx, R. Gross, and S. T. B. Goennenwein, “High Coopera- tivity in Coupled Microwave Resonator Ferrimagnetic Insul ator Hybrids,” Phys. Rev. Lett. 111, 127003 (2013) . [3] X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, “Strongly Coupled Magnons and Cavity Microwave Photons,” Phys. Rev. Lett. 113, 156401 (2014) . [4] R. Hisatomi, A. Osada, Y . Tabuchi, T. Ishikawa, A. Noguch i, R. Yamazaki, K. Usami, and Y . Nakamura, “Bidirectional conversion between microwave and light via ferromagnetic magnons,” Phys. Rev. B 93, 174427 (2016) . [5] A. Osada, R. Hisatomi, A. Noguchi, Y . Tabuchi, R. Yamazak i, K. Usami, M. Sadgrove, R. Yalla, M. Nomura, and Y . Naka- mura, “Cavity Optomagnonics with Spin-Orbit Coupled Pho- tons,” Phys. Rev. Lett. 116, 223601 (2016) . [6] X. Zhang, N. Zhu, C.-L. Zou, and H. X. Tang, “Optomagnonic Whispering Gallery Microresonators,” Phys. Rev. Lett. 117, 123605 (2016) . [7] J. A. Haigh, A. Nunnenkamp, A. J. Ramsay, and A. J. Fergu- son, “Triple-Resonant Brillouin Light Scattering in Magne to- Optical Cavities,” Phys. Rev. Lett. 117, 133602 (2016) . [8] J. A. Haigh, S. Langenfeld, N. J. Lambert, J. J. Baum- berg, A. J. Ramsay, A. Nunnenkamp, and A. J. Ferguson, “Magneto-optical coupling in whispering-gallery-mode re s- onators,” Phys. Rev. A 92, 063845 (2015) . [9] N. J. Lambert, A. Rueda, F. Sedlmeir, and H. G. L. Schwefel, “Coherent Conversion Between Microwave and Op- tical Photons—An Overview of Physical Implementations,” Adv. Quantum Technol. 3, 1900077 (2020) . [10] S. Sharma, Y . M. Blanter, and G. E. W. Bauer, “Optical Coo ling of Magnons,” Phys. Rev. Lett. 121, 087205 (2018) . [11] F. ˇSimi´ c, S. Sharma, Y . M. Blanter, and G. E. W. Bauer, “Coherent pumping of high-momentum magnons by light,” Phys. Rev. B 101, 100401(R) (2020) . [12] S. Sharma, B. Z. Rameshti, Y . M. Blanter, and G. E. W. Bauer, “Optimal mode matching in cavity optomagnonics,” Phys. Rev. B 99, 214423 (2019) .[13] N. Zhu, X. Zhang, X. Han, C.-L. Zou, C. Zhong, C.-H. Wang, L. Jiang, and H. X. Tang, “Waveguide cavity optomagnon- ics for broadband multimode microwave-to-optics conversi on,” arXiv:2005.06429 (2020), arXiv:2005.06429 . [14] S. Viola Kusminskiy, H. X. Tang, and F. Marquardt, “Coupled spin-light dynamics in cavity optomagnonics,” Phys. Rev. A 94, 033821 (2016) . [15] K. J. Vahala, “Optical microcavities,” Nature 424, 839 (2003) . [16] F. Heyroth, C. Hauser, P. Trempler, P. Geyer, F. Sy- rowatka, R. Dreyer, S. G. Ebbinghaus, G. Woltersdorf, and G. Schmidt, “Monocrystalline Freestanding Three- Dimensional Yttrium-Iron-Garnet Magnon Nanoresonators, ” Phys. Rev. Applied 12, 054031 (2019) . [17] H. Yu, Y . Peng, Y . Yang, and Z.-Y . Li, “Plasmon- enhanced light–matter interactions and applications,” Npj Comput. Mater. 5, 1 (2019) . [18] S. Schlauderer, C. Lange, S. Baierl, T. Ebnet, C. P. Schmid, D. C. Valovcin, A. K. Zvezdin, A. V . Kimel, R. V . Mikhaylovskiy, and R. Huber, “Temporal and spec- tral fingerprints of ultrafast all-coherent spin switching ,” Nature 569, 383 (2019) . [19] N. Maccaferri, I. Zubritskaya, I. Razdolski, I.-A. Chioar, V . Belotelov, V . Kapaklis, Peter M. Oppe- neer, and A. Dmitriev, “Nanoscale magnetophotonics,” Journal of Applied Physics 127, 080903 (2020) . [20] M. Trupke, E. A. Hinds, S. Eriksson, E. A. Curtis, Z. Mok- tadir, E. Kukharenka, and M. Kraft, “Microfabricated high- finesse optical cavity with open access and small volume,” Appl. Phys. Lett. 87, 211106 (2005) . [21] T. Steinmetz, Y . Colombe, D. Hunger, T. W. H¨ ansch, A. Ba loc- chi, R. J. Warburton, and J. Reichel, “Stable fiber-based Fab ry- P´ erot cavity,” Appl. Phys. Lett. 89, 111110 (2006) . [22] A. Muller, E. B. Flagg, M. Metcalfe, J. Lawall, and G. S. Solomon, “Coupling an epitaxial quan- tum dot to a fiber-based external-mirror microcavity,” Appl. Phys. Lett. 95, 173101 (2009) . [23] G. Cui, J. M. Hannigan, R. Loeckenhoff, F. M. Matinaga, M. G. Raymer, S. Bhongale, M. Holland, S. Mosor, S. Chat- terjee, H. M. Gibbs, and G. Khitrova, “A hemispherical, high-solid-angle optical micro-cavity for cavity-QED stu dies,” Opt. Express, OE 14, 2289 (2006) .7 [24] M. Trupke, J. Goldwin, B. Darqui´ e, G. Dutier, S. Erikss on, J. Ashmore, and E. A. Hinds, “Atom Detection and Pho- ton Production in a Scalable, Open, Optical Microcavity,” Phys. Rev. Lett. 99, 063601 (2007) . [25] R. Albrecht, A. Bommer, C. Deutsch, J. Reichel, and C. Becher, “Coupling of a Single Nitrogen-Vacancy Center in Diamond to a Fiber-Based Microcavity,” Phys. Rev. Lett. 110, 243602 (2013) . [26] D. Wang, H. Kelkar, D. Martin-Cano, T. Utikal, S. G¨ otzi nger, and V . Sandoghdar, “Coherent Coupling of a Sin- gle Molecule to a Scanning Fabry-Perot Microcavity,” Phys. Rev. X 7, 021014 (2017) . [27] S. Dufferwiel, S. Schwarz, F. Withers, A. a. P. Trichet, F. Li, M. Sich, O. Del Pozo-Zamudio, C. Clark, A. Nalitov, D. D. Solnyshkov, G. Malpuech, K. S. Novoselov, J. M. Smith, M. S. Skolnick, D. N. Krizhanovskii, and A. I. Tartakovskii, “Exc i- ton–polaritons in van der Waals heterostructures embedded in tunable microcavities,” Nat. Commun. 6, 8579 (2015) . [28] R. J. Barbour, P. A. Dalgarno, A. Curran, K. M. Nowak, H. J. Baker, D. R. Hall, N. G. Stoltz, P. M. Petroff, and R. J. Warburton, “A tunable microcavity,” Journal of Applied Physics 110, 053107 (2011) . [29] P. R. Dolan, G. M. Hughes, F. Grazioso, B. R. Patton, and J. M. Smith, “Femtoliter tunable optical cavity arrays, ” Opt. Lett. 35, 3556 (2010) . [30] Oxford HiQ https://www.oxfordhighq.com/ . [31] A. A. P. Trichet, P. R. Dolan, D. M. Coles, G. M. Hughes, an d J. M. Smith, “Topographic control of open-access microcavi ties at the nanometer scale,” Opt. Express, OE 23, 17205 (2015) . [32] T. Goto, Y . Eto, K. Kobayashi, Y . Haga, M. Inoue, and C. A. Ross, “Vacuum annealed cerium-substituted yttrium iron ga rnet films on non-garnet substrates for integrated optical circu its,” Journal of Applied Physics 113, 17A939 (2013) . [33] T. Yoshimoto, T. Goto, R. Isogai, Y . Nakamura, H. Takagi , C. A. Ross, and M. Inoue, “Magnetophotonic crystal with cerium substituted yttrium iron garnet and enhanced Farada y rotation angle,” Opt. Express, OE 24, 8746 (2016) . [34] M. Levy, R. M. Osgood, A. Kumar, and H. Bakhru, “Epitaxia l liftoff of thin oxide layers: Yttrium iron garnets onto GaAs ,” Appl. Phys. Lett. 71, 2617 (1997) . [35] Innovent Jena https://www.innovent-jena.de/ . [36] Ion implantation was carried out at University of Surre y Ion Beam Centre. [37] M. Levy, R. M. Osgood, A. Kumar, and H. Bakhru, “Crystal ion slicing of single-crystal magnetic garnet film s,” Journal of Applied Physics 83, 6759 (1998) . [38] M. Levy and R. M. Osgood, “Crystal ion-slicing of single - crystal films,” US Pat. , US6120597A (2000). [39] “CYCLOTENETM3000 Series Advanced Electronics Resins,” Dow Chem. Appl. Note (2012). [40] F. Niklaus, P. Enoksson, E. K¨ alvesten, and G. Stemme, “Low-temperature full wafer adhesive bonding,” J. Micromech. Microeng. 11, 100 (2001) . [41] T. M. Stokich, C. C. Fulks, M. T. Bernius, D. C. Burdeaux, P. E. Garrou, and R. H. Heistand, “Planarization with CycloteneTM 3022 (BCB) Polymer Coatings,” MRS Online Proc. Libr. Arch. 308(1993). [42] Y .-G. Zhao, W.-K. Lu, Y . Ma, S.-S. Kim, S. T. Ho, and T. J. Marks, “Polymer waveguides useful over a very wide wavelength range from the ultraviolet to infrared,” Appl. Phys. Lett. 77, 2961 (2000) .[43] G. Roelkens, J. Brouckaert, D. Van Thourhout, R. Baets, R. N¨ otzel, and M. Smit, “Adhesive Bonding of InP / InGaAsP Dies to Processed Silicon-On-Insulator Wafers using DVS-b is- Benzocyclobutene,” J. Electrochem. Soc. 153, G1015 (2006) . [44] “Processing Procedures for BCB adhesion,” Dow Chem. Ap pl. Note (2007). [45] R. V . Pisarev, I. G. Sinii, N. N. Kolpakova, and Yu M. Yako vlev, “Magnetic birefringence of light in iron garnets,” Sov Phys JETP 33, 1175 (1971). [46] N. Ismail, C. C. Kores, D. Geskus, and M. Poll- nau, “Fabry-P´ erot resonator: Spectral line shapes, gener ic and related Airy distributions, linewidths, finesses, and performance at low or frequency-dependent reflectivity,” Opt. Express, OE 24, 16366 (2016) . [47] J. A. Haigh, N. J. Lambert, S. Sharma, Y . M. Blanter, G. E. W. Bauer, and A. J. Ramsay, “Selection rules for cavity-enhanced Brillouin light scattering from magnetos tatic modes,” ArXiv180400965 Cond-Mat Physicsphysics (2018), arXiv:1804.00965 [cond-mat, physics:physics] . [48] S. Klingler, A. V . Chumak, T. Mewes, B. Khodadadi, C. Mewes, C. Dubs, O. Surzhenko, B. Hillebrands, and A. Conca, “Measurements of the exchange stiffness of YIG films using broadband ferromagnetic resonance techniques, ” J. Phys. D: Appl. Phys. 48, 015001 (2014) . [49] R. W. Damon and J. R. Eshbach, “Mag- netostatic modes of a ferromagnet slab,” Journal of Physics and Chemistry of Solids 19, 308 (1961) . [50] P. Andrich, C. F. de las Casas, X. Liu, H. L. Bretscher, J. R. Berman, F. J. Heremans, P. F. Nealey, and D. D. Awschalom, “Long-range spin wave medi- ated control of defect qubits in nanodiamonds,” Npj Quantum Inf. 3(2017), 10.1038/s41534-017-0029-z . [51] C. Dubs, O. Surzhenko, R. Linke, A. Danilewsky, U. Br¨ uc k- ner, and J. Dellith, “Sub-micrometer yttrium iron gar- net LPE films with low ferromagnetic resonance losses,” J. Phys. D: Appl. Phys. 50, 204005 (2017) . [52] F. J. Rachford, M. Levy, R. M. Osgood, A. Ku- mar, and H. Bakhru, “Magnetization and ferromag- netic resonance studies in implanted and crystal ion sliced bismuth-substituted yttrium iron garnet films,” Journal of Applied Physics 85, 5217 (1999) . [53] M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, “Cav ity optomechanics,” Rev. Mod. Phys. 86, 1391 (2014) . [54] V . A. S. V . Bittencourt, V . Feulner, and S. Viola Kus- minskiy, “Magnon heralding in cavity optomagnonics,” Phys. Rev. A 100, 013810 (2019) . [55] L. McKenzie-Sell, J. Xie, C.-M. Lee, J. W. A. Robinson, C . Cic- carelli, and J. A. Haigh, “Low-impedance superconducting m i- crowave resonators for strong coupling to small magnetic mo de volumes,” Phys. Rev. B 99, 140414(R) (2019) . [56] V . Ranjan, S. Probst, B. Albanese, T. Schenkel, D. Vion, D. Esteve, J. J. L. Morton, and P. Bertet, “Electron spin resonance spectroscopy with femtoliter detection volume, ” Appl. Phys. Lett. 116, 184002 (2020) . [57] V . V . Kruglyak, S. O. Demokritov, and D. Grundler, “Magn on- ics,” J. Phys. D: Appl. Phys. 43, 264001 (2010) . [58] J. Walowski and M. M¨ unzenberg, “Perspec- tive: Ultrafast magnetism and THz spintronics,” Journal of Applied Physics 120, 140901 (2016) . [59]https://doi.org/10.5281/zenodo.4012308 . | 2020-09-09 | Microwave-to-optical conversion via ferromagnetic magnons has so-far been
limited by the optical coupling rates achieved in mm-scale whispering gallery
mode resonators. Towards overcoming this limitation, we propose and demonstrate
an open magneto-optical cavity containing a thin-film of yttrium iron garnet
(YIG). We achieve a 0.1 pL (100 $\mu$m$^{3}$) optical mode volume, $\sim$50
times smaller than previous devices. From this, we estimate the magnon
single-photon coupling rate is $G\approx50$ Hz. This open cavity design offers
the prospect of wavelength scale mode volumes, small polarization splittings,
and good magneto-optical mode overlap. With achievable further improvements and
optimization, efficient microwave-optical conversion and magnon cooling devices
become a realistic possibility. | Sub-pico-liter magneto-optical cavities | 2009.04162v1 |
arXiv:1905.07278v1 [physics.optics] 17 May 2019Keywords : Optomagnonic Cavity, Voigt Geometry, Magnetostatic Spin Waves , Inelastic Light Scattering, Time Floquet Method High-efficiency triple-resonant inelastic light scatterin g in planar optomagnonic cavities Petros Andreas Pantazopoulos,∗Kosmas L. Tsakmakidis, Evangelos Almpanis, Grigorios P. Zouros, and Nikolaos Stefanou Section of Solid State Physics, National and Kapodistrian U niversity of Athens, Panepistimioupolis, GR-157 84 Athens, Greece (Dated: May 20, 2019) Abstract Optomagnonic cavities have recently been emerging as promi sing candidates for implementing coherent microwave-to-optical conversion, quantum memor ies and devices, and next generation quantum networks. A key challenge in the design of such cavit ies is the attainment of high efficien- cies, which could, e.g., be exploited for efficient optical in terfacing of superconducting qubits, as well as the practicality of the final designs, which ideally s hould be planar and amenable to on-chip integration. Here, on the basis of a novel time Floquet scatt ering-matrix approach, we report on the design and optimization of a planar, multilayer optomag nonic cavity, incorporating a Ce:YIG thin film, magnetized in-plane, operating in the triple-res onant inelastic light scattering regime. This architecture allows for conversion efficiencies of abou t 5%, under realistic conditions, which is orders of magnitude higher than alternative designs. Our re sults suggest a viable way forward for realizing practical information inter-conversion betwee n microwave photons and optical photons, mediated by magnons, with efficiencies intrinsically greate r than those achieved in optomechanics and alternative related technologies, as well as a platform for fundamental studies of classical and quantum dynamics in magnetic solids, and implementation of futuristic quantum devices. PACS numbers: 1I. INTRODUCTION Optomagnoniccavitiesarejudiciouslydesigneddielectricstructure sthatincludemagnetic materials capable of simultaneously confining light and spin waves in the same region of space. This confinement leads, under certain conditions, to stron g enhancement of the inherently weak interaction between the two fields, which allows for a n efficient microwave- to-optical transduction, enabling, e.g., optical interfacing of sup erconducting qubits1,2. Theoptomagnonic interactionis expected to belarger when theso- called triple-resonance condition is met, i.e., when the frequency of a cavity magnon matches a photon transition between two resonant modes. This implies that the cavity must supp ort two well-resolved optical resonances (in the hundred terahertz range) separate d by a few gigahertz, which requires quality factors at least of the order of 105, as schematically depicted in figure 1. A (sub)millimeter-sized sphere, made of a low-loss dielectric magnetic material, consti- tutes a simple realization of an optomagnonic cavity. The sphere sup ports densely spaced long-lifetime optical whispering gallery modes3–7, and infrared incident light evanescently coupled to these modes can be scattered by a uniformly precessing (so-called Kittel) spin wave toa neighbouring optical whispering gallery mode. Inthe prosp ect of achieving smaller modal volumes and larger spatial overlap between the interacting fi elds, higher-order mag- netostatic modes8–10, magnetically split optical Mie resonances in small spheres11, as well as particles of different shapes12have been proposed. However, these proposals currently face appreciable challenges in the fabrication of high-quality particle s and/or the efficient excitation of the spin waves. A promising alternative design of optomagnonic cavities is based on planargeometries, which can exhibit even stronger magnon-to-photon conversion effi ciencies13, while at the same time allowing integration into a hybrid opto-microwave chip using m odern nanofab- rication methods. To this end, optomagnonic cavities formed in a mag netic dielectric film bounded by two mirrors14–16, or in a defect layer in a dual photonic-magnonic periodic lay- ered structure17, have also been investigated. However, the studies reported so f ar refer to the Faraday configuration, with out-of-plane magnetized films, wh ere it is challenging to obtain two optical resonances in the required close proximity to eac h other. In this work we show that, by using in-plane magnetized films in the so- called Voigt con- figuration, wecanovercometheafore-describedshortcomingso fprevious schemes anddesign 2/g90out /g90in /g39/g90/g58 /g900/g900~ 10 -4 FIG. 1: Schematic of inelastic light scattering through mag non absorption in an optomagnonic cavity. The frequency of a cavity magnon, Ω, matches a photon transition between two resonant optical modes: Ω = ∆ ω≡ωout−ωin(triple-resonance condition). efficient optomagnonic cavities operating in the triple-resonance re gime. In section II we de- scribe our statically magnetized structure and discuss its optical r esponse. In section III we summarize our recently developed fully dynamic time Floquet metho d for layered opto- magnonic structures16and in section IV we present details of our attained numerical result s. The last section concludes the article. II. STRUCTURE DESIGN We propose a simple design of planar optomagnonic cavity, simultaneo usly confining light and spin waves in the same subwavelength region of space. It consis ts of an iron garnet thin film bounded symmetrically by two-loss, dielectric Bragg mirrors, in air , as schematically illustrated in figure 2(a). Iron garnets are ferrimagnetic materials exhibiting important func tionalities for bulk 3and thin-film device applications that require magnetic insulators, ow ing to their unique physical properties such as high optical transparency in a wide ran ge of wavelengths, high Curie temperature, ultra-low spin-wave damping, and strong magn eto-optical coupling18. In our work, we consider cerium-substituted yttrium iron garnet ( Ce:YIG) which, at the telecom wavelength of 1 .5µm, has a relative electric permittivity ǫ= 5.10+i4×10−4and a Faraday coefficient f=−0.0119, while its relative magnetic permeability equals unity. The Ce:YIG film extends from −d/2 tod/2 and is magnetically saturated to M0by an in-plane bias magnetic field H0oriented, say, along the xdirection. Therefore, the corresponding relative electric permittivity tensor, neglecting the small Cotton-M outon contributions, is of the form20 ǫ= ǫ0 0 0ǫ if 0−if ǫ . (1) We consider the Voigt geometry with light propagating in the y-zplane. The struc- ture in this geometry, with the magnetic field parallel to the surface and also perpendicular to the propagation direction, remains invariant under reflection wit h respect to the plane of incidence. Consequently, contrary to the Faraday configurat ion studied in our previous work15–17, the transverse magnetic (TM) and transverse electric (TE) pola rization modes, i.e., modes with the electric field oscillating in and normal to the plane of in cidence, respec- tively, are eigenmodes of the system. Interestingly, in the chosen geometry, the magnetic film behaves as isotropic, with permittivity ǫ−f2/ǫandǫfor TM- and TE-polarized waves, respectively. In other words, only TM-polarized light is affected by t he (magnetic) polariza- tion field. Each Bragg mirror consists of an alternate sequence of six SiO 2and six Si quarter-wave layers, i.e., dm/radicalig (2πnm/λ)2−q2y=π/2, where dm(m: SiO 2or m: Si) is the layer thickness andnmthe corresponding refractive index ( nSiO2= 1.47 andnSi= 3.5) at the operation wavelength λ≈1.5µm21,22. Due to translation invariance parallel to the x-yplane, the in-plane component of the wave vector, qy= 2πsinθ/λ, whereθis the angle of incidence, remains constant. Taking, for instance, qy= 3µm−1, which corresponds to an angle of incidence of about 45o, we obtain dSiO2= 290 nm and dSi= 110 nm. Accordingly, we choose a thickness d= 350 nm for the Ce:YIG film to satisfy the half-wave condition that corresponds to transmission maxima. 4M z /s48/s46/s48/s48/s46/s53/s49/s46/s48/s49/s46/s52/s57/s55/s51 /s49/s46/s52/s57/s55/s50 /s50/s48/s48/s46/s50/s50 /s50/s48/s48/s46/s50/s52/s40 /s109/s41 /s84/s69 /s47/s50 /s40/s84/s72/s122/s41/s84/s84/s77 /s32(a) (b) FIG. 2: (a) Schematic view of the optomagnonic cavity under s tudy. It is formed by a 350- nm-thick Ce:YIG film magnetized in-plane (along the xdirection), bounded symmetrically by two Bragg mirrors, each consisting of6periodsofalternating S iO2andSilayers ofthickness 290nmand 110 nm, respectively, grown along the zdirection. For light incident with qy= 3µm−1, the cavity supports two localized resonant modes, one of TM and the othe rs of TE polarization, manifested in the corresponding transmission spectrum shown in (b), with the dotted and solid curves referring to the lossless and lossy structure, respectively. A snapsh ot of the associated electric field profiles along the zdirection in the lossless case is illustrated in (a). This design provides two (one TM and one TE) high-quality-factor re sonances within the lowest Bragg gap, at a wavelength of about 1.5 µm, separated by a frequency difference ∆f= 9.5 GHz that matches the frequency of magnetostatic spin waves18,20. These resonant modes are strongly localized in the region of the Ce:YIG film, which can b e considered as a defect in the periodic stacking sequence of the Bragg mirrors. Abs orption losses reduce the transmittance peak. In particular, the long-lifetime TE resonance is strongly suppressed in the presence of dissipative losses, as shown by the solid line blue line in fi gure 2(b). It should be pointed out that the position and width of the optical re sonances can be tailored at will by appropriate selection of the materials, and by prop erly adjusting the geometric parameters of the structure and the angle of incidence . 5III. THEORY FOR LAYERED OPTOMAGNONIC STRUCTURES The magnetic Ce:YIG film supports magnetostatic spin waves where t he magnetization precesses in-phase, elliptically, throughout the film (uniform prece ssion mode) with angular frequencyΩ =/radicalbig ΩH(ΩH+ΩM), whereΩ H=γµ0H0andΩ M=γµ0M0,γbeingthegyromag- netic ratio and µ0the magnetic permeability of vacuum20. The corresponding magnetization field profile is given by M(r,t)/M0=/hatwidex+ηAysin(Ωt)/hatwidey+ηAzcos(Ωt)/hatwidez, (2) whereAy=/radicalbig (ΩH+ΩM)/(2ΩH+ΩM),Az=/radicalbig ΩH/(2ΩH+ΩM), andηis an amplitude factor that defines the magnetization precession angle. Under the action of the spin wave, the magnetic film and, consequen tly, the entire struc- ture can be looked upon as a periodically driven system because the m agnetization field, given by Eq. (2), induces a temporal perturbation15 δǫ(t) =1 2/bracketleftbig δǫexp(−iΩt)+δǫ†exp(iΩt)/bracketrightbig (3) in the permittivity tensor of the statically magnetized material, wher e δǫ=fη 0iAzAy −iAz0 0 −Ay0 0 . (4) The solutions of the underlying Maxwell equations are Floquet modes F(r,t) = Re{F(r,t)exp(−iωt)}, withF(r,t+T) =F(r,t),T= 2π/Ω, where by Fwe denote electric field, electric displacement, magnetic field, and magnetic indu ction, while ωis the Floquet quasi-frequency, similarly to the Floquet quasi-momentum ( or else the Bloch wave vector) when there is spatial periodicity23,24. Seeking Floquet modes in the form of plane waves with given qyand expanding all time-periodic quantities into truncated Fourier se ries in the basis of complex exponential functions exp( inΩt),n=−N,−N+1,...,N, leads to an eigenvalue-eigenvector equation, which has 4(2 N+ 1) physically acceptable solutions16. We characterize them by the following indices: s= +(−) that denotes waves propagating or decaying in the positive (negative) zdirection, p= 1,2 that indicates the two eigen- polarizations, and ν=−N,−N+ 1,···,Nwhich labels the different eigenmodes. These eigenmodes are polychromatic waves, each composed of 2 N+1 monochromatic components 6of angular frequency ω−nΩ,n=−N,−N+ 1,...,N16. We note that, in a static ho- mogeneous medium, the corresponding eigenmodes of the electrom agnetic (EM) field are monochromatic waves characterized by the indices s,p, andn. Scattering of an eigenmode occurs at an interface between two diff erent homogeneous media. For such a planar interface between a static and a time-perio dic medium, the relative complex amplitudes of the transmitted (reflected) waves, denote d byQI pν;p′n′(QIII pn;p′n′) for incidence intheforwarddirection or QIV pn;p′ν′(QII pν;p′ν′)for incidence inthebackward direction in the configuration shown in figure 3, are obtained in the manner des cribed in Ref.16. Primed indices refer to the incident wave. For an interface between two static homogeneous media, the Qmatrices relate monochromatic waves and are diagonal in n, which reflects frequency conservation. We note that, in order to evaluate the s cattering properties of layered optomagnonic structures in a straightforward manner, t he waves on each side of a given interface are expressed around different points, at a dista nce−d1andd2from the center of the interface (see figure 3), so that all backward a nd forward propagating or evanescent waves in the region between two consecutive interf aces refer to the same (arbitrary) origin. Of course, because of translation invariance p arallel to the x-yplane, the choice of the x-ycomponents of d1andd2are immaterial; thus, for simplicity, we choose d1 andd2along the zdirection. The transmission and reflection matrices of a pair of consecutive int erfaces, i and i+1, are obtained by properly combining those of the two interfaces so a s to describe multiple scattering to any order. This leads to the following expressions aft er summing up the infinite geometric series involved, as schematically illustrated in figure 3, i.e., QI(i,i+1)=QI(i+1)[I−QII(i)QIII(i+1)]−1QI(i) QII(i,i+1)=QII(i+1)+QI(i+1)QII(i)[I−QIII(i+1)QII(i)]−1QIV(i+1) QIII(i,i+1)=QIII(i)+QIV(i)QIII(i+1)[I−QII(i)QIII(i+1)]−1QI(i) QIV(i,i+1)=QIV(i)[I−QIII(i+1)QII(i)]−1QIV(i+1). (5) It should benotedthat thewaves onthe left (right) ofthepair of in terfaces arereferred toan origin at a distance −d1(i) [d2(i+1)] from the center of the i-th [(i+1)-th] interface. We also recall that, though the choice of d1andd2associated to each interface is to a certain degree arbitrary, it must besuch that d2z(i)+d1z(i+1) equals thethickness ofthelayer between the 7Dynamic Static QI pν ;p'n' QIV pn ;p'ν'QII pν ;p'ν'QIII pn ;p'n' O1 O2 d1d2 z+ + +... + + +...+ + +...+ + +... Q :I Q :II Q :III Q :IV FIG. 3: Left-hand diagram: Transmission and reflection matr ices for a planar interface between a static and a dynamic medium, defined with respect to appropri ate origins, O1andO2, at distances −d1andd2from a center at the interface, respectively. Right-hand di agram: The transmission and reflection matrices of two consecutive interfaces are ev aluated by summing up all relevant multiple-scattering processes. i-thand(i+1)-thinterfaces. It is obvious that onecan repeat the above process to obtainthe transmission and reflection matrices Qof three consecutive interfaces, by combining those of the pair of the first interfaces with those of the third interface , and so on, by properly combining the Qmatrices of component units, one can obtain the Qmatrices of a slab which comprises any finite number of interfaces23,24. This method applies to an arbitrary slab which comprises periodically time-varying layers, provided that a ll dynamic media have the same temporal periodicity. It is then straightforward to calcu late the transmittance, T, and reflectance, R, of the slab as the ratio of the transmitted and reflected, respec tively, energy flux to the energy flux associated with the incident wave. TandRare given by the sum of the corresponding quantities over all scattering channels ( p,n):T=/summationtext p,nTpnand R=/summationtext p,nRpn. It is worthnoting that, because of the time variation of the permit tivity tensor, the EM energy is not conserved even in the absence of diss ipative (thermal) losses. In this case, A= 1−T −R >0(<0) means energy transfer from (to) the EM to (from) 8the spin-wave field. We close this section by pointing out a useful polarization selection ru le, which can be readily derived in the linear-response approximation. To first order , the coupling strength associated to the photon-magnon scattering is proportional to t he overlap integral G= /an}bracketle{tout|δǫ|in/an}bracketri}ht, where/an}bracketle{trt|in/an}bracketri}ht=Ein(z)exp[i(q/bardbl·r−ωt)] and/an}bracketle{tout|r′t′/an}bracketri}ht=Eout⋆(z)exp[−i(q′ /bardbl· r−ω′t)] denote appropriate incoming and outgoing monochromatic time-h armonic waves in the static magnetic layered structure. Using Eq. (3) we obtain G= 4π3fηδ(q/bardbl−q′ /bardbl)/bracketleftbig δ(ω−ω′−Ω)g−+δ(ω−ω′+Ω)g+/bracketrightbig (6) whereg±=/integraldisplay dzu±·/bracketleftbig Eout⋆(z)×Ein(z)/bracketrightbig , withu±=∓Ay/hatwidey+iAz/hatwidez. The delta functions in Eq. (6) express conservation of in-plane momentum and energy in in elastic light scattering processes that involve emission and absorption of one magnon by a p hoton, as expected in the linear regime. Obviously, the amplitude of transition between two optical eigenmodes of the same polarization, TM or TE, is identically zero because the corre sponding eigenvectors are real. In other words, one-magnon processes change the linea r polarization state of a photon. IV. RESULTS AND DISCUSSION We now assume continuous excitation of a uniform-precession spin- wave mode in the magnetic film, with a relative amplitude η= 0.06, which induces a periodic time variation in thecorresponding electricpermittivitytensor, givenbyEq. (3). T heoptomagnonicstructure is illuminated from the left by TM-polarized light with qy= 3µm−1at the corresponding resonance frequency, which corresponds to an angle of incidence of about 45o. The dynamic optical response of the structure is calculated with sufficient accu racy by considering a cutoff ofN= 5 in the Fourier series expansions involved in our time Floquet scatte ring-matrix method outlined in section III. figure 4(a) shows the total (transmitted plus reflected) intensit ies,In=/summationtext p(Tpn+Rpn), as a function of the spin-wave frequency Ω /2π. It can be seen that inelastic light scattering is negligible when the allowed final photon states fall within a gap, wher e the optical density of states is very low, and we essentially have only the elastic outgoing beam. On the contrary, when the spin-wave frequency matches the frequenc y difference ∆ f= 9.5 GHz 9/s53 /s49/s48 /s49/s53/s48/s46/s48/s48/s46/s53/s49/s46/s48 /s45/s48/s46/s51/s48/s46/s48 /s53 /s49/s48 /s49/s53/s48/s46/s51/s48/s46/s52/s40/s100/s41/s40/s99/s41/s40/s98/s41 /s110 /s50/s110 /s49 /s32/s32/s73 /s110 /s47/s50 /s40/s71/s72/s122/s41/s110 /s48 /s32/s32 /s32/s49/s48/s52 /s65 /s32 /s32/s65 /s47/s50 /s40/s71/s72/s122/s41 FIG. 4: The structure of figure 2(a), under continuous excita tion of a uniform precession spin-wave mode of angular frequency Ω with a relative amplitude η= 0.06, is illuminated from the left by TM-polarized light with qy= 3µm−1at the corresponding resonance frequency [see figure 2(b)]. Variation of the dominant elastic and inelastic total outgo ing light intensities versus the spin-wave frequency (a) and corresponding optical absorption (b) and (c). Dotted and solid curves refer to the lossless and lossy structure, respectively. between the two optical resonances [see figure 2(a)], the triple-r esonance condition is fulfilled and one-magnon absorption processes are favoured, leading to e nhanced intensities of the corresponding ( n=−1) inelastically transmitted and reflected light beams, with conversio n efficiency of the order of 30% if dissipative losses are neglected. At t he same time, the elastic beam intensity is considerably reduced while the other inelastic proce sses are also resonantly affected, though to a much lesser degree, as shown in figure 4(a) a nd also in figure 5. Overall, there is an excess number of magnons absorbed, which can be accounted for by our fully dynamic time Floquet scattering-matrix method16. This is manifested as a small negative absorption peak [see figure 4(b)], which clearly indicates a r esonant energy transfer from the magnon to the photon field. Considering a saturation magnetization M0= 150 emu /cm3for Ce:YIG19, the triple- resonance condition (Ω /2π= 9.5 GHz) is achieved with a bias magnetic field H0= 2.5 kOe. Inthiscase, theconeangleofmagnetizationprecession(ellipticalin thechosenconfiguration) attains a maximum of 2 .75o, which is a tolerable value for linear spin waves. 10/s53 /s49/s48 /s49/s53/s49/s48/s45/s53/s49/s48/s45/s51/s49/s48/s45/s49/s49/s48/s48 /s53 /s49/s48 /s49/s53/s49/s48/s45/s57/s49/s48/s45/s55/s49/s48/s45/s53/s49/s48/s45/s51/s49/s48/s45/s49 /s40/s98/s41/s32 /s32/s73 /s84/s77/s59 /s110 /s40/s97/s41 /s47/s50 /s40/s71/s72/s122/s41 /s47/s50 /s40/s71/s72/s122/s41/s110 /s51/s32 /s32/s73 /s84/s69/s59 /s110/s110 /s48 /s110 /s50/s110 /s49 /s110 /s49 /s110 /s51 /s110 /s50 FIG. 5: Polarization-converting (a) and polarization-con serving (b) contributions to the spectrum of the figure 4(a). The peak in (a) indicated by the arrow corre sponds to the resonant transition when accomplished by absorption of three magnons. It is interesting to note that the triple-resonance condition can be accomplished by many- magnon absorption processes as well ( mΩ/2π= ∆f), provided that the number of magnons, m, isoddinordertochangethepolarizationstateofthephoton, fro mTMtoTE, asrequired in our case. We recall that our method of calculation is not restricte d to the first-order Born approximation and thus it can describe nonlinear effects that are us ually relatively weak. Forexample, such a three-magnonabsorptionprocess ismanifest ed asa peakintheintensity of then=−3 outgoing beam, for Ω /2π= ∆f/3≈3.2 GHz, as pointed out by the arrow in figure 5(a). As can be seen in figure 4(a), when dissipative losses are taken into a ccount, the elastic beamintensity is uniformlyby about 30%, inagreement withthe result s shown infigure 2(b) for the TM mode. Here, when the triple-resonance condition is satis fied, the corresponding drop in the n=−1 beam is considerably larger because of the longer lifetime of the fina l (TE) state but, nonetheless, the optical conversion efficiency is s till as high as 5%. We note that, because of the high quality factor of the final (TE) sta te and the presence of non-negligible losses in this case, we overall obtain resonant optical absorption (instead of gain in the lossless case), as shown in figure 4(c). 11V. CONCLUSIONS To conclude, we have presented a detailed analysis and optimization o f a planar opto- magnonic structure operating in the triple-resonance regime and a llowing for optical con- version efficiencies of the order of 5% [cf. figure 4(a)] under realist ic conditions, mediated by a uniformly precessing spin wave. The outlined time Floquet multiple-sc attering methodol- ogy was able to resolve absorption and emission of multiple magnons, in dicating that under special conditions the attained conversion efficiencies mediated by m ultiple magnons can be comparable to those mediated by a single magnon [cf. orange and pink dotted lines in fig- ure 5(a)]. We have also found that the absorption or emission of a ma gnon leads to a change in the polarization of the optical conversion process. An interestin g further objective would be to extend the current approach to the full spatio-temporal Floquet scattering-matrix methodology, which should allow for investigating, among others, su rface Dammon-Eshbach and backward volume waves with an in-plane propagation wave vecto r that can lead to more exotic physical behavior, including emergence of a paraxial ou tgoing scattered beam and bandgap formation. Acknowledgments P.A.P. was supported by the General Secretariat for Research an d Technology (GSRT) andtheHellenic FoundationforResearch andInnovation(HFRI) th rougha PhD scholarship (No. 906). K.L.T., E.A., and G.P.Z. were supported by HFRI and GSRT un der Grant 1819. References ∗Electronic address: pepantaz@phys.uoa.gr 1Tabuchi Y, Ishino S, Noguchi A, Ishikawa T, Yamazaki R, Usami K and Nakamura Y 2015 Coherent coupling between a ferromagnetic magnon and a supe rconducting qubit Science349, 405408 122Lachance-QuirionD,TabuchiY,IshinoY,NoguchiA,Ishikaw aT,Yamazaki RandNakamuraY 2017Resolvingquantaofcollective spinexcitations inami llimeter-sized ferromagnet Sci. Adv. 3, e1603150 3Osada A, Hisatomi R, Noguchi A, Tabuchi Y, Yamazaki R, Usami k , Sadgrove M, Yalla R, Nomura M and Nakamura Y 2016 Cavity optomagnonics with spin- orbit coupled photons Phys. Rev. Lett. 116, 223601 4Zhang X, Zhu N, Zou C.-L. and Tang H X 2016 Optomagnonic whispe ring gallery microres- onatorsPhys. Rev. Lett. 117, 123605 5Haigh J A, Nunnenkamp A, Ramsay A J and Ferguson A J 2016 Triple -resonant Brillouin light scattering in magneto-optical cavities Phys. Rev. Lett. 117, 133602 6Viola-Kusminskiy S, Tang H X and Marquard F 2016 Coupled spin -light dynamics in cavity optomagnonics Phys. Rev. A 94, 033821 7Sharma S, Blanter Y M and Bauer G E W 2017 Light scattering by ma gnons in whispering gallery mode cavities Phys. Rev. B 96, 094412 8Haigh J A, Lambert N J, SharmaS, Blanter Y M, Bauer G E W and Rams ay A J 2018 Selection rules for cavity-enhanced Brillouin light scattering from magnetostatic modes Phys. Rev. B 97, 214423 9Osada A, Gloppe A, Hisatomi R, Noguchi A, Yamazaki R, Nomura M , Nakamura Y and Usami K 2018 Brillouin light scattering by magnetic quasivo rtices in cavity optomagnon- icsPhys. Rev. Lett. 120, 133602 10Osada A, Gloppe A, Nakamura Y and Usami K 2018 Orbital angular momentum conservation in Brillouin light scattering withina ferromagnetic spher eNew J. Phys. 20, 103018 11Almpanis E 2018 Dielectric magnetic microparticles as phot omagnonic cavities: Enhancing the modulation of near-infrared light by spin waves Phys. Rev. B 97, 184406 12Graf J, Pfeifer H, Marquardt F, and Viola Kusminskiy S 2018 Ca vity optomagnonics with magnetic textures: Coupling a magnetic vortex to light Phys. Rev. B 98, 241406(R) 13Kostylev M and Stashkevich A A 2019 Proposal for a microwave p hoton to optical photon converter based on traveling magnons in thin magnetic films J. Magn. Magn. Mat. 484, 329– 344 14Liu T Y, Zhang X F, Tang H X and Flatt´ e M E 2016 Optomagnonics in magnetic solids Phys. Rev. B 94, 060405(R) 1315PantazopoulosPA,StefanouN,AlmpanisEandPapanikolaou N 2017Photomagnonicnanocav- ities for strong light–spin-wave interaction Phys. Rev. B 96, 104425 16Pantazopoulos P A and Stefanou N 2019 Layered optomagnonic s tructures: Time Floquet scattering-matrix approach Phys. Rev. B 99, 144415 17Pantazopoulos P A, Papanikolaou N and Stefanou N 2019 Tailor ing coupling between light and spin waves with dual photonic-magnonic resonant layered st ructures J. Opt.21, 015603 18ZvezdinAKandKotov VA1997 Modern Magnetooptics and Magnetooptical Materials (Bristol: Institute of Physics Publishing) 19Onbasli M C, Beran L, Zahradn´ ık M, Kuˇ cera M, Antoˇ s R, Mistr ´ ık J, Dionne G F, Veis M and Ross K A 2016 Optical and magneto-optical behavior of Cerium Yttrium Iron Garnet thin films at wavelengths of 200–1770 nm Sci. Rep. 6, 23640 20Stancil D D and Prabhakar A 2009 Spin Waves-Theory and Applications (Boston: Springer) 21Pierce D T and Spicer W E 1972 Electronic structure of amorpho us Si from photoemission and optical studies Phys. Rev. B 5, 307 22Gao L, Lemarchand F and Lequime M 2012 Exploitation of multip le incidences spectrometric measurements for thin film reverse engineering Opt. Express 20, 15734–15751 23Stefanou N, Yannopapas V and Modinos A 1998 Heterostructure s of photonic crystals: fre- quency bands and transmission coefficients Comput. Phys. Commun. 113, 49–77 24Stefanou N, Yannopapas V and Modinos A 2000 MULTEM 2: A new ver sion of the program for transmissionandband-structurecalculations ofphotonic crystalsComput. Phys. Commun. 132, 189–196 14 | 2019-05-17 | Optomagnonic cavities have recently been emerging as promising candidates for
implementing coherent microwave-to-optical conversion, quantum memories and
devices, and next generation quantum networks. A key challenge in the design of
such cavities is the attainment of high efficiencies, which could, e.g., be
exploited for efficient optical interfacing of superconducting qubits, as well
as the practicality of the final designs, which ideally should be planar and
amenable to on-chip integration. Here, on the basis of a novel time Floquet
scattering-matrix approach, we report on the design and optimization of a
planar, multilayer optomagnonic cavity, incorporating a Ce:YIG thin film,
magnetized in-plane, operating in the triple-resonant inelastic light
scattering regime. This architecture allows for conversion efficiencies of
about 5%, under realistic conditions, which is orders of magnitude higher than
alternative designs. Our results suggest a viable way forward for realizing
practical information inter-conversion between microwave photons and optical
photons, mediated by magnons, with efficiencies intrinsically greater than
those achieved in optomechanics and alternative related technologies, as well
as a platform for fundamental studies of classical and quantum dynamics in
magnetic solids, and implementation of futuristic quantum devices. | High-efficiency triple-resonant inelastic light scattering in planar optomagnonic cavities | 1905.07278v1 |
1 Efficient g eometrical control of spin waves in microscopic YIG waveguides S. R. Lake1 , B. Divinskiy2*, G. Schmidt1,3, S. O. Demokritov2, and V . E. Demidov2 1Institut für Physik, Martin -Luther -Universität Halle -Wittenberg, 06120 Halle, Germany 2Institute for Applied Physics, University of Muenster, 48149 Muenster, Germany 3Interdisziplinäres Zentrum für Materialwissenschaften , Martin -Luther -Universität Halle - Wittenberg, 06120 Halle, Germany We study experimentally and by micromagnetic simulations the propagation of spin waves in 100-nm thick YIG waveguides , where the width linearly decreases from 2 to 0.5 m over a transition region with varying length between 2.5 and 10 m. We show that this geometry result s in a down conversion of the wavelength , enabling efficient generation of waves with wavelengths down to 350 nm. We also find that th is geometry leads to a modification of the group velocity , allow ing for almost -dispersionless propagation of spin -wave pulses . Moreover , we demonstrate that the in fluence of energy concentration outweighs that of damping in these YIG waveguides , resulting in an overall increase of the spin -wave intensity during propagation in the transition region. These findings can be utilized to improve the efficiency and functio nality of magnonic devices which use spin waves as an information carrier. *Corresponding author, e -mail: b_divi01@uni -muenster.de 2 Spin waves propagating in microscopic magnetic waveguides present a flexib le and highly functional tool for transmission and processing of information on the nano -scale1-4. Among the most important advantages provided by spin waves is their controllabili ty by the magnetic field, which , for example, enables efficient control of t heir propagation characteristics by electric current s. This controllability also forms the basis of using spatially non -uniform , dipolar magnetic fields to manipulate spin waves5-7. Because these fields are determined by the waveguide’s geometry , varying its spatial parameters enable s different mode transformations and wavelength conversion7-15. Although tuning spin waves by using geometrical effects provides many opportunities for the implementation of magnonic devices, the functionality of this approach is strongly limited by the spatial attenuation. Indeed, in metallic waveguides with a small decay length, the passage through a conversion region can lead to a massive loss of spin-wave intensity16. The restrictions imposed by the fast spatial decay of spin waves can be overcome by using high-quality , nanometer s-thick films of the low-damping magnetic insulator , yttrium iron garnet (YIG)17-19, where the decay length of spin waves can surpass many tens of micrometers20-22. Recently it was shown that these films can be structured on the micrometer and the sub -micrometer scale without significantly increasing the magnetic damping23-25. Additionally, magnetic dynamics in these films can be controlled by sp in-torque effects, which can be used to enhance further the propagation characteristics26,27 and generate propagating spin waves by dc electric currents without the need to use energy -inefficient microwave excitation28. These features make ultrathin YIG fi lms an excellent candidate for magnonic applications where spin waves are steered via geometrical parameters. In this Letter , we study the control of spin -wave propagation characteristics in microscopic , ultrathin -YIG waveguides in which the width linearly decreases along the 3 propagation direction . By using spatially -, temporally -, and phase -resolved measurements and micromagnetic simulations , we show that the spatial variation of the demagnetizing field caused by the narrowing of the waveguide results in a robust decrease of the spin-wave wavelength. Due to the minimal spatial attenuation, this wavelength conversion occurs without the decrease of the spin-wave intensity during propagation in the transition region. On the contrary , due to spatial compression, the intensity exhibits a noticeable increase, which becomes particularly pronounced for shorter transition lengths . These effects can be utilized to implement highly efficient excitation of short -wavelength spin waves . Additionally, we show that the geometrical control can be used to tune the propagation velocity of spin -wave pulses and reach a regime where the velocity is almost independent of the spin -wave frequency . Our findings demonstrate a simple and robust method to control spin -wave propag ation which can enhance the functionality of nano scale magnonic devices. Figure 1(a) shows the schematics of our experiment. We study a microscopic spin-wave waveguide patterned from a 100-nm thick YIG film grown by pulsed -laser deposition (PLD) . The YIG film is characterized by a saturation magnetization of 4πM = 1.75 kG and a Gilbert damping constant α = 4×10-4, as determined from ferromagnetic -resonance measurements. The width of the waveguide, w, linearly decreases from 2 m to 0.5 m over a 10 - m long transition region. The spin waves are excited by using a 500 -nm wide and 150 -nm thick inductive Au antenna perpendicular to the waveguide, with its right -hand edge located at the beginning of the transition region. The structures were patterned on a GGG <111> substrate using a two -layer PMMA resist and subsequent electron beam lithography. After development in isopropanol , 110 nm of YIG was deposited via PLD, following a recipe published by Hauser et al. (Ref. 19). The sample was then 4 placed in aceto ne for lift -off of extraneous material and afterwards annealed in a pure oxygen atmosphere19. Next, 10 nm of YIG were etched using phosphoric acid in order to remove seams that can appear at the edges of the structures due to the mobility of the deposited atoms during PLD. Finally, the overlying antenna was patterned using a tri-layer PMMA resist, evaporation of Ti (10 nm) and Au (150 nm), and lift -off. The YIG waveguide is magnetized to saturation by an in-plane , static magnetic field , H0, applied along the Au antenna . Because of demagnetization effects, the internal magneti c field Hint is smaller than H0. It is not uniform across the waveguide width and strongly differs in the waveguide’s wide st and narrow est parts (see the distribution in Fig. 1(b) calculated by using the micromagnetic simulation software MuMax3 (Ref. 29)). At H0 = 1000 Oe, the maximum internal field is 945 and 785 Oe in the wide st and the narrow est part, respectively. As seen from Fig. 1(c), this difference results in a shift of approx imately 0.7 to 0.8 GHz in the dispersion curves . We note that the dispersion curves calculated by using MuMax3 (curves in Fig. 1(c)) coincide well with those obtained from phase -resolved measurements (symbols in Fig. 1(c)) described in detail below. This good agreement allows us to rely on results of simulations to obtain the information about the propagation of spin waves which cannot be obtained from direct measurements. To analyze the propagation of spin waves experimentally , we use the time- and phase - resolved micro -focus Brillouin light scattering (BLS) spectroscopy16. We focus the probing laser light with the wavelength of 473 nm and a power of 0.25 mW into a diffraction -limited spot on the surface of the YIG waveguide (see Fig. 1(a)) and analyze the light inelastically scattered from spin waves. The measured signal , or BLS intensity , is proportional to the intensity of spin waves at the position of the probing spot, which allows us to record two-dimensional spin-wave intensity 5 maps. Additionally, by using the interference of the scattered light with the probing light modulated by the microwave excitation signal, we measure the spatial maps of cos( ), where is the phase difference between the spin wave and the signal applied to the antenna. The Fourier analysis of the se maps provides direct information about the wavelength of spin waves at a given excitatio n frequency . Figures 2(a) and 2(b) show representative phase and intensity maps recorded at the excitation frequency f=4.5 GHz. The left edge of the maps corresponds to the position x=0.5 m which is selected to avoid measuring where the probing light is partially blocked by the antenna . The data of Fig. 2(a) indicate that the w avelength of spin waves gradually decreases during propagation in the transition region, reflecting the frequency shift in the dispersion spectrum due to the continuous reduction of the internal static magnetic field as seen in Figs. 1(b) and 1(c). We note that the phase profiles are slightly disturbed in the vicinity of the antenna, which is caused by the weak excitation of higher -order transverse waveguide mode s16. The slight periodic transverse modulation of the intensity distribution seen in Fig. 2(b) also demonstrates this effect . Analysis of the experimental maps shows that in the transition region the wavelength of the spin wave decreases from about 4 m to 0.5 m, i.e., by a factor of 8, (point -down triangles in Fig. 2(c)). This is in quantitative agreement with the results obtained from micromagnetic simulations ( point -up triangles in Fig. 2(c)). The wavelength -conversion process is characterized thoroughly in Fig. 2(d), which shows the spin -wave wavelength at the end of the transition region , OUT, as a function of the wavelength of the spin wave excited by the antenna , EXC. We note that the efficiency of the inductive excitation by the 500-nm wide antenna quickly decreases for waves with wavelength s smaller than 1 m (Ref. 16), limiting the interval of EXC accessible in the experiment . This restriction is a significant drawback of the inductive excitation mechanism, 6 which strongly limits its use in magnonic devices operating with short -wavelength spin waves. As seen from the da ta of Fig. 2(d), the observed conversion of the wavelength allows one to extend the range of usable wavelengths down to 350 nm. From the point of view of technical applications, the d emonstrated wavelength conversion is advantageous only if it is not acco mpanied by a strong decrease of the intensity of the spin wave in the transition region. On one hand, one expects a spatial decrease of the intensity due to the damping and/or wave reflections. On the other hand, the narrowing of the waveguide is expected to result in an increase of the wave’s intensity due to its energy being concentrated into a smaller cross section. To prove which of these mechanisms dominate s in the studied waveguide , we analyze spatial dependencies of the spin -wave intensity obtained from the measurements and micromagnetic simulations (Fig. 3(a)) . The experimental curve in Fig. 3(a) exhibits an almost constant intensity in the interval x=0.5-7 m. This indicates that the energy -concentration effect approximately compensates the effect s of the damping. However, at larger x, the intensity quickly decreases . We emphasize that this observation cannot be related to wave reflection s because the intensity profile shows no signature of the formation of a standing wave. We also note that the observed decrease is not reproduced in the micromagnetic simulations. The calculated intensity coincides well with the experimental one in the range x=0.5-7 m. However, contrary to experimental data, the calculated intensity noticeably increases in the vicinity of the end of the transition region. We associate this discrepancy with the wavelength -dependent sensitivity of the measurement apparatus. Indeed, the sensitivity of magneto -optical techniques is known to decrease with decreasing wavelength of spin waves , and vanish when the latter becomes equal to the diameter of the probing light spot, d. Assuming d=0.3 m and decreasing from 4 to 0.5 m, one can estimate that the experimental sensitivity decrease s approximately by a factor of 4. This is in 7 good agreement with the ratio between the calculated and experimental intensities at x>10 m in Fig. 3(a) . Taking this into account , we base our further analysis on the results of simulations. Considering the calculated intensity curve (Fig. 3(a)) and comparing the intensit y of spin waves at the beginning of the transition region with the intensity at a point 0.5 m beyond the end of this region, we conclude that the conversion process is accompanied by an increase of the spin - wave intens ity by approximately a factor of 1.5. This result clearly proves the suitability of the proposed conversion mechanism for practical applications. Additionally, as shown by the data of Fig. 3(b), the intensity enhancement can be further improved by reducing the length of the transition region. Note here, that this reduction also leads to stronger reflections of the wave at the end of the transition, as seen from the increasing intensity drop at that point (marked by arrows in Fig. 3(b)) . However, this advers e effect does not compromise the overall increase of the intensity enhancement (see the inset in Fig. 3(b)). Finally, we analyze the effects of the spatial variation of the waveguide geometry on the propagation of spin -wave pulses. As can be seen in Fig. 1(c), the reduction of the waveguide width affects not only the wavelength of a spin wave at a given frequency, but also the slope of the dispersion curve, which determines the group velocity. In other words, during the propagation in the transition region, the dispersion of a spin-wave pulse is expected to change as well. In order to address this phenomenon, we perform time-resolved BLS measurements using an excitation signal in the form of 20 -ns long pulses and determin e the temporal delay of the spin-wave pulses at different spatial positions. In agreement with the above arguments, the found dependence of the propagation delay (Fig. 4(a)) is not linear within the transition region and clearly exhibits a gradual decelera tion of the pulse. The observed deceleration can be used, for example, to implement controllable compression of the spin -wave pulses in the space domain. 8 From the local slope of the dependence shown in Fig. 4(a) , we determine the spin-wave group velocities at the beginning of the transition region ( x=0.5 m) and at its end ( x=10 m). Figure 4(b) summarizes the se results obtained at different excitation frequencies. These data show that the initial -stage group velocity depends strongly on the carrier frequen cy, f, while the velocity at x=10 m is almost independent of frequency. The latter observation indicates that the spin -wave pulses experience nearly dispersionless propagation in the narrow ( w=0.5 m) waveguide30. In narrow waveguides, the spectral region where the group velocity has weak frequency dependence corresponds to spin waves with relatively short wavelengths30. These are difficult to excite by an inductive antenna but, by the down conversion presented here , can be easily achieved . The data of Fig. 4(b) show that the demonstrated approach makes it possible to achieve propagation of spin - wave pulses over longer distances without pulse broadening by dispersion effects . In conclusion, we show that the geometrical control in ultrathin -YIG waveguides p rovides practical opportunities for spin-wave manipulation s, such as the down conversion of the wavelength and the tuning of the propagation velocity. We demonstrate that the intensity of spin waves can be maintained while passing through the control region , due to the small damping in YIG, and can in fact noticeably increase due to spatial compression. These findings can be used for implementation of energy -efficient magnonic devices that exploits sub-micrometer wavelengths. This work was supported in part by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – Project -ID 433682494 – SFB 1459 and TRR227 TP B02. 9 Data availability The data that support the findings of this study are available from the corresponding author upon reasonable request. References 1. S. Neusser and D. Grundler, Adv. Mater. 21, 2927 (2009). 2. V. V. Kruglyak, S. O. Demokritov, and D. Grundler, J. Phys. D: Appl. Phys. 43, 264001 (2010). 3. B. Lenk, H. Ulrichs, F. Garbs, and M. Münzenberg, Phys. Rep. 507, 107–136 (2011). 4. A. V. Chumak, V. I. Vasyuchka, A. A. Serga, and B. Hillebrands, Nat. Phys. 11, 453 (2015). 5. V. E. Demidov, S. O. Demokritov, K. Rott, P. Krzysteczko, and G. Reiss , Appl. Phys. Lett. 92, 232503 (2008). 6. J. Topp, J. Podbielski, D. Heitmann, and D. Grundler, Phys. Rev. B 78, 024431 (2008). 7. V. E. Demidov, J. Jersch, S.O. Demokritov, K. Rott, P. Krzysteczko, and G. Reiss , Phys. Rev. B 79, 054417 (2009). 8. V. E. Demidov, M. P. Kostylev , K. Rott, J. Münchenberger, G. Reiss, and S. O. Demokritov, Appl. Phys. Lett. 99, 082507 (2011). 9. K. Vogt, H. Schultheiss, S. Jain, J. E. Pearson, A. Hoffmann, S. D. Bader, and B. Hillebrands, Appl. Phys. Lett. 101, 042410 (2012). 10. C. S. Davies, A. Francis, A. V. Sadovnikov, S. V. Chertopalov, M. T. Bryan, S. V. Grishin, D. A. Allwood, Y. P. Sharaevskii, S. A. Nikitov, and V. V. Kruglyak, Phys. Rev. B 92, 020408(R) (2015). 10 11. Z. Zhang, M. Vogel, J. Holanda, M. B. Jungfleisch, C. Liu, Y. Li, J. E. Pearson, R. Diva n, W. Zhang, A. Hoffmann, Y. Nie, and V. Novosad, Appl. Phys. Lett. 115, 232402 (2019). 12. D. V. Kalyabin, A. V. Sadovnikov, E. N. Beginin, and S. A. Nikitov , J. Appl. Phys. 126, 173907 (2019). 13. S. Mieszczak, O. Busel, P. Gruszecki, A. N. Kuchko, J. W. Kłos, and M. Krawczyk, Phys. Rev. Applied 13, 054038 (2020 ). 14. A. Haldar and A . O. Adeyeye , Appl. Phys. Lett. 119, 060501 (2021) . 15. H. Yu, J. Xiao , and H. Schultheiss , Physics Reports , 905, 1 (2021). 16. V. E. Demidov and S. O. Demokritov, IEEE Trans. Mag. 51, 0800215 (2015). 17. Y. Sun, Y. Y. Song, H. Chang, M. Kabatek, M. Jantz, W. Schneider, M. Wu, H. Schultheiss, and A. Hoffmann, Appl. Phys. Lett. 101, 152405 (2012). 18. O. d’Allivy Kelly, A. Anane, R. Bernard, J. Ben Youssef, C. Hahn, A. H. Molpeceres, C. Carretero, E. Jacquet, C. Deranlot, P. Bortolotti, R. Lebourgeois, J. -C. Mage, G. de Loubens, O. Klein, V. Cros, and A. Fert, Appl. Phys. Lett. 103, 082408 (2013). 19. C. Hauser, T. Richter, N. Homonnay, C. Eisenschmidt, M. Qaid, H. Deniz, D. Hesse, M. Sawicki, S. G. Ebbinghaus, and G. Schmidt, Sci. Rep. 6, 20827 (2016). 20. H. Yu, O. d’Allivy Kelly, V. Cros, R. Bernard, P. Bortolotti, A. Anane, F. Brandl, R. Huber, I. Stasinopoulos, and D. Grundler, Sci. Rep. 4, 6848 (2014). 21. M. Collet, O. Gladii, M. Evelt, V. Bessonov, L. Soumah, P. Bortolotti, S. O. Demokritov, Y. Henry, V. Cros, M. Bailleul, V. E. Demidov, and A. Anane, Appl. Phys. Lett. 110, 092408 (2017). 11 22. C. Liu, J. Chen, T. Liu, F. Heimbach, H. Yu, Y. Xiao, J. Hu, M. Liu, H. Chang, T.Stueckler, S. Tu, Y. Zhang, Y. Zhang, P. Gao, Z. Liao, D. Yu, K. Xia, N. Lei, W. Zhao, and M. Wu, Nat. Commun. 9, 738 (2018). 23. S. Li, W. Zhang, J. Ding, J. E. Pearson, V. Novosad, and A. Hoffmann, Nanoscale 8, 388 (2016). 24. B. Heinz, T. Br ächer, M. Schneider, Q. Wang, B. L ägel, A. M. Friedel, D. Breitbach, S. Steinert, T. Meyer, M. Kewenig, C. Dubs, P. Pirro, and A. V. Chumak, Nano Letters 20, 4220−4227 (2020). 25. G. Schmidt , C. Hauser, P. Trempler, M. Paleschke, and E.T. Papaioannou, Phys. Stat. Sol. B 257, 1900644 (2020). 26. M. Evelt, V. E. Demidov, V. Bessonov, S. O. Demokritov, J. L. Prieto, M. Munoz, J. Ben Youssef, V. V. Naletov, G. de Loubens, O. Klein, M. Collet, K. Garcia -Hernandez, P. Bortolotti, V. Cros, and A. Anane, Appl. Phys. Lett. 108, 172406 (2016). 27. T. Wimmer, M. Althammer, L. Liensberger, N. Vlietstra, S. Geprägs, M. Weiler, R. Gross, and H. Huebl , Phys. Rev. Lett. 123, 257201 (2019). 28. M. Evelt, L. Soumah, A. B. Rinkevich, S. O. Demokritov, A. Anane, V. Cros, J. Ben Youssef, G. de Loubens, O. Klein, P. Bortolotti, and V. E. Demidov, Phys. Rev. Appl. 10, 041002 (2018). 29. A. Vanstee nkiste, J. Leliaert, M. Dvornik, M. Helsen, F. Garcia -Sanchez, and B. Van Waeyenberge, AIP Adv. 4, 107133 (2014). 30. B. Divinskiy, H. Merbouche, K. O. Nikolaev, S. Michaelis de Vasconcellos, R. Bratschitsch, D. Gouéré, R. Lebrun, V. Cros, J. Ben Youssef, P. Bortolotti, A. Anane, S. O. Demokritov, and V. E. Demidov, Phys. Rev. Appl. 16, 024028 (2021). 12 FIG. 1. (a) Schematics of the experiment. Inset shows the scanning -electron micrograph of the sample recorded under a n angle of 70°. (b) Distribution of the internal static magnetic field in the waveguide calculated by using micromagnetic simulations. (c) Dispersion curves of spin waves in the wide st (w=2 m) and the narrow est (w=0.5 m) parts of the waveguide obtained from micromagnetic si mulations (curves) and from phase -resolved measurements (symbols). The data are obtained at H0=1000 Oe. 13 FIG. 2. Representative maps of the spin -wave phase (a) and intensity (b) recorded by BLS at the excitation frequency f=4.5 GHz. The left edge of the maps corresponds to the displacement x=0.5 m from the edge of the antenna. (c) Spatial dependence of the wavelength of the spin wave with the frequency f=4.5 GHz . Vertical dashed line marks the end of the transition region. (d) Spin - wave wavelength at the end of the transition region , OUT, as a function of the wavelength of the spin wave excited by the antenna , EXC. Dashed curve – guide for the eye. In ( c) and ( d): point - down triangles – experimental data, point -up triangl es – results of micromagnetic simulations. The data are obtained at H0=1000 Oe. 14 FIG. 3. (a) Spatial dependences of the spin -wave intensity integrated over the width of the waveguide obtained from the measurements and micromagnetic simulations, as labelled. Vertical dashed line marks the end of the transition region. (b) Spatial dependence of the spin -wave intensity calculated for the waveguides with the transition length of 10 and 5 m, as labelled. Arrows mark the intensity drop du e to reflections at the end of the transition region. Inset shows the ratio between the intensity detected at a point 0 .5 m beyond the end of the transition region and the intensity detected at its beginning as a function of the transition length. The dat a are obtained at H0=1000 Oe and f=4.5 GHz . 15 FIG. 4. Spatial dependence of the propagation delay measured for a 20 -ns long spin -wave pulse at the carrier frequency f=4.5 GHz . Vertical dashed line marks the end of the transition region. Frequency dependence of the group velocity at the beginning of the transition region ( x=0.5 m) and at its end ( x=10 m), as labelled. Symbols – experimental data. Curves – guide for the eye. The data are obtained at H0=1000 Oe. | 2021-11-03 | We study experimentally and by micromagnetic simulations the propagation of
spin waves in 100-nm thick YIG waveguides, where the width linearly decreases
from 2 to 0.5 micrometers over a transition region with varying length between
2.5 and 10 micrometers. We show that this geometry results in a down-conversion
of the wavelength, enabling efficient generation of waves with wavelengths down
to 350 nm. We also find that this geometry leads to a modification of the group
velocity, allowing for almost-dispersionless propagation of spin-wave pulses.
Moreover, we demonstrate that the influence of energy concentration outweighs
that of damping in these YIG waveguides, resulting in an overall increase of
the spin-wave intensity during propagation in the transition region. These
findings can be utilized to improve the efficiency and functionality of
magnonic devices which use spin waves as an information carrier. | Efficient geometrical control of spin waves in microscopic YIG waveguides | 2111.02236v1 |
Tunable sign change of spin Hall magnetoresistance in Pt/NiO/YIG structures Dazhi Hou,1, 2Zhiyong Qiu,1, 2,Joseph Barker,3Koji Sato,1Kei Yamamoto,3, 4, 5Sa ul V elez,6Juan M. Gomez-Perez,6Luis E. Hueso,6, 7F elix Casanova,6, 7and Eiji Saitoh1, 2, 3, 8 1WPI Advanced Institute for Materials Research, Tohoku University, Sendai 980-8577, Japan 2Spin Quantum Rectication Project, ERATO, Japan Science and Technology Agency, Sendai 980-8577, Japan 3Institute for Materials Research, Tohoku University, Sendai 980-8577, Japan 4Institut f ur Physik, Johannes Gutenberg Universit at Mainz, D-55099 Mainz, Germany 5Department of Physics, Kobe University, 1-1 Rokkodai, Kobe 657-8501, Japan 6CIC nanoGUNE, 20018 Donostia-San Sebastian, Basque Country, Spain 7IKERBASQUE, Basque Foundation for Science, 48011 Bilbao, Basque Country, Spain 8Advanced Science Research Center, Japan Atomic Energy Agency, Tokai 319-1195, Japan Spin Hall magnetoresistance (SMR) has been investigated in Pt/NiO/YIG structures in a wide range of temperature and NiO thickness. The SMR shows a negative sign below a temperature which increases with the NiO thickness. This is contrary to a conventional SMR theory picture applied to Pt/YIG bilayer which always predicts a positive SMR. The negative SMR is found to presist even when NiO blocks the spin transmission between Pt and YIG, indicating it is governed by the spin current response of NiO layer. We explain the negative SMR by the NiO 'spin-
op' coupled with YIG, which can be overridden at higher temperature by positive SMR contribution from YIG. This highlights the role of magnetic structure in antiferromagnets for transport of pure spin current in multilayers. Magnetoresistance plays essential roles in providing both a fundamental understanding of electron transport in magnetic materials and in various technological ap- plications. Anisotropic magnetoresistance (AMR) [1, 2], giant magnetoresistance [3, 4], and tunneling magnetore- sistance [5{8] underpin technologies in sensors, memo- ries, and data storage. Recent studies of thin lm bi- layer systems comprised of a normal metal (NM) and a ferromagnetic insulator (FI) revealed a new type of mag- netoresistance called spin Hall magnetoresistance (SMR) [9{11], originating from the interplay between the spin accumulation at the NM/FI interface and the magnetiza- tion of the FI layer. When the NM layer has a signicant spin-orbit interaction, e.g. Pt, an in-plane charge current jcinduces a spin current via the spin Hall eect, which in turn generates a spin accumulation near the NM/FI interface. At the same time, this spin accumulation is aected by the orientation of the magnetization in the ferromagnet. The conductivity of the NM layer is thus subject to a magnetization dependent modication to the leading order in 2 SHE, whereSHEis the spin Hall angle in the NM layer. Since the discovery of SMR, experimental studies were instigated in various systems [12{19]. The amplitude of SMR is dened as the dierence of the resistivities with an applied eld, H, parallel (k) and perpendicular ( ?) tojc:SMR=k ?. This is predicted to be always positive because when Hkjc, the FI can absorb more spin current, by which the back
ow required to ensure the stationary state is reduced at the FI/NM interface, in turn causing less secondary forward charge current, and therefore gives : k> ?[9, 10]. Positive SMR is found in most experimental observations. Very recently, a negative SMR ( k<?) was reportedwhen an antiferromagnetic (AFM) insulator, in this case NiO, is inserted between Pt and YIG [20]. The negative SMR was also found to revert to the conventional positive sign at higher temperatures. Signal contamination from other magnetoresistances such as AMR was excluded by a systematic eld angle dependence measurement. This re- sult challenges the present understanding of SMR. Since the SMR does not change its sign in the Pt/YIG bilayer structure, the NiO layer must be the cause. However, it is not clear why NiO should give a negative SMR since antiferromagnets are thought only to aect the eciency of the spin communication between Pt and YIG [21{26]. In this letter, we report the temperature dependence of SMR in Pt/NiO/YIG structures with dierent thick- nesses of NiO. The temperature at which the SMR be- comes negative is found to depend on the NiO thickness. The anomalous negative SMR at low temperatures is ex- plained from a `spin-
op' conguration whereby the N eel order of the NiO is perpendicularly coupled to the mag- netization of YIG [27]. As the spin conductivity of NiO increases with increasing temperature [24{26], the mo- ments of the YIG beneath have an increasing in
uence on the total SMR signal. The positive SMR contribution from YIG competes the negative SMR from NiO. At the sign change, the competition leads to a vanishing SMR. Above, in the high temperature regime, the positive SMR of the YIG dominates. We introduce a phenomenological model to describe the competition between the positive and negative SMR contributions, which reproduces the NiO thickness dependent SMR sign change behaviors in Pt/NiO/YIG. An epitaxial YIG lm of thickness 3 m was grown on a gadolinium gallium garnet (111) substrate prepared by the liquid phase epitaxy. NiO lms of dierent thick-arXiv:1610.07362v2 [cond-mat.mes-hall] 8 Mar 20172 30 40 50 60 70 80 YIG(444) NiO(222)Intensity (a.u.) 2θ (deg)NiO(111) Pt NiO YIG 3 nm FIG. 1. X-ray diraction patterns of a 50 nm NiO lm on YIG(111). Inset shows the cross section TEM photo for a Pt/NiO/YIG trilayer measured in the transport experiment. nesses were grown by sputtering onto the YIG at 400C. The lm was then covered with 4 nm of sputtered Pt. The X-ray diraction patterns of a 50 nm NiO lm on YIG is plotted in Fig. 1, which only shows (111) and (222) NiO peaks of narrow line width. It suggests that the NiO lm is of high crystallinity and a (111) preferred orienta- tion. The inset in Figure 1 shows a representative cross- section TEM picture for a Pt/NiO/YIG sample, which conrms a good thickness uniformity and clean interface. Figure 2(a) shows the illustration of the magnetore- sistance (MR) measurement setup and the denition of magnetic eld angles. Standard four-probe method is employed for the MR observation at current density 108A/m2, and MR can be detected either by sweep- ingHalong a xed direction or by rotating Hof the same magnitude [9]. Figure 2(b) shows the MR measured by Hsweeping in a Pt/NiO(2.5 nm)/YIG sample at eld angle= 0for various temperatures. The range of magnetic eld over which the magnetoresistance occurs, coincides with that of the switching process of YIG [28]. The MR data for T >140 K is consistent with the predic- tionk> ?of the SMR theory. When T=140 K, the MR nearly vanishes. For T <140 K, a sign change of MR is observed and the MR amplitude increases with decreas- ing temperature. The MR data from the same sample at eld angle= 90is plotted in Fig. 2(c), which shows the same feature of the sign change. The SMR ratio SMR=xxextracted from Fig. 2(b) and 2(c) are plot- ted in Fig. 2(d). Figure 2(e) and 2(f) show the eld angle dependence of resistance in Pt/NiO(2 nm)/YIG at 260 K and 20 K, which not only reproduces the MR sign change behaviour, but conrms the SMR-type eld angle depen- dence symmetry as well [20]. Thus, it looks reasonable to claim that SMR is the dominant contribution for the MR in Pt on NiO/YIG, since other mechanisms such as anisotropic magnetoresistance will cause a dierent eld angle dependence [29]. However, the sign change of the magnetoresistance in the low temperature regime seems to be at odds with SMR which, conventionally, can only be positive [10]. 0 100 200 300 400-40-2002040 α = 0˚ ΔSMR/xx (10-6) T (K)α = 90˚ Rxx (Ω)(d) 2e-5ΔSMR/2xx= 0˚ αt = 2.5 nm NiO -400 -200 0 200 400 H (Oe)(b) 0 -400 -200 0 200 40 H (Oe) 20 K40 K 60 K100 K140 K180 K220 K260 K300 K340 K 2e-5= 90˚ α(c) 20 K40 K 60 K100 K140 K180 K220 K260 K300 K340 K/ xx(H)xx(400 Oe) / xx(H)xx(400 Oe) (a) Hβ xz (a) γ x y,J z H e(a) Hαx y,J ez y,J e(a) 98.29898.29998.30098.30198.302Rxx (Ω) 62.305262.305662.306062.3064 α β γ -90 0 90 180 270 α, β, γ (deg ) 20K260Kα β γ -90 0 90 180 270 α, β, γ (deg )(e) (f) Pt NiO YIGFIG. 2. (a), The illustration for the magnetoresistance mea- surement setup for various magnetic eld ( H) orientations. , and
are the eld angles dening the Hdirections when H is applied in the x-y,x-zandy-zplanes, respectively. (b), (c), Magnetoresistance measured by Hsweeping for a Pt/NiO(2.5 nm)/YIG at = 0and 90for various temperatures. (d), Temperature dependence of the SMR ratio SMR=xxfor Pt/NiO(2.5 nm)/YIG at = 0and 90. (e), (f), Field an- gle dependent resistance measured for Pt/NiO(2 nm)/YIG at 260 K and 20 K with jHj= 20000 Oe, which shows positive and negative SMR, respectively. Fig. 3(a) shows the temperature dependence of the SMR ratio measured in Pt/NiO/YIG devices with dif- ferent NiO thicknesses, dNiO. The change in sign of the SMR occurs at higher temperatures in larger dNiOsam- ples. ThedNiOdependence proves to be a key piece of in- formation for understanding the negative SMR. Further- more, the SMR ratios have (positive) maxima at higher temperatures for thicker NiO samples. These dNiOde- pendent characteristics show a quantitative eect of the NiO on the SMR modulation, rather than a nuanced in- terface eect [30]. To gain further insight into the temperature depen- dence of spin transport in NiO, we carried out spin pump- ing measurements for the same samples, in which spin current is injected from YIG through NiO to generate a voltage in Pt via the inverse spin Hall eect (ISHE) [22]. The Pt/NiO/YIG device is placed on a coplanar waveguide which serves as a 5 GHz microwave source at 14 dbm, and the details of the experimental setup can be found elsewhere [24]. The ISHE voltage VISHE from all the samples is plotted against Tin Fig. 3(b), the be- haviour of which is very similar to the result we found in Pt/CoO/YIG [24]: spin transmission is nearly zero for3 0 100 200 300 400024680 100 200 300 400-20020406080 T (K) 2.0 nm 2.2 nm 2.5 nm 2.7 nm 4.0 nm 5.4 nm 7.0 nm 15 nm 30 nm T (K) ΔSMR/xx (10-6 ) VISHE (μV) 0 200 40001V V(b)(a) FIG. 3. (a), The SMR ratio measured in Pt/NiO( dNiO)/YIG devices with dierent NiO thickness dNiOat various temper- atures, which shows that the SMR sign change temperature is lower for a thinner NiO sample. The SMR ratio peak posi- tions are marked by arrows. Negative SMR at low tempera- tures can be observed for all the NiO thickness except dNiO= 30 nm. The dashed curves are the tting based on Eq. (2). (b),VISHE in Pt/NiO/YIG devices versus temperature from spin pumping measurement. The peak positions are marked by arrows, which are found to be close to the SMR ratio peak positions marked in Figure 2a. The inset shows the normal- izedVISHE temperature dependence. low temperature limit and increases with temperature to reach the maximum around the N eel point. At room temperature, VISHE shows a non-monotonic dNiOdepen- dence, which is consistent with previous result. Fig. 3(b) inset shows the normalized VISHE temperature depen- dence, in which the data for dNiO= 5.4 nm, 7 nm and 15 nm collapse into a single curve. This conrms that the VISHE is governed by the NiO spin conductivity, which shows the same Tdependence when NiO is thick enough to exhibit bulk property. For dNiO= 30 nm,VISHE is below our measurement sensitivity 5 nV. An important conclusion can be drawn by combining the results from SMR and spin pumping measurements: the negative SMR does not rely on the spin transmis- sion between Pt and YIG, because it reaches the largest magnitude for the lowest temperature at which NiO spin conductivity vanishes. This argument can be further sup-ported by the fact that the negative SMR is present even fordNiO= 15 nm, where the NiO spin conductivity is nearly zero throughout the entire temperature range. It indicates that the negative SMR is not caused by the magnetic moment of the YIG layer but that of the NiO layer, which is beyond any model based on spin commu- nication between YIG and Pt [10, 31]. Let us next provide an explanation for the negative SMR. The SMR in the trilayer system in this experiment is governed by the spin current through the Pt/NiO in- terface, which also re
ects the eect of the presence of the NiO/YIG interface. The sign change and the thick- ness dependent behavior can be understood by assum- ing a `spin-
op' coupling between NiO and YIG [27, 32], which means the antiferromagnetic axis (N eel vector unit nAFM) in NiO is perpendicular to the YIG magnetization unit vectormFIas illustrated in Fig. 4(a). Although a perpendicular coupling has not yet been conrmed ex- perimentally for NiO on YIG, spin-
op coupling between NiO and other ferromagnets is quite common and well understood[27, 33, 34]. For dNiObelow the domain wall width of NiO (15 nm) [35], which is the case for nearly all the samples, nAFM tends to be uniform in NiO, which is strongly coupled with YIG and can be manipulated by magnetic eld [36]. Thus, nAFM is always perpendicu- lar toHbelow the N eel temperature, because the mFI is parallel to H. In the low temperature limit, e.g. 10 K, the spin current generated in Pt can not penetrate through the NiO, thus the SMR signal is only caused by the NiO layer. The NiO local moments perpendicu- lar toHgives rise to a 90-degree phase shift in the SMR eld angular dependence with respect to the conventional SMR [9]. Such a 90-degree phase shift in a four-fold SMR eld angular dependence is equivalent to a sign reversal in the conventional denition of MR, which explains the negative SMR in Pt/NiO/YIG at low temperatures. For dNiO= 30 nm which is beyond the domain wall width, nAFM at the Pt/NiO interface decouples with mFIand does not respond to H, which explains the vanishing of the negative SMR. At higher temperatures, but below the N eel point, antiferromagnetic order is maintained but the spin cur- rent from Pt has some transmission through NiO, which makes the eect of the YIG more visible as illustrated in Fig. 4(b). The negative SMR contribution from NiO and positive SMR contribution from YIG compete with each other. With increasing temperature, NiO becomes more transparent to the spin current, so the SMR con- tribution from YIG is enhanced. The SMR from NiO may also be suppressed because of the attenuation of the antiferromagnetic order at elevated temperatrues. As a result, the zero point of the SMR occurs at a temperature where the antiferromagnet is still in the ordered phase. Thinner NiO layers have a lower N eel point due to the - nite size eect [37], hence the SMR also changes the sign at lower temperatures in thinner-NiO samples, which is4 (d) ΔSMR/xx T0T( = 0)ΔSMR T 0 H H H(a) ) c ( ) b ( mFInAFM mFInAFM mFIPt YIGNiO FIG. 4. Illustrations for the magnetic structure and spin transport in Pt/NiO/YIG at dierent temperatures. The red and green arrows represent the phenomenologically described spin currents, j1and j2in Eq. (1), respectively. The length of the arrow describes the penetration depth of the spin current. (a),Tclose to the low temperature limit. (b), Tfar above the low temperature limit and lower than the N eel temperature. (c),Thigher than the N eel temperature. (d), Illustration of T-dependent SMR in which the temperatures corresponding to the conditions in Fig. 4(a), (b) and (c) are marked with red circle. in accordance with our observation shown in Fig. 3(a). Around the N eel point as illustrated in Fig. 4(c), the spin transparency of NiO are maximized [24], where the SMR contribution from YIG reaches its peak value and the SMR contribution from NiO vanishes. As ex- plained above, all the main features of the SMR data in Pt/NiO/YIG, such as negative SMR at low tempera- tures,dNiOdependent sign change temperature and peak temperature, can be interpreted by the `spin-
op' con- guration. Figure 4(d) shows an illustration of SMRtemperature dependence, in which the temperature cor- responding to these features are marked. We note that negative SMR has also been reported in bilayers of Pt on gadolinium iron garnet and Ar-sputtered YIG, in which the garnet interface moments can align perpendicularly toH[30, 38]. A simple phenomenological model based on the picture discussed above can also provide a quantitative descrip- tion of the observed SMR temperature dependence. Let us consider a NM/AFM/FI trilayer system. The key as- sumption is that we can describe the spin current through the NM/AFM interface by ejs=GAFnAFM(nAFMs) +t(T)mFI(mFIs) =ej1+ej2; (1) GAFis the real part of the spin mixing conductance at NM/AFM interface. sis the spin accumulation at the same interface. The rst term, which we denote by ej1, is what is expected for NM/AFM bilayer systems as seen in the case studied in Ref. [39]. We have introduced the second term, which is denoted by ej2, to phenomenologi- cally capture the eect of the FI layer. t(T) encapsulates the temperature dependent transparency of the AFM to the spin current. In the case that the AFM is completely transparent the NM/FI bilayer result mFI(mFIs) is recovered. The linear combination of the NM/AFM and NM/FI terms has been chosen in an attempt to em- ulate our SMR data in the NM/AFM/FI system, seen in Fig. 3(a), which seems to indicate a crossover from NM/AFM bilayer like behavior at low temperatures to NM/FI bilayer like behavior for higher temperatures. Once we admit the form of the interfacial spin current in Eq. (1), we can calculate the SMR by employing the diusion equation and the Onsagar principle, according to Refs. [10, 39]. The SMR contribution to the longitu- dinal resistivity then is given by 0=22 SHE2 N dNGAFcos2n+t(T) cos2m+t(T)GAFsin2(m n) 1 +GAF+t(T) +2t(T)GAFsin2(m n)tanh2dN 2N ; (2) where we dened = (2N=) coth(dN=N) withN andSHE being the spin diusion length and the spin Hall angle in NM, respectively, and = 1 0is the con- ductivity of the NM layer. Here, n(m)denotes the angle betweennAFM(mFI) and the applied current jcin NM. Now we set out a hypothesis that the crossover between the negative and positive SMR is of the same origin as the temperature dependence of the spin pumping signal (Fig. 3(b)). In order to support it, the temperature de- pendence of t(T) is obtained by tting to the spin pump-ing data. The resulting function is then used alongside the other parameters in Eq. (1) to t the SMR data to test the validity of our model. Based on the observation that the ISHE signal in Fig. 3(b) is roughly exponential in the intermediate tem- perature regime, we employ VISHE/t(T)/eaT 1 to re- produce the temperature dependence of both spin pump- ing and SMR. The exponential behavior may not apply near the N eel temperature and the data points near and above the N eel temperature have been excluded from the5 tting. Under these assumptions, acan be determined from the spin pumping data (TABLE I). We then t =jm=0 =jm==2based on Eq. (2) to the experimentally obtained SMR ratio SMR=xx in Fig. 3(a) using the tted value of afrom theVISHE data. We x N= 1:5nm,dN= 4:0nm,0= 1= 860 nm, andSHE = 0:05, which are taken to be relevant values to the present experiment, and we further deter- mineGAFandGFfrom the data, where the latter two are dened by t(T) =GF(eaT 1);n m==2, re- spectively. The temperature dependence of 0andSHE is ignored since they scale in some powers of T, which is wiped out by the exponential change in t(T). The t- ting curves can quantitatively reproduce the SMR sign change behavior as shown in Fig. 3(a), and the tting parameters are summarized in TABLE I. dNiOa[K 1]102GAF GF 2:0 1:830:22 3:580:3210128:390:571011 2:2 1:380:19 4:480:1710127:780:261011 2:7 1:420:10 3:670:0910123:010:081011 4:0 1:160:09 2:460:1310122:220:141011 TABLE I. The results of the tting with the data from the SMR and spin pumping signals. The parameters are dened in the main text. The units of the last two columns are both [ 1m 2]. Our result highlights the importance of magnetic struc- ture in AFM for spin transport, which suggests an al- ternative degree of freedom of spin manipulation. The NiO-induced SMR indicates that spin current response of AFM is anisotropic, which opens the possibility to use AFM insulator as a spin current valve or memory. Note added: |Recently, we became aware of similar results for the SMR sign change observed in Pt/NiO/YIG by W. Linetal: [31]. The NiO-thickness dependent SMR at room temperature was also reported by Yu-Ming Hung et al: [40]. qiuzy@imr.tohoku.ac.jp [1] I. A. Campbell, A. Fert, and O. Jaoul, J. Phys. C 3, S95 (1970). [2] T. McGuire and R. Potter, IEEE Trans. Magn. 11, 1018 (1975). [3] G. Binasch, P. Gr unberg, F. Saurenbach, and W. Zinn, Phys. Rev. B 39, 4828 (1989). [4] M. N. Baibich, J. M. Broto, A. Fert, F. N. Van Dau, F. Petro, P. Etienne, G. Creuzet, A. Friederich, and J. Chazelas, Phys. Rev. Lett. 61, 2472 (1988). [5] M. Julliere, Phys. Lett. A 54, 225 (1975). [6] T. Miyazaki and N. Tezuka, Journal of Magnetism and Magnetic Materials 139, L231 (1995). [7] S. Yuasa, T. Nagahama, A. Fukushima, Y. Suzuki, and K. Ando, Nat. Mater. 3, 868 (2004).[8] S. S. P. Parkin, C. Kaiser, A. Panchula, P. M. Rice, B. Hughes, M. Samant, and S.-H. Yang, Nat. Mater. 3, 862 (2004). [9] H. Nakayama, M. Althammer, Y.-T. Chen, K. Uchida, Y. Kajiwara, D. Kikuchi, T. Ohtani, S. Gepr ags, M. Opel, S. Takahashi, R. Gross, G. E. W. Bauer, S. T. B. Goennenwein, and E. Saitoh, Phys. Rev. Lett. 110, 206601 (2013). [10] Y.-T. Chen, S. Takahashi, H. Nakayama, M. Althammer, S. T. B. Goennenwein, E. Saitoh, and G. E. W. Bauer, Phys. Rev. B 87, 144411 (2013). [11] M. Althammer, S. Meyer, H. Nakayama, M. Schreier, S. Altmannshofer, M. Weiler, H. Huebl, S. Gepr ags, M. Opel, R. Gross, D. Meier, C. Klewe, T. Kuschel, J.-M. Schmalhorst, G. Reiss, L. Shen, A. Gupta, Y.-T. Chen, G. E. W. Bauer, E. Saitoh, and S. T. B. Goennenwein, Phys. Rev. B 87, 224401 (2013). [12] M. Weiler, M. Althammer, M. Schreier, J. Lotze, M. Pernpeintner, S. Meyer, H. Huebl, R. Gross, A. Kamra, J. Xiao, Y.-T. Chen, H. Jiao, G. E. W. Bauer, and S. T. B. Goennenwein, Phys. Rev. Lett. 111, 176601 (2013). [13] S. Meyer, M. Althammer, S. Gepr ags, M. Opel, R. Gross, and S. T. B. Goennenwein, Applied Physics Letters 104, 242411 (2014). [14] R. Iguchi, K. Sato, D. Hirobe, S. Daimon, and E. Saitoh, Applied Physics Express 7, 013003 (2014). [15] M. Schreier, T. Chiba, A. Niedermayr, J. Lotze, H. Huebl, S. Gepr ags, S. Takahashi, G. E. W. Bauer, R. Gross, and S. T. B. Goennenwein, Phys. Rev. B 92, 144411 (2015). [16] J. Lotze, H. Huebl, R. Gross, and S. T. B. Goennenwein, Phys. Rev. B 90, 174419 (2014). [17] S. R. Marmion, M. Ali, M. McLaren, D. A. Williams, and B. J. Hickey, Phys. Rev. B 89, 220404 (2014). [18] M. Isasa, A. Bedoya-Pinto, S. V elez, F. Golmar, F. Snchez, L. E. Hueso, J. Fontcuberta, and F. Casanova, Applied Physics Letters 105, 142402 (2014). [19] C. O. Avci, K. Garello, A. Ghosh, M. Gabureac, S. F. Alvarado, and P. Gambardella, Nat Phys 11, 570 (2015). [20] T. Shang, Q. F. Zhan, H. L. Yang, Z. H. Zuo, Y. L. Xie, L. P. Liu, S. L. Zhang, Y. Zhang, H. H. Li, B. M. Wang, Y. H. Wu, S. Zhang, and R.-W. Li, Applied Physics Letters 109, 032410 (2016). [21] C. Hahn, G. de Loubens, V. V. Naletov, J. B. Youssef, O. Klein, and M. Viret, EPL (Europhysics Letters) 108, 57005 (2014). [22] H. Wang, C. Du, P. C. Hammel, and F. Yang, Phys. Rev. Lett. 113, 097202 (2014). [23] T. Moriyama, S. Takei, M. Nagata, Y. Yoshimura, N. Matsuzaki, T. Terashima, Y. Tserkovnyak, and T. Ono, Applied Physics Letters 106, 162406 (2015). [24] Z. Qiu, J. Li, D. Hou, E. Arenholz, A. T. N'Diaye, A. Tan, K.-i. Uchida, K. Sato, S. Okamoto, Y. Tserkovnyak, Z. Q. Qiu, and E. Saitoh, Nature Communications 7, 12670 (2016). [25] L. Frangou, S. Oyarz un, S. Auret, L. Vila, S. Gam- barelli, and V. Baltz, Phys. Rev. Lett. 116, 077203 (2016). [26] W. Lin, K. Chen, S. Zhang, and C. L. Chien, Phys. Rev. Lett. 116, 186601 (2016). [27] N. C. Koon, Phys. Rev. Lett. 78, 4865 (1997). [28] K. Uchida, Z. Qiu, T. Kikkawa, R. Iguchi, and E. Saitoh, Applied Physics Letters 106, 052405 (2015).6 [29] X. Zhou, L. Ma, Z. Shi, W. J. Fan, J.-G. Zheng, R. F. L. Evans, and S. M. Zhou, Phys. Rev. B 92, 060402 (2015). [30] S. V elez, A. Bedoya-Pinto, W. Yan, L. E. Hueso, and F. Casanova, Phys. Rev. B 94, 174405 (2016). [31] W. Lin and C. L. Chien, Phys. Rev. Lett. 118, 067202 (2017). [32] T. Schulthess and W. Butler, Physical Review Letters 81, 4516 (1998). [33] I. P. Krug, F. U. Hillebrecht, M. W. Haverkort, A. Tanaka, L. H. Tjeng, H. Gomonay, A. Fraile- Rodr guez, F. Nolting, S. Cramm, and C. M. Schneider, Phys. Rev. B 78, 064427 (2008). [34] J. Li, A. Tan, S. Ma, R. F. Yang, E. Arenholz, C. Hwang, and Z. Q. Qiu, Phys. Rev. Lett. 113, 147207 (2014). [35] M. Fraune, U. Rdiger, G. Gntherodt, S. Cardoso, and P. Freitas, Applied Physics Letters 77, 3815 (2000). [36] W. Kim, E. Jin, J. Wu, J. Park, E. Arenholz, A. Scholl,C. Hwang, and Z. Q. Qiu, Phys. Rev. B 81, 174416 (2010). [37] D. Alders, L. Tjeng, F. Voogt, T. Hibma, G. Sawatzky, C. Chen, J. Vogel, M. Sacchi, and S. Iacobucci, Physical Review B 57, 11623 (1998). [38] K. Ganzhorn, J. Barker, R. Schlitz, B. A. Piot, K. Ollefs, F. Guillou, F. Wilhelm, A. Rogalev, M. Opel, M. Al- thammer, S. Gepr ags, H. Huebl, R. Gross, G. E. W. Bauer, and S. T. B. Goennenwein, Phys. Rev. B 94, 094401 (2016). [39] S. Takei, B. I. Halperin, A. Yacoby, and Y. Tserkovnyak, Phys. Rev. B 90, 094408 (2014); Y.-T. Chen, S. Taka- hashi, H. Nakayama, M. Althammer, S. T. B. Goennen- wein, E. Saitoh, and G. E. W. Bauer, Journal of Physics: Condensed Matter 28, 103004 (2016). [40] Y.-M. Hung, C. Hahn, H. Chang, M. Wu, H. Ohldag, and A. D. Kent, AIP Advances 7, 055903 (2017). | 2016-10-24 | Spin Hall magnetoresistance (SMR) has been investigated in Pt/NiO/YIG
structures in a wide range of temperature and NiO thickness. The SMR shows a
negative sign below a temperature which increases with the NiO thickness. This
is contrary to a conventional SMR theory picture applied to Pt/YIG bilayer
which always predicts a positive SMR. The negative SMR is found to persist even
when NiO blocks the spin transmission between Pt and YIG, indicating it is
governed by the spin current response of NiO layer. We explain the negative SMR
by the NiO 'spin-flop' coupled with YIG, which can be overridden at higher
temperatures by positive SMR contribution from YIG. This highlights the role of
magnetic structure in antiferromagnets for transport of pure spin current in
multilayers. | Tunable sign change of spin Hall magnetoresistance in Pt/NiO/YIG structures | 1610.07362v2 |
arXiv:2112.01727v1 [quant-ph] 3 Dec 2021Microwave Amplification in a PT-symmetric-like Cavity Magnomechanical System Hua Jin1, Zhi-Bo Yang1, Jing-Wen Jin1, Jian-Yu Liu1, Hong-Yu Liu1,∗and Rong-Can Yang2,3,4† 1Department of Physics, College of Science, Yanbian Univers ity, Yanji, Jilin 133002, China 2College of Physics and Energy, Fujian Normal University, Fujian Provincial Key Laboratory of Quantum Manipulation a nd New Energy Materials, Fuzhou, 350117, China 3Fujian Provincial Engineering Technology Research Center of Solar Energy Conversion and Energy Storage, Fuzhou, 350117, Chin a and 4Fujian Provincial Collaborative Innovation Center for Adv anced High-Field Superconducting Materials and Engineering, Fuzhou, 35011 7, China (Dated: December 6, 2021) We propose a scheme that can generate tunable magnomechanic ally induced amplification in a double-cavityparity-time-( PT-)symmetric-likemagnomechanical systemunderastrongco ntrol and weak probe field. The system consists of a ferromagnetic-mat erial yttrium iron garnet (YIG) sphere placed in a passive microwave cavity which is connected with another active cavity. We reveal that ideally induced amplification of the microwave probe signal may reach the maximum value 106when cavity-cavity, cavity-magnon and magnomechanical coupli ng strengths are nonzero simultaneously. The phenomenon might have potential applications in the fiel d of quantum information processing and quantum optical devices. Besides, we also find the phenom ena of slow-light propagation. In this case, group speed delay of the light can achieve 3 .5×10−5s, which can enhance some nonlinear effect. Moreover, due to the relatively flat dispersion curve , the proposal may be applied to sensitive optical switches, which plays an important role in storing p hotons and quantum optical chips. I. INTRODUCTION The interaction between light and matter is an impor- tant subject in the field of quantum optics. The study of light toward the perspective of quantum leads to some interesting phenomena different from classical ones. One of the most famous phenomenon is induced transparency (such as electromagnetically/optomechanically induced transparency)[ 1–16], aswellasinducedabsorption[ 3,14] and induced amplification [ 15–20] which has been widely studied in current decades. Besides, signal amplification whose aim is to increase signal-to-noise ratio is signifi- cantly crucial in the field of quantum information and quantum optics. It is known that optical amplification usually results from the inversion of particle numbers under the action of a pumping field and stimulated ra- diation. It can directly amplify optical signals without converting them into electrical ones so as to possess a high degree of transparency on the format and rate of signals, making the whole optical fiber communication transmission system more simple and flexible [ 15,16]. It is noted that there are many mechanisms of light ampli- fication, such as adding external drive and changing de- tuning conditions. Through the coupling effect of strong photon tunneling, double-cavityOMS not only showsthe characteristics of photomechanically-induced absorption, photomechanically-inducedamplificationandsimplenor- mal mode splitting (NMS), but also adjusts the photon tunneling intensity. The transformation from photome- chanical induced absorption to photomechanical induced amplification can be further realized. In this article we ∗liuhongyu@ybu.edu.cn †rcyang@fjnu.edu.cnbuild our mechanism by adding active cavities. In ad- dition, the added gain scheme is widely used in quan- tum information and quantum communication due to its excellent characteristics of convenience and easy adjust- ment and may very useful for optical and microwave am- plifiers [21]. Parity-time( PT) symmetry, the non-Hermitian Hami- tonian, which has a real spectra was proposed by Beb- der in 1998 firstly and attracted wide attention [ 22–30]. SincePT-symmetry requires a strict balance between loss and gain. However the balance condition may be too difficult for the realistic implementation, especially when tiny disturbances are inevitable. PT-symmetric- like system not requiring the strict balance can still fol- low the predictions of the PT-symmetry in many cases and thus attract considerable attention [ 31,32]. At the exceptional point, where the system undergoes the transition from the PT-symmetric-like phase and PT- symmetric-like broken, pairs of eigenvalues collide and become complex has manifested in various physical sys- tem, such as photonics, electronics, acoustics, phonon- ics. And the OMIA of the PT-symmetric OMS has been achieved in the whispering-gallery-mode microtoroidal cavities. Common PT symmetric systems have two- cavity systems, but there are also examples of single cavities achieving effective gain by introducing external drives or other means [ 27]. In the past few years, cavity magnonics, a new inter- discipline, attracted much attention. It mainly explores the interaction between confined electromagnetic fields and magnons, especially Yttrium iron garnet (YIG) [ 33– 41]. The reason is that the Kittel mode within YIG has a low damping rate and holds great magnonic nonlinear- ities [39]. In addition, the high spin density of magnons allows strong coupling between magnons and photons,2 Probe Field Control Field YIG may xz b a FIG. 1. Schematic of the setup studied in this paper. A cavity magnomechanical system consists of one ferromagnet ic yttrium iron garnet (YIG) sphere placed inside a passive mi- crowave cavity, which connected with an auxiliary cavity. A bias magnetic field is applied in the zdirection on the sphere toexcites themagnon modes, which are strongly coupled with the cavity field. In the YIG sphere, bias magnetic field ac- tivates the magnetostrictive interaction. The magnetic co u- pling strength of a magnon depends on the diameter of the sphere and the direction of the external bias field [ 45]. We assumed that the YIG’s magnomechanical interactions were directly enhanced by microwave driving (in the ydirection) its magnon mode. Cavity, phonon, and magnon modes are labeledai,b,m(i= 1,2). giving rise to quasiparticles, i.e. the cavity-magnon po- laritons. Then strong coupling between magnons and cavity photons can be observed at both low and room temperature. In this case, a large number of quantum- information-related problems have been studied by this method, including the coupling of magnons with super- conductingqubits, observationofbistability [ 42,43], cav- ity spintronics, energy level attraction of cavity mag- netopolaron, magnon dark modes. Other interesting phenomena including magneton-induced transparency (MIT), magnetically induced transparency (MMIT), and magnetically controlled slow light have also been stud- ied [44]. In this paper, we utilize a cavity-magnomechanical system, which consists of a YIG sphere placed inside a three-dimensional microwave cavity that is connected with anpassivecavityto realizemicrowaveamplification. Through the discussion the properties of absorption and transmission, we obtain the amplification in the context ofPT-symmetric-like cavity magnomechanical system. The remaining parts are organized as follows. In Sec.II,weintroducethemodelofourproposal. InSec. II, we plot the magnomechanically induced transparency window profiles. In Sec. III, we explore magnomechan- ically induced amplification of the PT-symmetric-like cavity magnomechanical system and slow light propega- tion Sec. IV, we present the conclusion of our work.II. MODEL AND HAMILTONIAN We use a hybrid cavity magnomechanical system that consists of one high-quality YIG sphere placed inside a microwave cavity which connects with another empty cavity, as shown in Fig. 1. The YIG sphere has 250 µm in diameter and ferric ions Fe+3of density ρ= 4.22× 1027m−3. This causes a total spin S= 5/2ρVm= 7.07×1014, where Vmis the volume of the YIG and Sis the collective spin operator which satisfies the alge- bra i.e., [ Sα,Sβ] =iεαβγSγ. A uniform bias magnetic field (along z direction) is applied on the sphere, exciting the magnon mode that is then coupled to the first cav- ity field via magnetic-dipole interaction. In addition, the excitation of the magnon mode (i.e. Kittel mode) inside the sphere leads to a variable magnetization that results in the deformation of its lattice structure. The magne- tostrictiveforcecausesvibrationsofthe YIG, resultingin magnon-phonon interaction within YIG spheres [ 45]. It is noted that the single-magnon magnomechanical cou- pling strength depended on sphere diameter and direc- tion of the external bias field is very weak. In this case, magnomechanical interaction of YIG can be enhanced by directly driving its magnon mode via an external mi- crowave field. Furthermore, the first cavity is not only coupled to the second cavity, but also driven by a weak probe field. With consideration of the situation, the Hamiltonian for the whole system reads [ 44,46] H//planckover2pi1=ωmˆm†ˆm+ωa1ˆa† 1ˆa1+ωa2ˆa† 2ˆa2+ωbˆb†ˆb +g1(ˆm†ˆa1+ ˆmˆa† 1)+g2ˆm†ˆm(ˆb+ˆb†) +J(ˆa† 1ˆa2+ˆa† 2ˆa1)+iΩ(ˆm†e−iωput−ˆmeiωput) +iεpr(ˆa† 1e−iωprt−ˆa1eiωprt),(1) where ˆa† j(j= 1,2), ˆm†andˆb†(ˆaj, ˆmandˆb) are the cre- ation (annihilation) operators of the jth cavity, magnon and phonon, respectively. They all satisfy the stan- dard commutation relations for bosons. ωaj,ωm,ωb represent the resonance frequencies for the jth cavity, magnonandphonon, respectively. g1(J)denotesthecou- pling strength between the first cavity mode and magnon (the second cavity), and g2is the coupling constant be- tween magnon and phonon. It is noted that the fre- quencyωmis determined by the gyromagnetic ratio γ and external bias magnetic field Hi.e.,ωm=γHwith γ/2π= 28GHz. In addition, Ω =√ 5/4γ√ NB0is the Rabi frequency, which is dependent of the coupling strength of the driving field with amplitude B0and fre- quencyωpu. Andωpris the probe field frequency having amplitude εpr=/radicalbig 2Ppκ1//planckover2pi1ωpr. It should be noted that wehaveignoredthenonlinearterm Kˆm†ˆm†ˆmˆminEq.(1) that may arise due to strongly driven magnon mode [ 43] so as to K|/angbracketleftm/angbracketright|3≪Ω. With the rotating wave approxi-3 mation, we can rewrite the whole Hamiltonian as H//planckover2pi1= ∆mˆm†ˆm+∆a1ˆa† 1ˆa1+∆a2ˆa† 2ˆa2+ωbˆb†ˆb +g1(ˆm†ˆa1+ ˆmˆa† 1)+g2ˆm†ˆm(ˆb+ˆb†) +J(ˆa† 1ˆa2+ˆa† 2ˆa1)+iΩ(ˆm†−ˆm) +iεpr(ˆa† 1e−iδt−ˆa1eiδt),(2) with ∆ aj=ωaj−ωpu(j= 1,2), ∆m=ωm−ωpu, and δ=ωpr−ωpu. In order to obtain the evolution of aj(t),m(t) andb(t), we use quantum Heisenberg-Langevin equations, which can be expressed by ˙ˆa1=−i∆a1ˆa1−ig1ˆm−κ1ˆa1+εpre−iδt +√2κ1ˆain 1(t)−iJˆa2, ˙ˆa2=−i∆a2ˆa2−κ2ˆa2+√ 2κ2ˆain 2(t)−iJˆa1, ˙ˆm=−i∆mˆm−ig1ˆa1−κmˆm−ig2ˆm(ˆb+ˆb†) +√2κmˆmin(t)+Ω, ˙ˆb=−iωbˆb−ig2ˆm†ˆm−κbˆb+√2κbˆbin(t)(3) whereκ1(κ2),κbandκmare the decay rates of the cav- ities, phonon and magnon modes, respectively. ˆ ain 1(t), ˆain 2(t),ˆbin(t) and ˆmin(t) are the vacuum input noise operators which have zero mean values and satisfies/angbracketleftbig ˆqin/angbracketrightbig = 0(q=a1,a2,m,b). The magnon mode m is strongly driven by a microwave field that causes a large steady-state amplitude corresponds to |/angbracketleftms/angbracketright| ≫1. Moreover, owing to the magnon coupled to the cavity modethroughthebeam-splitter-typeinteraction, thetwo cavity fields also exhibit large amplitudes |/angbracketleftajs/angbracketright| ≫1. Then we can linearize the quantum Langevin equations around the steady-state values and take only the first- order terms in the fluctuating operator:/angbracketleftBig ˆO/angbracketrightBig =Os+ ˆO+e−iδt+ˆO−eiδt[43], where ˆO=a1,a2,b,m.the steady- state solutions are given by a1s=−(ig1ms+iJa2s) i∆a1+κ1,a2s=−iJa1s i∆a2+κ2, bs=−ig2|ms|2 iωb+κb, ms=−ig1a1s+Ω i/tildewide∆m+κm, /tildewide∆m= ∆m+g2(bs+bs∗)(4) In ordertoachieveourmotivationofsignalamplification, we neglect off resonance terms to let ˆO−= 0, but ˆO+ safisfying the relations (iλ−κ1)ˆa1+−ig1ˆm+−iJˆa2++εpr= 0, (iλ−κ2)ˆa2+−iJˆa1+= 0, (iλ−κm)ˆm+−ig1ˆa1+−iGˆb+= 0, (iλ−κb)ˆb+−iG∗ˆm+= 0,(5)0 0.5 1 1.5 2 δ/ωb0100200300400|tp|2 (a) 0 0.5 1 1.5 2 δ/ωb05101520|tp|2 (b) 0 0.5 1 1.5 2 δ/ωb11.21.41.61.8|tp|2 (c) 0 0.5 1 1.5 2 δ/ωb11.21.41.6|tp|2 (d) FIG. 2. The transmission |tp|2spectrum of probe field as function of δ/ωbwhen only interaction between two cavities is nonzero, (a) J/2π= 0.6MHz, (b) J/2π= 0.8MHz, (c) J/2π= 2.0MHz and (d) J/2π= 6MHz. where we have set G=g2ms,λ=δ−ωb,ωai≫κi (i= 1,2), and ∆ a1= ∆a2=/tildewide∆m=ωb. In this case, we can easily obtain ˆa1+=εpr κ1−iλ+J2 κ2−iλ+g12 κm−iλ+|G|2 κb−iλ.(6) By use of the input-output relation for the cavity field εout=εin−2κ1/angbracketlefta1+/angbracketrightand setting εin= 0, the amplitude of the output field can be written as ε′ out=εout εpr=2κ1ˆa1+ εpr. (7) The real and imaginary parts of the output field are Re [ε′ out] =κ1(ˆa1++ ˆa∗ 1+)/εprand Im [ ε′ out] =κ1(ˆa1+− ˆa∗ 1+)/εpr. These factors describe the absorption and dis- persion of the systems, respectively. III. INDUCED AMPLIFICATION AND SLOW LIGHT PROPEGATION IN PT-SYMMETRIC-LIKE MAGNOMECHANICALLY SYSTEMS For the numerical calculation, we use parameters cho- sen from a recent experiment on a hybrid magnome- chanical system, where ωa1/2π=ωa2/2π= 10GHz, ωb/2π= 10MHz, κb/2π= 100Hz, ωm/2π= 10GHz, κ1/2π= 2.0MHz,κm/2π= 0.1MHz,g1/2π= 1.0MHz, G/2π= 3.5MHz,∆ a1= ∆a2=/tildewide∆m=ωb,ωd/2π= 10GHz are set [ 26,33,34].4 0 0.5 1 1.5 2 δ/ωb11.11.21.31.41.5|tp|2 (a) 0 0.5 1 1.5 2 δ/ωb11.11.21.31.41.5|tp|2 (b) 0 0.5 1 1.5 2 δ/ωb11.11.21.31.41.5|tp|2 (c) 0 0.5 1 1.5 2 δ/ωb123456|tp|2 (d) FIG. 3. The transmission |tp|2spectrum of probe field as function of δ/ωbwhen only coupling between magnon and phonon is absent means G= 0,J/2π= 3.0MHz (a) g1/2π= 1.0MHz, (b) g1/2π= 1.2MHz, (c) g1/2π= 1.5MHz and (d) g1/2π= 2.0MHz. At first, we consider the transmission rate |tp|2as a function of the probe detuning δ/ωbin the context of parity-time-( PT-) symmetric-like magnomechanical sys- tem. FromEq.( 7), the rescaledtransmissioncorrespond- ing to the probe field can be expressed as tp= 1−2κ1ˆa1+ εpr. (8) We first depict the transmission spectrum of the probe field against the scaled detuning δ/ωb, for different val- ues ofJin Fig.2, where the phonon-magnon coupling rate and photon-magnon interaction parameters are set to zero, i.e. G=g1= 0. From Fig. 2(a), we can observe that the transmission peak near δ=ωbwhich is asso- ciated with the coupling rate of two cavities can much be larger than 1. The reason is that the gain cavity can scatter photons into the dissipative cavity. From Fig2(a)-(d), transmission coefficient decreases with the increase of coupling strength between two cavities. And we got a downward dip with two peaks From Fig 2(c)- (d), amplification area becomes wider when Jgetting lager simultaneously. This means that we can adjust the transmission coefficient and the size of the amplification region by changing the coupling between the two cavities when the system is double-cavity PT-symmetric-likeand the cavity contains no magnon. Next, we introduced one more coupling constant only set the coupling between magnon-phonon G= 0. We got another peak near δ=ωbcompared with Fig 2(c)-(d), which was caused by coupling between magnon-photon in Fig3. This is because the magnon can scatter the0 0.5 1 1.5 2 δ/ωb012345|tp|2×106 (a) 0 0.5 1 1.5 2 δ/ωb01000300050007000|tp|2 (b) 0 0.5 1 1.5 2 δ/ωb040010001200|tp|2 (c) 0 0.5 1 1.5 2 δ/ωb0100200300400|tp|2 (d) FIG. 4. The transmission |tp|2spectrum of probe field as function of δ/ωbwhenG/2π= 2.0MHz,g1/2π= 6.0MHz (a) J/2π= 0.64MHz, (b) J/2π= 0.8MHz, (c) J/2π= 2MHz and (d)J/2π= 4MHz. photons of the active field into the probe field via indi- rect interaction. However, from Fig 3(b)-(d), the middle peak became taller when g1increases. Hence the effect of light amplification caused by the interaction of magnon- photon get better as g1increase. And with the increasing ofmiddlepeak, theheightoftwopeaksonbothsidesstay the same, that is, the light amplification caused by the coupling between two cavities not affected by g1. How- ever, the amplification effect is not ideal. We show the transmission spectrum when three cou- pling constants are nonzero simultaneously and coupling betweenmagnon-photonlagerthanmagnon-phonon g1> Gin Fig4(a)-(d). We got only one amplification peak when the coupling between two cavities J/2π= 0.64MHz, another upward peak appeared with the in- creasing of J, and the height of two peaks is the same and the amplification effect induced by the interaction of magnon-phonon and magnon-photon were superior at this time, this is because magnon and phonon can also scatter the photons of active cavity field into the probe field. And since the excited states of the cavity field are pumped into higher energy levels, they stay long enough can also be amplified by stimulated radiation. Amplifi- cation area becomes wider when Jgetting lager simul- taneously. These results provide an effective way to re- alize continuous optical amplification and have practical significance for the construction of quantum information processing enhancement signal based on cavity magnetic system. Finally, we plotted the transmission spectrum of the probe field against the scaled detuning δ/ωb, for different values of /tildewide∆m. From Fig 5(a)-(d), The obvious displace-5 0 0.5 1 1.5 2 δ/ωb020406080|tp|2 (a) 0 0.5 1 1.5 2 δ/ωb020406080|tp|2 (b) FIG. 5. The transmission |tp|2spectrum of probe field as function of δ/ωbwhen three coupling constants are nonzero, (a)/tildewide∆m= 0.5ωb, (b)/tildewide∆m= 1.5ωb. 0 0.5 1 1.5 2 δ/ωb-0.501233.5τg(s)×10-5 FIG. 6. The group delay τgas functions of δ/ωbwhenG= 2MHz,J/2π= 6.3MHz,g1/2π= 6.1MHz. ment ofthe twopeaksmeansthat wecannotonly change the value of amplification and the size of the amplifica- tion region by adjusting the coupling strength, but also flexibly change the location of the amplification region. Moreover, the phase φtof the output field can be givenas φt= arg[εout] (9) And the rapid phase dispersion of output field can cause the group delay, which can expressed as τg=∂φt ∂ωpr(10) From Fig. 6shows that the group delay τgas a func- tion of the detuning δ/ωbwhen three coupling constants are present. We can observe double peaks and double dips, peaks corresponding positive group delay i.e., slow light propagation, dips corresponding negative group de- lay means fast light propagation. And we can realize group speed delay of 3 .5×10−5s, a tunable switch from slow to fast can be achieved by adjusting the gain of ac- tive cavity or coupling constants. IV. CONCLUSION In conclusion, we study the transmission of probe field inthesituationof PT-symmetric-likeunderastrongcon- trol field in a hybrid magnomechanical system in the microwave regime and realized ideal induced amplifica- tion when three coupling constants are nonzero simulta- neously, which due to gain cavity, magnon and phonon can also scatter the photons into the dissipative cavity. Therefore, our results are not only providing rich scien- tific insight in terms of new physics but also potentially have important long-term technological implications, in- cluding the development of on-chip optical systems that support states of light that are immune to back scat- ter, are robust against perturbation and feature guar- anteed unidirectional transmission. Then we achieved a group delay of 3 .5×10−5seconds. Slowing down the en- ergyspeed oflight allowsphotonsto interact with matter enough to enhance some nonlinear effects. And because the dispersion curve is relatively flat, a small change in frequency will also cause a large change in photon mo- mentum, so it can be made into a more sensitive opti- cal switch. Finally, the slow light effect slows down the energy speed of light, which can play a role of storing photons and quantum optical chips. ACKNOWLEDGMENTS This work is supported by the National Natural Sci- ence Foundation of China (Grant No. 62165014) and the Fujian Natural Science Foundation (Grant No. 2021J01185). [1] S.E. Harris, J.E. Field, Phys.Rev.Lett. 64, 1107 (2019). [2] R. Thomas, C. Kupchak, G. S. Agarwal and A. I.Lvovsky, Opt. Express. 21, 6880 (2013). [3] H. Fredrik, S. Albert, J. K. Tobias and H. Hans, New. J.6 Phys.14, 123037 (2012). [4] H. Y. Ma and L. Zhou, Phys. Rev. B. 22, 024204 (2013). [5] K. J. Boller, A. Imamoglu and S. E. Harris, Phys. Rev. Lett.66, 2593 (1991). [6] M. Fleishhauer, A. Imamoglu and J. P. Marangos, Rev. Mod. Phys. 77, 633 (2005). [7] S. M. Huang and G. S. Agarwal, Phys. Rev. A. 83, 043826 (2011). [8] H. Jing, Phys. Rev. A, 90, 043839 (2014). [9] M. Karuza, Phys. Rev. A, 88, 013804 (2013). [10] Y. He, Phys. Rev. A, 91, 013827 (2015). [11] W. J. Nie, A. X. Chen and Y. H. Lan, Phys. Rev. A, 93, 023841 (2016). [12] H. Xiong and Y. Wu, Appl. Phys. Rev. 5, 031305 (2018). [13] L. W. Liu, D. J. Gengzang and X. J. An, Chin. Phys. B 27, 034205 (2018). [14] M. Karuza, C. Biancofiore, M. Bawaj, C. Molinelli, M. Galassi, R. Natali, P. Tombesi, G. Di Giuseppe, and D. Vitali, Phys. Rev. A. 88, 013804 (2013). [15] L. W. Liu, D. J. Gengzang and X. J. An, Chin. Phys. B 27, 034205 (2018). [16] X. B. Yan, W. Z. Jia and Y. Li, Frontiers of Physics. 10, 354-357(2015). [17] X. B. Yan, W. Z. Jia, Y. Li, J. H. Wu, X. L. Li and H. W. Mu, Front. Phys. 10, 104202 (2015). [18] T.Wang, M.H.Zheng, C.H.Bai, D.Y.Wang, A.D.Zhu, H. F. Wang and S.Zhang, Ann. Phys, Berlin1800228 (2018). [19] B. Wang, Z. X. Liu, C. Kong, H. Xiong, and Y. Wu, Opt. Express 26, 20248 (2018). [20] Q. Yang, B. P. Hou and D. G. Lai, Opt. Express. 25, (2017) [21] C. Jiang, L. N. Song and Y. Li, Phys. Rev. A. 99, 023823 (2019). [22] Y. L. Liu, R. B. Wu and J. Zhang, Phys. Rev. A. 95, 013843 (2017). [23] J. Zhang, B. Peng, ¸ S. K. ¨Ozdemir, Y.-x. Liu, H. Jing, X.-y. L¨ u, Y.-l. Liu, L. Yang, and F. Nori, Phys. Rev. B. 92, 115407 (2015). [24] X. Y. Lv, H Jing and J. Y. Ma, Phys. Rev. Lett. 114, 253601 (2015). [25] C. M. Bender and S. Boettcher, Phys. Rev. Lett. 70, 947(2007). [26] K. G. Makris, R. El-Ganainy, D. N. Christodoulides, andZ. H. Musslimani, Phys. Rev. Lett. 100, 103904(2008). [27] Z. Lin, H. Ramezani, T. Eichelkraut, T. Kottos, H. Cao, and D. N. Christodoulides, Phys. Rev. Lett. 106, 213901 (2011). [28] H. Jing, S. K. ¨Ozdemir, X.-Y. L¨ u, J. Zhang, L. Yang, and F. Nori, Phys. Rev. Lett. 113, 053604(2014). [29] X. Li and X.-T. Xie, Phys. Rev. A, 90, 033804 (2014). [30] S. N. Huai, Y. L. Liu, J. Zhang, L. Yang, and Y. X. Liu, Phys. Rev. A. 99, 043803 (2019). [31] S. Chakraborty and A. K. Sarma, Phys. Rev. A. 100, 063846 (2019). [32] Y. T. Chen, L. Du, Y. Zhang and J. H. Wu, Phys. Rev. A.103, 053712 (2020). [33] Y. Tabuchi, S. Ishino, T. Ishikawa, R. Yamazaki, K. Us- ami, and Y. Nakamura, Phys. Rev. Lett. 133, 083603 (2014). [34] M. Goryachev, W. G. Farr, D. L. Creedon, Y. Fan, M. Kostylev, and M. E. Tobar, Phys. Rev. Applied, 2, 054002 (2014). [35] H.Huebl,C.W.Zollitsch, J.Lotze, F.Hoche, M.Greifen - stein, A. Marx, R. Gross, and S. T. B. Goennenwein, Phys. Rev. Lett. 111, 127003 (2013). [36] X. Zhang, C. L. Zou, L. Jiang, and H. X. Tang, Phys. Rev. Lett. 113, 156401 (2014). [37] Y. Tabuchi, S. Ishino, T. Ishikawa, R. Yamazaki, K. Us- ami, and Y. Nakamura, Phys. Rev. Lett. 113, 083603 (2014). [38] D. Zhang, X. M. Wang, T. F. Li, X. Q. Luo, W. Wu, F. Nori, and J. Q. You, npj Quantum Inf. 1, 15014 (2015). [39] C. Kittel, Phys. Rev. 73, 155 (1948). [40] X. Y. Zhang, C. L. Zou, L. Jiang, and H. X. Tang, Sci. Adv.2, e1501286(2016). [41] S. N. Huai, Y. L. Liu, J. Zhang, L. Yang, and Y. X. Liu, Phys. Rev. A. 99, 043803 (2019). [42] Z. B. Yang, H. Jin, J. W. Jin, J. Y. Liu, H. Y. Liu, and R. C. Yng, Phys. Rev. Research 3, 023126 (2021). [43] Y. P. Wang, G. Q. Zhang, D. K. Zhang, T. F. Li, C. M. Hu, and J. Q. You, Phys. Rev. Lett. 120, 057202 (2018). [44] U. Kamran and M. Tahir Naseem, Phys. Rev. A 102, 033721 (2020). [45] X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, Sci. Adv. 2,e1501286 (2016) [46] J. Li and S.-Y. Zhu, New J. Phys. 21, 085001 (2019). | 2021-12-03 | We propose a scheme that can generate tunable magnomechanically induced
amplification in a double-cavity parity-time-(PT -) symmetric-like
magnomechanical system under a strong control and weak probe field. The system
consists of a ferromagnetic-material yttrium iron garnet (YIG) sphere placed in
a passive microwave cavity which is connected with another active cavity. We
reveal that ideally induced amplification of the microwave probe signal may
reach the maximum value 1000000 when cavity-cavity, cavity-magnon and
magnomechanical coupling strengths are nonzero simultaneously. The phenomenon
might have potential applications in the field of quantum information
processing and quantum optical devices. Besides, we also find the phenomena of
slow-light propagation. In this case, group speed delay of the light can
achieve 0.000035s, which can enhance some nonlinear effect. Moreover, due to
the relatively flat dispersion curve, the proposal may be applied to sensitive
optical switches, which plays an important role in storing photons and quantum
optical chips. | Microwave Amplification in a PT -symmetric-like Cavity Magnomechanical System | 2112.01727v1 |
Determination of the origin of the spin Seebeck eect - bulk vs. interface eects Andreas Kehlberger,Ren e R oser, Gerhard Jakob, and Mathias Kl aui Institute of Physics, University of Mainz, 55099 Mainz, Germany Ulrike Ritzmann, Denise Hinzke, and Ulrich Nowak Department of Physics, University of Konstanz, D-78457 Konstanz, Germany Mehmet C. Onbasli, Dong Hun Kim, and Caroline A. Ross Department of Materials Science and Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA Matthias B. Jung
eisch and Burkard Hillebrands Fachbereich Physik and Landesforschungszentrum OPTIMAS, Technische Universit at Kaiserslautern, Kaiserslautern 67663, Germany (Dated: November 2, 2021) The observation of the spin Seebeck eect in insulators has meant a breakthrough for spin caloritronics due to the unique ability to generate pure spin currents by thermal excitations in insulating systems without moving charge carriers. Since the recent rst observation, the under- lying mechanism and the origin of the observed signals have been discussed highly controversially. Here we present a characteristic dependence of the longitudinal spin Seebeck eect amplitude on the thickness of the insulating ferromagnet (YIG). Our measurements show that the observed behavior cannot be explained by any eects originating from the interface, such as magnetic proximity eects in the spin detector (Pt). Comparison to theoretical calculations of thermal magnonic spin currents yields qualitative agreement for the thickness dependence resulting from the nite eective magnon propagation length so that the origin of the eect can be traced to genuine bulk magnonic spin currents ruling out parasitic interface eects. PACS numbers: 72.20.Pa, 72.25.Mk, 75.30.Ds, 85.80.-b Graduate School Materials Science in Mainz, Staudinger Weg 9, 55128, GermanyarXiv:1306.0784v1 [cond-mat.mtrl-sci] 4 Jun 20132 INTRODUCTION In the fast evolving eld of spin caloritronics[1] many interesting discoveries have been made, such as the magneto Seebeck eect[2], and the spin Seebeck eect (SSE) in metals[3], and semiconductors[4]. One of the most interesting eects is the SSE in ferromagnetic insulators (FMI)[5], such as yttrium iron garnet (YIG). Even in insulators this eect oers the possibility to generate a pure spin current by just thermal excitation. Hence this spin current excited in an insulating system is not carried by moving charge carriers but by excitations of the magnetization, known as magnons. Common theories explain this magnonic SSE, being due to a dierence between the phonon- and magnon temperatures TNandTm[6, 7], while other theories rely on a strong local magnon-phonon[8] coupling. Up to now no experimental method has been capable of directly observing this temperature dierence of magnons and phonons[9] so that the origin of the genuine SSE in the transverse conguration is still unclear. Furthermore, in the transverse conguration a thermal gradient is generated in the lm plane, while the detection layer is on top of the ferromagnetic lm. Due to dierences in the thermal conductivity of the substrate and ferromagnetic lm and the thickness dierence between lm and substrate as well as the temperature dierences between sample and environment, it is challenging to generate only a pure in-plane thermal gradient without an out-of-plane component[10]. In conductors this out-of-plane component of the thermal gradient will unavoidably lead to parasitic eects, such as the anomalous Nernst eect, that superimpose with any genuine SSE signals in the transverse geometry. For insulators an alternative geometry provides a better controlled conguration. The so called longitudinal conguration[11] establishes the thermal gradient in the out-of-plane direction across the substrate, the ferromagnetic thin lm and the detection layer on top. This opens the possibility to study the SSE independently of the thermal conductivity of the substrate, which does not aect the direction of the thermal gradient. Another key point, which complicates the interpretation of the SSE experiments, is the detection method for the thermally excited spin currents. Most spin caloric experiments rely on the indirect measurement by the inverse spin Hall eect (ISHE)[12] to detect the spin current pumped by the SSE. The measured inverse spin Hall voltage[6, 13] is predicted to be: VISHE = SHElNN Ipump s SHElNN
~g"#kB MsVaA Tm TN : (1) Here, SHEis the spin Hall angle of the spin detector material, lNthe length between the voltage contacts, and N the resistance of detection layer. The underlying spin current Ipump s itself depends on the gyromagnetic ratio
, the saturation magnetization MS, the spin mixing conductance g"#, the Boltzmann constant kB, the coherence volume of the magnetic system Va, the contact area A, and the temperature dierence of the magnons of the FMI and the phonons of the normal metal (NM) at the interface Tm TN . Unfortunately the spin mixing conductance g"#and with that the pumped spin currents Ipump s are very sensitive to the interface quality[14]. This interplay of the interface and the ISHE makes it necessary to carefully maintain the properties of the interface if one wants to compare dierent samples. Furthermore parasitic eect may be caused by the detection layer itself: Most experiments use platinum3 Table I. Thickness of YIG and Pt, crystalline orientation and number of samples for each series. Series Number YIG (nm) Pt (nm) Orientation in-situ etching 1 70,130,200 8.5 100 yes 2 20,70,130,200,300 8.5 100 yes 3 40,80,100,130,150 10.3 100 no (Pt) for this layer, due to the high spin Hall angle, making it an ecient spin current detector[15]. Many experiments have shown that the paramagnetic Pt shows a measureable magnetoresistance eect in contact with YIG[16, 17]. The YIG/Pt interface has been investigated more closely by X-ray magnetic circular dichroism measurements that reveal possibly a small induced magnetic moment of the Pt[18]. In combination with a thermal gradient, this proximity eect can cause an additional parasitic thermoelectric eect, the anomalous Nernst eect. For this reason one needs even in insulating ferromagnets to clearly distinguish between such parasitic interface eects and the genuine spin Seebeck eects due to spin currents[19]. Other recent measurements[20, 21] attribute such magnetoresistance eects to a spin Hall magnetoresistance and observe no proximity eect[22]. So given these dierent contradicting claims there is a clear need to distinguish whether the observed signals originate from a parasitic interface eect or a real bulk spin Seebeck eect. Here we present a detailed study of the relevant length scales of the longitudinal SSE in in YIG/Pt by varying the thickness of the ferromagnetic insulator. The obtained results show an increasing and saturating SSE signal with increasing YIG lm thickness. By determining also the dependence of the magnetoresistive eect and the saturation magnetization on the thickness, we can exclude an interface eect as the source of the measured signal. By atomistic spin simulation of the propagation of exchange magnons in temperature gradients, we are able to explain this behavior as being due to a nite eective propagation length of the thermally excited magnons. RESULTS All Y 3Fe5O12samples presented in this paper were grown by pulsed laser deposition with lm thicknesses ranging from 20 nm to 300 nm as shown in table I. The samples have been sorted into three series, where the interface conditions are identical for samples within one series. Details are given in the Methods section. We rst determined the intrinsic magnetic properties of every sample by SQUID magnetometry. Fig. 1a shows the saturation magnetization ( MS) as a function of lm thickness. Apart from very thin lms (20 nm), we nd values of approximately 120 kA/m 25 kA/m, which is close to the literature value of 140 kA/m for YIG thin lm samples[23]. For the very thin lms, a decrease of the moment has been previously observed for other thin YIG lms produced by PLD[24]. To estimate the in
uence of the YIG/Pt interface coupling, which was previously claimed to be the origin of the measured SSE signals[16], we checked the magnetic eld dependence of the Pt resistivity. The magnetoresistive eect,4 05 0100150200250300040801201600 5 01001502002503000123 Series 1 Series 2 Series 3Ms (kA/m)Y IG thickness (nm) Series 2 Series 3Δρ/ρY IG thickness (nm)ba Figure 1. Thickness dependence of saturation magnetization and magnetoresistance eect. (a) Saturation magnetization ( MS) as a function of YIG lm thickness. Each series is marked in dierent colors. The literature value of 140 kA/m is indicated by a grey dashed line. The error of 25 kA/m takes into account that the active magnetic volume of the lm had to be estimated and the subtraction of the paramagnetic substrate signal. The uncertainty in the thickness determination translates into an error of the active magnetic volume and therefore an error of MS.(b)=as a function of YIG-layer thickness measured for series 2 and 3. The y-axis error represents one standard deviation combined with a systematic error considering the temperature variability of the measurement. which was observable in every sample, showed a dependence on the magnetization direction as well as an in-plane angular dependence that can be explained by the novel spin Hall magnetoresistance eect[20, 21]. To determine the correlation with YIG lm thickness, we measured the in-plane resistivity for = 90and= 0in a four-point contact conguration. From this data we calculated == 2 (0 90)=(0+90), as shown in Fig. 1b. For each series =remained constant, and largely independent of the YIG lm thickness. Due to the identical interface conditions for samples of one series, we can assume that the magnetoresistive eect amplitude exhibits no signicant dependence on the YIG-lm thickness as expected for an interface eect. The changes of the absolute magnetoresistance signal between the dierent series can be explained by the change of the Pt-thickness and a residual variation of the interface quality. With the knowledge of the thickness dependence of these material and interface-related parameters, we can now ascertain whether the spin Seebeck eect is correlated to one of those parameters. Three series of YIG lms have been investigated in terms of the spin-Seebeck coecient (SSC), covering a thickness range from 20 nm to 300 nm. A more detailed explanation of the SSC measurements is given in the Methods section. Fig. 2 shows the measured YIG-layer thickness dependence of the SSC for each series. Below 100 nm, lms of each series showed an increase of the signal amplitude with increasing thickness. For larger thicknesses the signal starts to saturate. This saturation behavior could be observed in all our series that consist of epitaxial single crystalline lms. The samples of series 3 generated signals a factor of two lower than the signals of the other series, due to no in-situ interface etching prior to the Pt deposition, which leads to a less transparent interface5 0501001502002503000,00,20,40,60,8 SSC (µV/K)Y IG thickness (nm) Series 1 Series 2 Series 3 Figure 2. Spin Seebeck coecient as a function of YIG layer thickness . SSC as a function of YIG-layer thickness. The samples are sorted into dierent series. Samples of one series have been processed under identical conditions. Data points of each series are connected for clarity. The error in y-axis corresponds to one standard deviation of the measurement data combined with a systematic error taking into account the uncertainty of the mechanical mounting. for the magnons and therefore a smaller spin mixing conductance[14]. This observation underlines the importance of the interface conditions for the comparison of dierent samples, but the absolute trend of the thickness dependence was not aected by this. DISCUSSION In order to understand the origin of the signal, we compare the thickness dependence of the SSC with the thickness dependence of possible underlying mechanisms: When comparing this thickness scaling with that of the saturation magnetization MS, shown in Fig. 1a, we can exclude a direct correlation. We would expect a constant SSC for lms thicker than 40 nm, since only lms below 40 nm showed a MSdependence on the YIG thickness. Secondly we compare the thickness dependence of the magnetoresistive eect, shown in Fig. 1b, with the one of the SSC. Again one would expect a constant contribution to the measured signal independent of the YIG lm thickness when comparing with the thickness dependence of the magnetoresistive eect. For this reason we can exclude that any interface coupling eect leads to the observed thickness dependence of the SSC. Even if the magnetoresistive eect in combination with a thermal gradient leads to a Nernst eect, the signals produced by it deliver a constant oset for each series, which cannot be the source of the signal with the thickness dependence that we observe. This is of major importance as it allows us to conclude that the source of the observed signals is not the currently discussed proximity eects at the6 interface[16]. The clear thickness dependence points to an origin in the bulk of the YIG. In the following analysis we assume that the role of the YIG thickness for the SSE might be due to a nite length scale for magnon propagation in the YIG material. In order to investigate this we simulate the propagation of thermally excited magnons in a temperature gradient using an atomistic spin model. The model is generic and not intended to describe YIG quantitatively. It contains a ferromagnetic nearest-neighbor exchange interaction Jand an uniaxial anisotropy with easy-axis along x-direction and anisotropy constant dx= 0:1J. We investigate a cubic system with 51288 spins, which are initialized parallel to the x-axis. The dynamics of the spin system is calculated by solving the stochastic Landau-Lifshitz-Gilbert equation numerically with the Heun-Method[25]. The phonons provide a heat-bath for the spin system where we assume a linear temperature gradient over the length Linx-direction as shown in Fig. 3. This temperature prole remains constant during the simulation. After an initial relaxation, the local, reduced magnetization m(x) depends on the space coordinate xand its prole is determined as an average over all spins Siin the corresponding y-z-plane and additionally as an average over time. Due to the temperature gradient, magnons propagate from the hotter towards the colder region of the system and this magnonic spin current leads to deviations of the local magnetization m(x) from its local equilibrium value m0(x) which would follow from the local temperature Tp(x) of the phonon system. A temperature dependent calculation of the equilibrium magnetization m0(T) for a system with constant temperature allows us to describe this deviation, which we dene as magnon accumulation m(x)[26] via m(x) =m(x) m0(x;Tp(x)) . (2) Fig. 3 shows this magnon accumulation mxas a function of space coordinate xin a system with a damping constant of = 0:01 and a temperature prole with a linear temperature gradient of T= 10 5J=(kBa), whereais the lattice constant of the cubic system, for two dierent lengths Lof the temperature gradient. At the hotter end of the gradient magnons propagate towards the cooler region of the system and this reduces the number of the local magnons and increases the local magnetization. On the other side at the cold end of the gradient magnons arriving from hotter parts of the system decrease the local magnetization. The resulting magnon accumulation is symmetric in space and changes its sign in the center of the temperature gradient. The spatial dependence of the magnon accumulation as well as the height of the peaks at the hot and cold end are aected by the mean eective propagation length of the magnons[26] in the system. If the length Lof the gradient is smaller than the mean propagation length of the magnons, the magnon accumulation depends linearly on the space coordinate x. For higher values of the length Lthe accumulation at the center of the gradient vanishes and appears only at the edges of the temperature gradient. This is in agreement with simulation by Ohe et al. of the transverse spin Seebeck eect[7]. In their simulation they modify the mean propagation length of the magnons by changing the damping constant and obtain comparable results. The eective mean propagation length of the magnons can be estimated by tting it to the function.7 -100- 500 5 01 00-3-2-10123T p (J/kB) Δm • 105x [# spins] 20a 100a0,000,01 Figure 3. Magnon accumulation in a spin system with a temperature gradient. Magnon accumulation mas a function of the space coordinate xfor a given phonon temperature Tpincluding a temperature gradient of two dierent lengths L= 20a;100a The magnitude of the magnon accumulation at the cold end of the gradient increases with increasing length Lup to a saturation value depending on the mean propagation length of the magnons. The magnon accumulation can be understood as the averaged sum of the magnons, which can reach the end of the gradient. As illustrated in Fig. 4 only those magnons from distances smaller than their propagation length contribute to the resulting magnon accumulation at the cold end of the temperature gradient. Xiao et al. showed that the resulting spin current from the ferromagnet into the non-magnetic material is proportional to the temperature dierence between the magnon temperature Tm in the ferromagnet and the phonon temperature of the non-magnetic material TN[6]. Here, for simplicity, we assume that the temperature of the non-magnetic material is TN= 0 K and no back
ow from the non-magnetic material exists. The magnon temperature Tmat the cold end of the gradient can be calculated from the local magnetization m(x). The resulting magnon temperature dependence on the length Lof the temperature gradient saturates due to the mean propagation length of the magnons as shown in Fig. 5 for two dierent damping constants . The variation of the damping constant leads to variation of the mean magnon propagation length and, consequently, dierent length scales where saturation for the magnon temperature sets in. Tcold m/ 1 exp L . (3)8 ferromagnetic insulator normal metal Figure 4. Origin of SSC thickness dependence. Illustration of the saturation eect of the measured voltage due to a nite propagation length of the excited magnonic spin currents. 05 01 001 502 000,00,20,40,60,81,00 1 002 003 004 000,00,20,40,60,81,0bTm/ Tc • 104L [# spins] α 0.01 α 0.05a normalized SSCY IG thickness (nm) Series 1 Series 2 Series 3 Figure 5. Comparison between the theoretical and the experimental results. (a) Magnon temperature Tmat the cold end of the temperature gradient as a function of the length Lof the temperature gradient for two dierent damping constants shows a saturation eect depending on the propagation length of the thermally excited magnons. (b)Normalized SSC data and corresponding t functions plotted against the YIG thickness. SSC data have been normalized to the saturation value for an innitely large system. From the t we obtain an eective magnon propagation length of 101 nm 5 nm for series 1, 127 nm 44 nm for series 2 and 89 nm 19 nm for series 3. The resulting ts are shown as solid lines in Fig. 5. The calculated values are comparable to other calculations of the mean propagation length of thermally induced magnons [26]. This mean propagation length depends on the9 frequency spectra of the excited magnons, with that on the model parameters and the damping process during their propagation. The latter depends on the damping constant as well as the frequency !and the group velocity of the magnons@!=@ q[27]. The proportionality between the magnon temperature and the measured ISHE-voltage[6] allows us now to evaluate the SSC data points using eq. 3. Each series was evaluated separately as can be seen in Fig. 2b. We obtain a mean eective magnon propagation length of 101 nm 5 nm for series 1, 127 nm 44 nm for series 2, and 89 nm 19 nm for series 3. By this we derive, independent of the dierent interface qualities between the series, an eective mean propagation length for thermally excited magnons of the order of 110 nm 16 nm from all our series. Based on our model, we can now explain the behavior of the SSC data qualitatively: The increase of the SSC with increasing YIG lm thickness below 110 nm can be attributed to an increasing magnon accumulation at the interface, while the accumulation starts to saturate and therefore the ISHE-voltage in thicker lms. Consequently we can assume that the magnon emitting source is the ferromagnetic thin lm and thus we can pinpoint the origin of the observed signal to the magnonic spin Seebeck eect. In conclusion, we have observed an increasing and saturating spin Seebeck signal with increasing YIG lm thickness. This behavior can neither be explained by the thickness dependence of the saturation magnetization nor magnetore- sistive eects in the Pt detection layer or any other interface eect. Instead we present a model based on atomistic simulations that attributes this characteristic behavior to a nite propagation length of thermally excited magnons, which are created in the bulk of the ferromagnetic material. From the evaluation of our data we obtain an eective mean propagation length of the order of 110 nm for thermally excited magnons, which is in agreement with other studies predicting a nite propagation length of thermally excited magnons of the order of 100 nm [27]. Our results thus clearly allow us to rule out parasitic interface eects and identify thermal magnonic spin currents as the source of the observed signals and thus identify unambiguously the longitudinal spin Seebeck eect. METHODS Thin lm Y 3Fe5O12samples were grown by pulsed laser deposition from a stoichiometric powder target, using a KrF excimer laser ( = 248 nm) with a
uence of 2 :6 J/cm2, and repetition rate of 10 Hz[28]. Monocrystalline 10 mm10 mm0:5 mm gadolinium gallium garnet (Gd 3Ga5O12,GGG) substrates in the (100) crystalline orientation were used to ensure an epitaxial growth of the lms, due to the small lattice mismatch below 0 :06 %. The optimal deposition conditions were found for a substrate temperature of 650C30C and an oxygen partial pressure of 6:6710 3mbar. In order to improve the crystallographic order and to reduce oxygen vacancies, every lm was ex-situ annealed at 820C30C by rapid temperature annealing for 300 s under a steady
ow of oxygen. X-ray re
ectometry (XRR) and prolometer (Tencor P-16 Surface Prolometer) measurements were done to determine the lm thickness, while the crystalline quality was measured by X-ray diraction. The samples are sorted into series to highlight the dierent platinum deposition conditions and therefore interface10 qualities, which have been used to study interface in
uence. This in
uence is minimized for samples within a series by sputtering and cleaning these samples at the same time. Therefore the Pt thickness and the interface preparation were kept identical, while the YIG lm thickness varied. Between the series the interface preparations and thus qualities dier and lead to dierent spin mixing conductance and thus dierent signal amplitudes for a given thermally excited spin current. For the Pt deposition the samples had been transferred at atmosphere and may therefore have suered from surface contamination. In order to enhance the interface quality, an in-situ low power ion etching of the YIG surface was performed for some series prior the deposition. DC-magnetron sputtering was used for a homogeneous deposition of the Pt-lm under an argon pressure of 1 10 2mbar at room temperature. XRR measurements were done afterwards to control the Pt thickness. In the last fabrication step, the Pt-layer was structured by optical lithography and ion etching in order to reduce in
uences on the ISHE-voltage by slight variations of the sample geometry. Fig. 6a shows a sketch of the nal sample stack. V 𝐻 𝛻𝑇 Θ 5 mm GGG (100) YIG (100) x z 500 µm -60 -30 0 30 60-8-6-4-202468 V Voltage (µV) H-Field (Oe) 11K 9K 7K 5K 3K 1K 036912-6-3036 T|| -z T|| -z T|| z VISHE (µV) Tz(K) T|| -za b Figure 6. Experimental conguration and measured signals. (a) Sketch of the sample conguration geometry. The structured grey layer, indicates the 4 mm long platinum stripes with 1 :2 mm large triangular shaped contact pads. The YIG layer is indicated by yellow, the GGG substrate by light grey. For a further understanding of the spin caloric measurements the direction of thermal gradient and the in-plane magnetic eld have also been marked. (b): Recorded voltage signals for the 200 nm thick YIG lm of series 2. Each color represents a dierent stable temperature dierence. The inset shows the evaluated ISHE-voltage VISHE for both directions of the thermal gradient Tz. For the thermoelectric transport measurements a setup was constructed that is able to generate a temperature dierence up to 15 K at room temperature in the parallel and anti-parallel out-of-plane direction. Two copper blocks can either serve as heat source or cooling bath to establish a temperature dierence between both blocks, while the sample is mounted in between. The relative temperature dierence, which was used in the graphs and for the calculation of the spin Seebeck coecient, was determined by the dierence between those two copper blocks. To ensure a good thermal connection, each sample was mounted with thermally conductive adhesive transfer tape. In addition the tape compensated misalignments of the sample mounting. Due to the pressure sensitive heat conduction of the tape, springs were used to mechanically press the two copper blocks together with a constant force in order to reproduce the same conditions for every measurement. Magnons, generated by the thermal gradient in the ferromagnetic layer, will now propagate, depending on the11 orientation of the thermal gradient, to the FMI/NM interface. An exchange interaction of the local moments of the FMI and the conduction electrons of the NM leads to a spin transfer torque, which creates spin-polarized charge carrier in the NM[13]. Due to the inverse spin-Hall eect in Pt, a charge carrier separation, based on the spin orientation, is taking place, leading to a measureable potential dierence at the edges of the stripe geometry[12, 13]. Our setup used a two point-conguration, shown in Fig. 6a, to detect this voltage by a nanovoltmeter (Keithley 2182A). By sweeping the in-plane magnetic eld with = 90with respect to the platinum stripes in addition to an out-of-plane thermal gradient, one is able to measure magnetic eld dependence of the measured voltage as shown as in Fig. 6b. To exclude in
uences of a ground oset, the ISHE-voltage VISHE needs to be extracted from these voltages signals by dividing the dierence of the voltage in saturation for positive and negative H-eld of the signals, V, by two. If one now plots the ISHE-Voltage against the corresponding out-of-plane thermal dierence Tz, as done in the inset of Fig. 6b the SSC can be derived from the slope by a linear t. Dependent on the direction of the thermal gradient the SSC will switch its sign, while the absolute value should be the same. The values of the order of 0 :54V=K (200 nm thick YIG lm with 8 :5 nm of Pt) derived for the SSC, in our setup and sample preparation, are similar to experiments of other groups[5, 8, 11, 17, 29]. A SSC data point for one sample, as shown in Fig. 2, is the result of an average of the measurements of the two stripes per sample and for both directions of the thermal gradient. [1] G. E. W. Bauer, E. Saitoh, and B. J. van Wees, Nature Mat. 11, 391 (2012). [2] M. Walter, J. Walowski, V. Zbarsky, M. M unzenberg, M. Sch afers, D. Ebke, G. Reiss, A. Thomas, P. Peretzki, M. Seibt, J. S. Moodera, M. Czerner, M. Bachmann, and C. Heiliger, Nature Mat. 10, 742 (2011). [3] K. Uchida, S. Takahashi, K. Harii, J. Ieda, W. Koshibae, K. Ando, S. Maekawa, and E. Saitoh, Nature 455, 778 (2008). [4] C. M. Jaworski, J. Yang, S. Mack, D. D. Awschalom, J. P. Heremans, and R. C. Myers, Nature Mat. 9, 898 (2010). [5] K. Uchida, J. Xiao, H. Adachi, J. Ohe, S. Takahashi, J. Ieda, T. Ota, Y. Kajiwara, H. Umezawa, H. Kawai, G. E. W. Bauer, S. Maekawa, and E. Saitoh, Nature Mat. 9, 894 (2010). [6] J. Xiao, G. E. W. Bauer, K. Uchida, E. Saitoh, and S. Maekawa, Phys. Rev. B 81, 214418 (2010). [7] J. Ohe, H. Adachi, S. Takahashi, and S. Maekawa, Phys. Rev. B 83, 115118 (2011). [8] H. Adachi, K. Uchida, E. Saitoh, J.-i. Ohe, S. Takahashi, and S. Maekawa, Appl. Phy. Lett. 97, 252506 (2010). [9] M. Agrawal, V. I. Vasyuchka, A. A. Serga, A. D. Karenowska, G. A. Melkov, and B. Hillebrands, arXiv:1209.3405. [10] S. Y. Huang, W. G. Wang, S. F. Lee, J. Kwo, and C. L. Chien, Phys. Rev. Lett. 107, 216604 (2011). [11] K. Uchida, H. Adachi, T. Ota, H. Nakayama, S. Maekawa, and E. Saitoh, Appl. Phy. Lett. 97, 172505 (2010). [12] E. Saitoh, M. Ueda, H. Miyajima, and G. Tatara, Appl. Phy. Lett. 88, 182509 (2006). [13] Y. Kajiwara, K. Harii, S. Takahashi, and J. Ohe, Nature 464, 262 (2010). [14] M. B. Jung
eisch, V. Lauer, R. Neb, A. V. Chumak, and B. Hillebrands, arXiv:1302.6697. [15] O. Mosendz, V. Vlaminck, J. E. Pearson, F. Y. Fradin, G. E. W. Bauer, S. D. Bader, and A. Homann, Phys. Rev. B 82, 214403 (2010). [16] S. Y. Huang, X. Fan, D. Qu, Y. P. Chen, W. G. Wang, J. Wu, T. Y. Chen, J. Q. Xiao, and C. L. Chien, Phys. Rev. Lett. 109, 107204 (2012).12 [17] M. Weiler, M. Althammer, F. Czeschka, H. Huebl, M. Wagner, M. Opel, I.-M. Imort, G. Reiss, A. Thomas, R. Gross, and S. Goennenwein, Phys. Rev. Lett. 108, 106602 (2012). [18] Y. M. Lu, Y. Choi, C. M. Ortega, X. M. Cheng, J. W. Cai, S. Y. Huang, L. Sun, and C. L. Chien, Phys. Rev. Lett. 110, 147207 (2013). [19] D. Qu, S. Y. Huang, J. Hu, R. Wu, and C. L. Chien, Phys. Rev. Lett. 110, 067206 (2013). [20] H. Nakayama, M. Althammer, T. Y. Chen, K. Uchida, Y. Kajiwara, D. Kikuchi, T. Ohtani, S. Gepr ags, M. Opel, S. Takahashi, R. Gross, G. E. W. Bauer, S. T. B. Goennenwein, and E. Saitoh, arXiv:1211.0098. [21] N. Vliestra, J. Shan, V. Castel, B. J. van Wees, and J. B. Youssef, arXiv:1301.3266v1. [22] S. Geprags, S. Meyer, S. Altmannshofer, M. Opel, F. Wilhelm, A. Rogalev, R. Gross, and S. T. B. Goennenwein, Appl. Phy. Lett. 101, 262407 (2012). [23] A. A. Serga, A. V. Chumak, and B. Hillebrands, J. Phys. D: Appl. Phys. 43, 264002 (2010). [24] N. Kumar, D. S. Misra, N. Vankataramani, S. Prasad, and R. Krishnan, J. Magn. Magn. Mater. 272-276 , e899 (2004). [25] U. Nowak, \Handbook of magnetism and advanced magnetic materials," (John Wiley & Sons, 2007) Chap. Classical Spin-Models. [26] U. Ritzmann, D. Hinzke, and U. Nowak, (in preparation). [27] A. A. Kovalev and Y. Tserkovnyak, Europhys. Lett. 97, 67002 (2012). [28] L. Bi, J. Hu, P. Jiang, D. H. Kim, G. F. Dionne, L. C. Kimerling, and C. A. Ross, Nature Photon. 5, 758762 (2011). [29] A. Kirihara, K. Uchida, Y. Kajiwara, M. Ishida, Y. Nakamura, T. Manako, E. Saitoh, and S. Yorozu, Nature Mat. 11, 686 (2012). Author contributions M.K. and G.J. conceived and supervised the research. A.K. and R.R. performed the experiments and analyzed the data. M.C.O., D.H.K. and C.A.R. provided and characterized the YIG lms A.K. and R.R. structured and characterized the samples. U.R., D.H. and U.N. worked on the theoretical modelling. B.J. and B.H. performed additional sample analysis. A.K. organized and wrote the paper with all authors contributing to the discussions and preparation of the manuscript. ACKNOWLEDGMENTS The authors would like to thank the Deutsche Forschungsgemeinschaft (DFG) for nancial support via SPP 1538 "Spin Caloric Transport" and the Graduate School of Excellence Materials Science in Mainz (MAINZ) GSC 266, the EU (IFOX, NMP3-LA-2012246102, MAGWIRE, FP7-ICT-2009-5 257707, MASPIC, ERC-2007-StG 208162) and the National Science Foundation. | 2013-06-04 | The observation of the spin Seebeck effect in insulators has meant a
breakthrough for spin caloritronics due to the unique ability to generate pure
spin currents by thermal excitations in insulating systems without moving
charge carriers. Since the recent first observation, the underlying mechanism
and the origin of the observed signals have been discussed highly
controversially. Here we present a characteristic dependence of the
longitudinal spin Seebeck effect amplitude on the thickness of the insulating
ferromagnet (YIG). Our measurements show that the observed behavior cannot be
explained by any effects originating from the interface, such as magnetic
proximity effects in the spin detector (Pt). Comparison to theoretical
calculations of thermal magnonic spin currents yields qualitative agreement for
the thickness dependence resulting from the finite effective magnon propagation
length so that the origin of the effect can be traced to genuine bulk magnonic
spin currents ruling out parasitic interface effects. | Determination of the origin of the spin Seebeck effect - bulk vs. interface effects | 1306.0784v1 |
Quantum drives produce strong entanglement between YIG samples without using intrinsic nonlinearities Jayakrishnan M. Prabhakarapada Nair1,and G. S. Agarwal1, 2,y 1Institute for Quantum Science and Engineering, Texas A &M University, College Station, TX 77843, USA 2Department of Biological and Agricultural Engineering, Department of Physics and Astronomy, Texas A &M University, College Station, TX 77843, USA (Dated: July 15, 2019) We show how to generate an entangled pair of yttrium iron garnet (YIG) samples in a cavity-magnon system without using any nonlinearities which are typically very weak. This is against the conventional wisdom which necessarily requires strong Kerr like nonlinearity. Our key idea, which leads to entanglement, is to drive the cav- ity by a weak squeezed vacuum field generated by a flux-driven Josephson parametric amplifier (JPA). The two YIG samples interact via the cavity. For modest values of the squeezing of the pump, we obtain significant en- tanglement. This is the principal feature of our scheme. We discuss entanglement between macroscopic spheres using several di erent quantitative criteria. We show the optimal parameter regimes for obtaining entanglement which is robust against temperature. We also discuss squeezing of the collective magnon variables. Yttrium iron garnet (YIG), an excellent ferrimagnetic sys- tem, has attracted considerable attention during the past few years. The Kittel mode [1] in YIG possesses unique properties including rich magnonic nonlinearities [2] and a low damping rate [3] and in addition the high spin density in YIG allows strong coupling between magnons and microwave cavity pho- tons giving rise to quasiparticles, namely the cavity-magnon polaritons [3–8]. Strong coupling between the YIG sphere and the cavity photons have been observed at both cryogenic and room temperatures [8]. Aided by these superior proper- ties, YIG is reckoned to be the key ingredient in future quan- tum information networks [9]. Thus a variety of intriguing phenomena have been investigated in the context of magnons. This include the observation of bistability [10], cavity spin- tronics [7, 11], level attraction for cavity magnon-polaritons [12], magnon dark modes [13], the exceptional point [14] etc. By virtue of the strong coupling among magnons, a multi- tude of quantum information aspects have been investigated including the coupling of magnons to a superconducting qubit [15] and phonons [16]. Other interesting phenomena involve magnon induced transparency [17], magnetically controllable slow light [18] etc. Owing to the diverse interactions of magnons with other in- formation carriers, YIG o er a novel platform in the analysis of macroscopic quantum phenomena. The coherent phonon- magnon interactions due to the radiation pressure like mag- netostrictive deformation [19] was studied. The nonlinear interaction between magnons and phonons can give rise to magnomechanical entanglement which further transfers to photon-magnon and photon-phonon subsystems, generating a tripartite entangled state [20]. Another recent work pro- posed a scheme to create squeezed states of both magnons and phonons in a hybrid magnon-photon-phonon system [21]. The squeezing generated in the cavity was transferred to the magnons via the cavity-magnon beamsplitter interaction. There is not much work on the coupling of two macroscopic YIG samples in a cavity. Recently the spin current genera- tion in a YIG sample due to excitation in another YIG sam- ple has been investigated [11]. This arises from the cavitymediated coupling between the two samples. It is thus nat- ural to consider the possibility of quantum entanglement be- tween two YIG samples as there has been significant interest in the study of quantum entanglement between macroscopic systems. Recently there has been remarkable success in the observation of quantum entanglement between macroscopic mechanical oscillators [22, 23] with photonic crystal cavities and with superconducting quibits. In addition entanglement between cavity field and mechanical motion has been reported [24]. The conventional wisdom of producing entanglement involves nonlinearities in the system. The well known nonlin- earities are the magnetostrictive interaction [16] and the Kerr eect [2]. The magnetostrictive force allows the magnons to couple to the phonons and can be used to generate magnon- phonon entanglement [20]. The Kerr nonlinearity arises from the magnetocrystalline anisotropy and has been used to pro- duce bistability in magnon-photon systems. In recent publica- tions, these nonlinearities have been used to produce entangle- ment between two magnon modes in a magnon-cavity system [25–27]. Here we present a scheme to generate an entangled pair of YIG spheres in a cavity-magnon system without using any nonlinearities. In addition, we also investigate the squeezed states of the coupled system of two YIG spheres. Two YIG spheres are coupled to the cavity field and the cavity is driven by a squeezed vacuum field [30, 31], resulting in a squeezed cavity field. A flux-driven Josephson parametric amplifier (JPA) is used to generate the squeezed vacuum microwave field. The squeezing in the cavity will be transferred to the two YIG samples due to the cavity-magnon beamsplitter in- teraction. Based on experimentally attainable parameters, we show that significant bipartite entanglement can be generated between the YIG samples. The entanglement is robust against temperature. Our results can be extended to other geometries of YIG. Further the method that we propose is quite generic and can be used for other macroscopic systems. We consider the cavity-magnon system [16, 19, 20] which consists of cavity microwave photons and magnons, as shown in figure 1. The magnons are quasiparticles, a collective ex-arXiv:1905.07884v2 [quant-ph] 11 Jul 20192 Signal Pump Output YIG YIG YIG FIG. 1: Two YIG spheres are placed inside a microwave cav- ity near the maximum magnetic field of the cavity mode, and si- multaneously in a uniform bias magnetic field. The cavity is driven by a week squeezed vacuum field generated by a flux-driven JPA. The magnetic field of the cavity mode is in the xdirec- tion and the bias magnetic field is applied along the zdirection. citation of a large number of spins in a YIG sphere. They are coupled to the cavity photons via the magnetic dipole interac- tion. The Hamiltonian of the system reads [32] H=~=!aaya+!m1my 1m1+!m2my 2m2 +gm1a(a+ay)(m1+my 1)+gm2a(a+ay)(m2+my 2);(1) where a(ay) are the annihilation (creation) operator of cav- ity mode, m1,m2(my 1,my 2) are the annihilation (creation) operators of the two magnon modes and they represent the collective motion of spins via the Holstein-Primako trans- formation [33] in terms of Bosons. The parameters !a,!mi (i=1,2) are the resonance frequencies of the cavity and the magnon modes. Hereafter, wherever we use a subscript ‘ i’ it can take values from 1 to 2. The magnon frequency is given by the expression !mi=
Hi, where
=2=28 GHz /T is the gyromagnetic ratio and Hiare the external bias magnetic fields. The gmiain Eq.(1) are the linear photon-magnon cou- pling strengths. The cavity is driven by a week squeezed vac- uum field generated by a flux driven JPA. JPAs can in prin- ciple amplify a single signal quadrature without adding any extra noise. The squeezed vacuum is generated by degenerate parametric down-conversion using the nonlinear inductance of Josephson junctions [34–44] and a squeezing down to 10% of the vacuum variance has been produced [36]. The oper- ation of generating squeezed vacuum is depicted in figure 1. Vacuum fluctuations are at the signal port and the pump field is applied at frequency 2 !s. The pump photon splits into a signal and an idler photon. Strong quantum correlations be- tween the signal and idler photons are generated which re- sult in squeezing. The output is at the frequency !s[37, 40]. The Hamiltonian described by Eq.(1) does not contain terms involving the input drive field. We use standard quantum Langevin formalism to model the system and the equations describing the evolution of the system operators will contain the input drive terms. Applying the rotating-wave approxima-tiongmia(a+ay)(mi+my i) becomes gmia(amy+aym) [3–7, 16]. In the rotating frame at the frequency !sof the squeezed vac- uum field, the quantum Langevin equations (QLEs) describ- ing the system can be written as follows ˙a= (ia+ka)a igm1am1 igm2am2+p 2kaain; ˙m1= (im1+km1)m1 igm1aa+p 2km1min 1; (2) ˙m2= (im2+km2)m2 igm2aa+p 2km2min 2; where a=!a !s,mi=!mi !s,kais the dissipation rate of the cavity, kmiare the dissipation rates of the magnon modes, andain,min iare the input noise operators of the cavity and magnon modes, respectively. The input noise operators are characterized by zero mean and the following correlation re- lations [45],hain(t)ainy(t0)i=(N+1)(t t0),hainy(t)ain(t0)i= N(t t0),hain(t)ain(t0)i=M(t t0),hainy(t)ainy(t0)i= M(t t0), whereN=sinh2r,M=eisinhrcosh rwith r andbeing the squeezing parameter and the phase of the input squeezed vacuum field, respectively. We have the other input correlations for the magnon as hmin i(t)miny i(t0)i= [Nmi(!mi)+1](t t0),hminy i(t)min i(t0)i=Nmi(!mi)(t t0), where Nmi(!mi)=[exp(~!mi kBT) 1] 1are the equlibrium mean thermal magnon numbers of the two magnon modes. We now show that the YIG spheres can be entangled by resonantly driving the cavity with a squeezed vacuum field. We write down the field operators as their steady state values plus the fluctuations around the steady state. The fluctuations of the system can be described by the QLEs ˙a= (ia+ka)a igm1am1 igm2am2+p 2kaain; ˙m1= (im1+km1)m1 igm1aa+p 2km1min 1; (3) ˙m2= (im2+km2)m2 igm2aa+p 2km2min 2: The quadratures of the cavity field and the two magnon modes are given by X=(a+ay)=p 2,Y=i(ay a)=p 2,xi= (mi+my i)=p 2 andyi=i(my i mi)p 2, and similarly for the input noise operators. The QLEs describing the quadrature fluctuations ( X;Y;x1;y1;x2;y2) can be written as ˙u(t)=Au(t)+n(t); (4) where u(t)=[X(t);Y(t);x1(t);y1(t);x2(t);y2(t)]T, n(t)=[p2kaXin;p2kaYin;p 2km1xin 1;p 2km1yin 1;p 2km2xin 2;p 2km2yin 2]Tand A=2666666666666666666666664 kaa 0 gm1a 0 gm2a a ka gm1a0 gm2a0 0 gm1a km1m1 0 0 gm1a0 m1 km10 0 0 gm2a 0 0 km2m2 gm2a0 0 0 m2 km23777777777777777777777775: (5) The system is a continuous variable (CV) three- mode Gaus- sian state and it can be completely described by a 6 6 covariance matrix (CM) Vdefined as V(t)=1 2hui(t)uj(t0)+3 (a) (b) FIG. 2: Density plot of bipartite entanglement Em1m2between the two magnon modes versus aandm1(a) with m2= m1, r=1,=0,T=20 mK, (b) with m2= m1,r=2, =0,T=20 mK. Other parameters are given in the text. 00.050.10.150.20.250.3 Temperature (K)0.40.50.60.70.80.9Em1m2 FIG. 3: Plot of bipartite entanglement Em1m2be- tween the two magnon modes against temperature with a= m1= m2=0,r=2 and=0: uj(t0)ui(t)i, (i,j=1, 2....6). The steady state CM Vcan be obtained by solving the Lyapunov equation [46, 47] AV+VAT= D; (6) where Dis the di usion matrix defined as hni(t)nj(t0)+ nj(t0)ni(t)i=2=Di j(t t0). We use logarithmic negativ- ity [48] as the quantitative measure to investigate the bi- partite entanglement Em1m2between the two magnon modes. It can be obtained from Em1m2=max[0; ln(2 )] where =min[eig(i P12VP 12)], = iyL iy,P12=1L z andy,zare the Pauli matrices [49]. Figure 2(a)-(b) shows the bipartite entanglement between the two magnon modes at two di erent squeezing parameters. We use a set of ex- perimentally feasible parameters [16]: !a=2=10 GHz, ka=2=5kmi=2=5 MHz, gm1a=gm2a=4kaandT=20 mK, Nm1=Nm20 at 20 mK. The YIG sphere has a diam- eter 250-mand the number of spins N3:51016. We have adopted the parameters so that the two magnon modes are identical. We observe that a= m1= m2=0, in other words!a=!s,!mi=!sare optimal for the entanglement between the two YIG samples. At resonance we observe the maximum amount of entanglement and it increases with the increase in the squeezing parameter. Figure 3 shows that the (a) (b) FIG. 4: (a)hM2 xi+hm2 yiagainst aandm1with m2= m1, r=2,=0,T=20 mK. (b)hM2 xiagainst aand squeez- ing parameter rwith m1= m2=0,=0, T=20 mK. bipartite entanglement is quite robust against temperature. We observe significant amount of entanglement even at T=0.5 K which is quite remarkable for the system of two YIG spheres. We have chosen identical coupling between photon and the two magnon modes. In the case of unequal coupling the en- tanglement goes down. Although we have chosen two identi- cal YIG spheres, one can have two cuboidal YIG samples as in [11] with an angle between the external magnetic field and the local microwave magnetic field at one YIG sample. This makes the resonance frequencies of the two samples di erent. To compare our results with the protocols using nonlinear methods, a recent work [26] produced an entanglement close to 0.25 between the magnon modes at a temperature 10 mK through a Kerr nonlinearity introduced by a strong classical drive. The use of a di erent kind of nonlinearity, namely the magnetostrictive interaction in one YIG sphere produces sim- ilar entanglement [25] at a temperature 10 mK. The entan- glement vanishes as the temperature approaches 20 mK. In contrast our scheme for entanglement generation produces a steady and strong entanglement between 0 to 100 mK and a significant amount of entanglement is present even at 500 mK. The mechanism of the entanglement generation will become clear from the discussion below. Next we discuss two di erent criteria for entanglement in a two mode CV system. The advantage of these criteria over logarithmic negativity is that the former can be easily exam- ined through experiments [22, 23], though in a qualitative way. The first inseparability condition proposed by Simon [49] and Duan et al. [50] is the su cient condition for entan- glement in a two mode CV system. We define a new set of operators M=(m1+m2)=p 2,m=(m1 m2)=p 2. The cri- terion suggests that if the two modes are separable then they should satisfy the following inequality hM2 xi+hm2 yi1; (7) whereMxandmyare the fluctuations in the quadratures Mx andmydefined as Mx=(M+My)=p 2,my=i(my m)=p 2. In other words, violation of the inequality in Eq.(7) means the existance of entanglement between the two YIG samples. Fig- ure 4(a) shows that there is region around a=0 and m1=0 (resonance) in which hM2 xi+hm2 yiis less than one and it is4 a clear manifestation of the entanglement present between the YIG samples. Mancini et al. [51] derived another inequality which is useful in characterizing separable states. It suggests that if the two mode CV system is separable, then it should satisfy the following inequality hM2 xihm2 yi1=4: (8) Hence the violation of Eq.(8) implies that the YIG samples are entangled. We use identical coupling strengths between the cavity and the two YIG samples. Therefore when m1= m2=0 the Hamiltonian of the system in the rotating frame of the drive can be written as H=~= aaya+p 2gm1a(a+ay)(M+My): (9) The Hamiltonian does not contain a term involving mandmy. Hence the fluctuations in mwill be equal to the fluctuations at time t=0. Since matt=0 is in the vacuum state (at low temperature 20 mK), we have hm2 yi=1=2. Figure 4(b) shows that there is a region close to resonance where the quan- tityhM2 xiis less than 1 =2. This violates the inequality in Eq.(8) and hence the two YIG samples are entangled. This further corroborates our results. As a byproduct of our results we investigate the squeezing of the two magnon modes and show that it can be acheived by resonantly driving the cavity with a squeezed vacuum field. We are interested in the vari- ances of the cavity and magnon mode quadratures and they are given by diagonal elements of the time-dependent CM V(t) as defined previously. The amount of squeezing in a mode quadrature Xcan be expressed in decibels (dB). It is obtained from the expression 10log10[hX(t)2i=hX(t)2ivac], wherehX(t)2ivac=1 2. As discussed in [21] when the cav- ity and the two magnon modes are decoupled, the cavity field is squeezed as a result of the squeezed driving field and the magnon modes possesses vacuum fluctuations. As we in- crease the coupling strength, squeezing is partially trans ered to the two identical YIG samples. The blue region in figure 5(a)-(b) represents the region of squeezing. For r=2 the input squeezing is about 17.35 dB. We observed a squeezing of about 2.27 dB for each of the two magnon modes at res- onance with T=20 mK. Note that figure 5 give the magnon quadrature when both the YIG samples are present. Figure 6(a) shows that the magnon squeezing is robust against tem- perature. We observe moderate squeezing for both spheres even at T=0.35 K. At resonance we also find a squeezing of about 7.28 dB for the Mxquadrature of the collective variable M. This is comparable to the results when one had only one YIG sample present and clearly manifested in figures 6(b)-(c). In conclusion, We have presented a scheme to generate an entangled pair of YIG samples in a cavity-magnon system. Entanglement of magnon modes can be generated through res- onantly driving the cavity by a squeezed vacuum field and it can be realized using experimentally attainable parameters. The entanglement produced is robust against temperature. We observe considerable amount of entanglement even at T=0:5K. We have also discussed possible strategies to mea- sure the generated entanglement. We have also showed that (a) (b) FIG. 5: (a) Variance of the first magnon quadrature hx1(t)2iver- susaandm1. (b) Variance of the first magnon quadrature against squeezing parameter rand phase. The other parameters in (a) are r=2,=0,m1= m2,T=20 mK. Other parameters in (b) are a= m1= m2=0 and T=20 mK.hx2(t)2iis identical tohx1(t)2i. (a) (b) (c) FIG. 6: (a) Variance of the first magnon quadrature hx1(t)2i against squeezing parameter rand temperature Twhen both YIG samples are present. (b) Variance of the first magnon quadrature hx1(t)2iagainst squeezing parameter rand temperature Twith only one YIG sample is present. (c) Variance hM2 xiof the col- lective variable Magainst squeezing parameter rand temperature T. The other parameters are a= m1= m2=0,=0. by employing the same method squeezed states of magnons in two di erent modes can be achieved. For an input squeez- ing of 17.35 dB we have observed a squeezing of about 2.27 dB for the magnon modes at T=20 mK. Our scheme for entangling YIG samples does not require any nonlinearities and hence goes against the conventional wisdom of producing entanglement. This provides an entirely new method for entangling macroscopic systems, which can be used in other macroscopic systems. ACKNOWLEDGMENTS Jayakrishnan would like to thank Jie Li for helpful discus- sions, carefully reading the article and providing constructive feedback.5 jayakrishnan00213@tamu.edu ygirish.agarwal@tamu.edu [1] C. Kittel, On the theory of ferromagnetic resonance absorption , Phys. Rev. 73, 155 (1948). [2] Y .-P. Wang, G.-Q. Zhang, D. Zhang, X.-Q. Luo, W. Xiong, S.-P. Wang, T.-F. Li, C.-M. Hu, and J. Q. You, Magnon Kerr e ect in a strongly coupled cavity-magnon system , Phys. Rev. B 94, 224410 (2016). [3] X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, Strongly cou- pled magnons and cavity microwave photons , Phys. Rev. Lett. 113, 156401 (2014). [4] Y . Tabuchi, S. Ishino, T. Ishikawa, R. Yamazaki, K. Usami, and Y . Nakamura, Hybridizing ferromagnetic magnons and mi- crowave photons in the quantum Limit , Phys. Rev. Lett. 113, 083603 (2014). [5] H. Huebl, C. W. Zollitsch, J. Lotze, F. Hocke, M. Greifenstein, A. Marx, R. Gross, and S. T. B. Goennenwein, High coopera- tivity in coupled microwave resonator ferrimagnetic insulator hybrids , Phys. Rev. Lett. 111, 127003 (2013). [6] M. Goryachev, Warrick G. Farr, Daniel L. Creedon, Yaohui Fan, Mikhail Kostylev, and Michael E. Tobar, High-cooperativity cavity QED with magnons at microwave frequencies , Phys. Rev. Appl. 2, 054002 (2014). [7] L. Bai, M. Harder, Y . P. Chen, X. Fan, J. Q. Xiao, and C.-M. Hu, Spin pumping in electrodynamically coupled magnon-photon systems , Phys. Rev. Lett. 114, 227201 (2015). [8] D. Zhang, X.-M. Wang, T.-F. Li, X.-Q. Luo, W. Wu, F. Nori, and J. Q. You, Cavity quantum electrodynamics with ferromag- netic magnons in a small yttrium-iron-garnet sphere , npj Quan- tum Information 1, 15014 (2015). [9] D. L.-Quirion, Y . Tabuchi, A. Gloppe, K. Usami, and Y . Naka- mura, Hybrid quantum systems based on magnonics , Applied Physics Express 12, 070101 (2019). [10] Y .-P Wang, G.-Q Zhang, D. Zhang, T.-F. Li, C.-M. Hu, and J. Q. You, Bistability of cavity magnon polaritons , Phys. Rev. Lett.120, 057202 (2018). [11] L. Bai, M. Harder, P. Hyde, Z. Zhang, C.-M. Hu, Y . P. Chen, and J. Q. Xiao, Cavity mediated manipulation of distant spin currents using a cavity-magnon-polariton , Phys. Rev. Lett. 118, 217201 (2017). [12] M. Harder, Y . Yang, B. M. Yao, C. H. Yu, J. W. Rao, Y . S. Gui, R. L. Stamps, and C.-M. Hu, Level attraction due to dissipative magnon-photon coupling , Phys. Rev. Lett. 121, 137203 (2018). [13] X. Zhang, C.-L Zou, N. Zhu, F. Marquardt, L. Jiang, and H. X. Tang, Magnon dark modes and gradient memory , Nat. Com- mun. 6, 8914 (2015). [14] D. Zhang, X.-Q. Luo, Y .-P. Wang, T.-F Li, and J. Q. You, Ob- servation of the exceptional point in cavity magnon-polaritons , Nat. Commun. 8, 1368 (2017). [15] Y . Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Yamazaki, K. Usami, and Y . Nakamura, Coherent coupling between a fer- romagnetic magnon and a superconducting qubit , Science. 349, 405 (2015). [16] X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, Cavity mag- nomechanics , Sci. Adv. 2, e1501286 (2016). [17] B. Wang, Z.-X. Liu, C. Kong, H. Xiong, and Y . Wu, Magnon- induced transparency and amplification in PT-symmetric cavity-magnon system , Opt. Express 26, 20248 (2018). [18] C. Kong, B. Wang, Z.-X. Liu, H. Xiong, and Y . Wu, Magneti- cally controllable slow light based on magnetostrictive forces , Opt. Express. 27, 5544 (2019).[19] C. Kittel, Interaction of spin waves and ultrasonic waves in fer- romagnetic crystals , Phys. Rev. 110, 836 (1958). [20] J. Li, S.-Y . Zhu, and G. S. Agarwal, Magnon-photon-phonon entanglement in cavity magnomechanics , Phys. Rev. Lett. 121, 203601 (2018). [21] J. Li, S.-Y . Zhu, and G. S. Agarwal, Squeezed states of magnons and phonons in cavity magnomechanics , Phys. Rev. A 99, 021801(R) (2019). [22] R. Reidinger, A. Wallucks, I. Marinkovi ´c, C. L ¨oschnauer, M. Aspelmeyer, S. Hong, and S. Gr ¨oblacher, Remote quantum en- tanglement between two micromechanical oscillators , Nature (London) 556, 473 (2018). [23] C. F. Ockeloen-Korppi, E. Damsk ¨agg, J.-M. Pirkkalainen, M. Asjad, A. A. Clerk, F. Massel, M. J Wooley, and M. A. Sil- lanp¨a¨a,Stabilized entanglement of massive mechanical oscilla- tors, Nature (London) 556, 478 (2018). [24] T. A. Palomaki, J. D. Teufel, R. W. Simmonds, and K. W. Lehn- ert,Entangling mechanical motion with microwave fields , Sci- ence342, 710 (2013). [25] J. Li and S.-Y . Zhu, Entangling two magnon modes via magne- tostrictive interaction , arXiv:1903.00221v1 [quant-ph]. [26] Z. Zhang, M. O. Scully, G. S. Agarwal, Quantum entan- glement between two magnon modes via Kerr nonlinearity , arXiv:1904.04167v1 [quant-ph]. [27] However these nonlinearities are weak and require strong drives, thus the realization of entanglement is di cult. There have been exceptions especially in the context of superconduct- ing qubits where strong Kerr nonlinearities have been used to produce a variety of cat states [28]. Theoretical description for Kerr nonlinearity produced CAT states is given in [29]. [28] G. Kirchmair, B. Vlastakis, Z. Leghtas, S. E. Nigg, H. Paik, E. Ginossar, M. Mirrahimi, L. Frunzio, S. M. Girvin, and R. J. Schoelkopf, Observation of quantum state collapse and re- vival due to the single-photon Kerr e ect, Nature 495, 205-209 (2013) [29] K. Tara, G. S. Agarwal, and S. Chaturvedi, Production of Schrdinger macroscopic quantum-superposition states in a Kerr medium , Phys. Rev. A. 47, 5024 (1993) [30] K. J ¨ahne, C. Genes, K. Hammerer, M, Wallquist, E. S. Polzik, and P Zoller, Cavity-assisted squeezing of a mechanical oscil- lator , Phys. Rev. A. 79, 063819 (2009). [31] S. Huang and G. S. Agarwal, Entangling nanomechanical os- cillators in a ring cavity by feeding squeezed light , New J. Phys. 11, 103044 (2009). [32] We consider the size of the YIG sphere much smaller than the microwave wavelength and hence the e ect of radiation pres- sure is negligible. The magnon-phonon coupling has been mea- sured [16] to be 8 or 9 orders smaller than the magnon-cavity coupling and hence we can neglect the e ect of magnetostric- tive forces. [33] T. Holstein and H. Primako ,Field dependence of the intrinsic domain magnetization of a ferromagnet , Phys. Rev. 581098 (1940). [34] B. Yurke, Squeezed-state generation using a Josephson para- metric amplifier , J. Opt. Soc. Am. B 4, 1551 (1987); B. Yurke, P. G. Kaminsky, R. E. Miller, E. A. Whittaker, A. D. Smith, A. H. Silver, and R. W. Simon, Observation of 4.2-K equilibrium- noise squeezing via a Josephson-parametric amplifier , Phys. Rev. Lett. 60, 764 (1988); B. Yurke, L. R. Corruccini, P. G. Kaminsky, L.W. Rupp, A. D. Smith, A. H. Silver, R. W. Simon, and E. A Whittaker, Observation of parametric amplification and deamplification in a Josephson parametric amplifier , Phys. Rev. A. 39, 2519 (1989).6 [35] R. Movshovich, B. Yurke, P. G. Kaminsky, A. D. Smith, A. H. Silver, R. W. Simon, and M. V . Schneider, Observation of zero-point noise squeezing via a Josephson-parametric ampli- fier, Phys. Rev. Lett. 65, 1419 (1990). [36] M. A Castellanos- Beltran, K. D. Irwin, G. C. Hilton, L. R. Vale, and K. W. Lehnert, Amplification and squeezing of quantum noise with a tunable Josephson metamaterial , Nat. Phys. 4, 929 (2008). [37] T. Yamamoto, K. Inomata, M. Watanabe, K. Matsuba, T. Miyazaki, W. D. Oliver, Y . Nakamura, and J. S. Tsai, Flux- driven Josephson parametric amplifier , Appl. Phys. Lett. 93, 042510 (2008). [38] F. Mallet, M. A. Castellanos-Beltran, H. S. Ku, S. Glancy, E. Knill, K. Irwin, G. C. Hilton, L. R. Vale, and K. W. Lehn- ert,Quantum State Tomography of an Itinerant Squeezed Mi- crowave Field , Phys. Rev. Lett. 106, 220502 (2011). [39] E. P. Menzel, R. Di Candia, F. Deppe, P. Eder, L. Zhong, M. Ihmig, M. Haeberlein, A. Baust, E. Ho mann, D. Ballester, K. Inomata, T. Yamamoto, Y . Nakamura, E. Solano, A. Marx, and R. Gross, Path entanglement of continuous-variable quantum microwaves , Phys. Rev. Lett. 109, 250502 (2012). [40] L. Zhong, E. P. Menzel, R. Di Candia, P. Eder, M. Ihmig, A. Baust, M. Haeberlein, E. Ho man, K. Inomata, T. Yamamoto, E. Solano, F. Deppe, A. Marx, and R. Gross, Squeezing with a flux-driven Josephson parametric amplifier , New. J. Phys. 15, 125013 (2013). [41] K. G. Fedorov et al .,Displacement of propagating squeezed microwave states , Phys. Rev. Lett. 117, 020502 (2016). [42] S. Kono et al.,Nonclassical photon number distribution in a superconducting cavity under a squeezed drive , Phys. Rev. Lett.119, 023602 (2017). [43] A. Bienfait et al .,Magnetic resonance with squeezed mi- crowaves , Phys. Rev. X. 7, 041011 (2017). [44] M. Malnou, D. A. Palken, Leila R. Vale, Gene C. Hilton, and K. W. Lehnert, Optimal operation of a Josephson parametric amplifier for vacuum squeezing , Phys. Rev. Applied 9044023 (2018). [45] C. W. Gardiner, Inhibition of atomic phase decays by squeezed light: a direct e ect of squeezing , Phys. Rev. Lett. 56, 1917 (1986). [46] D. Vitali, S. Gigan, A. Ferreira, H. R. B ¨ohm, P. Tombesi, A. Guerreiro, V . Vedral, A. Zeilinger, and M. Aspelmeyer, Op- tomechanical entanglement between a movable mirror and a cavity field , Phys. Rev. Lett. 98, 030405, (2007). [47] P. C. Parks and V . Hahn, Stability theory (Prentice Hall, Newyork, 1993). [48] J. Eisert, Ph.D. thesis, University of Potsdam, Potsdam, 2001; G. Vidal and R. F. Werner, Computable measure of entangle- ment , Phys. Rev. A 65, 032314 (2002); M. B. Plenio, Logarith- mic negativity: a full entanglement monotone that is not convex , Phys. Rev. Lett. 95, 090503 (2005). [49] R. Simon, Peres-Horodecki separability criterion for continu- ous variable systems , Phys. Rev. Lett. 84, 2726 (2000). [50] L.-M Duan, G. Giedke, J. I. Cirac, and P. Zoller, Inseparability criterion for continuous variable systems , Phys. Rev. Lett. 84, 2722 (2000). [51] S. Mancini, V . Giovannetti, D. Vitali and P. Tombesi, En- tangling macroscopic oscillators exploiting radiation pressure , Phys. Rev. Lett. 88, 120401 (2002). | 2019-05-20 | We show how to generate an entangled pair of yttrium iron garnet (YIG)
samples in a cavity-magnon system without using any nonlinearities which are
typically very weak. This is against the conventional wisdom which necessarily
requires strong Kerr like nonlinearity. Our key idea, which leads to
entanglement, is to drive the cavity by a weak squeezed vacuum field generated
by a flux-driven Josephson parametric amplifier (JPA). The two YIG samples
interact via the cavity. For modest values of the squeezing of the pump, we
obtain significant entanglement. This is the principal feature of our scheme.
We discuss entanglement between macroscopic spheres using several different
quantitative criteria. We show the optimal parameter regimes for obtaining
entanglement which is robust against temperature. We also discuss squeezing of
the collective magnon variables. | Quantum drives produce strong entanglement between YIG samples without using intrinsic nonlinearities | 1905.07884v2 |
Nonreciprocal Pancharatnam-Berry Metasurface for Unidirectional Wavefront Manipulation Hao Pan1, Mu Ku Chen2, Din Ping Tsai2, and Shubo Wang1,3 * 1Department of Physics, City University of Hong Kong, Tat Chee Avenue, Kowloon, Hong Kong, China. 2Department of Electrical Engineering, City University of Hong Kong, Tat Chee Avenue, Kowloon, Hong Kong, China. 3City University of Hong Kong Shenzhen Research Institute, Shenzhen, Guangdong 518057, China. *Corresponding author: shubwang@cityu.edu.hk ABSTRACT Optical metasurfaces have been widely used for manipulating electromagnetic waves due to their low intrinsic loss and easy fabrication. The metasurfaces employing the Pancharatnam- Berry (PB) geometric phase, called PB metasurfaces, have been extensively applied to realize spin-dependent functionalities, such as beam steering, focusing, holography, etc. The demand for PB metasurfaces in complex environments has brought about one challenging problem, i.e., the interference of multiple wave channels that limits the performance of PB metasurfaces. A promising solution is developing nonreciprocal PB metasurfaces that can isolate undesired wave channels and exhibit unidirectional functionalities. Here, we propose a mechanism to realize nonreciprocal PB metasurfaces of subwavelength thickness by using the magneto- optical effect of YIG material in synergy with the PB geometric phase of spatially rotating meta-atoms. Using full-wave numerical simulations, we show that the metasurface composed of dielectric cylinders and a thin YIG layer can achieve nearly 92% and 81% isolation of circularly polarized lights at 5.5 GHz and 6.5 GHz, respectively, attributed to the enhancement of the magneto-optical effect by the resonant Mie modes and Fabry-Pérot cavity mode. In addition, the metasurface can enable efficient unidirectional wavefront manipulations of circularly polarized lights, including nonreciprocal beam steering and nonreciprocal beam focusing. The proposed metasurface can find highly useful applications in optical communications, optical sensing, and quantum information processing. Ⅰ. INTRODUCTION The recent decade has witnessed significant progress in the design and fabrication of artificial optical structures working at microwave [1,2], terahertz [3,4], infrared [5,6], and visible optical bands [7,8], which can exhibit intriguing electromagnetic (EM) properties not existing in nature [9,10]. One important type is an ultrathin layer of structures known as metasurfaces. Metasurfaces can induce strong light-matter interaction in the nanoscale, and they benefit from small intrinsic loss and easy fabrication compared to conventional bulky metamaterials. By carefully designing the subwavelength elements (i.e., meta-atom) in each unit cell, metasurfaces can give rise to various fascinating wavefront-manipulation functionalities, such as perfect absorption [11-13], structural colors [14,15], anomalous reflection or refraction [16- 18], surface wave excitation [19,20], metalens [21-23], metaholograms [24-26], and many others [27,28]. Different from conventional diffractive optical devices, metasurfaces not only can employ the resonance phase and propagation phase (also called dynamic phase) but also can utilize the PB geometric phase derived from the spatial rotation of meta-atoms [29-33], leading to the so-called PB metasurfaces. Therefore, the PB metasurfaces can acquire an extra degree of freedom to control the wavefront of circularly polarized (CP) light besides the resonance and dynamic phases, giving rise to many intriguing phenomena such as photonic spin Hall effect [34,35], vortex beam generation [36,37], etc. Meanwhile, the PB metasurfaces can serve as a powerful platform for developing CP-light-associated applications, e.g., the CP wave control of the motions of biomolecules exhibiting chiral structures [38,39]. Consequently, the PB metasurfaces have great application potential in the next-generation photonic devices with multifunctionalities. Despite the PB metasurfaces’ unprecedented performance in wavefront manipulation, their functionalities are intrinsically restricted by the Lorentz reciprocity [40]. Introducing additional mechanisms to break the reciprocity of PB metasurfaces can generate new functionalities that are essential to many applications, such as invisible sensing, full-duplex communication, and noise-tolerant quantum computation, etc., where nonreciprocity can prevent the backscattering from defects or boundaries [41-43]. In addition, nonreciprocity allows metasurfaces to exhibit different properties for the opposite propagating waves, thus giving rise to Janus-type functionalities. One effective way to achieve nonreciprocity is by using gyroelectric or gyromagnetic materials sensitive to magnetic-field biasing, which exhibit asymmetric permittivity or permeability tensor accounting for the Faraday-magneto-optical (FMO) effect [44]. Qin et al. proposed a set of self-biased nonreciprocal magnetic metasurfaces to achieve bidirectional wavefront modulation based on the different hybrid resonant-dynamic phase profiles for bidirectional CP waves [45]. However, the local resonant phase can be easily affected by the coupling from neighboring meta-atoms, resulting in an undesired phase profile that limits the performance of the metasurface. In contrast, the PB geometric phase only determined by the rotation angle of meta-atoms is a better mechanism for realizing the stable nonreciprocal wavefront manipulation. Zhao et al. presented an interesting metadevice combining a PB metasurface and an anisotropic metasurface, which can simultaneously realize phase modulation and nonreciprocal isolation [46]. The metadevice involves a complex multilayer structure with a large thickness that inevitably affects its diffraction efficiency. Therefore, simple and thin PB metasurfaces capable of achieving high-efficiency nonreciprocal wavefront manipulation are highly desirable. In this article, we report a nonreciprocal PB metasurface composed of elliptical dielectric cylinders and a thin YIG layer to simultaneously realize PB-phase-based wavefront manipulation and microwave isolation. The thin YIG layer under a magnetic field bias can give rise to strong spin-selective isolation due to the FMO effect and Fabry-Pérot (FP) resonance. Meanwhile, the resonant coupling between the Mie modes in the dielectric cylinder and the FP mode in the YIG layer can effectively tune the nonreciprocal band by enhancing the FMO effect, thus achieving a high isolation ratio of nearly 92% and 81% at 5.5 GHz and 6.5 GHz, respectively. The PB phase of each meta-atom can be individually controlled via the corresponding optical-axis rotation and is unaffected by the FMO effect. Following a digital coding metasurface design methodology, the proposed nonreciprocal PB metasurfaces can offer multiple functionalities with high isolation, which are demonstrated by the tailored metadeflector with nonreciprocal beam steering and the metalens with nonreciprocal focusing. Our proposed all-dielectric nonreciprocal PB metasurface can find applications in multiple fields, e.g., EM wave isolation, nonreciprocal antennas, optical sensing, quantum information processing, etc. Ⅱ. RESULTS AND DISCUSSIONS A. Unidirectional spin-selective nonreciprocal metasurface The metasurface consists of subwavelength meta-atoms arranged in a square lattice with period l, as depicted in Fig. 1(a). Each meta-atom comprises a dielectric cylinder sitting on a YIG substrate of thickness t. Under the external biased magnetic field along the + z-direction, the YIG is characterized by a permeability tensor with asymmetric off-diagonal elements [44] 00 0 0 0 1r r r ri i , (1) where0 2 2 01m r m ,2 2 0m r , 0m sM , 0 0 0H i , γ = 1.579×1011 C/kg is the gyromagnetic ratio, 4π Ms = 1780 G is the saturation magnetization, B0 = μ0H0 = 0.05 T is the external magnetic field, α = 0.002 is the damping factor, and μ0 is the vacuum permeability. The relative permittivity of YIG is εr1 = 15. The dielectric cylinder has relative permittivity εr2 = 24 and relative permeability μr2 = 1. Its major and minor axis are a and b, respectively, and its height is h. The orientation of the cylinder is denoted by the angle θ. In this configuration, the metasurface exhibits different refractive indices for the incident CP waves with the wavevector parallel and antiparallel to the biased magnetic field, attributed to the FMO effect. Thus, the metasurface can give rise to spin- and direction-dependent manipulation of EM waves. To illustrate the physical mechanism, we first analyze the spin-selective transmission of normally incident CP waves in the thin infinite YIG layer with the + z magnetic field bias. The reflection and transmission coefficients for the left-hand circularly polarized (LCP) and right- hand circularly polarized (RCP) EM waves with wavevectors antiparallel to the biased magnetic field can be derived straightforwardly (See Appendix A for the derivations), and their expressions are 2 2 2 221 1 1 1f L f Li k t f Lf RLi k t f f L LY e r Y e Y , 0( ) 2 224 1 1f L f Li k k tf f L LLi k t f f L LY et Y e Y , 2 2 2 221 1 1 1f R f Ri k t f Rf LRi k t f f R RY e r Y e Y , 0( ) 2 224 1 1f R f Ri k k tf f R RRi k t f f R RY et Y e Y , (2) where 1f L r r rY and 1f R r r rY are the relative wave admittances for the LCP and RCP waves forward propagating in the YIG layer, 1 0( )f L r r rk k and 1 0( )f R r r rk k are the corresponding LCP and RCP wavevectors in the YIG layer, and k0 is the wavevector in free space. Here, the superscript “+” denotes the + z-biased magnetic field, “f” denotes forward incidence (i.e., - z-direction), and the subscript “ RL” (“LR”) denotes the CP conversion from LCP to RCP (RCP to LCP). It can be noted from Eq. (2) that the off-diagonal element κr in the permeability tensor results in the different impedances and wavevectors for the LCP and RCP waves, which leads to the differences in the co-polarized transmission (2| |f LLtand 2| |f RRt) and cross-polarized reflection (2| |f RLrand2| |f LRr). Figure 1(b) shows the transmission spectra given by Eq. (2) (denoted by the blue symbol lines). For the considered thin YIG layer, the FP cavity resonance can enhance the FMO effect and increase the transmission difference (i.e., 2 2| | | |f f RR LLt t ) for the LCP and RCP incident waves. This transmission difference can reach a maximum of nearly 94% around 6.8 GHz. To verify the analytical results, we conducted full-wave finite-element simulations by using COMSOL and computed the transmission spectra. The numerical results are denoted by the red symbol lines in Fig. 1(b), which is consistent with the analytical results. In addition, as the forward (− z- direction) normally incident LCP wave is equivalent to the backward (+ z-direction) normally incident RCP wave for the infinite YIG layer with the + z-biased magnetic field, Fig. 1(b) also indicates that the thin YIG layer can exhibit an evident nonreciprocal-transmission feature, i.e., the large transmission contrast between the forward and backward CP waves (See Appendix A for details). Figure 1(c) shows the transmission spectra of the PB metasurface under the incidence of the LCP and RCP plane waves. We set θ=0° for the rotation angle of all the dielectric cylinders. Due to the breaking of cylindrical symmetry by the meta-atoms, the helicity of the wave is not conserved, and the transmitted wave generally contains both LCP and RCP components. We notice that the transmission is dominated by the cross-polarized components f RLt and b LRt for the forward LCP and backward RCP incidence, respectively, which have two resonance peaks at 5.5 GHz and 6.5 GHz with differences (2 2| | | |f b RL LRt t ) of 92% and 81%, respectively. The large isolation ratio can be attributed to the FMO effect of the YIG layer enhanced by the Mie resonance in the dielectric cylinders. To understand the effect of the Mie resonant modes of the cylinders, we show in Fig. 1(d) the numerically calculated multipole decomposition of the cylinder scattering power under the excitation of the forward incident LCP wave (See Appendix C for the multipole decomposition). It is noted that the two resonances at 5.5 GHz and 6.5 GHz are mainly attributed to the magnetic dipole mode and the hybrid magnetic dipole- electric quadrupole mode, respectively. The resonant electric and magnetic field amplitudes are shown in the insets of Fig.1(d). We notice that the magnetic field inside the cylinder is strongly enhanced at 5.5 GHz, while both the electric and magnetic fields are strongly localized in the cylinder at 6.5 GHz due to the resonant electric quadrupole and magnetic dipole resonances. The resonant coupling between these hybrid Mie resonances and the FP cavity resonance in the YIG layer can enhance the interaction between the wave and the magnetic material, leading to the enhanced FMO effect and thus the strong nonreciprocity of the metasurface [47]. We further investigate the relationship between the nonreciprocal properties of the PB metasurface and various system parameters, including the cylinder height h, the biased magnetic field B0, and the incident angle. Figure 2(a) shows the numerically simulated isolation ratio 2 2| | | |f b RL LRt t as a function of the cylinder height h for the system in Fig. 1(a). As seen, the isolation peaks undergo redshift as h increases, which is expected since the eigenfrequencies of the Mie modes in the cylinder are generally inversely proportional to the geometric size of the cylinder. Specifically, as h varies, the spectral profile of the first resonance maintains a Lorentz shape, where the local maximum of the isolation remains above 85%. In contrast, the spectral profile of the second resonance undergoes dramatic variation due to the interference with other multipoles, as evidenced by the sharp transition of isolation from negative to positive values. Figure 2(b) shows the isolation ratio 2 2| | | |f b RL LRt t of the proposed metaisolator when the external magnetic field is B0 = 0.05 T, B0 = 0 T, and B0 = −0.05 T (corresponding to red, magenta, and blue symbol lines, respectively). We notice that the results for different biasing directions are nearly antisymmetric with respect to the case of B0 = 0 T which induces zero isolation. This can be understood as follows. The transmission coefficients follow the relationships f b RL RLt t and b f LR LRt t because the forward normally incident LCP (RCP) wave is converted to RCP (LCP) wave by the elliptical cylinder and the resulting RCP (LCP) wave is equivalent to the LCP (RCP) wave backward normally incident on the YIG layer (similar to the property of single YIG layer mentioned above). In addition, the magnetic field bias direction decides the spin-selective transmission of the metaisolator. For the opposite magnetic biasing, we can obtain the relationships b b RL LRt t and f f LR RLt t (See Appendix B for details). Consequently, we have the relationships f b RL LRt t and b f LR RLt t , and thus 2 2 2 2| | | | | | | |f b f b RL LR RL LRt t t t , i.e., reversing the direction of biased magnetic field leads to a sign change of the isolation value in Fig. 2(b). Figure 2(c) shows the dependence of the isolation on the magnitude of the external magnetic field. We notice that the isolation peaks at 5.5 GHz and 6.5 GHz are blue-shifted without obvious reduction of the isolation ratio, demonstrating the robust performance of the proposed metasurface isolator. We also investigate the effect of the incident angle of CP waves on the isolation. At large incident angles, higher-order diffractions can appear, and we only consider the isolation for the 0th-order cross-polarized transmission under the forward LCP and backward RCP wave incidence with the same incident angle. As depicted in Fig. 2(d), the isolation at the resonance frequency of 5.5 GHz will slightly shift with the increase of the incident angle. At large incident angles, the isolation at 5.5 GHz is reduced owing to the combined effect of the resonance shift and change of CP conversion efficiency in the elliptical cylinder. Notably, the isolation can still reach above 80% for the incident angle as large as 45°. Interestingly, the isolation ratio at 6.5 GHz is insensitive to the variation of the incident angle, and it can maintain a large value above 80% for the incident angle within [0°, 60°]. Therefore, the proposed nonreciprocal metaisolator can achieve a stable and high isolation ratio at the targeted frequencies for a wide range of incident angles, which lays the foundation for further nonreciprocal wavefront manipulations. In addition to manipulating the wave amplitude, the metasurface can also be applied to achieve unidirectional phase manipulation for the transmitted CP wave. This is done by varying the orientational angle θ of the dielectric cylinder to induce PB geometric phases, as shown by the inset in Fig. 3. For CP waves normally forward incident on the metasurface, the output waves can be expressed as 2 ( , ) 2 ( , )0 0out f i x y in L LR L out f i x y in R RL RE t e E E t e E , (3) where f LRt and f RLt are the cross-polarized transmission coefficients for the forward incident RCP and LCP waves, respectively. The superscript “±” denotes the direction of the external biased magnetic field B0. The dielectric cylinder can induce a PB phase shift φ = 2σθ, where σ = +1 (σ = –1) for the LCP (RCP) wave. Figure 3 shows the simulated amplitude of the transmitted electric field (blue symbol line) and the PB phase (red symbol line) for different orientation angles of the cylinder. As seen, the orientation angle θ of the cylinder has a negligible impact on the transmission amplitude, which is around 96% for different rotation angles. Meanwhile, the PB phase agrees with the relationship φ= 2σθ. The stable high CP transmission and the PB phase of 2π range lay the foundation for designing wavefront- manipulation metasurfaces. B. Nonreciprocal PB metadeflector for beam steering Owing to the superior nonreciprocal isolation under the large-angle incidence and the stable PB phase of the meta-atoms, it is possible to construct a nonreciprocal metadeflector with an on-demand phase profile to manipulate the propagation direction of the incident CP beam. Figure 4(a) schematically shows the concept of the nonreciprocal PB-phase-based metadeflector with the + z-biased magnetic field. The meta-atoms are invariant along y direction, but they are orientated differently in the x direction to induce the PB geometric phase profile. At 5.5 GHz, the metasurface can convert the forward incident LCP wave into the RCP wave and deflect it away from the normal direction. Meanwhile, the metasurface can isolate the backward RCP wave incident along the opposite deflection direction, i.e., the time-reversed wave of the deflected RCP wave. The transmitted wavevector and the incident wavevector satisfy the phase-matching condition in the periodic structure [48]: out in PBk k mk , (4) where kout = 2πsinθout/λ, kin = 2πsinθin/λ, kPB = 2π/P, θin and θout are the incident and deflected angles, respectively, λ is the incident wavelength, P is the period size of the supercell (covering 2π phase range) along the y-direction, and m is the deflection order. For the normally incident wave (θin = 0°), Eq. (4) can be simplified as sin θout = mλ/P where the supercell period P = Nl with N being the meta-atom number in the supercell and l being the meta-atom period. The discrete PB phase profile in the supercell can be expressed as φ(n) = 2πn/N where n denotes the n-th meta-atom in the supercell, thus requiring a rotation angle distribution θ(n) = πn/N. Following this principle, we design four different metadeflectors working at 5.5 GHz with the supercells consisting of 4, 6, 8, and 12 meta-atoms, respectively. These metasurfaces induce the 1st-order diffraction at the angles 74.64°, 40°, 28.82°, and 18.75°, respectively. Figure 4(b) shows the simulated electric field ( Ey) profiles at 5.5 GHz for the four metasurfaces. The deflection angles of the output beam are consistent with analytical values given by Eq. (4). Under the forward normal incidence, the 1st-order diffraction efficiency in these four cases is 66.44%, 92.1%, 95.59%, and 96.08%, respectively. Under the backward incidence, the transmission efficiency in the four cases is 6.64%, 0.79%, 0.026%, and 0.002%, respectively. Accordingly, the isolation ratios are 59.8%, 91.31%, 95.564%, and 96.078%, which demonstrate the highly efficient nonreciprocal beam steering function of the proposed metadeflectors. Additionally, we note that for 6-, 8-, and 12-cell cases, the deflected beams are mainly composed of the 1st-order diffraction, while higher-order diffraction components begin to appear in the output beam of the 4-cell case, which can be attributed to the large wavevector component parallel to the metasurface. The emergence of the higher-order diffractions in this case decreases the isolation ratio and leads to a complex output wavefront. C. Nonreciprocal PB metalens for beam focusing The PB-phase-based planar metalenses with excellent performance, e.g., high numerical aperture (NA), have been widely proposed and fabricated, generating broad applications in imaging [49,50], microscopy [51], and spectroscopy [52,53]. However, the effect of backscattering is usually neglected in conventional PB metalenses, thus limiting their applications in the platforms requiring anti-echo and anti-reflection functions. Introducing nonreciprocity to PB metalenses can be a solution to this problem. This corresponds to the concept of nonreciprocal PB metalens for unidirectional beam focusing, as illustrated in Fig. 5(a). The forward normally incident LCP wave passes through the nonreciprocal metalens with the +z-biased magnetic field and is focused into one spot, but the RCP wave radiated from the focusing spot, i.e., the time-reversal excitation, will be blocked by the metalens, thus realizing the nonreciprocal beam focusing. The PB phase profile φ(x,y) of the metalens should follow [49] 2 2 2 2( , )x y f x y f , (5) where λ is the wavelength, f is the focal length, x and y are the coordinates of each meta-atom. Similar to the metadeflector mentioned above, we consider the metalens with invariant phase profile in x-direction. The rotation angle profile of the meta-atoms in this case is 2 2( )y f y f , which has the discretized form 2 2 2( )n f n l f , where n denotes the n-th meta-atom, and l is the period of each meta-atom. To demonstrate the nonreciprocal focusing functionality, we design three metalenses with different focal lengths 1.5λ, 2λ, and 3λ (λ=54.5 mm at 5.5 GHz), respectively. We conduct numerical simulations for the nonreciprocal focusing realized by the three metalenses. Figure 5(b) depicts the simulated electric-field distributions in the yz-plane with the forward incident LCP (the upper panels) and the backward RCP radiation from the focal point (the bottom panels). It is noticed that the forward incident LCP waves are focused into spots at different focal points. The corresponding focal lengths are determined to be 80.62 mm, 108.85 mm, and 149.83 mm, respectively. The discrepancy between the theoretical and simulated focal lengths can be attributed to the coupling effect between the adjacent meta-atoms. Figures 5(c)-(e) show the intensity on the focal planes with the diffraction-limited ( λ/(2×NA)) full width at half-maximum (FWHM) of 30 mm, 30.27 mm, and 31.47 mm, respectively. The corresponding NA of the metalenses is 0.908, 0.9, and 0.866, respectively. To understand the nonreciprocity of the metalenses, we calculate the light transmission under the forward incidence, which reaches 79.9%, 85.1%, and 89.64% for the three cases, respectively. Meanwhile, the focusing efficiency is found to be 68.18%, 73.14%, and 74.51% for the three cases, respectively, where the focusing efficiency is defined as the fraction of the incident light that passes through a circular aperture in the focal plane with a diameter equal to three times of the FWHM spot size [54]. Additionally, we find that the backward RCP radiation from the focal point only gives rise to the transmission of 12.9%, 14.67%, and 10.9%, respectively. Therefore, the isolation ratios of the three metalenses are 55.28%, 58.47%, and 63.61%, respectively. The contrast between the focusing efficiency under forward incidence and the transmission under backward radiation clearly demonstrates the nonreciprocal focusing functionality of the designed PB metalenses. Ⅲ. CONCLUSION To summarize, we have demonstrated that high-performance nonreciprocal wavefront manipulation of CP beams can be achieved by using the magnetic-biased PB metasurfaces consisting of elliptical dielectric cylinders and a thin magnetic YIG layer. Due to the strong resonant coupling between the Mie modes in the cylinders and the FP cavity mode in the thin YIG layer, the FMO effect can be greatly enhanced near the resonant frequencies, thus giving rise to significant spin-selective nonreciprocal isolation. Meanwhile, the stable PB phase and the large isolation ratio over a wide range of incident angles can guarantee efficient nonreciprocal wavefront manipulation. By designing the PB phase gradient profile, we have demonstrated two types of nonreciprocal functional metasurfaces: the metadeflectors that can realize nonreciprocal beam steering with different deflection angles, and the high-NA metalenses that can realize nonreciprocal focusing with different focal lengths. The proposed nonreciprocal PB metasurfaces can simultaneously achieve high-efficiency wavefront manipulation and large isolation ratio, which pave the way to the applications in wave multiplexing for high-capacity communications and optical imaging with anti-reflection functions. ACKNOWLEDGEMENTS The work described in this paper was supported by grants from the Research Grants Council of the Hong Kong Special Administrative Region, China (Projects No. AoE/P-502/20 and No. CityU 11308223). APPENDIX A: SPIN-SELECTIVE TRANSMISSION OF AN INFINITE YIG LAYER The yttrium iron garnet (YIG) material is a common magnetic material that can show obvious asymmetry spin characteristics, e.g., the different propagation constants and impedances for the orthogonal CP states, due to the large off-diagonal elements in the permeability tensor under the external magnetic field biasing. For the considered thin YIG layer, the FP cavity resonance can enhance the FMO effect. To understand this property, we analytically determine the transmission and reflection for different CP waves propagating through the YIG layer. Consider an LCP wave backward (+ x direction) normally incident onto the YIG layer of thickness t and with the + x-biased magnetic field, as shown in Fig. 6, the electric and magnetic fields in regions 1, 2, and 3 can be expressed as: Region 1: 0 00 1ik xE e i incE, 0 0 00 1ik xY E i e incH, 0' '0 ik x y zE e E refE, 0' 0 '0 ik x z yY E e E refH. (A1) Region 2: 10 1b Lik x LE e i LCPE, 1 00 1b Lik x b L LE YY i e LCPH , 10 1b Rik x RE e i RCPE, 1 00 1b Rik x b R RE YY i e RCPH , (A2) 20 1f Lik x LE e i LCPE, 2 00 1f Lik x f L LE YY i e LCPΗ, 20 1f Rik x RE e i RCPE , 2 00 1f Rik x f R RE YY i e RCPH . (A3) Region 3: 0 " "0 ik x y zE e E tranE . (A4) where 0 0 0Y is the wave admittance in free space, 1( )b f L R r r rY Y and 1( )b f R L r r rY Y are the relative wave admittances for LCP and RCP waves normally forward (- x-direction) and backward (+ x-direction) propagating in the YIG material, 1 0( )b f L R r r rk k k and 1 0( )b f R L r r rk k k are the wave vectors of the LCP and RCP waves normally forward (- x-direction) and backward (+ x-direction) propagating in the YIG material, k0 is the wave vector in free space. Furthermore, according to the boundary conditions ( 20 n 1e H H , 0 n 1 2e Ε E ) between regions 1 and 2 ( x=0), and between regions 2 and 3 ( x=t), we can get eight equations to solve for the eight unknowns in Eq. (A1- A4), which can be expressed as below: At x=0: ' 0 1 1 2 2 ' 0 1 1 2 2 ' 0 0 1 0 1 0 2 0 2 0 ' 0 0 1 0 1 0 2 0 2 0( ) ( )y L R L R z L R L R b b f f z L L R R L L R R b b f f y L L R R L L R RE E E E E E iE E iE iE iE iE Y iE E iE YY iE YY iE YY iE YY Y E E E YY E YY E YY E YY (A5) At x=t: 0 0 0" 1 1 2 2 " 1 1 2 2 1 0 1 0 2 0 2 0 " 0 1 0f f b b L R L R f f b b L R L R f f b b L R L R b Lik t ik t ik t ik t ik t L R L R y ik t ik t ik t ik t ik t L R L R z ik t ik t ik t ik tb b f f L L R R L L R R ik t z ik tb L LE e E e E e E e E e iE e iE e iE e iE e E e iE YY e iE YY e iE YY e iE YY e Y E e E YY e 01 0 2 0 2 0 " 0f f b R L R ik t ik t ik tb f f R R L L R R ik t yE YY e E YY e E YY e Y E e (A6) By solving the Eq. (A5-A6), we can get the solutions: 2 0 122 2 1 2 0 222 2 22 ' 0 22 2 22 ' 0 22 2 " 02 ( 1) ( 1) ( 1) 0 0 2 ( 1) ( 1) ( 1) (1 )( 1) ( 1) ( 1) (1 )( 1) ( 1) ( 1) 4b L b L f R b L b L b L b Li k t b L Li k t b b L L R L f R Ri k t f f R R i k tb L yi k t b b L L i k tb L zi k t b b L L yE Y eE Y e Y E E E YE Y e Y E Y eE Y e Y E Y eE i Y e Y E YE 0 0( ) 22 2 ( ) " 0 22 2( 1) ( 1) 4 ( 1) ( 1)b L b L b L b Li k k tb L i k t b b L L i k k tb L zi k t b b L Le Y e Y E Y eE i Y e Y (A7) From Eq. (A7), we can observe that there exists the LCP wave along the + x-direction and the RCP wave along the - x-direction in the YIG layer, thus inducing coherent interference and the FP cavity resonance. Additionally, we note that the reflected and transmissive waves are always the RCP and LCP waves, respectively, and their coefficients can be represented as 22 22 2(1 )( 1) ( 1) ( 1)b L b Li k tb b L RLi k t b b L LY er Y e Y , 0( ) 22 24 ( 1) ( 1)b L b Li k k tb b L LLi k t b b L LY et Y e Y , (A8) where the superscript “+” indicates the + x-directional magnetic biasing, “ b” represents the backward normal incidence (+ x-direction), “ RL” symbolizes the CP state conversion from LCP to RCP, and of course “ LL” stands for the CP state conservation for LCP wave. Similarly, we also can get the solutions of the reflected and transmitted coefficients for the case of the RCP wave normally backward (+ x-direction) passing through the t-thick YIG layer with + x- directional magnetic biasing: 22 22 2(1 )( 1) ( 1) ( 1)b R b Ri k tb b R LRi k t b b R RY er Y e Y , 0( ) 22 24 ( 1) ( 1)b R b Ri k k tb b R RRi k t b b R RY et Y e Y . (A9) Comparing Eq. (A8) with Eq. (A9), it can be concluded that the difference in spin-dependent reflection and transmission is determined by the off-diagonal element κr. Furthermore, when the external magnetic field reverses the direction, the off-diagonal element in the permeability tensor will change from κr to –κr, thus the intrinsic admittance and wave vector of the LCP (RCP) wave in the case of + x-biased magnetic field case will be equal to those of the RCP (LCP) wave in the case of – x-biased magnetic field. Therefore, the corresponding reflection and transmissive coefficients can be expressed by , ,b b b b RL LR LR RL b b b b LL RR RR LLr r r r t t t t (A10) Meanwhile, due to the symmetry feature of YIG layer relative to the yz-plane, the reverse of the applied magnetic field is equivalent to the reverse of the incident direction of CP waves, thus getting b f RL LRr r , b f LR RLr r , b f LL RRt t , and b f RR LLt t . Thus, the spin-selective transmission and reflection also depend on the incident direction in addition to the magnetic field biasing. APPENDIX B: DEPENDENCE OF TRANSMISSION ON THE DIRECTIONS OF INCIDENCE AND BIASED MAGNETIC FIELD FOR METAISOLATOR The relationship between the CP states, wave propagation direction, and the biased magnetic field direction for the proposed metaisolator is numerically verified in Fig. 7. It can be noted in Fig. 7(a) that the cross-polarized transmissions for normally forward and backward incident LCP waves are nearly identical (i.e., f b RL RLt t ). This is also true for the RCP wave (i.e., f b LR LRt t ). A similar phenomenon can also be found in the case of − z-biased magnetic field shown in Fig. 7(b). This can be understood as follows. For the forward LCP wave passing through the metasurface with the + z-biased magnetic field, the cross-polarized transmission can be expressed as f L R f RL c Rt t t , where L R ctdenotes the conversion from LCP wave to RCP wave and f Rt is the forward RCP transmission for the YIG layer under the resonant coupling of the dielectric cylinder. Similarly, for the backward LCP wave, the cross-polarized transmission can be represented by b b L R RL L ct t t , where b Lt is the backward LCP transmission for the YIG layer under the resonant coupling of the dielectric cylinder. It should be noted that the forward RCP transmission f Rt is equal to the backward LCP transmission b Lt for the YIG layer in the presence of the Mie resonances of the cylinder, similar to property of the single YIG layer (discussed in Appendix A). Since the Mie resonances in the elliptical cylinder are spin-independent, the efficiency of its coupling to the YIG layer is unaffected by the CP state. Therefore, f b RL RLt t can be concluded. Meanwhile, their co-polarized transmission can also be expressed as (1 )f L R f LL c Lt t t and (1 )b b L R LL L ct t t , respectively. Since f b L Lt t owing to the nonreciprocal characteristic of YIG, we can obtain f b LL LLt t . Additionally, the magnetic-biased direction determines the spin-selective property of the metaisolator due to the electromagnetic characteristic of YIG. As demonstrated by the equal transmission of different CP states passing through the metaisolators with the opposite magnetic biasing, i.e., f f LR RLt t and f f RL LRt t . To summarize, these relationships can be described by f b f b RL RL LR LRt t t t and f b f b LR LR RL RLt t t t . APPENDIX C: ELECTROMAGNETIC MULTIPOLE EXPANSION The external field can induce the charge density ρ and current density J in the metasurface, which give rise to electromagnetic multipoles. Therefore, the resonance response of the metastructure can be understood based on the multipole decompositions. The multipole moments can be evaluated using the current density J(r) within the unit cell ( α, β, γ=x, y, z) as [55-57]: 31d rip J , (C1) 3 1 2d rc m r J , (C2) 2 3 1210r d rc T r J r J , (C3) 3 , ,1 2[ ( )]2 3eQ r J r J d ri r J, (C4) 3 3 ,1[( ) ] [( ) ]3mQ r d r r d rc r J r J , (C5) 2 2 3 , ,1[4 ( ) 5 ( ) 2 ( )]28TQ r J r r J r J r d rc r J r J , (C6) where p, m, T, Qe, Qm, and QT represent the electric dipole, magnetic dipole, toroidal dipole, electric quadrupole, magnetic quadrupole, and toroidal quadrupole, respectively, c is the light speed. The total scattered power Is of the metasurface can be expressed as [58] 4 4 5 2 2 3 3 4 6 6 62 2 2 , , 5 5 5 52 2 4 3 3 3 2 1 3 5 40s e mIc c c Q Q Oc c c c p m p T T (C7) We evaluated each term on the right-hand side of Eq. (C7) for the metaisolator, and the results are shown in Fig. 1(d). REFERENCES [1] K. Chen, Y. Feng, F. Monticone, J. Zhao, B. Zhu, T. Jiang, L. Zhang, Y. Kim, X. Ding, S. Zhang, A. Alù, and C. W. Qiu, A reconfigurable active Huygens' metalens, Adv. Mater. 29, 1606422 (2017). [2] Y. Liu, J. Sun, Y. Che, K. Qi, L. Li, and H. Yin, High numerical aperture microwave metalens, Opt. Lett. 45, 6262 (2020). [3] Y. Xu, J. Gu, Y. Gao, Q. Yang, W. Liu, Z. Yao, Q. Xu, J. Han, and W. Zhang, Broadband achromatic terahertz metalens constituted by Si-SiO2-Si hybrid meta-Atoms, Adv. Funct. Mater. 33, 2302821 (2023). [4] X. Zang, H. Ding, Y. Intaravanne, L. Chen, Y. Peng, J. Xie, Q. Ke, A. V. Balakin, A. P. Shkurinov, X. Chen, Y. Zhu, and S. Zhuang, A multi-foci metalens with polarization- rotated focal points, Laser Photonics Rev. 13, 1900182 (2019). [5] Y. Wang, Q. Chen, W. Yang, Z. Ji, L. Jin, X. Ma, Q. Song, A. Boltasseva, J. Han, V. M. Shalaev, and S. Xiao, High-efficiency broadband achromatic metalens for near-IR biological imaging window, Nat. Commun. 12, 5560 (2021). [6] G. Yoon, K. Kim, S. Kim, S. Han, H. Lee, and J. Rho, Printable nanocomposite metalens for high-contrast near-infrared imaging, ACS Nano 15, 698 (2021). [7] S. Wei, G. Cao, H. Lin, X. Yuan, M. Somekh, and B. Jia, A varifocal graphene metalens for broadband zoom imaging covering the entire visible region, ACS Nano 15, 4769 (2021). [8] J. Kim, J. Seong, W. Kim, G. Y. Lee, S. Kim, H. Kim, S. W. Moon, D. K. Oh, Y. Yang, J. Park, J. Jang, Y. Kim, M. Jeong, C. Park, H. Choi, G. Jeon, K. Lee, D. H. Yoon, N. Park, B. Lee, H. Lee, and J. Rho, Scalable manufacturing of high-index atomic layer-polymer hybrid metasurfaces for metaphotonics in the visible, Nat. Mater. 22, 474 (2023). [9] N. Meinzer, W. Barnes, and I. Hooper, Plasmonic meta-atoms and metasurfaces, Nature Photon. 8, 889 (2014). [10] H. Chen, A. J. Taylor, and N. Yu, A review of metasurfaces: physics and applications, Rep. Prog. Phys. 79, 076401 (2016). [11] Y. Huang, K. Kaj, C. Chen, Z. Yang, S. R. Ul Haque, Y. Zhang, X. Zhao, R. D. Averitt, and X. Zhang, Broadband terahertz silicon membrane metasurface absorber, ACS Photonics 9, 1150 (2022). [12] W. Yu, Y. Lu, X. Chen, H. Xu, J. Shao, X. Chen, Y. Sun, J. Hao, and N. Dai, Large-area, broadband, wide-angle plasmonic metasurface absorber for midwavelength infrared atmospheric transparency window, Adv. Optical Mater. 7, 1900841 (2019). [13] S. Jia, T. Fu, J. Peng, and S. Wang, Broadband and large-area optical chirality generated by an achiral metasurface under achiral excitation, Phys. Rev. A 108, 053504 (2023). [14] W. Yue, S. Gao, S. Lee, E. Kim, and D. Choi, Highly reflective subtractive color filters capitalizing on a silicon metasurface integrated with nanostructured aluminum mirrors, Laser Photonics Rev. 11, 1600285 (2017). [15] A. Tittl, A. Leitis, M. Liu, F. Yesilkoy, D. Y. Choi, D. N. Neshev, Y. S. Kivshar, and H. Altug, Imaging-based molecular barcoding with pixelated dielectric metasurfaces, Science 360, 1105 (2018). [16] M. R. Akram, G. Ding, K. Chen, Y. Feng, and W. Zhu, Ultrathin single layer metasurfaces with ultra-sideband operation for both transmission and reflection, Adv. Mater. 32, 1907308 (2020). [17] Y. Zheng, K. Chen, W. Yang, L. Wu, K. Qu, J. Zhao, T. Jiang, and Y. Feng, Kirigami reconfigurable gradient metasurface, Adv. Funct. Mater. 32, 2107699 (2022). [18] N. K. Grady, J. E. Heyes, D. R. Chowdhury, Y. Zeng, M. T. Reiten, A. K. Azad, A. J. Taylor, D. A. R. Dalvit, and H. T. Chen, Terahertz metamaterials for linear polarization conversion and anomalous refraction, Science 340, 1304 (2013). [19] W. Sun, Q. He, S. Sun, and L. Zhou, High-efficiency surface plasmon metacouplers: concept and microwave-regime realizations, Light Sci. Appl. 5, e16003 (2016). [20] Y. Chen, X. Zheng, X. Zhang, W. Pan, Z. Wang, S. Li, S. Dong, F. Liu, Q. He, L. Zhou, and S. Sun, Efficient meta-couplers squeezing propagating light into on-chip subwavelength devices in a controllable way, Nano Lett. 23, 3326 (2023). [21] M. Y. Shalaginov, S. An, Y. Zhang, F. Yang, P. Su, V. Liberman, J. B. Chou, C. M. Roberts, M. Kang, C. Rios, Q. Du, C. Fowler, A. Agarwal, K. A. Richardson, C. R. Baleine, H. Zhang, J. Hu, and T. Gu, Reconfigurable all-dielectric metalens with diffraction-limited performance, Nat. Commun. 12, 1225 (2021). [22] L. Li, Z. Liu, X. Ren, S. Wang, V. C. Su, M. K. Chen, C. H. Chu, H. Y. Kuo, B. Liu, W. Zang, G. Guo, L. Zhang, Z. Wang, S. Zhu, and D. P. Tsai, Metalens-array-based high- dimensional and multiphoton quantum source, Science 368, 1487 (2020). [23] T. Li, C. Chen, X. Xiao, J. Chen, S. Hu, and S. Zhu, Revolutionary meta-imaging: from superlens to metalens, Photonics Insights 2, R01 (2023). [24] P. Georgi, Q. Wei, B. Sain, C. Schlickriede, Y. Wang, L. Huang, and T. Zentgraf, Optical secret sharing with cascaded metasurface holography, Sci. Adv. 7, eabf9718 (2021). [25] X. Li, X. Zhang, R. Zhao, G. Geng, J. Li, L. Huang, and Y. Wang, Independent light field manipulation in diffraction orders of metasurface holography, Laser Photonics Rev. 16, 2100592 (2022). [26] H. Ren, X. Fang, J. Jang, J. Bürger, J. Rho, and S. A. Maier, Complex-amplitude metasurface-based orbital angular momentum holography in momentum space, Nat. Nanotechnol. 15, 948 (2020). [27] A. H. Dorrah, N. A. Rubin, A. Zaidi, M. Tamagnone, and F. Capasso, Metasurface optics for on-demand polarization transformations along the optical path, Nat. Photonics 15, 287 (2021). [28] A. Tripathi, A. Zalogina, J. Liao, M. Wurdack, E. Estrecho, J. Zhou, D. Jin, S. S. Kruk, and Y. Kivshar, Metasurface-controlled photonic Rashba effect for up-conversion photoluminescence, Nano Lett. 23, 2228 (2023). [29] W. Pan, Z. Wang, Y. Chen, X. Zheng, S. Li, X. Tian, Q. He, L.Zhou, and S. Sun, Efficiently controlling near-field wavefronts via designer metasurfaces, ACS Photonics 10, 2423 (2023). [30] Q. Song, M. Odeh, J. Z. Pérez, B. Kanté, and P. Genevet, Plasmonic topological metasurface by encircling an exceptional point, Science 373, 1133 (2021). [31] M. V. Berry, Quantal phase factors accompanying adiabatic changes, Proc. Roy. Soc. A Math. Phys. Eng. Sci. 392, 45 (1984). [32] S. Pancharatnam, Generalized theory of interference and its applications, Proc. Indian Acad. Sci. Sect. A 44, 398 (1956). [33] Y. Guo, M. Pu, F. Zhang, M. Xu, X. Li, X. Ma, and X. Luo, Classical and generalized geometric phase in electromagnetic metasurface, Photonics Insights 1, R03 (2022). [34] R. Jin, L. Tang, J. Li, J. Wang, Q. Wang, Y. Liu, and Z. G. Dong, Experimental demonstration of multidimensional and multifunctional metalenses based on photonic spin Hall effect, ACS Photonics 7, 512 (2020). [35] S. Liu, S. Qi, Y. Li, B. Wei, P. Li, and J. Zhao, Controllable oscillated spin Hall effect of Bessel beam realized by liquid crystal Pancharatnam-Berry phase elements, Light Sci. Appl. 11, 219 (2022). [36] Q. Zhou, M. Liu, W. Zhu, L. Chen, Y. Ren, H. J. Lezec, Y. Lu, A. Agrawal, and T. Xu, Generation of perfect vortex beams by dielectric geometric metasurface for visible light, Laser Photonics Rev. 15, 2100390 (2021). [37] H. Zhang, X. Sha, Q. Chen, J. Cheng, Z. Ji, Q. Song, S. Yu, and S. Xiao, All-dielectric metasurface-enabled multiple vortex emissions, Adv. Mater. 34, 2109255 (2022). [38] M. Tonouchi, Cutting-edge terahertz technology, Nat. Photonics 1, 97 (2007). [39] P. U. Jepsen, D. G. Cooke, and M. Koch, Terahertz spectroscopy and imaging modern techniques and applications, Laser Photonics. Rev. 5, 124 (2011). [40] X. Guo, Y. Ding, Y. Duan, and X. Ni, Nonreciprocal metasurface with space-time phase modulation, Light Sci. Appl. 8, 123 (2019). [41] Z. Yang, Y. Cheng, N. Wang, Y. Chen, and S. Wang, Nonreciprocal light propagation induced by a subwavelength spinning cylinder, Opt. Express 30, 27993 (2022). [42] H. Shi, Z. Yang, C. Zhang, Y. Cheng, Y. Chen, and S. Wang, Robust exceptional point of arbitrary order in coupled spinning cylinders, Opt. Express 29, 29720 (2021). [43] H. Shi, Y. Cheng, Z. Yang, Y. Chen, and S. Wang, Optical isolation induced by subwavelength spinning particle via spin-orbit interaction, Phys. Rev. B 103, 094105 (2021). [44] D. M. Pozar, Microwave engineering , John Wiley & Sons, USA 2011. [45] W. Yang, J. Qin, J. Long, W. Yan, Y. Yang, C. Li, E. Li, J. Hu, L. Deng, Q. Du, and L. Bi, A self-biased non-reciprocal magnetic metasurface for bidirectional phase modulation, Nat. Electron. 6, 225 (2023). [46] D. Zhao, F. Fan, J. Liu, Z. Tan, H. Wang, Q. Yang, Q. Wen, and S. Chang, Terahertz magneto-optical metadevice for active spin-selective beam steering and energy distribution with nonreciprocal isolation, Optica 10, 1295 (2023). [47] S. Xia, D. O. Ignatyeva, Q. Liu, J. Qin, T. Kang, W. Yang, Y. Chen, H. Duan, L. Deng, D. Long, M. Veis, V. I. Belotelov, and L. Bi, Enhancement of the Faraday effect and magneto- optical figure of merit in all-dielectric metasurfaces, ACS Photonics 9, 1240 (2022). [48] D. Zhao, Z. Tan, H. Zhao, F. Fan, and S. Chang, Active terahertz beam deflection and nonreciprocal spin chirality selection based on magneto-optical P-B metasurface with stacked-graphene layers, Opt. Lett. 47, 818 (2022). [49] S. Wang, P. C. Wu, V. C. Su, Y. C. Lai, M. K. Chen, H. Y. Kuo, B. H. Chen, Y. H. Chen, T. T. Huang, J. H. Wang, R. M. Lin, C. H. Kuan, T. Li, Z. L. Wang, S. Zhu, and D. P. Tsai, A broadband achromatic metalens in the visible, Nat. Nanotechnol. 13, 227 (2018). [50] R. J. Lin, V. C. Su, S. Wang, M. K. Chen, T. L. Chung, Y. H. Chen, H. Y. Kuo, J. W. Chen, J. Chen, Y. T. Huang, J. H. Wang, C. H. Chu, P. C. Wu, T. Li, Z. Wang, S. Zhu, and D. P. Tsai, Achromatic metalens array for full-colour light-field imaging, Nat. Nanotechnol. 14, 227 (2019). [51] Y. Luo, C. H. Chu, S. Vyas, H. Y. Kuo, Y. H. Chia, M. K. Chen, X. Shi, T. Tanaka, H. Misawa, Y. Y. Huang, and D. P. Tsai, Varifocal metalens for optical sectioning fluorescence microscopy, Nano Lett. 21, 5133 (2021). [52] C. Chen, W. Song, J. W. Chen, J. H. Wang, Y. H. Chen, B. Xu, M. K. Chen, H. Li, B. Fang, J. Chen, H. Y. Kuo, S. Wang, D. P. Tsai, S. Zhu, and T. Li, Spectral tomographic imaging with aplanatic metalens, Light Sci. Appl. 8, 99 (2019). [53] G. Cai, Y. Li, Y. Zhang, X. Jiang, Y. Chen, G. Qu, X. Zhang, S. Xiao, J. Han, S. Yu, Y. Kivshar, and Q. Song, Compact angle-resolved metasurface spectrometer, Nat. Mater. 23, 71 (2024). [54] A. Arbabi, Y. Horie, A. J. Ball, M. Bagheri, and Andrei Faraon, Subwavelength-thick lenses with high numerical apertures and large efficiency based on high-contrast transmitarrays, Nat. Commun. 6, 7069 (2015). [55] E. E. Radescu, and G. Vaman, Exact calculation of the angular momentum loss, recoil force, and radiation intensity for an arbitrary source in terms of electric, magnetic, and toroid Multipoles, Phys. Rev. E 65, 046609 (2002). [56] V. Savinov, V. A. Fedotov, and N. I. Zheludev, Toroidal dipolar excitation and macroscopic electromagnetic properties of metamaterials, Phys. Rev. B 89, 205112 (2014). [57] G. N. Afanasiev, and Y. P. Stepanovsky, The electromagnetic field of elementary time- dependent toroidal sources, J. Phys. A: Gen. Phys. 28, 4565 (1995). [58] P. C. Wu, C. Y. Liao, V. Savinov, T. L. Chung, W. T. Chen, Y. W. Huang, P. R. Wu, Y. H. Chen, A. Q. Liu, N. I. Zheludev, and D. P. Tsai, Optical Anapole Metamaterial, ACS Nano 12, 1920 (2018). FIG. 1. The PB metasurface isolator and its nonreciprocal properties. (a) The schematics of the metaisolator and the meta-atom. The meta-atom is composed of an elliptical dielectric resonator and a YIG layer with geometric parameters l=14 mm, h=20 mm, a=12 mm, b=6 mm, and t=4.5 mm. The external biased magnetic field B0 is along + z direction. (b) The simulated (red symbol lines) and analytical (blue symbol lines) nonreciprocal transmission spectra of the YIG layer with B0=0.05 T pointing in the + z direction. 2| |f LLtand 2| |f RRt are the co-polarized transmission for the forward (- z-direction) normally incident LCP and RCP waves, respectively. (c) The transmission spectra of the metaisolator with B0=0.05 T pointing in the + z direction. 2| |f LLtand 2| |f RLt are the co- and cross-polarized transmission for the forward (- z-direction) normally incident LCP waves, and 2| |b RRtand 2| |b LRtare those for the backward (+ z-direction) normally incident RCP waves. (d) The normalized multipole scattering power of the dielectric elliptical cylinder under the excitation of the forward-incident LCP wave. p, m, T, Q e, and Q m are the electric dipole, magnetic dipole, toroidal dipole, electric quadrupole, and magnetic quadrupole, respectively. The inner image shows the corresponding electric and magnetic fields in the meta-atom at the resonant frequencies 5.5 GHz and 6.5 GHz. FIG. 2. The nonreciprocal characteristics of the PB-phase-based metaisolator. (a) The isolation ratio (i.e. 2 2| | | |f b RL LRt t ) of the metaisolator as a function of the frequency and the height of the dielectric elliptical cylinder. (b) The isolation ratio of the metasurface with the ± z-biased magnetic field of 0.05 T or without the magnetic field for the normally incident CP light (i.e., 2 2| | | |f b RL LRt t ). (c) The isolation ratio (i.e., 2 2| | | |f b RL LRt t ) as a function of the external + z- biased magnetic field strength B0 and the frequency of normally incident CP waves. (d) The isolation ratio (i.e., 2 2| | | |f b RL LRt t ) as a function of the incident angle and the frequency of the CP waves. The external magnetic field is in the + z-direction with a magnitude of 0.05 T. FIG.3. The cross-polarized transmission amplitude and phase shift for the metaisolator composed of elliptical dielectric cylinders with different orientation angles. The incident wave is LCP working at 5.5 GHz, and it normally incidents on the metasurface. FIG. 4. Nonreciprocal metadeflector for beam steering. (a) The schematic for the nonreciprocal beam steering by the PB metadeflector with the magnetic-field biasing pointing in + z-direction. (b) The simulated normalized electric-field profiles for the supercells with different meta-atoms. The incident wave is LCP for the upper-row panels and RCP for the bottom-row panels, and their propagation directions are denoted by the white arrows. The frequency is at 5 GHz. The metadeflectors with the supercells consisting of 4, 6, 8, and 12 cells can achieve the deflection angle of 74.64°, 40°, 28.82°, and 18.75°. FIG. 5. The nonreciprocal PB metalens for beam focusing. (a) The schematic for the nonreciprocal focusing of the proposed PB metalens. The incident LCP beam is converted to the RCP beam and focused at one point, while the RCP radiation from the focal point cannot pass through the metalens. (b) The normalized electric-field distribution in the yz-plane when the forward incident LCP beam passes through the metalenses with different focal lengths 1.5 λ, 2λ, and 3λ (the upper panels), and when the RCP wave radiated from the focal points backward propagates into the metalenses (the bottom panels). The normalized intensity distribution at the focal planes z=−1.5λ (c), z=−2λ (d), and z=−3λ (e), respectively, corresponding to the three cases in the upper panels of (b). FIG. 6. The schematic for the thin YIG layer with + x-biased magnetic field under the normal incidence of a LCP plane wave. FIG. 7. The comparison of cross-polarized transmission of different CP waves with different propagation directions through the metaisolator with magnetic biasing along +z (a) and -z (b) directions shown in the insets. The magnitude of the magnetic field is 0.05 T. The subscript “LR” (“RL”) represents the CP state conversion from RCP to LCP (LCP to RCP). The superscripts “+” and “−” indicate the + z- and −z-biased magnetic field; “ f” and “b” stand for the forward (− z-direction) and backward (+ z-direction) incidence. | 2024-01-19 | Optical metasurfaces have been widely used for manipulating electromagnetic
waves due to their low intrinsic loss and easy fabrication. The metasurfaces
employing the Pancharatnam-Berry (PB) geometric phase, called PB metasurfaces,
have been extensively applied to realize spin-dependent functionalities, such
as beam steering, focusing, holography, etc. The demand for PB metasurfaces in
complex environments has brought about one challenging problem, i.e., the
interference of multiple wave channels that limits the performance of PB
metasurfaces. A promising solution is developing nonreciprocal PB metasurfaces
that can isolate undesired wave channels and exhibit unidirectional
functionalities. Here, we propose a mechanism to realize nonreciprocal PB
metasurfaces of subwavelength thickness by using the magneto-optical effect of
YIG material in synergy with the PB geometric phase of spatially rotating
meta-atoms. Using full-wave numerical simulations, we show that the metasurface
composed of dielectric cylinders and a thin YIG layer can achieve nearly 92%
and 81% isolation of circularly polarized lights at 5.5 GHz and 6.5 GHz,
respectively, attributed to the enhancement of the magneto-optical effect by
the resonant Mie modes and Fabry-P\'erot cavity mode. In addition, the
metasurface can enable efficient unidirectional wavefront manipulations of
circularly polarized lights, including nonreciprocal beam steering and
nonreciprocal beam focusing. The proposed metasurface can find highly useful
applications in optical communications, optical sensing, and quantum
information processing. | Nonreciprocal Pancharatnam-Berry Metasurface for Unidirectional Wavefront Manipulation | 2401.10772v2 |
arXiv:2306.04390v1 [physics.optics] 7 Jun 2023Gain assisted controllable fast light generation in cavity magnomechanics Sanket Das,1Subhadeep Chakraborty,2and Tarak N. Dey1,∗ 1Department of Physics, Indian Institute of Technology Guwa hati, Guwahati-781039, Assam, India. 2Centre for Quantum Engineering Research and Education, TCG Centres for Research and Education in Science and Techno logy, Sector V, Salt Lake, Kolkata 70091, India (Dated: June 8, 2023) We study the controllable output field generation from a cavi ty magnomechanical resonator system that consists of two coupled microwave resonators. T he first cavity interacts with a ferromagnetic yttrium iron garnet (YIG) sphere providing t he magnon-photon coupling. Under passive cavities configuration, the system displays high ab sorption, prohibiting output transmission even though the dispersive response is anamolous. We replac e the second passive cavity with an active one to overcome high absorption, producing an effecti ve gain in the system. We show that the deformation of the YIG sphere retains the anomalous disp ersion. Further, tuning the exchange interaction strength between the two resonators leads to th e system’s effective gain and dispersive response. As a result, the advancement associated with the a mplification of the probe pulse can be controlled in the close vicinity of the magnomechanical r esonance. Furthermore, we find the existence of an upper bound for the intensity amplification a nd the advancement of the probe pulse that comes from the stability condition. These finding s may find potential applications for controlling light propagation in cavity magnomechanics. I. INTRODUCTION Cavity magnonics [1, 2], has become an actively pur- sued field of research due to its potential application in quantum information processing [3, 4]. The key con- stituent to such systems is a ferrimagnetic insulator with high spin density and low damping rate. It also supports quantized magnetization modes, namely, the magnons [5, 6]. With strongly coupled magnon-photon modes, cavity magnonics is an excellent platform for studying all the strong-coupling cavity QED effects [7]. Besides originating from the shape deformation of the YIG, the magnon can also couple to a vibrational or phonon mode [5]. This combined setup of magnon-photon-phonon modes, namely the cavity magnomechanics, has already demonstrated magnomechanically induced transparency [5], magnon-induced dynamical backaction [8], magnon- photon-phonon entanglement [9, 10], squeezed state gen- eration [11], magnomechanical storage and retreival of a quantum state [12]. Recently, PT-symmetry drew extensive attention to elucidate the dynamics of a coupled system character- ized by gain and loss [13, 14]. Here, Pstands for the parity operation, that results in an interchange between the twoconstituentmodes ofthe system. Thetime rever- sal operator Ttakesito−i.PT-symmetry demands the Hamiltonian is commutative with the joint PToperators i.e.,[H,PT] = 0. This system possesses a spectrum of entirely real and imaginary eigenvalues that retain dis- tinguishable characteristics [15]. The point separating these two eigenvalues is the exceptional point (EP) [16] where the two eigenvalues coalesce, and the system de- generates. Anaturaltestbed for PT-symmetricHamilto- ∗tarak.dey@iitg.ac.innian is optical as well as quantum optical systems [17–19] which alreadyled to the demonstrationof someofthe ex- oticphenomena,likenonreciprocallightpropagation[20], unidirectional invisibility [21, 22], optical sensing and light stopping [23]. Very recently, a tremendous effort has been initiated to explore non-Hermitian physics in magnon assisted hybrid quantum systems. The second- order exceptional point is detected in a two-mode cavity- magnoic system, where the gain of the cavity mode is ac- complished by using the idea of coherent perfect absorp- tion [24]. The concept of Anti- PTsymmetry has been realized experimentally [25], where the adiabatic elim- ination of the cavity field produces dissipative coupling between two magnon modes. Beyond the unique spectral responses, these non-Hermitian systems can manipulate the output microwavefield transmission[26, 27]. Theun- derlying mechanism behind such an application is mag- netically induced transparency[5, 28], where the strong magnon-photoncouplingproducesanarrowspectral hole inside the probe absorption spectrum. Further studies in this direction establish the importance of the weak magnon-phonon coupling to create double transmission windows separated by an absorption peak. Moreover, manipulating the absorption spectrum is also possible by varying the amplitude and phase of the applied magnetic field [29]. It is well established over the past decade that op- tomechanically induced transparency (OMIT) [30–32] is an essential tool for investigating slow light [33] and light storage [34, 35] in cavity. In addition, incorporating PT- symmetry in optomechanical systems, provides a better controllability of light transmission [36, 37] and produces subluminal to superluminal light conversion. Nonethe- less, their proposals may find experimental challenges as the gain of the auxiliary cavity can lead the whole system to instability [38]. An eminent advantage of the2 magnomechanical system over the optomechanical sys- tem is that it offers strong hybridization between the magnon-photon mode. The magnomechanical systems offer better tunability as an external magnetic field can vary the magnon frequency. Exploiting these advan- tages, aPT-symmetry-like magnomechanicalsystem can be constructed by resonantly driving the YIG sphere to an active magnon mode [39]. The controllable sideband generation with tunable group delay can be feasible by changing the power of the control field. This paper investigates a controllable advancement and transmission of the microwave field from a coupled cavity magnomechanical system. Optical coupling be- tween a passive cavity resonator containing YIG sphere and a gain-assisted auxiliary cavity can form a coupled cavity resonator. An external drive has been used to deform the YIG sphere’s shape, resulting in the magnon- phonon interaction in the passive cavity. We show how the gain of the auxiliarycavityhelps to overcomeabsorp- tive behaviourin our hybrid system. As a result, the out- put microwave field amplifies at the resonance condition. Moreover, the weak magnon-phonon interaction exhibits anomalous dispersion accompanied by a gain spectrum, demonstrating superluminal light. We also examine how the slope of the dispersion curve can be controlled by tuning the photon hopping interaction strength between the two cavities. The paper is organized as follows. In Section II, a the- orical model for the coumpound cavity magnomechanical system with PT-symmetric resonator is described. The Heisenserg equations of motion to govern the expecta- tion values of operators of every system are derived in this Section. In Section IIIA, we analyse the stability criteria of the model system and examine the effect of the auxiliary cavity gain on the absorptive and disper- sive response of the system in Section IIIB. Section IIIC discusses the output probe field transmission. Further, the group velocity of the optical probe pulse has been studied analytically and verified numerically in Section IIID. Finally, we draw our conclusions in Section IV. II. THEORETICAL MODEL Recently, there has been a growing interest in real- izing a gain in different components of cavity magnon- ics systems [24, 39]. In this work, we investigate the effect of medium gain on the probe response and its transmission. The system under consideration is a hy- brid cavity magnomechanical system that consists of two coupled microwave cavity resonators. One of the res- onators is passive and contains a YIG sphere inside it. We refer to this resonator as a cavity magnomechanical (CMM) resonator. Applying a uniform bias magnetic field to the YIG sphere excites the magnon mode. The magnon mode, in turn, couples with the cavity field by the magnetic-dipole interaction. Nonetheless, the exter- nalbiasmagneticfieldresultsinshapedeformationofthea1a2J εc,ωl εp,ωp κ1 κ2B0 FIG. 1. The schematic diagram of a hybrid cavity mag- nomechanical system. The system consists of two coupled microwave cavities. One of them is passive, and another one is active. The passive cavity contains a ferromagnetic YIG sphere inside it. The applied bias magnetic field pro- duces the magnetostrictive interaction between magnon and phonon. The coupling rates between the magnon-photon and magnon-phonon are gmaandgmb, respectively. Strong con- trol field of frequency ωland a weak probe field of frequency ωpare applied to the passive cavity. YIG sphere, leading to the magnon-phonon interaction. The second resonator (degenerate with the first one) is coupled to the first resonator via optical tunnelling at a rateJ. Two input fields drive the first resonator. The amplitude of the control, εl, and probe fields, εp, are given byεi=/radicalbig Pi/ℏωi,(i∈l,p) withPiandωibeing the power and frequency of the respective input fields. The Hamiltonian of the combined system can be written as H=ℏωca† 1a1+ℏωca† 2a2+ℏωmm†m+ℏωbb†b +ℏJ(a† 1a2+a† 2a1)+ℏgma(a† 1m+a1m†) +ℏgmbm†m(b†+b)+iℏ/radicalbig 2ηaκ1εl(a† 1e−iωlt−a1eiωlt) +iℏ/radicalbig 2ηaκ1εp(a† 1e−iωpt−a1eiωpt), (1) wherethe firstfourtermsoftheHamiltoniandescribethe free energy associated with each system’s constituents. The constituents of our model are characterized by their respective resonance frequencies: ωcfor the cavity mode, ωmfor the magnon mode, ωbfor the phonon mode. The annihilationoperatorsforthecavity,magnonandphonon modes are represented by ai, (i= 1,2),mandb, respec- tively. The fifth term signifies the photon exchange in- teraction between the two cavities with strength, J. The sixth term of the Hamiltonian corresponds to the inter- action between the magnon and photon modes, charac- terized by a coupling rate gma. The interaction between the magnon and phonon modes is described by the sev- enth term of the Hamiltonian and the coupling rate be- tween magnon and phonon mode is gmb. Finally, the last two terms arise due to the interaction between the cav- ity field and two input fields. The cavity, magnon and phonon decay rates are characterized by κ1,κmandκb, respectively. The coupling between the CMM resonator and the output port is given by ηa=κc1/2κ1, where3 κc1is the cavity external decay rate. In particular, we will consider the CMM resonator to be working in the critical-coupling regime where ηais 1/2. At this point, it is convenient to move to a frame rotating at ωl. Fol- lowing the transformation Hrot=RHR†+iℏ(∂R/∂t)R† withR=eiωl(a† 1a1+a† 2a2+m†m)t, the Hamiltonian in Eq. (1) can be rewritten as Hrot=/planckover2pi1∆a(a† 1a1+a† 2a2)+/planckover2pi1∆mm†m+/planckover2pi1ωbb†b +/planckover2pi1J(a† 1a2+a† 2a1)+/planckover2pi1gma(a† 1m+a1m†) +ℏgmbm†m(b†+b)+i/planckover2pi1/radicalbig 2ηaκ1εl(a† 1−a1) +i/planckover2pi1/radicalbig 2ηaκ1εp(a† 1e−iδt−h.c), (2) where ∆ a=ωc−ωl(∆m=ωm−ωl) andδ=ωp−ωlare, respectively, the cavity (magnon) and probe detuning. The mean response of the system can be obtained by the Heisenberg-Langevinequationas /angbracketleft˙O/angbracketright=i//planckover2pi1/angbracketleft[Hrot,O]/angbracketright+ /angbracketleftN/angbracketright. Further, we consider the quantum fluctuations ( N) as white noise. Then starting form Eq. 2, the equations of motion of the system can be expressed as /angbracketleft˙a1/angbracketright= (−i∆a−κ1)/angbracketlefta1/angbracketright−igma/angbracketleftm/angbracketright−iJ/angbracketlefta2/angbracketright +/radicalbig 2ηaκ1εl+/radicalbig 2ηaκ1εpe−iδt, /angbracketleft˙m/angbracketright= (−i∆m−κm)/angbracketleftm/angbracketright−igma/angbracketlefta1/angbracketright −igmb/angbracketleftm/angbracketright(/angbracketleftb†/angbracketright+/angbracketleftb/angbracketright), /angbracketleft˙b/angbracketright= (−iωb−κb)/angbracketleftb/angbracketright−igmb/angbracketleftm†/angbracketright/angbracketleftm/angbracketright, /angbracketleft˙a2/angbracketright= (−i∆a+κ2)/angbracketlefta2/angbracketright−iJ/angbracketlefta1/angbracketright, (3) whereκ2andκbrespectively denote the gain of the sec- ond resonator and phonon damping rates. We note that κ2>0 corresponds to a coupled passive-active CMM resonators system and κ2<0 describes a passive-passive coupled CMM resonators system. Assuming the control field amplitude εlto be larger than the probe field εp, each operator expectation values /angbracketleftO(t)/angbracketrightcan be decom- posedintoitssteady-statevalues Osandasmallfluctuat- ing termδO(t). The steady-state values of each operator are a1s=(−i∆a+κ2)(−igmams+√2ηaκ1εl) (i∆a+κ1)(−i∆a+κ2)−J2,(4a) ms=−igmaa1s i∆′m+κm, (4b) bs=−igmb|ms|2 iωb+κb, (4c) a2s=iJa1s (−i∆a+κ2). (4d) While the fluctuating parts of Eq. 3can be expressed as δ˙a1=−(i∆a+κ1)δa1−iJδa2−igmaδm +/radicalbig 2ηaκ1εpe−iδt, δ˙m=−(i∆′ m+κm)δm−igmaδa1−iGδb−iGδb†, δ˙b=−(iωb+κb)δb−iGδm†−iG∗δm, δ˙a2=−(i∆a−κ2)δa2−iJδa1, (5)where∆′ m= ∆m+gmb(bs+b∗ s)istheeffectivemagnonde- tuning and G=gmbmsis the enhanced magnon-phonon coupling strength. For simplicity, we express these fluc- tuation equations as id dt|ψ/angbracketright=Heff|ψ/angbracketright+F, (6) where the fluctuation vector |ψ/angbracketright= (δa1,δa† 1,δa2,δa† 2,δb,δb†,δm,δm†)T, input field F= (√2ηaκ1εpe−iδt,√2ηaκ1εpeiδt,0,0,0,0,0,0)T. Next, we adopt the following ansatz to solve Eq. 5: δa1(t) =A1+e−iδt+A1−eiδt, δm(t) =M+e−iδt+M−eiδt δb(t) =B+e−iδt+B−eiδt, δa2(t) =A2+e−iδt+A2−eiδt. (7) HereAi+andAi−correspond to the ithcavity generated probe field amplitude and the four-wavemixing field am- plitude, respectively. By considering h1=−i∆a+iδ− κ1, h2=−i∆a−iδ−κ1, h3=−i∆a+iδ+κ2, h4= −i∆a−iδ+κ2, h5=−iωb+iδ−κb, h6=−iωb−iδ− κb, h7=−i∆′ m+iδ−κm, h8=−i∆′ m−iδ−κm, we obtainA1+which corresponds to the output probe field amplitude from the CMM resonator as A1+(δ) =C(δ) D(δ), (8) where C(δ) =−/radicalbig 2ηaκaεph3(h5h7h∗ 6(J2h∗ 8+h∗ 4(g2 ma+h∗ 2h∗ 8)) +|G|2(h5−h∗ 6)(J2(h7−h∗ 8)−h∗ 4(gma2+h∗ 2(h∗ 8−h∗ 7)))), D(δ) =h5h∗ 6(g2 mah3+h7(h1h3+J2)) (J2h∗ 8+h∗ 4(g2 ma+h∗ 2h∗ 8))+|G|2(h5−h∗ 6) (J2(g2 mah3+(h1h3+J2)(h7−h∗ 8))−h∗ 4 ((h1h3+J2)(g2 ma−h∗ 2(h7−h∗ 8))−h∗ 2h3g2 ma)).(9) The output field from the CMM resonator is obtained by the cavity input-output relation εout=/radicalbig 2ηaκ1/angbracketlefta1/angbracketright−εl−εpe−iδt.(10) By substituting Eq. 7into Eq. 10, we obtain the normal- izedoutputprobefieldintensityfromtheCMMresonator as T=|tp|2=/vextendsingle/vextendsingle/vextendsingle/vextendsingle√2ηaκ1A1+ εp−1/vextendsingle/vextendsingle/vextendsingle/vextendsingle2 . (11) In order to numerically simulate the transmitted output probefield spectrum, weusethe followingexperimentally realizable set of parameter values [5, 40]. The degenerate microwave cavities of frequency ωc/2π= 7.86 GHz. The decay rate of the first cavity is κ1/π= 3.35 MHz. The spin density ρ= 4.22×1027m−3and the diameter of the4 YIG sphere D= 25µm. It results in 3 ×1016number of spins (Nm) present in the YIG sphere. The phonon mode has frequency ωb/2π= 11.42 MHz with decay rateκb/π= 300 Hz, and the magnon-phonon coupling strengthgmb/2πis 1 Hz. The Kittel mode frequency of the YIG sphere is ωm=γeB0,i, with gyromagnetic ratio,γe/2π= 28 GHz/T and B0,iis the input bias magnetic field amplitude. The magnon decay rate is κm= 3.52 MHz. Magnon-photon coupling strength gma=γeBvac√5Nm/2 can be controlled by changingthe vacuum magnetic field amplitude as Bvac=/radicalbig 2π/planckover2pi1ωc/V. III. RESULTS A. Stability Analysis Initially we consider the two coupled cavities which are operating under a balanced gain-loss condition. The Hamiltonian describing such coupled resonator system (gma=gmb= 0) can be written as Hcav=/planckover2pi1(∆a−iκ1)δa† 1δa1+/planckover2pi1(∆a+iκ1)δa† 2δa2 +/planckover2pi1J(δa† 1δa2+δa† 2δa1). (12) The eigenvalues of Hcavareλ±= ∆a±/radicalbig J2−κ2 1. Note that the above Hamiltonian remains invariant under the simultaneous parity P:a1↔a2and time-reversal operation T:i→ −ioperations, and, its eigenvalues are entirely real and complex for J > κ 1andJ < κ 2. The pointJ=κ1, which marks this transition from PT symmetric to the PTbreaking phase, is known as the exceptional point (EP). One must understand the com- petitive behaviour between the inter-cavityfield coupling and the loss/gain rates to get insight into this transition. ForJ > κ 1the intracavity field amplitudes can be coherently exchanged and thus give rise to a coherent oscillation between the field amplitudes. However, for J < κ 1the intracavity field can not be transferred to the other one, resulting in a strong field localization or in other words exponential growth. A quick look at Eq. 4(a) also suggests such gain-induced dynamic instability ina1atJ=κ1for ∆a= 0. This situation becomes more complicated in the presence of magnon-photon coupling. Now, the combined system ( gma,gmb/negationslash= 0) ceases to become PTsymmetric. However, the effect of an additional gain cavity ( κ2>0) can be understood by analyzing the stability diagram of the whole system. In the following, we derive the stability condition by invoking the Routh-Hurwitz criterion which requires all the eigenvalues of Heffhave negative real parts. The magenta region of Fig. 2suggests that when gma is small the instability threshold remains close to the J=κ1(the conventional EP for a binary PTsymmetric system). However, with increasing gmathe system reaches instability at a larger exchange interaction J. Such a restriction over the choice of the photon exchange FIG. 2. The stable and unstable regions are determined as a function of normalized evanescent coupling strength ( J/κ1) and the cavity-magnon coupling strength ( gma) when the loss of the CMM-resonator is perfectly balanced by the gain of the auxiliary cavity ( κ1=κ2). We consider the control field intensity to be 10 mW. The other parameters are ωc= 2π× 7.86 MHz, ωb= 2π×11.42 MHz, ∆ a= ∆′ m=ωb= 2π×11.42 MHz,κ1=κ2=π×3.35 MHz, κm=π×1.12 MHz, κb= π×300 Hz, and gmb= 2πHz. rateparameter Jwill be followedthroughoutthis paper. B. Absorption and dispersion spectrum Themagnomechanicalsystemunder considerationcor- responds to the level diagram of Fig. 3. Application of a probe field excites the passive cavity mode and allows the transition between |1/angbracketrightand|2/angbracketright. The exchange in- teraction,J, couples two degenerate excited states |2/angbracketright and|5/angbracketright. The presence of the strong control field dis- tributes the population between the two states |2/angbracketrightand |3/angbracketright. The magnon-phonon coupling, gmb, couples both the metastable ground states |3/angbracketrightand|4/angbracketright. Here, we con- sider both the microwave cavities to be passive ( κ2<0) under a weak magnon-photon coupling strength, gma. In this situation, the magnon-photon hybridization is in- significant. The absorptive and dispersive response can be quantified by the real and imaginary components of (tp+1) that will be presented as αandβ, respectively. In Fig.4(a), we present the absorptive response of the system as a function of normalized probe detuning. The black-solid curve depicts a broad absorption spectrum of the probe field when the exchange interaction is much weaker than the cavity decay rate. One can explain it by considering the level diagram of Fig. 3, where the ini- tial population stays in the ground state |1/angbracketright. Applying a probefieldtransfersthepopulationfromthegroundstate to the excited state, |2/angbracketright. In addition, the weak magnon-5 εp,ωpεl,ωl gmb /C2 |1/angbracketright|2/angbracketright |3/angbracketright |4/angbracketright|5/angbracketright |n1,n2,m,b/angbracketright|n1+1,n2,m,b/angbracketright|n1,n2+1,m,b/angbracketright |n1,n2,m+1,b/angbracketright |n1,n2,m,b+1/angbracketright FIG. 3. Level diagram of the model system. |ni/angbracketright,|m/angbracketrightand|b/angbracketright represents the photon number state of ithcavity, magnon mode and phonon mode, respectively. The application of strong control field to the CMM resonator couples |n1+ 1,n2,m,b/angbracketrightand|n1,n2,m+ 1,b/angbracketright, whereas, the pres- ence of a weak probe field increases the photon number of CMM resonator by unity. gmbcouples |n1,n2,m+ 1,b/angbracketrightand|n1,n2,m,b+ 1/angbracketright. The hopping interaction be- tween the two cavities directly couples |n1+ 1,n2,m,b/angbracketright ↔ |n1,n2+1,m,b/angbracketright. photon coupling (with respect to κ1) restricts a signifi- cant transition from |2/angbracketrightto|3/angbracketright. As a result, it allows the transfer of a fraction of the excited state’s population by invoking the exchange interaction J. The increase in the exchange interaction strength causes a gradual decrease in|2/angbracketright’s population. It reduces the absorption coefficient around the resonance condition except for δ=ωb. This phenomenon is shown by the red-dashed and the blue dotted-dashed curve of Fig. 4(a). We observe a nar- row absorption peak inside the broad absorption peak forJ= 0.4κ1. The sharp absorption peak, exactly at δ=ωb, occurs due to the magnomechanical resonance. Furtherincreasingtheexchangeinteractionvirtuallycuts off the population distribution from |2/angbracketrightto|3/angbracketright. As a result, the effect of magnon-phonon resonance also de- creases, and the absorption peak at δ=ωbeventually diminishes. In Fig. 4(b), we present the dispersion spec- trum as a function of normalized detuning δ/ωb. For the time being, we neglect the effect of magnomechanical couplingandobservetheoccurrenceofanomalousdisper- sion around δ=ωbforJ= 0.4κ1. Further, increasing the exchange interaction strength more significant than the cavity decay rate can alter the dispersive response from anomalous to normal, as shown by the red-dashed and blue-dot-dashed curves. In the inset of Fig. 4(b), we plot the slope of the temporal dispersion dβ/dδat the extreme vicinity of the magnon-phonon resonance con- dition. The positive values of the slope of the temporal dispersion signify anomalous dispersion due to the mag- nomechanical coupling. However, the steepness of the dispersion curve can be reduced by increasing the ex- change interaction strength, as shown by the red-dashed and blue-dot-dashed curves. Note that this dispersion curve is accompanied by absorption. Output transmis- sionoftheprobefieldisprohibitedinthepresenceofhuge0.5 0.75 1 1.25 1.5 δ/ωb00.20.40.60.81αJ = 0.4 κ1 J = 1.1 κ1 J = 1.3 κ1(a) 0.5 0.75 1 1.25 1.5 δ/ωb-0.4-0.200.20.4βJ = 0.4 κ1 J = 1.1 κ1 J = 1.3 κ1 0.999 1.0008δ/ωb 0123 dβ/dδ(b) x10-6 FIG. 4. (a) Absorption and (b) dispersion spectrum of the model system. The slope of the dispersion curve is shown in the inset. Here we consider both the microwave cavities are passive, with identical decay rates ( κ1=−κ2). The magnon- photon couplong strength, gmais taken as 2 MHz. All the other parameters are mentioned earlier. absorption. Therefore, reducing absorption or introduc- ing the gain to the system is mandatory for observingthe group velocity phenomena. To achieve reasonable transmission at the output, we replace the auxiliary passive cavity with an active one where the second cavity’s gain ( κ2>0) completely bal- ances the first cavity’s loss. In this scenario, the sta- bility criterion for the hybrid system allows us to con- sider the exchange interaction strength Jto be greater than 1.053κ1forgma= 2 MHz. We present the ab- sorptive response of the model system in Fig. 5(a). The black solid curve of Fig. 5(a) illustrates the occurrence of a double absorption peak spectrally separated by a broad gain regime. The graphical nature is determined by the roots of D(δ), which are, in general, complex. The real parts of the roots determine the spectral peak position, and the imaginary parts correspond to their widths. To illustrate this, we consider J= 1.30κ1with all other parameters remaining the same as earlier. The real parts of the root of D(δ) present two distinct nor- mal mode positions at δ/ωbvalues 0.88 and 1.12. The other two normal modes are spectrally located at the same position δ/ωb= 1. The interference between these6 0.8 0.9 1 1.1 1.2 δ/ωb-40-2002040α J = 1.30 κ1 J = 1.10 κ1 J = 1.07 κ1(a) 0.8 1 1.2 δ/ωb-40-2002040Im (tp+1) J = 1.30 κ1 J = 1.10 κ1 J = 1.07 κ10.96 11.04δ/ωm -25025Im (tp+1) (b) FIG. 5. (a) Absorption and (b) dispersion spectrum of the model system. The slope of the dispersion curve is shown in the inset. Here we consider the second cavity as a gain cavity, with κ2=κ1. All the other parameters are the same as before. two normal modes becomes significant while approach- ing the stability bound as depicted by the red dashed and blue dot-dashed curve of Fig. 5(a). In turn, it re- duces the overall gain of the composite system. Further, we investigate the effect of a gain-assisted auxiliary cav- ity on the medium’s dispersive response in Fig. 5(b). For J= 1.30κ1, the two absorption peaks produce two dis- tinct anomalous dispersion regions separated by a broad normal dispersive window. Weakening the exchange in- teraction strength reveals prominent normal dispersion around the resonance condition except for δ=ωb, and the window shrinks. In the inset of Fig. 5(b), we present the slope of the dispersive response due to the magnome- chanical resonance. The black solid curve of Fig. 5(b) suggests the occurrence of anomalous dispersion at the magnon-phonon resonance condition. Moreover, one can increase the steepness of the dispersion curve by simply approaching the instability threshold, as delineated by the red-dashed and blue-dotted-dashed curve of the inset of Fig.5(b). In the consecutive section, we will discuss how the change in the dispersion curve can produce con- trollable group velocity of the light pulses through the medium and investigate the role of the exchange interac- tion.C. Output probe transmission The output probe intensity from the system depends onitsabsorptiveresponse. Equation 11dictatethetrans- mission of the probe field and is presented in Fig. 6. For Fig.6(a), we consider both the microwave cavities as passive ones with identical decay rates, i.e.,κ1=−κ2. The black solid curve shows a broad absorptive response forJ= 0.40κ1. Increasing the exchange interaction strengthcausesgradualenhancementintheoutputprobe transmission, as delineated by the red-dashed and blue dot-dashed curve of Fig. 6(a), and the absorption win- dow splits into two parts. A precise observation confirms the presence of extremely weak transmission dip exactly atδ=ωbfor all the three exchange interaction strengths under consideration. In Fig. 6(b), we present the advan- tage of using a gain-assistedauxiliary cavity along with a CMM resonator to obtain a controllable amplification of the output probe field. We begin our discussion consid- ering the photon hopping interaction, J= 1.30κ1. The black solid curve of Fig. 6(b) estimates the normalized probe transmission of 6 .03. Here the normalization is done with respect to the input probe field intensity. By decreasingthe parameter J, weapproachthe unstablere- 0.5 0.75 1 1.25 1.5 δ/ωb00.20.40.60.81TJ = 0.4 κ1 J = 1.1 κ1 J = 1.3 κ1(a) 0.96 0.98 1 1.02 1.04 δ/ωb0200400600800 TJ = 1.30 κ 1 J = 1.10 κ1 J = 1.07 κ1(b) FIG. 6. Exchange interaction Jdependent normalized out- put probe transmission is plotted as a function of normalize d detuning between the control and the probe field when (a) both the cavities are passive ones, and (b) one is active and another one is passive.7 gion and observe the occurrence of a double transmission peak separated by a sharp and narrow transmission dip. The amplitude of the double transmission peak demon- stratestheprobepulseamplificationbyafactorof830,as presented by the blue dotted-dashed curve. However, an explicit observation suggests the output probe field am- plification by a factor of 67 at the resonance condition δ=ωb. The physics behind the probe field amplifica- tion can be well understood as: Introduction of gain to the second cavity compensates a portion of losses in the first cavity through J. This leads to an enhanced field amplitude in the first cavity. In the presence of mod- erate magnon-photon coupling it increases the effective magnon-photon coupling strength. Hence, we observe a higher transmission at the two sidebands but also find a large transmission dip at δ=ωb. D. Group delay Controllable group delay has gained much attention due to its potential application in quantum information processing and communication. The dispersive nature of the medium is the key to controlling the group delay of the light pulse under the assumption of low absorptionor gain. The pulse with finite width in the time domain is produced by superposing severalindependent waveswith different frequencies centered around a carrier frequency (ωs). The difference in time between free space propa- gation and a medium propagation for the same length can create a group delay. The analytical expression for the group delay can be constructed by considering the envelope of the optical pulse as f(t0) =/integraldisplay∞ −∞˜f(ω)e−iωt0dω, where˜f(ω) corresponds to the envelope function in the frequency domain. Accordingly, the reflected output probe pulse can be expressed as fR(t0) =/integraldisplay∞ −∞tp(ω)˜f(ω)e−iωt0dω, (13) =e−iωst0/integraldisplay∞ −∞tp(ωs+δ)˜f(ωs+δ)e−iδt0dδ, =tp(ωs)e−iωsτgf(t0−τg). (14) This expression can be obtained by expanding tp(ωs+δ) in the vicinity of ωsby a Taylor series and keeping the terms upto first order in δ. An expression for time-delay is obtained as [31, 41] τg= Re/bracketleftBigg −i tp(ωs)/parenleftbiggdtp dω/parenrightbigg/vextendsingle/vextendsingle/vextendsingle/vextendsingle ωs/bracketrightBigg , (15) which can be further simplified as τg=(α(ωs)−1)dβ dω/vextendsingle/vextendsingle/vextendsingle/vextendsingle ωs−β(ωs)dα dω/vextendsingle/vextendsingle/vextendsingle/vextendsingle ωs |tp(ωs)|2.(16)0.98 0.99 1 1.01 1.02 δ/ωb-18-12-60612τg (µs) J = 1.30 κ1 J = 1.10 κ1 J = 1.07 κ1 FIG. 7. Time delay of the probe pulse for different evanescent coupling strength Jhave been plotted against the normalized probe detuning δ/ωb, while the control power is 10 mW. All other parameters are taken as the same as in Fig. 5. From Eq. 16, the slope of the medium’s absorption and dispersion curves determine the probe pulse propagation delay or advancement. However, Fig. 5(b) suggests that the value of βis negligibly small near the magnomechan- ical resonance. Hence, the group delay depends on the first term of the numerator of Eq. 16. In Fig. 7, we ex- amine the effect of photon-photon exchange interaction on the probe pulse propagation delay when both cavities operate under balanced gain-loss condition. The system produces anomalous dispersion accompanied by a gain response. The black solid curve of Fig. 7depicts the probe pulse advancement of 2 .4µs for the photon hop- ping interaction strength, J= 1.3κ1. Moreover, one can enhance the effective gain and the slope of the anomalous dispersion curve by approaching the instability thresh- old. That, in turn, brings out the super luminosity of the output probe pulse, characterized by the advancement of 17.9µs as shown by the blue dotted-dashed curve of Fig. 7. To verify the above results, we consider a Gaussian probepulsewith afinite width aroundthe resonancecon- dition,i.e.,δ=ωb, and numerically integrate it by using Eq.13. The shape of the input envelope is considered as, ˜f(ω) =εp√ πΓ2e−(ω−ωs)2 Γ2, (17) where Γ is the spectral width of the optical pulse. We consider Γ to be 7 .17 kHz, such that the Gaussian enve- lope is well-contained inside the gain-window around the resonance condition ( δ=ωb), as depicted in Fig. 5(a). The dispersive, absorptive as well as gain response of the system can be demonstrated by examining the effect of the probe transmission coefficient ( tp) on the shape of the input envelope. The gain of the auxiliary cavity manipulates the probe transmission coefficient in such a way that it amplifies the intensity of the output probe pulse. The black solid curve depicts the output probe pulse amplification of 6 .2 for photon-hopping interaction strength,J= 1.30κ1. A decrease in the Jvalue8 gradually enhances the effective gain in the system. It -3 -1.5 0 1.5 3 Normalized time Γt020406080|tp|2J = 1.30 κ1 J = 1.10 κ1 J = 1.07 κ1-0.2 -0.1 0 0.1 0.2Γt 0.9960.99750.999 Normalized intensity FIG. 8. The relative intensity of the output probe pulse is plotted against the normalized time (Γt) for different photo n- photon exchange interaction strength when both cavities ar e operating under balanced gain-loss condition. amplifies the output probe transmission as presented by the red dashed and blue dotted-dashed curves of Fig. 8. We observe that the output field amplification can reach to a factor of 65.3 while considering the exchange interaction strength to be 1 .07κ1. Further decreasing the exchange interaction will lead to dynamical insta- bility in our model system. Interestingly, the temporal width of the probe pulse is almost unaltered during the propagation through the magnon-assisted double cavity system. This numerical result agrees with our analytical results for the output probe transmission, as shown in Fig. 6(b). Moreover, the importance of the photon-photon exchange interaction on the probe pulse propagation advancement can be observed from the inset of Fig. 8. The peak separation between the input pulse ( t= 0) and the output pulse for J= 1.30κ1estimates the probe pulse advancement of 2 .34µs. The red dashed, and blue dashed-dotted curve of the inset estimates the probe pulse advancement of 8 .75 µsand 13.30µsforJ= 1.10κ1and 1.07κ1, respectively. IV. CONCLUSION In conclusion, we have theoretically investigated the controllable output field transmission from a critically coupled cavity magnomechanical system. We drive the first cavity with a YIG sphere inside it, establishing the magnon-photon coupling. The photon exchange interac- tion connects the second microwave cavity with the first. An external magnetic field induces the deformation ef- fect of the YIG sphere. In this study, the interaction between the magnon and photon modes lies under the weak coupling regime. The medium becomes highly ab- sorbent when both cavities are passive, and the output probe transmission is prohibited. We introduce a gain to the auxiliary cavity to overcome this situation. It is noteworthy that the instability threshold must be close to the conventional exceptional point for a binary PT- symmetric system. At the magnomechanical resonance, the auxiliary cavity produces an effective gain associ- ated with anomalous dispersion. Further, decreasing the photon exchange interaction strength causes gradual en- hancement of the effective gain and the steepness of the dispersion spectrum. As a result, we observe a control- lable superluminal microwave pulse propagation associ- atedwithamplificationbyafactorof67. Bystudyingthe propagationdynamics ofa Gaussianprobepulse ofwidth 7.17 kHz, we confirm that the numerical study is con- sistent with the analytical results. Our study may find potential applicationsin weaksignalsensingand commu- nication purposes in the newly emerging field of cavity magnomechanics. [1]B.Z. Rameshti, S.V.Kusminskiy,J. A.Haigh, K.Usami, D. Lachance-Quirion, Y. Nakamura, C.-M. Hu, H. X. Tang, G. E. Bauer, and Y. M. Blanter, Physics Reports 979, 1 (2022). [2]F. Schlawin, D. M. Kennes, and M. A. Sentef, Applied Physics Reviews 9, 011312 (2022). [3]D. Lachance-Quirion, Y. Tabuchi, A. Gloppe, K. Usami, and Y. Nakamura, Applied Physics Express 12, 070101 (2019) . [4]R. Hisatomi, A. Osada, Y. Tabuchi, T. Ishikawa, A. Noguchi, R. Yamazaki, K. Usami, and Y. Nakamura, Phys. Rev. B 93, 174427 (2016) . [5]X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, Science Advances 2, e1501286 (2016) , https://www.science.org/doi/pdf/10.1126/sciadv.1501 286. [6]Y. Tabuchi, S. Ishino, T. Ishikawa, R. Ya- mazaki, K. Usami, and Y. Nakamura, Phys. Rev. Lett. 113, 083603 (2014) .[7]X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, Phys. Rev. Lett. 113, 156401 (2014) . [8]C. A. Potts, E. Varga, V. A. S. V. Bitten- court, S. V. Kusminskiy, and J. P. Davis, Phys. Rev. X 11, 031053 (2021) . [9]J. Li, S.-Y. Zhu, and G. S. Agarwal, Phys. Rev. Lett. 121, 203601 (2018) . [10]J. Li and S. Gr¨ oblacher, Quantum Science and Technology 6, 024005 (2021) . [11]J. Li, S.-Y. Zhu, and G. S. Agarwal, Phys. Rev. A 99, 021801(R) (2019) . [12]B. Sarma, T. Busch, and J. Twamley, New Journal of Physics 23, 043041 (2021) . [13]C. M. Bender and S. Boettcher, Phys. Rev. Lett. 80, 5243 (1998) . [14]C. M. Bender, D. C. Brody, and H. F. Jones, Phys. Rev. Lett. 89, 270401 (2002) .9 [15]R. El-Ganainy, K. G. Makris, M. Khajavikhan, Z. H. Musslimani, S. Rotter, and D. N. Christodoulides, Nature Physics 14, 11 (2018) . [16]W.D.Heiss, Journal of Physics A: Mathematical and Theoretical 45, 444016 (2012) . [17]K.G. Makris, R. El-Ganainy, D. N. Christodoulides, and Z. H. Musslimani, Phys. Rev. Lett. 100, 103904 (2008) . [18]Z. H. Musslimani, K. G. Makris, R. El- Ganainy, and D. N. Christodoulides, Phys. Rev. Lett. 100, 030402 (2008) . [19]A. Guo, G. J. Salamo, D. Duchesne, R. Morandotti, M. Volatier-Ravat, V.Aimez, G.A. Siviloglou, andD.N. Christodoulides, Phys. Rev. Lett. 103, 093902 (2009) . [20]B. Peng, S ¸. K. ¨Ozdemir, F. Lei, F. Monifi, M. Gianfreda, G. L. Long, S. Fan, F. Nori, C. M. Bender, and L. Yang, Nature Physics 10, 394 (2014) . [21]S.Longhi, Journal of Physics A: Mathematical and Theoretical 44, 485302 (2011) . [22]L. Feng, Y.-L. Xu, W. S. Fegadolli, M.-H. Lu, J. E. B. Oliveira, V. R. Almeida, Y.-F. Chen, and A. Scherer, Nature Materials 12, 108 (2013) . [23]T. Goldzak, A. A. Mailybaev, and N. Moiseyev, Phys. Rev. Lett. 120, 013901 (2018) . [24]D. Zhang, X.-Q. Luo, Y.-P. Wang, T.-F. Li, and J. Q. You,Nature Communications 8, 1368 (2017) . [25]J. Zhao, Y. Liu, L. Wu, C.-K. Duan, Y.-x. Liu, and J. Du,Phys. Rev. Appl. 13, 014053 (2020) . [26]F. Wang and C. Gou, Opt. Lett. 48, 1164 (2023) . [27]C. Kong, H. Xiong, and Y. Wu, Phys. Rev. Appl. 12, 034001 (2019) . [28]K. Ullah, M. T. Naseem, and ¨O. E. M¨ ustecaplio˘ glu, Phys. Rev. A 102, 033721 (2020) . [29]X. Li, W.-X. Yang, T. Shui, L. Li, X. Wang, and Z. Wu, Journal of Applied Physics 128, 233101 (2020) , https://doi.org/10.1063/5.0028395 .[30]G. S. Agarwal and S. Huang, Phys. Rev. A 81, 041803(R) (2010) . [31]A. H. Safavi-Naeini, T. P. M. Alegre, J. Chan, M. Eichen- field, M. Winger, Q. Lin, J. T. Hill, D. E. Chang, and O. Painter, Nature472, 69 (2011) . [32]S. Weis, R. Rivi` ere, S. Del´ eglise, E. Gavartin, O. Arcizet, A. Schliesser, and T. J. Kippenberg, Science330, 1520 (2010) , https://www.science.org/doi/pdf/10.1126/science.119 5596. [33]L. V. Hau, S. E. Harris, Z. Dutton, and C. H. Behroozi, Nature397, 594 (1999) . [34]V. Fiore, Y. Yang, M. C. Kuzyk, R. Barbour, L. Tian, and H. Wang, Phys. Rev. Lett. 107, 133601 (2011) . [35]V. Fiore, C. Dong, M. C. Kuzyk, and H. Wang, Phys. Rev. A 87, 023812 (2013) . [36]H. Jing, S ¸. K. ¨Ozdemir, Z. Geng, J. Zhang, X.-Y. L¨ u, B. Peng, L. Yang, and F. Nori, Scientific Reports 5, 9663 (2015) . [37]W. Li, Y. Jiang, C. Li, and H. Song, Scientific Reports 6, 31095 (2016) . [38]G. Vanderhaegen, C. Naveau, P. Szriftgiser, A. Kudlinski, M. Conforti, A. Mussot, M. Ono- rato, S. Trillo, A. Chabchoub, and N. Akhmediev, Proceedings of the National Academy of Sciences 118, e2019348118 (2021) , https://www.pnas.org/doi/pdf/10.1073/pnas.201934811 8. [39]S.-N. Huai, Y.-L. Liu, J. Zhang, L. Yang, and Y.-x. Liu, Phys. Rev. A 99, 043803 (2019) . [40]J. M. P. Nair, D. Mukhopadhyay, and G. S. Agarwal, Phys. Rev. B 105, 214418 (2022) . [41]S. Das and T. N. Dey, J. Opt. Soc. Am. B 39, 859 (2022) . | 2023-06-07 | We study the controllable output field generation from a cavity
magnomechanical resonator system that consists of two coupled microwave
resonators. The first cavity interacts with a ferromagnetic yttrium iron garnet
(YIG) sphere providing the magnon-photon coupling. Under passive cavities
configuration, the system displays high absorption, prohibiting output
transmission even though the dispersive response is anamolous. We replace the
second passive cavity with an active one to overcome high absorption, producing
an effective gain in the system. We show that the deformation of the YIG sphere
retains the anomalous dispersion. Further, tuning the exchange interaction
strength between the two resonators leads to the system's effective gain and
dispersive response. As a result, the advancement associated with the
amplification of the probe pulse can be controlled in the close vicinity of the
magnomechanical resonance. Furthermore, we find the existence of an upper bound
for the intensity amplification and the advancement of the probe pulse that
comes from the stability condition. These findings may find potential
applications for controlling light propagation in cavity magnomechanics. | Gain assisted controllable fast light generation in cavity magnomechanics | 2306.04390v1 |
Design of an optomagnonic crystal: towards optimal magnon-photon mode matching at the microscale Jasmin Graf,1,2Sanchar Sharma,1Hans Huebl,3,4,5and Silvia Viola Kusminskiy1,2 1Max Planck Institute for the Science of Light, Staudtstraße 2, 91058 Erlangen, Germany 2Department of Physics, University Erlangen-Nürnberg, Staudtstraße 7, 91058 Erlangen, Germany 3Walther-Meißner-Institut, Bayerische Akademie der Wissenschaften, Walther-Meissner-Straße 8, 85748 Garching, Germany 4Physik-Department, Technische Universität München, James-Franck-Straße 1, 85748 Garching, Germany 5Munich Center for Quantum Science and Technology (MCQST), Schellingstraße 4, 80799 München, Germany We put forward the concept of an optomagnonic crystal: a periodically patterned structure at the microscale based on a magnetic dielectric, which can co-localize magnon and photon modes. The co-localization in small volumes can result in large values of the photon-magnon coupling at the single quanta level, which opens perspectives for quantum information processing and quantum conversionschemeswiththesesystems. Westudytheoreticallyasimplegeometryconsistingofaone- dimensional array of holes with an abrupt defect, considering the ferrimagnet Yttrium Iron Garnet (YIG) as the basis material. We show that both magnon and photon modes can be localized at the defect, and use symmetry arguments to select an optimal pair of modes in order to maximize the coupling. We show that an optomagnonic coupling in the kHz range is achievable in this geometry, and discuss possible optimization routes in order to improve both coupling strengths and optical losses. I. INTRODUCTION Progress in fundamental quantum physics has by now established a basis for developing new technologies in the fields of information processing, secure communication, and quantum enhanced sensing. In order to perform these tasks, physical systems are needed which are ca- pable of processing, storing, and communicating infor- mation in a quantum coherent manner and with a high fidelity. Similar to the classical realm, accomplishing this goal requires different degrees of freedom and efficient couplings between them, giving rise to hybridsystems. In this context, systems at the mesoscopic scale (with di- mensions ranging from nanometers to microns) are spe- cially interesting since their collective degrees of freedom can be tailored [1]. An important and successful exam- ple of these mesoscopic hybrid systems are optomechan- ical systems [2], where light couples to mechanical mo- tion. Seminal experiments in these systems have demon- strated extra-sensitive optical detection of small forces and displacements [3–10], manipulation and detection of mechanical motion in the quantum regime with light [11– 13], and the creation of nonclassical light and mechanical motion states [11, 14, 15]. In the recent years, the family of hybrid quantum sys- tems has been extended by incorporating magnetic ma- terials, where the collective spin degree of freedom can be exploited. For example in spintronics [16], information is carried by spins (as opposed to electrons) in order to re- move Ohmic losses and to increase memory and process- ing capabilities [17, 18]. An ultimate form of spintronics is the new field of quantum magnonics [19, 20], where superconducting quantum circuits couple, via microwave yxz(b) (c)(a)zxyd2rDefect areaaMagnetic modeOptical modeG YIG-SiN-heterostructure YIGSiNSiN |Hz| 01|δmz| Optical modeMagnetic modeFigure 1. Investigated geometry: (a) Optomagnonic crystal with an abrupt defect at its center for localizing an optical and a magnon mode at the same spot in the defect area. (b) Optomagnonic crystal from the side representing a het- erostructure. (c) Mode profiles of the localized optical and magnon mode discussed in the main text. Note: all mode shape plots are normalized to their corre- sponding maximum value. fields in a cavity, coherently to magnetic collective ex- citations (magnons) [21, 22]. Such systems are promis- ing for generating and characterizing non-classical quan- tum magnon states [19, 23–25], quantum thermometry protocols [26], and for developing microwave-to-optical quantum transducers for quantum information process- ing [27, 28]. The coherent coupling of magnons to op- tical photons has also been demonstrated in recent ex-arXiv:2012.00760v2 [cond-mat.mes-hall] 25 Mar 20212 periments [29–33], in what have been denominated op- tomagnonic systems [28, 34–40]. In current experiments exploring optomagnonics, the ferrimagnetic dielectric Yttrium Iron Garnet (YIG) is used as the magnetic element, since YIG presents low absorption and a large Faraday constant in the infrared (= 0:069 cm 1andF= 240 deg=cm@1:2µm[29, 30, 41–43]). The coupling between spins and optical pho- tons is an second order processes involving spin-orbit coupling and it is generally small. This can be enhanced using a well polished sphere that acts as an optical cav- ity, trapping the photons by total internal reflection in order to effectively enhance the spin-photon coupling. The coupling, however, still remains too small for ap- plications. This is due to the large size of the used YIG spheres (the coupling increases as the volume of the cavity decreases [34]), with radius of the order of hun- dreds of microns, and, concomitantly, the large differ- ence between the optical and the magnetic mode volume (VmagVopt), by which most of the magnetic mode vol- ume does not participate in the coupling. This can be partially mitigated by making smaller cavities [44, 45], but care has to be taken both to obtain a good mode matching and to retain a good confinement of the optical mode in order to minimize radiation losses. Recent pro- posals have investigated one dimensional layered struc- tures to this end [37, 46]. In order to tackle these issues, we propose an opto- magnonic array at the microscale, which acts simultane- ously as a photonic crystal [47], determining the opti- cal properties of the structure, and as a magnonic crys- tal [48–50] with tailored magnetostatic modes. Our pro- posal is inspired in the success of this approach for op- tomechanical crystals, which can be designed such as to enhance the phonon-photon coupling by many orders of magnitude [51–74]. In our case, we use similar concepts in order to design the coupling between photonic and magnonic modes. Although similar conceptually, mag- netic materials present new challenges for the design, due to the complexity of the magnon modes. Photonic crystals are the basis for many novel appli- cations in quantum information, and are of high interest due to their ability to guide [75–79] and confine [80–85] light, allowing for example to enhance non-linear optical interactions [86–89]. In turn, magnonic crystals can be designed to create reprogrammable magnetic band struc- tures [90], to act as band-pass or band-stop filters, or to create single-mode and bend waveguides [91–94]. Addi- tionally these crystals can be used for spin wave comput- ing via logical gates [95–97]. An advantage of magnonic crystalsistheirscalability, theirlowenergyconsumption, and possibly faster operation rates [49, 97, 98]. Together withthestateoftheartinoptomagnonicsdetailedabove, thisprovidesagreatincentivetoexplorethepossibilityof anoptomagnonic crystal, combining both photonic and magnetic degrees of freedom.Specifically, we consider an optomagnonic crystal con- sisting of a dielectric magnetic slab (YIG in our simula- tions) with a periodic array of holes along the slab and with an abrupt defect in the middle. The repeated holes at each side of the defect act as a Bragg mirror for both optical and magnetic modes, localizing them in the re- gion of the defect (see Fig. 1(a)+(c)). We show that this structure can co-localize photonic and magnonic modes, and explore how the symmetry of the modes can be used to optimize their coupling. We find that coupling rates in the range of kHz are achievable in these structures, and that optimization of the geometry can lead to higher coupling values, indicating the promise of this approach. Further optimization is nevertheless needed to improve the decay rates, in particular the optical quality factor is low compared to the state of the art in non-magnetic structures (where Silicon is used as the dielectric). This manuscript is organized as follows. In Sec. II we derive the general expression for the coupling of magnons to optical photons and discuss the normalization of the modes required to find the photon-magnon coupling at the single quanta level, denominated optomagnonic cou- pling. The remaining sections refer to the numerical method for evaluating this coupling. For our simula- tions we choose YIG as the magnetic material, in line with the material of choice in experiments. In Sec. III we discuss the properties of the proposed structure as a photonic crystal. In Sec. IV, in turn, we investigate its properties as a magnonic crystal. Sec. V combines the re- sults in order to numerically evaluate the optomagnonic coupling for appropriately chosen confined modes. For concreteness, we focus on the coupling between one sin- gle magnon mode and one single optical mode. Sec. VI is devoted to a discussion on how the structure can be op- timized and presents results for an optimized geometry. TheconclusionsandanoutlookarepresentedinSec.VII. The supplementary material contains further details of the analytic calculations and of the simulations. II. OPTOMAGNONIC COUPLING In this section we derive the theoretical expression for the coupling rate between magnons and photons. The instantaneous electromagnetic energy is [99] Etot em=1 2 dr D(r;t)E(r;t) +B(r;t)H(r;t) (1) withDthe displacement field, Ethe electric field, B the magnetic induction, and Hthe magnetic field. In complex representation D= (D+D)=2andE= (E+E)=2, and similar for the magnetic induction and field. The effect of the magnetization Mis to modify the displacement field as D(r;t) = "[M(r;t)]E(r;t)where3 the components of the permittivity tensor "are [100, 101] "ij(M) ="0 "rij ifFX kijkMk+fCMiMj! ; (2) with"0the vacuum permittivity, "rthe relative per- mittivity,ijkthe Levi-Cevita tensor, and ffF;fCgma- terial dependent magneto-optical constants. At opti- cal frequencies the second term in Eq. (1) can be ne- glected [101, 102], being smaller than the first by the fine structure constant squared, and the permeability of the material can be set to the vacuum permeability 0. The magneto-optical constants can be related to the Faraday rotation Fand the Cotton-Mouton ellip- ticityCper unit length as F=!=(2cp"r)fFMsand C=!=(2cp"r)fCM2 s, withcthe speed of light in vac- uum andMsthe saturation magnetization. We are interested in how light couples to the fluctua- tionsofthemagnetizationaroundthestaticgroundstate. We consider norm-preserving small fluctuations, M(r;t) =M0(r)s 1 M(r;t) Ms2 +M(r;t);(3) where the ground state satisfies M0M0=M2 sand the fluctuations are perpendicular to local equilibrium magnetization MM0= 0. In complex notation M= [M+M]=2. The correction to the electromag- netic energy stemming from the interaction between the light field and the magnetization can be rewritten as Eem=1 8 dr E(r;t)D(r;t)+E(r;t)D(r;t) ;(4) ignoring E(r;t)D(r;t)andE(r;t)D(r;t)in the rotating wave approximation. Inserting the relation be- tween the displacement and the electric field along with the permittivity in Eq. (2) gives Eem=EF em+EC emwhere EF em="0fF 8 dr[i(EE)M+h.c.](5) is the Faraday contribution and EC em=0fC 8 dr[E(MM 0+M0M)E+h.c.] (6) the Cotton-Mouton one. We have used the dyadic no- tation and neglected all terms that represent a constant energy shift or that are higher order in M. Quantizing this expression leads to the optomagnonic coupling Hamiltonian. By assuming that the magnetic material acts as an optical cavity, the electric field of the light can be quantized by using the annihilation (cre- ation) operator ^a(y) of one photon E(r;t)!2iX E(r) ^a(t); (7)withE() the mode shape, and the mode index. We note that we identified E(r;t)with 2E+(r;t)from the well known quantization expression of the electric field [103] E(r;t) =E+(r;t) +E (r;t) =iX h E(r) ^a(t) E (r) ^ay (t)i :(8) In order to find the coupling per photon, we normalize the electromagnetic field amplitude to one photon over the electromagnetic vacuum [104] ~! 2"0= dr"r(r)jE(r)j2: (9) The spin waves can be quantized as M(r;t)!2MsX
m
(r)^b
(t);(10) where ^b(y)
annihilates (creates) one magnon, m()
is the mode shape, and
the mode index. We note that as in the optical case we identified M(r;t)with 2M+(r;t) from the magnetic quantization expression M(r;t) =M+(r;t) +M (r;t) =MsX
h m
(r)^b
(t) +m
(r)^by
(t)i : (11) In order to normalize the amplitude of the magnetic fluc- tuations to one magnon, we use the following expression derived in the supplementary material (see appendix A) gB Ms= drim0 m
m
(12) withgthe g-factor, Bthe Bohr magneton, and m0= M0=Ms. This expression is valid for arbitrary magnetic textures and it is consistent with the normalization de- rived previously for a uniform ground state [105]. The quantized optomagnonic energy, neglecting the constant energy shifts, leads to the coupling Hamiltonian ^Hom=~X
h G
^ay ^a^b
+h.c.i (13) with the coupling constant G
=GF
+GC
, where GF
= i"0"r ~Fn dr E E m
;(14) GC
="0"r ~Cn dr[m0E] [E m
] +"0"r ~Cn dr[m0E ] [Em
](15)4 are, respectively, the Faraday and Cotton-Mouton com- ponents of the optomagnonic coupling constant, being n the light wavelength inside the material. The coupling between optical photons and magnons, as can be seen from Eq. 13, involves a three-particle pro- cess in which a magnon is created or annihilated by a two-photon scattering process. This is an example of parametric coupling, andreflectsthefrequencymismatch between the excitations. The coupling can be enabled by a triple-resonance, where the frequency of the magnon matches the frequency difference between two photonic modes [29–31, 35], or, in the case of scattering with a single photon mode, by an external driving laser at the right detuning [34]. If the laser is red (blue) detuned by the magnon frequency, implying a lower (higher) driving frequency than the photon resonance, it will annihilate (create) magnons. In the red detuned regime this can be used, for example, to actively cool the magnon mode to its ground state [2, 40, 106]. In this work we focus on the coupling between a given magnon mode and a given optical mode hosted by the 1D optomagnonic crystal shown in Fig. 1(a)+(b). Hence we set=and drop the indices in the following. For con- creteness, we focus on GFas the analysis for GCis anal- ogous.GFis proportional to the overlap of the magnon’s spatial distribution with the electric component of the optical spin density defined as [107] Sopt(r) ="0 2i!opt EE : (16) The optical spin density is finite only for fields with cer- tain degree of circular polarization, and points perpen- dicular to the plane of polarization. III. PHOTONIC CRYSTAL Photonic crystals are engineered structures which, by proper shape design, can confine light to a specific re- gion. These are formed by low-loss media exhibiting a periodic dielectric function "(r), with a discrete transla- tional symmetry "(r) ="(r+R)for any R=nawith nan integer and athe lattice constant given by the im- posed periodicity. Photonic band gaps arise at the edges of the Brillouin zone (BZ) k==adue to the periodicity imposed by the susceptibility of the crystal on the electric field, with wavelength = 2a(correspondingtotheedgeoftheBZ). For example, in a 1D photonic crystal (see Fig. 2(a)) the symmetry of the unit cell around its center implies that the nodes of the standing light wave must be centered either at each low- "layer or at each high- "layer. The latter necessarily has lower energy than the former, re- sulting in a band gap. The position of the photonic band gap is given by the mid-gap frequency at the BZ edge. In the case of two materials with refractive indices n1 (b)(a)1D photonic crystal1D photonic crystal with defectɛ2ɛ2ɛ2ɛ2ɛ1ɛ1ɛ1ɛ2ɛ2ɛ2ɛ2ɛ1ɛ1ɛ1localization area“mirror”“mirror”d1d2aFigure 2. General structure of a 1D photonic crystal and mode localization at a defect: (a) 1D photonic crystal con- sisting of periodic layers alternated by the lattice constant awith different dielectric constants "1> " 2and widths d1 andd2. (b) A defect breaks the symmetry and can pull a band-edge mode into the photonic band gap. Since a mode in the band gap cannot propagate into the structure, the light is Bragg-reflected and is thus localized (see e.g. [47]). andn2and thicknesses d1andd2=a d1, the normal incidence gap is maximized for n1d1=n2d2. In this case the mid-gap frequency is given by [47] !mg=n1+n2 4n1n22c a(17) withn1=p"1,n2=p"2. The corresponding vacuum wavelength mg= (2c)=!mgthereby satisfies the rela- tionsmg=n1= 4d1andmg=n2= 4d2meaning that the individual layers are a quarter-wavelength thick. An input at frequencies within the photonic band gap is reflected entirely except for an exponentially decaying tail inside the crystal. Thus, two of such crystals can be used to create a Fabry-Perot like cavity. More concretely, as shown in Fig. 2(b), a defect in the form of a layer with a different width breaking the symmetry of the crystal may permit localized modes in the band gap by consec- utive reflection on both sides. Since the light is localized in a finite region, the modes are quantized into discrete frequencies. We note that the degree of localization is the largest for modes with frequencies near the center of the gap [47]. For our purposes, we consider a geometry in which the permittivity can take two distinct values, attained by holes carved into a dielectric slab (see Fig. 1(a)). The typical material used for photonic crystals is silicon due toitshighrefractiveindexatopticalfrequencies, "r= 12. We use instead YIG for our study, which is a dielectric magnetic material transparent in the infrared range with "r= 5[108]. The lower dielectric constant reduces the confinement of the optical modes along the height of the slab, which is reflected in low optical quality factors as discussed below. This structure is a 1D photonic crystal, periodic in one direction (chosen to be the ^x-direction), with a band gap along this direction and which confines light through index guiding [47] (a generalization of total internal reflection) in the remaining directions. In order to localize an optical mode in this structure we create a defect by increasing the spacing between the two middle holes, which pulls a mode into the band gap. We note that due to the (discrete) periodicity, the crystal only5 possesses an incomplete band gap and the localized mode can scatter to air modes [47]. We search for a localized mode in the infrared fre- quency range where YIG is transparent and presents low absorption [42, 109]). Thus, the geometrical parameters of the crystal need to be chosen in such a way that the band gap lies in the desired frequency range. We choose a lattice constant of a= 450 nm which gives a mid- gap frequency of !mg= 2240 THz (corresponding to1250 nm), using Eq. (17) with refractive indices of YIGn1=nYIG=p 5and of airn2=nair= 1. Note that we choose a lattice constant that allows us to work in the transparency frequency range for the optics, and which at the same time is small enough in order to reduce the computational cost of the micromagnetic simulations of the corresponding magnonic crystal in the next section. Using the relation dairnair=dYIGnYIGfor a maximized normal incidence gap we find the optimal radius of the air holes as rair=nYIG nYIG+ 1a 2(18) withdair= 2rair=a dYIG, from which we obtain rair= 155:25 nm. In order to find the mode with the least losses, the defect width is optimized in order to lo- calize the desired mode most effectively to the defect. We find the optimal defect size, defined as the center- to-center distance between the two bounding holes (see Fig. 1(a)), to be d= 731 nm , obtained by evaluating the transmission spectra as a function of the defect size (we used the electromagnetic simulation tool MEEP to this end [110]). In order to get a good quality factor of the localized mode we need to insert as many air holes as possible. For creating a compromise between short com- putational evaluation time (especially important for the magnetic simulations discussed later) and a good quality factor we chose N= 12holes at each side of the defect. Therefore the investigated crystal is in total l= 11:75µm long. The overall width of the wave guide is w= 600 nm and its height is h= 60 nm, again to keep the magnetic simulations (which we detail in the next section) feasible. Such a thin slab will not be good at confining the opti- cal modes along its height, since it is much smaller than the light wavelength in the material. In order to increase the optical quality factor without influencing the mag- netics we sandwich the crystal with two Si 3N4layers (see Fig. 1(b)) with a height of hSi3N4= 200 nm as proposed in [111]. Si 3N4has an index of refraction similar to YIG (nSi3N4=p 4), so that the combined structure acts ap- proximately as a single cavity for the light and its height is roughly half a wavelength, enough to provide a rea- sonable confinement. The presented simulations include these two extra layers. We now turn to categorizing the photonic modes of the crystal by using its three mirror symmetry planes (see Fig. 3(a)). This imposes several restrictions on the (b)(a)z=0TE-likeTM-like Mirror symmetry (y=0)Mirror symmetry (x=0)Symmetries of the photonic crystalMirror symmetry (z=0)Symmetry of TE and TM modesyzEEFigure 3. Symmetries of the optical modes in a periodic waveguide: (a) Symmetry planes of the investigated 1D pho- tonic crystal shown in Fig. 1(a). (b)Symmetry of a transverse electric (TE)-like and a transverse magnetic (TM)-like opti- cal mode in a thin 3D structure. The red arrows indicate the electric field vector Ewhich forz= 0(middle of the crystal along the height) lie in plane for TE-like modes and point out of plane for TM-like modes. For z6= 0this is not fulfilled anymore (see e.g. [47]). mode shape and the mode polarization. We define the three mirror symmetry operations ^E zE(r) =0 @Ex(x;y; z) Ey(x;y; z) Ez(x;y; z)1 A; ^E yE(r) =0 @Ex(x; y;z) Ey(x; y;z) Ez(x; y;z)1 A; ^E xE(r) =0 @ Ex( x;y;z ) Ey( x;y;z ) Ez( x;y;z )1 A:(19) In the following we restrict the discussion to transverse electric modes with an in-plane electric field, which are the modes of interest for the magnetic configuration we consider, as it will be clear from the next section. We note that structures made of a high- "material with air holesfavourabandgapfortransverseelectricmodes[47], which is advantageous for our purposes. Unlike in 2D, in three dimensions we cannot generally distinguish be- tween transverse electric (TE) and transverse magnetic (TM) modes. However, provided that the crystal has a mirrorsymmetryalongitsheight(under ^E z), andthatits thickness is smaller than the mode wavelength, the fields are mostly polarized in TE-like and TM-like modes [47] (see Fig. 3(b)). Since a TE-like mode has a non-zero elec- tric field in the plane of the crystal ( xy-plane), both Ex andEycannot be odd as a function of z(see Fig. 3(b)). From Eq. (19), this implies that the mode must be even under ^E z. Similar symmetry considerations [47] show that a TE-like mode must satisfy ^E zE(r) =E(r); ^E yE(r) = E(r); ^E xE(r) = E(r):(20)6 zxyzxy -11(b)(a) |Ey||Ex||Hz|Mode shape 01Optical spin density ~ i(E* x E)z Band diagramSymmetriesRe[Ex]Re[Ey]Extended states 0𝞹/akxfno defectDiscrete states1. band2. band ~ 200 Hz~ 270 HzDefect state07 x 10-3kx [1/nm] -11 -11 01(c) Re[δmz]Im[δmz]Mode profileBand diagram|δmz| SymmetriesLocalized mode @ 13.12 GHz016f [GHz]FT[mz]normMode spectrumkx = 𝞹/aδmzkx [1/nm] f [GHz]016𝛅mzdefect07 x 10-3 0𝞹/akxzxy Figure 4. Optical (a and b) and magnetic modes (c): (a) Band diagram (obtained with MEEP) for TE-like modes within the irreducible BZ with a state that was pulled into the gap from the upper band-edge state by the insertion of a defect (note that the gap-state was not obtained by band diagram simulations). The bands in the green shaded area representing the light cone are leaky modes which couple with radiating states inside the light cone [112]. From the shape of the localized defect mode (obtained with Comsol) with a frequency of !opt= 2246 THz (middle layer in the xy-plane) we see that this mode is odd with respect to x= 0andy= 0(and even with respect to z= 0). (b) Optical spin density (middle layer in the xy-plane) of the localized mode, fulfilling the same symmetries, see main text. (c) Band diagram of backward volume waves within the irreducible BZ showing magnetic modes with extended k-values but preferring wave vectors at the edge of the BZ. The highest excited localized mode has a frequency of !mag= 213:12 GHzand is odd along the mirror symmetry planes for x= 0 andy= 0(and additionally even with respect to the plane for z= 0). The dashed line in the middle inset shows the mode spectrum in case of no defect. Note: all mode shape plots are normalized to their corresponding maximum value. For evaluating the optical modes we used the two finite elementtoolsMEEP[110]andComsolMultiphysics[113] (see appendix B). The simulated band structure for TE- like modes in the considered photonic crystal shows a broad band gap in the infra-red frequency range with a nicely pulled defect band (see Fig. 4(a)) which is ex- tended in frequency space resulting from the confinement in real space. The defect mode in the gap at the edge of the BZ has a frequency of !opt2246 THz (obtained by Comsol, 205 THz/1550 nm according to MEEP, note that the difference is due to a reduction of the sim- ulation geometry to 2D in order to save simulation time) with a damping factor of opt20:1 THzwhich gives a linewidth (FWHM) of
opt= 2opt20:2 THz. Using the values obtained by Comsol this gives an optical quality factor ofQ=!opt=(4opt) = 1250 , which is in the expected range for this kind of geometry [114] (note that MEEP gives a roughly three times larger value due to the 2D simulation which effectively resembles a sim- ulated system of infinite height). The obtained quality factor is however small compared to 1D crystals made of silicon with a smooth defect, where quality factors in the order of 104 106can be achieved [53, 55, 57]. The corresponding mode shape in real space is shown in Fig. 4(a). We see that the mode is nicely localized at the defect. Furthermore the Excomponent is even (odd) as a function of x(y), whereas the Eyis even as a function of both xandyfulfilling the symmetry require- ments for a TE-like mode given in Eqs. (20). Due to thissymmetry, we can disregard the Ezcomponent here since Ez0. FortheFaradaycomponentoftheoptomagnonic coupling, the relevant quantity is the electric component of the optical spin density, Sopt/EE[see Eq. (14)]. Soptpoints mostly along z-direction, is odd as a function ofxandy, and is even along z(see Fig. 4(b) and 6(b)). IV. MAGNONIC CRYSTAL Asphotoniccrystalscontroltheflowoflight, magnonic crystals can be used to manipulate the spin wave dynam- ics in magnetic materials. In general a magnonic crystal is made of a magnetic material with a periodic distribu- tion of material parameters. Examples include the mod- ulation of the saturation magnetization or the magne- tocrystalline anisotropy, a periodic distribution of differ- ent materials, or the modulation by external parameters, such as an applied magnetic field [48–50]. Historically, magnonic crystals precede photonic crystals [115, 116]. Unlike in photonic or phononic crystals, the band struc- ture in magnonic crystals depends not only on the peri- odicity of the crystal but also on the spatial arrangement of the ground state magnetization, resulting in an ad- ditional degree of freedom. Hence the band structure depends on the applied external magnetic field, the rela- tive direction of the wave vector, the shape of the mag- net, and the magnetocrystalline anisotropy of the mate- rial [48–50]. In this section we study the properties of the crystalpresentedinSec.III(seeFig.3(a)), asamagnonic7 crystal. In the following we consider magnetic excitations which are non homogeneous in space, and we focus only on systems in the presence of an external magnetic field saturating the magnetization in a chosen direction. In this case spin waves can be divided into three classes: if all spins precess uniformly in phase, the mode is homo- geneous and denominated the Kittel mode. If the dis- persion is dominated by dipolar interactions (which is usually the case for wavelengths above 100 nm) the exci- tationsarecalleddipolarspinwaves. Forwavelengthsbe- low100 nmthe exchange interaction dominates instead, giving rise to exchange spin waves. The frequencies of the dipolar spin waves lie typically in the GHz-regime, whereas the exchange spin waves have frequencies in the THz-regime. Since the size of the structure considered in this work is in the micrometre range, we will focus on dipolarspinwaves. Forthiscase, themodescanbeclassi- fiedfurtherbytheirpropagationdirectionwithrespectto the magnetization. For an in-plane magnetic field, modes with a frequency higher than the frequency of the uni- form precession tend to localize at the surface and have a wave vector pointing perpendicular to the static magne- tization M0and thus the external field, k?M0kHext (seeFig.5(a)). ThesemodesarecalledsurfaceorDamon- Eshbach modes [117, 118]. If the wave vector is parallel to the external field such that kkM0kHextholds, the waves are called backward volume waves and their fre- quency is smaller than the frequency of the Kittel mode (see Fig. 5(a)). Finally, if the external field and the mag- netization are normal to the crystal’s plane and the wave vector lies in plane k?M0kHextthese waves are called forward volume waves (see Fig. 5(a)) [50, 119]. In the fol- lowing we restrict the discussion to external fields which are applied in the plane of the crystal. Similar to light modes in photonic crystals, magnon modes can also be localized within a certain region in the magnonic crystal. It is well known that the two di- mensional periodic modification of a continuous film, for example by the insertion of holes (denominated antidot arrays) can drastically change the behavior of the spin waves [120, 121]. In this case the modes have either a localized or extended character. The localized mode is a consequence of non-uniform demagnetization fields cre- ated by the antidots. These fields change abruptly at the edges of the antidots and act as potential wells for the spin waves [50]. Thus, the above designed crystal, which localizes the optical mode by the insertion of a defect, is also a good candidate for acting as a magnonic crys- tal localizing magnetic modes via the holes. Although the geometry of the crystal is optimized for the optics, it should be able to host and localize magnetic modes due to its shape and material (YIG). Therefore we do not change the crystal further and use this structure as a proof of principle. This implies that we expect con- siderable room of improvement with respect to the opto- (b)(a)BWVWHextkFWVWkHextHextkSWDipolar spin waves 𝞹-rotation symmetry (around x-axis)Mirror symmetry (x=0)Symmetries of the magnonic crystalFigure 5. Dipolar spin wave types and symmetries in the magnonic crystal: (a) Dipolar spin waves can be divided into three types: backward volume waves (BWVW) with their wave vector parallel to the external field which both lie in the plane of the structure ( kkm0kHext). Forward volume waves (FWVW) with their wave vector in plane and perpen- dicular to the external field which lies normal to the struc- ture’s plane ( k?m0kHext). Surface waves are also forward volume waves but they have their wave vector in plane and perpendicular to the external field which also lies in plane of the structure. ( k?m0kHext). (b) Symmetries of the inves- tigated 1D magnonic crystal shown in Fig. 1(a). Since the external magnetic field breaks two mirror symmetry planes only the mirror symmetry plane normal to the saturation di- rection remains. Additionally a -rotation symmetry around the saturation axis is present. magnonic coupling rates obtained in this structure. YIG is a good choice for magnonics since it has the lowest spin wave damping when compared to other materials commonly used [49]. It is however difficult to pattern at the microscale, but recent advances in fabrication show great promise in this respect [122, 123]. For concreteness, in the following we proceed to design the Faraday part of the optomagnonic coupling GF, see Eq. (14). Since GFis proportional to the overlap inte- gral between the optical spin density and the magnon mode, we search for a magnon mode with the same symmetries as the optical spin density, in order to get the highest possible overlap. Like in the optical case, the magnonic crystal has three mirror symmetry planes (z= 0; y= 0; x= 0). However, the external applied magnetic field saturating the magnetization breaks two of these symmetries and thus only the mirror symmetry w.r.t. the plane perpendicular to the external field re- mains (see Fig. 5(b)). Note that the magnetization is a pseudo vector and its components perpendicular to the mirror does not change. Thus, the mirror operation is inverted from Eq. (19), ^M xm(r) =0 @mx( x;y;z ) my( x;y;z ) mz( x;y;z )1 A=m(r):(21) Since the optical spin density pointing along ^zis odd as a function of x, we require mzto be odd as well and consequently mto be even under ^M x. Additionally, a8 -rotationaroundthe ^x-axissymmetryremainsunbroken ^R xm(r) =0 @mx(x; y; z) my(x; y; z) mz(x; y; z)1 A=m(r):(22) Invoking again the symmetries of the optical spin den- sity (odd as a function of yand even with z) we consider modes with even rotational symmetry. We note that due to the different symmetries respected by the photon and magnon modes, we choose the symmetries of the modes in such a way that they preferably match in the xy-plane, which is the most relevant dimension for thin structures. In this case, the symmetries of the optical and the mag- netic mode along the height do not necessarily match. For thin films however they do, see Fig. 6(b). Since spin waves are excited by an external magnetic pulse which controls the direction of the wave vector k, the pulse also breaks the mirror symmetries of the crys- tal. Therefore we focus on a setup which conserves the relevant mirror symmetry, and only excite backward vol- ume waves where the external saturation field and the wave vector of the mode are parallel and lie in the plane of the crystal. We note that this configuration is also the most favourable one from an experimental point of view, and additionally the configuration most likely used in magnonic devices [119]. We evaluated the magnetization dynamics numerically bymeansofthefinitedifferencetoolMuMax3[124]which solves for the Landau-Lifshitz-Gilbert equation of motion for the local magnetization vector (see appendix C). In order to excite magnon modes with the desired symme- try, we use a 2D antisymmetric sinc-pulse which should moreover avoid spurious effects in the spectrum [125] Hpulse =Hpulsesin2(!ct) !ctsin2(kcx) kcxsin2(kcy) kcyey;(23) pointing along the ^y-direction in order to excite back- ward volume waves [119]. The cut-off frequency was chosen to be !c= 216 GHzand the cut-off wave vector to be kc==ain order to concentrate all the excitation energy in the first BZ. Since this pulse is cen- tered in the middle of the crystal, we only excite modes around the crystal’s center. The external saturation field was set to H ext= 400 mT (found by hysteresis) and the pulse field to H pulse = 0:4 mT. We note that the pulse strength should be a small perturbation of the saturation field in order to minimize non linear effects. We used the material parameters for YIG, Ms= 140 kA=m(sat- uration magnetization), Aex= 2 pJ=m(exchange con- stant),Kc1= 610 J=m3(anisotropy constant) with the anisotropy axis along ^z[126]. In order to accelerate the simulations, we used an increased Gilbert damping pa- rameter
= 0:008(compare to
10 5 10 4for YIG) [127, 128]. In the following considerations we focus only on the mzcomponent of the magnetization dynamics, since the yxzzxy -11 01 -11(b)(a) Different symmetries(E*xE)zMirror symmetryRe[δmz] Rotational symmetrySpatial shape of the couplingRe[G]Im[G]GFigure 6. Spatial shape of the coupling and different sym- metries of the optical and the magnetic mode: (a) Spatial shape of the coupling similar to the magnon mode shape. (b) Different symmetries along the crystal’s height of the opti- cal spin density and the magnon mode. Due to the external magnetic field the mirror symmetry along the height is broken andonlya-rotationsymmetryremains, resultingindifferent mode shapes along the height of the crystal. For thin films this difference is rather small. Note: all mode shape plots are normalized to their corre- sponding maximum value. optical spin density of a TE-like mode mostly points into the^z-direction, rendering mxandmyirrelevant for GF (see Eq. (14)). We find that the optical defect also acts as confinement of the magnetic mode, resulting in the defect like dispersion relation presented in Fig. 4(c). The obtained band structure shows modes around the edge of the BZ with extended wave vector character, imply- ing that the modes are highly localized in space. The frequency of the highest excited localized mode at the BZ edge is !mag= 213:12 GHzwith an estimated linewidth (FWHM) of mag=
!mag= 2131:2 kHz whereweusedtherealGilbertdampingofYIG
= 10 5. Note that the simulated linewidth shown in Fig. 4c is larger due to the different choice of the Gilbert damping in order to speed up the simulation. As we see from its mode shape, this mode is nicely localized at the holes attached to the defect and is odd with respect to x= 0andy= 0, and hence has the same symmetry as the optical spin density as we aimed for (see Fig. 4c). V. OPTOMAGNONIC CRYSTAL As shown above, the crystal in Fig. 1a can host both optical and magnetic modes and therefore can be consid- ered anoptomagnonic crystal . In this section, we eval- uate the optomagnonic coupling GFgiven in Eq. (14) (GCis briefly discussed at the end of the section) for the modes found in Secs. III and IV shown in Fig. 4. Numerically evaluating Eq. (14) gives a Faraday con- tribution to the optomagnonic coupling per magnon and per photon ofjGF numj= 20:5 kHz(spatial shape of the coupling see Fig. 6a). In order to gauge this value we want to compare it to the analytical estimate derived in [39]. In the optimal case, the magnetic mode volume and the optical mode volume coincide, VmagVopt. In9 this case, we estimate the coupling as jGF optimalj=Fn 2!optrgB Ms1p Vmag;(24) which evaluates to jGF optimalj= 20:6 MHzusing the material parameters of YIG ( (Fn)=(2) = 4 10 5; Ms= 140 kA=m) and the optical frequency found in Sec. III, !opt= 2249 THz. The magnetic mode volume is defined as the one where the magnon intensity is above a certain threshold, giving Vmag= 2:810 2µm3 (see appendix D). The coupling is bounded by the magnon mode volume, since in the investigated structure itissmallerthantheopticalmodevolume(seeFig.4). In order to take the mismatch in the mode volume into ac- count, we introduce the following overlap measure which is also known as filling factor O=Voverlap Vopt; (25) whereVoverlaprepresents the volume where the magnon and photon modes overlap. The volumes are estimated similar to the case of magnons to be Voverlap = 9:7 10 3µm3andVopt= 0:7µm3(see appendix D). Note that for the optical volume it was taken into account that the mode leaks out of YIG into the Si 3N4layer and air, which is not shown in Fig. 4 (a)+(b). Thus the over- lap measure evaluates to O= 0:01, shrinking the opti- mal coupling toOjGF optimalj26 kHz. Hence, even though the optomagnonic crystal localizes both modes in the same region, the overlap measure is rather small due to the much larger optical mode volume (see Fig. 4 and Fig 6(a)), which is detrimental for the coupling strength. Furthermore by looking at the fine structure of the op- tical spin density and the magnon mode we see that the amplitude peaks of both do not coincide (see Fig. 7): the magnonic peaks are localized nearer to the center than the optical ones. This results in a smaller overlap volume which would be Vmagif the peaks of the modes would be at the same position. Sincethecouplingalsostronglydependsontherelative direction between the vectors of the modes, we addition- ally introduce a ‘directionality’ measure D= drm(r)[E(r)E(r)] drjm(r)jjE(r)E(r)j(26) evaluating toD= 51%using the numerical results pre- sented above. As we see, although the symmetries of the optical spin density and the magnon mode match, the vectors of the modes do not perfectly align in the defect area (see Figs. 4). Taking also this sub-optimal alignment into account the coupling estimate reduces to jGF expectedj=ODjGF optimalj= 23 kHzwhichcoincides well with the numerically obtained value. We conclude that the coupling in the investigated structure is mostly Fine structure |δm|[E*×E]0L01Normalized strengthLengthFigure 7. Fine structure of the optical spin density and the magnon mode along the length of the crystal for a fixed height and width. affected by the large difference between the optical and the magnetic mode volumes, shrinking the coupling value by two orders of magnitude. We remind the reader that the obtained values are for a proof of principle struc- ture which has been only partially optimized, since we started from a fixed photonic crystal structure. In the next section we discuss a possible optimization from the magnonics side. We now proceed to briefly discuss the Cotton-Moutton effect for the results found in Secs. III and IV. For YIG, the Cotton-Moutton coefficient (Cn)=(2) = 2 10 5[41] is of the same order of magnitude as the corre- sponding Faraday coefficient, determined by F. Since in the Voigt configuration both effects are of leading order in the magnetization fluctuations (see Eqs. (14) and (15)), it is important to take its contribution into account. Moreover, since the coefficients GFandGCare complex, it is difficult to estimate a priori the total cou- plingjGF+GCj, duetotheunknownpossibleinterference effects. Numerically evaluating Eq. (15) gives an interac- tion value ofjGC numj= 21:6 kHz. This large value can be explained by the symmetry of the integrand which re- duces tomx 0[ExE ymy+E xEymy]due to the backward volume wave setup and the TE-like character of the opti- calmode. Thisintegrandisfullyevensince ExE yhasthe same symmetry as my. The full optomagnonic coupling jGnumj=GF num+GC num (27) is found to be Gnum= 21:3 kHz. Compared to the optomechanical coupling in simi- lar 1D crystals, where coupling values (per photon and phonon) up to 2950 kHzcan be obtained [53–58], the optomagnoniccouplingobtainedhereisstillrathersmall. However, this is large compared to other optomagnonic systems. As we argued above, the coupling is limited10 Height dependence of the coupling|G|𝓓304050607080902.02.22.42.62.83.0 49.049.550.050.551.051.552.0|Optomagnonic coupling|/2𝛑 [kHz]Overlap [%] Height [nm] Figure 8. Height dependence of the Faraday component of the optomagnonic coupling: The coupling shows ap Vmag dependence since the optical mode volume in the YIG and the Si 3N4slab is constant. The decrease with larger height can be explained by the shrinking directionality measure (see Eq. 26) between the optical and the magnetic mode. by the imperfect spatial matching of magnons and pho- tons with overlap O= 0:01while it is enhanced due to small volumes, Vmag0:01µm3andVopt1µm3. In the standard setups involving spheres [29–31], typically optical volumes are very large 105µm3with low op- tomagnonic overlap 10 3, resulting in low couplings 1 Hz. It was theoretically shown that >75%over- lap in such systems is achievable [129] but the couplings would still be2500 Hz. The miniaturization of an optical cavity to 100µm3was demonstrated in [44], where the coupling is however still small, 250 Hz, in this case due to the large magnon volume involved. Animportantprerequisiteforapplicationsinthequan- tum regime such as magnon cooling, wavelength conver- sion, and coherent state transfer based on optomagnonics isahighcooperativity. Thecooperativityperphotonand magnon is an important figure of merit which compares the strength of the coupling to the lifetime of the coupled modes, and is given by C0=4G2 num
opt mag; (28) where
optis the optical linewidth (FWHM), and mag is the magnonic linewidth (FWHM). To evaluate the theoretical cooperativity of the struc- ture proposed in this manuscript, we use mag=
!mag where
= 10 5is the Gilbert constant and !mag= 213:12 GHz. The optical linewidth is found from sim- ulations to be
opt= 20:2 THz. Using the corresponding parameters the cooperativity per photon and magnon of the optomagnonic crystal is Ccrystal 02:510 10. The single-particle cooperativity can be enhanced by the photon number in the cavity,C=nphC0. Experimentally, there is a bound on the photon density that can be supported by the cavity with- out undesired effects due to heating, and it is empirically given by 5104photons per m3[130]. In our structure, considering the effective mode volume Voptthis gives an enhanced cooperativity at maximum photon density of Ccrystal110 5, whichistwoordersofmagnitudelarger than the current experimental state of the art [44, 130]. Since our model does not account for fabrication im- perfections, this number is expected to be lower in a physical implementation, indicating that optimization is needed. Results for similar 1D optomechanical crys- tals indicate that optimization can lead to larger co- operativity values (at maximum photon density), e.g. 10[54]. The small cooperativity obtained in our struc- ture is a combination of a reduced coupling due to mode- mismatch, plus the very modest quality factor of the op- tical mode in this simple geometry. For boosting the coupling strength we investigate briefly in the following the influence of the optomagnonic crystal’sheightonthecoupling, asproposedinRef.[111]. Therefore we increase the height of the YIG layer from 30 nmto90 nmwithout changing the other parameters of the geometry (including the Si 3N4layer in the optical simulations). As we see from the result (see Fig. 8) the coupling exhibits ap Vmagdependence. We find that the optical mode volume does not change substantially in the modified geometry, and therefore the observed behavior is consistent with the expectedp Vmag=Voptdependence for a constant optical mode volume. The slight decrease forlargerheightscanbeexplainedbytheshrinkingofthe directionality measure D, stemming from the difference in symmetries obeyed by the magnetization (rotational) and the electric field (mirror). VI. OPTIMIZATION So far, we optimized the crystal in order to minimize optical losses for the given geometry. In this section we investigate how to optimize the geometry for magnonics. The optical optimization was achieved by fixing the hole radius and intra-hole distance, which are both along the zxyd2raOptimized crystal Figure 9. Optimization of the geometry: Through increasing the parameters along the width of the crystal we create more spaceforthemodeswithouttouchingtheopticaloptimization of the original crystal (dashed line). We note that we also increased the defect size, not shown here.11 zxyzxy 01 -11 -11 (b)(a)|Ey||Ex||Hz|Mode shape Optical spin density ~ i(E* x E)zBand diagramSymmetriesRe[Ex]Re[Ey]Extended states 0𝞹/akxfno defectDiscrete states1. band2. band ~ 230 Hz~ 310 HzDefect state07 x 10-3kx [1/nm] -11 01 (c) Re[δmz]Im[δmz]Mode profileBand diagram|δmz| SymmetriesLocalized mode @ 13.17 GHz016f [GHz]FT[mz]normMode spectrumkx = 𝞹/aδmz kx [1/nm]f [GHz]016𝛅mzdefect07 x 10-3 0𝞹/akxzxy Figure 10. Optical (a and b) and magnetic modes (c) of the optimized crystal: (a) Band diagram for TE-like modes within the irreducible BZ with a defect mode in the photonic band gap which was pulled from the upper band-edge state into the gap by the insertion of a defect. From the mode shape of the localized mode with a frequency of !opt=2= 279 THz (middle layer in thexy-plane) we see that this mode is odd with respect to x= 0andy= 0(and even with respect to (z=0)). (b) Optical spin density of the localized mode (middle layer in the xy-plane) which is odd with respect to x= 0andy= 0(and even with respect to z= 0). (c) Band diagram of backward volume waves within the irreducible BZ showing magnetic modes with extended k-values but preferring wave vectors at the edge of the BZ. The highest excited localized mode has a frequency of !mag= 213:17 GHzand is odd along the mirror symmetry planes for x= 0andy= 0(and additionally even with respect to the plane for z= 0). The dashed line in the middle inset shows the mode spectrum in case of no defect. Note: all mode shape plots are normalized to their corresponding maximum value. length of the crystal. In the following we tune instead only the parameters along the width of the crystal ( ^y- direction), in order to perturb as little as possible the optical optimization. We found a promising structure by increasing the width of the crystal and considering ellip- tical holes, see Fig. 9. From a set of trials, we found that a width of w= 900 nm and a radius of the holes along the width of rw= 380 nm give the highest coupling. An increased defect size of d= 1201:5 nmis also beneficial for decreasing the optical losses in this case, it nicely lo- calizes the optical defect mode in the middle of the band gap and thus does not drastically change the localization behavior of the photonic crystal. For evaluating the photonic band structure and the optical modes we use the same procedure as described in Sec. III. We obtain a similar band structure for TE- like modes and also a similar localized mode with a frequency of !opt= 2279 THz (obtained by Com- sol,235 THz according to MEEP) and a damping of opt= 23 THzwhich gives an optical linewidth (FWHM) of
opt= 26 THz(see Fig. 10(a)). Us- ing the values obtained by Comsol this results in a re- duced optical quality factor of Q= 93(note that MEEP gives a twice as large value). This rather low optical quality factor is a trade off for the magnetic optimiza- tion achieved by elliptical holes. Moreover, the optical spin density compared to the original crystal is mostly localized within the defect which is advantageous for our purposes (see Fig. 10(b)). Similarly, for evaluating the magnon modes we used the parameters and procedurespresented in Sec. IV. In the following we focus on the Faraday part of the optomagnonic coupling and there- fore consider only the mzcomponent of the magnon modeduetothestructureoftheopticalspindensity. The Cotton-Mouton term is discussed briefly at the end of the section. The simulated band diagram for backward vol- ume waves again shows extended magnon modes but in thiscaseweobtainonebroadband, mostlikelystemming from a fusion of several bands due to the larger width of the crystal (see Fig. 10(c)). The frequency of the highest excited localized mode is !mag= 213:17 GHzwith an estimated linewidth of
!mag= 2131:7 kHzwhere we used the Gilbert damping of YIG
= 10 5. As in the previouscase, thesimulatedlinewidthislargerduetothe larger Gilbert damping used in the simulations. As we see from its mode shape, this mode is nicely localized at the holes attached to the defect and has approximately the same shape and symmetry as the optical spin density (see Fig. 10(c)). Using the results discussed above, the Faraday com- ponent of the optomagnonic coupling of Eq. 14 for the optimized crystal evaluates to jGF numj= 22:9 kHz. Therefore the optimized coupling is one order of mag- nitude larger than in the crystal discussed in Sec. V. As before we want to gauge this value by comparing it to the analyticalestimategiveninEq.24. Theoptimalcoupling in the optimized crystal is jGF optimalj= 20:5 MHz. Again the magnetic mode volume bounds the coupling due to the smaller size of the magnetic mode compared to the optical mode which also extends to the Si 3N4lay-12 ers. This results in a overlap measure (see Eq. 25) of O= 0:04. Therefore the mode overlap is increased by 25%compared to the un-optimized crystal. Evaluating the directionality measure given in Eq. 26 gives D= 53% whichisjustslightlylargerthanintheun-optimizedcase. Taking both measures into account the analytical cou- pling estimate shrinks to jGF expectedj=ODjGF optimalj 210 kHzwhich lies slightly above the numerically ob- tained value. Although the fine structure peaks of the optical spin density and the magnon mode still do not coincide (see Fig. 11), the coupling values are improved by “pulling" the optical and magnetic modes completely into the defect area by the insertion of elliptical holes, creating an overlap area with high density of both modes. The Cotton-Moutton effect in this structure evaluates tojGC numj= 21 kHzand results in a total coupling of jGnumj=jGF num+GC numj= 22 kHz. We can conclude that in this case both effects interfere constructively for the total coupling. The cooperativity per photon and magnon in this case isCop 0210 11, which can be enhanced to Cop0:510 6by the number of photons trapped in the cavity. Thus the cooperativity at maximum photon density is slightly lower as in the crystal presented above, a consequence of the reduced quality factor of the optical mode. VII. CONCLUSION We proposed an optomagnonic crystal consisting of a one-dimensional array with an abrupt defect. We showed that this structure acts as a Bragg mirror both for pho- ton and magnon modes, leading to co-localization of the modes at the defect. By proper design and taking into account the required symmetries of the modes in order to optimize the coupling, we showed that coupling values in the kHz range are possible in these structures. This value is orders of magnitude larger than the experimental state of the art in the field, but still rather small compared to the theoretically predicted optimal value for micron sized structures, which is in the range of 10 1MHz[34]. We showed that the strength of the coupling in our proposed structure is still limited largely by the sub- optimal mode overlap, <5%. Further optimization in design is moreover needed in order to boost the coop- erativity value, which is limited mainly by the optical losses. The simultaneous optimization is challenging, due to the complexity of the demagnetization fields in patterned geometries. Whereas it is well known that a tapered defect (that is, a smooth defect) can highly in- crease the optical quality factors, its effect on the mag- netic modes is non-trivial and is disadvantageous for localizing the magnon modes of the kind used in this work. Other magnon modes, however, could be explored in this case. More complex geometries, including one- Fine structure |δm|[E*×E] 0L01Normalized strengthLengthFigure 11. Fine structure of the optical spin density and the magnon mode along the length of the optimized crystal for a fixed height and width. dimensional crystals combining tapering and an abrupt defect, or two-dimensional crystals, are good candidates to be explored in order to improve quality factors and coupling. The first results shown in this work point to the promise of designing the collective excitations in op- tomagnonic systems via geometry, in order to boost the coupling strength and minimize losses, paving the way for applications in the quantum regime. VIII. ACKNOWLEDGEMENTS We thank Clinton Potts and Tahereh Parvini for in- sightful discussions. J.G. acknowledges financial support from the International Max Planck Research School - Physics of Light (IMPRS-PL). H.H. acknowledges fund- ing from the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excel- lence Strategy EXC-2111-390814868. S.S. and S.V.K. acknowledge funding from the Max Planck Society through an Independent Max Planck Research Group. S.V.K also acknowledges funding from the Deutsche Forschungsgemeinschaft (DFG, German Research Foun- dation) through Project-ID 429529648 – TRR 306 QuCoLiMa (“Quantum Cooperativity of Light and Mat- ter"). APPENDIX A: NORMALIZATION OF MAGNON MODES In this section, we discuss the normalization of magnonsoverageneralmagnetizationtexture. Themag- netization satisfies the Landau-Lifshitz (LL) equation dM dt= gB0 ~M(Hex+He[M]);(29)13 where Hexis an external field and Heis a linear func- tional which can be interpreted as the effective field gen- erated by spin-spin interactions such as exchange, dipo- lar, etc. Let the static solution, i.e. putting dM=dt= 0, beMsm0(r)with saturation magnetization Msand unit vector m0m0= 1. This magnetization generates an effective field of the form He[Msm0(r)] =H0(r)m0(r) Hex(r);(30) where the function H0(r)depends on the nature of spin- spin interactions. The magnon modes m
(r)e i!
tare found by the linearized LL equation i!
m
=gB0 ~[m0h
+H0m
m0];(31) whereh
=He[m
]. The LL equation can be derived from the Hamiltonian H= 0 dVMHe[M] 2+HexM :(32) Up to quadratic terms in m, we expand the magnetiza- tion MMs 1 mm 2 m0+Msm;(33) and the effective field He[M] 1 mm 2 (H0m0 Hex) +h;(34) where A=X
A
+A
; (35) withAbeing morhand
being magnon amplitudes, i.e. classical counterpart of ^b
defined in Eq. (10) in the main text. The above form ensures MM=M2 sup to second order in m. The Hamiltonian becomes (ignoring a constant term) H= 0Ms 2 dV H0mm+mh +mHex+m0h :(36) The last two terms are linear in m
and thus should be zero for m
to correspond to a magnon mode. We can simplify the second term by finding the component ofh
perpendicular to m0using Eq. (31), h
m0(m0h
) =H0m
i~!
gB0m0m
:(37) Usingthisandignoringthelinearterms, theHamiltonian simplifies to H=0Ms 2X
i~! gB0 dVm0h m m m
+m
i : (38)As the eigenmodes should diagonalize the Hamiltonian toP~!
j
j2, we should have dVm0(m
m) = 0; (39) and iMs dVm0 mm
=gB
:(40) For=
, this gives the normalization for magnons. For circularlypolarizedmagnonswith m=m(y+iz)=p 2 andm0=x, the normalization becomes dVm2=gB Ms: (41) APPENDIX B: NUMERICAL SETTINGS - OPTICAL SIMULATIONS In the following we shortly discuss how the optical band structure and the optical modes can be evaluated numerically. In our work we used two different computa- tional methods: for calculations done in the time domain we use the electromagnetic simulation tool MEEP [110], whereas for calculations done in the frequency domain we use the finite element solver Comsol [113]. We use two simulation tools since with MEEP it is much easier to ob- tain the band structure of the crystal, and with Comsol the exact mode shape. 1. MEEP MEEP in general solves for Maxwell’s equations in the time domain within some finite composite volume. Therefore it essentially performs a kind of numerical ex- periment [110]. We use MEEP for simulating the band structure of a YIG crystal without defect in order to find its band gap and its corresponding mid-gap frequency. Furthermore we use a transmission spectrum simulation to optimize the defect size in order to get the least lossy localized mode. For finding the exact frequency of the localized mode we also simulate its spatial shape in the time domain. For simplicity we simulate the YIG crystal in a 2D model (see [47, 112]). Its material parameters are set by the relative permittivity of "= 5. In order to account for the leakage of the electromagnetic field the investigated crystal is surronded by a finite size air region large enough so that the leaking electric field de- cays before it reaches the boundaries, in order to avoid spurious reflection effects. This is achieved by choosing the distance between the surface of the crystal and the boundary of the air region as dair= 30where0is the vacuum wavelength ( dair= 33:6µmin our case). Fur- thermore, we need boundary conditions along the out- side of the air region that are transparent to the leaking14 field such that the truncated air region represents a rea- sonable approximation of free space. Therefore we use a perfectly matched layer at the boundaries of the air region which absorbs all outgoing waves. The thickness of this layer should be at least a vacuum wave length [131]. The whole geometry is meshed by one single reso- lution parameter which discretizes the structure in time and space and gives the number of pixels per distance unit. For all band simulations we used a resolution of 40 pixels, whereas in case of the transmission spectrum we used a resolution of 20pixels and a resolution of 50pixels in case of the mode shapes [112]. Band structure simulations For obtaining the band structure we use a YIG crystal without defect ( d=a) and therefore we can simulate only one single unit cell with a side length of acontain- ing one air hole, and apply an infinite repetition of this cell at each side in ^x-direction. Since we expect the mid- gap frequency of the crystal with defect to be around 240 THz, we excite the crystal with a gaussian pulse with a center-frequency of 225 THz and a width of 450 THz to cover all modes around the band gap. We center the pulse peak at an arbitrary postion (x= 0:00123;y= 0) in order to couple the pulse to an arbitrary mode. Since we want to simulate only TE-like modes, in order to save computational time the pulse only has a Hzcomponent. For decreasing the computation time even more, we ap- ply an odd mirror symmetry plane for y= 0. The mirror symmetry for x= 0is broken by a boundary condition for0<kx<[112]. Transmission spectrum simulations For optimizing the defect size we simulate a transmission spectrum for frequencies at the band gap by measuring the flux at the end of the waveguide stemming from a source at the other end. The measured flux then is nor- malized to the flux of a waveguide without holes. We therefore simulate the transmission spectrum as a func- tion of different defect sizes and use the defect size which gives the highest transmission. In order to consider only TE-like modes where the electric field lies in plane we need to excite the system with a Jy-current source trans- verse to the propagation direction which is achieved by a gaussian pulse with only a Ey-component. Its cen- ter frequency thereby is 222 THz (simulated mid-gap fre- quency) and its width is 90 THz(>band width). Also in this case we apply an odd mirror symmetry for y= 0for decreasing the simulation time. We note that the mirror symmetry for x= 0is broken by the source since it is located at the edge of the waveguide [112]. Mode shape simulations For evaluating the mode frequency of the localized mode within the band gap we simulate the time evolution of this single mode by exciting it by a gaussian pulse with a center frequency of 203 THz (frequency of the peak in thetransmission spectrum) and a width of 15 THz. Since in this simulation no symmetry is broken we also apply an odd mirror symmetry for x= 0andy= 0for obtaining only a TE-like mode [112]. 2. Comsol We use Comsol to find the spatial mode shape. There- fore we use the “Electromagnetic waves, Frequency do- main" package of COMSOL’s “RF module" which solves for the Helmholtz equation of the form r1 r(rE) k2 0 "r i !" E= 0;(42) wherek0indicates the vacuum wave number, !the an- gular frequency, rthe relative permeability and "0the vacuum permittivity. Contrary to the MEEP simula- tions above, we simulate the full geometry composite of a YIG layer sandwiched by two Si 3N4layers. The used material parameters thereby are "YIG= 5,"Si= 4, YIG=Si= 0, andYIG=Si= 0withthe relative permeabilty and the conductivity. Again we also need to simulate an truncated air region around the crystal which is able to absorb the outgoing radiation. The cor- responding material parameters are "air=air= 1and air= 0. Besides perfectly matched layers we also can use second order scattering boundary conditions at the air surfaces given by the expression [131] nrEz+ik0Ez i 2k0r2 tEz= 0 (43) withnthe normal vector to the considered plane. For large enough air regions both approaches are almost equivalent as long as the leaking field is propagating normal to the air surfaces. In order to account for a large enough air region we choose the distance between the surfaces of the crystal and the air boundaries as 4:5µm. For reducing the simulation time we use the symmetry requirements of a TE-like mode. Therefore, we cut the geometry into an eighth of the whole struc- ture and apply perfect electric conductor boundary con- ditions ( nE= 0) at the cut surfaces along x= 0and y= 0and a perfect magnetic conductor boundary condi- tion (nH= 0) at the cut surface along z= 0. The full solution is then obtained by using the symmetry require- ments of a TE-like mode. The whole geometry is meshed by a physics-controlled tetrahedral mesh with a maxi- mum element size of 0=50:3µm[132]. We note that in case of a physics-controlled mesh Comsol automati- cally meshes the material areas of interest with a finer mesh and uses a coarser mesh e.g. for the air regions.15 APPENDIX C: NUMERICAL SETTINGS - MAGNETIC SIMULATIONS In this section we briefly discuss how the magnetic band structure and magnetic mode shape is obtained nu- merically. For evaluating the magnetization dynamics we use the finite difference tool MuMax3 [124] which solves for the Landau-Lifshitz-Gilbert equation of the form @m @t=g1 1 +2[mBeff+(m(mBeff))](44) withm=M=Msthe local reduced magnetization of one simulation cell, gthe gyromagnetic ratio, the damp- ing parameter, and Beffan effective field which contri- butions can be found in [124]. As material parameters we used the parameters for YIG, Ms= 140 kA=m(sat- uration magnetization), Aex= 2 pJ=m(exchange con- stant),Kc1= 610 J=m3(anisotropy constant) with the anisotropy axis along ^z. In order to accelerate the simu- lations, weusedanincreasedGilbertDampingparameter = 0:008(compare to 10 5for YIG) [126]. The used meshgrid had (1024;50;5)cells in the (^x;^y;^z)- direction what guarantees to take the exchange interac- tion into account ( lex13 nmfor YIG). In general, in all our simulations the spin wave dy- namics is excited via an external pulse field and the time evolution is recorded for all three magnetization com- ponents. For post processing the output of the form mi(x;y;z )withi= (x;y;z )saved for all simulated time steps separately we create for each magnetization com- ponenti= (x;y;z )a 4D-array of the form mi(t;x;y;z ). Band structure simulations In order to obtain the band structure along a spe- cific direction jwithj= (x;y;z )e.g. chosen to be the^x-direction, we reduce the four dimensional array to a two dimensional array of the form mi(t;x) =Pny mPnz nmi(t;x;ym;zn)and perform a 2D Fourier transform on this array mi(f;kx) =FT2D[mi(t;x)]re- sulting in the band diagram along the chosen direction. For increasing the resolution in the band diagram we plot the quantityp jmi(t;x)j=max (jmi(t;x)j). Mode shape simulations In order to obtain the mode shape we perform a space- dependent Fourier transform in time on each array entry separatelymi(f;x;y;z ) =FT1D[mi(t;x;y;z )]. APPENDIX D: EVALUATION OF THE MODE VOLUMES For evaluating the mode volume numerically we first, due to numerical errors, need to identify all cells of the simulated array (either containing m(r)in case of the magnetic mode volume or E(r)in case of the opticalmode volume) which contribute to the volume by a high enough mode density. This means we need to define a threshold which determines if a cell should contribute to the mode volume or not. We define this threshold by T=max(jxj) min(jxj) 5; (45) where xis the content of the cell (either x=morx= E). For being able to count the cells which contribute to the volume we create an additional array matching the simulated arrays in size. This array then contains ones if the absolute value of the cell ( jxj) in the original array is larger than the threshold given in Eq. (45) and zeros ifjxjis smaller. The mode volume is then obtained by summing over this array (giving the number of cells contributing to the volume) and by multiplying this such calculated number by the volume of one cell sxsysz. For evaluating the overlap volume between the mag- netic mode ( x=m) and in this case the optical spin density ( EE) we create the same additional arrays as above identifying the cells which contribute to their cor- responding mode volume. For identifying the cells where the magnon mode and the optical spin density overlap we create a third array again matching the size of the original array. But this array now contains ones if the corresponding cells of the “threshold arrays" both in the magnetic and optical case contain a one, otherwise we set the cell value to zero. The overlap volume is then obtained by summing over this third array (giving the number of cells contributing to the overlap) and by mul- tiplying this such calculated number by the volume of one cellsxsysz. [1] G. Kurizki, P. Bertet, Y. Kubo, K. Mølmer, D. Pet- rosyan, P. Rabl, and J. Schmiedmayer, Quantum tech- nologies with hybrid systems, Proceedings of the Na- tional Academy of Sciences 112, 3866 (2015). [2] M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, Cavity optomechanics, Rev. Mod. Phys. 86, 1391 (2014). [3] M. Metcalfe, Applications of cavity optomechanics, Ap- plied Physics Reviews 1, 031105 (2014). [4] Y.-H. Zhou, Q.-S. Tan, X.-M. Fang, J.-F. Huang, and J.-Q. Liao, Spectrometric detection of weak forces in cavity optomechanics, Opt. Express 28, 28620 (2020). [5] P. Weber, J. Güttinger, A. Noury, J. Vergara-Cruz, and A.Bachtold,Forcesensitivityofmultilayergrapheneop- tomechanical devices, Nat. Commun. 7, 12496 (2016). [6] S. Schreppler, N. Spethmann, N. Brahms, T. Botter, M. Barrios, and D. M. Stamper-Kurn, Optically mea- suring force near the standard quantum limit, Science 344, 1486 (2014). [7] M. Toroš and T. S. Monteiro, Quantum sensing and cooling in three-dimensional levitated cavity optome- chanics, Phys. Rev. Research 2, 023228 (2020). [8] S. Ghosh, D. Carney, P. Shawhan, and J. M. Taylor,16 Backaction-evading impulse measurement with mechan- ical quantum sensors, Phys. Rev. A 102, 023525 (2020). [9] P. J. de Groot, A review of selected topics in interfero- metricopticalmetrology,ReportsonProgressinPhysics 82, 056101 (2019). [10] T. Liu, F. Pagliano, R. van Veldhoven, V. Pogoretskiy, Y. Jiao, and A. Fiore, Integrated nano-optomechanical displacement sensor with ultrawide optical bandwidth, Nat. Commun. 11, 2407 (2020). [11] V. Bergholm, W. Wieczorek, T. Schulte-Herbrüggen, and M. Keyl, Optimal control of hybrid optomechanical systems for generating non-classical states of mechani- cal motion, Quantum Science and Technology 4, 034001 (2019). [12] H. Xu, L. Jiang, A. A. Clerk, and J. G. E. Harris, Non- reciprocal control and cooling of phonon modes in an optomechanical system, Nature 568, 65 (2019). [13] G. A. Brawley, M. R. Vanner, P. E. Larsen, S. Schmid, A. Boisen, and W. P. Bowen, Nonlinear optomechanical measurement of mechanical motion, Nat. Commun. 7, 10988 (2016). [14] B. Q. Baragiola and J. Twamley, Generating nonclas- sical states of motion using spontaneous emission, New Journal of Physics 20, 073029 (2018). [15] Y. Yan, J. pei Zhu, and G. xiang Li, Preparation of a nonlinear coherent state of the mechanical resonator in an optomechanical microcavity, Opt. Express 24, 13590 (2016). [16] F. Pulizzi, Spintronics, Nature Matter 11, 367 (2012). [17] A. Hirohata, K. Yamada, Y. Nakatani, I.-L. Prejbeanu, B. Diény, P. Pirro, and B. Hillebrands, Review on spin- tronics: Principles and device applications, Journal of Magnetism and Magnetic Materials 509, 166711 (2020). [18] T. Wimmer, M. Althammer, L. Liensberger, N. Vliet- stra, S. Geprägs, M. Weiler, R. Gross, and H. Huebl, Spin transport in a magnetic insulator with zero effec- tive damping, Phys. Rev. Lett. 123, 257201 (2019). [19] Y. Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Yamazaki, K. Usami, and Y. Nakamura, Quantum magnonics: The magnon meets the superconducting qubit, Comptes Rendus Physique 17, 729 (2016). [20] H. Huebl, C. W. Zollitsch, J. Lotze, F. Hocke, M. Greifenstein, A. Marx, R. Gross, and S. T. B. Goen- nenwein, High cooperativity in coupled microwave res- onator ferrimagnetic insulator hybrids, Phys. Rev. Lett. 111, 127003 (2013). [21] A. A. Clerk, K. W. Lehnert, P. Bertet, J. R. Petta, and Y. Nakamura, Hybrid quantum systems with cir- cuit quantum electrodynamics, Nature Physics 16, 257 (2020). [22] Y. Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Ya- mazaki, K.Usami,andY.Nakamura,Coherentcoupling betweenaferromagneticmagnonandasuperconducting qubit, Science 349, 405 (2015). [23] D.Lachance-Quirion,Y.Tabuchi,S.Ishino,A.Noguchi, T. Ishikawa, R. Yamazaki, and Y. Nakamura, Resolving quantaofcollectivespinexcitationsinamillimeter-sized ferromagnet, Science Advances 3, e1603150 (2017). [24] D.Lachance-Quirion, S.P.Wolski, Y.Tabuchi, S.Kono, K. Usami, and Y. Nakamura, Entanglement-based single-shot detection of a single magnon with a super- conducting qubit, Science 367, 425 (2020). [25] S. Sharma, V. A. S. V. Bittencourt, A. D. Karenowska, and S. V. Kusminskiy, Spin cat states in ferromagneticinsulators, Phys. Rev. B 103, L100403 (2021). [26] C. A. Potts, V. A. S. V. Bittencourt, S. V. Kusminskiy, and J. P. Davis, Magnon-phonon quantum correlation thermometry, Phys. Rev. Applied 13, 064001 (2020). [27] R. Hisatomi, A. Osada, Y. Tabuchi, T. Ishikawa, A. Noguchi, R. Yamazaki, K. Usami, and Y. Nakamura, Bidirectional conversion between microwave and light via ferromagnetic magnons, Phys. Rev. B 93, 174427 (2016). [28] D.Lachance-Quirion, Y.Tabuchi, A.Gloppe, K.Usami, and Y. Nakamura, Hybrid quantum systems based on magnonics, Applied Physics Express 12, 070101 (2019). [29] J. A. Haigh, A. Nunnenkamp, A. J. Ramsay, and A. J. Ferguson, Triple-resonant brillouin light scattering in magneto-optical cavities, Phys. Rev. Lett. 117, 133602 (2016). [30] A. Osada, R. Hisatomi, A. Noguchi, Y. Tabuchi, R. Ya- mazaki, K. Usami, M. Sadgrove, R. Yalla, M. Nomura, and Y. Nakamura, Cavity optomagnonics with spin- orbit coupled photons, Phys. Rev. Lett. 116, 223601 (2016). [31] X. Zhang, N. Zhu, C.-L. Zou, and H. X. Tang, Op- tomagnonic whispering gallery microresonators, Phys. Rev. Lett. 117, 123605 (2016). [32] A. Osada, A. Gloppe, R. Hisatomi, A. Noguchi, R. Ya- mazaki, M. Nomura, Y. Nakamura, and K. Usami, Bril- louinlightscatteringbymagneticquasivorticesincavity optomagnonics, Phys. Rev. Lett. 120, 133602 (2018). [33] J. A. Haigh, N. J. Lambert, S. Sharma, Y. M. Blanter, G. E. W. Bauer, and A. J. Ramsay, Selection rules for cavity-enhanced brillouin light scattering from magne- tostatic modes, Phys. Rev. B 97, 214423 (2018). [34] S. Viola Kusminskiy, H. X. Tang, and F. Marquardt, Coupled spin-light dynamics in cavity optomagnonics, Phys. Rev. A 94, 033821 (2016). [35] T. Liu, X. Zhang, H. X. Tang, and M. E. Flatté, Optomagnonics in magnetic solids, Phys. Rev. B 94, 060405(R) (2016). [36] S. Sharma, Y. M. Blanter, and G. E. W. Bauer, Light scattering by magnons in whispering gallery mode cav- ities, Phys. Rev. B 96, 094412 (2017). [37] P. A. Pantazopoulos, N. Stefanou, E. Almpanis, and N. Papanikolaou, Photomagnonic nanocavities for strong light–spin-wave interaction, Phys. Rev. B 96, 104425 (2017). [38] A. Osada, A. Gloppe, Y. Nakamura, and K. Usami, Or- bital angular momentum conservation in brillouin light scattering within a ferromagnetic sphere, New Journal of Physics 20, 103018 (2018). [39] S. V. Kusminskiy, Cavity optomagnoics, arXiv:1911.11104 (2019). [40] V. A. S. V. Bittencourt, V. Feulner, and S. V. Kusmin- skiy, Magnon heralding in cavity optomagnonics, Phys. Rev. A 100, 013810 (2019). [41] D. D. Stancil and A. Prabhakar, Spin waves , 1st ed. (Springer, 2009). [42] A. Antonov, A. Agranovskaya, G. Petrova, and A. Titova, Optical properties of yttrium iron garnet, Journal of Applied Spectroscopy 11, 1225 (1969). [43] S. H. Wemple, S. L. Blank, J. A. Seman, and W. A. Biolsi, Optical properties of epitaxial iron garnet thin films, Phys. Rev. B 9, 2134 (1974). [44] J. Haigh, R. Chakalov, and A. Ramsay, Subpicoliter magnetoptical cavities, Phys. Rev. Applied 14, 04400517 (2020). [45] E. Almpanis, G. P. Zouros, P. A. Pantazopoulos, K. L. Tsakmakidis, N. Papanikolaou, and N. Stefanou, Spher- ical optomagnonic microresonators: Triple-resonant photon transitions between zeeman-split mie modes, Phys. Rev. B 101, 054412 (2020). [46] P.-A.Pantazopoulos,N.Papanikolaou,andN.Stefanou, Tailoring coupling between light and spin waves with dual photonic–magnonic resonant layered structures, Journal of Optics 21, 015603 (2018). [47] J. D. Joannopoulos, S. G. Johnson, J. N. Winn, and R. D. Meade, Photonic Crystals - Molding the Flow of Light,2nded.,Vol.8(PrincetonUniversityPress,2008). [48] S. O. Demokritov and A. N. Slavin, Magnonics - From Fundamentals to Applications , 1st ed., Vol. 125 (Springer-Verlag Berlin Heidelberg, 2013). [49] M. Krawczyk and D. Grundler, Review and prospects of magnonic crystals and devices with reprogrammable band structure, Journal of Physics: Condensed Matter 26, 123202 (2014). [50] B. Lenk, H. Ulrichs, F. Garbs, and M. Münzenberg, The building blocks of magnonics, Physics Reports 507, 107 (2011). [51] M.Eichenfield, J. Chan, R. M. Camacho, K. J. Vahala, and O. Painter, Optomechanical crystals, Nature 462, 78 (2009). [52] M. Eichenfield, J. Chan, A. H. Safavi-Naeini, K. J. Va- hala, and O. Painter, Modeling dispersive coupling and losses of localized optical and mechanical modes in op- tomechanical crystals, Opt. Express 17, 20078 (2009). [53] J. Chan, T. P. M. Alegre, A. H. Safavi-Naeini, J. T. Hill, A. Krause, S. Gröblacher, M. Aspelmeyer, and O. Painter, Laser cooling of a nanomechanical oscillator into itsquantum ground state, Nature 478, 89 (2011). [54] A. H. Safavi-Naeini, T. P. M. Alegre, J. Chan, M. Eichenfield, M. Winger, Q. Lin, J. T. Hill, D. E. Chang, and O. Painter, Electromagnetically induced transparency and slowlight with optomechanics, Nature 472, 69 (2011). [55] J. Chan, A. H. Safavi-Naeini, J. T. Hill, S. Meenehan, and O. Painter, Optimized optomechanical crystal cav- ity with acoustic radiation shield, Applied Physics Let- ters101, 081115 (2012). [56] J. T. Hill, A. H. Safavi-Naeini, J. Chan, and O. Painter, Coherent optical wavelength conversion via cavity op- tomechanics, Nat. Commun. 3, 1196 (2012). [57] A. H. Safavi-Naeini, J. Chan, J. T. Hill, S. Gröblacher, H. Miao, Y. Chen, M. Aspelmeyer, and O. Painter, Laser noise in cavity-optomechanical cooling and ther- mometry, New Journal of Physics 15, 035007 (2013). [58] A. G. Krause, J. T. Hill, M. Ludwig, A. H. Safavi- Naeini, J. Chan, F. Marquardt, and O. Painter, Nonlin- ear radiation pressure dynamics in an optomechanical crystal, Phys. Rev. Lett. 115, 233601 (2015). [59] M. J. Burek, J. D. Cohen, S. M. Meenehan, N. El- Sawah, C. Chia, T. Ruelle, S. Meesala, J. Rochman, H. A. Atikian, M. Markham, D. J. Twitchen, M. D. Lukin, O. Painter, and M. Lončar, Diamond optome- chanical crystals, Optica 3, 1404 (2016). [60] M. H. Matheny, Enhanced photon-phonon coupling via dimerization in one-dimensional optomechanical crys- tals, Applied Physics Letters 112, 253104 (2018). [61] A. H. Safavi-Naeini and O. Painter, Design of optome- chanical cavities and waveguides on a simultaneousbandgap phononic-photonic crystal slab, Opt. Express 18, 14926 (2010). [62] A. H. Safavi-Naeini, T. P. M. Alegre, M. Winger, and O. Painter, Optomechanics in an ultrahigh-q two- dimensional photonic crystal cavity, Applied Physics Letters 97, 181106 (2010). [63] D. E. Chang, A. H. Safavi-Naeini, M. Hafezi, and O. Painter, Slowing and stopping light using an op- tomechanical crystal array, New Journal of Physics 13, 023003 (2011). [64] T. P. M. Alegre, A. Safavi-Naeini, M. Winger, and O. Painter, Quasi-two-dimensional optomechanical crystals with a complete phononic bandgap, Opt. Ex- press19, 5658 (2011). [65] A. H. Safavi-Naeini, J. T. Hill, S. Meenehan, J. Chan, S. Gröblacher, and O. Painter, Two-dimensional phononic-photonic band gap optomechanical crystal cavity, Phys. Rev. Lett. 112, 153603 (2014). [66] H. Ren, M. H. Matheny, G. S. MacCabe, J. Luo, H. Pfeifer, M. Mirhosseini, and O. Painter, Two- dimensional optomechanical crystal cavity with high quantum cooperativity, Nat. Commun. 11, 3373 (2020). [67] M. Eichenfield, R. Camacho, J. Chan, K. J. Vahala, and O. Painter, A picogram- and nanometre-scale photonic- crystaloptomechanical cavity, Nature 459, 550 (2009). [68] J. Chan, M. Eichenfield, R. Camacho, and O. Painter, Optical and mechanical design of a “zipper” photonic crystal optomechanical cavity, Opt. Express 17, 3802 (2009). [69] A. H. Safavi-Naeini, S. Gröblacher, J. T. Hill, J. Chan, M. Aspelmeyer, and O. Painter, Squeezed light from a silicon micromechanical resonator, Nature 500, 185 (2013). [70] S. M. Meenehan, J. D. Cohen, S. Gröblacher, J. T. Hill, A. H. Safavi-Naeini, M. Aspelmeyer, and O. Painter, Silicon optomechanical crystal resonator at millikelvin temperatures, Phys. Rev. A 90, 011803(R) (2014). [71] A. G. Krause, T. D. Blasius, and O. Painter, Optical read out and feedback cooling of a nanostring optome- chanical cavity, arXiv:1506.01249v1 (2015). [72] T. K. Paraïso, M. Kalaee, L. Zang, H. Pfeifer, F. Mar- quardt, and O. Painter, Position-squared coupling in a tunable photonic crystal optomechanical cavity, Phys. Rev. X 5, 041024 (2015). [73] H. Pfeifer, T. Paraïso, L. Zang, and O. Painter, De- sign of tunable ghz-frequency optomechanical crystal resonators, Opt. Express 24, 11407 (2016). [74] G. S. MacCabe, H. Ren, J. Luo, J. D. Cohen, H. Zhou, A. Sipahigil, M. Mirhosseini, and O. Painter, Nano- acoustic resonator with ultralong phonon lifetime, Sci- ence370, 840 (2020). [75] J. Joannopoulos, P. Villeneuve, and S. Fan, Photonic crystals: putting a new twist on light, Nature 386, 143 (1997). [76] S.-Y. Lin, E. Chow, V. Hietala, P. R. Villeneuve, and J. D. Joannopoulos, Experimental demonstration of guiding and bending of electromagnetic waves in a pho- tonic crystal, Science 282, 274 (1998). [77] R. F. Cregan, B. J. Mangan, J. C. Knight, T. A. Birks, P. S. J. Russell, P. J. Roberts, and D. C. Allan, Single- modephotonicbandgapguidanceoflightinair,Science 285, 1537 (1999). [78] M. Lončar, D. Nedeljković, T. Doll, J. Vučković, A. Scherer, and T. P. Pearsall, Waveguiding in pla-18 nar photonic crystals, Applied Physics Letters 77, 1937 (2000). [79] J. C. Knight and P. S. J. Russell, New ways to guide light, Science 296, 276 (2002). [80] J. Joannopoulos, P. R. Villeneuve, and S. Fan, Photonic crystals, Solid State Communications 102, 165 (1997). [81] P. Villeneuve, S. Fan, and J. Joannopoulos, Three- dimensional photon confinement in photonic crystals of low-dimensional periodicity, IEE Proceedings - Opto- electronics 145, 384 (1998). [82] H. Benisty, C. Weisbuch, D. Labilloy, M. Rattier, C. J. M. Smith, T. F. Krauss, R. M. de la Rue, R. Houdre, U. Oesterle, C. Jouanin, and D. Cassagne, Optical and confinement properties of two-dimensional photonic crystals, Journal of Lightwave Technology 17, 2063 (1999). [83] A. Scherer, O. Painter, J. Vuckovic, M. Loncar, and T. Yoshie, Photonic crystals for confining, guiding, and emitting light, IEEE Transactions on Nanotechnology 1, 4 (2002). [84] K. J. Vahala, Optical microcavities, Nature 424, 839 (2003). [85] Y. Akahane, T. Asano, B.-S. Song, and S. Noda, Fine- tuned high-q photonic-crystal nanocavity, Opt. Express 13, 1202 (2005). [86] V. Berger, Nonlinear photonic crystals, Phys. Rev. Lett. 81, 4136 (1998). [87] M. Soljačić, S. G. Johnson, S. Fan, M. Ibanescu, E. Ip- pen, and J. D. Joannopoulos, Photonic-crystal slow- lightenhancementofnonlinearphasesensitivity,J.Opt. Soc. Am. B 19, 2052 (2002). [88] J.-J. Li, Z.-Y. Li, and D.-Z. Zhang, Second harmonic generation in one-dimensional nonlinear photonic crys- tals solved by the transfer matrix method, Phys. Rev. E75, 056606 (2007). [89] N. Segal, S. Keren-Zur, N. Hendler, and T. Ellenbo- gen, Controlling light with metamaterial-based nonlin- ear photonic crystals, Nature Photonics 9, 180 (2015). [90] J.Topp, D.Heitmann, M.P.Kostylev,andD.Grundler, Making a reconfigurable artificial crystal by ordering bistable magnetic nanowires, Phys. Rev. Lett. 104, 207205 (2010). [91] S.-K. Kim, K.-S. Lee, and D.-S. Han, A gigahertz-range spin-wavefiltercomposedofwidth-modulatednanostrip magnonic-crystal waveguides, Applied Physics Letters 95, 082507 (2009). [92] H. Al-Wahsh, A. Akjouj, B. Djafari-Rouhani, and L. Dobrzynski, Magnonic circuits and crystals, Surface Science Reports 66, 29 (2011). [93] X. Xing, S. Li, X. Huang, and Z. Wang, Engineering spin-wave channels in submicrometer magnonic waveg- uides, AIP Advances 3, 032144 (2013). [94] J. W. Kłos, D. Kumar, M. Krawczyk, and A. Barman, Magnonic Band Engineering by Intrinsic and Extrin- sic Mirror Symmetry Breaking in Antidot Spin-Wave Waveguides, Sci. Rep. 3, 2444 (2013). [95] T.Schneider,A.Serga,B.Hillebrands,andM.Kostylev, Spin-wave ferromagnetic film combiner as a not logic gate, Journal of Nanoelectronics and Optoelectronics 3, 69 (2008). [96] T. Schneider, A. A. Serga, B. Leven, B. Hillebrands, R. L. Stamps, and M. P. Kostylev, Realization of spin- wave logic gates, Applied Physics Letters 92, 022505 (2008).[97] A.Khitun, M.Bao,andK.L.Wang,Magnoniclogiccir- cuits, Journal of Physics D: Applied Physics 43, 264005 (2010). [98] A. Khitun, Multi-frequency magnonic logic circuits for parallel data processing, Journal of Applied Physics 111, 054307 (2012). [99] J. D. Jackson, Classical electrodynamics , 3rd ed. (Wiley New York, 1999). [100] W. Wettling, M. G. Cottam, and J. R. Sandercock, The relation between one-magnon light scattering and the complex magneto-optic effects in YIG, Journal of Physics C: Solid State Physics 8, 211 (1975). [101] A. Borovik-Romanov and N. Kreines, Brillouin- mandelstam scattering from thermal and excited magnons, Physics Reports 81, 351 (1982). [102] P. A. Fleury and R. Loudon, Scattering of light by one- and two-magnon excitations, Phys. Rev. 166, 514 (1968). [103] P. A. M. Dirac and N. H. D. Bohr, The quantum theory of the emission and absorption of radiation, Proceedings of the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical Character 114, 243 (1927). [104] M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, Cavity Optomechanics (Springer-Verlag Berlin Heidel- berg, 2014). [105] D. Mills, Quantum theory of spin waves in finite sam- ples, Journal of Magnetism and Magnetic Materials 306, 16 (2006). [106] S. Sharma, Y. M. Blanter, and G. E. W. Bauer, Opti- cal cooling of magnons, Physical Review Letters 121, 087205 (2018). [107] S. V. Kusminskiy, Quantum Magnetism, Spin Waves, and Optical Cavities (Springer, 2019). [108] D. L. Wood and J. P. Remeika, Effect of impurities on the optical properties of yttrium iron garnet, Journal of Applied Physics 38, 1038 (1967). [109] G. B. Scott, D. E. Lacklison, and J. L. Page, Absorption spectra of y3fe5o12(yig) and y3ga5o12:fe3+, Phys. Rev. B10, 971 (1974). [110] A. F. Oskooi, D. Roundy, M. Ibanescu, P. Bermel, J. Joannopoulos, and S. G. Johnson, Meep: A flexible free-softwarepackageforelectromagneticsimulationsby the fdtd method, Computer Physics Communications 181, 687 (2010). [111] J. Graf, H. Pfeifer, F. Marquardt, and S. Viola Kus- minskiy, Cavity optomagnonics with magnetic textures: Coupling a magnetic vortex to light, Phys. Rev. B 98, 241406(R) (2018). [112] /https://meep :readthedocs :io/en/latest/Python_ Tutorials/Resonant_Modes_and_Transmission_in_a_ Waveguide_Cavity/ (accessed November 20, 2020). [113] /https://www :comsol:eu/(accessed November 20, 2020). [114] T.-R. Lin, C.-H. Lin, and J.-C. Hsu, Strong Op- tomechanical Interaction in Hybrid Plasmonic-Photonic Crystal Nanocavities with Surface Acoustic Waves, Sci. Rep.5, 13782 (2015). [115] E. Yablonovitch, Inhibited spontaneous emission in solid-state physics and electronics, Phys. Rev. Lett. 58, 2059 (1987). [116] S. John, Strong localization of photons in certain disor- dered dielectric superlattices, Phys. Rev. Lett. 58, 2486 (1987).19 [117] J. R. Eshbach and R. W. Damon, Surface magnetostatic modes and surface spin waves, Phys. Rev. 118, 1208 (1960). [118] A. A. Serga, A. V. Chumak, and B. Hillebrands, YIG magnonics, Journal of Physics D: Applied Physics 43, 264002 (2010). [119] G. Venkat, D. Kumar, M. Franchin, O. Dmytri- iev, M. Mruczkiewicz, H. Fangohr, A. Barman, M. Krawczyk, and A. Prabhakar, Proposal for a stan- dard micromagnetic problem: Spin wave dispersion in a magnonic waveguide, IEEE Transactions on Magnetics 49, 524 (2013). [120] R. Huber and D. Grundler, Ferromagnetic nanodisks for magnonic crystals and waveguides, Proc SPIE 8100, 81000D (2011). [121] L. Bai, M. Kohda, and J. Nitta, Observation of spin wave modes depending on a tunable periodic magnetic field, Applied Physics Letters 98, 172508 (2011). [122] P. Trempler, R. Dreyer, P. Geyer, C. Hauser, G. Woltersdorf, and G. Schmidt, Integration and char- acterization of micron-sized yig structures with very low gilbert damping on arbitrary substrates, Applied Physics Letters 117, 232401 (2020). [123] G. Schmidt, C. Hauser, P. Trempler, M. Paleschke, and E. T. Papaioannou, Ultra thin films of yttrium iron gar- net with very low damping: A review, physica status solidi (b) 257, 1900644 (2020). [124] A. Vansteenkiste and B. V. de Wiele, MuMax: A new high-performance micromagnetic simulation tool, Jour- nal of Magnetism and Magnetic Materials 323, 2585 (2011). [125] D. Kumar, O. Dmytriiev, S. Ponraj, and A. Bar-man, Numerical calculation of spin wave dispersions in magnetic nanostructures, J. Phys. D: Appl. Phys. 45, 015001 (2011). [126] J. E. Losby, F. F. Sani, D. T. Grandmont, Z. Diao, M. Belov, J. A. J. Burgess, S. R. Compton, W. K. Hiebert, D. Vick, K. Mohammad, E. Salimi, G. E. Bridges, D. J. Thomson, and M. R. Freeman, Torque- mixing magnetic resonance spectroscopy, Science 350, 798 (2015). [127] V. Cherepanov, I. Kolokolov, and V. L’vov, The saga of yig: Spectra, thermodynamics, interaction and relax- ation of magnons in a complex magnet, Physics Reports 229, 81 (1993). [128] S. Klingler, H. Maier-Flaig, C. Dubs, O. Surzhenko, R. Gross, H. Huebl, S. T. B. Goennenwein, and M. Weiler, Gilbert damping of magnetostatic modes in a yttrium iron garnet sphere, Applied Physics Letters 110, 092409 (2017). [129] S.Sharma, B.Z.Rameshti, Y.M.Blanter,andG.E.W. Bauer, Optimal mode matching in cavity optomagnon- ics, Phys. Rev. B 99, 214423 (2019). [130] N. Zhu, X. Zhang, X. Han, C.-L. Zou, C. Zhong, C.-H. Wang, L. Jiang, and H. X. Tang, Waveguide cavity op- tomagnonics for microwave-to-optics conversion, Optica 7, 1291 (2020). [131] /https://www :comsol:de/blogs/using-perfectly- matched-layers-and-scattering-boundary- conditions-for-wave-electromagnetics-problems/ (accessed November 20, 2020). [132] /https://www :comsol:de/blogs/automatic-meshing- for-electromagnetic-simulations/ (accessed November 20, 2020). | 2020-12-01 | We put forward the concept of an optomagnonic crystal: a periodically
patterned structure at the microscale based on a magnetic dielectric, which can
co-localize magnon and photon modes. The co-localization in small volumes can
result in large values of the photon-magnon coupling at the single quanta
level, which opens perspectives for quantum information processing and quantum
conversion schemes with these systems. We study theoretically a simple geometry
consisting of a one-dimensional array of holes with an abrupt defect,
considering the ferrimagnet Yttrium Iron Garnet (YIG) as the basis material. We
show that both magnon and photon modes can be localized at the defect, and use
symmetry arguments to select an optimal pair of modes in order to maximize the
coupling. We show that an optomagnonic coupling in the kHz range is achievable
in this geometry, and discuss possible optimization routes in order to improve
both coupling strengths and optical losses. | Design of an optomagnonic crystal: towards optimal magnon-photon mode matching at the microscale | 2012.00760v2 |
1 Competing effects at Pt/YIG interfaces: spin Hall magnetoresistance, magnon excitations and magnetic frustration Saül Vélez1,*, Amilcar Bedoya -Pinto1,‡, Wenjing Yan1, Luis E. Hueso1,2, and Fèlix Casanova1,2,† 1CIC nanoGUNE, 20018 Donostia -San Sebastian, Basque Country, Spain 2IKERBASQUE, Basque Foundation for Science, 48011 Bilbao, Basque Country, Spain ‡ Present address: Max Planck Institute of Microstructure Physics, D -06120 Halle, Germany * s.velez@nanogune.eu † f.casanova@nanogune.eu We study the spin Hall magnetoresistance (SMR) and the magnon spin transport (MST) in Pt/ Y3Fe5O12(YIG) -based devices with intentionally modified interfaces. Our measurements show that the surface treatme nt of the YIG film results in a slight enhancement of the spin -mixing conductance and an extraordinary increase in the efficiency of the spin -to-magnon excitations at room temperature . The surface of the YIG film develop s a surface magnetic frustration at low temperatures, causing a sign chang e of the SMR and a dramatic suppression of the MST. Our results evidence that SMR and MST could be used to explore magnetic properties of surfaces, including those with complex magnetic textures , and stress the critical importance of the non -magnetic/ferro magnetic interface properties in the performance of the resulting spintronic devices. I. Introduction Insulating spintronics [1] has emerged as a promising, nove l technological platform based o n the integration of ferroma gnetic insulators (FMIs) in devices as a media to generate, process and transport spin information over long distances [1–30]. The advantage of using FMIs against metall ic ones is that the flow of c harge currents is avoided, thus preventing ohmic losses or the emergence of undesired spurious effects. Some phenomena explored in insulating spintronics include the spin pumping [2–5], the spin Hall magnetoresistance (SMR) [5–15], the spin Seebeck effect [5,16 –18], the spin Peltier effect [19], the magnetic gating of pure spin currents [20,21] or the magnon spin transport (MST) [2,22 –30]. The fundamental building block structure employed to explore these phenomena is formed by a FMI layer –typically Y 3Fe5O12 (YIG) due to its small damping, soft ferrimagnetism and negligible magnetic anisotropy – and a non -magnetic (NM) metal with strong spin -orbit coupling (SOC) such as Pt or Ta placed next to it, which is essentially used to either generate or detect spin currents via the spin Hall effect (SHE) or its inverse [ 31–35]. Since these spintronic phenomena are based on the transfer of spin currents a cross the NM/FMI interface, it plays a key role in the properties and the performance of the resulting devices. It is well established that the most relevant parameter that determines the spin -current transport across the interface is the spin -mixing conductance 𝐺↑↓=(𝐺𝑟+ 𝑖𝐺𝑖) [5,36,37] . However, it is still under debate whether other interface effects could 2 also be relevant in these hybrid systems. Some examples are the magnetic proximity effect (MPE) [38–43], the Rashba -Edelstein effect [44–47], the anomalous Nerns t effect [38,48,49] or the spin -dependent interfacial scattering [50]. Therefore, understanding the role of the NM/FMI interface and the impact of its proper ties on the resulting spintronic phenomena is of outmost importance. In this work, we show that different spin -dependent phenomena in Pt/YIG -based devices (SMR and MST) are dramatically altered when the YIG surface is treated with a soft Ar+-ion milling. At room temperature, while the SMR effect in the treated samples is slightly larger than in the non -treated ones , the MST signal is fourfold increased. This extraordinary increase in the MST amplitude indicates that the spin-to-magnon conversion in Pt /YIG interfaces is strongly dependent on the magnetic details of the atomic layer of the YIG beyond the change in 𝐺↑↓. In addition, at low temperature, we observe a sign change of the SMR and a strong suppression of the MST signal in the treated samples, indicating the emergence of a surface magnetic frus tration of the treated YIG at low temperature. Our experimental res ults point out SMR and MST to be powerful tools to explore magnetic properties of surfaces and show that care should be taken when treating the surface of YIG, especially when used for studying spin - dependent phenomena originati ng at interfaces. II. Exper imental details Two different types of device structures were studied. In the first design, Pt/YIG samples were prepared by patterning a Pt Hall bar (width W=100 m, length L=800 m and thickness dN=7 nm ) on top of a 3.5 -m-thick YIG film [51] via e-beam lithography, sputtering deposition of Pt and lift -off, as fabricated in Ref. 52. In some samples, the YIG top surface was treated with a gentle Ar+-ion milling [53] prior the Pt deposition (Pt/YIG+ samples). In the second design, non-local NL -Pt/YIG and NL-Pt/YIG+ lateral nano structures were prepared on top of a 2.2 -m-thick YIG film [51] by patterning two long Pt strip lines ( W=300 nm, L1=15.0 m, L2=12.0 m and dN=5 nm) separated by a gap of ~500 nm –similar to the device structure used in Refs. 25 and 29–, following the same fabrication procedure used for the Hall bar . For each device structure , the Pt for both treated and non -treated samples was deposited in the same run . Here, for the sake of clarity, we present data taken for one sample of each type (Pt/YIG, Pt/YIG+, NL- Pt/YIG and NL-Pt/YIG+), although more samples were fabricated and measured, all showing reproducible results . Magnetotransport measurements wer e performed using a Keithley 6221 sourcemeter and a Keithley 2182 A nanovoltmeter operating in the dc -reversal method [54–56]. These measureme nts were performed at different temperatures between 10 and 300 K in a liquid -He cryostat that allows applying magnetic fields H of up to 9 T and to rotate the sample by 360º degrees . No difference in the magnetic properties between YIG and YIG+ substrates were observed via VSM magnetometry measurements (not shown). III. Results and Discussion IIIa. Spin Hall magnetoresistance First, we explore the angular -dependent magnetoresistance (ADMR) in Pt/YIG and Pt/YIG+ at room temperature. Figures 1( a)-1(c) show the longitudinal ( RL) ADMR 3 curves obtained for both samples in the three relevant H-rotation planes. The transverse (RT) ADMR curves taken in the plane are plotted in Fig. 1(d). The measurement configuration, the definition of the axes, and the rotation angles ( ) are defined in the sketches next to each panel. Note that for the magnetic fields applied, the magnetization of the YIG film is saturated [see Ref. 52 for the characterization of the YIG films]. The angular dependences are the same in both milled and non -milled samples and show the expected behaviour for the SMR effect, in agreement with measurements reported earlier in Pt/YIG bilayers [5–7,11,52] . FIG. 1 (color online). (a) -(c) Longitudinal ADMR measurements performed in Pt/YIG (dashed lines) and Pt/YIG+ (solid lines) samples at 300 K in the three relevant H-rotation planes (). (d) Transverse ADMR measurements taken in the same samples and temperature in the plane. Sketches on the right side indicate the definition of the angles, the axes, and the measurement configuration. The applied magnetic field is denoted in each panel . RL0 and RT0 are the subtracted base resistances. The SMR arises from the interaction of the spin currents generated in the NM layer due to the SHE with the magnetic moments of the FMI. According to the SMR theory [8,52] , the longitudinal and transverse resistivities of the Pt layer are given by 𝜌𝐿=𝜌0+∆𝜌0+∆𝜌1 (1−𝑚𝑦2), 𝜌𝑇=∆𝜌1𝑚𝑥𝑚𝑦+∆𝜌2𝑚𝑧, (1) where 𝐦(𝑚x,𝑚y,𝑚z)=𝐌/𝑀s are the normalized projections of the magnetization of the YIG film to the three main axes, 𝑀s is the saturated magnetization of the YIG and 𝜌0 is the Drude resistivity. ∆𝜌0 accounts for a number of corrections due to the SHE [52,57,58] , ∆𝜌1 is the main SMR term, and ∆𝜌2 accounts for an anomalous Hall - like contribution. Considering that these magnetoresistance (MR) corrections are very small, we identify the base resistivity of our longitudinal ADMR measurements as 𝜌𝐿0(𝑚𝑦=1)=𝜌0+∆𝜌0≃𝜌0. Since H is rotated in the plane of the film in our transverse measurements, the ∆𝜌2 contribution does not appear. Note that , in ADMR measurements, the amplitude of 𝜌𝐿(𝛽), 𝜌𝐿(𝛼) and 𝜌𝑇(𝛼) are equal and given by ∆𝜌1. Therefore, these measurements are equivalent when only the SMR contributes to the MR. The SMR term is quantified by 4 Δ𝜌1 𝜌0=𝜃𝑆𝐻2 λ 𝑑𝑁Re 2λ𝐺↑↓𝜌0tanh2(𝑑𝑁 2λ⁄ ) 1+2λ𝐺↑↓𝜌0coth (𝑑𝑁 λ⁄) , (2) where isthe spin diffusion length and SH the spin Hall angle of the Pt layer. According to Eq s. (1) and (2), the difference in the SMR amplitude observed between the two samples (see Fig. 1) can be interpreted as an enhanced 𝐺↑↓ at the Pt/YIG+ interface –with respect to Pt/YIG – due to the Ar+-ion milling process . Note that the spin transport properties for both Pt layers are expected to be the same because the measured resistivity is the same [59–61]. As the spin relaxation is governed by the Elliott -Yafet mechanism in Pt [59–61], we can calculate its spin diffusion length using the relation ×10-15Ωm2)/ρ [61]. Following Ref. 61, the spin Hall angle in the moderately dirty regime can be calculated using the intrinsic spin Hall conductivity 𝜎𝑆𝐻𝑖𝑛𝑡 (𝜃𝑆𝐻= 𝜎𝑆𝐻𝑖𝑛𝑡𝜌), which for Pt is 1600 Ω-1cm-1 [61,62] . In our films, L0 ~ 63 cm at 300 K, which thus corresponds to ~1.0 nm and SH~0.097. Using these andSH value s, dN=7 nm, Δ𝜌𝐿/𝜌0~5.310-5 and ~7.0610-5 (for Pt/YIG and Pt/YIG+, respectively, at 300 K) , and that Gi<<Gr [63], Eq. (2) yields Gr ~3.31013 -1m-2 and ~4.41013 -1m-2 for the Pt/YIG and Pt/YIG+ samples , respectively, which is within the range of values reported using the same bilayer structure [2,5–7,9–11,52,64,65] . This increase in Gr is in agreement with previous studies, where it was shown that an Ar+-ion milling process can improve the NM/YIG int erface quality by removing residues that might remain over the YIG substrate before the deposition of the NM layer [65,66] . However, it has been observed that an Ar+-ion milling process might also affect the YIG structure [49,64] . In the following, we proceed to study the temperature dependence of the SMR effect in these samples. Figures 2(a) and 2(b) show the measured temperature dependence of RT() for Pt/YIG and Pt/YIG+, respectively, in the angular range 0 -180º and for H=0.1 T. In both samples, the angular dependence predicted by the SMR effect is preserved when decreasing the temperature, following a sin( )cos() dependence [see Eq. (1)]. However, the polarity of the ADMR amplitude reverses the sign for Pt/YIG+ at low temperatures (crossing zero around T~45 K ), which is a completely unexpected behavior. According to the SMR theory , this amplitude is given by the term Δ𝜌1/𝜌0 in Eq. (2), which is a positive magnitude by definition. In Fig. 2(c), we plot the temperature dependence of the normalized amplitude of the transverse ADMR T/0T/L0=[[RT(45º)-RT(135º)]/RL0]·[L/W] for Pt/YIG (black squares) and Pt/YIG+ (red circles) . The weak temperature dependence of the SMR effect observed in our Pt/YIG sample is very similar to the one reported by other s using the same bilayer structure and it can be well understood with the temperature evolution of the spin transpor t properties in Pt [13,14,59,61] . In contrast, the different temperature dependen ce observed in Pt/YIG+ [see red dashed line in Fig. 2(c), which shows a scaling of the MR measured in Pt/YIG], having a sharp drop below 140 K and even a sign change at low temperatures, suggests the emergence of an additiona l interface effect. Systematic ADMR measurements are required to address its origin. Figure 2(d) show s the temperature dependence of the normalized amplitude of the longitudinal ADMR L/0L/L0=[RL(0º)-RL(90º)]/ RL0 measured in Pt/YIG+ for the three relevant H-rotation planes at H=1 T. We can see that both L()/0 and 5 L()/0 follow the same trend and that L()/0 remains zero, except for T~10 K. At very low temperatures, weak anti -localization effects emerge in Pt thin films [52,67 – 69], resulting in an extra out -of-plane vs in-plane MR, giving an explanation for the very small signal detected at 10 K. These measurements show that the sudden drop and the change in sign of the MR observed in Pt/YIG+ when decreasing temperature preserve the symmetry given by the polarization ( s) of the spin current produced in the Pt layer via the SHE , i.e., the measured MR has the symmetry of the SMR effect, which is distinct to the anisotropic MR that would appear if MPE were present. Therefore, this excludes MPE to be at the origin of the sign change of the MR at low temperatures in Pt/YIG+. FIG. 2 (color online). (a), (b) Transverse ADMR curves measured in Pt/YIG and Pt/YIG+, respectively, at different temperatures for H=0.1 T in the plane (see sketch). Data in the 180º - 360º range reproduce the same curves. RT0 is the subtracted base resistance at the corresponding temperature. (c) Temperature dependence of the normalized amplitude of the transverse ADMR, T/0, for the Pt/YIG (black squares) and Pt/YIG+ (red circles) samples extracted from (a) and (b), respectively. The red dashed line in (c) is a scaling of the temperature dependence of the amplitude measured in Pt/YIG to overlap with the amplitude obtained in Pt/YIG+ in the high temperature range (from ~150 to 300 K). (d) Temperature dependence of the normalized amplitude of the longitudinal ADMR, L/0, obtained in Pt/YI G+ at H=1 T and for the three H- rotation planes ( ). (e), (f) Transverse magnetic -field-dependent MR curves measured in Pt/YIG and Pt/YIG+, respectively, at 10 K with H in the plane of the film and for =45º and =135 º [see sketch in (e) for the color code of the magnetic field direction]. The vertical dashed lines show the saturation field of the YIG film obtai ned via magnetometry measurements . It is important to point out that , in hybrid systems of this kind , the interaction of s with the magneti zation M of the FMI leads to a resistance modulation not only due to the SMR , but also due to the excitation of magnons [25,29] . While the amplitude of the SMR is maximum when s and M are perpendicular, the resistance modulation due to magnon excitation is maximized when s and M are collinear. This implies that the MR modulation obtained in NM/FMI hybrids via ADMR measurements must actually be the result of the competition of these two spin-dependent MR effects, having the same angular dependences, but with reversed polarity. However, the MR expected from magnon excitation s is much smaller than from the SMR for the range of temperatures explored here. It has been estimated to be ~16 % at room temperature with respect to the SMR [19,21,25] , and that it should vanish at zero temperature [29]. Therefore, this rules out the excitation of magnons as responsible for the unexpected MR measured in H Jc H x y z VT H 0 50100150200250300-404812 H = 0.1 TPt/YIG Pt/YIG+T /(10-5) T(K)-3.0-1.50.01.53.0 0 45 90 135 180-4.0-2.00.02.04.0 0 50100150200250300-404812 60 K 30 K 10 K RT-RT0 (m) 300 K 200 K 100 K Pt/YIG+ 40K 30K 10K 80K 70K 60K 50K 300K 200K 130K 100K c)Pt/YIG Pt/YIG+ Pt/YIG/(10-5) T(K)0 50100150200250300-404812 H = 1 T Pt/YIG+L/0 (10-5 T (K)c) d) RT -RT0 (m) e) -240-120 0 120240-2.0-1.00.01.02.0 Pt/YIG H (Oe) = 45 = 135T=10 K -240-120 0 120240-1.0-0.50.00.51.0 Pt/YIG+ 45 135 H (Oe)T=10 KH Jc H x y z VT f) a) b) Fig. 2 6 Pt/YIG+ at low temperatures [see Fig. 2(b) and 2c)]. However, note that the excitation of magnons may lead to a larger correction in the ADMR amplitude at very high temperatures. This could give an alternative explanation to the measured temperature dependence of the MR in Pt/YIG bilayers close to the Curie temperature of the YIG film [15]. Figure s 2(e) and 2(f) show the magnetic -field-dependent MR curves measured in Pt/YIG and Pt/YIG+, respectively, at 10 K with the magnetic field applied in the plane of the film and along two representative directions ( =45º and =135º). The peaks and dips correspond to the magnetization reversal of the YIG film as reported earlier [6,9– 11]. Note that the saturation field of the YIG film obtained via magnetometry measurements (denoted as vertical dashed line s) matches perfectly with the one obtained through MR measurements in both samples. Moreover, the signs of the MR signals (for 45º and 135º) are reversed in Pt/YIG+ with respect to the ones measured in Pt/YIG, which is in agreement with the sign change observed in the ADMR at low temperatures [see Figs. 2(b) and 2(c)]. Because the SMR effect is basically sensitive to the magnetic properties of the first magnetic layer, having an estimated penetration depth of just a few Å [36], all previous measurements indicate that the ma gnetic moments of the surface of the YIG+ film are perpendicularly coupled to the ones of the bulk at low temperatures. The emergence of this surface magnetic frustration in our treated samples could be caused by a competing ferromagnetic and antiferromagn etic coupling of the modified complex stoichiometry of the YIG film due to the Ar+-ion milling process. In fact, magnetic frustration has already been observed in some ferrimagnets at low temperatures [70–73]. The angle between the magnetic moments of the surface and the bulk magnetization would be maximum (up to 90º) at low temperatures . The fact t hat the external magnetic field H aligns the bulk M but the SMR is sensitive to the magnetic moments of the surface yields a negative amplitude of the ADMR. A rise in the temperature would lead to a reduction of the angle due to the increase of the thermal energy in the magnetically coupled system . Considering our measurements, both surface and bulk magnetizations would lie together above ~140 K , recovering the expected positive amplitude of the ADMR. According to this physical picture, when the magnetic field (with H>HS) rotates in a particular H-rotation plane, the magnetic frustration forces the surface magnetization to point to a perpendicular direction. Due to the degeneracy in the orientation where the surface magnetization could point to, the angular dependences of the ADMR signals are preserved . As for the magnetic -field-dependent MR curves, when H<Hs, our YIG bulk film breaks in domains [74–76], resulting in the peaks and dips observed [see Fig. 2(e)]. The fact that the estimated HS of the surface magnetization via MR measur ements is the same for both samples [see Figs. 2(e) and 2(f)] and correlates with the measured HS of the film indicates that the magnetic moments of the surface of the YIG+ must be coupled to the bulk. The fact that the peaks and dips in the MR curves are reversed confirms that the angle between the magnetizations of the frustrated surface and the bulk should approach 90º at very low temperatures. In this scenario, one may think that , by applying a large enough magnetic field , we should be able to exert enough canting to the frustrated surface magnetic moments to shift the ADMR amplitude to positive values (i.e., reduce . Positive ADMR values have actually been measur ed for H>2T at low temperatures. However, the large Hanle 7 magnetoresistance (HMR) effect [52] present in our samples (the measured HMR amplitude at 300 K and 9 T is L/0~16·10-5) dominates the MR at large fields , preventing us from quantifying the canting exerted to the frustrated magnetic moments via MR measurements. An alternative interpretation of the temperature dependence of the SMR, motivated by the results obtained exploring a Pt/NiO/YIG system [77], is that the magnetic moments of the treated YIG+ surface are perpendicularly coupled to the magnetization of the YIG film at any temperature. In this situation, the frustrated magnetization of the surface dominates the SMR at low temperature, which is negative. When increasing the temperature, the frustrated surface becomes more transparent to the spin currents due to the thermal fluctuations and the YIG magnetization progressively dominates the SMR, which becomes positive. In other words, the spin current generate d by the Pt reaches the bulk YIG and the usual SMR in Pt/YIG is detected. This competition would lead to a decrease in the SMR amplitude below ~140 K , a comp ensation at an intermediate temperature (i.e., zero SMR amplitude, which oc curs around 45 K in our system), and a negative amplitude at low temperature s, when the frustrated Pt/YIG+ interface dominates. Our model allows us to qualitatively show that the emergence of a surface magnetic frustration can be well identified via SMR measurements. Note that magnetic frustration at the first atomic layer of a film cannot be detected by means of standard surface techn iques such as ma gneto -optical Kerr effect , magnetic force microscopy, or X -ray magnetic circular dichroism because of the relatively long penetration depth. Other surface sensitive techniques such as spin -polarized scanning tunneling microscopy or scanning electron microscopy with polarization analysis cannot be used in magnetic insulators either. Only complex, depth sensitive techniques such as polarized neutron reflectometry might resolve the surface magnetization independently of bulk. In other words, t he magnetic properties of the very first layer of a n insulating film will generally remain hidden by the large magnetic response of its bulk. Remarkabl y, unlike other techniques, the SMR can be applied to FMI films, is sensitive to only the first atomic layer [36], and its response is associated to the relative direction of the magnetic moments of the FM with respect to the spins of the NM layer (whether they are parallel or perpendicular), but not to their orientation (up or down). This highlights the potential of the SMR to explore complex surface magnetic properties [78]. IIIb. Magnon spin transport We now move to study the magnon spin transport in the non-local NL-Pt/YIG and NL- Pt/YIG+ samples. Figure 3(a) shows an optical image of one of the devices fabricated. In these samples, the current is injected in the central wire and both the local resistance (RL=VL/I) and the non -local resistance ( RNL=VNL/I) are measured as schematically drawn in Fig. 3(a). Note that RNL is measured using the dc -reversal method [54–56], which is equivalent to the first harmonic signal in ac lock -in type measurements [79]. 8 FIG. 3 (color online). (a) Optical image of the NL -Pt/YIG sample. Grey wires are the Pt stripes and the yellow areas correspond to additional Au pads. The black background is the surface of the YIG film. Both the local and non -local measurement configurations are schematically shown. (b) and (c) are the local ( RL) and non -local ( RNL) ADMR signals, respectively, measured in the NL -Pt/YIG sample at 150 K and for H=1 T rotating in the plane. Note that, along this rotation angle, M changes its relative orientation with s (being parallel for =90º and 270º and perpendicular for =0º and 180º). In (b), the bias current was 100 A. In (c), non -local ADMR measurements performed at I=100 (black line) and 300 (red line) are shown. The arrows in (b) and (c) indicate the sign convention used for the amplitude of the local ( RL) and non -local (RNL) resistance plotted in Figs. 4(a) and 4(b), respectively. Figure s 3(b) and 3(c) show an example of the local and non -local ADMR measurements, respectively, performed in our samples. The data correspond to the NL- Pt/YIG sample measured at 150 K with H=1 T rotating in the plane [see Fig. 1(b) for the definition]. Simila r ADMR curves were obtained in the NL-Pt/YIG+ sample. The local resistance RL [Fig. 3(b)] shows the expected cos2() dependence for the SMR effect . Taking into account that in these samples 𝜌𝐿0(300 K)~54 𝜇Ωcm –which according to Ref. 61 corresponds to ~1.2 nm and SH~0.083 for the Pt film –, that the measured SMR amplitudes at the same temperature are Δ𝜌𝐿/𝜌0~6.210-5 and ~7.610-5 (for the NL -Pt/YIG and NL-Pt/YIG+ samples , respectively) , dN=5nm, and that Gi<<Gr [63], Eq. (2) yields Gr ~3.21013-1m-2 and ~4.01013 -1m-2 for the Pt/YIG and Pt/YIG+ interfaces , respectively, which is in very good agreement with our previous results. 9 The non -local resistance RNL [Fig. 3(c)] shows a sin2() dependence, which is expected for the excitation, transpor t and detection of magnon spin information through the YIG film [25,29,30] . The physical description of this phenomenon is the following. The current applied in the central Pt wire (injector) produces a transverse spin current ( via the SHE ) that flows along the z axis [being s parallel to the y axis; see Fig. 3(a) for the definition of the axes]. When these spins reach the Pt/YIG interface, they can excite (annihilate ) magnons in the YIG film when s is parallel (antiparallel) to M [25], which produce a change in the magnon population below the Pt injector. These non - equilibrium magnons diffuse throug h the YIG film and, when they reach the nearby Pt wire (detector), the reciprocal process takes place. Therefore, the non -equilibrium magnons below the Pt detector transform into a non -equilibrium spin imbalance at the Pt/YIG interface, which produces the flow of a pure spin current perpendicular to the interface that is ultimately converted into a perpendicular charge current (along the Pt wire) via the ISHE. The combination of all these processes generates the non -local resistance RNL shown in Fig. 3(c) [80]. The angular dependence observed in Fig. 3(c) confirms that the excitation and absorption of propagating magnons in the YIG film are maxima when s and M are collinear, which occurs for =90º and 270º (note that the sign of the signal captured agrees with th e sign convention chosen for our experiments [25,29] ). Moreover, VNL should be linear with I for moderate applied currents [25]. This is confirmed in Fig. 3(c), where it is shown that the same RNL() curve is obtained for I=100 (black) and 300 A (red). The amplitude of the RNL() curve measured in our sample is consistent with results reported using YIG films with similar thicknesses [29]. Figure 4 shows the temperature dependence of the amplitude of (a) the SMR and (b) the MST measured in both the NL-Pt/YIG (black squares) and NL-Pt/YIG+ (red circles) samples. The sign of the amplitude of the SMR (local) and the MST (non -local measurements) is indicated with the arrows drawn in Figs. 3(b) and 3(c), respectively. The SMR data is presented normalized to the base resistance, followin g the same procedure used in the previous case . In Fig. 4(a), we see that the temperature dependence of the SMR i n these samples is qualitatively similar to the one observed in the previous experiments [see Figs. 2(c) and 2(d)], which confirms once again the emergence of a surface magnetic frustration in the treated YIG+ substrate at low temperatures. Interestingly, while the amplitude of the SMR in the temperature range ~150 -300 K is only slightly larger in the NL-Pt/YIG+ sample than in the NL-Pt/YIG one (i.e., slight enhancement of 𝐺r), the amplitude of the MST is about four times larger [see Fig. 4( b)]. This indicates that in this temperature range the efficiency of the spin -to-magnon conversion (and its reciprocal process) in the treated Pt/YIG+ interface is much higher than in the non-treated Pt/YIG interface , but not related to the change in 𝐺r. Instead , it must be associated to the different magnetic properties of the treated YIG+ surface compared to the YIG bulk for temperatures above the emergence of the magnetic frustration. Further studies will be needed in order to fully understand the role of this surface enhancement . 10 FIG. 4 (color online). Temperature dependence of the amplitude of (a) the SMR and (b) the MST measured in NL -Pt/YIG (black squares) and NL -Pt/YIG+ (red circles). The amplitude is extracted from ADMR measurements performed in the plane at H=1 T. Measurements in (a) and (b) are independent of I (at least) to up to 300 A. The inset in panel (b) shows a zoom of the measured RNL at low temperatures. Black solid line is a fit to the experimental points to the power law dependence T3/2. The temperature dependence of the amplitude of the MST follows a remarkabl y different trend than the SMR, which is in agreement with recent report s [29]. In fact, we found that the MST amplitude in the NL-Pt/YIG sample at low temperatures follows a ~T3/2 dependence [see inse t in Fig. 4(b)], expected for thermally induced diffusive magnons in the limit of large magnon diffusion lengths (i.e., weak magnon -phonon interactions) [27,29,81,82] . Importantly, the temperature dependence decays more abruptly for the NL-Pt/YIG+ sample, and no MST signal is detected at low temperatures (within the noise level), evidencing that the emergence of the surface magnetic frustration r esults in the suppression of non -equilibrium diffusive magnons at the surface of the YIG+ film. In other words, the frustrated magnetic surface, which may host a magnon dispersion relation different from the YIG bulk, is preventing the efficient spin -to-magnon conversion (and viceversa) at the Pt/YIG+ interface. IV . Conclusions We demonstrate via SMR and MST measurements in Pt/YIG -based devices that an Ar+- ion milling treatment of the YIG surface has a profound impact in the resulting spintronic phenomena. Beyond a slight increase in the spin-mixing conductance observed for the treated samples at room temperature, which accounts for a better interface quality, we show tha t the MST is fourfold increased. This elucidates the higher sensitivity of the magnon excitations to fine details in the magnetic properties of the magnetic surface. Moreover, we show that the treated surface of YIG develops a magnetic frustration at low temperature, which makes the SMR signal to reverse the 11 sign below ~45K and dramatically suppresses the spin -to-magnon excitations in these interfaces. Our results give new insights on the interactions between the spi ns in a NM material with the magnetic moments of a FM at interfaces free from MPE, and show the potential of SMR and MST to explore the magnetic properties of materials with complex magnetic textures and surfaces. ACKNOWLEDGEMENTS We thank E. Sagasta fo r fruitful discussions. This work was supported by the European Union 7th Framework Programme under the European Research Council (257654 - SPINTROS) and the Marie Curie Ac tions (607904 -13-SPINOGRAPH) , by the Spanish MINECO under Project s No. MAT2012 -37638 and MAT2015 -65159 -R, and by the Basque Government under Project No. PC2015 -1-01. REFERENCES [1] A. Brataas, Physics 6, 56 (2013). [2] Y. Kajiwara, K. Harii, S. Takahashi, J. Ohe, K. Uchida, M. Mizuguchi, H. Umezawa, H. Kawai, K. Ando, K. Takanashi, S. Maekawa, and E. Saitoh, Nature 464, 262 (2010). [3] Y. Tserkovnyak, A. Brataas, and B. I. Halperin, Rev. Mod. Phys. 77, 1375 (2005). [4] E. Saitoh, M. Ueda, H. Miyajima, and G. Tatara, Appl. Phys. Lett. 88, 182509 (2006). [5] M. Weiler, M. Althammer, M. Schreier, J. Lotze, M. Pernpeintner, S. Meyer, H. Huebl, R. Gross, A. Kamra, J. Xiao, Y. -T. Chen, H. J. Jiao, G. E. W. Bauer, and S. T. B. Goennenwein, Phys. Rev. Lett. 111, 176601 (2013). [6] H. Nakayama, M. Althammer, Y. -T. Chen, K. Uchida, Y. Kajiwara, D. Kikuchi, T. Ohtani, S. Geprägs, M. Opel, S. Takahashi, R. Gross, G. E. W. Bauer, S. T. B. Goennenwein, and E. Saitoh, Phys. Rev. Lett. 110, 206601 (2013). [7] C. Hahn, G. de Loubens, O. Klein, M. Viret, V. V. Naletov, and J. Ben Youssef, Phys. Rev. B 87, 174417 (2013). [8] Y.-T. Chen, S. Takahashi, H. Nakayama, M. Althammer, S. T. B. Goennenwein, E. Saitoh, and G. E. W. Bauer, Phys. Rev. B 87, 144411 (2013). [9] N. Vlietstra, J. Shan, V. Castel, B. J. van Wees, and J. Ben Youssef, Phys. Rev. B 87, 184421 (2013). [10] N. Vlietstra, J. Shan, V. Castel, J. Ben Youssef, G. E. W. Bauer, and B. J. van Wees, Appl. Phys. Lett. 103, 032401 (2013). [11] M. Althammer, S. Meyer, H. Nakayama, M. Schreier, S. Altmannshofer, M. Weiler, H. Huebl, S. Geprägs, M. Opel, R. Gross, D. Meier, C. Klewe, T. Kuschel, J. -M. Schmalhorst, G. Reiss, L. Shen, A. Gupta, Y. -T. Chen, G. E. W. Bauer, E. Saitoh, and S. T. B. G oennenwein, Phys. Rev. B 87, 224401 (2013). 12 [12] M. Isasa, A. Bedoya -Pinto, S. Vélez, F. Golmar, F. Sanchez, L. E. Hueso, J. Fontcuberta, and F. Casanova, Appl. Phys. Lett. 105, 142402 (2014). [13] S. Meyer, M. Althammer, S. Gepr, and S. T. B. Goennenwein, Appl. Phys. Lett. 104, 242411 (2014). [14] S. R. Marmion, M. Ali, M. McLaren, D. A. Williams, and B. J. Hickey, Phys. Rev. B 89, 220404 (2014). [15] K. Uchida, Z. Qiu, T. Kikkawa, R. Iguchi, and E. Saitoh, Appl. Phys. Lett. 106, 052405 (2015). [16] K. Uch ida, J. Xiao, H. Adachi, J. Ohe, S. Takahashi, J. Ieda, T. Ota, Y. Kajiwara, H. Umezawa, H. Kawai, G. E. W. Bauer, S. Maekawa, and E. Saitoh, Nat. Mater. 9, 894 (2010). [17] S. Geprägs, A. Kehlberger, F. Della Coletta, Z. Qiu, E. -J. Guo, T. Schulz, C. Mix, S. Meyer, A. Kamra, M. Althammer, H. Huebl, G. Jakob, Y. Ohnuma, H. Adachi, J. Barker, S. Maekawa, G. E. W. Bauer, E. Saitoh, R. Gross, S. T. B. Goennenwein, and M. Kläui, Nat. Commun. 7, 10452 (2016). [18] M. Schreier, N. Roschewsky, E. Dobler, S. Meyer, H. Huebl, R. Gross, and S. T. B. Goennenwein, Appl. Phys. Lett. 103, 242404 (2013). [19] J. Flipse, F. K. Dejene, D. Wagenaar, G. E. W. Bauer, J. Ben Youssef, and B. J. van Wees, Phys. Rev. Lett. 113, 027601 (2014). [20] E. Villamor, M. Isasa, S. Vélez, A . Bedoya -Pinto, P. Vavassori, L. E. Hueso, F. S. Bergeret, and F. Casanova, Phys. Rev. B 91, 020403 (2015). [21] F. K. Dejene, N. Vlietstra, D. Luc, X. Waintal, J. Ben Youssef, and B. J. van Wees, Phys. Rev. B 91, 100404 (2015). [22] D. D. Stancil and A. Prabhakar, Spin Waves: Theory and Applications (Springer, 2009). [23] A. V. Chumak, V. I. Vasyuchka, A. A. Serga, and B. Hillebrands, Nat. Phys. 11, 453 (2015). [24] A. V Chumak, A. A. Serga, and B. Hillebrands, Nat. Commun. 5, 5700 (2014). [25] L. J. Cornelissen, J. Liu, R. A. Duine, J. Ben Youssef, and B. J. van Wees, Nat. Phys. 11, 1022 (2015). [26] S. S.-L. Zhang and S. Zhang, Phys. Rev. Lett. 109, 096603 (2012). [27] S. S.-L. Zhang and S. Zhang, Phys. Rev. B 86, 214424 (2012). [28] S. A. Bender, R. A. Duine, and Y. Tserkovnyak, Phys. Rev. Lett. 108, 246601 (2012). [29] S. T. B. Goennenwein, R. Schlitz, M. Pernpeintner, K. Ganzhorn, M. Althammer, R. Gross, and H. Huebl, Appl. Phys. Lett. 107, 172405 (2015). [30] L. J. Cornelissen a nd B. J. van Wees, Phys. Rev. B 93, 020403 (2016). 13 [31] M. I. Dyakonov and V. I. Perel, Phys. Lett. A 35, 459 (1971). [32] J. E. Hirsch, Phys. Rev. Lett. 83, 1834 (1999). [33] A. Hoffmann, IEEE Trans. Magn. 49, 10 (2013). [34] S. O. Valenzuela and M. Tinkh am, Nature 442, 176 (2006). [35] M. Morota, Y. Niimi, K. Ohnishi, D. H. Wei, T. Tanaka, H. Kontani, T. Kimura, and Y. Otani, Phys. Rev. B 83, 174402 (2011). [36] X. Jia, K. Liu, K. Xia, and G. E. W. Bauer, Europhys. Lett. 96, 17005 (2011). [37] A. Brataas, Y. V. Nazarov, and G. E. W. Bauer, Phys. Rev. Lett. 84, 2481 (2000). [38] S. Y. Huang, X. Fan, D. Qu, Y. P. Chen, W. G. Wang, J. Wu, T. Y. Chen, J. Q. Xiao, and C. L. Chien, Phys. Rev. Lett. 109, 107204 (2012). [39] Y. M. Lu, J. W. Cai, S. Y. Huang, D. Qu , B. F. Miao, and C. L. Chien, Phys. Rev. B 87, 220409 (2013). [40] B. F. Miao, S. Y. Huang, D. Qu, and C. L. Chien, Phys. Rev. Lett. 112, 236601 (2014). [41] Y. Yang, B. Wu, K. Yao, S. Shannigrahi, B. Zong, and Y. Wu, J. Appl. Phys. 115, 17C509 (2014). [42] T. Lin, C. Tang, and J. Shi, Appl. Phys. Lett. 103, 132407 (2013). [43] T. Lin, C. Tang, H. M. Alyahayaei, and J. Shi, Phys. Rev. Lett. 113, 037203 (2014). [44] V. L. Grigoryan, W. Guo, G. E. W. Bauer, and J. Xiao, Phys. Rev. B 90, 161412 (2014). [45] I. M. Miron, K. Garello, G. Gaudin, P. -J. Zermatten, M. V. Costache, S. Auffret, S. Bandiera, B. Rodmacq, A. Schuhl, and P. Gambardella, Nature 476, 189 (2011). [46] K. Garello, I. M. Miron, C. O. Avci, F. Freimuth, Y. Mokrousov, S. Blügel, S. Auffret, O. Boulle, G. Gaudin, and P. Gambardella, Nat. Nanotechnol. 8, 587 (2013). [47] G. Allen, S. Manipatruni, D. E. Nikonov, M. Doczy, and I. A. Young, Phys. Rev. B 91, 144412 (2015). [48] S. Y. Huang, W. G. Wang, S. F. Lee, J. Kwo, and C. L. Chien, Phys. Rev. L ett. 107, 216604 (2011). [49] B. F. Miao, S. Y. Huang, D. Qu, and C. L. Chien, Phys. Rev. Lett. 111, 066602 (2013). [50] S. S.-L. Zhang and G. Vignale, Phys. Rev. Lett. 116, 136601 (2016). [51] The YIG film was grown by liquid phase epitaxy on a (111) gadolinium gallium garnet single crystal at Innovent e.V. (Jena, Germany). The YIG films used in this work are 3.5 and 2.2 m thick for the first and second design, respectively. 14 [52] S. Vélez, V. N. Golovach, A. Bedoya -Pinto, M. Isasa, E. Sagasta, M. Abad ia, C. Rogero, L. E. Hueso, F. S. Bergeret, and F. Casanova, Phys. Rev. Lett. 116, 016603 (2016). [53] Samples with different milling times ranging from 10 to 30 s were explored and showed similar results. The milling was performed in a 4Wave Load Lock Etch System under the following conditions: PBN emission 0.5 A, PBN Ar flow 8 sccm, ion source Ar flow 15 sccm, beam voltage 300 V, beam current 50 mA, accelerator voltage 50 V, angle incidence 10º normal to sample surface and spin speed 15 rpm. [54] F. Ca sanova, A. Sharoni, M. Erekhinsky, and I. K. Schuller, Phys. Rev. B 79, 184415 (2009). [55] M. Erekhinsky, A. Sharoni, F. Casanova, and I. K. Schuller, Appl. Phys. Lett. 96, 022513 (2010). [56] M. Erekhinsky, F. Casanova, I. K. Schuller, and A. Sharoni, Ap pl. Phys. Lett. 100, 212401 (2012). [57] M. I. Dyakonov, Phys. Rev. Lett. 99, 126601 (2007). [58] A. V. Khaetskii, Fiz. Tekh. Poluprovodn 18, 1744 (1984) [Sov. Phys. Semicond. 18, 1091 (1984)] . [59] M. Isasa, E. Villamor, L. E. Hueso, M. Gradhand, and F. C asanova, Phys. Rev. B 91, 024402 (2015). [60] M.-H. Nguyen, D. C. Ralph, and R. A. Buhrman, Phys. Rev. Lett. 116, 126601 (2016). [61] E. Sagasta, Y. Omori, M. Isasa, M. Gradhand, L. E. Hueso, Y. Niimi, Y. Otani, and F. Casanova, Phys. Rev. B 94, 060412(R) (2016) . [62] L. Wang, R. J. H. Wesselink, Y. Liu, Z. Yuan, K. Xia, and P. J. Kelly, Phys. Rev. Lett. 116, 196602 (2016). [63] By measuring the anomalous Hall effect -like contribution to the SMR [ 2 in Eq. (1), not shown] , we estimate that 2/ in our samples. This result indicates that Gi/Gr~0.05, which is in agreement with other values reported usi ng Pt/YIG structures (see Ref. 11 ) and with the expected theoretical value for NM/FMI bil ayers (see Ref. 36). [64] Z. Qiu, K. Ando, K. Uchida, Y. Kajiwara, R. Takahashi, H. Nakayama, T. An, Y. Fujikawa, and E. Saitoh, Appl. Phys. Lett. 103, 092404 (2013). [65] M. B. Jungfleisch, V. Lauer, R. Neb, A. V. Chumak, and B. Hillebrands, Appl. Phys. Lett. 103, 022411 (2013). [66] C. Burrowes, B. Heinrich, B. Kardasz, E. A. Montoya, E. Girt, Y. Sun, Y. -Y. Song, and M. Wu, Appl. Phys. Lett. 100, 092403 (2012). [67] H. Hoffmann, F. Hofmann, and W. Schoepe, Phys. Rev. B 25, 5563 (1982). [68] G. Bergmann, Phys. Rev. Lett. 48, 1046 ( 1982). 15 [69] Y. Niimi, D. Wei, H. Idzuchi, T. Wakamura, T. Kato, and Y. Otani, Phys. Rev. Lett. 110, 016805 (2013). [70] M. E. Lines and M. Eibschutz, Phys. Rev. B 32, 3067 (1985). [71] T. Sarkar, V. Pralong, V. Caignaert, and B. Raveau, Chem. Mater. 22, 2885 (2010). [72] Y. S. Hou, H. J. Xiang, and X. G. Gong, Sci. Rep. 5, 13159 (2015). [73] A. K. Paul, M. Reehuis, V. Ksenofontov, B. Yan, A. Hoser, D. M. Többens, P. M. Abdala, P. Adler, M. Jansen, and C. Felser, Phys. Rev. Lett. 111, 167205 (2013). [74] C. B. Braus, Magnetic Properties of Nanocrystalline Materials for High Frequency Applications, University of Groningen, 2003. [75] W. X. Xia, Y. S. Chun, S. Aizawa, K. Yanagisawa, K. M. Krishnan, D. Shindo, and A. Tonomura, J. Appl. Phys. 108, 123919 (2010). [76] L. Pietrobon, L. Fallarino, A. Berger, A. Chuvilin, F. Casanova, and L. E. Hueso, Small 11, 6242 (2015). [77] D. Hou, Z. Qiu, J. Barker, K. Sato, K. Yamamoto, S. Vélez, J. M. Gomez -perez, L. E. Hueso, F. Casanova, and E. Saitoh. In Preparation . [78] M. Isasa, S. Vélez, E. Sagasta, A. Bedoya -Pinto, N. Dix, F. Sánchez, L. E. Hueso, J. Fontcuberta, and F. Casanova, Phys. Rev. Appl. 6, 034007 (2016). [79] F. L. Bakker, A. Slachter, J. -P. Adam, and B. J. van Wees, Phys. Rev. Lett. 105, 136601 (2010). [80] Note that in this physical description we are not taking into account any spin transport contribution coming from finite thermal gradients that might be present across the Pt/YIG interfaces (which would be induced by the joule heat dissipated in the inject or). These effects do not contribute to our measurements . [81] S. A. Bender and Y. Tserkovnyak, Phys. Rev. B 91, 140402 (2015). [82] J. Xiao and G. E. W. Bauer, arXiv:1508.02486 (2015). | 2016-06-09 | We study the spin Hall magnetoresistance (SMR) and the magnon spin transport
(MST) in Pt/Y3Fe5O12(YIG)-based devices with intentionally modified interfaces.
Our measurements show that the surface treatment of the YIG film results in a
slight enhancement of the spin-mixing conductance and an extraordinary increase
in the efficiency of the spin-to-magnon excitations at room temperature. The
surface of the YIG film develops a surface magnetic frustration at low
temperatures, causing a sign change of the SMR and a dramatic suppression of
the MST. Our results evidence that SMR and MST could be used to explore
magnetic properties of surfaces, including those with complex magnetic
textures, and stress the critical importance of the non-magnetic/ferromagnetic
interface properties in the performance of the resulting spintronic devices. | Competing effects at Pt/YIG interfaces: spin Hall magnetoresistance, magnon excitations and magnetic frustration | 1606.02968v2 |
1 Planar Hall effect in Y3Fe5O12 (YIG) /IrMn films X. Zhang1 1 Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China Correspondence and requests for materials should be addressed to X. Z. (zhangxu1986@ iphy.ac.cn ) ABSTART The planar Hall effect of IrMn on an yttrium iron garnet (YIG = Y 3Fe5O12) was measured in the magnetic field rotating in the film plane . The magnetic field angular dependence of planar Hall resistance (PHR) has been observed in YIG/IrMn bilayer at different temperatures , while the GGG/IrMn (GGG= Gd3Ga5O12) shows constant PHR for different magnetic field angles at both 10 K and 300 K . This provides evidence that IrMn has interfacial spins which can be led by FM in YIG/IrMn structure. A hysteresis can be observed in PHR- magnetic field angle loop of YIG/IrMn films at 10 K , indicating the irreversible switching of IrMn interfacial spins at low temperature. I. INTRODUCTION The coupling between a ferromagnetic (FM )-antiferromagnetic ( AFM ) bilayer at the interface can lead to a shift of the hysteresis loop along the magnetic field axis due to unidirectional anisotropy. This phenomenon is referred as exchange bias (EB) , which often accompanies with the enhancement of coercivity. Due to the pinning e ffect of magnetic moment in the adjacent FM layer, exchange bias effect has become an integral part in modern spintronics devices such as giant magnetoresistive heads 1. Although EB was discovered more than 60 years ago2, the mechanism of it is still a controversial and attracting object to be comprehended as one fundamental physic al issue. To 2 understand the mechanism of EB effect, much effort has been dedicated to investigate the behaviors of FM layer in EB systems ,3-5 while the attention on the essential feature of AFM is relatively less due to the notorious difficulties of requirement of large -scale facilities, such as neutron diffraction or XMCD. The theory has predicted that the unidirectional anisotropy in EB system derives from the interfacial uncompensated AFM spin, which does not reverse with the magnetization reversal of the FM layer. However, there is still no solid evidence to prove the existence of it even in the intensively studied Co (-Fe)/IrMn (111) in -plane exchange -biased system .6-8 Compared to the measuring methods above, there is a more common way to investigate EB system , which is through the electronic transport properties. In the past years, EB effect measured by planar Hall effect has become a more interesting subject due to higher signal -to-noise ratio compared with the magnetoresistance or spin valve configuration .3,9,10 In a FM metallic film, the magneto resistance can be separated into longitudinal and transverse part due to the anisotropic scattering of conduction elections. For a single domain model, the electric field can be described as :11 2( ) cosxEj jρ ρρ α⊥⊥=+− (1) ( )sin cosyEjρρ α α⊥= − (2) where the current density j is assumed along the x- axis direction, the magnetization of the single domain is at angle α with respect to x -axis, while ρand ρ⊥are the resistivity parallel and perpendicular to the magnetization, respectively. Equ. (1) is for anisotropic magnetoresistance (AMR) , while Equ. (2) is f or planar Hall effect ( PHE ). By using AMR effect and PHE, Guohong Li .etc have performed significant work on investigating the NiFe/NiMn EB multilayer and spin- valve.10 And recently, researches have shown that the reversal of AFM magnetization can lead to the AMR effect in the tunnel junction 3 structures .12-14 One significant work is the observation of large AF M-tunneling anisotropic magnetores istance (TAMR ) in the MTJs based on NIFe/IrMn/MgO/Pt structure by B.G. Park group,12 which emphasizes the significant role of AFM in generating the TAM R. However, these works mentioned above contains both conductive FM and AFM, suggesting that the magnetoresistance signal is consisted of the contribution from both layers. In this condition, the FM may provide an important part of the AMR signal since the reversing of its magnetization can cause a large variation of magnetoresistance. This prevent s us from directly observing the magnetization state of AFM layer and the proportion of AMR contribut ed by AFM is also not cle ar. In spite of this , these works shed the light on studying the AFM properties in EB system by AMR and PHE . In this work , in order to study the behavior of AFM with other disturbance excluded, an almost isolated FM material Y 3Fe5O12 (YIG) was chosen as FM layer. Through this way, all the magnetoresistance signal would derive from IrMn, thus the result would give out the more specific and directly information of AFM IrMn. By usin g isolated FM layer in the FM -AFM structure, it provide s evidence that there exists uncompensated spins at the interface of IrMn that can be reversed by FM in the YIG/IrMn system at room temperature, and the anisotropy of A FM interfacial spins get enhanced down to the low temperature region. II. Experimental In this study, the YIG substrates is consist of 4 μm thick single crystalline (111) YIG layers grown by liquid phase epitaxy on (111) Gd 3Ga5O12 (GGG) substrates , and single crystalline (111) GGG substrates are also used in this work . Films with structure of YIG/ Ir25Mn 75 (1.8-15 nm)/Ta (2 nm) and GGG/Ir 25Mn 75 (5 nm)/Ta (2 nm) were deposited by DC magnetron sputtering, where (111) texture of the substrates is to promote (111) texture in Ir 25Mn 75 layer , and Ta serves as a cap layer to protect the sample from oxidation/contamination. The vacuum of the sputtering system was better than 4×10-5 Pa, and the working Ar pressure was 0.5 Pa. Sets of up to 18 samples were prepared at a time. Composite Ir -Mn target by placing Mn chips symmetrically on the Ir target 4 was used to deposit Ir 25Mn 75 film. The composition was determined by inductively coupled plasma- atomic emissio n spectroscopy (ICP -AES). The samples were patterned into four terminal Hall bars. For simplicity, Ir 25Mn 75 hereinafter is denoted as IrMn. M -H, PHR –θ and PHR- H curves were measured by a vibrating sample magnetometer (VSM) and physical property measurement system (PPMS) at different temperature s, respectively , where PHR is planar Hall resistivity and θ is t he external magnetic field angl e. III. RESULTS AND DISCUSSION Magnetic hysteresis loops measured by VSM s hows that the YIG films are magnetically soft and isotropic in the film plane. As shown in the uppe r panel of Fig. 1(a ), at room temperature, the YIG film has an in -plane saturation field of about 60 Oe , and the loop remains the same as the sample rotates 90° in the same plane (not shown) . The YIG film has a spontaneous magnetization (4πM S) of 1.73 kG, which can be obtained in the out -of-plane loop due to the shape anisotropy , as indicated in the bottom panel of Fig. 1(a) . These magnetic properties is in agreement with the YIG films reported before .15 Fig. 1(b) shows the hysteresis loop of a typical exchange bias structure of CoFe (8 nm)/IrMn (10 nm), which displays a saturation field smaller than 300 Oe and an exchange bias field of about 120 Oe, proving that the IrMn deposited by our composite target is reliable antiferromagnetic material . Fig. 1(c ) and ( d) exhibit the magnetic field angular dependence of PH R for mon o-CoFe (8 nm) layer and CoFe (8 nm)/IrMn (10 nm), respectively. In the angular -dependent measurement, the sample was rotated first from θ=0° to 360° and backwards from to θ= 360° to 0° in a fixed magnetic field. The data in Fig. 1 ( c) was measured under 2 kOe magnetic field at 300 K, which is larger than the saturation field of CoFe . The behavior of CoFe layer illustrates that the PHR of single FM layer would exhibit a sinusoidal dependence of magnetic field angular with period of 180º, and also no difference between the clockwise and anticlockwise curve is observed. Note that the curve is not symmetric of zero resistance point as expected for uniaxial anisotropy of FM material s, which 5 can be explained by the misalignment of the Hall bar that measures the voltage, leading to an additional resistance along the current direction . When coupled to IrMn, the CoFe (8 nm)/IrMn (10 nm) (referred as CoFe/IrMn hereinafter ) structure develops exchange bias effect due to the coupling between FM and AFM layer , and an interfacial unidirectional anisotropy would be induced into the system as a result . Fig. 1( b) shows the hysteresis loop of CoFe/IrMn , which displays a saturation field smaller than 300 Oe and an exchange bias field of about 120 Oe. In Fig. 1( d), for the magnetic field below 300 Oe , the angular -dependent curve of CoFe/IrMn shows a distortion compared to mono- CoFe layer data . Since 300 Oe is larger than the saturation field of CoFe/IrMn structure based on the loops i n Fig. 1(b) , the PHR- theta curve of 300 Oe indicat es that the unidirec tional anisotropy would lead to a magnetic moment rotating processes different to that of uni axial anisotropy . When magnetic field increases to 2 kOe, which is much larger than the saturation field obta ined in Fig. 1( b), although no obvious distortion can be observed in the PHR curve, the re is still a slightly deviation when fitting PHR- θ data by a sine function. This suggests that most of the CoFe layer synchronously rotates with the external magnetic field of 2 kOe. It is worth mentioning that because of the coupling between CoFe and IrMn, some of the interfacial IrMn spins should rotates certain degree with CoFe, which also should have contribution to PHR . However , owning to the much larger signal of 8 nm thick CoFe layer, this PHR from IrMn can hardly be confirmed . In addition, it is worth noting that the PHR -θ curve under 50 Oe field shows a hysteresis behavior. Since 50 Oe is not large enough to reverse the CoFe spins pinned by IrMn, and also PHR- θ curve at this field shows a smaller amplitude when compared with the curves of 300 Oe and 2 kOe , which suggests that the FM moments still rotates a small angle and this spin flo pping follow s a different route in the clockwise and anticlockwise magnetic field cycling procedure , i.e. the FM moment shows an irreversible switching under 50 Oe field. Recently, several researches have reported the AMR derives from AFM in the FM-AFM EB systems, proposing a new possibility to obtain large magnetoresistance in relatively small fields .12-14 However, these studies probed AFM with conductive FM layer, which 6 would lea d to difficulty of distinguishing AFM signal from FM signal, thus the specific behavior of AFM is still not clearly revealed. In order to investigate the interfacial magnetization of AFM in a n FM-AFM coupled system more independently , antiferromagnetic material IrMn was deposit on the isolated ferromagnetic material YIG. Figure 2 shows the angular dependence of PHE for YIG/ IrMn (1.8 nm) and ( 5 nm) at different temperatures . The sample was first ly field cooling to 10 K under a constant field of 3 kOe in the same direction as the magnetic field applied during the film growth , and then measured at increasing temperatures in the magnetic field of 2 kOe, which is larger than the saturation field of common exchange bias system . At each temperature, t he measurement s were also performed form 0 ° to 360° and cycling back to 0° for both samples . Clearly , magnetic field angle dependence of PHR can be observed in both YIG/IrMn sampl es from 10 K to 300 K, indicating an anisotropic magnetoresist ance. Since 2 kOe is not large enough to rotate bulk magnetic moments of IrMn as well as YIG is a nearly isolated material , the anisotropy signal c an only derive from the rotation of IrMn interfacial moment, confirm ing that there are uncompensated spins existing in the interface of IrMn that can be rotated by a small magnetic field. In Fig. 2(a) and (e), a hysteresis of magnetic field angle can be observed for both samples at 10 K, suggesting that the IrMn interfacial moment cannot be fully reversed by the magnetic field of 2 kOe at this temperature. It was reported that in an EB structure of Co/YMnO 3 with i nsulated YMnO 3 as antiferromagnetic layer, a similar hysteresis can be observed in the angular -dependent magnetic field measurement of Co at low temperatures5, where EB effect is more remarkable. However, in contrast to Co/YMnO 3 system, the contributions to PHE are all from the AFM layer in YIG/IrMn film , which indicates that the coupled spins of AFM may exhibit the same behavior as adjacent FM spins. Furthermo re, according to CoFe/IrMn in Fig. 1(c) and Co/YMnO 3 result s, this irreversible of IrMn interfacial magnetization implie s a combination of uniaxial and unidirectional anisotr opy in IrMn interfacial moment at low temperature s. On the other hand, at 7 higher temperatures, the hysteresis and distortion gradually disappear, and the curves are sinusoidal at 300 K, indicating that the IrMn interfacial moment is uniaxial at room temperature. Thus, it can be deduced that the unidirectional anisotropy is the cause of not -fully reversed moment at 10 K, which is analogous to the CoFe/IrMn- 50 Oe curve in Fig. 1(d) . Similar to the behavior of NiFe/IrMn stack ,13 the fast loss of the unidirectional anisotropy is probably due to the decrease of IrMn anisotropy as temperature approaches to the Neel temperature of IrMn .16 Therefore, it can be conjectured that the uncompensated spins at IrMn interface has relatively large coupling with IrMn spins around it at low temperature, which leads to the unidirectional anisotropy. When the temperature increases, this exchange interaction becomes weak and it vanishes at room temp erature. Due to the coupling interaction between FM and AFM, the rotation of interfacial magnetic moment is supposed to be le d or promoted by FM moments. To verify this scenario , the angular dependence of GGG/IrMn (5 nm) sample was papered , where GGG is a non- magnetic insulated substrate of (111) Gd 3Ga5O12. As shown in Fig ure 2(i) and (j) , the sample with GGG substrate shows isotropy of PH R for the field below 2 kOe at both 10 K and 300 K , this proves that the promotion of FM to the reversing of interfacial moment of AFM. Therefore , the sinusoidal curves of YIG/IrMn samples at 300 K suggests that the coupling between of YIG and IrMn still exist s at room temperature , leading to the rotation of some IrMn interfacial spins with uniaxial anisotropy behavior . Besid es, in Fig. 2(d) and (h), the curves of 200 Oe are almost identical to those measured under 2 kOe field, indicating that both IrMn- 1.8 nm and IrMn -5 nm samples have reached saturation under field as low as 200 Oe. This phenomenon also suggests a very small anisotropy of interfacial IrMn moments , which is different to the large unidirectional anisotropy observed in CoFe/IrMn at low magn etic field (Fig. 1(c)) . Whereas there is no solid evidence to explain this contradiction, one possible assumption is that the roughness of IrMn interfaces may be the reason for this difference. Owning to the fact that the series of YIG/IrMn samples are not in- situ grown, which could induce defects and impurities at the interface to cause 8 the interface of YI G become much rougher than that of CoFe. Therefore, a small part of uncompensated spins of the net magnetic moment at IrMn interface may has relatively weak coupling with the bulk than the other s due to the roughness of the interface , resulting in the presentation of a smaller anisotropy, these spins are referred as free spins (F spins) . At low temperature, the F spins exhibit unidirectional anisotropy due to their exchange interaction of the bulk IrMn, which decreases at room temperature, leading to the YIG/IrMn sinusoidal dependence at 300 K. And the rest of uncompensated pinned spins (P spins) of IrMn net magnetic moment still has strong coupling with IrMn since the Neel (blocking) temperature is much higher than the room tem perature , these P spins cannot be rotated or can only be rotated for a small angle under 2 kOe field . And because CoFe/IrMn is grown in-situ, the roughness of the CoFe/IrMn interface should be much smaller than YIG/IrMn interface. Thus the EB effect of CoFe/IrMn are mostly developed by t he coupling between P spins and FM layer , which provides the unidirectional anisotropy for CoFe/IrMn at room temperature. Another evident phenomenon is that both the IrMn- 1.8 nm and IrMn- 5 nm sample s show dramatic deducti on at 90 K and 130 K, as well as the overall trend of low field curve phase shifts 90° as the temperature increases, indicating the shift of easy axis. This phenomenon can be explained by the existence of strongly coupled Fe3+ and weakly coupled Y3+ in YIG, of which Y3+ and the ferrimagnetic component of Fe3+ magnetic moments are antiparallel to each other17. Because the interactions between Y3+ ions are weak, Y3+ shows paramagnetic properties in the exchange field generated by Fe3+ spins. In low temperature region, the total magnetic moments parallel to the Y3+ ions due to its large magnetic moments. As the temperature increases, the magnetic moments of the sub -lattice of Y3+ quickly decrease, leading to the domination of Fe3+ magnetic moments at high temperature. Therefore, the decreasing of PHR at 90 K and 130 K is probably as a result of the competition of the two kinds of moments, and the change of easy axis is because the exchange interaction only exits between Fe 3+ in the high temperature region. Additionally, as shown in Fig ure 2, apart from the larger amplitude of PHR, the YIG/IrMn 9 (1.8 nm) sample shows basically the same behavior of the YIG/IrMn (5 nm) at different temperatures . Since the anisotropy of AFM decreases with decreasing thickness in ultrathin AFM film,18 it can be argued that the spins in IrMn is easier to be reversed by FM, which leads to a larger PHE of YIG/IrMn (1 .8 nm). By a ssuming the interface of IrMn is a single domain film structure , this observed angular dependence of the PHR at 10 K can be described by 2 01 2 3 cos ( ) sin[2( )] cos( )Mj Mj Mj PHR R R R R ϕϕ ϕϕ ϕϕ =+ −+ − + − (3) where 𝜑𝑀 and 𝜑𝑗 are the angle s of IrMn interfacial magnetization and current direction relative to the sample position of zero degree, respectively . The first and second term represent the ARM caused by the misaligning of Hall bar which lead s to a voltage difference along the current direction , while the third and fourth term are classic PH R and an extra PH R generating from unidirectional anisotropy, respectively. The curve of IrMn -1.8 nm and IrMn- 5 nm fitted by Eq. (3) is shown in Figure 3, exhibiting good agreement with the experiment da ta. The parameters obtained from fitting are shown in Table. I . It should be noted that R3 become smaller as the temperature increases, which consists with the decreasing tendency of unidirectional anisotropy observed in Figure 2. TABLE I . Fitting parameters for YIG/IrMn (1.8 nm) and YIG/IrMn (5 nm) film by Equ. (3) at 10 K and 300 K Sample Temperature (K) R0 (ohm) R1 (ohm) R2 (ohm) R3 (ohm) IrMn -1.8nm 10 5.91278 3.93735× 10-11 0.00413 0.00101 300 4.94969 3.023× 10-10 0.00541 0.00013 IrMn -5nm 10 4.3297 -5.47 ×10-4 -0.00171 0.000469 300 3.97303 1×10-5 0.00361 6.09181× 10-5 A further investigation on the IrMn thickness dependence of Hall angle is summarized in Fig. 4(a). T he measurement was performed at 300 K under 200 Oe, and Δρxy and the ρ xx was obtained by the PHR -H curve while the angle between magnetic field and current is 45°. The PHR- H loops of representative samples, IrMn -1.8 nm and IrMn- 5 nm are shown in Fig. 4(b) and 10 (c). Despite the IrMn -1.8 nm sample shows a larger PH R amplitude than IrMn -5 nm sample in Figure 2, it can be observed in Fig. 4(a) that the Hall angle almost remains constant when IrMn is not thicker than 5 nm. When the thickness of IrMn is larger than 5 nm, the Hall angle begin to decrease dramatically and reaches a constant as IrMn is thicker than 10 nm. This confirms the interface nature of PHE in IrMn, and th is interfacial effect does not quickly decrease until IrMn is not thick enough. IX. CONCLUSION In conclusion, the PHR of YIG/IrMn samples with different IrMn thickness es were investigated in this work. The magnetic f ield angle dependence of PHR can be observed from 10 K to 300 K below 2 kOe , suggesting the existence of uncompensated spins at YIG/IrMn interface. And also a hysteresis derived from unidirectional anisotropy can be observed in the PHR- θ curves at 10 K in YIG/IrMn films . By comparing with GGG/IrMn sample, which PHR- θ curve shows isotropic behavior at both 10 K and 300 K, it can be considered that the interfacial IrMn spins are led by the magnetic moment of YIG. ACKNOWLEDGEMENTS This work was supported by the National Key Basic Research Program of China under Grant No. 2014CB921002, and the National Natural Science Foundation of China under Grant Nos.51171205 and 11374349. REFERENCES 1 C. Chappert, A. Fert, and F. N. Van Dau, Nature Mater. 6, 813 (2007). 2 W. H. Meiklejohn and C. P . Bean, Phys. Rev. 105, 904 (1957). 3 J. Barzola -Quiquia, A. Lessig, A. Ballestar, C. Zandalazini, G. Bridoux, F. Bern, and P . Esquinazi, J. Phys. : Condens. M atter 24, 366006 (2012). 4 J. Sort, V. Baltz, F. Garcia, B. Rodmacq, and B. Dieny, Phys. Rev. B 71, 054411 (2005). 5 Y . F. Liu, J. W. Cai, and a. S. L. He, J. Phys. D: Appl. Phys. 42, 115002 (2009). 11 6 H. Ohldag, A. Scholl, F. Nolting, E. Arenholz, S. Maat, A. T. Young, M. Carey, and J. Stöhr, Phys. Rev. Lett. 91, 017203 (2003). 7 S. Doi, N. Awaji, K. Nomura, T. Hirono, T. Nakamura, and H. Kimura, Appl. Phys. Lett. 94 (2009). 8 H. Takahashi, Y . Kota, M. Tsunoda, T. Nakamura, K. Kodama, A. Sakuma, and M. Taka hashi, J. Appl. Phys. 110 (2011). 9 G. Li, Z. Lu, C. Chai, H. Jiang, and W. Lai, Appl. Phys. Lett. 74, 747 (1999). 10 G. Li, T. Yang, Q. Hu, H. Jiang, and W. Lai, Phys. Rev. B 65, 134421 (2002). 11 T. R. McGuire and R. I. Potter, IEEE Trans. Magn. 11, 1018 (1975). 12 B. G. Park, J. Wunderlich, X. Martí, V. Holý, Y . Kurosaki, M. Yamada, H. Yamamoto, A. Nishide, J. Hayakawa, H. Takahashi, A. B. Shick, and T. Jungwirth, Nature Mater. 10, 347 (2011). 13 X. Martí, B. G. Park, J. Wunderlich, H. Reichlová, Y . Kurosaki, M. Yamada, H. Yamamoto, A. Nishide, J. Hayakawa, H. Takahashi, and T. Jungwirth, Phys. Rev. Lett. 108, 017201 (2012). 14 Y . Y . Wang, C. Song, B. Cui, G. Y . Wang, F. Zeng, and F. Pan, Phys. Rev . Lett. 109, 137201 (2012). 15 Y . M. Lu, Y . Choi, C. M. Ortega, X. M. Cheng, J. W. Cai, S. Y . Huang, L. Sun, and C. L. Chien, Phys. Rev. Lett. 110, 147207 (2013). 16 G. Vallejo -Fernandez, L. E. Fernandez -Outon, and K. O’Grady, Appl. Phys. Lett. 91 (2007). 17 L. Neel, Comp. Rend. 239, 8 (1954). 18 Y . Xu, Q. Ma, J. W. Cai, and L. Sun, Phys. Rev. B 84, 054453 (2011). Figure captions: FiG. 1. (a) The in-plane (top panel) and out -of-plane (bottom panel) hysterias loops of YIG, and (b) the in-plane hysterias loop of CoFe (8 nm)/IrMn (10 nm), and (c) the PHR-θ curve of CoFe measured under 2 kOe magnetic field at room temperature, and (d) the PHR -θ of CoFe (8 nm)/IrMn (10 nm) measured under different magnetic fields at room temperature. FIG. 2. PHR of (a -d) YIG/IrMn (1.8 nm) and (e -h) YIG/IrMn (5 nm) sample when rotating in the applied magnetic field of 2 kOe at different temperatures. The measurements were also performed under 200 Oe for both YIG/IrMn samples at 300 K. (i) and (j) are the magnetic field angu lar dependence of GGG/IrMn (5 nm) film at 10 K and 300 K, respectively. At each temperature, the PHR -θ curves were measured under different magnetic fields of 200 Oe and 2 kOe, both of which exhibit no obvious variation with magnetic field angular θ. FIG. 3. PHR -θ cure and its fitting curve of YIG/IrMn (1.8 nm) sample at (a) 10 K and (b) 300 K, and PHR -θ cure and its fitting curve for YIG/IrMn (5 nm) sample at (c) 10 K and (d) 300 K. The fitting curve uses the data of magnetic field clockwisely rotating from 0° to 360°. The data of 10 K and 300 K are fitted by Equ. (3). FIG. 4. (a) The IrMn thickness dependence of Hall angle, and the PHR-H curve for (b) YIG/IrMn (1.8 nm) and (c) YIG/IrMn (5 nm) measured at 300 K. The angle between magnetic field and curre nt is 45°. Figures: 12 FIG.1. 13 FIG. 2. FIG. 3. 14 FIG. 4. | 2014-10-05 | The planar Hall effect of IrMn on an yttrium iron garnet (YIG = Y3Fe5O12) was
measured in the magnetic field rotating in the film plane. The magnetic field
angle dependence of planar Hall resistance (PHR) has been observed in YIG/IrMn
bilayer at different temperatures, while the GGG/IrMn (GGG= Gd3Ga5O12) shows
constant PHR for different magnetic field angles at both 10 K and 300 K. This
provides evidence that IrMn has interfacial spins which can be led by FM in
YIG/IrMn structure. A hysteresis can be observed in PHR-magnetic field angle
loop of YIG/IrMn films at 10 K, indicating the irreversible switching of IrMn
interfacial spins at low temperature. | Planar Hall effect in Y3Fe5O12(YIG)/IrMn films | 1410.1112v1 |
1 Thermally Driven Long Range Magnon Spin Currents in Yttrium Iron Garnet due to Intrinsic Spin Seebeck Effect Brandon L. Giles1, Zihao Yang2, John S. Jamison1, Juan M. Gomez -Perez4, Saül Vélez4, Luis E. Hueso4,5, Fèlix Casanova4,5, and Roberto C. Myers1-3 1Department of Materials Science and Engineering, The Ohio State University, Columbus, OH, 43210, USA 2Department of Electrical and Computer Engineering, The Ohio State University, Columbus, OH, 43210, USA 3 Department of Physics, The Ohio State University, Columbus, OH, USA 4CIC nanoGUNE, 20018 Donostia- San Sebastian, Basque Country, Spain 5IKERBASQUE, Basque Foundation for Science, 40813 Bilbao, Basque Country, Spain Email: myers.1079@osu.edu , Web site: http://myersgroup.engineering.osu.edu The longitudinal spin Seebeck effect refers to the generation of a spin current when heat flows across a normal metal/magnetic insulator interface. Un til recently, most explanations of the spin Seebeck effect use the interfacial temperature difference as the conversion mechanism between heat and spin fluxes. However, recent theoretical and experimental works claim that a magnon spin current is generated in the bulk of a magnetic insulator even in the absence of an interface. This is the so -called intrinsic spin Seebeck effect. Here, by utilizing a non -local spin Seebeck geometry, we provide additional evidence that the total magnon spin current in the ferrimagnetic insulator yttrium iron g arnet (YIG ) actually contains two distinct terms: one proportional to the gradient in the magnon chemical potential (pure magnon spin diffusion) , and a second proportional to the gradient in magnon 2 temperature (𝛁𝛁𝑻𝑻𝒎𝒎). We observe two characteristic decay lengths for magnon spin currents in YIG with distinct temperature dependences: a temperature independent decay length of ~ 10 𝝁𝝁m consistent with earlier measurements of pure ( 𝛁𝛁𝑻𝑻𝒎𝒎=𝟎𝟎) magnon spin diffusion, and a long er decay length ranging from about 20 𝝁𝝁m around 250 K and exceeding 80 𝝁𝝁m at 10 K. The coupled spin- heat transport processes are modeled using a finite element method revealing that the longer range magnon spin current is attributable to the intrinsic spin Seebeck effect (𝛁𝛁𝑻𝑻𝒎𝒎≠𝟎𝟎), whose length scale increases at lower temperatures in agreement with our experimental data . Recently, s ignificant effort s have focused on understanding magnon spin diffusion arising from the spin Seebeck effect [1,2] . In particular , the effective magnon spin diffusion length in YIG has been experimentally measured using many different methods, including the systematic variation of YIG sample thickness to observe the effect on the longitudinal spin Seebeck signal [3 – 5], and by the use of a non- local geometry to directly measure the magnon spin diffusion length of electrically and thermally excited magnons [6–8]. Both methods demonstrated that the magnon spin diffusion length in YIG is only minimally dependent on film thickness and also that the magnon spin diffusion length is around 10 𝜇𝜇m at low temperatures. However, the studies report contradictory results near room temperature. T he thickness dependence study carried out by Kehlberger et. al. [3] found that the magnon spin diffusion length gradually decreases from 10 to 1 𝜇𝜇m as the temperature is increased to room temperature, while the non- local measurement carried out by Cornelissen et. al. [7] found that the magnon spin diffusion length is only very slightly dependent on temperature . These discrepancies might be expected due to variation in the temperature profile between experiments with different sample sizes and geometries, and the variation in the relative impact of the intrinsic (bulk) spin Seebe ck effect. The need to include the se 3 bulk temperature gradient driven magnon currents to fully explain room temperature nonlocal spin transport in thin film YIG has recently been discussed in detail in Ref. [8]. In this Rapid Communication, we further demonstrate the central role of the intrinsic spin Seebeck effect in the generation of long -range spin signals in bulk YIG that emerge at low temperatures . For this purpose, we carry out two independent experiment s to measure diffusive magnon spin currents in bulk single crystal YIG as a function of temperature using the nonlocal opto- thermal [9] and the nonlocal electro -thermal [6] techniques . For both measurements , magnons carrying spin angular momentum are thermally excited beneath a Pt injector resulting in a measureable voltage induced in an electrically isolated Pt spin detect or. In both the opto- thermal and electro -thermal measureme nts, two independent magnon spin current decay lengths are observed. The shorter decay length ~ 10 𝜇𝜇m is roughly temperature independent and in agreement with Cornelissen et al . [6]. In addition to this shorter decay length, we also identify a longer range magnon spin decay length at lower temperatures that reaches values in excess of 80 𝜇𝜇m at 10 K . The longer magnon spin decay length originates from magnon s generat ed by heat flow within the bulk YIG itself, and represents the intrinsic spin Seebeck effect. Finite element modeling (FEM) is used to solve coupled spin- heat transport equations in YIG that describe both the pure magnon spin diffusion that is driven by a gradient in the magnon chemical potential, ∇𝜇𝜇𝑚𝑚, and also the magnon spin current that is driven by a thermal gradient in the YIG itself, ∇𝑇𝑇𝑚𝑚. Microscope image s of typical devices used for opto- thermal measurements and electro - thermal measurements are shown in Fig. 1(a) and Fig. 1(c) . The opto- thermal device consists of 10 nm of Pt that was sputter deposited onto a 500 𝜇𝜇m <100> single crystal YIG that was purchased commercially from Princeton Scientific . Standard lithography techniques we re used to pattern the Pt into a 50×50 𝜇𝜇m detection pad surrounded by electrically isolated 5× 5 𝜇𝜇m injector pads with 3 4 𝜇𝜇m between them . The electro -thermal device consists of 5 nm of Pt that was sputter deposited onto a 500 𝜇𝜇m <100> single crystal YIG from the same wafer . Each electro -thermal device was fabricated via high-resolution e-beam lithography using a negative resist and Ar -ion milling to pattern one Pt injector and two Pt detectors (width W = 2.5 µm and length L = 500 𝜇𝜇 m). Injector - detector distances range from 12 to 100 𝜇𝜇 m. FIG 1. Optical images of the devices used in the opto- thermal and electro -thermal measurements. (a) In the opto- thermal measurement, a laser is used to thermally excite magnons in YIG beneath a Pt injector. The magnons diffuse laterally and are con verted into a measureable voltage in the Pt detector. (b) A typical hysteresis loop showing the measured voltage as a function of magnetic field. 𝑉𝑉𝑁𝑁𝑁𝑁,𝑂𝑂 is defined as the magnitude of the hysteresis loop. (c) In the electro -thermal measurement, current flowing through the injector causes resistive heating, resulting in the excitation of magnons into YIG. The non- equilibrium magnons produced diffuse to the re gion beneath a non- local Pt detector, where can be detected due to the inverse spin Hall voltage induced. (d) The measured voltage depends sinusoidally on the angle α of the applied in- plane magnetic field. The maximum detected voltage is defined as 𝑉𝑉 𝑁𝑁𝑁𝑁,𝐸𝐸. 𝑑𝑑 represents the distance the magnons have diffused from the injection to the detection site. In the opto- thermal experiment a diffraction -limited 980-nm-wavelength laser is used to thermally excite magnons beneath a Pt injector whose center is located at a distance d from the closest edge of the Pt detector . The experiments were carried out in a Montana Instruments C2 5 cryostat at temperatures between 4 and 300 K. The laser is modulated at 10 Hz and a lock- in amplifier referenced to the laser choppin g frequency is used to measure t he inverse spin Hall effect voltage , defined as 𝑉𝑉𝐼𝐼𝐼𝐼𝐼𝐼𝐸𝐸 ,𝑂𝑂, across the detector . An in -plane magnetic field is applied along the x axis and is swept from - 200 mT to 200 mT while 𝑉𝑉𝐼𝐼𝐼𝐼𝐼𝐼𝐸𝐸 ,𝑂𝑂 is continuously recorded. A representative hysteresis loop taken at 89.5 K and for d = 21 𝜇𝜇m is shown in Fig. 1(b). The detector signal proportional to nonlocal magnon spin diffusion, defined as 𝑉𝑉𝑁𝑁𝑁𝑁,𝑂𝑂, is obtained by taking half the difference between saturated 𝑉𝑉𝐼𝐼𝐼𝐼𝐼𝐼𝐸𝐸 ,𝑂𝑂 values at positive and negative fields, i.e. the height of the hysteresis loop. For the electro -thermal experiment, m agnetotransport measurements were carried out using a Keithley 6221 sourcemeter and a 2182A nanovoltmeter operating in delta mode . In contrast to the standard current -reversal method, where one obtains information about the electrically excited magnons in devices of this kind [10] , here a dc- pulsed method is used where the app lied current is continuously switched on and off at a frequency of 20 Hz. This measurement provides equivalent information as the second harmonic in ac lock- in type measurements [11] , i.e., it provides information about the thermally excited magnons. A current of I = 300 µA was applied to the injector. The experiments were carried out in a liquid- He cryostat at temperatures between 2.5 and 10 K. A magnetic field of H = 1 T was applied in the plane of the sample and rotated (defined by the angle 𝛼𝛼 ) while the resulting voltage VISHE,E was measured in one of the detectors . Fig. 1 (d) shows a representative measurement . The signal obtained is proportional to sin 𝛼𝛼 , which is indicative of the diffusive magnon spin current [12] . The magnitude of the signal is defined as 𝑉𝑉𝑁𝑁𝑁𝑁,𝐸𝐸 [see Fig. 1 (b)]. The magnon spin current decays exponentially with d [13] . Therefore, the VNL measured in our devices is given by 6 𝑉𝑉𝑁𝑁𝑁𝑁= 𝐴𝐴𝑜𝑜𝑒𝑒−𝜆𝜆𝑆𝑆∗ 𝑑𝑑, (1) where A0 is a pre- factor that is independent of d and 𝜆𝜆𝐼𝐼∗, is the effective magnon spin diffusion length . The experimental data obtained for both the opto- thermal and the electro -thermal magnon spin excitation are shown in Fig. 2 and analyzed using Eq. (1). At high temperatures, the data fits very well to a single exponential as expected. Surprisingly , at low temperatures, the fit analysis reveals that there must actually be two different decay lengths. For inst ance, for the opto- thermal case, it is observed that the quality of the fit rapidly decreases below a correlation coefficient of r 2=0.985 when the distances considered range from the smallest measured (5.5 𝜇𝜇m) to greater than 37.5 𝜇𝜇m. This indicates that the application of the spin decay model is only appropriate up to 37.5 𝜇𝜇 m. If distances greater than 37.5 𝜇𝜇 m are considered and the data is fit to Eq. (1) , a lower r2 factor is obtained, indicating a low quality fit . This observation inspires us to separate the 𝑉𝑉𝑁𝑁𝑁𝑁,𝑂𝑂 data into two distinct regions defined as the 𝜆𝜆1 and 𝜆𝜆2 region s [see Fig. 2( a)]. Equation (1) is fit to each individual region. Th e effective magnon spin diffusion length 𝜆𝜆𝐼𝐼∗ is extracted for each region separately and plotted in Fig. 3 . The same FIG 2. (a) 𝑉𝑉𝑁𝑁𝑁𝑁,𝑂𝑂 as a function of 𝑑𝑑 with the measurement shown at different temperatures. The measurement results are divided into two regions defined as 𝜆𝜆1 and 𝜆𝜆2. Dotted lines represent single exponential fits of the data to Eq. (1) in each region. The decay in 𝜆𝜆1 is shorter, while it appears to be much longer in 𝜆𝜆2. (b) 𝑉𝑉𝑁𝑁𝑁𝑁,𝐸𝐸 as a function of 𝑑𝑑 with the measurement shown at multiple temperatures. Dividing the data also into the 𝜆𝜆1 and 𝜆𝜆2 regions confirms the existence of the two different characteristic decay lengths. Dashed li nes are fits to Eq. (1) in each region. 7 analysis was performed for the electro - thermal measurements and the existence of two different decay lengths was confirmed (See Fig . 2(b)). Fig. 3 shows the extracted values of the magnon spin diffusion lengths in each of the two regions as a function of temperature for both the opto- thermal and electro -thermal measurements. At low temperature, both measurements indicate an effective spin diffusion length of about 10 𝜇𝜇m in the 𝜆𝜆 1 region, which is in excellent agreement with p reviously reported values and temperature dependence of the magnon spin diffusion length [7] . Note that in the earlier opto -thermal study [9] the data indicated only a single exponential decay, which was interpreted as the spin diffusion length. In the opto- thermal measurements reported here, the improved signal to noise ratio of the experiment reveals the double exponential character of the spin decay profile. The current data can still be fitted to a single exponential decay at 23 K of 47 µm, consistent with the earlier report, however the improved data set in the current study demonstrates that a double exponential decay fit is far better quality. A larger 𝜆𝜆 𝐼𝐼∗ in the 𝜆𝜆 2 region is observed in both the opto- thermal and electro -thermal measurements . At temperatures above 10 K in the electro -thermal measurement, the non- local signal magnitude strongly decreased and could not be measured at enough values of d in order to make a meaningful exponential fit to extract 𝜆𝜆𝐼𝐼∗ in the 𝜆𝜆 2 region . The effective magnon spin FIG 3. The extracted decay parameters 𝜆𝜆𝐼𝐼∗ from the 𝜆𝜆1 and 𝜆𝜆2 regions as a function of temperature and for both experiments. 𝜆𝜆𝐼𝐼∗ values reported in Ref. 7 are included for comparison. Inset: zoomed view of low temperature data. 8 diffusion length in the 𝜆𝜆2 region is approximately one order of magnitude larger than in the 𝜆𝜆 1 region at low temperatures and decreases monotonically with increasing temperature. The maximum value of 83.03 𝜇𝜇 m occurs at 9.72 K and the minimu m value of 14.05 𝜇𝜇 m at 247.5 K. A zoom of the data at low T is shown in the inset to Fig. 3. In the electro -thermal measurements, the maximum value of 𝜆𝜆2 is not at the lowest temperature, but at ~10 K in agreement with the optothermal measurements . This is consistent with the origin of 𝜆𝜆2 as from intrinsic SSE associated with the temperature profile in YIG since as T approaches 0 K, thermal conductivity becomes negligible . To justify the existence of the long range spin current persisting well beyond the intrinsic magnon spin diffusion length, t he measurements are compared to a simulation of the diffusive transport of thermally generated magnons , which is obtained using three dimensional (3D) finite element modeling (FEM). The simulation is solved using COMSOL Multiphysics and is based on the spin and heat transport formalism that is developed in [14,15] . In the simulation , the length scale of the inelastic phonon and magnon scattering is assumed to be small, implying that the phonon temperature , 𝑇𝑇𝑝𝑝, is equal to the magnon temperature 𝑇𝑇𝑚𝑚 over the length s of interest . In addition, the simulation neglects the spin Peltier effect . Thus, the spin and heat transport equations are only partially coupled. The simplified spin transport equation that is used to model the magnon spin current within YIG is 𝜎𝜎∇2𝜇𝜇+ 𝜍𝜍∇2𝑇𝑇=𝑔𝑔𝜇𝜇 (2) and t he Pt/ YIG interfacial boundary condition states 𝑗𝑗𝑚𝑚,𝑧𝑧=𝜎𝜎∇𝜇𝜇𝑧𝑧+𝜍𝜍∇𝑇𝑇𝑧𝑧= 𝐺𝐺𝐼𝐼𝜇𝜇 (3) 9 where 𝑗𝑗𝑚𝑚,𝑧𝑧 is the simulated spin current perpendicular to the Pt/YIG interface, 𝜎𝜎 is the spin conductivity in the YIG , 𝜇𝜇 is the magnon chemical potential, 𝜍𝜍 is the intrinsic spin Seebeck coefficient, 𝑔𝑔 describes the magnon relaxation, 𝑇𝑇=𝑇𝑇𝑝𝑝~𝑇𝑇𝑚𝑚 is the temperature in YIG , 𝐺𝐺𝐼𝐼 is the interfacial magnon spin conductance, and ∇𝜇𝜇𝑧𝑧 and ∇𝑇𝑇𝑧𝑧 represent the gradient of the magnon chemical potential and temperature along the direction perpendicular to the Pt/YIG interface , respectively. We first solve for the temperature profile in a simulated Pt/YIG system using the parameters listed in Table I. The geometry of the model is the same as the experimental geometry of the opto- thermal measurement including the Pt absorbers . As previously stated, d is defined as the distance from the edge of the Pt detector to the center of the (simulated) laser heat source at the center of the absorber . Table I – Parameters used in the 3D FEM modeling. 𝜎𝜎 and 𝐺𝐺𝐼𝐼 are calculated based on data reported in [15] . 𝜅𝜅𝑌𝑌𝐼𝐼𝑌𝑌 is taken from [19] and 𝜅𝜅𝑃𝑃𝑃𝑃 is from [20] . 𝑇𝑇(K) 𝜎𝜎(JmV⁄ ) 𝐺𝐺𝐼𝐼(Sm2⁄ ) 𝜅𝜅𝑌𝑌𝐼𝐼𝑌𝑌 (WmK)⁄ 𝜅𝜅𝑃𝑃𝑃𝑃 (WmK)⁄ 10 3.10×10−8 5.84×1010 60.00 1214.98 70 8.32×10−8 1.08×1012 37.59 91.82 175 1.32×10−7 4.27×1012 11.41 75.56 300 1.73×10−7 9.60×1012 6.92 73.01 The decay profile for the interfacial spin current 𝑗𝑗𝑚𝑚,𝑧𝑧 is obtained by using the calculated temperature profile as an input in Eq. (3) . We report the total interfacial spin current that reaches the detector 𝑗𝑗𝑚𝑚,𝑧𝑧 by evaluating the surface integral ∬𝑗𝑗𝑚𝑚,𝑧𝑧(𝑥𝑥,𝑦𝑦)𝑑𝑑𝐴𝐴 beneath the detector. The decay profile is calculated as a function of simulated laser position , at multiple different temperatures, 10 ranging from 5 – 300 K. The values of the physical parameters used in the model are recorded in Table I. From Eq. (3) one can see that 𝒋𝒋𝑚𝑚,𝑧𝑧 can be broken up into two components 𝑗𝑗𝑚𝑚,𝑧𝑧 ∇𝜇𝜇, which is a component that is proportional to the interfacial gradient of the magnon chemical potential , and 𝑗𝑗𝑚𝑚,𝑧𝑧 ∇𝑇𝑇, which is a component that is proportional to the interfacial gradient of the magnon temperature. The decomposition of the simulated spin current at the detector is shown in Fig. 4(a) , which depicts a representative plot of the total 𝒋𝒋𝑚𝑚,𝑧𝑧 as a funct ion of 𝑑𝑑 at 70 K , as well as the components 𝑗𝑗𝑚𝑚,𝑧𝑧 ∇𝜇𝜇 and 𝑗𝑗𝑚𝑚,𝑧𝑧 ∇𝑇𝑇 . By analyzing the decay lengths of these individual components of 𝒋𝒋𝑚𝑚,𝑧𝑧 separately , it is possible to qualitatively understand the existence of the experimentally observed short and long range decay lengths . As shown in Fig. 4(a), the component of 𝒋𝒋𝑚𝑚,𝑧𝑧 that is proportional to ∇ 𝜇𝜇 decays much more rapidly than the component of 𝒋𝒋𝑚𝑚,𝑧𝑧 that is proportional to ∇ 𝑇𝑇. This indicates that the total spin current that reaches the Pt det ector should consist of a short er decay component and a long er decay component. We hypothesize that the driving force of the short er range component is the gradient of the magnon chemical potential, ∇𝜇𝜇 and that the driving force of the long er range component is the gradient of the magnon temperature ∇𝑇𝑇. To verify this conjecture, the plot of the simulated 𝒋𝒋𝑚𝑚,𝑧𝑧 vs. 𝑑𝑑 is divided into the same 𝜆𝜆 1 and 𝜆𝜆2 regions as in the opto- thermal experimental measurement (where the 𝜆𝜆2 region is defined as 𝑑𝑑 > 37.5 𝜇𝜇m). Equation ( 1) is fit independently to the simulated 𝑗𝑗𝑚𝑚,𝑧𝑧 ∇𝜇𝜇 within the 𝜆𝜆 1 region , where the short er range driving force is expected to dominate , and to the simulated 𝑗𝑗𝑚𝑚,𝑧𝑧 ∇𝑇𝑇 with in the 𝜆𝜆 2 region where the long er range driving force will be most prevalent , as shown in the representative 70 K plot in Fig. 4(a). The decay parameters of these fits, 𝜆𝜆∇𝜇𝜇∗ and 𝜆𝜆∇𝑇𝑇∗, are extracted and plotted as a function of temperature 11 in Fig. 4(b). The intrinsic spin diffusion length, 𝜆𝜆∇𝜇𝜇∗, is relatively constant as a function of temperature, implying that ∇𝜇𝜇 is responsible for the short er range spin current observed in the 𝜆𝜆1 region (Fig. 3) . On the other hand, the bulk generated magnon current, characterized by 𝜆𝜆∇𝑇𝑇∗, decays monotonically with temperature, in agreement with the observed long er decay in the 𝜆𝜆2 region (Fig. 3) , thus implying that ∇𝑇𝑇 is the driving force for the long range spin current . Since it is the temperature profile within YIG that determines 𝜆𝜆∇𝑇𝑇∗, it will vary with the thermal boundary conditions. This explains why the long range spin current manifests in bulk YIG at low temperatu re [9], but not in YIG/GGG thin films [7]. It should be noted that while the monotonic decay with temperature of the simulated 𝜆𝜆 ∇𝑇𝑇∗ agrees with the measured opto- thermal and electro -thermal long range decay in the λ2 region , the simulated magnitude of 𝜆𝜆∇𝑇𝑇∗ is smaller than the one obtained experimentally . This is attributed to uncertainties in the temperature dependence of the inputs to the FEM modeling, particularly of the magnon scattering time 𝜏𝜏 , which is used to calcul ate 𝜎𝜎𝑚𝑚. At low temperatures magnon relaxation FIG 4. 3D FEM modeling simulation of the opto-thermal measurement. (a) Dashed lines represent the total spin current (black), the component of spin current proportional to ∇𝜇𝜇 (green) and the component of spin current proportional to ∇𝑇𝑇 (pink). Solid lines represent individual exponential fits to the corresponding component of the spin current in each of the distinct 𝜆𝜆1 and 𝜆𝜆2 regions (blue and red respectively). (b) The magnon spin diffusion lengths 𝜆𝜆∇𝜇𝜇∗ and 𝜆𝜆∇T∗ extracted for each region are plotted as a function of temperature. 12 is primarily governed by magnon- phonon interactions that create or annihilate spin waves by magnetic disorder and 𝜏𝜏 ~ ℏ𝛼𝛼𝑌𝑌𝑘𝑘𝐵𝐵𝑇𝑇⁄ where 𝛼𝛼𝑌𝑌= 10−4 [16] . This leads to calculated values of 𝜎𝜎𝑚𝑚 that vary with experimental measurements by orders of magnitude [15] . Such discrepancies may be explained by recent works that attribute the primary contributors to the SSE as low -energy subthermal magnons [5,17] , however an analysis of the complete temperature dependence of effective magnon scattering time based on the spectral dependence of the dominant magnons involved in SSE is outside the scope of this work. Another source of uncertainty in the simulations is the role of spin sinking into the Pt absorbers (present in the opto- thermal measurements) on the spin current decay profile . To test this, identical simulations, as described above, are carried out but with the Pt absorber pads removed. The absorbers cause a decrease in 𝜆𝜆∇𝜇𝜇∗ of 1-2 µm , while the 𝜆𝜆∇𝑇𝑇∗ shows no significant change within the uncertainty. During the review of this paper, we became aware of a related paper discussing the role of intrinsic spin Seebeck in the nonlocal spin currents decay profile [18] . In conclusion, opto- thermal and electro -thermal measurements independently demonstrate the existence of a longer range magnon spin current at low temperatures persisting well beyond the intrinsic spin diffusion length. By representing the total magnon spin current by its individual components , one of which is proportional to the gradient in magnon chemical potential and the other of which is proportional to the gradient in magnon temperature, the driving force of the longer range magnon spin diffusion can be attributed to the gradient in magnon temperature , i.e. the intrinsic spin Seebeck effect . The authors thank B art van Wees, Ludo Cornel issen, Yaroslav Tserkovnyak and Benedetta Flebus for valuable discussions. This work was primarily supported by the Army Research Office MURI 13 W911NF -14-1-0016. J.J. acknowledges the Center for Emergent Materials at The Ohio State University, an NSF MRSEC (Award Number DMR -1420451), for providing partial funding for this research. The work at CIC nanoGUNE was supported by the Spanish MINECO (Project No. MAT2015- 65159- R) and by the Regional Council of Gipuzkoa (Project No. 100/16). J.M.G.- P. thanks the Spanish MINECO for a Ph.D. fellowship (Grant No. BES -2016- 077301). [1] K. Uchida, M. Ishida, T. Kikkawa, A. Kirihara, T. Murakami, and E. Saitoh, J. Phys. Condens. Matter 26, 343202 (2014). [2] A. Prakash, J. Brangham, F. Yang, and J. P. Heremans, Phys. Rev. B 94, 014427 (2016). [3] A. Kehlberger, U. Ritzmann, D. Hinzke, E.- J. Guo, J. Cramer, G. Jakob, M. C. Onbasli, D. H. Kim, C. A. Ross, M. B. Jungfleisch, B. Hillebrands, U. Nowak, and M. Kläui, Phys. Rev. Lett. 115, 096602 ( 2015). [4] E.-J. Guo, J. Cramer, A. Kehlberger, C. A. Ferguson, D. A. MacLaren, G. Jakob, and M. Kläui, Phys. Rev. X 6, 031012 (2016). [5] T. Kikkawa, K. Uchida, S. Daimon, Z. Qiu, Y. Shiomi, and E. Saitoh, Phys. Rev. B 92, 064413 (2015). [6] L. J. Cornelissen, J. Liu, R. A. Duine, J. B. Youssef, and B. J. van Wees, Nat Phys 11, 1022 (2015). [7] L. J. Cornelissen, J. Shan, and B. J. van Wees, Phys. Rev. B 94, 180402 (2016). [8] J. Shan, L. J. Cornelissen, N . Vlietstra, J. Ben Youssef, T. Kuschel, R. A. Duine, and B. J. van Wees, Phys. Rev. B 94, 174437 (2016). [9] B. L. Giles, Z. Yang, J. S. Jamison, and R. C. Myers, Phys. Rev. B 92, 224415 (2015). [10] S. Vélez, A. Bedoya -Pinto, W. Yan, L. E. Hueso, and F. Casanova, Phys. Rev. B 94, 174405 (2016). [11] F. L. Bakker, A. Slachter, J.- P. Adam, and B. J. van Wees, Phys. Rev. Lett. 105, 136601 (2010). [12] S. R. Boona, R. C. Myers, and J. P. Heremans, Energy Environ. Sci. 7, 885 (2014). [13] T. Valet and A. Fert, Phys. Rev. B 48, 7099 (1993). [14] B. Flebus, S. A. Bender, Y. Tserkovnyak, and R. A. Duine, Phys. Rev. Lett. 116, 117201 (2016). [15] L. J. Cornelissen, K. J. H. Peters, G. E. W. Bauer, R. A. Duine, and B. J. van Wees, Phys. Rev. B 94, 014412 (2016). [16] S. Hoffman, K. Sato, and Y. Tserkovnyak, Phys. Rev. B 88, 064408 (2013). [17] I. Diniz and A. T. Costa, New J. Phys. 18, 052002 (2016). [18] J. Shan, L. J. Cornelissen, J. Liu, J. B. Youssef, L. Lian g, and B. J. van Wees, ArXiv170906321 Cond- Mat (2017). [19] S. R. Boona and J. P. Heremans, Phys. Rev. B 90, 064421 (2014). [20] J. E. Jensen, W. A. Tuttle, H. Brechnam, and A. G. Prodell, Brookhaven National Laboratory Selected Cryogenic Data Notebook (Brookhaven National Laboratory, New York, 1980). | 2017-08-06 | The longitudinal spin Seebeck effect refers to the generation of a spin
current when heat flows across a normal metal/magnetic insulator interface.
Until recently, most explanations of the spin Seebeck effect use the
interfacial temperature difference as the conversion mechanism between heat and
spin fluxes. However, recent theoretical and experimental works claim that a
magnon spin current is generated in the bulk of a magnetic insulator even in
the absence of an interface. This is the so-called intrinsic spin Seebeck
effect. Here, by utilizing a non-local spin Seebeck geometry, we provide
additional evidence that the total magnon spin current in the ferrimagnetic
insulator yttrium iron garnet (YIG) actually contains two distinct terms: one
proportional to the gradient in the magnon chemical potential (pure magnon spin
diffusion), and a second proportional to the gradient in magnon temperature
($\nabla T_m$). We observe two characteristic decay lengths for magnon spin
currents in YIG with distinct temperature dependences: a temperature
independent decay length of ~ 10 ${\mu}$m consistent with earlier measurements
of pure ($\nabla T_m = 0$) magnon spin diffusion, and a longer decay length
ranging from about 20 ${\mu}$m around 250 K and exceeding 80 ${\mu}$m at 10 K.
The coupled spin-heat transport processes are modeled using a finite element
method revealing that the longer range magnon spin current is attributable to
the intrinsic spin Seebeck effect ($\nabla T_m \neq 0$), whose length scale
increases at lower temperatures in agreement with our experimental data. | Thermally Driven Long Range Magnon Spin Currents in Yttrium Iron Garnet due to Intrinsic Spin Seebeck Effect | 1708.01941v3 |
arXiv:1603.03578v2 [cond-mat.str-el] 11 Apr 2016Investigation of anomalous-Hall and spin-Hall effects of an tiferromagnetic IrMn sandwiched by Pt and YIG layers T. Shang,1,a)H. L. Yang,1Q. F. Zhan,1,b)Z. H. Zuo,1Y. L. Xie,1L. P. Liu,1S. L. Zhang,1Y. Zhang,1H. H. Li,1B. M. Wang,1Y. H. Wu,2S. Zhang,3,c)and Run-Wei Li1,d) 1)Key Laboratory of Magnetic Materials and Devices &Zhejiang Province Key Laboratory of Magnetic Materials and Application Technology, Ningbo Institute of Material Tech nology and Engineering, Chinese Academy of Sciences, Ningbo, Zhejiang 315201, China 2)Department of Electrical and Computer Engineering, Nation al University of Singapore, 4 Engineering Drive 3 117583, Singapore 3)Department of Physics, University of Arizona, Tucson, Ariz ona 85721, USA (Dated: 14 June 2021) We report an investigation of temperature and IrMn layered thickn ess dependence of anomalous-Hall resistance (AHR), anisotropic magnetoresistance (AMR), and ma gnetization on Pt/Ir 20Mn80/Y3Fe5O12 (Pt/IrMn/YIG) heterostructures. The magnitude of AHR is dram atically enhanced compared with Pt/YIG bilayers. The enhancement is much more profound at higher temper atures and peaks at the IrMn thickness of 3 nm. The observed spin-Hall magnetoresistance (SMR) in the te mperature range of 10-300 K indicates that the spin current generated in the Pt layer can penetrate the entire thickness of the IrMn layer to interact with the YIG layer. The lack of conventional anisotropic magnetore sistance (CAMR) implies that the inser- tion of the IrMn layer between Pt and YIG efficiently suppresses the magnetic proximity effect (MPE) on induced Pt moments by YIG. Our results suggest that the dual role s of the IrMn insertion in Pt/IrMn/YIG heterostructures are to block the MPE and to transport the spin current between Pt and YIG layers. We discuss possible mechanisms for the enhanced AHR. I. INTRODUCTION Antiferromagnts (AFMs) are promising candidates for spintronic applications.1Compared to ferromagnetic (FM) materials, the AFMs exhibit unique advantages, e.g., zero net magnetization, insensitivity to the exter- nal magnetic perturbation, lack of stray field, and ac- cess to extremely high frequency. Recently, the gener- ation and transmission of spin current in AFMs have attracted great attention. The spin pumping studies on (Pt, Ta)/(NiO, CoO)/Y 3Fe5O12(YIG) heterostruc- turesdemonstratethatthespincurrentgeneratedinYIG layer can pass through the antiferromagnetic (AFM) in- sulator NiO or CoO layer and can be detected in Pt or Ta layer by inverse spin-Hall effect (ISHE).2–5Sim- ilar results were also revealed in (Pt, Ta)/IrMn/CoFeB or Pt/NiO/FeNi heterostructures by spin-torque ferro- magnetic resonance(ST-FMR) technique, where the spin current generated by spin-Hall effect (SHE) in Pt or Ta layer can propagate through IrMn or NiO layer and change the FMR linewidth.6–8The spin current gen- erated by spin pumping or spin Seebeck was also ob- served in IrMn/YIG, Cr/YIG, and XMn/Py ( X= Fe, Pd, Ir, and Pt) bilayers through ISHE.9–13Moreover, the IrMn/YIG, Pt/Cr 2O3, and Pt/MnF 2exhibit spin- a)Present address: Swiss light source & Laboratory for Scient ific Developments and Novel Materials, Paul Scherrer Institut, CH- 5232 Villigen PSI, Switzerland b)Electronic mail: zhanqf@nimte.ac.cn c)Electronic mail: zhangshu@email.arizona.edu d)Electronic mail: runweili@nimte.ac.cnHall magnetoresistance (SMR) and large ISHE voltage, respectively, implying that the AFMs can be both spin- current detector and generator.14–16These investigations open up new opportunities in developing the AFMs- based spin-current devices. The IrMn alloy, which have been widely used to pin an adjacent FM layer in spin valve devices via exchange bias,17demonstrates large ISHE voltage when in con- tacts with YIG.9Recently, a large SHE and anomalous- Hall effect (AHE) have been theoretically proposed in Cr, FeMn, and IrMn AFMs owing to their large spin- orbit coupling (SOC) or Berry phase of the non-collinear spintextures.18–20Thesetheoreticalpredictionswerealso found to be valid for other cubic non-collinear AFMs, e.g., SnMn 3and GeMn 3, where the calculations have beenrepeatedwithcomparableresults.21Theexperimen- tal investigation of AHE and SHE on the AFMs could be helpful from both fundamental and practical viewpoints for AFMs spintronics. As previously revealed in Cr/YIG bilayers, the large anomalous-Hall resistance (AHR) in thin unprotected Cr film is likely caused by the surface FM Cr oxides.11Similar situation is expected in unpro- tected IrMn/YIG bilayers. Since the Pt/YIG bilayer is well studied,22in this study, we choose the Pt as cap layer to protect the IrMn from oxidation to investigate the AHE and SHE of IrMn by measuring the spin trans- port properties in Pt/IrMn/YIG heterostructures. II. EXPERIMENTAL DETAILS The Pt/IrMn/YIG heterostructures were prepared in a combined ultra-high vacuum (10−9Torr) pulsed laser/s50/s48 /s51/s48 /s52/s48 /s53/s48 /s54/s48 /s55/s48 /s56/s48/s40/s98/s41/s40/s52/s52/s52/s41 /s32/s32/s73/s110/s116/s101/s110/s115/s105/s116/s121/s32/s40/s97/s46/s117/s46/s41 /s40/s100/s101/s103/s114/s101/s101 /s41/s40/s50/s50/s50/s41 /s50/s48/s54/s48 /s50/s48/s56/s48 /s50/s49/s48/s48 /s50/s49/s50/s48 /s50/s49/s52/s48 /s50/s49/s54/s48 /s32/s32/s100/s73 /s70/s77/s82/s47/s100/s72/s32/s40/s97/s46/s117/s46/s41 /s70/s105/s101/s108/s100/s32/s40/s79/s101/s41/s72/s126/s56/s79/s101/s40/s100/s41/s52/s57/s46/s53 /s53/s48/s46/s48 /s53/s48/s46/s53 /s53/s49/s46/s48 /s53/s49/s46/s53 /s53/s50/s46/s48 /s53/s50/s46/s53/s40/s97/s41 /s71/s71/s71 /s32/s32/s73/s110/s116/s101/s110/s115/s105/s116/s121/s32/s40/s97/s46/s117/s46/s41 /s50 /s32 /s40/s100/s101/s103/s114/s101/s101 /s41/s89/s73/s71 /s45/s48/s46/s54/s110/s109/s80/s116/s47/s73/s114/s77 /s110/s47/s89/s73/s71 /s40/s99/s41/s48/s46/s54/s110/s109 /s53/s46/s48 /s109/s48/s46/s48 FIG. 1. (Color online) (a) A representative 2 θ-ωXRD pat- terns for YIG/GGG film near the (444) peaks of GGG sub- strate and YIG film. (b) The full range of XRD patterns from 20 to 80 degree. (c) Atomic force microscope surface topogra - phy of Pt/IrMn(3)/YIG heterostructure over an area of 5 µm ×5µm. (d) A FMR derivative absorption spectrum of a 60 nm YIG film with an in-plane magnetic field; the line-width is estimated to be 8 Oe. deposition (PLD) and magnetron sputter system. The high quality epitaxial YIG films were deposited on (111)- orientated single crystalline Gd 3Ga5O12(GGG) sub- strate via PLD technique as described elsewhere.23The Ir20Mn80(IrMn) and Pt films were sputtered at room temperature in argon atmosphere in an in situprocess. The thickness and crystal structure of films were char- acterized by Bruker D8 Discover high-resolution x-ray diffractometer (HRXRD). The thickness was estimated by using the software package LEPTOS (Bruker AXS). The surface topography of the films was measured in a Bruker Icon atomic force microscope. The ferromagnetic resonance (FMR) was measured by Bruker electron spin resonance spectrometers. The measurements of trans- verse Hall resistance, longitudinal resistance, and mag- netization were carried out in a Quantum Design physi- cal properties measurement system (PPMS) with a rota- tionoptionandmagneticpropertiesmeasurementsystem (MPMS), respectively. III. RESULTS AND DISCUSSION Figure 1(a) plots a representative room-temperature 2θ-ωXRD pattern of epitaxial YIG/GGG film near the (444)reflections. ClearLaueoscillationsindicatetheflat- ness and uniformity of the epitaxial YIG film. As shown in the Fig. 1(b), only the (222) and (444) reflections can FIG. 2. (Color online) (a)-(c) Schematic plot of longitudin al resistance and transverse Hall resistance measurements. T he magnetic fields are applied in the xy,xz, andyzplanes with anglesθxy,θxz, andθyzrelative to the y-,z-, andz-axes. The electric current is applied along the x-axis. Anomalous- Hall resistance R AHRfor Pt/YIG (d) and Pt/IrMn(1)/YIG (e) as a function of magnetic field at different temperatures. (f) Temperature dependence of the ρAHRfor Pt/IrMn/YIG with various IrMn thicknesses. The ρAHRare replotted as a function of IrMn thickness at various temperatures in (g). A ll ρAHRare averaged by[ ρAHR(70 kOe)- ρAHR(-70 kOe)]/2. The error bars are the results of subtracting OHR in different fiel d ranges be observed, and no indication of impurities or misorien- tation was detected in the full range of 2 θ-ωscan. In this study, the thicknesses of YIG and Pt films, determined by simulation of the x-rayreflectivity (XRR) spectra, are approximately 60 nm and 3 nm, respectively, while the IrMn thickness ranges from 0 nm to 8 nm. The atomic force microscope surface topography of Pt/IrMn(3)/YIG heterostructure over an area of 5 µm×5µm in Fig. 1(c) reveals a root-mean-squaresurface roughness of 0.18 nm, indicating atomical flat of prepared films. The other films show similar surface roughness. The number in the brackets represents the thickness of IrMn layer in nm unit. A representative FMR derivative absorption spec- trum of YIG film (60 nm) shown in Fig. 1(d) exhibits a line width ∆H = 8 Oe, which was measured at radio frequency 9.39 GHz and power 0.1 mW with an in-plane magnetic field at room temperature. The above proper- ties indicate excellent quality of our prepared films. 2/s48 /s49/s48/s48 /s50/s48/s48 /s51/s48/s48/s48/s50/s52/s54/s56 /s48 /s49/s48/s48 /s50/s48/s48 /s51/s48/s48/s48/s51/s48/s54/s48/s57/s48 /s40/s103/s41/s32/s80/s116/s47/s73/s114/s77/s110/s47/s89/s73/s71/s48/s32/s40/s49/s48/s45/s53 /s41 /s84/s32/s40/s75/s41 /s84/s32/s40/s75/s41 /s47 /s48/s40/s49/s48/s45/s53 /s41 /s120 /s121 /s120 /s122 /s121/s122 /s40/s104/s41/s32/s80/s116/s47/s89/s73/s71/s52/s50/s48/s45/s50/s45/s52/s48/s51/s54/s57 /s48 /s57/s48 /s49/s56/s48 /s50/s55/s48 /s51/s54/s48/s45/s57/s45/s54/s45/s51/s48/s48/s50/s48/s52/s48/s54/s48 /s45/s50/s48/s45/s49/s48/s48 /s48 /s57/s48 /s49/s56/s48 /s50/s55/s48 /s51/s54/s48/s45/s54/s48/s45/s51/s48/s48/s40/s98/s41/s48/s32/s40/s49/s48/s45/s53 /s41/s80/s116/s47/s73/s114/s77/s110/s47/s89/s73/s71 /s32/s49/s48/s75 /s32/s53/s48 /s32/s49/s48/s48 /s32/s49/s53/s48 /s32/s50/s48/s48 /s32/s50/s53/s48 /s32/s51/s48/s48/s40/s97/s41 /s40/s99/s41 /s40/s100/s101/s103/s114/s101/s101/s41/s80/s116/s47/s89/s73/s71/s40/s100/s41 /s40/s101/s41 /s47 /s48/s40/s49/s48/s45/s53 /s41 /s40/s102/s41 /s32/s40/s100/s101/s103/s114/s101/s101/s41 FIG. 3. (Color online) Anisotropic magnetoresistance for Pt/IrMn(1)/YIG at various temperatures down to 10 K with the magnetic field varied within xy(a),xz(b), and yz(c) planes. The results of Pt/YIG are shown in (d)-(f). Temper- ature dependence of AMR amplitudes for Pt/IrMn(1)/YIG (g) and Pt/YIG (h) heterostructures. The cubic, circle and triangle symbols standfor the θxy,θxz,θyzscans, respectively. A. Anomalous-Hall resistance As shown in the top panel of Fig. 2, in order to measure the transverse Hall resistance and longitudinal resistance, all the Pt/IrMn/YIG heterostructures were patterned into Hall-bar configuration (central area: 0.3 mm×10 mm; electrode 0.3 mm ×1 mm). The trans- verse Hall resistance R xyof Pt/IrMn/YIG was measured in the temperature range of 10 K to 300 K with per- pendicular magnetic field ranging from -70 to 70 kOe. In metal thin film, the ordinary-Hall resistance (OHR) ROHRis subtracted from the measured R xy, i.e., R AHR = Rxy- ROHR×µ0H, where R AHRis AHR. As shown in Figs. 2(d)-(e), the resulting R AHRas a function of magnetic field for Pt/YIG and Pt/IrMn(1)/YIG are pre- sented. It is noted that the Pt becomesmagnetic when in contacts with the YIG due to its proximity to the stoner ferromagnetic instability, i.e., magnetic proximity effect (MPE), as previously shown experimentally by x-ray magnetic circular dichroism (XMCD) and theoretically by first-principles calculation.24,25The magnetized Pt shares some common features as magnetic YIG film, i.e., strong anisotropy.23Thus, when the magnetic field ap- proaches zero, the magnetized Pt moments are randomlydistributed, the R AHRexhibits irregular M-shaped be- haviorclosetozerofield. However,forPt/IrMn/YIG,the RAHRcontinuously decreases as approaching zero field, implying that the Pt/IrMn and IrMn/YIG interfaces are freeofMPE,beingconsistentwith theabsenceofconven- tional anisotropic magnetoresistance (CAMR) (see be- low). We summarize the derived anomalous-Hall resis- tivityρAHRofPt/IrMn/YIGheterostruturesasfunctions of temperature ( T) and IrMn thickness ( tIrMn) in Figs. 2(f)-(g). The ρAHR(T) for all the Pt/IrMn/YIG exhibits rich characteristics whose magnitude and sign are highly non-trivial, which were also found in Pt/LaCoO 3bilay- ers.26. As shown in Fig. 2(h), the magnitude of ρAHR decrease with temperature and then it increases again below 100 K. Simultaneously, the ρAHRchange its sign at the temperature which is independent of IrMn thick- ness. We also replotted all the ρAHRas a function of IrMn thickness in Fig. 2(g). In the studied temperature range, as increasing the tIrMn, theρAHRalso increases and reaches a maximum around tIrMn= 3 nm, which excludes the interfacial origin of the observed AHR. B. Spin-Hall magnetoresistance The anisotropic magnetoresistance (AMR) for Pt/IrMn/YIG was also measured down to low temper- atures. As an example, the AMR of Pt/IrMn(1)/YIG and Pt/YIG for three different field scans are presented in top panel of Fig. 3. When the magnetic field scans within the xyplane [Fig. 3(a)(d)], both the CAMR and SMR contribute to the total AMR; for the xzplane [Fig. 3(b)(e)], the resistance changes are attributed to the MPE-induced CAMR; for the yzplane [Fig. 3(c)(f)], the CAMR is zero, and only the SMR are expected.29,30 As shown in Fig. 3(b), the θxzscan shows negligible AMR and the resistance is almost independent of θxz, indicating the extremely weak MPE at the interface even down to low temperatures. However, the MPE is significant at Pt/YIG interface [see Fig. 3(e)]: the max- imum amplitude of CAMR is around 2.2 ×10−4, which is comparable to the SMR. Thus, the IrMn can be used as clean spin current detector and generator, similar to the normal Rh or AFM Cr metals.11,23Since the CAMR is negligible in Pt/IrMn(1)/YIG, the SMR dominates the AMR when the magnetic field is varied within the xyplane, the amplitudes of θxyscan are almost identical toθyzscan. While for Pt/YIG, due to the MPE-induced CAMR, none of the amplitudes is identical to each other. The temperature dependence of the AMR amplitudes for allθxy,θxz, andθyzscans are summarized in Fig. 3(g) and Fig. 3(h) for Pt/IrMn(1)/YIG and Pt/YIG, respectively. Upon decreasing the temperature, the SMR persists down to 10 K, with the amplitudes monotonically decreasing from 7.5 ×10−5(300 K) to 3.0 ×10−5(10 K) in Pt/IrMn(1)/YIG. For Pt/IrMn/YIG, the amplitudes of SMR are almost an order smaller than that of the Pt/YIG due to the smaller spin-Hall angle, 3/s45/s49/s50/s48 /s45/s54/s48 /s48 /s54/s48 /s49/s50/s48/s45/s49/s46/s48/s45/s48/s46/s53/s48/s46/s48/s48/s46/s53/s49/s46/s48 /s48 /s49/s48/s48 /s50/s48/s48 /s51/s48/s48/s48/s45/s50/s48/s45/s52/s48/s45/s54/s48 /s48 /s49/s48/s48 /s50/s48/s48 /s51/s48/s48/s48/s49/s48/s50/s48/s51/s48/s52/s48/s32/s50/s75 /s32/s49/s48 /s32/s50/s48 /s32/s51/s48 /s32/s52/s48 /s32/s53/s48 /s32/s54/s48 /s32/s55/s48 /s32/s56/s48 /s32/s57/s48 /s32/s49/s48/s48 /s32/s49/s50/s48 /s32/s49/s53/s48 /s32/s49/s56/s48 /s32/s51/s48/s48/s77/s47/s77 /s115 /s70/s105/s101/s108/s100/s32/s40/s79/s101/s41/s40/s97/s41 /s32/s80/s116/s47/s73/s114/s77/s110/s40/s49/s32/s110/s109/s41/s47/s89/s73/s71 /s72 /s69/s32/s40/s79/s101/s41/s32/s49/s110/s109 /s32/s51 /s32/s53/s40/s98/s41 /s32 /s84/s32/s40/s75/s41/s72 /s67/s32/s40/s79/s101/s41 /s84/s32/s40/s75/s41 FIG. 4. (Color online) (a) Field dependence of normalized magnetization M/Msfor Pt/IrMn(1)/YIG at various temper- atures down to 2 K. The magnetic field is applied parallel to the film surface. The paramagnetic background of the GGG substrate has been subtracted. (b) The in-plane exchange bias field HEversus temperature. The arrows indicate the AFM block temperatures Tb. The inset plots the coercivity fieldHCas a function of temperature for Pt/IrMn(1)/YIG. shorter spin diffusion length, and larger electrical re- sistivity of IrMn.9,12,29The temperature characteristics of SMR amplitudes in Pt/IrMn/YIG are significantly different from the Pt/YIG or Pd/YIG bilayers, where the SMR amplitudes exhibit nonmonotonic tempera- ture dependence and acquire a maximum around 100 K.31,32For Pt/YIG, the temperature dependence of SMR amplitude can be described by a single spin- relaxation mechanism.31The spin diffusion length is defined as λ=/radicalbigDτsf, whereDandτsfare diffusion constant and spin-flip relaxation time, respectively. Within the Elliot-Yafet spin-orbit scattering model, bothDandτsfare proportional to the reciprocal of temperature dependence of the resistivity 1/ ρ(T).33,34 In Pt metal, the electrical resistivity mainly comes from phonon-electron scattering at high temperature, thenλ∝1/T. However, the extra magnetic electron scattering need to be considered in Pt/IrMn/YIG heterostructures, the assumption of λ∝1/Tis invalid. It is noted that the heterostructures with different IrMn thicknesses exhibit similar temperature dependent characteristics with different numerical values compared to the Pt/IrMn(1)/YIG heterostructure shown here. For example, the Pt/IrMn(3)/YIG exhibits the SMR amplitude of 6.8 ×10−5at room temperature. The sizable SMR observed in Pt/IrMn/YIG heterostructures indicates that the spin current can transport through IrMn layer.C. Magnetization Since the magnetic transitions of very thin AFMs are expected to be well below the ordering temper- ature of bulk forms, we measured the field depen- dence of magnetization down to low temperatures, from which we can track the AFM blocking temperature Tbfor Pt/IrMn/YIG heterostructures. As an exam- ple, the normalized magnetic hysteresis loops M/Msfor Pt/IrMn(1)/YIGatvarioustemperaturesafterfieldcool- ing from 300 K are presented in Fig. 4(a). The derived exchange bias field HEversus temperature are summa- rized in Fig. 4(b), from which the Tbare approximately estimated to be 150 K, 180 K, and 220 K for 1 nm, 3 nm and 5 nm IrMn, respectively, as the arrow indicated. Similar blocking temperatures were previously reported in IrMn/MgO/Ta tunnel junctions and IrMn/NiFe bi- layer.35,36Moreover, the coercivity HCalso exhibits a step-like increase near the blocking temperature, as the arrow shown in the inset of Fig. 4(b), indicating the strongly enhanced exchange coupling between IrMn and YIG layer below Tb. D. Discussion Based on the above experimental results, we discuss the origins of the significant AHR in Pt/IrMn/YIG het- erostructures and the effect of AFM order on spin trans- port properties. There are at least four contributions to the observed AHR in Pt/IrMn/YIG: MPE, spin-Hall based SMR, spin-dependent interface scattering, and in- trinsic properties of IrMn metal. In contrast to the Pt/YIG, the negligible CAMR in Pt/IrMn/YIG indi- cates the extremely weak MPE at Pt/IrMn or IrMn/YIG interfaces, which is different from the previous studied of IrMn/YIG bilayer.14The SMR model based on SHE also predicts an anomalous-Hall-like resistance,29whose magnitude and sign are determined by the spin diffu- sion length and spin-Hall angle of the metal and the imaginary part of the spin mixing conductance, respec- tively. Though the thickness dependence of the AHR in Pt/IrMn/YIG can be described by the SMR model, it fails to explain the AHR by the following reasons: (i) An arbitrary temperature dependence of the imag- inary part of the spin mixing conductance parameter is required to qualitatively describe the temperature- dependent AHR data, i.e., signreversal; (ii) Accordingto the spin pumping studies, both the spin-Hall angle and the spin diffusion length of IrMn are smaller than Pt, which cannot explain the enhancement of AHR by in- creasing the IrMn thickness.9,12,13Spin-dependent scat- tering at the interface, combined with the conventional skew-scatteringand side-jump mechanisms, can also give rise to AHR.37Again, the enhancement of AHR by in- creasing the IrMn thickness excludes the interfacial ori- gin. Finally, the theoretical calculations predict a large AHE and SHE in IrMn metal not only attributed to the 4large SOC of heavy Ir atoms which is transferred to the magnetic Mn atoms by hybridization effect but also the Berry phase of the non-collinear spin structures.18–20We conclude that the large AHR observed in Pt/IrMn/YIG is likely associated with SOC and non-collinear magnetic structure of IrMn. However, the non-trivial temperature dependence of AHR demands further theoretical and ex- perimental investigations. Now we discuss the possible interplay between AFM order and spin transport properties. As shown in Fig. 2 and Fig. 3, there is no clear anomalous in AHR or SMR near the blocking temperatures of IrMn, imply- ing weak correlations between the AHE or SHE and the AFM order in IrMn. Similar results were also observed in Cr/YIG bilayers, where the ISHE voltage and AHR is also independent of AFM ordering temperature.11Ac- cording to our magnetization results (Fig. 4), the AFM ordering temperatures of our IrMn films are well below room temperature. However, the enhancement of AHR in Pt/IrMn/YIG happens in the whole studied tempera- ture range [see Fig. 2(g)]. There are two possible reasons for this phenomenon, one is that the AHE and SHE at- tributed to non-collinear magnetism is generated on a length scale of nanometer and is a local property not relying on long range magnetic order, i.e., regardless of how IrMn grains are orientated, as reported previously in Mn 5Si3film.38The second one is that the strength of SOC is independent of AFM order in IrMn metal, which is mainly determined by the Ir atoms. IV. CONCLUSIONS In summary, we report an investigation of AHE and SHE by measuring the AHR and SMR in Pt/IrMn/YIG heterostrucutres. The significant AHR in Pt/IrMn/YIG islikelyassociatedwiththestrongSOCandnon-collinear magnetic structure of IrMn, and the sizable SMR in- dicates that the spin current can transport through IrMn. The observed non-trivial temperature dependence of AHR cannot be consistently explained by the existing theories, further investigations are needed to clarify this issue. Moreover, both the AHR and SMR are uncoupled to the AFM order of IrMn metal. The negligible MPE at Pt/IrMn or IrMn/YIG interface and large ISHE indi- cate that IrMn can be another model system to explore physics and devices associated with antiferromagnetism and pure spin current. ACKNOWLEDGMENTS We thank the high magnetic field laboratory of Chi- nese Academy of Sciences for the FMR measurements. This work is financially supported by the National Nat- ural Science foundation of China (Grants No. 11274321, No. 11404349, No. 51502314, No. 51522105) and the Key Research Program of the Chinese Academy of Sci-ences (Grant No. KJZD-EW-M05). S. Zhang was par- tiallysupportedbytheU.S.NationalScienceFoundation (Grant No. ECCS-1404542). 1A. H. MacDonald and M. Tsoi, Phil. Trans. R. Soc. A 369, 3098 (2011). 2C. Hahn, G. de. Loubens, O. Klein, M. Viret, V. V. Naletov, and J. B. Youssef, Europhys. Lett. 108, 57005 (2014). 3H. L. Wang, C. H. Du, P. C. Hammel, and F. Y. Yang, Phys. Rev. Lett. 113, 097202 (2014). 4Z. Qiu, J. Li, D. Hou, E. Arenholz, A. T. NDiaye, A. Tan, K. Uchida, K. Sato, Y. Tserkovnyak, Z. Q. Qiu, and E. Saitoh, arXiv: 1505.03926. 5W. Lin, K. Chen, S. Zhang, and C. L. Chien, arXiv: 1603.00931. 6T. Moriyama, M. Nagata, K. Tanaka, K-J. Kim, H. Almasi, W. Wang, T. Ono, arXiv: 1411. 4100. 7H. Reichlov´ a, D. Kriegner, V. Hol´ y, K. Olejn´ ık, V. Nov´ ak , M. Yamada, K. Miura, S. Ogawa, H. Takahashi, T. Jungwirth, and J. Wunderlich, Phys. Rev B 92, 165424 (2015). 8T. Moriyama, S. Takei, M. Nagata, Y. Yoshimura, N. Matsuzaki , T. Terashima3, Y. Tserkovnyak, and T. Ono, Appl. Phys. Lett. 106, 162406 (2015). 9J. B. S. Mendes, R. O. Cunha, O. Alves Santos, P. R. T. Ribeiro, F. L. A. Machado, R. L. Rodr´ ıguez-Su´ arez, A. Azevedo, and S . M. Rezende, Phys. Rev B 89, 140406(R) (2014). 10L. Frangou, S. Oyarz´ un, S. Auffret, L. Vila, S. Gambarelli, a nd V. Baltz, Phys. Rev. Lett. 116, 077203 (2016). 11D. Qu, S. Y. Huang, and C.L. Chien, Phys. Rev B 92, 020418(R) (2015). 12W. Zhang, M. B. Jungfleisch, W. Jiang, J. E. Pearson, A. Hoff- mann, F. Freimuth, and Y. Mokrousov, Phys. Rev. Lett. 113, 196602 (2014). 13J. Sinova, S. O. Valenzuela, J. Wunderlich, C. H. Back, and T. Jungwirth, Rev. Mod. Phys. 87, 1213 (2015) and reference therein. 14X. Zhou, L. Ma, Z. Shi, W. J. Fan, J. G. Zheng, R. F. L. Evans, and S. M. Zhou, Phys. Rev. B 92, 060402(R) (2015). 15S. Seki, T. Ideue, M. Kubota, Y. Kozuka, R. Takagi, M. Naka- mura, Y. Kaneko, M. Kawasaki, and Y. Tokura, Phys. Rev. Lett. 115, 266601 (2015). 16S. M. Wu, W. Zhang, A. KC, P. Borisov, J. E. Pearson, J. S. Jiang, D. Lederman, A. Hoffmann, and A. Bhattacharya, Phys. Rev. Lett. 116, 097204 (2016). 17K. Hoshino, R. Nakatani, H. Hoshiya, Y. Sugita, and S. Tsunashima, Jpn. J. Appl. Phys. 35, 607 (1996). 18H. Chen, Q. Niu, and A. H. MacDonald, Phys. Rev. Lett. 112, 017205 (2014). 19R. Shindou and N. Nagaosa, Phys. Rev. Lett. 87, 116801 (2001). 20F. Freimuth, S. Bl¨ ugel, and Y. Mokrousov, Phys. Rev. Lett. 105, 246602 (2010). 21J. K¨ ubler and C. Felser, Europhys. Lett. 108, 67001 (2014). 22M. Z. Wu and A. Hoffmann, Recent Advances in Magnetic Insulators-From Spintronics to Microwave Applications (Aca- demic Press, San Diego, Vol 64, 2013). 23T. Shang, Q. F. Zhan, H. L. Yang, Z. H. Zuo, Y. L. Xie, H. H. Li, L. P. Liu, B. M. Wang, Y. H. Wu, S. Zhang, and Run-Wei Li, Sci. Rep. 5, 17734 (2015). 24Y. M. Lu, Y. Choi, C. M. Ortega, X. M. Cheng, J. W. Cai, S. Y. Huang, L. Sun, and C. L. Chien, Phys. Rev. Lett. 110, 147207 (2013). 25D. Qu, S. Y. Huang, J. Hu, R. Q. Wu, and C. L. Chien, Phys. Rev. Lett. 110, 067206 (2013). 26T.Shang, Q.F.Zhan, H.L.Yang, Z.H.Zuo, Y.L.Xie, Y.Zhang, L. P. Liu, B. M. Wang, Y. H. Wu, S. Zhang, and Run-Wei Li, Phys. Rev. B 92, 165114 (2015). 27B. F. Miao, S. Y. Huang, D. Qu, and C. L. Chien, Phys. Rev. Lett.112, 236601 (2014). 28M. Althammer, S. Meyer, H. Nakayama, M. Schreier, S. Alt- mannshofer, M. Weiler, H. Huebl, S. Gepr¨ ags, M. Opel, R. Gro ss, D. Meier, C. Klewe, T. Kuschel, J. M. Schmalhorst, G. Reiss, L . 5M. Shen, A. Gupta, Y. T. Chen, G. E. W. Bauer, E. Saitoh, and S. T. B. Goennenwein, Phys. Rev. B 87, 224401 (2013). 29Y. T. Chen, S. Takahashi, H. Nakayama, M. Althammer, S. T. B. Goennenwein, E. Saitoh, and G. E. W. Bauer, Phys. Rev. B 87, 144411 (2013). 30M. Isasa, A. B. Pinto, S. V´ elez, F. Golmar, F. S´ anchez, L. E. Hueso, J. Fontcuberta, and F. Casanova, Appl. Phys. Lett. 105, 142402 (2014). 31S. R. Marmion, M. Ali, M. McLaren, D. A. Williams, and B. J. Hickey, Phys. Rev. B 89, 220404(R) (2014). 32T. Lin, C. Tang, H. M. Alyahayaei, and J. Shi, Phys. Rev. Lett. 113, 037203 (2014).33J. Fabian and S. Das Sarma, Phys. Rev. Lett. 81, 5624 (1998). 34L. Q. Liu, R. A. Buhrman, and D. C. Ralph, arXiv:1111.3702. 35D. Petti, E. Albisetti, H. Reichlov´ a, J. Gazquez, M. Varela , M. Molina-Ruiz, A. F. Lopeand´ ıa, K. Olejn´ ık, V. Nov´ ak, I. Fi na, B. Dkhil, J. Hayakawa, X. Marti, J. Wunderlich, T. Jungwirth, a nd R. Bertacco, Appl. Phys. Lett. 102, 192404 (2013). 36V. Tshitoyan, C. Ciccarelli, A. P. Mihai, M. Ali, A. C. Irvine , T. A. Moore, T. Jungwirth, and A. J. Ferguson, Phys. Rev. B 92, 214406 (2015). 37N. Nagaosa, J. Sinova, S. Onoda, A. H. MacDonald, and N. P. Ong, Rev. Mod. Phys. 82, 1539 (2010) and reference therein. 38C. S¨ urgers, W. Kittler, T. Wolf, and H. v. Loehneysen, arXiv : 1601.01840. 6 | 2016-03-11 | We report an investigation of temperature and IrMn layered thickness
dependence of anomalous-Hall resistance (AHR), anisotropic magnetoresistance
(AMR), and magnetization on Pt/Ir20Mn80/Y3Fe5O12 (Pt/IrMn/YIG)
heterostructures. The magnitude of AHR is dramatically enhanced compared with
Pt/YIG bilayers. The enhancement is much more profound at higher temperatures
and peaks at the IrMn thickness of 3 nm. The observed spin-Hall
magnetoresistance (SMR) in the temperature range of 10-300 K indicates that the
spin current generated in the Pt layer can penetrate the entire thickness of
the IrMn layer to interact with the YIG layer. The lack of conventional
anisotropic magnetoresistance (CAMR) implies that the insertion of the IrMn
layer between Pt and YIG efficiently suppresses the magnetic proximity effect
(MPE) on induced Pt moments by YIG. Our results suggest that the dual roles of
the InMn insertion in Pt/IrMn/YIG heterostructures are to block the MPE and to
transport the spin current between Pt and YIG layers. We discuss possible
mechanisms for the enhanced AHR. | Investigation of anomalous-Hall and spin-Hall effects of antiferromagnetic IrMn sandwiched by Pt and YIG layers | 1603.03578v2 |
EPJ manuscript No. (will be inserted by the editor) Building instructions for a ferromagnetic axion haloscope Nicol o Crescini1 Univ. Grenoble Alpes, CNRS, Grenoble INP, Institut Nel, 38000 Grenoble, France Received: date / Revised version: date Abstract. A ferromagnetic haloscope is a rf spin-magnetometer used for searching Dark Matter in the form of axions. A magnetic material is monitored searching for anomalous magnetization oscillations which can be induced by dark matter axions. To properly devise such instrument one rst needs to understand the features of the searched-for signal, namely the eective rf eld of dark matter axions Baacting on electronic spins. Once the properties of Baare dened, the design and test of the apparatus may start. The optimal sample is a narrow linewidth and high spin-density material such as Yttrium Iron Garnet (YIG), coupled to a microwave cavity with almost matched linewidth to collect the signal. The power in the resonator is collected with an antenna and amplied with a Josephson Parametric amplier, a quantum-limited device which, however, adds most of the setup noise. The signal is further amplied with low noise HEMT and down-converted for storage with an heterodyne receiver. This work describes how to build such apparatus, with all the experimental details, the main issues one might face, and some solutions. 1 Introduction The axion is an hypothetical beyond the Standard Model particle, rst introduced in the seventies as a consequence of the strong CP problem of QCD. Axions can be the main constituents of the galactic Dark Matter halos. Their experimental search can be carried out with Earth-based instruments immersed in the Milky Way's halo, which are therefore called \haloscopes". Nowadays haloscopes rely on the inverse Primako eect to detect axion-induced excesses of photons in a microwave cavity under a static magnetic eld. This work describes the process leading to the successful operation of a ferromagnetic axion haloscope, which does not exploit the axion-to-photon conversion but its interaction with the electron spin. The study of the axion-spin interaction and of the Dark Matter halo properties yields the features of the axionic signal, and is fundamental to devise a proper detector. A scheme of a realistic ferromagnetic haloscope is drawn to realize the challenges of its development. It emerges that there are a number of requirements for a this setup to get to the sensitivity needed for a QCD-axion search. These are kept in mind when designing the prototypes, to overcome the problems without compromising other requirements. A state-of-the-art sensitivity to rf signals allows for the detection of extremely weak signals as the axionic one. The number of monitored spins is necessarily large to increase the exposure of the setup, thus its scalability is a key part of the design process. A ferromagnetic haloscope consists in a transducer of the axionic signal, which is then measured by a suitable detector. The transducer is a hybrid system formed by a magnetic material coupled to a microwave cavity through a static magnetic eld. Its two parts are separately studied to nd the materials which match the detection conditions imposed by the axion-signal. The detector is an amplier, an HEMT or a JPA, reading out the power from the hybrid system collected by an antenna coupled to the cavity. A particular attention is given to the measurement of the noise temperature of the amplier. As it measures variation in the magnetization of the sample, the ferromagnetic haloscope is congured as a spin-magnetometer. The present haloscope prototype [1] works at 90 mK and reaches the sensitivity limit imposed by quantum mechan- ics, the Standard Quantum Limit, and can be improved only by quantum technologies like single photon counters. The haloscope embodies a large quantity of magnetic material, i. e. ten 2 mm YIG spheres, and is designed to be further up-scaled. This experimental apparatus meets the expected performances, and, to present knowledge, is the most sensitive rf spin-magnetometer existing. The minimum detectable eld at 10.3 GHz results in 5 :510 19T for 8 h integration, and corresponds to a limit on the axion-electron coupling constant gaee1:710 11. This result is the best limit on the DM-axions coupling to electron spins in a frequency span of about 150 MHz, corresponding to an axion mass range from 42 :4eV to 43:1eV. The eorts to enhance the haloscope sensitivity include improvements in both the hybrid system and the detector. The deposited axion power can be increased by means of a larger material volume, possibly with a narrower linewidth.arXiv:2201.04081v1 [hep-ex] 11 Jan 2022#### Page 2 of 18 Eur. Phys. J. Plus ##################### To overcome the standard quantum limit of linear ampliers one must rely on quantum counters. Novel studies on microwave photon counters, together with some preliminary results, are reported. Other possible usages of the spin- magnetometer are eventually discussed. 2 Overview on axions A long-standing puzzle of beyond the Standard Model physics consists in the dark matter (DM) problem. In 1933 Fritz Zwicky used two dierent techniques to estimate the mass of the Coma and Virgo clusters, one was based on the luminosity of the galaxies in the clusters, while the other used the velocity dispersion of individual galaxies. These two independent estimations did not agree by orders of magnitude [2]. It is only in the seventies that this discrepancy started to be studied systematically. In particular, Vera Rubin studied the rotation curves of spiral galaxies and observed a violation of the second Kepler's law which can be explained assuming that the mass prole does not vanish beyond the stars [3,4]. This was an early indication that spiral galaxies could be surrounded by an halo of DM. Despite these evidences, the nature of DM is still unknown. Its possible composition could be baryonic or non-baryonic. The former considers matter similar to the one already known, while the latter comprehends hypothetical particles of beyond the Standard Model (BSM) physics. The case of a non-baryonic DM is where cosmology meets particle physics. Approaching this problem, physicists glimpse the possibility of merging dierent questions which are apparently uncorrelated. New theories, remarkably su- persymmetric DM [5], triggered experimental searches in dierent forms and with various techniques. Low-background laboratory experiments aim at a direct detection [6], accelerators could produce such particles and observe their miss- ing energy and momentum [7], while indirect evidences are based on their decay or annihilation [8]. The detection of BSM particles would shed light on fundamental question like DM or the unication of all forces. Up to now the results of the LHC and experiments therein showed no evidence of new physics up to the 10 TeV scale. On the other hand, there are signicant hints for physics at the sub-eV scale, like neutrino oscillation or the vacuum energy density of the Universe [9]. The physics case of weakly interacting sub-eV particles (WISPs) is motivated by the fact that any theory introducing a high-energy global symmetry breaking implies a light particle by the Nambu-Goldstone theorem [10]. Among other WISPs, the axion appears as a well-motivated BSM particle. Originally introduced to account for a ne-tuning issue in the SM known as the \strong CP problem" of quantum chromodynamics (QCD), it quickly became a prominent DM candidate. The existence of axions is a very attractive perspective, since its addition to the Standard Model would solve two major problems of modern physics in a single shot [11]. QCD is a non-Abelian SU(3) cgauge theory which describes the strong interactions. Its Lagrangian LQCD contains a CP-violating term which is compatible with all symmetries of the SM gauge group. However, since there is no experimental sign of CP violation in strong interactions, one needs to unnaturally suppress of this term. The rst SU(3) ctheory was proposed as CP-conserving to agree with experimental observations, but had an issue at low energy known as Weinberg's U(1) Amissing meson problem. The strong CP problem arises following the solution of the missing meson problem proposed by t'Hooft [12, 13]. This solution brings on CP violation in QCD, parametrized by =QCD+2Y, two angles relative to QCD which, according to the theory, are independent. However, to conserve CP either should be zero or one of the quarks should be massless. Among the measurable observables containing and thus CP violation, the neutron electric dipole moment results [14,15,16]. Recent results by the nEDM collaboration constrain the parameter even more 10 10[17,18,19]. Since it is unlikely that nature chose very small YandQCD or a ne tuning among them, a naturalness problem arises. It is normally denoted as the strong CP problem, which is an important hint of BSM physics. A solution to this problem is a scenario where is promoted from a parameter to an actual particle. This was realized, albeit in a dierent way, by Peccei and Quinn [20], who introduced a new U(1) PQsymmetry to the SM to dynamically interpret . The idea was further developed by, among others, Weinberg and Wilczek [21,22,23,24,25], who realized that it implies the existence of a new light pseudo-Nambu Goldstone boson which was called axion. The minimization of the meson potential adjusts the axion vacuum expectation value to cancel any eect of CP violation, addressing the strong CP problem. The axion mass eigenstate can then be computed from the masses of the pion m, of the up and down quarks, mu andmd, and from the decay constants of the pion and of the axion itself, fandfa, resulting in the axion mass m2 a'=mumd (mu+md)2m2 f2 f2a: (1) The energy scale fais the PQ-symmetry breaking scale, and as the axion is the pseudo-Goldstone boson arising from this process, its mass and couplings are proportional to f 1 amaking it very light and weakly-interacting. The so-called \invisible-axion models" consider fa'1012GeV, and evade current experimental limits [26]. There are two main classes of invisible axions whose archetype are the Kim-Shifman-Vainshtein-Zakharov (KSVZ) and Dine-Fischler- Srednicki-Zhitnitsky (DFSZ) models [27,28,25,23,29], where the main dierence is the coupling to SM particles.Eur. Phys. J. Plus ##################### #### Page 3 of 18 It is now possible to analyze the axion as a constituent of DM. Cold DM particles must be present in the Universe in a sucient quantity to account for the observed DM abundance and they have to be eectively collisionless, i. e. to have only signicant long-range gravitational interactions. The axion satises both these criteria. Even if it is very light, the axion population is non-relativistic since it is produced out of equilibrium by vacuum realignment, string decay or domain wall decay [30,31,29,32,33,34,35,36]. Being the main cold axions production mechanism, vacuum realignment is basically explained hereafter. In a time when the Universe cools down to a temperature lower than the axion mass, the axion eld is sitting in a random point of its potential, and not necessarily at its minimum. As a consequence the eld starts to oscillate and, since the axion has extremely weak couplings, it has no way to dissipate its energy. This relic energy density is a form of cold Dark Matter [37,38]. Lattice QCD calculation can be used together with the present Dark Matter density to give an estimation of the QCD axion mass [39,40,41,42,43,44] The axion has to be framed in the context of present physical theories, since one can wonder if the presence of light scalars may in
uence the behavior of already studied physical systems. Several constraints come from tting the axion theory into astrophysical and cosmological observations. As other weakly interacting low-mass particles, they can contribute to the cooling of stars and be produced in astrophysical plasmas and in the Sun, contribute to stellar evolution and even aect supernovae [45,46,47,48,49,50,51,52,53,54,55,56,57]. To sum up, these observations suggest that ma10 meV. Cosmology provides both upped and lower limits for the axion mass, but being the upper ones weaker than the ones already described, the focus will be on lower bounds. These limits on the axion mass come from the production of DM-axions in the early Universe [30,58,59,60,61,62,63,64,65,66]. In particular the axion mass must be higher than 6 eV to avoid the overclosure problem, i. e. an axion density exceeding the observed DM density. Lighter masses are still possible within the so-called \anthropic axion window". A general case of BSM particles are the so-called \axion-like particles" (ALPs). The interest in ALPs relies on the fact that its mass and coupling constants can be unrelated (other than for axions), thus they do not necessarily solve the strong-CP problem but still can account for the whole DM density of the Universe [67]. Any experimental search not reaching the axion sensitivity is still a probe of ALPs. 2.1 Experimental searches In the last decades several experimental techniques have been proposed to detect axions and ALPs [9,68]. Most experiments do not reach the axion-required sensitivity, but the physics result of these measurements is to limit the ALPs parameter space. The most tested eects of axions are related to their coupling to photons, being this one the strongest and thus most accessible parameter. These limits mostly rely on the inverse Primako eect: in a strong static magnetic eld it is possible to convert an itinerant axion into a photon that can be detected. Amongst all the experiment proposed or realised to detect axions, only haloscopes are treated in some details. As already discussed axions may constitute DM, and if existing at least a fraction of DM have to be composed of axions. DM is an interesting source of axions and triggered multiple experimental searches. Instruments searching for DM-axions composing the Milky Way's halo are called haloscopes. Haloscopes are particularly interesting in the scope of this work, which is devoted to the study of a ferromagnetic one. In 1983, Sikivie proposed new ways to detect the axion by resonantly converting them into microwave photons inside a high quality factor ( Q) cavity under a static magnetic eld [69]. The resonance condition implies that the apparatus is sensitive to axions in a very narrow frequency range. The frequency of the axion signal is related to its mass and its width depends on the virial DM velocities in the galaxy. These kinds of experiments need to change resonant frequency to scan for dierent masses. The Axion Dark Matter eXperiment (ADMX) reached the sensitivity of KSVZ axions in the range 1 :9eV 3:3eV [70,71] assuming virialized axions composing the whole DM density %DM= 0:45 GeV=cm3. The setup was improved by using SQUID ampliers [72], and then reached the line of the DFSZ model [73,74]. The HAYSTAC experiment searched for heavier axions by operating a setup similar to the ADMX one but using a Josephson Parametric Amplier (JPA), and achieving quantum limited sensitivity [75] and beyond [76]. The collaborations UF and RBF also reached remarkable limits, and the ORGAN experiment operated a pathnding haloscope at 110 eV [77,78]. Several new concepts have been proposed to search for DM axions with next-generation haloscopes based not only on the Primako eect but also on axion-induced electric dipole moments or on the axion-spin interaction [79,80,81,82,83,84,85,86]. 3 The eective magnetic eld of DM axions The coupling between axions and electron spins can be used for axion detection as an alternative to the coupling to photons [87,88,89,90]. Being it weaker than the axion-photon coupling it was not immediately exploited, but recently new experimental schemes were presented. Besides the axion discovery, the axion-electron coupling is interesting for distinguishing between dierent axion models. The possibility of detecting galactic axions by means of converting them into collective excitations of the magnetization (magnons) was considered by Barbieri et al. in [87], laying the founda- tions to the Barbieri Cerdonio Fiorentini Vitale (BCFV) scheme and to the following experimental proposal [84]. The#### Page 4 of 18 Eur. Phys. J. Plus ##################### original idea is to use the large de Broglie wavelength of the galactic axions to detect the coherent interaction between the axion DM cloud and the homogeneous magnetization of a macroscopic sample. To couple a single magnetization mode to the axion eld, the sample is inserted in a static magnetic eld. The interaction yields a conversion rate of axions to magnons which can be measured by monitoring the power spectrum of the magnetization. The form of the interaction is calculated hereafter in terms of an eective magnetic eld. Such a eld is the searched-for signal. Its features are derived and characterized as follows. The axion derivative interaction with fermions is invariant under a shift of the axion eld a!a+a0and reads L =C 2fa
5 @a (2) where is the spinor eld of a fermion of mass m , andC is a model-dependent coecient. The dimensionless couplings can be dened as ga =C m =fa (3) and play the role of Yukawa couplings, while the ne structure constant of the interaction is a =g2 a =4. The tree-level coupling coecient to the electrons of the DFSZ model is [27,28] Ce= cos20=3, where tan 0=vd=vu, the ratio of the vacuum expectation values of the Higgs eld. The axion-electron derivative part of the interaction can be expressed as Le=gaee 2me@a(x) e(x)
5e(x) ' igaeea(x)e(x)
5e(x); (4) where the last term is an equivalent Lagrangian obtained by using Dirac equation and neglecting quadridivergences. The Feynman diagram of this interaction is reported in Fig. 1 and suggests how the process happens: an axion is absorbed and causes the fermion to
ip its spin, and the macroscopic eect is a change in the magnetization of the sample containing the spin. aigaeeγ5e− e− Fig. 1. Feynman diagrams of the axion-fermion interaction, showing how the eect of the axion is to be absorbed and cause the spin
ip of the fermion. The corresponding interaction Lagrangian is reported in Eq. (4). By taking the non-relativistic limit of the Euler{Lagrange equation, the time evolution of a spin 1/2 particle can be described by the usual Schroedinger equation i~@' @t= ~2 2mer2 gaee~ 2meera '; (5) whereeis the Pauli matrices spin vector. The rst term on the right side of Eq. (5) is the usual kinetic energy of the particle, while the second one is analogous to the interaction between a spin and a magnetic eld. One can notice that gaee~ 2meera= 2e~ 2meegaee 2e ra= 2eegaee 2e ra; (6) sinceeis the magnetic moment of the particle, it can be both Bohr magneton or a nuclear magneton, depending on the considered fermion. From Eq. (6) it is clear that the eect of the axion is the one of a magnetic eld, but since it does not respect Maxwell's equations, calling it an eective magnetic eld is more appropriate Bagaee 2e ra: (7) This denition is useful to quantify the performances of a ferromagnetic haloscope in terms of usual magnetometers sensitivity. In the BCFV case, the signal is given by the electrons' magnetization. An intuitive connection between theEur. Phys. J. Plus ##################### #### Page 5 of 18 macroscopic magnetization and the spin is given by M/BNSwhereBis Bohr magneton and NSis the number of spins which take part to the magnetic mode [91]. In such a way, the spin-
ip can be classically considered a variation of the magnetization at a frequency given by the axion eld. It is now necessary to understand which are the features of Bato design a proper detector. The isothermal model of the Milky Way's DM halo predict a local density of %DM'0:45 GeV=cm3[92]. An Earth-based laboratory is thus subjected to an axion-wind with a speed va'300 km/s, that is the relative speed of Earth through the Milky Way. Using the vector notation, the value of vi acan be calculated from the speed of the galactic rest frame. The speed on Earthvi Eis given by the sum of vi S,vi Oandvi R, which are respectively the Sun velocity in the galactic rest frame (magnitude 230 km/s), the Earth's orbital velocity around the Sun (magnitude 29.8 km/s), and the Earth's rotational velocity (magnitude 0.46 km/s). The observed axion velocity is then vi a= vi E, which follow a Maxwell-Boltzmann distribution. As will be shown hereafter, the eect of this motion is a non-zero value of the axion gradient, and a modulation of the signal with a periodicity of one sidereal day and one sidereal year [47,93,94]. The numeric axion density in the DM halo depends on the axion mass and results na'31012(10 4eV=ma) cm 3. The coherence length of the axion eld is related to the de Broglie wavelength of the particles, which is given by a=h mava'1410 4eV ma m: (8) Such wavelength allows for the use of macroscopic samples to detect the variation of the magnetization. The large occupation number na, coherence length a, anda=va=c'10 3permit to treat Baas a classical eld1. The coherent interaction of a(x) with fermions has a mean value a(x) =a0eip ax=a0ei(p0 at pi axi); (9) wherepi a=mavi Eandp0 a=p m2a+jpiaj2'ma+jpi aj2=(2ma). The production of DM axions is discussed in Section 2, where it is shown that they are indeed cold DM since their momentum is orders of magnitude smaller than the mass. The axion kinetic energy is expected to be distributed according to a Maxwell-Boltzmann distribution, with a mean relative to the rest mass of 7 10 7and a dispersion about the mean of MB'510 7[47,93,95]. The eect of the mean is a negligible shift of the resonance frequency with respect to the axion mass. The consequence of the dispersion on the eective magnetic eld is a natural gure of merit Qa=1 MB'ma hpiai2 =1 2a'2106: (10) To calculate the eld amplitude a0, the momentum density of the axion eld is equated to the mean DM momentum density yielding a2 0p0 1pi a=nahpi ai=namava)a0=p na=ma: (11) For calculation purposes natural units are dropped and the Planck constant ~and speed of light care restored. The eective magnetic eld associated to the mean axion eld reads Bi a=gaee 2ena~ mac1=2 pi asinp0 act+pi axi ~ : (12) From Eq. (12), the frequency and amplitude of the axionic eld interacting with electrons result Ba=gaee 2ena~ mac1=2 mava'510 23ma 50eV T; !a 2'cp0 a ~=mac2 ~'12ma 50eV GHz:(13) As the equivalent magnetic eld is not directly associated to the axion eld but to its gradient, the corresponding correlation length and coherence time must be corrected to [84] ra'0:74a= 0:74~ mava'2050eV ma m; ra'0:68a= 0:682~ mav2a'4650eV maQa 1:9106 s:(14) The nature of the DM axion signal is now well-dened: an eective magnetic eld of amplitude Ba, frequency fa= !a=2, and quality factor Qawith values dened by Eq.s (13) and (14). 1The average speed vacalso justies the approximation of Eq. (5), i. e. the use of the non-relativistic limit of Euler- Lagrange equations.#### Page 6 of 18 Eur. Phys. J. Plus ##################### 4 The axion-to-electromagnetic eld transducer A viable experimental scheme must be designed to detect the eld Bi a, whose features are dened by Eq.s (13) and (14). A magnetic sample with a high spin density nSand a narrow linewidth
m= 2=T 2(i. e. long spin-spin relaxation timeT2) can be used as a detector. The magnetic eld Badrives a coherent oscillation of the magnetization over a maximum volume of scale ra. The sensitivity increases with the sample volume Vsup to (ra)3. For electrons
e'(2)28 GHz/T, so the corresponding magnetic eld B0is of order 1 T and experimentally readily obtainable. The electrons' spins of a magnetic sample under a uniform and constant magnetic eld result in a magnetization M(x;t) that can be divided in magnetostatic modes. The space-independent mode of uniform precession is called Kittel mode. The axionic eld couples to the components of Mtransverse to the external eld, depositing power in the material. More power is deposited if the axion eld is coherent with the Kittel mode for a longer time. The best-case scenario is a material with a quality factor Qm=
eB0=
mwhich matches Qa, so that the coherent interaction between spins and DM-axions lasts for ra. For this reason the magnetic eld uniformity over the sample must be1=Qmto avoid inhomogeneous broadening of the ESR. According to these considerations, it is possible to detect an axion-induced oscillation of the magnetisation by monitoring a large sample with an precise magnetometer. However, the limit of this scheme lies in the short coherence time of the magnetic sample. In fact at high frequency, i. e. above 1 GHz, the rate of dipole emission becomes higher than the intrinsic material dissipation, this eect is know as radiation damping [96]. Since radiation damping is related to the sample dipole emission, a possible way to reduce its contribution is to limit the phase-space of the radiated light by working in a controlled environment like a resonant cavity [96,97,98,99]. By housing the sample in a mw cavity and tuning the static magnetic eld such that !m=
eB0'!c, where!cis the resonance frequency of a cavity mode with linewidth
c, one obtains a photon-magnon hybrid system (PMHS). An exact description of the system is given by the Tavis-Cummings model [99]. It discusses the interaction of NStwo-level systems with a single mw mode, and predicts a scaling of the cavity-material coupling strength gcm/pNS. The single-spin coupling is gs=
e 2s 0~!m Vc; (15) whereVcis the cavity volume, 0is the vacuum magnetic permeability and a mode-dependent form factor [100], with this relation gcm=gspNS. Such scaling has been veried experimentally down to mK temperatures for an increasing number of spins NS[100,101,102]. For a quantity of material such that gcm
c, the single cavity mode splits into two hybrid modes with frequencies !+,! and 2gcm=!+ ! . For!m=!cthe linewidths of the hybrid modes are the average of the cavity mode linewidth
cand of the material one
m, namely
h= (
c+
m)=2. The coupling gcmis in fact a conversion rate of the material magnetization quanta (magnons) to cavity photons and viceversa. If gcm>
h the system is in the strong-coupling regime, meaning that for a magnon (photon) it is more likely to be converted than to be dissipated. In this way, magnetisation
uctuations, which might be induced by axions, are continuously converted to electromagnetic radiation that be collected with an antenna coupled to the cavity mode, as schematically shown in Fig. 2. a ωam ωmc ωcgam≪1 gcm∼1 Fig. 2. The coupled harmonic oscillators are reported in orange, green and blue for cavity c, materialmand axionarespectively. The uncoupled normal-modes frequencies of the HOs are !c,!mand!aand the couplings are gamandgcm, represented by springs. The aim of a PMHS devised for an axion haloscope is to maximise the axionic signal, and it eectively works as an axion-to-photon transducer. A rendering of the resulting device is reported in Fig. 3. An important result for designing the transducer is that multiple spheres can be coherently coupled to a single cavity mode [103,102]. The measurements demonstrate that all the spins participate in the interaction, thus the samples act as a single oscillator. This is guaranteed by the fact that the static eld is uniform over the spheres and that the rf eld is degenerate over the axis of the cavity where they are placed. Several tests were performed to understand dierent features of the system in the light of the two properties mentioned before. To understand the results of the dierent measurements one can use a simple oscillators model as is done in [102]. The PMHS can beEur. Phys. J. Plus ##################### #### Page 7 of 18 described by introducing two magnon modes and two cavity modes, hereafter the photon modes are labeled as cand dwhile the magnon modes are mandn. In the matrix form, the system can be modeled by the hamiltonian Hcdmn =0 B@!c i
c=2gcdgcm gcn gcd!d i
d=2gdm gdn gcmgdm!m i
m=2gmn gcngdngmn!n i
n=21 CA; (16) where!,
andgare the frequencies, linewidth and coupling of the dierent modes. The autofunction of the system can be calculated as the determinant of !I4 H cdmn thus the function used to show the anticrossing curve reads fcdmn(!) = det !I4 H cdmn : (17) To ideally have a coherent coupling, one needs to let the spins of dierent spheres cooperate and make them indistinguishable, so they need to be uncoupled and their resonant frequencies must be the same. These conditions translate to gmn= 0 and!m=!n, which clearly can be extended to an arbitrary number of oscillators (in this case, the ten spheres). The interaction between two spheres yields non-zero value of gmn, and its eect is to introduce other resonances besides the two main ones of the PMHS. This eect needs to be avoided to have control over the system and couple all the spins of the samples to the cavity mode, avoiding magnons bouncing between dierent magnetic modes and eventually being dissipated before their photon conversion. Fig. 3. Rendering of the whole system, constituted by the cavity and the pipe with ten YIG spheres, ready to be tested at milli-Kelvin tempera- tures. The external part shows the superconducting magnet (in brown) which surrounds the cavity and the spheres to provide a magnetic eld with uniformity better than 7 ppm. The magnet is immersed in the liquid helium bath outside the vacuum chamber of the dilution unit. The cavity is at the centre of the magnet, is anchored to the mixing chamber of the dilution refrigerator with two copper bars and is equipped with two an- tennas, one is xed and weakly coupled, while the second one is movable and is used to extract the signal. The YIG spheres are inside the cavity, held by a fused silica pipe lled with helium and separated by thin PTFE spacers. The cap used to seal the pipe is made of copper and is anchored to the cavity body to ensure the thermalisation of the exchange helium and therefore of the YIG spheres. YIG sphere were produced on site with a technique described in [102]. This open the possibility of studying spheres of dierent diameters coupled to the same mode. One of the ndings is that, trying to couple spheres with dierent diameter to the same mode, the volume of the sample is linearly related to the oset eld [102]. The axion- to-electromagnetic eld transducer of a ferromagnetic haloscope. The constraints to remember for its design are in the following, and were tested with a room temperature setup consisting in a 10.7 GHz cavity with conical endcaps and a fused silica pipe holding the YIG spheres. The magnetic eld is given by a SC magnet which, to perform quick tests, it is equipped with a room temperature bore allowing the magnet to be in a liquid helium bath during operation. First the minimum separation between two spheres is tested by gradually increasing the distance between them and verifying that a usual anticrossing curve is reproduced. The minimum distance between 2 mm spheres results in 3 mm. A YIG sample then occupies 5 mm of space, and since the cylindrical part of the cavity is 6 cm it can house a maximum of twelve samples. Ten spheres are inserted in the pipe for them not to be too close to the conical part of the cavity. Multiple spheres of dierent diameters were fabricated and rened to verify that they hybridize with the cavity for the same value of the magnetic eld.#### Page 8 of 18 Eur. Phys. J. Plus ##################### The setup must ensure a proper thermalization of the cavity and of the YIG spheres, the preparation of the fused silica pipe is as follows. A vacuum system is designed in such a way to empty the pipe from air which is then immersed in a 1 bar helium controlled atmosphere. This way the pipe is lled with helium, and can be sealed by using a copper plug and Stycast. First the sealing is tested without the samples by measuring the shift of the TM110 mode of the cavity-pipe system with and without helium. The frequency is measured with the helium-lled pipe, which is then immersed in liquid nitrogen and again placed in the cavity. Re-measuring the same frequency excludes the presence of leaks. The used cavity is made of oxygen-free high-conductivity copper, and features a cylindrical body with two conical endcaps, as shown in Fig. 3. The central body is not a perfect cylinder but it has two
at surfaces used to remove the angular degeneration of the mode. This creates two modes rotated of =2 with dierent frequencies, which is the second cavity mode in Eq. (16). The function fcdmn(!) is tted to the measured PMHS dispersion relation to extract the parameter of our setup, and in particular the hybridization results 638 MHz which is compatible with the single 1 mm sphere since 638 MHz =p810 = 71 MHz. This value of the single sphere coupling is compatible with what previously obtained in simpler PMHS, indicating that the measured spin density of YIG is consistent both with the previous results and with the values reported in the literature. Remarkably, the lower frequency resonance is almost unaected by the behaviour of the rest of the PMHS, in the sense that its frequency does not dier from the one of a usual anticrossing curve, thus it can possibly be safely used for a measurement [104]. Since haloscopes need to scan multiple frequencies to search for axions, the resonant frequency of the PMHS mode used for the measurement need to be changed. The tuning is made extremely easy by the fact that it is controlled only by means of the external magnetic eld. A high stability of B0is necessary to perform long measurements over a single frequency band. This is set by the linewidth of the hybrid mode, which in this case is 2 MHz, and is tuned to cover a range close to 100 MHz [1,104]. Thanks to the anticrossing curve it is easy to identify the frequency of the correct mode to study. The hybrid mode is not aected by disturbances caused by other modes in a range that largely exceeds ten times its linewidth. These clean frequencies are selected for the measurements whenever it is possible to match them with the working frequencies of the amplier described in the next Section. 5 Quantum-limited amplication chain The PMHS described previously in this Chapter acts as a transducer of the axionic signal. The power coming from the PMHS must be measured and acquired with a suitable detection chain, and, as it is extremely weak, needs to be amplied. The intrinsic noise of an haloscope is essentially related to the temperature of the setup, and since axionic and Johnson power have the same origin it is the ultimate limit of the SNR. The amplication process inevitably introduces a technical noise which, for these setups, is useful to quantify in terms of noise temperature Tnto compare to the Johnson noise. This stage of the measurement is setting the overall sensitivity of the apparatus since, as shown hereafter, for very low working temperatures the noise temperature is higher than the thermodynamic one. Minimizing Tnis a key part of the development of an haloscope, and is complementary to the maximization of the axion deposited power. The mw ampliers used for precision measurement are mostly high electron mobility transistors (HEMT), since they have high gain and low noise, of the order of 4 K. The most sensitive amplier available is the Josephson parametric amplier (JPA), which reaches the quantum noise limit [105,106,107,108,109,110,111]. This type of amplier is used in the present haloscope, and its performances can be overcome only by using a photon counter. HEMT are eld eect transistors based on an heterojunction, i. e. a PN junction of two materials with dierent band gaps [112]. The proper doping prole and band alignment gives rise to extremely high electron mobilities, and thus to ampliers which can have high gain, very low noise temperature, and working frequency in the microwave domain. Even if their noise temperature is low, HEMTs are not the most sensitive ampliers available. At a frequency 10 GHz the SQL of linear ampliers is close to 0.5 K, which is about one order of magnitude lower than theTnof HEMTs. Such remarkably low Tnis achieved by JPAs, resonant ampliers with a narrow bandwidth but with quantum-limited noise. This feature makes them the ideal tool to measure faint rf signals, and thus to be implemented in ferromagnetic or Primako haloscopes. The non-linear mixing is given by a Josephson-RLC circuit with a quadratic time-dependent Hamiltonian, which can be degenerate or non-degenerate depending on whether the signal and idler waves are at the same frequency or not [113]. A non-degenerate device consists in a three-modes, three-input circuit made of four Josephson junctions forming a Josephson ring modulator. It eectively is a three-wave purely dispersive mixer which can be used for parametric amplication [114]. It can be computed that the non-linear mixing process appears as a linear scattering, conguring the JPA is a linear amplier. As such, it is quantum limited and its noise temperature depends on the working frequency. Being based on resonant phenomena, the JPA has a narrow working band of tens of MHz. To use the amplier in a wide frequency range, a bias eld is applied to the ring and the resonance frequencies of the signal, idler and pump mode are tuned. This is achieved with a small SC coil placed below the ring, biased with a current Ib. The implementation of a JPA in a ferromagnetic haloscope is shown and described in Fig. 4.Eur. Phys. J. Plus ##################### #### Page 9 of 18 Fig. 4. Rendering of the implementation of this JPA in a ferromagnetic haloscope. The golden pipe is connected to the mixing chamber of the dilution refrigerator used to cool down the setup, the circulators are only in thermal contact with this last stage, as is the shielding cage of the JPA (also drawn in gold). The blue component is a switch, present in one of the possible congurations of a ferromagnetic haloscope; attenuators are drawn in blue as well. The JPA is inside two concentric cans, the external one is made of Amuneal and the external is of aluminum. The rst is useful to reduce the Earth magnetic eld in which the superconducting parts (shields and junctions) undergo the transition, while the second screens from external disturbances. Everything is attached to the mixing chamber plate of a dilution refrigerator with a base temperature of about 90 mK. This image corresponds to the conguration reported in Fig. 5b. The characterisation the rf chain used for the measurements is described hereafter. It will focus on the setup described by Fig. 5a, as is the one used in [1] as it was found to be more reproducible and in general more reliable than 5b. The conguration of the electronics allows the testing of both the JPA and the PMHS. Transmission measurements of the PMHS can be performed by turning o the JPA (i.e. no bias eld and no pump) to re
ect the signal on it, the input is the SO line and the output is the readout line. The JPA can be tested with the help of the Aux line, by uncoupling the antenna from the cavity and re
ecting the incoming signal. Some rf is still absorbed at the cavity modes frequency but this does not compromise the measurement. The external static eld of the PMHS does not aect the resonances of the Josephson ring modulator as no dierence has been detected between the measurements with and without eld. Runs are performed with bias currents Ib'170A and 460A at frequencies ranging from 10.26 GHz to 10.42 GHz. Using the SO line and critically coupling the antenna to the hybrid mode, a signal is injected in the system and read with the whole amplication chain. The rst test is to verify the linearity of the JPA (and of the whole chain) using signals of growing intensity until the system saturates. These measurements show the linear and saturate behavior of the amplier. It is possible to calibrate the gain of the JPA by using a signal large enough to be measured with the JPA o but also not to saturate it once it is turned on. This is useful to know the gain of the amplier at the dierent working points to have a preliminary calibration of the system and to understand whether an output noise with higher amplitude is due to the JPA or to something else. Since the electronics above the 4 K line was already characterized for the previous prototype, the baseline noise with JPA o is roughly the amplier noise temperature T(hemt) n'10 K. The measured noise spectra with the JPA turned o is white in a bandwidth of several hundreds of MHz, when the parametric amplier is turned on its resonance exceeds this noise of roughly 10 dB. Since it is possible to calibrate the gain of the JPA GJPA'20 dB, the amplied noise level can be extracted as T(JPA) n =T(hemt) n=10(GJPA 10 dB)=10'1 K, which is the noise temperature amplied by the JPA. The value of 1 K is reasonable, since a single quantum at this frequency is 0.5 K such noise corresponds to two quanta. Even if this procedure is somewhat correct, it is not a proper calibration of the setup and something better is explained hereafter. Since some problems were encountered in the noise calibration with hot load (see Fig. 5b), the rf setup of Fig. 5a is designed to calibrate all the dierent lines with the help of the variable antenna coupling. By moving the antenna one can arbitrarily choose the coupling to a mode, if it is weakly coupled a test signal from the Aux port gets re
ected and goes to the JPA, while if the antenna is critically coupled to the mode, a signal from SO is transmitted through the cavity and than to the JPA. Almost the same result can be obtained by slightly changing the frequency of the test signal to be within the JPA band but out of the cavity resonance. The critical coupling can be reached by doubling the linewidth of the mode or equivalently by minimizing the re
ected signal from the Aux line to the Readout line. The procedure to calibrate all the lines is: 1. with the weakly coupled antenna or by detuning the mode the losses of the Aux-Readout line LARare measured;#### Page 10 of 18 Eur. Phys. J. Plus ##################### (a) (b) Fig. 5. Two possible electronics layout for a ferromagnetic haloscope. The blue lines show the temperature ranges, the crossed rectangles are the magnet, and the orange rectangle is the cavity with black YIG circles inside. The boxed numbers are attenuators and the red circled Ts are the thermometers. At the top of the cavity are located the weakly coupled antenna (empty dot) and the variably-coupled antenna (full dot). The weakly coupled antenna is connected to an attenuator and then to the source oscillator SO. In conguration (a) the the variable antenna is connected to the JPA through a circulator, whose other input is used for auxiliary measurements. The output of the JPA is further amplied by two HEMTs A1 and A2. Conguration (b) is basically the same as (a), where the input can be switched from the cavity antenna to a matched load with variable temperature regulated by a current Ih, and used for calibration. In both (a) and (b) the A2 output is down-converted and acquired. 2. the antenna is critically coupled to the mode and a signal is sent through the Aux-SO line to get LAS; 3. with the same critical coupling the transmission of the SO-Readot line LSRis acquired. At this point a signal of power Ainis injected in the SO line, the fraction of this power getting into the cavity through the weakly coupled antenna is Acal=AinLSO. The attenuation of the line can be calculated as LSO'p LSRLAS=LAR, which gives the power collected by the critically coupled antenna. Since Acalis eectively a calibrated signal, it can be used to measure gain and noise temperature of the Readout line. Dierent Ainare used to get increasingly large signals to be detected by the JPA-based chain. This calibration has some minor biases, the rst is given by the cable from the cavity to the rst circulator which is accounted for two times in the Aux-Readout line. This contribution can be safely neglected as the cable is superconducting, making its losses negligible. Another bias is related to the antenna coupling, which is not perfect. With a proper antenna coupling the re
ected signal is reduced of 10 dB, so there is a bias of a factor 10% intrinsic to the measurement which will be accounted for when calculating the error. As the calibration procedure is long it is not repeated for every run, however no important dierences are expected when changing the JPA frequency. As an example a run at 10.409 GHz is considered. The gain of the JPA at this frequency results GJPA'18 dB, and its bandwidth is 8 MHz, the hybrid mode is tuned the its central frequency and the calibration procedure is carried out,resulting in a noise temperature of Tn= 1:0 K and the total gain of the whole amplication chain is Gtot'120:4 dB. The value of Tnis compatible with the one estimated previously, and corresponds to two quanta. The coupling of the antenna with the hybrid mode is checked for every run. It is controlled by moving the dipole antenna in and out the cavity volume, the critical coupling is reached when the uncoupled linewidth of the mode is doubled. To verify the proper antenna positioning one may rely on the fact that depending on the temperature dierence between the cavity and a 50 , some power may be absorbed or added to the load thermal noise. The load under consideration is the hottest between the rst JPC isolator and the 20 dB attenuator of the Aux line. The hybrid resonance has a critical linewidth of about 2 MHz, so the depth will not be as narrow as the one of the cavity. In that case the temperature dierence is about 10%, which is about 10 mK, and if the temperature of the load and cavity are precisely measured the spectra can be used to get a two-points calibration. The selected calibration procedure was notEur. Phys. J. Plus ##################### #### Page 11 of 18 this one because the temperatures of loads and HS are not easily accessible. With two dedicated thermometers the temperatures of the loads could be measured, but it is not trivial to measure the temperature of the cavity and of the spheres with the needed precision. Since a small temperature dierence is expected, the measurement with the antenna coupled to the hybrid mode should be dierent from the uncoupled one. As reported in [1], there is a dierence between the two measurements and it is compatible with the thermal noise of the hybrid mode at a temperature slightly higher then the loads one. 6 Data acquisition and analysis A ferromagnetic haloscope's scientic run consists in several measurements in the common bands between the frequen- cies of the lower hybrid mode unaected by disturbances, and the JPA working range. The low temperature electronics is described in the previous section, and is completed by its following part hereafter. The room temperature electronics consists of a HEMT amplier (A2) followed by an IQ mixer used to down-convert the signal with a local oscillator (LO). In principle, it is possible to acquire the signal coming from both hybrid modes using two mixers working at f+andf . In this case it is chosen to work only with f+, thus setting the LO frequency tofLO=f+ 0:5 MHz. The amplied antenna output at the hybrid mode frequency is down-converted in the 0 - 1 MHz band, allowing to eciently digitize the signal. The phase and quadrature outputs are fed to two low frequency ampliers (A3 I;Q), with a gain of G3'50 dB each, and are acquired by a 16 bit ADC sampling at 2 MS/s (see [103]). A dedicated DAQ software is used to control the oscillators and the ADC, and veries the correct positioning of the LO with an automated measurement of the hybrid mode transmission spectrum. Some other online checks include a threshold monitor of the average amplitude, as well as of the peak amplitude, which
ags the le if some unexpected large signal is present. The ADC digitizes the time-amplitude down-converted signal coming from A3 Iand A3Qand the DAQ software stores collected data binary les of 5 s each. The software also provides a simple online diagnostic, extracting 1 ms of data every 5 s, and showing its 512 bin FFT together with the moving average of all FFTs. The signal is down-converted in its in-phase and quadrature components fngandfqng, with respect to the local oscillator, that are sampled separately. Fig. 6. Second amplication stage of the setup, and rst room temperature amplier. The image shows the top part of the vacuum vessel containing the dilution fridge stages, the cavity and the electronics. Just outside it, the rst amplier is A1 (reported in yellow), while the second one is already at room temperature. The blue box corresponds to the variable temperature load of conguration (b). The stability of the measurement is tested by injecting a signal in the SO line slightly o resonance with the PMHS peak, and with an amplitude guaranteeing a large SNR. Monitoring its amplitude is a way to continuously check the peak position. In this setup the stability results well below the percent level, which is more than enough for the purpose of the experiment, thanks to the lower and more stable working temperature and to an extremely stable current generator produced in the Padua University electronic workshop. The signal is analysed using a complex FFT on the combination of phase and quadrature fsng=fng+ifqngto get its power spectrum s2 !with positive frequencies for f >f LOand negative frequencies for f <f LO, in a total bandwidth#### Page 12 of 18 Eur. Phys. J. Plus ##################### of 2 MHz which contains the whole hybrid linewidth. Some bins are found to be aected by disturbances resulting in systematic noise, but they can be easily removed by analyzing the data with longer FFT. Using 32768 points the single frequency resolution is 61 Hz
aand thus single bins which do not respect the
uctuation dissipation theorem can be substituted with the average of the ten nearest neighbors. This procedure articially reduces the variance of the data, but the number of corrupted bins is negligible and so is the eect on the variance. This process does not cut the signal since it aects only known or single bins, while the axion signal is expected to be distributed over many. For debugging purposes a simulated signal is injected into the analysis code. It is created by generating a high- frequency noise to which are added a simulated axion signal and some disturbances. For creating the in-phase and in-quadrature component it is multiplied by a sine or to a cosine wave, to then extract only one point every 104and simulate the mixing and down-conversion processes. The FFT of this signal produces a white noise with some peaks (the axion signal plus disturbances), and is eventually integrated. The analysis procedure is veried to remove bad bins and to preserve the signal and SNR. Thanks to the stability and to the tests on the acquisition and analysis, all the les of a single run can be safely RMS averaged together. The calibration gives a noise temperature Tn= 1:0 K, which includes all the amplier noises, the cavity thermodynamic noise, and the losses. This can be used to calibrate the setup by setting the mean value of the FFT to kBTnf. As for the calibration errors, the
uctuations of the cavity temperature are of order 10 mK, thus they can change Tnof a fraction close to 1%. A larger contribution is intrinsic to the procedure, which requires the coupling of the antenna to be changed from weak to critical. This is believed to be the larger contribution to the uncertainty of the measurement, and even if in principle this is not a
uctuation it will be used to estimate an error. The fact that the coupling is close to critical is also supported by a thermal noise measurement [1]. The control over this parameter can be estimated by injecting a signal through the Aux port in the weak-coupling position, and then by reducing the re
ected amplitude by increasing the coupling. Typically the signal power can be decreased of more than 8 dB, this value can be used to estimate the larger uncertainty on the coupling resulting in a 16% error. To estimate the sensitivity to the axion eld the resolution bandwidth is set to 5 kHz, which at this frequency is the value producing the best SNR. The spectrum is tted with a degree ve polynomial to extract the residuals, whose standard deviation is the sensitivity of the apparatus in terms of power. For every run, by using the run length, the RMS power is compared with the estimated sensitivity obtained with Dicke radiometer equation. The measured
uctuations are close to the expectations for almost every run, and are always compatible when considering the 16% error which was previously calculated. This shows that the measured power is compatible thermal noise and that the integration is eective over the whole measurement time, since it follows the 1 =p time trend. The best sensitivity reached isP= 5:110 24W for an integration time of 9.3 h, corresponding to an excess rate .10 photons/s per bin. As is discussed in Section 1 this setup is a spin-magnetometer, as it is sensitive to variations of the sample magnetization. In this sense, it is interesting to determine the performances of this setup to get some physical intuition on its sensitivity and thus on all the possible phenomena that could be measured. The eld sensitivity can be estimated for a eld whose coherence length stretches on all the YIG spheres, and whose linewidth is narrower than the PMHS one. In the case of the present apparatus the threshold coherence length and time are then 10 cm and 100 ns, respectively. Iff and are the frequency and coherence time of the hybrid mode and NSis the number of spins, the magnetic eld sensitivity of the setup is B=P 4
eBf NS1=2 = 5:510 19h10:4 GHz f 83 ns 1:01021spins NSi1=2 T; (18) which is the minimum eective magnetic eld detectable by the spin magnetometer for a unitary SNR. A 0.5 aT sensitivity is remarkable by itself, and shows the high potential of HS-based magnetometers. The results of the previous section are used to extract a limit on the axion-electron coupling, which is a possible eect that modulates the sample magnetization and produces an excess of photons. The expected power deposited by DM-axions in the PMHS is Pout= 710 33ma 43eV3NS 1:01021spinsh 83 ns W: (19) The 10.4 GHz photon rate corresponding to this power is ra= 10 9Hz. By comparing the rate rawith(3) P=(~! ) one obtains that there are ten orders of magnitude to get the sensitivity required to detect the axion, which corresponds to ve in terms of eld (and thus of coupling constant). As discussed in Section 1, even if an instrument is not capable of limiting the QCD-axion parameter space it can still probe the presence of ALPs. These can also constitute the totality of DM and their mass and coupling are unrelated. The present measurements are used to extract a limit on the coupling of ALPs with electrons reported in [1]. No evidence of a signal due to axions has been detected, and theEur. Phys. J. Plus ##################### #### Page 13 of 18 measured spectra are compatible with noise. The 2 (95% C. L.) upper limit on the coupling reads gaee>e mavas 2tac(3) P 2B
naNS ; (20) wheretacis a frequency dependent coecient that takes into account that the axion-deposited power is not uniform in the haloscope operation range, as is discussed in [115]. All the experimental parameters used to extract the limits are measured within every run, making the measurement highly self-consistent. The limit on the ALP-electron coupling described by Eq. 20 is calculated for every run using the corresponding measured parameters. This result is compared with other techniques used for testing the axion-electron couling constant Fig. 7. Overview of axion searches based on their coupling with electrons [116]. The result obtained with the ferromagnetic haloscpe described in this work is labelled as \QUAX". The analysis is repeated by shifting the bins of half the RBW to exclude the possibility of a signal divided into two bins. The best limit obtained, and corresponding to P, isgaee<1:710 11. The improvement of a longer integration time is not much, since the limit on the coupling scales as the fourth root of time, and to improve the current best limit of a factor 2 the needed integration time is six days. 7 Conclusions Low-energy measurements, precisely testing known physical laws, are a powerful probe of BSM physics, mainly com- plementary to accelerator physics. As shown in Fig. 8, thanks to Nambu-Goldstone theorem, extremely high energy scales can be explored by measuring faint eects at the limit of present technology. Eventually, new instruments and devices can be built to push the current technological limits to new levels and hopefully help not only fundamental physics but many other elds. Among these, haloscopes play a pivotal role while searching for Dark Matter axions. The scope of this work is to illustrate the construction and outline the operation of the rst ferromagnetic axion haloscopes. Such instruments can be used to measure the DM-axion wind which blows on Earth, as this last one is moving through the halo of the Milky Way. The axions interact with the spin of electrons causing spin
ips that are, macroscopically, oscillations of a sample magnetization. The model of the isothermal galactic halo and of the axion yield the features of the searched for signal, namely its linewidth, frequency and amplitude. A frequency of 10 GHz, a linewidth of 5 kHz and an amplitude of about 10 23T are expected for axion masses of order 40 eV. A proper haloscope has a transducer that converts the axion
ux into rf power, followed by a sensitive detector to measure it. For a ferromagnetic haloscope, the transducer consists of a magnetic material containing the electron spins with which the axions interact. In order to maximize the axion- deposited power, the sample should have a large spin density and a narrow linewidth. The detector is a rf amplication chain based on a JPA. The described instrument features a power sensitivity limited by quantum
uctuations, in this sense no linear amplier is or will be able to improve the haloscope. In future setups only bolometers or quantum#### Page 14 of 18 Eur. Phys. J. Plus ##################### 10−6eV 104GeVLHC 1012GeVNambu-Goldstonetheorem Precision tests Symmetry breaking scale Fig. 8. The usage of Nambu-Goldstone theorem to infer on physics at energy scales inaccessible to accelerators. counters can yield better results. For example, recent developments on quantum technologies [117,118] demonstrated the detection of
uorescence photons emitted by an electron spin ensemble, and could be adopted for axion searches. Thermodynamic
uctuations are already negligible due to the extremely low working temperature, so it is not necessary to decrease them by orders of magnitude. As the rate of thermal photons of a cavity mode is exponentially decreasing with temperature, the present dilution refrigeration technology is enough to reach the axion-required noise level. As discussed in Section 1, to get a rate of axion-induced photons which can be measured in a reasonable amount of time, a much increased quantity of material and a narrower linewidth are required. This setup features 0.05 cc of YIG, such volume must be increased by three orders of magnitude to get the required rate. This large quantity can be achieved by increasing the quantity of material in a single cavity and the number of cavities. In conclusion, the successful operation of an ultra cryogenic quantum-limited prototype demonstrates the possibility of scaling up the setup of orders of magnitude without compromising its sensitivity. To further increase the axionic signal there are two parameters to work on: the hybrid mode linewidth and the sample spin-density and volume. To nally achieve the sensitivity required by a QCD axion search, it is necessary to use a photon counter. The upgrades planned until now are implemented and result eective, as the apparatus behaves as expected. No showstoppers were identied so far. 8 Acknowledgment N.C. is thankful INFN and the Laboratori Nazionali di Legnaro for hosting and encouraging the experiment. The help and support of Giovanni Carugno and Giuseppe Ruoso is deeply acknowledged. References 1. N. Crescini, D. Alesini, C. Braggio, et al. Axion search with a quantum-limited ferromagnetic haloscope. Phys. Rev. Lett. , 124:171801, May 2020. 2. F. Zwicky. Die Rotverschiebung von extragalaktischen Nebeln. Helvetica Physica Acta , 6:110{127, 1933. 3. V. C. Rubin, N. Thonnard, and W. K. Ford, Jr. Extended rotation curves of high-luminosity spiral galaxies. IV - Systematic dynamical properties, SA through SC. Astrophysical Journal , 225:L107{L111, November 1978. 4. V. C. Rubin, W. K. Ford, Jr., and N. Thonnard. Rotational properties of 21 SC galaxies with a large range of luminosities and radii, from NGC 4605 /R = 4kpc/ to UGC 2885 /R = 122 kpc/. Astrophysical Journal , 238:471{487, June 1980. 5. Gerard Jungman, Marc Kamionkowski, and Kim Griest. Supersymmetric dark matter. Physics Reports , 267(5):195 { 373, 1996. 6. Giorgio Arcadi, Ma ra Dutra, Pradipta Ghosh, et al. The waning of the wimp? a review of models, searches, and constraints. The European Physical Journal C , 78(3):203, Mar 2018. 7. Vasiliki A Mitsou. Overview of searches for dark matter at the LHC. Journal of Physics: Conference Series , 651:012023, nov 2015. 8. Nicolao Fornengo. Status and perspectives of indirect and direct dark matter searches. Advances in Space Research , 41(12):2010 { 2018, 2008. 9. Joerg Jaeckel and Andreas Ringwald. The low-energy frontier of particle physics. Annual Review of Nuclear and Particle Science , 60(1):405{437, 2010. 10. Steven Weinberg. Approximate symmetries and pseudo-goldstone bosons. Phys. Rev. Lett. , 29:1698{1701, Dec 1972. 11. Roberto D. Peccei. The Strong CP Problem and Axions , pages 3{17. Springer Berlin Heidelberg, Berlin, Heidelberg, 2008. 12. G. 't Hooft. Symmetry breaking through bell-jackiw anomalies. Phys. Rev. Lett. , 37:8{11, Jul 1976.Eur. Phys. J. Plus ##################### #### Page 15 of 18 13. G. 't Hooft. Computation of the quantum eects due to a four-dimensional pseudoparticle. Phys. Rev. D , 14:3432{3450, Dec 1976. 14. P. Di Vecchia. G. Veneziano R.J. Crewther and E. Witten. Chiral estimate of the electric dipole moment of the neutron in quantum chromodynamics. Physics Letters B , 91(3{4):487, 1980. 15. Maxim Pospelov and Adam Ritz. Electric dipole moments as probes of new physics. Annals of Physics , 318(1):119 { 169, 2005. Special Issue. 16. Norman F. Ramsey. Electric-dipole moments of particles. Annual Review of Nuclear and Particle Science , 32(1):211{233, 1982. 17. J. H. Smith, E. M. Purcell, and N. F. Ramsey. Experimental limit to the electric dipole moment of the neutron. Phys. Rev., 108:120{122, Oct 1957. 18. C. A. Baker et al. An Improved experimental limit on the electric dipole moment of the neutron. Phys. Rev. Lett. , 97:131801, 2006. 19. J. M. Pendlebury, S. Afach, N. J. Ayres, et al. Revised experimental upper limit on the electric dipole moment of the neutron. Phys. Rev. D , 92:092003, Nov 2015. 20. R. D. Peccei and Helen R. Quinn. CP conservation in the presence of pseudoparticles. Phys. Rev. Lett. , 38:1440{1443, Jun 1977. 21. Frank Wilczek. Problem of strong p and t invariance in the presence of instantons. Phys. Rev. Lett. , 40(5):279, 1978. 22. Steven Weinberg. A new light boson? Phys. Rev. Lett. , 40(4):223, 1978. 23. Jihn E. Kim. Weak-interaction singlet and strong CP invariance. Phys. Rev. Lett. , 43:103{107, Jul 1979. 24. Jihn E. Kim. A review on axions and the strong cp problem. AIP Conference Proceedings , 1200(1):83{92, 2010. 25. M.A. Shifman, A.I. Vainshtein, and V.I. Zakharov. Can connement ensure natural cp invariance of strong interactions? Nuclear Physics B , 166(3):493 { 506, 1980. 26. M. Tanabashi et al. Review of Particle Physics. Phys. Rev. , D98(3):030001, 2018. 27. A. R. Zhitnitsky. The weinberg model of the cp violation and t odd correlations in weak decays. Sov. J. Nucl. Phys. , 31:529{534, 1980. [Yad. Fiz.31,1024(1980)]. 28. Michael Dine, Willy Fischler, and Mark Srednicki. A simple solution to the strong cp problem with a harmless axion. Physics Letters B , 104(3):199 { 202, 1981. 29. Michael Dine and Willy Fischler. The not-so-harmless axion. Physics Letters B , 120(1):137 { 141, 1983. 30. Laurence F Abbott and P Sikivie. A cosmological bound on the invisible axion. Physics Letters B , 120(1-3):133{136, 1983. 31. John Preskill, Mark B. Wise, and Frank Wilczek. Cosmology of the invisible axion. Physics Letters B , 120(1):127 { 132, 1983. 32. Richard Lynn Davis. Goldstone bosons in string models of galaxy formation. Phys. Rev. D , 32:3172{3177, Dec 1985. 33. R.L. Davis. Cosmic axions from cosmic strings. Physics Letters B , 180(3):225 { 230, 1986. 34. S. Chang, C. Hagmann, and P. Sikivie. Studies of the motion and decay of axion walls bounded by strings. Phys. Rev. D , 59:023505, Dec 1998. 35. David H. Lyth. Estimates of the cosmological axion density. Physics Letters B , 275(3):279 { 283, 1992. 36. Michiyasu Nagasawa and Masahiro Kawasaki. Collapse of axionic domain wall and axion emission. Phys. Rev. D , 50:4821{ 4826, Oct 1994. 37. David J.E. Marsh. Axion cosmology. Physics Reports , 643:1 { 79, 2016. 38. Pierre Sikivie. Invisible axion search methods. Rev. Mod. Phys. , 93:015004, Feb 2021. 39. Bonati, Claudio, D'Elia, Massimo, Mariti, Marco, et al. Recent progress on qcd inputs for axion phenomenology. EPJ Web Conf. , 137:08004, 2017. 40. Florian Burger, Ernst-Michael Ilgenfritz, Maria Paola Lombardo, Michael M uller-Preussker, and Anton Trunin. Topology (and axion's properties) from lattice qcd with a dynamical charm. Nuclear Physics A , 967:880 { 883, 2017. The 26th International Conference on Ultra-relativistic Nucleus-Nucleus Collisions: Quark Matter 2017. 41. Evan Berkowitz, Michael I Bucho, and Enrico Rinaldi. Lattice qcd input for axion cosmology. Physical Review D , 92(3):034507, 2015. 42. Szabolcs Bors anyi, Z Fodor, J Guenther, et al. Calculation of the axion mass based on high-temperature lattice quantum chromodynamics. Nature , 539(7627):69, 2016. 43. Peter Petreczky, Hans-Peter Schadler, and Sayantan Sharma. The topological susceptibility in nite temperature qcd and axion cosmology. Physics Letters B , 762:498{505, 2016. 44. Giovanni Grilli di Cortona, Edward Hardy, Javier Pardo Vega, and Giovanni Villadoro. The qcd axion, precisely. Journal of High Energy Physics , 2016(1):34, Jan 2016. 45. Georg G Raelt. Stars as laboratories for fundamental physics: The astrophysics of neutrinos, axions, and other weakly interacting particles . University of Chicago press, 1996. 46. Georg G. Raelt. Astrophysical methods to constrain axions and other novel particle phenomena. Physics Reports , 198(1):1 { 113, 1990. 47. Michael S Turner. Windows on the axion. Physics Reports , 197(2):67{97, 1990. 48. Paolo Gondolo and Georg G. Raelt. Solar neutrino limit on axions and kev-mass bosons. Phys. Rev. D , 79:107301, May 2009. 49. H. Schlattl, A. Weiss, and G. Raelt. Helioseismological constraint on solar axion emission. Astroparticle Physics , 10(4):353 { 359, 1999.#### Page 16 of 18 Eur. Phys. J. Plus ##################### 50. Javier Redondo. Solar axion
ux from the axion-electron coupling. Journal of Cosmology and Astroparticle Physics , 2013(12):008{008, dec 2013. 51. N. Viaux, M. Catelan, P. B. Stetson, et al. Neutrino and axion bounds from the globular cluster m5 (ngc 5904). Phys. Rev. Lett. , 111:231301, Dec 2013. 52. Georg G. Raelt. Axion constraints from white dwarf cooling times. Physics Letters B , 166(4):402 { 406, 1986. 53. Alejandro H. C orsico, Alejandra D. Romero, Leandro G. Althaus, et al. An asteroseismic constraint on the mass of the axion from the period drift of the pulsating DA white dwarf star l19-2. Journal of Cosmology and Astroparticle Physics , 2016(07):036{036, jul 2016. 54. J. Engel, D. Seckel, and A. C. Hayes. Emission and detectability of hadronic axions from sn 1987a. Phys. Rev. Lett. , 65:960{963, Aug 1990. 55. Lev B. Leinson. Axion mass limit from observations of the neutron star in cassiopeia a. Journal of Cosmology and Astroparticle Physics , 2014(08):031{031, aug 2014. 56. Jochen Keller and Armen Sedrakian. Axions from cooling compact stars: Pair-breaking processes. Nuclear Physics A , 897:62 { 69, 2013. 57. Armen Sedrakian. Axion cooling of neutron stars. Phys. Rev. D , 93:065044, Mar 2016. 58. Patrick Fox, Aaron Pierce, and Scott Thomas. Probing a qcd string axion with precision cosmological measurements. arXiv preprint hep-th/0409059 , 2004. 59. Kyu Jung Bae, Ji-Haeng Huh, and Jihn E Kim. Updating the axion cold dark matter energy density. Journal of Cosmology and Astroparticle Physics , 2008(09):005, sep 2008. 60. Olivier Wantz and E. P. S. Shellard. Axion cosmology revisited. Phys. Rev. D , 82:123508, Dec 2010. 61. Max Tegmark, Daniel J. Eisenstein, Michael A. Strauss, et al. Cosmological constraints from the sdss luminous red galaxies. Phys. Rev. D , 74:123507, Dec 2006. 62. Mar a Beltr an, Juan Garc a-Bellido, and Julien Lesgourgues. Isocurvature bounds on axions revisited. Phys. Rev. D , 75:103507, May 2007. 63. Mark P. Hertzberg, Max Tegmark, and Frank Wilczek. Axion cosmology and the energy scale of in
ation. Phys. Rev. D , 78:083507, Oct 2008. 64. J Hamann, S Hannestad, G.G Raelt, and Y.Y.Y Wong. Isocurvature forecast in the anthropic axion window. Journal of Cosmology and Astroparticle Physics , 2009(06):022{022, jun 2009. 65. Planck Collaboration, Ade, P. A. R., Aghanim, N., et al. Planck 2013 results. xvi. cosmological parameters. A&A , 571:A16, 2014. 66. Planck Collaboration, Ade, P. A. R., Aghanim, N., et al. Planck 2015 results - xiii. cosmological parameters. A&A , 594:A13, 2016. 67. Paola Arias, Davide Cadamuro, Mark Goodsell, et al. WISPy cold dark matter. Journal of Cosmology and Astroparticle Physics , 2012(06):013{013, jun 2012. 68. Igor G. Irastorza and Javier Redondo. New experimental approaches in the search for axion-like particles. Progress in Particle and Nuclear Physics , 102:89 { 159, 2018. 69. P. Sikivie. Experimental tests of the "invisible" axion. Phys. Rev. Lett. , 51:1415{1417, Oct 1983. 70. S. J. Asztalos, R. F. Bradley, L. Duy, et al. Improved rf cavity search for halo axions. Phys. Rev. D , 69:011101, Jan 2004. 71. L. D. Duy, P. Sikivie, D. B. Tanner, et al. High resolution search for dark-matter axions. Phys. Rev. D , 74:012006, Jul 2006. 72. S. J. Asztalos, G. Carosi, C. Hagmann, et al. Squid-based microwave cavity search for dark-matter axions. Phys. Rev. Lett., 104:041301, Jan 2010. 73. N. Du, N. Force, R. Khatiwada, et al. Search for invisible axion dark matter with the axion dark matter experiment. Phys. Rev. Lett. , 120:151301, Apr 2018. 74. C. Bartram, T. Braine, R. Cervantes, et al. Axion dark matter experiment: Run 1b analysis details. Phys. Rev. D , 103:032002, Feb 2021. 75. B. M. Brubaker, L. Zhong, Y. V. Gurevich, et al. First results from a microwave cavity axion search at 24 eV. Phys. Rev. Lett. , 118:061302, Feb 2017. 76. K. M. Backes, D. A. Palken, S. Al Kenany, et al. A quantum enhanced search for dark matter axions. Nature , 590(7845):238{ 242, February 2021. 77. S. DePanlis, A. C. Melissinos, B. E. Moskowitz, et al. Limits on the abundance and coupling of cosmic axions at 4:5ma5:0ev.Phys. Rev. Lett. , 59:839{842, Aug 1987. 78. Ben T. McAllister, Graeme Flower, Eugene N. Ivanov, et al. The organ experiment: An axion haloscope above 15 ghz. Physics of the Dark Universe , 18:67 { 72, 2017. 79. Petrakou, Eleni and for CAPP/IBS. Haloscope searches for dark matter axions at the center for axion and precision physics research. EPJ Web Conf. , 164:01012, 2017. 80. Allen Caldwell, Gia Dvali, B ela Majorovits, et al. Dielectric haloscopes: a new way to detect axion dark matter. Phys. Rev. Lett. , 118(9):091801, 2017. 81. Dmitry Budker, Peter W. Graham, Micah Ledbetter, Surjeet Rajendran, and Alexander O. Sushkov. Proposal for a cosmic axion spin precession experiment (casper). Phys. Rev. X , 4:021030, May 2014. 82. Antoine Garcon, Deniz Aybas, John W Blanchard, et al. The cosmic axion spin precession experiment (CASPEr): a dark-matter search with nuclear magnetic resonance. Quantum Science and Technology , 3(1):014008, dec 2017.Eur. Phys. J. Plus ##################### #### Page 17 of 18 83. G Ruoso, A Lombardi, A Ortolan, et al. The quax proposal: a search of galactic axion with magnetic materials. Journal of Physics: Conference Series , 718(4):042051, 2016. 84. R. Barbieri, C. Braggio, G. Carugno, et al. Searching for galactic axions through magnetized media: The quax proposal. Physics of the Dark Universe , 15:135 { 141, 2017. 85. Jonathan L. Ouellet, Chiara P. Salemi, Joshua W. Foster, et al. First results from abracadabra-10 cm: A search for sub- eV axion dark matter. Phys. Rev. Lett. , 122:121802, Mar 2019. 86. F. Caspers and Y. Semertzidis. Ferrimagnetic resonance, magnetostatic waves and open resonators for axion detection. In Cosmic axions. Proceedings, Workshop, Upton, USA, April 13-14, 1989 , pages 0173{183, 1989. 87. R. Barbieri, M. Cerdonio, G. Fiorentini, and S. Vitale. Axion to magnon conversion. a scheme for the detection of galactic axions. Physics Letters B , 226(3):357 { 360, 1989. 88. AI Kakhidze and IV Kolokolov. Antiferromagnetic axion detector. Zh. Eksp. Teor. Fiz , 99:1077{1081, 1991. 89. PV Vorobyov, AN Kirpotin, ME Rovkin, and AP Boldyrev. Ferromagnetic detectors of axions in rf (s-x) band. arXiv preprint hep-ph/9506371 , 1995. 90. Lawrence Krauss, John Moody, Frank Wilczek, and Donald E. Morris. Calculations for cosmic axion detection. Phys. Rev. Lett., 55:1797{1800, Oct 1985. 91. L.D. Landau, E.M. Lif sic, E.M. Lifshitz, et al. Statistical Physics: Theory of the Condensed State . Course of theoretical physics. Elsevier Science, 1980. 92. Planck Collaboration, Ade, P. A. R., Aghanim, N., et al. Planck 2015 results - xiii. cosmological parameters. A&A , 594:A13, 2016. 93. Michael S. Turner. Periodic signatures for the detection of cosmic axions. Phys. Rev. D , 42:3572{3575, Nov 1990. 94. Michael S. Turner and Frank Wilczek. Positron line radiation as a signature of particle dark matter in the halo. Phys. Rev. D , 42:1001{1007, Aug 1990. 95. Erik W. Lentz, Thomas R. Quinn, Leslie J. Rosenberg, and Michael J. Tremmel. A new signal model for axion cavity searches from n -body simulations. The Astrophysical Journal , 845(2):121, 2017. 96. N. Bloembergen and R. V. Pound. Radiation damping in magnetic resonance experiments. Phys. Rev. , 95:8{12, Jul 1954. 97. Charles Kittel. Interpretation of anomalous larmor frequencies in ferromagnetic resonance experiment. Phys. Rev. , 71:270{ 271, Feb 1947. 98. M. Sparks and C. Kittel. Ferromagnetic relaxation mechanism for mzin yttrium iron garnet. Phys. Rev. Lett. , 4:232{234, Mar 1960. 99. Michael Tavis and Frederick W. Cummings. Exact solution for an n-molecule|radiation-eld hamiltonian. Phys. Rev. , 170:379{384, Jun 1968. 100. Xufeng Zhang, Chang-Ling Zou, Liang Jiang, and Hong X. Tang. Strongly coupled magnons and cavity microwave photons. Phys. Rev. Lett. , 113:156401, Oct 2014. 101. Yutaka Tabuchi, Seiichiro Ishino, Toyofumi Ishikawa, et al. Hybridizing ferromagnetic magnons and microwave photons in the quantum limit. Phys. Rev. Lett. , 113:083603, Aug 2014. 102. N. Crescini, C. Braggio, G. Carugno, A. Ortolan, and G. Ruoso. Coherent coupling between multiple ferrimagnetic spheres and a microwave cavity at millikelvin temperatures. Phys. Rev. B , 104:064426, Aug 2021. 103. Crescini, N., Alesini, D., Braggio, C., et al. Operation of a ferromagnetic axion haloscope at ma= 58eV.Eur. Phys. J. C, 78(9):703, 2018. 104. N. Crescini. Towards the development of the ferromagnetic axion haloscope, PH. D. thesis , padova university, 2019. 105. B. Yurke, L. R. Corruccini, P. G. Kaminsky, et al. Observation of parametric amplication and deamplication in a josephson parametric amplier. Phys. Rev. A , 39:2519{2533, Mar 1989. 106. Manuel Castellanos Beltran and K W. Lehnert. Widely tunable parametric amplier based on a superconducting quantum interference device array resonator. Applied Physics Letters , 91:083509{083509, 08 2007. 107. M. Sandberg, C. M. Wilson, F. Persson, et al. Tuning the eld in a microwave resonator faster than the photon lifetime. Applied Physics Letters , 92(20):203501, 2008. 108. Erik A. Thol en, Adem Erg ul, Evelyn M. Doherty, et al. Nonlinearities and parametric amplication in superconducting coplanar waveguide resonators. Applied Physics Letters , 90(25):253509, 2007. 109. T. Yamamoto, K. Inomata, M. Watanabe, et al. Flux-driven josephson parametric amplier. Applied Physics Letters , 93(4):042510, 2008. 110. B. Abdo, O. Suchoi, E. Segev, et al. Intermodulation and parametric amplication in a superconducting stripline resonator integrated with a dc-SQUID. EPL (Europhysics Letters) , 85(6):68001, mar 2009. 111. J. Y. Mutus, T. C. White, R. Barends, et al. Strong environmental coupling in a josephson parametric amplier. Applied Physics Letters , 104(26):263513, 2014. 112. T. Mimura. The early history of the high electron mobility transistor (hemt). IEEE Transactions on Microwave Theory and Techniques , 50(3):780{782, March 2002. 113. Ananda Roy and Michel Devoret. Introduction to parametric amplication of quantum signals with josephson circuits. Comptes Rendus Physique , 17(7):740 { 755, 2016. Quantum microwaves / Micro-ondes quantiques. 114. N. Bergeal, R. Vijay, V. E. Manucharyan, et al. Analog information processing at the quantum limit with a josephson ring modulator. Nature Physics , 6:296 EP {, Feb 2010. Article. 115. N. Crescini, C. Braggio, G. Carugno, et al. Magnon-driven dynamics of a hybrid system excited with ultrafast optical pulses. Communications Physics , 3(1):164, Sep 2020. 116. Ciaran O'HARE. cajohare/axionlimits: Axionlimits, July 2020.#### Page 18 of 18 Eur. Phys. J. Plus ##################### 117. Rapha el Lescanne, Samuel Del eglise, Emanuele Albertinale, et al. Irreversible qubit-photon coupling for the detection of itinerant microwave photons. Phys. Rev. X , 10:021038, May 2020. 118. Emanuele Albertinale, L eo Balembois, Eric Billaud, et al. Detecting spins by their
uorescence with a microwave photon counter. Nature , 600(7889):434{438, Dec 2021. | 2022-01-11 | A ferromagnetic haloscope is a rf spin-magnetometer used for searching Dark
Matter in the form of axions. A magnetic material is monitored searching for
anomalous magnetization oscillations which can be induced by dark matter
axions. To properly devise such instrument one first needs to understand the
features of the searched-for signal, namely the effective rf field of dark
matter axions $B_a$ acting on electronic spins. Once the properties of $B_a$
are defined, the design and test of the apparatus may start. The optimal sample
is a narrow linewidth and high spin-density material such as Yttrium Iron
Garnet (YIG), coupled to a microwave cavity with almost matched linewidth to
collect the signal. The power in the resonator is collected with an antenna and
amplified with a Josephson Parametric amplifier, a quantum-limited device
which, however, adds most of the setup noise. The signal is further amplified
with low noise HEMT and down-converted for storage with an heterodyne receiver.
This work describes how to build such apparatus, with all the experimental
details, the main issues one might face, and some solutions. | Building instructions for a ferromagnetic axion haloscope | 2201.04081v1 |
Evidence for exchange Dirac gap in magneto-transport of topological insulator-magnetic insulator heterostructures S. R. Yang1#, Y . T. Fanchiang2#, C. C. Chen1, C. C. Tseng1, Y . C. Liu1, M. X. Guo1, M. Hong2, S. F. Lee3*, and J. Kwo1* 1Department of Physics, National Tsing Hua University, Hsinchu 30013, Taiwan 2Department of Physics, National Taiwan University, Taipei 10617, Taiwan 3Institute of Physics, Academia Sinica, Taipei 11529, Taiwan # Authors who have equal contributions to this work * Corresponding authors Abstract: Transport signatures of exchange gap opening because of magnetic proximity effect (MPE) are reported fo r bilayer structures of Bi 2Se3 thin films on yttrium iron garnet (YIG) and thulium iron garnet (TmI G) of perpendicular magnetic anisotropy (PMA). Pronounced negative magnetoresistanc e (MR) was detected, and attributed to an emergent weak localization (WL) effect superimposing on a weak antilocalization (WAL). Thickness-dependent study shows that the WL originates from the time-reversal-symmetry breaking of topological surface states by interfacial exchange coupling. The weight of WL declined when the interfacial magnetization was aligned toward the in-plane direction, which is understood as the effect of tuning the exchange gap size by varying the perpendicular magnetization component. Importantly, magnetotransport study revealed anomalous Hall effect (AHE) of square loops and anisotropic magnetoresistance (AMR) char acteristic, typifyi ng a ferromagnetic conductor in Bi 2Se3/TmIG, and the presence of an interfacial ferromagnetism driven by MPE. Coexistence of MPE-induced ferromagnetism and the finite exchange gap provides an opportunity of realizing zero magnetic-field dissipation-less transport in topological insulator/ferromagne tic insulator heterostructures. Breaking time-reversal symmetry (TRS) in topological insulators (TIs) leads to several exotic phenomenon such as quantum anomalous Hall effect (QAHE), topological magnetoelectric effect, and magnetic monopole [1,2] . A prerequisite of these novel quantum state is an energy gap opened at the Dirac surface state induced by exchange interaction with magnetic elements [3]. Magnetic doping is a prevalent way of introducing ferromagnetism in TIs [4-7] . Study of TRS breaking in magnetically doped TIs was ignited by the direct observation of an exchange gap opening of topological surface states (TSS s) via angle-resolved photoemission spectroscopy (ARPES) [5], and culminated with the real ization of QAHE in Cr-doped (Bi,Sb) 2Te3 [8]. Although magnetic doping is proven to be effective in breaking TRS, the observation temperature of QAHE reported so far was less than 2 K [8-12] , order-of-magnitude lower than the ferromagnetic Curie temperature ( ܶC). It is suggested that the disorder created by dopants, as well as the small exchange gap size induced by low doping concentration, poses a limit of raising the QAHE temperature [12,13] . Recently, magnetic proximity effect (M PE) of TI/ferromagnetic insulator (FI) heterostructures was demonstrated as another promising route of breaking TRS [14-17] . Besides the benefit of much higher ܶC, the induced interfacial magnetization is uniform, free of crystal defects. A room -temperature ferromagnetism by MPE is directly observed in epitaxial EuS/Bi 2Se3 by polarized neutron reflectometry [16] . Moreover, robust anomalous Hall (AH) resistances up to 400 K has been detected in (Bi,Sb) 2Te3 films on TmIG with perpendicular magnetic anisotropy (PMA) [17] . Despite the clear observations of ferromagnetism and presumably pronounced TRS breaking, the experimental indications of exchange gap opening following MPE in these cases are still vague. Unlike magnet ically doped-TIs where the gapped surface can be exposed to the probe of ARPES technique [5], the gapped surface state caused by MPE is buried at the interface, making it difficult to investigate using typical ARPES. Attempts to detect MPE-induced exchange gap by transport measurements have been made by various groups [18-20] . One signature of exchange gap opening is an emerging weak localization (WL) taking the form of negative magnetoresistance (MR) accompanied by a suppressed weak antilocalization (WAL) [21]. However, negative MR in TI/FI was hitherto observed in samples comparable or beyond the Ioffe-Regel limit (sheet resistance ܴୱ݁/݄ ଶ) [18-20] . Above the limit, the Anderson (strong) localization due to disorder c ould also give rise to a negative MR [22,23] , which cannot be easily distinguished from the one due to an exchange gap. Therefore, a definite transport signature of MPE-induced exchange gap remains elusive. Despite the observation of negative MR and suppressed WAL [24-26], robust MPE-induced ferromagnetism in TIs, whic h should be best manifested as large remnant magnetization ܯ୰ pointing out-of-plane and hysteric AHE, has not been observed in those systems. Conversely, although clear ܯ୰ was detected in (Bi,Sb) 2Te3/TmIG, the negative MR and transport signature of gap opening were not simultaneously presented [17]. To achieve QAHE in TI/FI, both ingredients, robust ferromagnetism of TSS and exchange gap opening, must be fulfilled concurrently, which is a condition yet to be demonstrated for TI/FI. In pursuing high-temperature QAHE through MPE, it is important to identify the transport signature of the exchange gap associated with MPE-induced ferromagnetism, before one moves on to the ultimate goal of QAHE. Moreover, although AHE is often taken as an evidence of interfacial ferromagnetism, the possible spin current effect in TI/FI, such as spin Hall MR [27] , can also lead to similar AH resistances. Additional transport study is demanded to elucidate possible modulations of magneto-transport by spin current effects. In this work we report a pronounced WL that competes with the WAL in Bi2Se3/YIG and Bi 2Se3/TmIG bilayers. The ferrimagnetic garnet films were chosen because of their high ܶC above 500 K [28] , good thermal stability in conjunction with TIs, and their technological importance [29] . The emergent WL effect is strong enough to manifest a negative magnetoresistance (MR), showing systematic changes with the perpendicular magnetization controlled by tilted external fields. This strongly suggests that the observed WL arose from a finite MPE-induced exchange gap, whose size could be further tuned by the external field directions. Most importantly, the breaking of TRS by MPE is corroborated by the observation of AHE up to 180 K, together with clear anisotropic magnetoresistance (AMR), confirming ferromagnetic TSS. Our study thus presents a coherent picture of the long sought MPE-induced ferromagnetism and exchange gap in electron transport. Our YIG and TmIG thin films are fabricated using off-axis sputtering [30,31] . YIG is a ferrimagnetic insulator with ultralow magnetic damping ideal for spin wave and spin current transport study [29] , while TmIG films, when under a tensile strain, exhibit robust and tunable perpendicular magnetic anisotropy (PMA) [31-33] which is essential for exchange gap of TSSs. Bi 2Se3 thin films of 6 – 40 quintuple layer (QL) were deposited on the garnet substrates by molecular beam epitaxy (MBE). High quality Bi 2Se3 thin films and sharp Bi 2Se3/garnet interface were obtained by the invention of a novel growth procedure, a key factor for conveying strong exchange interaction of the localized magnetic moments of garnets layers and TSSs [34] . For transport measurements, Bi 2Se3/garnet bilayer samples were made into Hall bars (650 µm ൈ 50 µm ) by standard photolithography. Four points measurement was carried out in a 9 T Quantum Design physical property measurement system with a 10 µA DC current. Figure 1(a) shows the temperature dependence of sheet resistance ( ܴୱ) of Bi2Se3/YIG bilayers with Bi 2Se3 thickness of 6, 10, 16, and 40 QL. All samples exhibit metallic behavior of decreasing ܴୱ when the samples were cooled down from room temperature. The sheet carrier density of these samples is in the range of ሺ1.5 െ 3 ሻൈ1 0ଵଷcmିଶ, indicating that the bulk carriers of Bi 2Se3 participate in the electron transport. Due to an increasing surface scattering, the ܴୱ tends to be larger in thinner Bi 2Se3 [35]. Note that the maximum ܴୱ of these sample were well below h/݁ଶሺൎ 25.8 kΩ ሻ satisfying the condition of transport regime, thus quantum interference effects of 2D electron systems in TI thin films, such as WAL and possible WL, can be described by well-developed theories [22] . Figure 1(b) displays the MR data taken at 2 K under a perpendicular applied field for the four Bi 2Se3/YIG bilayers and, for comparison, a 9 QL Bi 2Se3 grown on Al 2O3. For Bi 2Se3/Al 2O3, a sharp cusp feature at low fields, characteristic of WAL effect, was observed, and the MR stayed positive up to 9 T. WAL in thin Bi 2Se3 is generally attributed to destructive interference because of difference of Berry’s phase 2ߨ accumulated by the Dirac fermion travelling in two time-reversed paths [1]. In contrast, notable negative MR were observed in 6 and 10 QL Bi 2Se3/YIG. Specifically, at low fields ( ൏1 T ), a weakened positive MR or suppressed WAL was observed in all the four Bi 2Se3/YIG bilayers. At intermediate fields ( 1െ4 T ), MR becomes negative for the thinner two samples. While the MR of 16 and 40 QL Bi 2Se3/YIG remained positive, that from 16 QL Bi 2Se3/YIG did show the much weaker positive MR compared to that of Bi2Se3/sapphire. When the external field exceeded 4 T, all the samples exhibited positive MR that results from the Lorentz force on moving electrons. The suppressed WAL effect of Bi 2Se3/YIG suggests that an additional transport mechanism shows up that contributes to a negative MR and competes with the WAL. Since the film quality of Bi 2Se3 grown on YIG is comparable to that on sapphire, the distinct MR behavior of Bi 2Se3/YIG is most likely originated from the interaction between the bottom surface of Bi 2Se3 and the YIG layer. As the YIG is ferrimagnetic at 2 K [36] , a sizable interfacial exchange coupling should exist in Bi 2Se3/YIG [15,25,30] , which is otherwise negligible for the non-magnetic sapphire substrate. Therefore, the stark contrast in the MR behaviors suggest that the suppressed WAL and negative MR are the indication on transport properties of TRS-breaking in TIs. Other possible mechanisms for the negative MR of TIs are defect-induced hopping transport [37] , hybridization gap of TSSs [21], and bulk subbands in thin TIs [38] . Each scenario can be excluded straightfo rwardly, as discussed in details in Supplemental Materials [39] . When the TSS is subjected to an exchange field, the Dirac fermion becomes massive, as expressed by an effective Hamiltonian ܪൌെ i ݒ ிሺෝൈܢොሻڄસෝ ଶۻڄ [21], where ݒி is the Fermi velocity, ෝ is Pauli matrix, ܬ is the exchange coupling constant and ۻ is the magnetization unit factor. The resulting energy dispersion is ܧൌേ ඨ൬ݒ ி݇௫ܬ 2ܯ௬൰ଶ ൬ ݒ ி݇௬െܬ 2ܯ௫൰ଶ ൬ܬ 2ܯ௭൰ଶ (1) with an exchange gap size of ܯܬ . It has been shown that electrons travelling a closed path would acquire a Berry’s phase as ߨሺ1െܯܬ ௭/2ܧ Fሻ, where E F is Fermi energy measured from the Dirac point [21] . The modulated Berry’s phase weakens the associated destructive interference, or even induces a crossover from WAL to WL. In the case of Bi 2Se3/YIG, interfacial exchange coupling can induce a finite gap through MPE, which further leads to competing WL. To quantitatively describe the negative MR, we calculated the longitudinal conductivity ܩ୶୶ from the tensor relation ܴ୶୶/ ሺܴ ୶୶ଶܴ ୷୶ଶሻ. The competition between WAL and WL can be described using the modified Hikami-Larkin-Nagaoka (HLN) equation [21,40] , Δܩ ൌ ܩ ୶୶ሺܤሻെܩ ୶୶ሺ0ሻൌ ߙ ୧ቆ݁ଶ ݄ߨቇቈ ߰ቆ 4݈݁୧ଶܤ1 2ቇെl nቆ 4݈݁୧ଶܤቇଵ ୧ୀܤߚ ଶ (2) , where ߰ is the digamma function, ߙ represents the weights of WL ( iൌ0 ) or WAL ( iൌ1 ), ݈୧ is the corresponding effective phase coherence length. The ܤߚଶ term primarily results from the Lorentz deflection of carriers [39]. To clearly reveal the presence of the WL component, the MC curves subtracted by the ܤߚଶ background are plotted in Fig. 1(c) . Positive MC was observed for 6, 10, and 16 QL Bi2Se3/YIG. Figure 1(d) shows the Bi 2Se3 thickness dependence of ߙ and ߙଵ. For 6 and 10 QL Bi 2Se3/YIG bilayers, large ߙ values of 0.7 and 2.7 were extracted, respectively. A crossing of the magnitudes of ߙ and ߙଵ was observed in thicker Bi2Se3 as ߙ decreased substantially for 16 QL Bi 2Se3 and became vanishing for 40 QL Bi 2Se3. Meanwhile, the ߙଵ value in general remains lower than -0.5, showing a slight decrease toward thinner Bi 2Se3. The smaller ߙଵ value suggests suppressed WAL channels in the bulk-surface-coupled Bi 2Se3 [23,35] . The extracted ߙ’s and ߙଵ’s thus reveal the competitive behavior between WL and WAL, whose interfacial origin are indicated by the stronger WL and weaker WAL in thinner Bi 2Se3. It is noteworthy that Eq. 2 is derived from a model considering an effective Hamiltonian of a single Dirac surface state. In reality, the Bi 2Se3 films have two conducting surfaces interacting through the bulk carriers in transport [23]. The complexity may give rise to the unexpectedly large ߙ in thin Bi 2Se3/YIG. Nevertheless, Eq. 2 and Fig. 1(d) do capture the concept of emergent WL from TRS breaking at the interfaces [13]. TmIG films with PMA are more desirable for exchange gap opening at zero applied field due to their robust ܯ ୰. To further verify the relation between WL and exchange gap, we tilted the applied field from the z direction. Based on Eq. (1), tilted field should effectively vary the size of ܯ and thus tune the exchange gap size ܯܬ . In Bi 2Se3/TmIG, ܯ stands for the z component of (i) magnetization of TmIG near the interface or (ii) MPE-driven magnetization on the TI side ଵ. Figure 2(c) -(f) shows the MR results of Bi 2Se3/Al 2O3 and Bi 2Se3/YIG under applied fields of different angles ߠ୷ൌ 90°, 60°, and 30° . For Bi 2Se3/Al 2O3, although the MR curves change with ߠ୷ in Fig. 2(c) , when plotted as a function of perpendicular field ܤ൫ൌ ܤsinߠ ୷൯ in Fig. 2(d) , the curves collapse into one. The observation implies that MR here is sensitive to ܤ only and can be well explained by an ordinary WAL effect in the Bi 2Se3/Al 2O3 where ܬൎ0 . In sharp contrast, unusual MR behaviors were seen for Bi 2Se3/TmIG. Firstly, Bi2Se3/TmIG also shows clear negative MR in the intermediate fields as Bi 2Se3/YIG (Fig. 2(e) ). Note that it is difficult to directly measure the ܤ-dependent magnetization of TmIG films because of large low- ܶ paramagnetic background from GGG [30,41] . The total anisotropy field of TmIG is ~0.07 T at room temperature [31], which should not increase dramatically at low ܶ as it is compromised by increasing saturation magnetization of TmIG. At fields ܤ 1.2 ܶ where negative MR starts to appear, we expect that the magnetization of TmIG has been saturated by ܤ .Secondly, as shown in Fig. 2(f) , the ܤ dependence of MR systematically changes with ߠ୷. At low ܤ, the MR curves for different ߠ୷’s coincide well because WAL is governed primarily by ܤ. The MR curves split when ܤ1 . 2 ܶ and possess weaker negative MR for smaller ߠ୷. The correspondence between negative MR and ߠ୷ can be best explained by a tunable exchange gap. As illustrated in Fig. 2(a) , when the applied field was sufficient to align ۻ at interface, the exchange gap size is tuned by re-orienting ۻ .Since the exchange gap ܯܬ ܯן ןs i n ߠ ୷, it follows that the negative MR of Bi 2Se3/TmIG, or the weight of WL, is in positive correlation with the exchange gap size. Because the exchange gap size determines the deviations of the Berry’s phase from ߨ ,the effect of gap tuning is manifested as the variable negative MR with ߠ୷. In principle, an exchange gap can be induced locally by individual magnetic impurities [42] . Although magnetic impurities deposited on a TI surface can acquire ferromagnetism via Ruderman–Kittel–Kasuya–Yosidas (RKKY) type interaction mediated by Dirac fermions [42], this may not be the leading mechanism for ferromagnetism in a TI/FI system, where interlayer exchange coupling plays the major role. To realize QAHE, ferromagnetism needs be establishe d for an exchange gap opened macroscopically without an applied field [13]. In the following, we show that the Bi 2Se3/TmIG does meet the criterion. Figure 3(a) shows a representative curve of AHE at 100 K. A square hysteresis loop of Hall resistance was observed after the contribution from the ordinary Hall effect was subtracted, based on the empirical formula ܴ୷୶ൌܴ Hሺܤሻܴ AHሺܤሻ, where ܴH is the ordinary Hall resistance, and ܴAH represents the AH resistance. Since TmIG layer is insulating, the AH resistance dominantly comes from the TI layer. The hysteresis loop resembles that of TmIG magnetization [31] . As displayed in Fig. 3(b) , the switching field of the hysteresis loops ܤୡ increases rapidly as temperature was lowered. The enhanced ܪୡ is likely associated with the larger strain in TmIG at low temperatures [33] . Moreover, the effect of the stray field on ܴAH was negligible as we did not observe an AH resistance in Bi 2Se3/Al 2O3/TmIG, in which the interfacial exchange coupling is greatly suppressed by nonmagnetic Al 2O3 (see Supplemental materials, Fig. S2(a), (b) [39]). The above observations indicate that a spontaneous magnetization, ଵ, has developed at Bi 2Se3/TmIG interface because of MPE, with the magnetized Bi 2Se3 bottom surface effectively acting as a magnetic conductor. The AH resistance can be further transformed to AH conductance using ߪ AHߩ؆ AH/ߩ୶୶. Figure 3(b) shows the temperature dependence of AH conductance amplitude ߪAH. ߪAH decays moderately with increasing ܶ below 50 K, and persists up to 180 K. In the weakly disorder limit, the exchange gap size can be estimated from total ߪAH using ߪAHൎమ ቀ ாFቁଷ , taking into account of extrinsic AH conductivity [43]. With ܧF0 . 1 5 eV for bulk-conductive Bi 2Se3 [23] , a lower bound of exchange gap size ~7.7 meV at 10 K is determined. The gap size is in good agreement with 9 meV obtained from density-functional theory calculations for EuS/Bi 2Se3 [44]. The gap is one order of magnitude smaller than that observed in magnetically doped TIs of ~100 meV [13,45] . However, the very large surface state gap in doped TIs is likely nonmagnetic and caused by resonant states induced by impurities near the Dirac point [45]. To clarify the role of MPE-induced magn etization, the MR measurements were conducted at 100 K to preclude low-fiel d quantum interference effects of TSS. Figure 4(a) shows the field-dependent resistance of our samples with longitudinal ( צܴ ,) transverse ( ܴT), and perpendicular ( ܴୄ) fields. Distinct turnings of צܴ and ܴT were observed at ~ 0.5 T, which were attributed to the field needed to fully saturate the perpendicular magnetization in Bi 2Se3/TmIG toward in-plane direction. Below the saturation field, צܴ and ܴT progressively increase with the increasing in-plane field, and צܴܴ T in particular. Meanwhile, ܴୄ is parabolic because of ܤ-induced Lorentz deflection (see also Figs. 4(c) and (d)). We recognize the MR behaviors in Fig. 4(a) as features of AMR caused by MPE. Regardless of the domain configuration, an in-plane field promotes (diminishes) the average of in-plane (perpendicular) MPE-induced magnetization ݉ۃଵ୶,ଵ୷ ۄ( ݉ۃଵۄ )until the saturation field was reached. The subsequent increase of צܴ and ܴT implies that ݉ۃଵ୶ۄ and ݉ۃଵ୷ۄ contribute larger resistances than ݉ۃଵۄ does, i.e. ܴ,צܴTܴ ୄ. Indeed, in the regime |ܤ|൏0 . 7 ܶ where ܴୄ was not overwhelmingly enhanced by ܤ ,the AMR relation of magnetic thin films, צܴܴ Tܴ ୄ [46], has already appeared. Furthermore, Figure 4(a) also rules out SMR to be the dominant source of the AH resistances because SMR features צܴൎܴ ୄܴ T [27] . The AMR amplitude צܴെܴ T continues to build up as ܤ went larger, further justified by the ሺcosሻଶ dependence on ߶୶୷ shown in Fig. 4(b) . Corresponding planar Hall effect characteristic of fe rromagnetic conductors was also detected (Supplemental materials, Fig. S4(a), (b) [39] ). Alternatively, צܴെܴ T can be extracted from the resistance difference of ߠ୷ and ߙ୶ scans displayed in Fig. 4(c) , despite the large contributions of Lorentz deflection in ܴୄ. As a comparison, nearly no ߶୶୷ dependence of resistances was detected in Bi 2Se3/Al 2O3 (Fig. 4(d) ). We notice that the field-enhanced צܴെܴ T was also reported in Pt/YIG, where the authors stated that the large-field צܴെܴ T mostly came from the "hybrid MR" of MPE [47] , which exhibits the same relation צܴൎܴ ୄܴ T as that of SMR. Notwithstanding the similarity between the two systems, we point out that Bi 2Se3 cannot be simply treated as a heavy metal with strong spin-orbit-coupling even at an elevated temperature. The MPE in Bi 2Se3/TmIG involves hybridization between Fe d-orbitals and the paramagnetic TSS arising from the Bi and Se p-orbitals, as opposed to Pt/YIG or other Pt/FM structures where d-d interaction is responsible for MPE. The distinct MR behaviors, AMR in Bi 2Se3/TmIG contrary to SMR/hybrid MR in Pt/YIG, may be an important clue to the microsc opic transport property of MPE in TI/FI. To summarize, a competing WL along with a suppressed WAL has been observed in Bi 2Se3/YIG and Bi 2Se3/TmIG. Bi 2Se3 thickness dependence study suggests that the WL comes from an exchange gap of TSSs opened at interface. In addition, the weight of WL evolves with tilted MPE-induced magnetization. Such angular dependence consolidates the exchange gap as the origin of WL, and the variable WL strength signifies tunability of the gap. Moreover, the well-defined square ܴ AH loops in Bi 2Se3/TmIG unambiguously point to a long-range ferromagnetic order at the interface, and thus ensure a macroscopic and uniform exchange gap at zero field. The MPE-induced ferromagnetism in Bi 2Se3/TmIG is doubly evidenced by typical AMR characteristics, alleviating the concern of spin current effects on the magneto-transport. The simultaneous presence of the MPE-induced long-range ferromagnetic order at the surface and the exchange gap is thus realized in the prototypical TI Bi 2Se3, pending the Fermi-level tuning to deplete the bulk conduction. Lastly, by circumventing the inhomogeneous magnetic doping and the impurity-induced resonance state problems that have been encountered in magnetically doped Bi 2Se3 [13,45] , our study demonstrates that MPE could be a more viable way of introducing ferromagnetism in various TI systems. Acknowledgment Technical support from Nano group Public Laboratory, Institute of Physics, at Academia Sinica in Taiwan, is acknowledged. The work is supported by MoST 105-2112-M-007-014-MY3, 106-2112-M- 002-010-, 107-2622-8-002-018, and 105-2112-M-001-031-MY3 of the Ministry of Science and Technology in Taiwan. References [1] M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010). [2] X. L. Qi and S. C. Zhang, Rev. Mod. Phys. 83, 1057 (2011). [3] X. L. Qi, T. L. Hughes, and S. C. Zhang, Phys. Rev. B 78, 195424 (2008). [4] R. Yu, W. Zhang, H. J. Zhang, S. C. Zhang, X. Dai, and Z. Fang, Science 61, 329 (2010). [5] Y . L. Chen, J. H. Chu, J. G. Analytis, Z. K. Liu, K. Igarashi, H. H. Kuo, X. L. Qi, S. K. Mo, R. G. Moore, D. H. Lu , M. Hashimoto, T. Sasagawa, S. C. Zhang, I. R. Fisher, Z. Hussain, and Z. X. Shen, Science 329, 659 (2010). [6] Y . S. Hor, P . Roushan, H. Beidenkopf, J. Seo, D. Qu, J. G. Checkelsky, L. A. Wray, D. Hsieh, Y . Xia, S. Y . Xu, D. Qian, M. Z. Hasan, N. P. Ong, A. Yazdani, and R. J. Cava, Phys. Rev. B 81 (2010). [7] C. Z. Chang, J. Zhang, M. Liu, Z. Zhang, X. Feng, K. Li, L. L. Wang, X. Chen, X. Dai, Z. Fang, X. L. Qi, S. C. Zhang, Y . Wang, K. He, X. C. Ma, and Q. K. Xue, Adv. Mater. 25, 1065 (2013). [8] C. Z. Chang, J. Zhang, X. Feng, J. Shen, Z. Zhang, M. Guo, K. Li, Y . Ou, P . Wei, L. L. Wang, Z. Q. Ji, Y . Feng, S. Ji, X. Chen, J. Jia, X. Dai, Z. Fang, S. C. Zhang, K. He, Y . Wang, L. Lu, X. C. Ma, and Q. K. Xue, Science 340, 167 (2013). [9] X. Kou, S. T. Guo, Y . Fan, L. Pan, M. Lang, Y . Jiang, Q. Shao, T. Nie, K. Murata, J. Tang, Y . Wang, L. He, T. K. Lee, W. L. Lee, and K. L. Wang, Phys. Rev. Lett. 113, 137201 (2014). [10] C. Z. Chang, W. Zhao, D. Y . Kim, H. Zhang, B. A. Assaf, D. Heiman, S. C. Zhang, C. Liu, M. H. Chan, and J. S. Moodera, Nat. Mater. 14, 473 (2015). [11] A. Kandala, A. Richardella, S. Kempinger, C. X. Liu, and N. Samarth, Nat. Commun. 6, 7434 (2015). [12] M. Mogi, R. Yoshimi, A. Tsukazaki, K. Yasuda, Y . Kozuka, K. S. Takahashi, M. Kawasaki, and Y . Tokura, Appl. Phys. Lett. 107, 182401 (2015). [13] C. Z. Chang, P. Tang, Y . L. Wang, X. Feng, K. Li, Z. Zhang, Y . Wang, L. L. Wang, X. Chen, C. Liu, W. Duan, K. He, X. C. Ma, and Q. K. Xue, Phys. Rev. Lett. 112, 056801 (2014). [14] P. Wei, F. Katmis, B. A. Assaf, H. Steinberg, P. Jarillo-Herrero, D. Heiman, and J. S. Moodera, Phys. Rev. Lett. 110, 186807 (2013). [15] M. Lang, M. Montazeri, M. C. Onbasli, X. Kou, Y . Fan, P. Upadhyaya, K. Yao, F . Liu, Y . Jiang, W . Jiang, K. L. W ong, G. Y u, J. T ang, T . Nie, L. He, R. N. Schwartz, Y . Wang, C. A. Ross, and K. L. Wang, Nano Lett. 14, 3459 (2014). [16] F. Katmis, V . Lauter, F. S. Nogueira, B. A. Assaf, M. E. Jamer, P. Wei, B. Satpati, J. W. Freeland, I. Eremin, D. Heiman, P. Jarillo-Herrero, and J. S. Moodera, Nature 533, 513 (2016). [17] C. Tang, Chang, C. Z., G. Zhao, Y . Liu, Z. Jiang, C. X. Liu, M. R. McCartney, D. J. Smith, T. Chen, J. S. Moodera, and J. Shi, Sci. Adv. 3, e1700307 (2017). [18] G. Zheng, N. Wang, J. Yang, W. Wang, H. Du, W. Ning, Z. Yang, H. Z. Lu, Y . Zhang, and M. Tian, Sci. Rep. 6, 21334 (2016). [19] Q. I. Yang, M. Dolev, L. Zhang, J. F. Zhao, A. D. Fried, E. Schemm, M. Liu, A. Palevski, A. F. Marshall, S. H. Risbud, and A. Kapitulnik, Phys. Rev. B 88, 081407(R) (2013). [20] Q. I. Yang and A. Kapitulnik, Phys. Rev. B 98, 081403((R) (2018). [21] H. Z. Lu, J. Shi, and S. Q. Shen, Phys. Rev. Lett. 107, 076801 (2011). [22] J. Liao, Y . Ou, X. Feng, S. Yang, C. Lin, W. Yang, K. Wu, K. He, X. Ma, Q. K. Xue, and Y . Li, Phys. Rev. Lett. 114, 216601 (2015). [23] M. Brahlek, N. Koirala, N. Bansal, and S. Oh, Solid State Commun. 215, 54 (2015). [24] Z. Jiang, F. Katmis, C. Tang, P. Wei, J. S. Moodera, and J. Shi, Appl. Phys. Lett. 104, 222409 (2014). [25] X. Che, K. Murata, L. Pan, Q. L. He, G. Yu, Q. Shao, G. Yin, P. Deng, Y . Fan, B. Ma, X. Liang, B. Zhang, X. Han, L. Bi, Q. H. Yang, H. Zhang, and K. L. Wang, ACS Nano 12, 5042 (2018). [26] S. Y . Huang, C. W. Chong, Y . Tung, T. C. Chen, K. C. Wu, M. K. Lee, J. C. Huang, Z. Li, and H. Qiu, Sci. Rep. 7, 2422 (2017). [27] H. Nakayama, M. Althammer, Y . T. Ch en, K. Uchida, Y . Kajiwara, D. Kikuchi, T. Ohtani, S. Geprags, M. Opel, S. Takahashi, R. Gross, G. E. Bauer, S. T. Goennenwein, and E. Saitoh, Phys. Rev. Lett. 110, 206601 (2013). [28] P. E. Seiden, Phys. Rev. 133, A728 (1964). [ 2 9 ] Y . K a j i w a r a , K . H a r i i , S . T a k a h a s h i , J . O h e , K . U c h i d a , M . M i z u g u c h i , H . Umezawa, H. Kawai, K. Ando, K. Takanashi, S. Maekawa, and E. Saitoh, Nature 464, 262 (2010). [30] Y . T. Fanchiang, K. H. M. Chen, C. C. Tseng, C. C. Chen, C. K. Cheng, S. R. Yang, C. N. Wu, S. F. Lee, M. Hong, and J. Kwo, Nat. Commun. 9, 223 (2018). [31] C. N. Wu, C. C. Tseng, Y . T. Fanchiang, C. K. Cheng, K. Y . Lin, S. L. Yeh, S. R. Yang, C. T. Wu, T. Liu, M. Wu, M. Hong, and J. Kwo, Sci. Rep. 8, 11087 (2018). [32] C. Tang, P. Sellappan, Y . Liu, Y . Xu, J. E. Garay, and J. Shi, Phys. Rev. B 94, 140403(R) (2016). [33] A. Quindeau, C. O. Avci, W. Liu, C. Sun, M. Mann, A. S. Tang, M. C. Onbasli, D. Bono, P. M. V oyles, Y . Xu, J. Robinson, G. S. D. Beach, and C. A. Ross, Adv. Electron. Mater. 3, 1600376 (2017). [34] C. C. Chen, K. H. M. Chen, Y . T. Fanchiang, C. C. Tseng, S. R. Yang, C. N. Wu, M. X. Guo, C. K. Cheng, C. T. Wu, M. Hong, and J. Kwo, arXiv:1809.04513 (2018). [35] Y . S. Kim, M. Brahlek, N. Bansal, E. Edrey, G. A. Kapilevich, K. Iida, M. Tanimura, Y. Horibe, S.-W. Cheong, and S. Oh, Phys. Rev. B 84 (2011). [36] E. E. Anderson, Phys. Rev. 134, A1581 (1964). [37] K. Banerjee, J. Son, P. Deorani, P. Ren, L. Wang, and H. Yang, Phys. Rev. B 90, 235427 (2014). [38] H. Z. Lu and S. Q. Shen, Phys. Rev. B 84, 125138 (2011). [39] Supplemental Materials. [40] S. Hikami, A. I. Larkin, and Y . Nagaoka, Prog. Theor. Phys. 63, 707 (1980). [41] J. C. Gallagher, A. S. Yang, J. T. Brangham, B. D. Esser, S. P. White, M. R. Page, K. Meng, . Y ., S. Yu, R. Adur, W. Ruane, S. R. Dunsiger, D. W. McComb, F. Yang, and P. C. Hammel, Appl. Phys. Lett. 109, 072401 (2016). [42] Q. Liu, C. X. Liu, C. Xu, X. L. Qi, and S. C. Zhang, Phys. Rev. Lett. 102, 156603 (2009). [43] I. A. Ado, I. A. Dmitriev, P. M. Ostrovsky, and M. Titov, Europhys. Lett. 111, 37004 (2015). [44] A. T. Lee, M. J. Han, and K. Park, Phys. Rev. B 90, 155103 (2014). [45] J. Sanchez-Barriga, A. Varykhalov, G. Springholz, H. Steiner, R. Kirchschlager, G. Bauer, O. Caha, E. Schierle, E. Weschke, A. A. Unal, S. Valencia, M. Dunst, J. Braun, H. Ebert, J. Minar, E. Golias, L. V . Y ashina, A. Ney, V . Holy, and O. Rader, Nat. Commun. 7, 10559 (2016). [46] T. G. S. M. Rijks, S. K. J. Lenczowski, R. Coehoorn, and W. J. M. de Jonge, Phys. Rev. B 56, 362 (1997). [47] B. F. Miao, S. Y . Huang, D. Qu, and C. L. Chien, Phys. Rev. Lett. 112, 236601 (2014). Figure 1. Transport properties of Bi 2Se3/YIG of various Bi 2Se3 thickness and one Bi2Se3/Al 2O3 bilayer. (a) Sheet resistance ܴୱ vs temperature ܶ( .b) Magnetoresistance (MR) measured at 2 K. (c) The magnetoconductance (MC) obtained by subtracting the contribution from the ܤߚଶ term of Eq. (2). Inset: decomposition of the MC curve into the WL, the WAL, and the B2 components for the 10 nm Bi 2Se3/YIG samples. The MC curve can be we ll-fitted to Eq. (2). (d) Thickness dependence of ߙ and ߙଵ extracted by curve fitting to Eq. (2). Figure 2. (a) Illustration of exchange gap opening and its size dependence on the direction of ( .b) Configurations of three different angular dependent resistance with field rotating in the xy-plane ( ߶ ୶୷-scan), yz-plane ( ߠ୷-scan), and xz-plane (ߙ୶-scan). (e)-(f) MR measurement at 2 K with magnetic field applied at ߠ୷ൌ 90°, 60° and 30°. In (c) and (e), resistances are plotted as a function of the magnetic field strength for Bi 2Se3/Al 2O3 and Bi 2Se3/TmIG, respectively. The field data are further transformed by ܤsinߠ ௬ൌܤ to show the ܤ dependence of MR in (d) and (f). Figure 3. (a) A representative ܴAHെܤ hysteresis loop of Bi 2Se3/TmIG at 100 K. (b) Amplitude of AH conductance ΔߪAH and ܤୡ as a function of temperature. Figure 4. Field- and angular dependent resistances of Bi 2Se3/TmIG and Bi 2Se3/Al 2O3 at 100 K. (a) Field-dependent צܴ ,ܴT, and ܴୄ of Bi 2Se3/TmIG. (b) ߶୶୷-, (c) ߠ୷-, and ߙ୶-dependent resistances of Bi 2Se3/TmIG. Here MR is defined as ൫ܴୱሺ݅ሻെ ܴୱሺ90°ሻ൯/ܴୱሺ90°ሻ with ݅ൌ߶ ୶୷,ߠ୷, and ߙ୶. (d) ߶୶୷-, ߠ୷-, and ߙ୶-dependent resistances of Bi 2Se3/Al 2O3. Fig. S1. Dependence of the fitted parameters on the curve fitting range . The figures in each column display data belonging to a specific sample indicated by the column headlines . The top row displays the MC curve s of each sample , where the insets show the MC curves at 𝐵<0.5 T. The 2nd, 3rd, and 4th rows show the extracted 𝛼0 and 𝛼1, 𝑙0 and 𝑙1, and the electron mobility obtained by curve fitting ( 𝜇L) and Hall measurement s (𝜇H), respectively. Error bars represent the standard errors of the fitted parameters. Fig. S2. Magneto -transport data of 6 nm Bi 2Se3/3 nm Al 2O3/TmIG . (a) The MR curve measured at 2 K. (b) The anomalous Hall resistance 𝑅AH taken at 100 K. Black and redy symbols represent magnetic field sweeping up and down, respectively. Fig. S3 . In-plane MR curve of (a) 9 nm Bi 2Se3/Al 2O3 and (b) 9 nm Bi 2Se3/TmIG bilayers measured at 2 K. Fig. S4 . Planar Hall effect data of 9 nm Bi 2Se3/TmIG. (a) 𝜙xy-dependent 𝑅yx for various 𝐵. Solid lines are fitted curves using Eq. (S2) . (b) Comparison of the extracted AMR and PHE amplitude 𝑅∥−𝑅T as a function of 𝐵. Fig. S5 . 𝑇-dependent MR curves of (a) 7 nm Bi 2Se3/YIG and (b) 9 nm Bi 2Se3/TmIG. 1. Discussion s of other possible origins of the negative MR at intermediate field Negative MR in TI thin films could also result from physical mechanisms other than TRS -breaking of surface states . Below we describe these mechanisms and discuss their possibilities in our Bi 2Se3/YIG and Bi 2Se3/TmIG samples. (i) Defect -induced hopping transport It has been shown that when crystal defects are deliberately introduced into TI thin films, a negative MR shows up as the applied fi eld increases. As described in R ef. [37], the samples that underwent ion milling treatme nts show considerably modified MR characteristics. At low fields, positive MR still dominates indicating the robustness of TSSs and the WAL ef fect against defects. However, a negative MR starts to take over when 𝐵>3 T and shows a quadratic dependence up to 𝐵=9 T. The large - 𝐵 negative MR was attributed to field -dependent hopping probabilities among defect states in transport. In disordered semiconductors, localized spins hos ted by defects lead to spin-dependent scattering of electrons. Such a process is suppressed when a large 𝐵 is applied that effectively align s the localized spins, giving rise to a negative MR. Here we note a key difference of the negative MR of Bi 2Se3/YIG and Bi2Se3/TmIG from that reported in R ef. [37]. The negative MR presented in our work occur s at smaller fields ( 3 T>𝐵>0.3 T), and at large field s the classical positive MR from electron cyclotron motion dominates. This is in sharp contrast to the defect -induced negative MR sho wing quadratic behaviors, who se magnitude is large enough to overcome the contribution of Lorentz deflection . Hence, the defect -induced ho pping transport is unlikely to be the origin of negative MR in Bi 2Se3/YIG and Bi 2Se3/TmIG. (ii) Formation of hybridization gap of top and bottom TSSs It is well -known that in a 3D TI , when its thickness approach 2D limits, the overlap of wave functions of the top and bottom TSSs causes a hybridization gap opened at Dirac point [48]. The 2D limit is 6 QL for Bi2Se3. The negative MR due to the hybridization gap -induced WL has been detected in bulk-insulating (Bi 0.57Sb0.43)2Te3 thin films [49]. The negative MR due to hybridization gap observed in Ref. [49] is actually similar to that in Bi 2Se3/YIG and Bi 2Se3/TmIG and primarily locate at even small er 𝐵< 0.3 T. The emergence of the WL effect upon hybridization gap opening can also be understood as a result of Berry phase 𝜋(1−∆H/2𝐸F) deviating from 𝜋, where ∆H denotes the size of the hybridization gap. In this aspect, both hybrid ization - and TRS -breaking -induced gaps give rise to WL effect . In our experiments, the WL was also detected in Bi 2Se3 thicker than 6 QL grown on YIG and TmIG . The positive MC component can be extracted for Bi2Se3 as thick as 16 QL as shown in Fig. 3(c) . As a comparison, the 9 QL Bi2Se3/Al 2O3 sample show s a very sharp negative MC cusp of WAL. Hence, we can conclude that for Bi2Se3 films at 9 QL or thicker, hybridization between top and bottom surfaces should not be a concern in interpreting the WL in Bi 2Se3/YIG and Bi 2Se3/TmIG. Nevertheless, the similarities between the negative MR reported in Ref. [49] and our work suggests that a Dirac gap is indeed opened, despite of a completely different physical origin. (iii) Quantized 2D bulk bands in thin TIs As studied in Ref. [38], a WL could also arise from quantized 2D subbands in ultrathin TI films. We again compare the MC of 9 QL Bi2Se3/Al 2O3 and 10 QL Bi 2Se3/YIG in Fig. 3(c) . Since the two samples exhibit comparable carrier concentration 𝑛2D, 𝑅s and thickness , the contribution of quantized bulk bands participating in transport in one sample should not differ significantly from the other. Obviously, the negative MR is absent in Bi 2Se3 9 QL/Al 2O3. Hence , we rule out such subbands as the main source of the WL in Bi 2Se3/YIG and Bi 2Se3/TmIG. 2. Curve fitting detail s of Fig. 1 The cusp-like feature at small fields resulted from WAL or WL is usually described by the one -component HLN equation. For our sample s, the negative MR at intermediate fields implies a much larger dephasing field of WL than that of WAL. To characterize the WL component, the curve fitting range is extended , and anoth er component of HLN equation including 𝛽𝐵2 (Eq. (2) ) is introduced into the fitting function. It has been shown that a 𝛽𝐵2 term is necessary to perform the curve fitting in wide ranges of field and tempera ture [50]. The first two terms of Eq. (2) present the competition effects of WL and WAL. The 𝛽𝐵2 term account for the classical cyclotronic moti on and other terms of quantum correction s in the conventional HLN equation [50]. In the following , we discuss the analyses of the MC data and the curve fitting at small and large fiel ds separately. (i) Small -field regime ( 𝐵<1 𝑇): suppressed WAL In this regime the weight of the 𝐵2 term is negligible, so the fitting involves four independent parameters, 𝛼0, 𝛼1, 𝑙0, and 𝑙1 in the beginning . We found that four -parameter fittings do not render reliable results. Instead, it is possible to obtain several sets of fitted parameters that all give reasonably good fits by manually adjust the parameter values. Since WL and WAL terms share the sam e mat hematical form, the curve fitting is valid only when the effective phase coherence lengths 𝑙0 and 𝑙1 differ to some extent s and the fitting range exceeds dephasing fields , otherwise an unique set of fitted parameters cannot be found . In this regime, we thus set 𝛼0 and 𝑙0 of WL to be zero. The results are shown in the 2nd and 3rd rows of Fig. S1. Therefore, only a suppressed WAL term can be concluded in this regime for our sample s. (ii) Larger field regime (𝐵>1 𝑇): negative MR and WL The presence of WL is seen in negative MR located at intermediate -field regime (Fig. 1(b) ). To disclose the characteristics of the WL effect, the curve fittin g range is extended to several T eslas. The MC curves within 4 T are shown in the first row of Fig. S1, and they exhibit a parabolic 𝐵 dependence toward 9 T. In this regime, five -parameter fitting, including the 𝛽𝐵2 term, has been performed . 𝛽 is composed of the classical cyclotronic part 𝛽c and the quantum correction one 𝛽q from the other two term s of the original HLN equation : 𝛽q𝐵2≈−𝑒2 24𝜋ℎ[𝐵 𝐵SO+𝐵𝑒]2 + 3𝑒2 48𝜋ℎ[𝐵 (4/3)𝐵SO+𝐵𝜙]2 , where 𝐵SO and 𝐵𝜙 are characteristic fields of the spin-orbit scattering length 𝑙SO and phase -coherence length 𝑙𝜙. Here, 𝛽<0 due to the negative MC at l arge fields. The five -parameters fitting results are shown in the 2nd to 4th rows of Fig. S1. The MC curves of all samples can be well fitted to Eq. (2), except the one of 40 nm Bi 2Se3/YIG which deviates the most from a 2D electron system . Since fitted param eters depend on the data range selected for the curve fitting , we di splay them as a functio n of the curve fitting range. The fitted pa rameters show a moderate ~10 % variations with respect to the fitting ranges. Throughout th e fitting ranges, the magnitudes of 𝛼0, 𝛼1, 𝑙0, and 𝑙1 can be compared without ambiguity. The reliability of the five - parameter fit s of the data is further justi fied by the following three observations . First ly, the 𝛼1’s and 𝑙1’s of the WAL component obtained from two -parameter fitting s at small fields are in good agreement with those from five -parameter fitting s. This implies the suitability of Eq. (2) for the MC behavior of Bi 2Se3/YIG and Bi 2Se3/TmIG in a wide range of 𝐵. Secondly, from the dephasing field 𝐵i calculated from ℏ/(4𝑒𝑙i2), we note that 𝐵0 is much larger than 𝐵1, which agrees with the observation that the negative MR shows up at larger fields . The notable difference of 𝑙0 and 𝑙1 causes the WL and WAL to manifest themselves in different regimes of 𝐵. Thirdly , if we set 𝛽q≈0, the electron mobility calculated by 𝜇L= √−𝛽c𝑅s overlap well with that calculated by our Hall measurement data 𝜇H=1/(𝑒𝑛2D𝑅xx), indicated by the blue dashed line. 𝛽q≈0 corresponds to a very large 𝐵SO or small 𝑙SO. 3. Data of the controlled sample Bi2Se3/Al 2O3/TmIG A trilayer sample 6 nm Bi 2Se3/3 nm Al 2O3/TmIG has been fabricated to test the effect of stray fields of TmIG. Here, the 3 nm Al 2O3 layer was deposited using atomic layer deposition (ALD). The nonmagnetic Al 2O3 insertion layer ought to suppress the interlayer exchange coupling of Bi 2Se3 and TmIG, while allow s the stray field to penetrate . Fig. S2(a) show s the MR of Bi 2Se3/Al 2O3/TmIG at 2 K. A cusp-like positive MR of the WAL effect was observed, and no negative MR was detected. Fig. S2(b) displays the 𝑅AH data as a function of 𝐵. No hysteresis loop was detected. Therefore, the data in Fig. S2 implies that stray field s are not the root cause of the negative MR and hysteric 𝑅AH loops observed in the Bi2Se3/TmIG. 4. In-plane MR data at 2 K Fig. S3(a) and (b) show the MR data under an in-plane applied (𝜃yz=0) field taken at 2 K for 9 nm Bi 2Se3/Al 2O3 and 9 nm Bi 2Se3/TmIG, respectively. For the Bi2Se3/Al 2O3 bilayer, a positive MR is detected . The MR induced by an in -plane field in Bi 2Se3 has been studied extensively in R ef. [51]. In short, the applica tion of in-plane fields force s the electron to scatter between top and bottom surface s, and the presence of bulk state is essential in understanding the in-plane MR. It was demonstrated that no existing theory can well -describe the distinct transport prope rties o f TIs under an in -plane field, thus highlighting the important role of bulk-surface coupling of TIs . The qualitatively difference between perpendicular and in -plane MR of Bi 2Se3 can already been seen by comparing Fig. 2(d) and S3(a) . While the MR with tilted field angles can be very well explained by WAL governed by 𝐵z, it can be inferred that when 𝑩 is rotated across a critical angle of 𝜃yz (< 30°), another physical picture of magneto -transport that dictates the in -plane MR comes into play in this regime . For Bi 2Se3/TmIG, the in -plane MR exhibit distinct feature s from those of Bi2Se3/Al 2O3: the MR is positive at 𝐵<3 T and becomes negative when 𝐵 goes larger. We may differentiate the physical origin of the in -plane negative MR from that of perpendicular one. From Eq. (1), we see that a gap in the TSS can only be opened by a perpendicular magnetization 𝑀z, while 𝑀x and 𝑀y shift the gapless Dirac cone in the momentum space. Although it is argued that an in -plane magnetic field can also break TRS of TIs when the field is aligned with a certain cry stal axis [52], this should not be of importance since our Bi 2Se3 films grown on gar net substrates contain randomly oriented in-plane domains. It is beyond the scope of this work to clarify the in -plane negative MR in Bi 2Se3/TmIG, especially when the magnetic scatt ering due to MPE adds to the complexity of the syste m. However, we emphasize that the observation of in -plane negative MR does not pose a major problem of our interpretations of the negative MR under tilted fields. As in the case of Bi 2Se3/Al 2O3, a different scheme of physical model is needed for the in -plane MR. 5. Planar Hall effect (PHE) in Bi 2Se3/TmIG For a measurement configuration defined Fig. 2(b), the anisotropic resistivity tensor induced by an in -plane field can be reduced to two elements , 𝑅s (or 𝑅xx) and 𝑅yx, with respect to the sample coordinate. Phenomenologically , the field - angle -dependent 𝑅s and 𝑅yx are identified as AMR and planar Hall resistances, respectively when they are expressed as, AMR : 𝑅s(𝜙xy)=𝑅T+(𝑅∥−𝑅T)cos2𝜙xy (S1) PHE : 𝑅yx(𝜙xy)=(𝑅∥−𝑅T)sin𝜙xycos𝜙xy (S2) Fig. S4(a) shows the 𝑅yx(𝜙xy) data of 9 QL Bi 2Se3/TmIG . The 𝑅yx satisfies the angular dependence sin𝜙xycos𝜙xy of PHE . By fitting the data in Fig. 4(b) and Fig. S4(a) to Eq. (S1) and (S2) respectively, we extract the c oefficient of the angular terms 𝑅∥−𝑅T for 𝑅s and 𝑅yx data. Fig. S 4(b) compares the 𝑅∥−𝑅T obtained from 𝑅s and 𝑅xy data at various fields . One immediately sees a good agreement between the two sets of data . PHE in TI has been previously observed in non-magnetic (Bi,Sb) 2Te3 films [53] and EuS/(Bi,Sb) 2Te3 [54]. In Ref. [53], the PHE results from anisotropic scattering of Dirac fermions due to T RS broken by an i n-plane field. T he PHE amplitude can be altered dramati cally by dual -gating, showing a unique two -peak profile as the Fermi level moves across the Dirac point. We are not able to completely preclude such a scenario in Bi 2Se3/TmIG bilayer, where a similar effect could also be caused by the interfacial e xchange effective field. Fermi -level dependent measurements enabled by top -gating will be performed to investigate this kind of PHE . In Ref. [54], an unconventional PHE was detected, whose angular dependence cannot be described by Eq. (S2). The autho rs argue that a non -linear Hall response defined a s 𝑗y=𝜎yxx𝐸x2 should be considered. The proposed possible origins of the non-linea r Hall response includes current -induced spin -orbit torques from TSSs, asymmetric scattering of electrons by magnons in magnetic TIs, and interband transitions between the two branches of the Dirac surface states. These scenario play minor roles , if any, in Bi 2Se3/TmIG because such an unconventional PHE was not observed in this work . Reference: [48] Y . Zhang, K. He, C. -Z. Chang, C. -L. Song, L. -L. Wang, X. Chen, J. -F. Jia, Z. Fang, X. Dai, W. -Y . Shan, S. -Q. Shen, Q. Niu, X. -L. Qi, S. -C. Zhang, X. -C. Ma, and Q. -K. Xue, Nat. Phys. 6, 584 (2010). [49] Z. Tang, E. Shikoh, H. Ago, K. Kawahara, Y . Ando, T. Shinjo, and M. Shiraishi, Physical Review B 87 (2013). [50] B. A. Assaf, T. Cardinal, P. Wei, F. Katmis, J. S. Moodera, and D. Heiman, Appl. Phys. Lett. 102, 012102 (2013). [51] C. J. Lin, X. Y . He, J. L iao, X. X. Wang, V . S. Iv, W. M. Yang, T. Guan, Q. M. Zhang, L. Gu, G. Y . Zhang, C. G. Zeng, X. Dai, K. H. Wu, and Y . Q. Li, Phys. Rev. B 88, 041307(R) (2013). [52] C. K. Chiu, J. C. Y . Teo, A. P. Schnyder, and S. Ryu, Rev. Mod. Phys. 88 (2016). [53] A. A. Taskin, H. F. Legg, F. Yang, S. Sasaki, Y . Kanai, K. Matsumoto, A. Rosch, and Y . Ando, Nat. Commun. 8, 1340 (2017). [54] D. Rakhmilevich, F. Wang, W. Zhao, M. H. W. Chan, J. S. Moodera, C. Liu, and C. Z. Chang, Phys. Rev. B 98, 094404 (2018). | 2018-11-02 | Transport signatures of exchange gap opening because of magnetic proximity
effect (MPE) are reported for bilayer structures of Bi2Se3 thin films on
yttrium iron garnet (YIG) and thulium iron garnet (TmIG) of perpendicular
magnetic anisotropy (PMA). Pronounced negative magnetoresistance (MR) was
detected, and attributed to an emergent weak localization (WL) effect
superimposing on a weak antilocalization (WAL). Thickness-dependent study shows
that the WL originates from the time-reversal-symmetry breaking of topological
surface states by interfacial exchange coupling. The weight of WL declined when
the interfacial magnetization was aligned toward the in-plane direction, which
is understood as the effect of tuning the exchange gap size by varying the
perpendicular magnetization component. Importantly, magnetotransport study
revealed anomalous Hall effect (AHE) of square loops and anisotropic
magnetoresistance (AMR) characteristic, typifying a ferromagnetic conductor in
Bi2Se3/TmIG, and the presence of an interfacial ferromagnetism driven by MPE.
Coexistence of MPE-induced ferromagnetism and the finite exchange gap provides
an opportunity of realizing zero magnetic-field dissipation-less transport in
topological insulator/ferromagnetic insulator heterostructures. | Evidence for exchange Dirac gap in magneto-transport of topological insulator-magnetic insulator heterostructures | 1811.00689v1 |
Magnon contribution to unidirectional spin Hall magnetoresistance in ferromagnetic-insulator/heavy-metal bilayers W.P. Sterk and D. Peerlings Institute for Theoretical Physics, Utrecht University, Princetonplein 5, 3584 CC Utrecht, The Netherlands R.A. Duine Institute for Theoretical Physics, Utrecht University, Princetonplein 5, 3584 CC Utrecht, The Netherlands and Department of Applied Physics, Eindhoven University of Technology, PO Box 513, 5600 MB Eindhoven, The Netherlands (Dated: March 1, 2019) We develop a model for the magnonic contribution to the unidirectional spin Hall magnetore- sistance (USMR) of heavy metal/ferromagnetic insulator bilayer films. We show that diffusive transport of Holstein-Primakoff magnons leads to an accumulation of spin near the bilayer interface, giving rise to a magnoresistance which is not invariant under inversion of the current direction. Un- like the electronic contribution described by Zhang and Vignale [Phys. Rev. B 94, 140411 (2016)], which requires an electrically conductive ferromagnet, the magnonic contribution can occur in fer- romagnetic insulators such as yttrium iron garnet. We show that the magnonic USMR is, to leading order, cubic in the spin Hall angle of the heavy metal, as opposed to the linear relation found for the electronic contribution. We estimate that the maximal magnonic USMR in Pt|YIG bilayers is on the order of 10 8, but may reach values of up to 10 5if the magnon gap is suppressed, and can thus become comparable to the electronic contribution in e.g. Pt|Co. We show that the magnonic USMR at a finite magnon gap may be enhanced by an order of magnitude if the magnon diffusion length is decreased to a specific optimal value that depends on various system parameters. PACS numbers: 73.43.Qt, 75.76.+j I. INTRODUCTION Thetotalmagnetoresistanceofmetal/ferromagnethet- erostructures is known to comprise several independent contributions, including but not limited to anisotropic magnetoresistance (AMR) [1], giant magnetoresistance (GMR, in stacked magnetic multilayers) [2] and spin Hall magnetoresistance (SMR) [3]. A common characteristic of these effects is that they are linear; in particular, this means the measured magnetoresistance is invariant un- der reversal of the polarity of the current. In 2015, however, Avci et al.[4] measured a small but distinct asymmetry in the magnetoresistance of Ta|Pt and Co|Pt bilayer films. Due to its striking similarity to the current-in-plane spin Hall effect (SHE) and GMR, save for its nonlinear resistance/current characteristic, this effect was dubbed unidirectional spin Hall magne- toresistance (USMR). In the years following its discovery, USMR has been detected in bilayers consisting of magnetic and nonmag- netictopologicalinsulators[5], andthedependenceofthe USMR on layer thickness has been investigated experi- mentally for Co|Pt bilayers [6]. Additionally, Avci et al. [7] have shown that USMR may be used to distinguish between the four distinct magnetic states of a ferromag- net|normal metal|ferromagnet trilayer stack, highlighting itspotentialapplicationinmultibitelectricallycontrolled memory cells. Although USMR is ostensibly caused by spin accumu- lationattheferromagnet|metalinterface, acompletethe- oretical understanding of this effect is lacking. In bilayerfilms consisting of ferromagnetic metal (FM) and heavy metal (HM) layers, electronic spin accumulation in the ferromagnet caused by spin-dependent electron mobility provides a close match to the observed results [8]. It re- mains unknown, however, whether this is the full story; indeed, this model’s underestimation of the USMR by a factor of two lends plausibility to the idea that there may be additional, as-yet unknown contributions provid- ing the same experimental signature. Additionally, the electronic spin accumulation model cannot be applied to bilayers consisting of a ferromagnetic insulator (FI) and a HM, as there will be no electric current in the ferro- magnet to drive accumulation of spin. Kim et al.[9] have measured the USMR of Py|Pt (where Py denotes for permalloy) bilayer and claim, us- ing qualitative arguments, that a magnonic process is in- volved. Likewise, for Co|Pt and CoCr|Pt, more recent re- sults by Avci et al.[10] argue in favor of the presence of a magnon-scatteringcontributionconsistingoftermslinear and cubic in the applied current, and having a magnitude comparable to the electronic contribution of Zhang and Vignale [8]. Although these experimental results provide a great deal of insight into the underlying processes, a theoretical framework against which they can be tested is presently lacking. In this work, we aim to take first steps to developing such a framework, by considering an accumulation of magnonic spin near the FI|HM bilayer interface, which we describe by means of a drift-diffusion model. The remainder of this article is structured as follows: in Sec. II, we present our analytical model as genericallyarXiv:1810.02610v2 [cond-mat.mes-hall] 28 Feb 20192 as possible. In Sec. III we analyze the behavior of our model using parameters corresponding to a Pt|YIG (YIG being yttrium iron garnet) bilayer as a basis. In par- ticular, in Sec. IIIA we give quantitative predictions of the magnonic USMR in terms of the applied current and layer thicknesses, and in Sec. IIIB we take into account the effect of Joule heating. In the remainder of Sec. III, we investigate the influence of various material param- eters. Finally, in Sec. IV we summarize our key results and present some open questions. II. MAGNONIC SPIN ACCUMULATION To develop a model of the magnonic contribution to the USMR, we focus on the simplest FI|HM heterostruc- ture: a homogeneous bilayer. We treat the transport of magnonic and electronic spin as diffusive, and solve the resulting diffusion equations subject to a quadratic boundary condition at the interface. In this approach, valid in the opaque interface limit, current-dependent spin accumulations—electronic in the HM and magnonic in the FI—form near the interface. In particular, the use of a nonlinear boundary condition breaks the invariance of the SMR under reversal of the current direction, i.e. it produces USMR. We consider a sample consisting of a FI layer of thick- nessLFIdirectlycontactingaHMlayerofthickness LHM. We take the interface to be the xyplane, such that the FI layer extends from z= 0toLFIand the HM layer from z= LHMto 0. The magnetisation is chosen to lie in the positive y-direction, and an electric field E=E^ xis applied in the x-direction. The set-up is shown in Fig. 1. The extents of the system parallel to the interface are taken to be infinite, and the individual layers com- pletely homogeneous. This allows us to treat the system as quasi-one-dimensional, in the sense that we will only consider spin currents that flow in the z-direction. We account for magnetic anisotropy only indirectly through the existence of a magnon gap. We further assume that our system is adequately described by the Drude model (suitably extended to include spin effects[11]), and that the interface between layers is not fully transparent to spin current, i.e., has a finite spin-mixing conductance [12]. For simplicity, we assume electronic spin and charge transport may be neglected in the ferromagnet, as is the case for ferromagnetic insulators. Wedescribethetransferofspinacrosstheinterfacemi- croscopically by the continuum-limit interaction Hamil- tonian Hint= Z d3rd3r0J(r;r0)h by(r0)cy #(r)c"(r) +b(r0)cy "(r)c#(r)i ; wherecy (r)[c(r)] are fermionic creation [annihilation] operators of electrons with spin 2f";#gat position rin the HM, and by(r0)[b(r0)] is the bosonic creation FIG. 1. Schematic depiction of our system. The magnetiza- tionMof the FI layer lies in the +ydirection, an electric field of magnitude Eis applied to the heavy metal layer (HM) in thexdirection, and the interface between the layers lies in thexyplane. [annihilation] operator of a circularly polarized Holstein- Primakoff magnon [13] at position r0inside the ferro- magnet. We leave J(r;r0)to be some unknown coupling between the electrons and magnons, which is ultimately fixed by taking the classical limit [14, 15]. Transforming to momentum space and using Fermi’s golden rule, we obtain the interfacial spin current jint s, which can be expressed in terms of the real part of the spin mixing conductance per unit area g"# ras [14, 16] jint s=g"# r sZ d"g(")(" ) nB" kBTe nB" m kBTm :(1) (Similar expressions were derived by Takahashi et al.[17] and Zhang and Zhang [18], although these are not given in terms of the spin-mixing conductance.) Here,sis the saturated spin density in the FI layer, g(")is the magnon density of states, nB(x) = [ex 1] 1 is the Bose-Einstein distribution function, kBis Boltz- mann’s constant, and TmandTeare the temperatures of the magnon and electron distributions, respectively, which we do not assume a priori to be equal (al- though the equal-temperature special case will be our primary interest). Of crucial importance in Eq. (1) are the magnon effective chemical potential m—which we shall henceforth primarily refer to as the magnon spin accumulation—and the electron spin accumulation " #, which we define as the difference in chem- ical potentials for the spin-up and spin-down electrons. (In both cases, a positive accumulation means the major- ity of spin magnetic moments point in the +ydirection.) We employ the magnon density of states g(") =p " 42J3 2s(" ): Here,Jsis the spin wave stiffness constant, (x)is the Heavisidestepfunction, and isthemagnongap, caused3 by a combination of external magnetic fields and in- ternal anisotropy fields in ferromagnetic materials [19]. In our primary analysis of a Pt|YIG bilayer, we take B1 TkB0:67 KwithBthe Bohr mag- neton, in good agreement with e.g. Cherepanov et al. [20], and in Sec. IIIE we specifically consider the limit of a vanishing magnon gap. To treat the accumulations on equal footing, we now redefinem!mand!, expand Eq. (1) to second order in , and set= 1to obtain jint s' " kBTmI0+Ie+Imm+Iee kBTe()2 +Imm kBTm2 m+Ime kBTmm# g"# r(kBTm)3 2 43J3 2ss:(2) Here, the Iiare dimensionless integrals given by Eqs. (A.1) in the Appendix. All Iiare functions of Tm and, andI0,IeandIeeadditionally depend on Te. In the special case where Tm=Te,I0vanishes,Im= Ie, andIee= (Imm+Ime). Inadditionto jint s, thespinaccumulationsandtheelec- tric driving field Egive rise to the following spin currents in thezdirection: je s=h 2e 2e@ @z SHE ;(3a) jm s= m h@m @z: (3b) Hereje sandjm sare the electron and magnon spin cur- rents, respectively. is the electrical conductivity in the HM,mis the magnon conductivity in the ferromagnet, eis the elementary charge, and SHis the spin Hall angle. In line with Cornelissen et al.[21] and Zhang and Zhang [22], we assume the spin accumulations mand obey diffusion equations along the z-axis: d2m dz2=m l2m;d2 dz2= l2e; wherelmandleare the magnon and electron diffusion lengths, respectively. We solve these equations analyti- cally subject to boundary conditions that demand con- tinuity of the spin current across the interface and con- finement of the currents to the sample: jm s(0) =je s(0) =jint s(0); jm s(LFI) =je s( LHM) = 0: This system of equations now fully specifies the magnonic and electronic spin accumulations mand, the latter of which enters the charge current jcvia the spin Hall effect: jc(z) =E+SH 2e@(z) @z: (4)The measured resistivity at some electric field strength Eis then given by the ratio of the electric field and the averaged charge current: (E) =E 1 LHMR0 LHMdzjc(z): (5) Finally, we define the USMR Uas the fractional differ- ence in resistivity on inverting the electric field: U(E) ( E) (E)=1 +R0 LHMdzjc(z;E) R0 LHMdzjc(z; E): It should be noted that the even-ordered terms in the expansion of the interface current are vital to the appear- anceofunidrectionalSMR.Supposeoursystemhasequal magnon and electron temperature, such that the interfa- cial spin Seebeck term I0vanishes (see Section IIIB), and we ignore the quadratic terms in Eq. (2). Then be- cause the only term in the spin current equations (3) that is independent of the accumulations is hSH 2eEin Eq. (3a), we have that /m/E. Then by Eqs. (4) and (5),jc/Eand(E)/E E, such thatU= 0. Con- versely, with quadratic terms in the interfacial spin cur- rent,(E)E E+E2, and likewise if I0does not vanish, (E)E 1+E. BothcasesgivenonvanishingUSMR.Phys- ically, one can say that the spin-dependent electron and magnon populations couple together in a nonlinear fash- ion (namely, through the Bose-Einstein distributions in Eq. (1)), leading to a nonlinear dependence on the elec- tric field. III. RESULTS A. Equal-temperature, finite gap case Although our model can be solved analytically (up to evaluation of the integrals Ii), the full expression ofUis unwieldy and therefore hardly insightful. To get an idea of the behavior of a real system, we use a set of parameters—listed in Table I—corresponding to a Pt|YIG bilayer as a starting point. (Unless otherwise specified, all parameters used henceforth are to be taken from this table.) Fig. 2 shows the magnonic USMR of a Pt|YIG bilayer versus applied driving current ( E) whenTm=Te=T, at the temperature of liquid nitrogen ( 77 K, blue), room temperature ( 293 K, green) and the Curie temperature of YIG ( 560 K[20], red). FI and HM layer thicknesses used are90 nmand3 nm, respectively, in line with experimen- tal measurements by Avci et al.[23]. In all cases the magnonic USMR is proportional to the applied electric current—that is, the cubic term found by Avci et al.[10] is absent—and at room temperature has a value on the order of 10 9at typical measurement currents [4]. This is roughly four orders of magnitude weaker than the USMR obtained—both experimentally4 0.0 0.2 0.4 0.6 0.8 1.0 σE(A m−2)×10120246U×10−9 T(K) 77 293 5600 600T(K)0.00.5×10−8 FIG. 2. USMRUversus driving current Efor a Pt|YIG bilayer at liquid nitrogen temperature ( 77 K, blue), room temperature ( 293 K, green) and the YIG Curie temperature (560 K, red). Inset: USMR versus system temperature Tat fixed current E= 11012A m 2. and theoretically—for FM|HM hybrids [4, 6, 8, 23], and is consistent with the experimental null results obtained for this system by Avci et al.[23]. Note, however, that the thickness of the FI layer used by these authors is significantly lower than the magnon spin diffusion length lm= 326 nm , which results in a suppressed USMR. Furthermore, it can be seen in the inset of Fig. 2 that the magnonic USMR is, to good approximation, linear in the system temperature, in agreement with observations by Kim et al.[9] and Avci et al.[10]. 1 2 3 4 5 LFI(µm)1020304050LHM(nm) 0.00.51.01.52.02.53.03.54.0 U×10−8 FIG. 3. Pt|YIG USMR UatTm=Te= 293 K versus FI layer thickness LFIand HM layer thickness LHM. A driving currentE= 11012A m 2is used. A maximal USMR of 4:210 8is reached at LHM= 4:5 nm,LFI= 5µm. In Fig. 3 we compute the USMR at E = 11012A m 2as a function of both LFIandLHM. A maximum is reached around LHM4:5 nm, while in terms ofLFI, a plateau is approached within a few spindiffusion lengths. By varying the layer thicknesses, a maximal USMR of 4:210 8can be achieved, an im- provement of one order of magnitude compared to the thicknesses used by Avci et al.[23]. B. Thermal effects We take into account a difference between the electron and magnon temperatures TeandTmby assuming these parameters are equal to the temperatures of the HM and FI layers, respectively, which we take to be homogeneous. We assume that the HM undergoes ohmic heating and dissipates this heat into the ferromagnet, which we take to be an infinite heat bath at temperature Tm. We only take into account the interfacial (Kapitza) thermal resis- tanceRthbetween the HM and FI layers, leading to a simple expression for the HM temperature Te: Te=Tm+RthE2LHM: Using this model, we still find a linear dependence in the electric field, U 'uE(Tm)E, but the coefficient uE(Tm)increases by three orders of magnitude compared to the case where the electron and magnon temperatures are set to be equal. The overwhelming majority of this increase can be attributed to an interfacial spin Seebeck effect (SSE) [21, 24]: it is caused by the accumulation- independent contribution I0(Eq. (A.1a)) in the interface current. When I0is artificially set to 0, uE(Tm)changes less than 1% from its equal-temperature value. Furthermore, the overall magnitude of the interfacial SSE in our system can be attributed to the fact that we have a conductor|insulator interface: the current runs through the HM only, resulting in inhomogeneous Joule heating of the sample and a large temperature disconti- nuity across the interface. C. Spin Hall angle The electronic spin accumulation at the interface in the standard spin Hall effect is linear in the electric fieldEand spin Hall angle SH[3]. From the linearity inE, we may conclude that the terms in Eq. (2) that are linear in have a suppressed contribution to the USMR. Thus, the contribution of the interface current is of order2 SH. Furthermore, enters the charge cur- rent (Eq. (4)) with a prefactor SH, leaving the magnonic USMR predominantly cubic in the spin Hall angle. In- deed, in the special case Tm=Te, expanding the full expression forU(which spans several pages and is there- fore not reproduced within this work) in SHreveals that the first nonzero coefficient is that of 3 SH. This suggests a small change in SHpotentially has a large effect on the USMR. In Fig. 4 we plot the USMR for a Pt|YIG bilayer— once again using Tm=Te= 293 K—consisting of 4:5 nm5 0.0 0.1 0.2 0.3 0.4 0.5 θSH01234U×10−6 Computed Cubic fit FIG. 4. USMRUatTm=Te= 293 K versus spin Hall angle SH. A driving current E= 11012A m 2and FI and HM layer thicknesses LFI= 5µmandLHM= 4:5 nmare used. Blue curve: computed value. Dashed green curve: fit of the formU=u3 SH, withu'3:110 4. of Pt and 5µmof YIG, in which we sweep the spin Hall angle. Included is a cubic fit U=u3 SH, where we findu'3:110 4. Here it can be seen that the magnonic USMR in HM|FI bilayers can, as expected, po- tentiallyacquiremagnitudesroughlycomparabletothose in HM|FM systems, provided one can find or engineer a metal with a spin Hall angle several times greater than that of Pt. This suggests that very strong spin-orbit coupling (SOC) is liable to produce significant magnon- mediated USMR in FI|HM heterostructures, although we expect our model to break down in this regime. D. A note on the magnon spin diffusion length Although we use the analytic expression for the magnon spin diffusion length[18, 21, 22], lm=vthr 2 3mr —wherevthis the magnon thermal velocity, is the com- binedrelaxationtime, and mristhemagnonicrelaxation time (see Table I)—this is known to correspond poorly to reality, being at least an order of magnitude too low in the case of YIG [21]. Artificially setting the magnon spin-diffusion length to the experimental value of 10µm (while otherwise continuing to use the parameters from Table I) results in a drop in USMR of some 4 orders of magnitude. It follows directly that there exists some optimal value oflm(which we shall label lm;opt) that maximizes the USMR, which we plot as a function of the FI layer thicknessLFIin Fig. 5, at LHM= 4:5 nmandE= 11012A m 2, and for various values of the magnon- phonon relaxation time mp, which is the shortest and therefore most important timescale we take into account.For the physically realistic value of mp= 1 ps(blue curve), the optimal magnon spin diffusion length is just 24 nm. Although lm;optitself depends on mp, the con- ditionlm=lm;optacts to cancel the dependence of the USMR on the magnon-phonon relaxation time. Curi- ously, the USMR additionally loses its dependence on LFI, reaching a fixed value of 4:1410 7for our param- eters. 0 5 10 15 LFI(µm)0.00.51.01.52.02.53.0lm,opt (µm)τmp(s) 10−12 10−11 10−10 10−9 10−8 ∞ FIG. 5. Value of the magnon spin diffusion length lmthat maximizes the USMR, as a function of FI layer thickness LFI, at various values of the magnon-phonon relaxation time mp. We further find that lm;optis independent of the spin Hall angle and driving current, and shows a weak de- creasewithincreasingtemperatureprovidedthemagnon- phonon scattering time is sufficiently short. A significant increase in the optimal spin diffusion length is only found at low temperatures and large mp. Similarly, a weak de- pendence on the Gilbert damping constant is found, becoming more significant at large mp, with lower val- ues ofcorresponding to larger lm;opt. Whenis swept, again the USMR at lm=lm;optacquires a universal value of4:1410 7for our system parameters. E. Effect of the magnon gap We have thus far utilised a fixed magnon gap with a valueof =B= 1 TforYIG.Althoughthisisreasonable for typical systems, it is possible to significantly reduce the gap size by minimizing the anisotropy fields within the sample, e.g. using a combination of external fields [25], optimized sample shapes [19, 26] and temperature [27, 28]. This leads us to consider the effect a decreased or even vanishing gap may have on our results. Fig. 6 shows the USMR Ufor a Pt|YIG system ( 4:5 nm of Pt and 5µmof YIG) at room temperature, plotted against the driving current E, now for different values of the magnon gap . Here it can be seen that while U is linear in Efor large gap sizes and realistic currents,6 0.00 0.25 0.50 0.75 1.00 σE(A m−2)×10120.00.51.01.5U×10−5 ∆/µB(T) 10−9 10−6 10−5 10−4 10−3 100 FIG. 6. USMRUof a Pt( 4:5 nm)|YIG( 5µm) bilayer at room temperatureversusappliedcurrent Eatvariousvaluesofthe magnon gap . For large gaps, linear behavior is recovered at realistic currents, while for smaller gap sizes, the USMR saturates as the current is increased. it shows limiting behavior at smaller gaps, becoming in- dependent of the electric current above some threshold (provided one neglects the effect of Joule heating). At low current and intermediate magnon gap, the current dependenceisnonlinearat O(I2)asopposedtothe O(I3) behavior found by Avci et al.[10]. Note also that the saturation value of the USMR is two to three orders of magnitude greater than the values found previously in our work, and of the same magni- tude as the electronic contribution found by Zhang and Vignale [8]. The maximal value of the USMR that can be achieved may be found by considering the full analytic expression forUin terms of the generic coefficients Iirepresenting the dimensionless integrals given by Eqs. (A.1) in the Appendix. In the gapless limit !0and at equal magnonandelectrontemperature( Tm=Te), thesecond- order coefficients ImmandImediverge, while their sum takes the constant value Imm+Ime'0:323551at room temperature. Ieedoes not diverge, and obtains the value . Now working in the thick-ferromagnet limit ( LFI! 1), we substitute Ime! Imm+and take the limits E!1andImm! 1. By application of l’Hôpital’s rule in the latter, all coefficients Iidrop out of the ex- pression forU. This leaves only the asymptotic value, which, after expanding in SH, reads Umax=4e2l2 s2 SHmtanh2 LHM 2ls h2lmLHM+4lse2LHMmcoth LHM ls+O(4 SH): (6) Whereas the linear-in- Eregime of the magnonic USMR growsas3 SH,wethusfindthattheleading-orderbehavior 10−1710−1410−1110−810−510−2101 ∆/µB(T)10−810−710−610−510−4UσE(A m−2) 107 108 109 1010 1011 1012FIG. 7. USMRUof a Pt( 4:5 nm)|YIG( 5µm) bilayer at room temperature as function of the magnon gap size , for various valuesofthebasechargecurrent E. Notethelog-logscaling. Solid colored lines: computed USMR. Dashed colored lines: continuationsofthehigh-gaptailsofthecorrespondingcurves accordingtotheone-parameterfit U=u0=p . Dashedblack line: asymptotic value of the USMR as given by Eq. (6). of the asymptotic value is only 2 SH, and the third-order term vanishes completely. Physically, this can be ex- plained by the fact that the asymptotic magnonic USMR is purely a bulk effect: all details about the interface van- ish, while parameters originating from the bulk spin- and charge currents remain. The appearance of lmin the de- nominator and its absence in the numerator of Eq. (6) once again highlights that a large magnon spin diffusion length acts to suppress the USMR. Fig. 7 is a log-log plot of the USMR versus gap size at various values of the driving current E. Here the valueUmaxis shown as a dashed black line, indicating that this is indeed the value to which Uconverges in the gapless limit or at high current. Moreover, it shows that for given E, one can find a turning point at which the USMR switches relatively abruptly from being nearly constant to decreasing as 1=p . A (backwards) continuation of the decreasing tails is included in Fig. 7 as dashed lines following the one- parameter fitU=u0=p , and we define the threshold gapthas the value of where this continuation in- tersectsUmax. We then find that thscales asE2, or conversely, that the driving current required to saturate the USMR scales as the square root of the magnon gap. We note that although the small-gap regime is math- ematically valid (even in the limit !0, asmay be brought arbitrarily close to 0 in a continuous manner), it does not necessarily correspond to a physical situa- tion: when the anisotropy vanishes, the magnetization of the FI layer may be reoriented freely, which will break our initial assumptions. Nevertheless, in taking the gap- less limit, we are able to predict an upper limit on the magnonic USMR.7 IV. CONCLUSIONS Using a simple drift-diffusion model, we have shown that magnonic spin accumulation near the interface be- tween a ferromagnetic insulator and a heavy metal leads to a small but nonvanishing contribution to the unidirec- tional spin Hall magnetoresistance of FI|HM heterostruc- tures. Central to our model is an interfacial spin current originating from a spin-flip scattering process whereby electronsintheheavymetalcreateorannihilatemagnons in the ferromagnet. This current is markedly nonlinear in the electronic and magnonic spin accumulations at the interface, and it is exactly this nonlinearity which gives rise to the magnonic USMR. For Pt|YIG bilayers, we predict that the magnonic USMRUis at most on the order of 10 8, roughly three orders of magnitude weaker than the measured USMR in FM|HM hybrids (where electronic spin accumulation is thought to form the largest contribution). This is fully consistent with experiments that fail to detect USMR in Pt|YIG systems, as the tiny signal is drowned out by the interfacial spin Seebeck effect, which has a similar ex- perimental signature and is enhanced compared to the FM|HM case due to inhomogeneous Joule heating. We have shown that the magnon-mediated USMR is approximately cubic in the spin Hall angle of the metal, suggesting that metals with extremely large spin Hall angles may provide a significantly larger USMR than Pt. It is therefore plausible that a large magnonic USMR can exist in systems with very strong spin-orbit coupling, even though our model would break down in this regime. The magnonic USMR depends strongly on the magnon spin diffusion length lmin the ferromagnet. Motivated by a large discrepancy between experimental values and theoretical predictions of lm, we have shown that a sig- nificant increase in USMR can be realized if a method is found to engineer this parameter to specific, optimal values that, for realistic values of the magnon-phonon relaxation time mp(on the order of 1 psfor YIG), are significantly shorter than those measured experimentally or computed theoretically. We further find that when the magnon spin diffusion length has its optimal value, the USMR becomes independent of the ferromagnet’s thick- ness and Gilbert damping constant. Although in physically reasonable regimes, themagnonic USMR is to very good approximation linear in the applied driving current E, it saturates to a fixed valuegivenextremelylargecurrentsorastronglyreduced magnon gap . The transition from linear to constant behavior in the driving current is heralded by a turn- ing point which is proportional to the square root of the magnon gap. The asymptotic behavior of the USMR be- yond the turning point is governed by the bulk spin- and charge currents, and is completely independent of the de- tails of the interface. While a vast reduction in is required to bring the saturation current of a Pt|YIG bilayer within experimen- tally reasonable regimes, the magnonic USMR scales as 1=p at currents below the turning point, suggesting that highly isotropic FI|HM samples are most likely to produce a measurable magnonic USMR. The increase in magnonic USMR at low gaps (and large currents) is in good qualitative agreement with the recent experimental work of Avci et al.[10], as is the linear dependence on system temperature. A notable disagreement with the experimental data of Avci et al.[10] is found in the scaling of the current de- pendence, which in our results lacks an O(I3)term at large magnon gaps and contains an O(I2)term at inter- mediate gaps. It is still unclear whether this discrepancy can be explained by system differences, such as the fi- nite electrical resistance of Co or the presence of Joule heating. Finally, we note that while our results apply to fer- romagnetic insulators, it is reasonable to assume a magnoniccontributionalsoexistsinHM|FMheterostruc- tures, although the possibility of coupled transport of magnons and electrons makes such systems more diffi- cult to model. Additionally, various extensions of our model may be considered, such as the incorporation of spin-momentum locking [5], ellipticity of magnons, heat transport and nonuniform temperature profiles [21], di- rectional dependence of the magnetization, etc. V. ACKNOWLEDGEMENTS R.A.D. is member of the D-ITP consortium, a pro- gram of the Dutch Organisation for Scientific Research (NWO) that is funded by the Dutch Ministry of Educa- tion, Culture and Science (OCW). This work is funded by the European Research Council (ERC). [1] T. McGuire and R. Potter, IEEE Transactions on Mag- netics11, 1018 (1975). [2] G. Binasch, P. Grünberg, F. Saurenbach, and W. Zinn, Physical Review B 39, 4828 (1989). [3] Y.-T. Chen, S. Takahashi, H. Nakayama, M. Althammer, S. T. B. Goennenwein, E. Saitoh, and G. E. W. Bauer, Physical Review B 87, 144411 (2013), arXiv:1302.1352 [cond-mat.mes-hall].[4] C. O. Avci, K. Garello, A. Ghosh, M. Gabureac, S. F. Alvarado, and P. Gambardella, Nature Physics 11, 570 (2015), arXiv:1502.06898 [cond-mat.mes-hall]. [5] K. Yasuda, A. Tsukazaki, R. Yoshimi, K. S. Takahashi, M. Kawasaki, and Y. Tokura, Physical Review Letters 117, 127202 (2016), arXiv:1609.05906 [cond-mat.mtrl- sci].8 [6] Y. Yin, D.-S. Han, M. C. H. de Jong, R. Lavrijsen, R. A. Duine, H. J. M. Swagten, and B. Koopmans, Applied Physics Letters 111, 232405 (2017), arXiv:1711.06488 [cond-mat.mtrl-sci]. [7] C. O. Avci, M. Mann, A. J. Tan, P. Gambardella, and G. S. D. Beach, Applied Physics Letters 110, 203506 (2017). [8] S. S.-L. Zhang and G. Vignale, Physical Review B 94, 140411 (2016), arXiv:1608.02124 [cond-mat.mes-hall]. [9] K. J. Kim, T. Moriyama, T. Koyama, D. Chiba, S. W. Lee, S. J. Lee, K. J. Lee, H. W. Lee, and T. Ono, ArXiv e-prints (2016), arXiv:1603.08746 [cond-mat.mtrl-sci]. [10] C. O. Avci, J. Mendil, G. S. D. Beach, and P. Gam- bardella, Physical Review Letters 121, 087207 (2018). [11] E. M. Chudnovsky, Physical Review Letters 99, 206601 (2007), arXiv:0709.0725 [cond-mat.mes-hall]. [12] W. Zhang, W. Han, X. Jiang, S.-H. Yang, and S. S. P. Parkin, Nature Physics 11, 496 (2015), arXiv:1504.07929 [cond-mat.mes-hall]. [13] T. Holstein and H. Primakoff, Physical Review 58, 1098 (1940). [14] S. A. Bender and Y. Tserkovnyak, Physical Review B 91, 140402 (2015), arXiv:1409.7128 [cond-mat.mes-hall]. [15] Y. Tserkovnyak, A. Brataas, and G. E. Bauer, Physical Review Letters 88, 117601 (2002), cond-mat/0110247. [16] S. A. Bender, R. A. Duine, and Y. Tserkovnyak, Physi- cal Review Letters 108, 246601 (2012), arXiv:1111.2382 [cond-mat.mes-hall]. [17] S. Takahashi, E. Saitoh, and S. Maekawa, in Journal of Physics Conference Series ,JournalofPhysicsConference Series, Vol. 200 (IOP Publishing, 2010) p. 062030.[18] S. S.-L. Zhang and S. Zhang, Physical Review B 86, 214424 (2012), arXiv:1210.2735 [cond-mat.mes-hall]. [19] C. Tang, M. Aldosary, Z. Jiang, H. Chang, B. Madon, K. Chan, M. Wu, J. E. Garay, and J. Shi, Applied Physics Letters 108, 102403 (2016). [20] V. Cherepanov, I. Kolokolov, and V. L’vov, Physics Re- ports229, 81 (1993). [21] L. J. Cornelissen, K. J. H. Peters, G. E. W. Bauer, R. A. Duine, and B. J. van Wees, Physical Review B 94, 014412 (2016), arXiv:1604.03706 [cond-mat.mes-hall]. [22] S. S.-L. Zhang and S. Zhang, Physical Review Let- ters109, 096603 (2012), arXiv:1208.5812 [cond-mat.mes- hall]. [23] C. O. Avci, K. Garello, J. Mendil, A. Ghosh, N. Blasakis, M. Gabureac, M. Trassin, M. Fiebig, and P. Gam- bardella, Applied Physics Letters 107, 192405 (2015). [24] M. Schreier, A. Kamra, M. Weiler, J. Xiao, G. E. W. Bauer, R. Gross, and S. T. B. Goennenwein, Physi- cal Review B 88, 094410 (2013), arXiv:1306.4292 [cond- mat.mes-hall]. [25] M. G. Pini, P. Politi, and R. L. Stamps, Physical Review B72, 014454 (2005), cond-mat/0503538. [26] H. Skarsvåg, C. Holmqvist, and A. Brataas, Physical Re- viewLetters 115,237201(2015),arXiv:1506.06029[cond- mat.mes-hall]. [27] J. F. Dillon, Physical Review 105, 759 (1957). [28] G. P. Rodrigue, H. Meyer, and R. V. Jones, Journal of Applied Physics 31, S376 (1960). [29] ASM Handbook Committee, ASM handbook , Vol. 2 (ASM International, Materials Park, Ohio, 1990). Appendix: Interfacial spin current integrals The following dimensionless integrals appear in the second-order expansion of the interfacial spin current to the spin accumulations, Eq. (2): I0=Z1 kBTmdxr x kBTmx nB(x) nBTm Tex ; (A.1a) Ie=Z1 kBTmdxr x kBTm nBTm Tex nB(x) Tm TexeTm Tex nBTm Tex2! ; (A.1b) Im=Z1 kBTmdxr x kBTmxex[nB(x)]2; (A.1c) Iee=Z1 kBTmdxr x kBTm eTm Tex nBTm Tex3 eTm Tex 1 Tmx 2Te eTm Tex+ 1! ; (A.1d) Imm=Z1 kBTmdxr x kBTmx 2ex[ex+ 1] [nB(x)]3; (A.1e) Ime= Z1 kBTmdxr x kBTmex[nB(x)]2: (A.1f)9 Description Symbol Expression Value at T= 293 K Ref. YIG spin-wave stiffness constant Js 8:45810 40J m2[21] YIG spin quantum number per unit cell S 10 [21] YIG lattice constant a 1:2376 nm [21] YIG Gilbert damping constant 110 4[21] YIG spin number density s Sa 35:27541027m 3[21] YIG magnon gap 9 :310 24J[20] YIG magnon-phonon scattering time mp 1 ps[21] YIG magnon relaxation time mrh 2kBTm130 ps[21] Combined magnon relaxation time 1 mr+1 mp 1 1 ps[21] Magnon thermal de Broglie wavelength q 4Js kBTm1:62 nm[21] Magnon thermal velocity vth2p JskBT h35:1 km s 1[21] Magnon spin diffusion length lmvthq 2 3mr 326 nm [21] Magnon spin conductivity m 3 22Js 3 1:3510 24J s m 1[21] Real part of spin-mixing conductance g"# r 51018m 2[16] Pt electrical conductivity 1107S m 1[29]a Pt spin Hall angle SH 0.11 [21] Pt electron diffusion length ls 1:5 nm[21] Pt|YIG Kapitza resistance Rth 3:5810 9m2K W 1[24] aThe conductivity of Pt is approximately inverse-linear in temperature over the regime we are considering. However, as we are not interested in detailed thermodynamic behavior, we use the fixed value = 1107S m 1throughout this work. TABLE I. System parameters for a Pt|YIG bilayer film. | 2018-10-05 | We develop a model for the magnonic contribution to the unidirectional spin
Hall magnetoresistance (USMR) of heavy metal/ferromagnetic insulator bilayer
films. We show that diffusive transport of Holstein-Primakoff magnons leads to
an accumulation of spin near the bilayer interface, giving rise to a
magnoresistance which is not invariant under inversion of the current
direction. Unlike the electronic contribution described by Zhang and Vignale
[Phys. Rev. B 94, 140411 (2016)], which requires an electrically conductive
ferromagnet, the magnonic contribution can occur in ferromagnetic insulators
such as yttrium iron garnet. We show that the magnonic USMR is, to leading
order, cubic in the spin Hall angle of the heavy metal, as opposed to the
linear relation found for the electronic contribution. We estimate that the
maximal magnonic USMR in Pt|YIG bilayers is on the order of $10^{-8}$, but may
reach values of up to $10^{-5}$ if the magnon gap is suppressed, and can thus
become comparable to the electronic contribution in, e.g., Pt|Co. We show that
the magnonic USMR at a finite magnon gap may be enhanced by an order of
magnitude if the magnon diffusion length is decreased to a specific optimal
value that depends on various system parameters. | Magnon contribution to unidirectional spin Hall magnetoresistance | 1810.02610v2 |
Magnons at low excitations: Observation of incoherent coupling to a bath of two-level systems Marco Prrmann,1,Isabella Boventer,1, 2Andre Schneider,1Tim Wolz,1Mathias Kl aui,2Alexey V. Ustinov,1, 3and Martin Weides1, 4,y 1Institute of Physics, Karlsruhe Institute of Technology, 76131 Karlsruhe, Germany 2Institute of Physics, Johannes Gutenberg-University Mainz, 55099 Mainz, Germany 3Russian Quantum Center, National University of Science and Technology MISIS, 119049 Moscow, Russia 4James Watt School of Engineering, University of Glasgow, Glasgow G12 8LT, United Kingdom (Dated: November 26, 2019) Collective magnetic excitation modes, magnons, can be coherently coupled to microwave photons in the single excitation limit. This allows for access to quantum properties of magnons and opens up a range of applications in quantum information processing, with the intrinsic magnon linewidth rep- resenting the coherence time of a quantum resonator. Our measurement system consists of a yttrium iron garnet sphere and a three-dimensional microwave cavity at temperatures and excitation powers typical for superconducting quantum circuit experiments. We perform spectroscopic measurements to determine the limiting factor of magnon coherence at these experimental conditions. Using the input-output formalism, we extract the magnon linewidth m. We attribute the limitations of the coherence time at lowest temperatures and excitation powers to incoherent losses into a bath of near-resonance two-level systems (TLSs), a generic loss mechanism known from superconducting circuits under these experimental conditions. We nd that the TLSs saturate when increasing the excitation power from quantum excitation to multiphoton excitation and their contribution to the linewidth vanishes. At higher temperatures, the TLSs saturate thermally and the magnon linewidth decreases as well. Strongly coupled light-spin hybrid systems allow for coherent exchange of quantum information. Such sys- tems are usually studied either classically at room tem- perature [1] or at millikelvin temperatures approaching the quantum limit of excitation [2{4]. The eld of cav- ity magnonics [5{9] harnesses the coherent exchange of excitation due to the strong coupling within the system and is used to access a new range of applications such as quantum transducers and memories [10]. Nonlinear- ity in the system is needed to gain access to the control and detection of single magnons. Because of experimen- tal constraints regarding required light intensities in a purely optomagnonic system [11], hybridized systems of magnon excitations and non-linear macroscopic quantum systems such as superconducting qubits [12, 13] are used instead, which opens up new possibilities in the emerging eld of quantum magnonics [14, 15]. An ecient interac- tion of magnonic systems and qubits requires their life- times to exceed the exchange time. Magnon excitation losses, expressed by the magnon linewidth m, translate into a lifetime of the spin excitation. Identifying its lim- iting factors is an important step toward more sophisti- cated implementations of hybrid quantum systems using magnons. Studies in literature show the losses in magnon excitations from room temperature down to about liquid helium temperatures [9]. The main contribution changes with temperature from scattering at rare-earth impu- rities [16, 17] to multi-magnon scattering at imperfect sample surfaces [18, 19]. For a typical environment of superconducting quantum circuit experiments, temper- atures below 100 mK and microwave probe powers com- parable to single-photon excitations, temperature sweepsshow losses into TLSs [3]. In this paper, we present both temperature- and power-dependent measurements of the magnon linewidth in a spherical yttrium iron gar- net (YIG) sample in the quantum limit of magnon ex- citations. We extract the critical saturation power and present on- and o-resonant linewidth that is mapped to the ratio of magnon excitation in the hybrid system. For large detuning, the fundamental linewidth can be extracted, thereby avoiding unwanted saturation eects from the residual cavity photon population. This renders the o-resonant linewidth a valuable information on the limiting factors of spin lifetimes. The magnetization dynamics inside a magnetic crys- tal is described by bosonic quasiparticles of collective spin excitation, called magnons. These magnons man- ifest as the collective precessional motion of the partic- ipating spins out of their equilibrium positions. Their energies and spatial distribution can be calculated ana- lytically using the Walker modes for spherical samples [20, 21]. We focus on the uniform in-phase precession mode corresponding to the wave vector k= 0, called the Kittel mode [22], treating it equivalently to one sin- gle large macro spin. The precession frequency (magnon frequency) of the Kittel mode in a sphere changes lin- early with a uniform external magnetic bias eld. The precessional motion is excited by a magnetic eld os- cillating at the magnon frequency perpendicular to the bias eld. We use the conned magnetic eld of a cav- ity photon resonance to create magnetic excitations in a macroscopic sample, biased by a static external magnetic eld. Tuning them into resonance, the magnon and pho- ton degree of freedom mix due to their strong interaction.arXiv:1903.03981v3 [quant-ph] 25 Nov 20192 185.5 186.25 187 187.75 188.5 Ext. magnetic field (mT) 5.205.225.245.265.28Probe frequency / (GHz) 5.20 5.22 5.24 5.26 5.28 Probe frequency / (GHz) 50 48 46 44 42 40 38 36 Reflection amplitude || (dB) / = . 1.8 2.0 2.2 2.4Coil current (A) a b Refl. amplitude || (arb. units) FIG. 1. (a) Color coded absolute value of the re
ection spectrum plotted against probe frequency and applied current at T= 55 mK and P= 140 dBm. The resonance dips show the dressed photon-magnon states forming an avoided level crossing with the degeneracy point at I0= 2:09 A, corresponding to an applied eld of B0= 186:98 mT (dashed vertical line). The inset displays the squared gradient of the zoomed-in amplitude data. The kink in the data represents a weakly coupled magnetostatic mode. This was also seen in Ref. [9]. (b) Raw data of the cross section at the center of the avoided level crossing and t to the input-output formalism. The t gives a magnon linewidth of m=2= 1:820:18 MHz. The data are normalized by the eld independent background before tting and is multiplied to the t to display it over the raw data. This creates hybridized states described as repulsive cav- ity magnon polaritons, which are visible as an avoided level crossing in the spectroscopic data with two reso- nance dips at frequencies !(see Supplemental Material [23]) appearing in the data cross section. The interaction is described by the macroscopic magnon-photon coupling strengthg. The system is probed in re
ection with mi- crowave frequencies using standard ferromagnetic reso- nance techniques [24]. We use the input-output formal- ism [25] to describe the re
ection spectrum. The complex re
ection parameter S11, the ratio of re
ected to input energy with respect to the probe frequency !p, reads as S11(!p) = 1 +2c i (!r !p) +l+g2 i(!m !p)+m;(1) with the cavity's coupling and loaded linewidths cand l, and the internal magnon linewidth m(HWHM). For our hybrid system we mount a commercially avail- able YIG (Y 3Fe5O12) sphere with a diameter d= 0:5 mm [26] inside a three-dimensional (3D) rectangular cavity made of oxygen-free copper and cool the device in a dilu- tion refrigerator down to millikelvin temperatures (see gure in Supplemental Material [23]). YIG as a ma- terial is particularly apt for microwave applications, as it is a ferrimagnetic insulator with a very low Gilbert damping factor of 10 3to 10 5[27{29] and a high net spin density of 2 :11022B=cm3[30]. The single crys- tal sphere comes pre-mounted to a beryllium oxide rod along the [110] crystal direction. The 3D cavity has a TE102mode resonance frequency of !bare r=2= 5:24 GHzand is equipped with one SMA connector for re
ection spectroscopy measurements. For low temperatures and excitation powers, we nd the internal and coupling quality factors to be Qi=!r=2i= 712597 and Qc=!r=2c= 543929, combining to a loaded quality factorQl= (1=Qi+ 1=Qc) 1= 308424 (see Supple- mental Material [23]). We mount the YIG at a magnetic anti-node of the cavity resonance and apply a static mag- netic eld of about 187 mT perpendicular to the cavity eld to tune the magnetic excitation into resonance with the cavity photon. The magnetic eld is created by an iron yoke holding a superconducting niobium-titanium coil. Additional permanent samarium-cobalt magnets are used to create a zero-current oset magnetic eld of about 178 mT. The probing microwave signal is provided by a vector network analyzer (VNA). Microwave attenu- ators and cable losses account for 75 dB of cable atten- uation to the sample. We apply probe powers between 140 dBm and 65 dBm at the sample's SMA port. To- gether with the cavity parameters, this corresponds to an average magnon population number hmifrom 0:3 up to the order of 107[23] in the hybridized case. The probe signal is coupled capacitively to the cavity photon using the bare inner conductor of a coaxial cable positioned in parallel to the electric eld component. The temperature of the sample is swept between 55 mK and 1 :8 K using a proportional-integral-derivative (PID) controlled heater. After a change in temperature, we wait at least one hour for the sample to thermalize before measuring. All data acquisition and analysis are done via qkit [31].3 0.1 1 Temperature (K) 0.81.01.21.41.61.82.02.2Magnon linewidth / (MHz) = = 140 120 100 80 60 Probe power (dBm) b a = = 1 10Temperature / (GHz) 110102103104105106107Average magnon number FIG. 2. (a) Temperature dependence of the magnon linewidth mat the degeneracy point. For low probe powers, m follows a tanh (1 =T) behavior (crosses), while for high probe powers (circles) the linewidth does not show any temperature dependence. (b) Power dependence of the magnon linewidth mforT= 55 mK and 200 mK at the degeneracy point. Both temperature curves show a similar behavior. At probe powers of about 90 dB mmdrops for both temperatures, following the (1 +P=P c) 1=2trend of the TLS model. All linewidth data shown here are extracted from the t at matching frequencies. A typical measurement is shown in Fig. 1(a), mea- sured atT= 55 mK with an input power level of P= 140 dBm. Figure 1(b) shows the raw data and the t of the cavity-magnon polariton at matching resonance fre- quencies for an applied external eld of B0= 186:98 mT. We correct the raw data from background resonances and extract the parameters of the hybridized system by t- ting to Eq. (1). The coupling strength g=2= 10:4 MHz of the system exceeds both the total resonator linewidth l=2=!r 2Ql=2= 0:85 MHz and the internal magnon linewidthm=2= 1:82 MHz, thus being well in the strong coupling regime ( gl; m) for all temperatures and probe powers. The measured coupling strength is in good agreement with the expected value gth=
e 2r 0~!r 2Vap 2Nss; (2) with the gyromagnetic ratio of the electron
e, the mode volumeVa= 5:40610 6m3, the Fe3+spin number s= 5=2, the spatial overlap between microwave eld and magnon eld , and the total number of spins Ns[9]. The overlap factor is given by the ratio of mode volumes in the cavity volume and the sample volume [1]. We nd for our setup the overlap factor to be = 0:536. For a sphere diameter of d= 0:5 mm we expect a total number ofNs= 1:371018spins. We nd the expected coupling strengthgth=2= 12:48 MHz to be in good agreement with our measured value. Even for measurements at high powers, the number of participating spins of the order of 1018is much larger than the estimated number of magnon excitations (107). We therefore do not expect to seethe intrinsic magnon nonlinearity as observed at excita- tion powers comparable to the number of participating spins [32]. The internal magnon linewidth decreases at higher temperature and powers (Fig. 2) while the coupling strength remains geometrically determined and does not change with either temperature or power. This behav- ior can be explained by an incoherent coupling to a bath of two-level systems (TLSs) as the main source of loss in our measurements. In the TLS model [33{36], a quantum state is conned in a double-well potential with dierent ground-state energies and a barrier in-between. TLSs be- come thermally saturated at temperatures higher than their frequency ( T&~!TLS=kB). Dynamics at low tem- peratures are dominated by quantum tunneling through the barrier that can be stimulated by excitations at sim- ilar energies. This resonant energy absorption shifts the equilibrium between the excitation rate and lifetime of the TLSs and their in
uence to the overall excitation loss vanishes. Loss into an ensemble of near-resonant TLSs is a widely known generic model for excitation losses in solids, glasses, and superconducting circuits at these ex- perimental conditions [37]. We t the magnon linewidth to the generic TLS model loss tangent m(T; P ) =0tanh ( ~!r=2kBT)p 1 +P=P c+o: (3) Directly in the avoided level crossing we nd 0=2= 1:050:15 MHz as the low temperature limit of the linewidth describing the TLS spectrum within the sam- ple ando=2= 0:910:11 MHz as an oset linewidth4 186.25 186.75 187.25 187.75 Ext. magnetic field (mT) 0.81.21.62.0Magnon linewidth (MHz) 186.25 186.75 187.25 187.75 Ext. magnetic field (mT) 01 Ratiophoton magnon 01 Ratio5.22 5.23 5.24 5.25 5.26Magnon frequency / (GHz) 5.22 5.23 5.24 5.25 5.26Magnon frequency / (GHz) a b = = FIG. 3. Magnetic eld (magnon frequency) dependence of the the magnon linewidth mfor dierent probe powers at T= 55 mK (a) and T= 200 mK (b). The shown probe powers correspond to the ones at the transition in Fig. 2 (b). The number of excited magnons depends on the detuning of magnon and photon frequency. At matching frequencies (dashed line) the magnon linewidth has a minimum, corresponding to the highest excited magnon numbers and therefore the highest saturation of TLSs. A second minimum at about 187 :25 mT corresponds to the coupling to an additional magnetostatic mode within the sample [inset of Fig. 1 (a)]. The insets show the ratio of excitation power within each component of the hybrid system. At matching frequencies, both components are excited equally. The magnon share drops at the plot boundaries to about 20 %. The coupling to the magnetostatic mode is visible as a local maximum in the magnon excitation ratio. The xaxes are scaled as in the main plots. The legends are valid for both temperatures. added as a lower boundary without TLS contribution. The critical power Pc= 816:5 dBm at the SMA port describes the saturation of the TLS due to res- onant power absorption, corresponding to an average critical magnon number of hmci= 2:4105. Using nite-element simulations, we map the critical excita- tion power to a critical AC magnetic eld on the or- der ofBc310 10T at the position of the YIG sam- ple. Looking at the linewidths outside the anti-crossing at constant input power, we nd a minimum at match- ing magnon and photon frequencies (dashed lines in Fig. 3). Here, the excitation is equally distributed between photons and magnons, reaching the maximum in both magnon excitation power and TLS saturation, respec- tively. At detuned frequencies the ratio between magnon and photon excitation power changes, less energy excites the magnons (insets in Fig. 3), and therefore less TLSs get saturated. The magnon linewidth increases with de- tuning, matching the low power data for large detunings. This eect is most visible at highest excitation powers. We calculate the energy ratios by tting the resonances in each polariton branch individually and weight the stored energy with the eigenvalues of the coupling Hamiltonian [38]. For higher powers a second minimum at about 187:25 mT can be seen at both studied temperatures. We attribute this to the coupling to a magnetostatic mode within the YIG sample and therefore again an increased number of excited magnons [see inset of Fig. 1 (a)]. Thiscan also be seen in the inset gures as a local magnon excitation maximum. We attribute the TLS-independent losseso=2= 0:910:11 MHz to multi-magnon scat- tering processes on the imperfect sphere surface [18, 19]. As described in Ref. [9], we model the surface of the YIG with spherical pits with radii of2 3of the size of the polish- ing material (2 =30:05µm) and estimate a contribution of about 2 1 MHz that matches our data. We attribute the slight increase in the linewidth visible in the high- power data (circles) in Fig. 2 (a) to the rst in
uence of rare-earth impurity scattering, dominating the linewidth behavior of the TLS-saturated system at higher temper- atures [9, 16, 39]. In principle, loss due to TLS can also be determined indirectly by weak changes of the reso- nance frequency [40{43] while keeping the eld constant. Our system, however, operates at xed frequency and magnetic remanence within the magnetic yoke leads to uncertainties in absolute magnetic eld value beyond the required accuracy. In this work, we studied losses in a spherical YIG sample at temperatures below 2 K and excitation powers down from 107photons below a single photon. We iden- tify incoherent coupling to a bath of two-level systems as the main source of excitation loss in our measurements. The magnon linewidth m=2at the degeneracy point ts well to the generic loss tangent of the TLS model with re- spect to temperature and power. It decreases from about 1:8 MHz in
uenced by TLSs to about 1 MHz with satu-5 rated TLSs. The magnon linewidth shows a minimum at maximum magnon excitation numbers, again corre- sponding with TLS saturation with increasing excitation power. While TLSs are a common source of loss in super- conducting circuits, their microscopic nature is still not fully understood. Possible models for TLS origin include magnetic TLSs in spin glasses [44{47] that manifest in crystalline samples in lower concentration, surface spins [48, 49] that in
uence the eective number of spins or magnon-phonon and subsequent phonon losses into TLSs [50] (see also Supplemental Material [23]. Improving the surface roughness and quality of the YIG crystal can lead to lower overall losses and lower TLS in
uence which can lead to longer coherence lifetimes for application in quan- tum magnonic devices. Note added in proof - Recently, a manuscript study- ing losses in thin lm YIG that independently observed comparable results and reached similar conclusions was published by Kosen et al. [51]. This work was supported by the European Research Council (ERC) under the Grant Agreement 648011 and the Deutsche Forschungsgemeinschaft (DFG) within Project INST 121384/138-1 FUGG and SFB TRR 173. We acknowledge nancial support by the Helmholtz In- ternational Research School for Teratronics (M.P. and T.W.) and the Carl-Zeiss-Foundation (A.S.). A.V.U ac- knowledges partial support from the Ministry of Educa- tion and Science of Russian Federation in the framework of the Increase Competitiveness Program of the National University of Science and Technology MISIS (Grant No. K2-2017-081). marco.prrmann@kit.edu ymartin.weides@glasgow.ac.uk [1] X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, Phys. Rev. Lett. 113, 156401 (2014). [2] H. Huebl, C. W. Zollitsch, J. Lotze, F. Hocke, M. Greifen- stein, A. Marx, R. Gross, and S. T. B. Goennenwein, Phys. Rev. Lett. 111, 127003 (2013). [3] Y. Tabuchi, S. Ishino, T. Ishikawa, R. Yamazaki, K. Us- ami, and Y. Nakamura, Phys. Rev. Lett. 113, 083603 (2014). [4] M. Goryachev, W. G. Farr, D. L. Creedon, Y. Fan, M. Kostylev, and M. E. Tobar, Phys. Rev. Appl. 2, 054002 (2014). [5] D. Zhang, X.-M. Wang, T.-F. Li, X.-Q. Luo, W. Wu, F. Nori, and J. You, npj Quantum Inf. 1, 15014 EP (2015). [6] L. Bai, M. Harder, Y. Chen, X. Fan, J. Xiao, and C.-M. Hu, Phys. Rev. Lett. 114, 227201 (2015). [7] Y. Cao, P. Yan, H. Huebl, S. T. B. Goennenwein, and G. E. W. Bauer, Phys. Rev. B 91, 094423 (2015). [8] M. Goryachev, S. Watt, J. Bourhill, M. Kostylev, and M. E. Tobar, Phys. Rev. B 97, 155129 (2018). [9] I. Boventer, M. Prrmann, J. Krause, Y. Sch on, M. Kl aui, and M. Weides, Phys. Rev. B 97, 184420(2018). [10] X. Zhang, C.-L. Zou, N. Zhu, F. Marquardt, L. Jiang, and H. X. Tang, Nat. Commun. 6, 8914 EP (2015). [11] S. Viola Kusminskiy, H. X. Tang, and F. Marquardt, Phys. Rev. A 94, 033821 (2016). [12] Y. Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Ya- mazaki, K. Usami, and Y. Nakamura, Science 349, 405 (2015). [13] Y. Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Ya- mazaki, K. Usami, and Y. Nakamura, C. R. Phys. 17, 729 (2016). [14] D. Lachance-Quirion, Y. Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Yamazaki, and Y. Nakamura, Sci. Adv. 3, e1603150 (2017). [15] D. Lachance-Quirion, Y. Tabuchi, A. Gloppe, K. Us- ami, and Y. Nakamura, Appl. Phys. Express 12, 070101 (2019). [16] E. G. Spencer, R. C. LeCraw, and R. C. Linares, Phys. Rev.123, 1937 (1961). [17] P. E. Seiden, Phys. Rev. 133, A728 (1964). [18] M. Sparks, R. Loudon, and C. Kittel, Phys. Rev. 122, 791 (1961). [19] J. Nemarich, Phys. Rev. 136, A1657 (1964). [20] L. R. Walker, Phys. Rev. 105, 390 (1957). [21] P. Fletcher, I. H. Solt, and R. Bell, Phys. Rev. 114, 739 (1959). [22] C. Kittel, Phys. Rev. 73, 155 (1948). [23] See Supplemental Material at http://link.aps.org/ supplemental/10.1103/PhysRevResearch.1.032023 for further information about tting procedure and excita- tion number callibration.l. [24] S. S. Kalarickal, P. Krivosik, M. Wu, C. E. Patton, M. L. Schneider, P. Kabos, T. J. Silva, and J. P. Nibarger, J. Appl. Phys. 99, 093909 (2006). [25] D. F. Walls and G. F. Milburn, Quantum Optics (Springer-Verlag GmbH, 2008). [26] Ferrisphere Inc., www.ferrisphere.com. [27] Y. Kajiwara, K. Harii, S. Takahashi, J. Ohe, K. Uchida, M. Mizuguchi, H. Umezawa, H. Kawai, K. Ando, K. Takanashi, S. Maekawa, and E. Saitoh, Nature 464, 262 (2010). [28] B. Heinrich, C. Burrowes, E. Montoya, B. Kardasz, E. Girt, Y.-Y. Song, Y. Sun, and M. Wu, Phys. Rev. Lett. 107, 066604 (2011). [29] H. Kurebayashi, O. Dzyapko, V. E. Demidov, D. Fang, A. J. Ferguson, and S. O. Demokritov, Nature Materials 10, 660 EP (2011). [30] M. A. Gilleo and S. Geller, Phys. Rev. 110, 73 (1958). [31] \Qkit - a quantum measurement suite in python," https://github.com/qkitgroup/qkit. [32] J. A. Haigh, N. J. Lambert, A. C. Doherty, and A. J. Ferguson, Phys. Rev. B 91, 104410 (2015). [33] P. W. Anderson, B. I. Halperin, and C. M. Varma, Phi- los. Mag. 25, 1 (1972). [34] W. A. Phillips, J. Low Temp. Phys. 7, 351 (1972). [35] S. Hunklinger and W. Arnold, Ultrasonic Properties of Glasses at Low Temperatures , Physical Acoustics, Vol. 12 (Academic Press, 1976) pp. 155 { 215. [36] W. A. Phillips, Reports on Progress in Physics 50, 1657 (1987). [37] C. M uller, J. H. Cole, and J. Lisenfeld, Reports on Progress in Physics 82, 124501 (2019). [38] M. Harder and C.-M. Hu, Cavity Spintronics: An Early Review of Recent Progress in the Study of Magnon-6 Photon Level Repulsion , edited by R. E. Camley and R. L. Stamps, Solid State Physics, Vol. 69 (Academic Press, 2018) pp. 47 { 121. [39] E. G. Spencer, R. C. LeCraw, and A. M. Clogston, Phys. Rev. Lett. 3, 32 (1959). [40] D. P. Pappas, M. R. Vissers, D. S. Wisbey, J. S. Kline, and J. Gao, Appl. Phys. Lett. 21, 871 (2011). [41] J. Gao, M. Daal, A. Vayonakis, S. Kumar, J. Zmuidzinas, B. Sadoulet, B. A. Mazin, P. K. Day, and H. G. Leduc, Appl. Phys. Lett. 92, 152505 (2008). [42] S. Kumar, J. Gao, J. Zmuidzinas, B. A. Mazin, H. G. LeDuc, and P. K. Day, Appl. Phys. Lett. 92, 123503 (2008). [43] J. Burnett, L. Faoro, I. Wisby, V. L. Gurtovoi, A. V. Chernykh, G. M. Mikhailov, V. A. Tulin, R. Shaikhaidarov, V. Antonov, P. J. Meeson, A. Y. Tza- lenchuk, and T. Lindstr om, Nat. Commun. 5, 4119 EP (2014). [44] M. Continentino, Solid State Communications 38, 981 (1981). [45] M. A. Continentino, Journal of Physics C: Solid State Physics 14, 3527 (1981). [46] E. F. Wassermann and D. M. Herlach, Journal of Applied Physics 55, 1709 (1984). [47] D. Wesenberg, T. Liu, D. Balzar, M. Wu, and B. L. Zink, Nature Physics 13, 987 EP (2017). [48] S. E. de Graaf, A. A. Adamyan, T. Lindstr om, D. Erts, S. E. Kubatkin, A. Y. Tzalenchuk, and A. V. Danilov, Phys. Rev. Lett. 118, 057703 (2017). [49] S. E. de Graaf, L. Faoro, J. Burnett, A. A. Adamyan, A. Y. Tzalenchuk, S. E. Kubatkin, T. Lindstr om, and A. V. Danilov, Nature Communications 9, 1143 (2018). [50] S. Streib, N. Vidal-Silva, K. Shen, and G. E. W. Bauer, Phys. Rev. B 99, 184442 (2019). [51] S. Kosen, A. F. van Loo, D. A. Bozhko, L. Mihalceanu, and A. D. Karenowska, APL Materials 7, 101120 (2019). [52] S. Probst, F. B. Song, P. A. Bushev, A. V. Ustinov, and M. Weides, Review of Scientic Instruments 86, 024706 (2015). [53] M. Newville, T. Stensitzki, D. B. Allen, and A. Ingar- giola, \LMFIT: Non-Linear Least-Square Minimization and Curve-Fitting for Python," (2014). [54] M. A. Continentino, Phys. Rev. B 27, 4351 (1983). [55] M. A. Continentino, Journal of Physics C: Solid State Physics 16, L71 (1983). [56] C. Arzoumanian, A. De Go er, B. Salce, and F. Holtzberg, J. Physique Lett. 44, 39 (1983). [57] R. N. Kleiman, G. Agnolet, and D. J. Bishop, Phys. Rev. Lett. 59, 2079 (1987). Cavity-Magnon coupling The frequencies of both arms of the avoided level cross- ing!are tted to the energy eigenvalues of a 2 2 matrix describing two coupled harmonic oscillators, one with constant frequency and one with a linearly changing frequency, !=!bare r+!I=0 m 2s!barer !I=0m 22 +g2:(S1)We use the current dependent data taken at T= 55 mK andP= 140 dBm to obtain the bare cavity fre- quency!bare r, the zero-current magnetic excitation fre- quency!I=0 m, and the coupling strength g. The fre- quencies of the anticrossing !were obtained by track- ing the minima in the amplitude data. From the t we obtain the bare cavity frequency !bare r=2= 5:239 020:000 02 GHz and the zero-current magnetic excitation frequency !I=0 m=2= 4:98170:0002 GHz due to the oset magnetic eld by the permanent mag- nets. The magnon-cavity coupling strength stays nearly constant for all temperatures and excitation powers at g=2= 10:390:17 MHz. Magnon number estimation Using the cavity's resonance frequency, quality factors, and the input power Pinwe estimate the total number of magnon and photon excitation within the cavity hNeiin units of ~!r, hNei= 4Q2 l Qc1 ~!2rPin: (S2) Note that the input power Pinis in units of watts and not to be confused with the probe power (level) Pin units of dBm. For the strongly coupled system the exci- tation energy at matching frequencies is stored in equal parts in photons and magnons, hni=hmi=hNei 2. We measure the re
ection signal of the cavity reso- nance at 55 mK at probe powers between 140 dBm and 65 dBm at zero current applied to the magnetic coil. The complex data is then tted using a circle t al- gorithm [52] to determine the power dependent qual- ity factors (coupling quality factor Qc=!r 2c, inter- nal quality factor Qi=!r 2i, and loaded quality factor Ql= (1=Qi+ 1=Qc) 1) and resonance frequencies as shown in Fig. S1 (a-d). Besides an initial shift of the quality factors of less than 1 % going from 140 dBm to 130 dBm of input power the quality factors show no power dependence, varying only in a range below 0 :15 %. The tted (dressed) frequencies are shifted compared to the bare cavity frequency due to the residual magnetic eld. From Eq. (S1) we expect a zero-current dressed frequency of !I=0 r= 5:239 4520:000 002 GHz. The cir- cle t gives a resonance frequency at the lowest power of !CF r= 5:239 4740:000 002 GHz. We calculate the av- erage total excitation for all probe powers using Eq. (S2) and t a line to the logarithmic data (Fig. S1 (e)), hNei= 62:046P1:0003 in fW 1: (S3) The t agrees well with the data and is used throughout all data evaluation to map probe powers Pto average excitation numbers. This results in average magnon ex- citation numbers at matching frequencies for our experi- ments between 0 :31 and 9:85106.7 140 120 100 80 60 Probe power (dBm) 110102103104105106107108Average total excitation number 1 0123/ (kHz) 140 120 100 80 60 Probe power (dBm) 70807120716072007240Internal Q factor 307830823086309030943098Loaded Q factor 542054255430543554405445Coupling Q factor a b cd e FIG. S1. (a-c) Loaded, coupling, and internal quality factors of the cavity resonance against probe power. The data was taken atT= 55 mK with zero current applied to the magnetic coil and does not show a power dependent behavior. (d) Shift of the tted cavity frequencies !=2= [!r(P) !r(P= 140 dB m)]=2with compared to the measurement at lowest probe power at zero current. Similar as with the quality factors, the cavity frequency does not show a power dependence. (e) Calculated average photon number in cavity against probe power. The t shows a linear dependence of the photon number calculated with Eq. (S2) to the input power. Note that this plot features a log-log scale, making the t linear again. The errors on the average photon number are estimated to be smaller than 0 :35 % and are not visible on this plot. Extracting the internal magnon linewidth We extract the internal magnon linewidth mby tting the re
ection amplitude jS11(!)jusing the input-output formalism [25]. jS11(!p;I)j= 1 +2c i (!r !p) +l+g2 i(!m(I) !p)+m;(S4) with the probe frequency !p, the magnon frequency !m, and the loaded, coupling and magnon linewidths l,c, andm(HWHM). Before tting, we normalize the databy the current independent baseline similar to Ref. [9]. We estimate the background value for each probe fre- quency by calculating a weighted average over all entries along the current axis, neglecting the areas around the dressed cavity resonances. The amplitude data is divided by this baseline to account for losses in the measurement setup. The normalized data together with the t results of Eq. (S1) and the circle tted cavity resonance at zero current are then tted to Eq. (S4) using the Python pack- age lmt [53].8 VNA3K 0.75K 55mK -20dB -20dB -20dB40dB32dBb aYIG coupler 3 mm FIG. S2. (a) Photograph of the sample in the cavity. The top half of the cavity resonator was removed and can be seen in the background. (b) Schematic diagram of the experimen- tal setup. The cavity holding the YIG sphere and the magnet providing the static eld are mounted at the mixing cham- ber plate of a dry dilution refrigerator. The microwave input signal is attenuated to minimize thermal noise at the sample. The attenuation of the complete input line to the input port of the cavity is 75 dB at the cavity resonance frequency. The output signal is amplied by a cryogenic amplier operating at 3 K and an amplier at room temperature. Two magneti- cally shielded microwave circulators protect the sample from amplier noise. Possible TLS origin The microscopic origins of TLSs is still unclear and part of ongoing research. Possible models include mag- netic TLSs proposed with analog behavior to the electricdipolar coupled TLSs [44, 45, 54, 55] and measured in spin glasses by thermal conductivity, susceptibility and magnetization measuements at low temperatures [46, 56]. With amorphous YIG showing spin glass behavior [47] it seems plausible to observe these eects in our crystalline YIG sample where in addition to the observed rare earth impurities [9] we can assume structural crystal defects. This is based on materials with electric dipolar coupled TLSs, where TLSs appear largely in disorderd crystals but also in single crystals with smaller density [57]. Another possibility could be surface spins leading to strong damping that were observed as an important loss mechanism in cQED experiments [48, 49]. We evaluated the coupling strength to nd a power or temperature de- pendence on the participating number of spins, see Fig. S3. We nd an increase in the coupling strength of about 1 % at the saturation conditions for the TLSs. With g/p Nthis translates to an increase in the number of participating spins of the order of 2 %, e.g. due to the in- creased participation of now environmentally decoupled surface spins. This should not be enough to explain the decrease in mby a factor of 2. A loss mechanism by magnon-phonon coupling and subsequent phonon losses due to TLS coupling can be neglected since for k= 0 magnons in YIG these magnon losses are proposed to be much smaller than the Gilbert damping [50].9 0.1 1 Temperature (K) 10.2510.3010.3510.4010.4510.50coupling strength / (MHz) = = 140 120 100 80 60 probe power (dBm) = = FIG. S3. (a) Temperature and (b) power dependence of the coupling strength evaluated at the same conditions as Fig. (2) in the main text. We nd a increase of the coupling strength of about 1 % going to higher powers that decreases at higher temperatures. This indicates an increase in participating spins on the order of 2 %. | 2019-03-10 | Collective magnetic excitation modes, magnons, can be coherently coupled to
microwave photons in the single excitation limit. This allows for access to
quantum properties of magnons and opens up a range of applications in quantum
information processing, with the intrinsic magnon linewidth representing the
coherence time of a quantum resonator. Our measurement system consists of a
yttrium iron garnet (YIG) sphere and a three-dimensional (3D) microwave cavity
at temperatures and excitation powers typical for superconducting quantum
circuit experiments. We perform spectroscopic measurements to determine the
limiting factor of magnon coherence at these experimental conditions. Using the
input-output formalism, we extract the magnon linewidth $\kappa_\mathrm{m}$. We
attribute the limitations of the coherence time at lowest temperatures and
excitation powers to incoherent losses into a bath of near-resonance two-level
systems (TLSs), a generic loss mechanism known from superconducting circuits
under these experimental conditions. We find that the TLSs saturate when
increasing the excitation power from quantum excitation to multi-photon
excitation and their contribution to the linewidth vanishes. At higher
temperatures, the TLSs saturate thermally and the magnon linewidth decreases as
well. | Magnons at low excitations: Observation of incoherent coupling to a bath of two-level-systems | 1903.03981v3 |
Control of spin current by a magnetic YIG substrate in NiFe/Al nonlocal spin valves F. K. Dejene,1,N. Vlietstra,1D. Luc,2X. Waintal,2J. Ben Youssef,3and B. J. van Wees1 1Physics of Nanodevices, Zernike Institute for Advanced Materials, University of Groningen, 9747AG, Groningen, The Netherlands 2CEA-INAC/UJF Grenoble 1, SPSMS UMR-E 9001, Grenoble F-38054, France 3Universit de Bretagne Occidentale, Laboratoire de Magnetisme de Bretagne CNRS, 6 Avenue Le Gorgeu, 29285 Brest, France (Dated: June 9, 2021) We study the eect of a magnetic insulator (Yttrium Iron Garnet - YIG ) substrate on the spin transport properties of Ni 80Fe20/Al nonlocal spin valve (NLSV) devices. The NLSV signal on the YIG substrate is about 2 to 3 times lower than that on a non magnetic SiO 2substrate, indicating that a signicant fraction of the spin-current is absorbed at the Al/YIG interface. By measuring the NLSV signal for varying injector-to-detector distance and using a three dimensional spin-transport model that takes spin current absorption at the Al/YIG interface into account we obtain an eective spin-mixing conductance G"#'5 81013 1m 2. We also observe a small but clear modulation of the NLSV signal when rotating the YIG magnetization direction with respect to the xed spin polarization of the spin accumulation in the Al. Spin relaxation due to thermal magnons or roughness of the YIG surface may be responsible for the observed small modulation of the NLSV signal. The coupled transport of spin, charge and heat in non- magnetic (N) metals deposited on the magnetic insulator Y3Fe5O12(YIG) has led to new spin caloritronic device concepts such as thermally driven spin currents, the gen- eration of spin angular momentum via the spin Seebeck eect (SSE) [1], spin pumping from YIG to metals [2], spin-orbit coupling (SOC) induced magnetoresistance ef- fects [3, 4] and the spin Peltier eect, i.e., the inverse of the SSE that describes cooling/heating by spin cur- rents [5]. In these spin caloritronic phenomena, the spin- mixing conductance G"#of the N/YIG interface controls the transfer of spins from the conduction electrons in N to the magnetic excitations (magnons) in the YIG, or vice versa [6{10]. The interconversion of spin current to a voltage employs the (inverse) spin Hall eect in heavy- metals such as Pt or Pd. The possible presence of prox- imity induced magnetism in these metals is reported to introduce spurious magnetothermoelectric eects [11, 12] or enhance G"#[7]. Owing to the short spin-diusion lengthin these large SOC metals, the applicability of the diusive spin-transport model is also questionable. Experimental measurements that alleviate these concerns are however scarce and hence are highly required. In this article, we investigate the interaction of spin current (in the absence of a charge current) with the YIG magnetization using the NLSV geometry [13{15]. Using a metal with low SOC and long spin-diusion length allows to treat our experiment using the diu- sive spin-transport model. We nd that the NLSV signal on the YIG substrate is two to three times lower than that on the SiO 2substrate, indicating signicant spin- current absorption at the Al/YIG interface. By vary- ing the angle between the induced spin accumulation and the YIG magnetization direction we observe a small but clear modulation of the NLSV signal. We also nd that modifying the quality of the Al/YIG interface, us-ing dierent thin-lm deposition methods [4], in
uences G"#and hence the size of the spin current
owing at the Al/YIG interface. Recently, a low-temperature mea- surements of a similar eect was reported by Villamor et al.[16] in Co/Cu devices where G"#1011 1m 2was estimated, two orders of magnitude lower than in the lit- erature [4, 8]. Here, we present a room-temperature spin- transport study in transparent Ni 80Fe20(Py)/Al NLSV devices. Figure 1 depicts the concept of our experiment. A non- magnetic metal (green) deposited on the YIG connects the two in-plane polarized ferromagnetic metals F1and F2, which are used for injecting and detecting spin cur- rents, respectively. A charge current through the F1/Al interface induces a spin accumulation s(~ r) = (0;s;0)T that is polarized along the ^ ydirection, parallel to the magnetization direction of F1. This non-equilibrium s, the dierence between the electrochemical potentials for spin up and spin down electrons, diuses to both +^ x and ^xdirections of F1/Al interface with an exponen- tial decay characterized by the spin diusion length N. Spins arriving at the detecting F2/Al interface give rise to a nonlocal voltage Vnlthat is a function of the rela- tive magnetic conguration of F1andF2, being minimum (maximum) when F1andF2are parallel (antiparallel) to each other. For NLSV devices on a SiO 2substrate, spin relaxation proceeds via electron scattering with phonons, impuri- ties or defects present in the spin transport channel, also known as the Elliot-Yafet (EY) mechanism. The situa- tion is dierent for a NLSV on the magnetic YIG sub- strate where additional spin relaxation due to thermal magnons in the YIG and/or interfacial spin orbit cou- pling can be mediated by direct spin-
ip scattering or spin-precession. Depending on the magnetization direc- tion ^mof the YIG with respect to sspins incident atarXiv:1503.06108v1 [cond-mat.mes-hall] 20 Mar 20152 on SiO 2 on YIG sµ 0ˆmAl YIGspin waves ˆm FIG. 1. (Color online) Concept of the experiment for ^ mks. (a) A charge current through the F 1/Al interface creates a spin accumulation sin the Al. The diusion of sto the F 2/Al interface is aected by spin-
ip relaxation at the Al/YIG interface. Scattering of a spin up electron ( s=~=2) into spin down electron ( s= ~=2) is accompanied by magnon emission (s=~) creating a spin current that is minimum (max- imum) when ^ sis parallel (perpendicular) to the magnetiza- tion of the YIG. (b) Prole of salong the Al strip on a SiO 2 (red) and YIG (blue) substrate. The spin accumulation at the F 2/Al is lower for the YIG substrate compared to that on SiO2. the Al/YIG surface are absorbed ( ^ m?s) or re
ected ( ^mks) thereby causing a spin current density js(~ r) through the Al/YIG interface [9] js( ^m)jz=0=Gr^m( ^ms)+Gi( ^ms)+Gss:(1) Here ^m= (mx;my;0)Tis a unit vector parallel to the in-plane magnetization of the YIG, Gr(Gi) is the real (imaginary) part of the spin-mixing conductance per unit area andGsis a spin-sink conductance that can be in- terpreted as an eective spin-mixing conductance that quanties spin-absorption (
ip) eects that is indepen- dent of the angle between ^ mands. When ^mkssome of the spins incident on the YIG are re
ected back into the Al while some fraction is ab- sorbed by the YIG. The absorption of the spin-current in this collinear case is governed by a spin-sinking eect either due to (i) the thermal excitation of the YIG mag- netization (thermal magnons) or (ii) spin-
ip processes due to interface spin orbit eects or magnetic impurities present at the interface. This process can be character- ized by an eective spin-mixing interface conductance Gs which, at room temperature, is about 20% of Gr[5]. Be- cause of this additional spin-
ip scattering, the maximum NLSV signal on the YIG substrate should also be smaller than that on the SiO 2. When ^m?sspins arriving at the Al/YIG interface are absorbed. In this case all threeterms in Eq. (1) contribute to a maximum
ow of spin current through the interface. The nonlocal voltage mea- sured at F 2is hence a function of the angle between ^ m andsand should re
ect the symmetry of Eq. 1. Fig. 2(a) shows the scanning electron microscope im- age of the studied NLSV device that was prepared on a 200-nm thick single-crystal YIG, having very low coercive eld [2, 4, 17], grown by liquid phase epitaxy on a 500 m thick (111) Gd 3Ga5O12(GGG) substrate. It consists of two 20-nm thick Ni 80Fe20(Py) wires connected by a 130-nm thick Al cross. A 5 nm-thick Ti buer layer was inserted underneath the Py to suppress direct exchange coupling between the Py and YIG. We studied two types of devices, hereafter named Type-A and Type-B devices. In Type-A devices (4 devices), prior to the deposition of the Al (by electron beam evaporation), Ar ion milling of the Py surface was performed to ensure a transparent Py/Al interface. This process, however, introduces un- avoidable milling of the YIG surface thereby introducing disordered Al/YIG interface with lower G"#[18]. To cir- cumvent this problem, in Type-B devices (2 devices), we rst deposit a 20 nm-thick Al strip (by DC sputtering) between the injector and detector Py wires. Sputtering is reported to yield a better interface [4]. Next, after Ar ion milling of the Py and sputtered-Al surfaces, a 130 nm-thick Al layer was deposited using e-beam evapora- tion. Similar devices prepared on SiO 2substrate were also investigated. All measurements were performed at room temperature using standard low frequency lock-in measurements. The NLSV resistance Rnl=Vnl=Ias a function of the applied in-plane magnetic eld (along ^ y) is shown in Fig. 2(b), both for SiO 2(red and orange) and YIG (blue) samples. Note that the magnetizations of the injector, detector and YIG are all collinear and hence no initial transverse spin component is present. The spin valve sig- nal, dened as the dierence between the parallel RPand anti-parallel RAPresistance values, RSV=RP RAP on the YIG substrate is about two to three times smaller than that on the SiO 2substrate. This reduction in the NLSV signal indicates the presence of an additional spin- relaxation process even for ^ mks. Assuming an iden- tical spin injection eciency in both devices, this means that spin relaxation in the Al on the YIG substrate occurs on an eectively shorter spin relaxation length N. To properly extract Nwe performed several measurements for varying distance between the Py wires, as shown in Figure 2(c) both on SiO 2(red diamond) and YIG (blue square) substrates. Also shown are dashed-line ts us- ing the expression for the nonlocal spin valve signal RSV obtained from a one-dimensional spin transport theory given by [14] RSV=2 FRNe d=2N (RF RN+ 1)[RF RNsinh(d=2N) + cosh(d=2N)]:(2)3 (a) AlYIG xy1 23 4 V(b) (c) 300 nm FIG. 2. (Color online) (a) Scanning electron microscopy image of the measured Type-A device. Two Py wires (indicated by green arrows) are connected by an Al cross. A charge current Ifrom contact 1 to 2 creates a spin accumulation at the F1/Al interface that is detected as a nonlocal spin voltage Vnlusing contacts 3 and 4. (b) The NLSV resistance Rnl=Vnl=I for representative YIG (blue) and SiO 2(red and orange) NLSV samples. For comparison, a constant background resistance has been subtracted from each measurement. (c) Dependence of the NLSV signal on the spacing dbetween the injecting and detecting ferromagnetic wires together with calculated spin signal values using a 1D (dashed lines) and 3D (solid lines) spin-transport model. For each distance dbetween the injector and detector several devices were measured, with the error bars indicating the spread in the measured signal. HereRF= (1 2 F)F FandRN=N Nare spin area re- sistance of the ferromagnetic (F) and non-magnetic (N) metals, respectively. NandFare the corresponding spin diusion lengths, F(N) is the electrical conduc- tivity of the F (N), Fis the spin polarization of Fandd is the distance between the injecting and detecting ferro- magnetic electrodes. Fitting the SiO 2data using Eq. (2), we extract F=0.32 and N;SiO 2=320 nm, which are both in good agreement with reported values [13{15]. A similar tting procedure for the YIG data, assuming an identical spin injection eciency, yields an eectively shorter spin-diusion length N;YIG=190 nm due to the additional spin-
ip scattering at the Al/YIG interface. This value of N;YIGtherefore contains important infor- mation regarding an eective spin-mixing conductance Gsthat can be attributed to the interaction of spins with thermal magnons in the YIG. When spin precession, due to the applied external eld as well as the eective eld due toGiis disregarded, we can now estimate Gsby relatingN;YIGtoN;SiO 2viaGsas (see Supplemental Material [19], Sec. I): 1 2 N;YIG=1 2 N;SiO 2+1 2r; (3) with 2 r= 2Gs=tAlN[19]. Using the extracted values from the t, N=2107S/m andtAl=130 nm, we extract Gs'2:51013 1m 2, which is about 25% of the maximumGr1014 1m 2reported for Pt/YIG [4, 7] and Au/YIG [8] interfaces. To quantify our results we performed three- dimensional nite element simulations using COMSOL Multiphysics (3D-FEM) [19, 20] that uses a set ofequations that are equivalent to the continuous random matrix theory in 3 dimensions (CRMT3D) [21]. The charge current j c(~ r) and spin current j s(~ r), (where 2x;y;z ), are linked to their corresponding driving forces via the electrical conductivity as j c(~ r) j s(~ r) = F F ~rc ~rs (4) wherec= ("+#)=2 ands= (" #)=2 are the charge and spin accumulation chemical potentials, re- spectively. We supplement Eq. (4) by the conservation laws for charge (rj c(~ r) = 0) and spin current ( rjs= (1 2 F) s=2+~ !Ls ) where~ !L=gB~B=~with g= 2 is the Larmor precession frequency due to spin precession in an in-plane magnetic eld ~B= (Bx;By;0)T andBis the Bohr magneton (see Supplemental Mate- rial [19], Sec. II). To include spin-mixing at the Al/YIG interface we impose continuity of the spin current jsat the interface using Eq. (1). The input material param- eters such as ,andFare taken from Refs. 22 and 23. The calculated spin signals obtained from our 3D-FEM are shown in Fig. 2(c) for samples on SiO 2(red solid line) and YIG (blue solid line) substrates. By matching the experimentally measured NLSV signal on the SiO 2sub- strate with the calculated values in the model we obtain F= 0:3 andN=350 nm. Using these two values and settingGs'51013 1m 2well reproduces the mea- sured spin signal on the YIG substrate. This value of Gs obtained here is consistent with that extracted from our 1D analysis based on Eq. 2. Hence, the interaction of spins with the YIG magnetization, as modeled here, can4 α ˆysµˆm SV S FIG. 3. (Color online) (a) Nonlocal spin valve resistance R nlof a Type-B device with d=500 nm between injecting and detecting Py wires and tAl=130 nm. A constant background resistance of 117 m was subtracted from the original data. (b) Angular dependence of the NLSV signal in the parallel and antiparallel congurations. The AP curve is average of 10 measurements and that of the P state is a single scan. Both resistance states exhibit a cos(2 ) dependence on the angle between ^ mands. The black solid lines are calculated using the 3D-FEM model for Gr= 11013 1m 2that show a percentage modulation of only 12% corresponding to the green curve in (c) RSV=RSVis plotted. The angular dependent measurement in (b) is from a device for which complete set of measruements were peformed. A spin valves measurement as in (a) was also performed for another device with d= 300 nm. capture the concept of spin-mixing conductance being responsible for the observed reduction in the spin signal. In the following we investigate the dependence of Rnl on the angle betweensand ^m. We rotate the sam- ple under the application of a very low in-plane mag- netic eld B5 mT, enough to saturate the low-coercive (0:5 mT) YIG magnetization [4, 5] but smaller than the coercive elds of F1andF2(20 mT). This con- dition is important to maintain xed polarization axes ofs, along the magnetization direction of the injecting ferromagnet, and also have a well dened . The re- sult of such measurement in a Type-B device is shown in Fig. 3(b) for d= 400 nm between F 1and F 2. Al- though the measured NLSV signal [Fig.3(a)] is smaller than in Type-A devices, possibly due to a better Al/YIG interface,Rnlexhibits a cos(2 ) behavior with a maxi- mum (minimum) for = 0 (==2), consistent with Eq. (1). However, the maximum change (modulation) of the signalRs=Rnl(= 0) Rnl(==2)) is only 12% of the total spin signal RSV, which is at odds with the large spin-mixing conductance estimated from Fig. 2(b). From anistropic magnetoresistance measurements we ex- clude the possibility of any rotation of the magnetization of the injector and detector as the cause for the observed modulation in the NLSV signal (see Supplemental Mate- rial [19], Sec. III-B). Using the 3D-FEM we calculated the angular depen- dence ofRSVfor various values of Grwhere the percent- age modulation Rs=RSVis plotted as a function of , as shown in Fig. 3(c). The Grvalue of 11013 1m 2 extracted from the NLSV signal modulation experiment is one order of magnitude less than reported elsewhere [4]. This can be possibly caused by the presence of disor- dered Al/YIG interface with r.m.s. roughness of 0.8 nm (as measured by AFM), which is close to the magnetic co-herence volume3pVc'1:3 nm [6] of the YIG. This length scale determines the eective width of the Al/YIG inter- face and also the extent to which spin current from the Al is felt by the YIG magnetization [6, 24]. Furthermore, the fact that there exists a nite spin-mixing when = 0, as discussed above, can also explain the observed small modulation. It is important to note that in our experi- ments the non-equilibrium spin accumulation induced by electrical spin injection into Al has a spin-polarization strictly along the direction of the magnetization of F1, which lies along the ^ yaxis. In the measurement results shown in Figs. 1(b) and 2(b) the magnetization of the F2is always kept either parallel or antiparallel to the de- tectorF1. This ensures that it is only the ^ ycomponent of the spin accumulation that is measured in our exper- iments as it is insensitive to other two spin-polarization /s48 /s57/s48 /s49/s56/s48 /s50/s55/s48 /s51/s54/s48/s45/s52/s48/s45/s50/s48/s48/s50/s48/s52/s48 /s48 /s57/s48 /s49/s56/s48 /s50/s55/s48 /s51/s54/s48/s45/s54/s45/s51/s48/s51/s54/s32/s82 /s110/s108/s32/s40 /s41 /s32/s40/s100/s101/s103/s114/s101/s101/s115/s41/s120/s45/s99/s111/s109/s112/s111/s110/s101/s110/s116 /s40/s97/s41 /s32/s32/s82 /s110/s108/s32/s40 /s41 /s32/s40/s100/s101/s103/s114/s101/s101/s115/s41/s32/s80 /s32/s65/s80/s32 /s122/s45/s99/s111/s109/s112/s111/s110/s101/s110/s116 /s40/s98/s41 FIG. 4. (Color online) Calculated NLSV signals showing the (a)x-component and (b) z-component of the NLSV signal Rnlin the parallel (red) and antiparallel (blue) magnetiza- tion congurations of the injector and detector ferromagnetic contacts for Gr= 11013 1m 2andGi= 0:1Gr. Even if the injected spin accumulation is polarized along the magne- tization direction of the injecting electrode F 1, its interaction with the magnons via the spin-mixing conductance induces these spin accumulation components.5 directions. It is however possible that the interaction of the initially injected spin accumulation with the YIG magnetization, via G"#, to induce a nite NLSV signal with components polarized along the ^ x- and ^z-directions. Figure 4 shows the angular dependence of the ^ x and ^z component of the NLSV signal as calculated using our 3D-FEM. While the ^ zcomponent exhibits a sin( ) de- pendence, the ^ xcomponent shows a sin(2 ) dependence which is consistent with Eq. (1). The size of the mod- ulation is determined by Grfor the ^x component and byGifor the ^z component. In a collinear measure- ment conguration these transverse spin accumulation components can induce local magnetization dynamics by exerting a spin transfer torque to the YIG. Separately measuring these spin accumulation using ferromagnetic contacts magnetized along the ^ xand ^zdirections can be an alternative way to extract G"#. In summary, we studied spin injection and relaxation at the Al/YIG interface in Ni 80Fe20/Al lateral spin valves fabricated on YIG. The samples on the YIG substrate yield NLSV signals that are two to three times lower than those grown on standard SiO 2substrates, indicating spin- current absorption by the magnetic YIG substrate. We also observed a small but clear modulation of the mea- sured NLSV signal as a function of the angle between the spin accumulation and magnetization of the YIG. The presence of a disordered Al/YIG interface combined with a spin-
ip (sink) process due to thermal magnons or interface spin-orbit eects can be accounted for this small modulation. Using nite element magnetoelectronic cir- cuit theory as well as additional control experiments, we establish the concept of collinear (eective) spin mixing conductance due to the thermal magnons in the YIG. Our result therefore calls for the inclusion of this term in the analysis of spintronic and spin caloritronic phenom- ena observed in metal/YIG bilayer systems. The authors thank M. de Roosz and J.G. Holstein for technical assistance. This work is part of the research program of the Foundation for Fundamental Research on Matter (FOM) and is supported by NanoLab NL, EU- FET Grant InSpin 612759 and the Zernike Institute for Advanced Materials.6 SUPPLEMENTAL MATERIAL I. Derivation for the eective spin relaxation length in the collinear case The spin accumulation s, with polarization parallel to the magnetization direction of F1(see Fig. S5), injected in the Al is governed by the Valet-Fert spin diusion equation [25] @2 x+@2 y+@2 z s=s=2 N, which can be re-arranged to give @2 xs=s=2 N @2 zs: (5) Here we assume that, for a homogeneous system, the spin current along the ^ y-direction is zero. As discussed in the main text, when the YIG magnetization direction ^mksthe spin current jz=0 sat the Al/YIG interface, in the ^z-direction, is governed by the spin sink term Gsin Eq. 1 of the main text. Applying spin current continuity condition at the Al/YIG interface we nd that N 2@zs=Gss (6) whereNis the conductivity of the normal metal. Now after re-arranging Eq. (6) to obtain @zs, dierentiat- ing it once and using @zs=s=tAl, wheretAlis the thickness of the Al, we obtain @2 zs= 2Gss NtAl: (7) Substituting Eq. (7) into Eq. (5) we obtain a modi- ed VF-spin diusion equation that contains two length scales @2 xs=s=2 N+ 2Grs=NtAl; (8a) =s 2 N+s 2r; (8b) where we dened a new length scale 2 r= 2Gs=NtAl that, together with the N, re-denes an eective spin relaxation length 2 e= 2 N+ 2 r. This eective spin relaxation length in the Al channel is weighted by the spin-mixing conductance Gsof the Al/YIG interface. The modulation of the NLSV signal observed in our mea- surements is hence determined by the interplay between these two length scales, Nandr. While the rst quan- ties the eective spin-conductance of the Al channel (GN=NAN=Al) over the spin relaxation length, the second is a measure of the quality of the Al/YIG inter- face and is set by Gs. For the devices investigated in this work, using AN=tAlwAlwith the width of the Al channelwAl= 100nm and Al= 2107S/m, we ob- tainA 1 NGN'61013 1m 2, which is close to the Gs obtained in our experiments. This highlights the impor- tance of spin-relaxation induced by the thermal motion of the YIG magnetization, as discussed in the main text.Geometrical enhancement of the modulation can be ob- tained by reducing tAl, as shown in Fig. S5(d), thereby maximizing spin-absorption at the Al/YIG interface [16]. II. Three dimensional (3D) spin transport model Here we describe the our 3D spin transport model used to analyze our data. It is similar to that described in Ref. 20 for collinear spin transport with the possibility of studying spin-relaxation eects (i) due to the spin-mixing conductance at the Al/YIG interface as well as (ii) Hanle spin-precession due to the in-plane magnetic eld [see Sec. III below for detail]. The charge current j c(~ r) and spin currentj s(~ r), forj c(~ r) (where2x;y;z ), are re- lated to the charge c(~ r) and spin potentials sas j c(~ r) j s(~ r) = F F ~rc ~rs (9) whereis the bulk conductivity and Fis the bulk spin polarization of the conductivity. The device geometry we model is shown in Fig. S5(a), showing schematic source- drain congurations as well as voltage contacts. We im- pose charge
ux at contact 1 and drain it at 2. The nonlocal voltage, due to spin diusion, is obtained by taking the dierence between the surface integrated c at contacts 3 and 4 , both for the parallel (P) or antipar- allel (AP) magnetization congurations. To solve Eq. (9), we use conservation laws for charge ( rj c(~ r) = 0) and spin current (rjs= (1 2 F)s=2) with spin pre- cession due to the in-plane applied eld also included in the model. By dening an angle betweensand the YIG magnetization ^ mand allowing for a boundary spin current at the Al/YIG interface using Eq. 1 of the main text, we can study the transport of spins in NLSV de- vices and their interaction with the YIG magnetization. The material parameters for the model, ,Fandsare taken from Ref. [23]. Our modeling procedure involves, rst, tting of the measured NLSV signal on a SiO 2sub- strate by varying Fand usingN= 350nm. Next, we aim to nd Gsof the Al/YIG interface that properly quanties spin transport properties of the YIG sample. Figure S5(b) shows the dependence of the NLSV signal onGs. As expected, when Gsvery low, the NLSV signal is not aected by the presence of the YIG as spins are not lost to the substrate. For Gs'51013 1m 2we obtain the experimentally measured NLSV signal (shown in red dashed line). For even larger Gsvalues, the eect is maximum with the NLSV signal falling by almost one order of magnitude. It is important to remember that the value of Gsthat is extracted here is a simple mea- sure of spin-
ip processes at the Al/YIG interface due to thermal
uctuation of the YIG magnetization or disorder induced eects. At the temperatures of our experiment it is dicult to distinguish which one of the two processes is dominant.7 13 2 13 2G 8 10 m and G 5 10 mr s− −= × Ω = × Ω(c)Gs(b) 12 4 3cj/arrowrightnosp xz y(a) (d) SV S FIG. 5. (a) Geometry of the modeled device showing the measurement conguration with a 3D prole and the y-component of the spin accumulation. (b) The dependence of the NLSV signal on the eective (collinear) spin mixing conductance Gs. To reproduce the experimentally observed decrease in the spin signal from SiO 2to the YIG substrate, an eective spin mixing conductance of Gs= 51013 1m 2is required. (c) The dependence of the NLSV signal on the angle between ^ mands forGs= 51013 1m 2. (d) The dependence of the spin signal modulation amplitude on the thickness of the Al channel signifying the interplay between the spin-mixing conductance and the spin-conductance in the Al channel. For the angular dependent simulation we only vary the anglebetweensand ^mwhile keeping all other param- eters constant (such as F,NGs= 51013 1m 1 andGr= 81013 1m 1). As shown in Fig. S5(b) our simulation as described above reproduces the cos2() de- pendence observed in our experiments as well as by Vil- lamor et al. [16]. For the extracted values of Grfrom our analysis, the experimentally observed modulation of the NLSV signal by the rotating magnetization direction of the YIG is small. Possible ways to enhance the modulation are to 1) maximize the spin-mixing conductance via controlled interface engineering of the Al/YIG interface or 2) reduce the thickness of the spin transport channel. In the latter, for a xed Gr, the eect of decreasing the thickness of the spin transport channel is to eectively reduce the spin conductance GNalong the channel thereby maximizing the spin current through the Al/YIG interface. Fig- ure S5(c) shows the thickness dependence of the modula- tion of the spin signal Rs=Rs(= 00) Rs(= 900)normalized by Rsas a function of the thickness tAl, with the inset showing that for the P and AP congurations. As the thickness of the Al channel increases the spin cur- rent absorption at the Al/YIG interface decreases or vice versa. III. Investigation of possible alternative explanations for the observed modulation It can be argued that the experimentally observed modulation of the NLSV signal can be fully explained by (i) the Hanle spin-precession and/or (ii) the rotation of the magnetizations of the injector/detector electrodes due to the 5 mT in-plane magnetic eld. Below, we show that even the combined eect of both mechanisms is too small to explain the experimentally observed modulation of the NLSV signal.8 (b) (c)(a) AntiparallelParallel Injector Detector FIG. 6. (a) Modulation of the NLSV when only considering the Hanle eect due to the in-plane magnetic eld in the P (dashed lines) and AP (solid lines) at 5 mT (red), 50 mT (blue) and 100 mT (black). see text for more details. (b) Anisotropic magnetoresistance (AMR) measurement for the injector (left) and detector (right) ferromagnets at two dierent magnetic elds. The insets show the full-scale plot of the measurements at 5mT. A. Hanle spin-precession induced modulation of the NLSV signal Spins precessing around an in-plane magnetic eld ~Bwould acquire an average spin precession angle of =!LD, where!L=gB~B=~is the Larmor precession frequency,D=L2=2Dc= 25 ps is the average diusion time an electron takes to traverse the distance Lbetween the injector and the detector and Dc= 0:005m2/s is the diusion coecient [26]. For an applied eld of 5 mT andL=500 nm, we obtain = 1:25o, giving us a max- imum contribution of 1 cos=0.02% [see Eq. (10)] to the experimentally observed signal (compared to the 12% in Fig. 3(b) of the main text). This is expected be- cause the spin-precession frequency ! 1 L(8 ns) at such magnetic elds is three orders of magnitude slower than D. This simple estimate is further supported by our 3D - nite element model as we show next. Figure S6(a) shows the angle dependence of the nonlocal signal due to an in-plane magnetic eld when we only consider the Hanleeect both for the AP (solid lines) and P (dashed lines) congurations at three dierent magnetic eld values of 5 mT (red), 50 mT (blue) and 100 mT (black). The maximum modulation of the NLSV signal that the Hanle eect presents is only 0.001% at the measurement eld of 5 mT and only become relevant at high elds. There- fore, the Hanle eect alone can not explain the results presented in the main text. B. Magnetization rotation induced modulation of the NLSV signal The in-plane rotation of the sample under an applied magnetic eld of 5 mT might induce rotations in the ma- gentization of the injector/detector electrodes. In such a case, a relative angle rbetween the magnetization direc- tion of the injector and detector electrodes would result in a modulation of the NLSV signal given by Rnl Rnl(r= 0)=Rnl(r= 0) Rnl(r) Rnl(r= 0)=j1 cosrj; (10)9 with +( ) corresponding to the P (AP) conguration. Using Eq. (10), we nd that a relative angle r'28o between the magnetization directions of the injector and detector is required in order to explain the experimentally observed modulation. To determine the eld induced in- plane rotation of the magnetization by the applied mag- netic eld, we carried out angle dependent anisotropic magnetoresistance (AMR) measurements both for the in- jector and detector electrodes, using a new set of devices with identical dimensions. The AMR measurements were repeated for dierent magnetic eld strengths, at 5 mT and at higher magnetic elds of 100 mT and 300 mT. Figure S6(b) and (c) show the two-probe AMR mea- surement of the injector and detector electrodes, respec- tively, at two dierent magnetic elds. For the injector electrode in Fig. S6(b), at an applied eld of 100 mT (red line), an AMR response R=Rk R?= 0:6 is observed, where Rk(R?) is the resistance of the ferro- magnet when the angle between the applied eld and the easy axis is = 0o(= 90o). For the same electrode, at an applied eld of 5 mT (blue line, see also the inset), the AMR response is only 0.025 . Now, by comparing these two measurements we conclude that the eect of the 5 mT eld would be to rotate the magnetization of this electrode by a maximum angle 1= 15ofrom the easy axis. A similar analysis for the detector electrode, using the AMR responses of 2 (at 300 mT) and 0.025 (at 5 mT) in Fig. S6(c), yields a maximum magneti- zation rotation 2= 10o. Relevant here is the net rel- ative magnetization rotation between the two electrodes r=1 2= 5oand, using Eq. (10), we conclude that it would only cause a modulation of 0.4 %, which is much smaller than the 12% observed in our experiments. Our analysis based on the AMR eect is equivalent to that in Ref. 16 where magneto-optical Kerr eect measurements were used to exclude a possible in-plane magnetization rotation as the origin for the observed modulation in the nonlocal spin valve signal [16]. To summarize this section, the Hanle eect and the magnetization rotation induced by the in-plane magnetic eld neither separately nor when combined are sucient to explain the experimentally observed modulation. Only after including the eect of the spin-mixing interaction viaG"#that it is possible to reproduce the modulation observed in the experiments. e-mail:f.k.dejene@gmail.com [1] K. Uchida, J. Xiao, H. Adachi, J. Ohe, S. Takahashi, J. Ieda, T. Ota, Y. Kajiwara, H. Umezawa, H. Kawai, et al. , Nature materials 9, 894 (2010). [2] V. Castel, N. Vlietstra, J. Ben Youssef, and B. J. van Wees, Applied Physics Letters 101, 132414 (2012). [3] M. Althammer, S. Meyer, H. Nakayama, M. Schreier,S. Altmannshofer, M. Weiler, H. Huebl, S. Gepr ags, M. Opel, R. Gross, D. Meier, C. Klewe, T. Kuschel, J.-M. Schmalhorst, G. Reiss, L. Shen, A. Gupta, Y.-T. Chen, G. Bauer, E. Saitoh, and S. Goennenwein, Phys. Rev. B 87, 224401 (2013). [4] N. Vlietstra, J. Shan, V. Castel, B. J. van Wees, and J. Ben Youssef, Physical Review B 87, 184421 (2013). [5] J. Flipse, F. K. Dejene, D. Wagenaar, G. E. W. Bauer, J. B. Youssef, and B. J. van Wees, Phys. Rev. Lett. 113, 027601 (2014). [6] J. Xiao, G. E. W. Bauer, K.-c. Uchida, E. Saitoh, and S. Maekawa, Physical Review B 81, 214418 (2010). [7] X. Jia, K. Liu, K. Xia, and G. E. W. Bauer, EPL (Eu- rophysics Letters) 96, 17005 (2011). [8] B. Heinrich, C. Burrowes, E. Montoya, B. Kardasz, E. Girt, Y.-Y. Song, Y. Sun, and M. Wu, Physical Re- view Letters 107, 066604 (2011). [9] Y.-T. Chen, S. Takahashi, H. Nakayama, M. Althammer, S. T. B. Goennenwein, E. Saitoh, and G. E. W. Bauer, Physical Review B 87, 144411 (2013). [10] M. Weiler, M. Althammer, M. Schreier, J. Lotze, M. Pernpeintner, S. Meyer, H. Huebl, R. Gross, A. Kamra, J. Xiao, Y.-T. Chen, H. Jiao, G. Bauer, and S. Goennenwein, Phys. Rev. Lett. 111, 176601 (2013). [11] S. Huang, X. Fan, D. Qu, Y. Chen, W. Wang, J. Wu, T. Chen, J. Xiao, and C. Chien, Phys. Rev. Lett. 109, 107204 (2012). [12] T. Kikkawa, K. Uchida, S. Daimon, Y. Shiomi, H. Adachi, Z. Qiu, D. Hou, X.-F. Jin, S. Maekawa, and E. Saitoh, Physical Review B 88, 214403 (2013). [13] F. J. Jedema, M. V. Costache, H. B. Heersche, J. J. A. Baselmans, and B. J. van Wees, Applied Physics Letters 81(2002). [14] F. J. Jedema, M. S. Nijboer, A. T. Flip, and B. J. van Wees, Physical Review B 67, 085319 (2003). [15] T. Kimura, T. Sato, and Y. Otani, Physical Review Let- ters100, 066602 (2008). [16] E. Villamor, M. Isasa, S. V elez, A. Bedoya-Pinto, P. Vavassori, L. E. Hueso, F. S. Bergeret, and F. Casanova, Phys. Rev. B 91, 020403 (2015). [17] V. Castel, N. Vlietstra, B. J. van Wees, and J. B. Youssef, Physical Review B 86, 134419 (2012). [18] Z. Qiu, K. Ando, K. Uchida, Y. Kajiwara, R. Takahashi, H. Nakayama, T. An, Y. Fujikawa, and E. Saitoh, Ap- plied Physics Letters 103, 092404 (2013). [19] See supplemental material for the derivation of Eq. 3, detailed procedure of the nite element simulation and additional control experiments. [20] A. Slachter, F. L. Bakker, and B. J. van Wees, Phys. Rev. B 84, 174408 (2011). [21] V. S. Rychkov, S. Borlenghi, H. Jares, A. Fert, and X. Waintal, Phys. Rev. Lett. 103, 066602 (2009). [22] F. L. Bakker, A. Slachter, J.-P. Adam, and B. J. van Wees, Phys. Rev. Lett. 105, 136601 (2010). [23] F. K. Dejene, J. Flipse, G. E. W. Bauer, and B. J. van Wees, Nature Physics 9, 636 (2013). [24] M. Schreier, A. Kamra, M. Weiler, J. Xiao, G. E. W. Bauer, R. Gross, and S. T. B. Goennenwein, Physical Review B 88, 094410 (2013). [25] T. Valet and A. Fert, Physical Review B 48, 7099 (1993). [26] F. J. Jedema, M. V. Costache, H. B. Heersche, J. J. A. Baselmans, and B. J. van Wees, Applied Physics Letters 81(2002). | 2015-03-20 | We study the effect of a magnetic insulator (Yttrium Iron Garnet - YIG)
substrate on the spin transport properties of Ni$_{80}$Fe$_{20}$/Al nonlocal
spin valve (NLSV) devices. The NLSV signal on the YIG substrate is about 2 to 3
times lower than that on a non magnetic SiO$_2$ substrate, indicating that a
significant fraction of the spin-current is absorbed at the Al/YIG interface.
By measuring the NLSV signal for varying injector-to-detector distance and
using a three dimensional spin-transport model that takes spin current
absorption at the Al/YIG interface into account we obtain an effective
spin-mixing conductance $G_{\uparrow\downarrow}\simeq 5 - 8\times
10^{13}~\Omega^{-1}$m$^{-2}$. We also observe a small but clear modulation of
the NLSV signal when rotating the YIG magnetization direction with respect to
the fixed spin polarization of the spin accumulation in the Al. Spin relaxation
due to thermal magnons or roughness of the YIG surface may be responsible for
the observed small modulation of the NLSV signal. | Control of spin current by a magnetic YIG substrate in NiFe/Al nonlocal spin valves | 1503.06108v1 |
arXiv:2202.03774v1 [cond-mat.mes-hall] 8 Feb 2022AIP/123-QED Modulation of Spin Seebeck Effect by Hydrogenation K. Ogata,1T. Kikkawa,2E. Saitoh,2, 3and Y. Shiomi1, 4 1)Department of Integrated Science, University of Tokyo, Meg uro, Tokyo 153-8902, Japan 2)Department of Applied Physics, University of Tokyo, Bunkyo , Tokyo 113-8656, Japan 3)Institute for AI and Beyond, University of Tokyo, Bunkyo, To kyo 113-8656, Japan 4)Department of Basic Science, University of Tokyo, Meguro, T okyo 153-8902, Japan (*Electronic mail: yukishiomi@g.ecc.u-tokyo.ac.jp) (Dated: 9 February 2022) We demonstrate the modulation of spin Seebeck effect (SSE) b y hydrogenation in Pd/YIG bilayers. In the presence of 3% hydrogen gas, SSE voltage dec reases by more than 50 % from the magnitude observed in pure Ar gas. The modulation o f the SSE voltage is reversible, but the recovery of the SSE voltage to the prehyd rogenation value takes a few days because of a long time constant of hydrogen desorption. We also demonstrate that the spin Hall magnetoresistance of the identical sample reduce s significantly with hydrogen exposure, supporting that the observed modulation of spin c urrent signals originates from hydrogenation of Pd/YIG. 1Hydrogen is an energy carrier which can be produced by an envi ronmentally clean process and therefore has a positive impact on decarbonization1. To utilize hydrogen as a clean and renewable alternative to carbon-based fuels, hydrogen safety sensor s are also critical to assure the develop- ment of hydrogen systems2. Metal-hydride systems have been widely studied for the pot ential of solid-state hydrogen storage and sensing. In particular, P d is frequently used as a catalyst for hy- drogen dissociation and adsorption. A hydrogen molecule de composes into independent hydrogen atoms when the molecule approaches a Pd surface due to the str ong interaction between Pd and H atoms. As the smallest single atom, a H atom can easily diffus e into the interstices of the Pd lattice and cause lattice expansion. As a result, hydrogen adsorpti on changes the density of states of Pd near the Fermi energy3,4, significantly modulating its electrical and optical prope rties. The hydrogenation of Pd films also impacts spintronic effect s. Pd is known to exhibit a strong spin-orbit coupling and has been used in many spintro nic experiments. Magnetic multi- layers and super-lattices which include Pd layers have been of particular interest. It was reported that in Co/Pd bilayers which possess a strong interfacial pe rpendicular magnetic anisotropy, the magnetic anisotropy and ferromagnetic resonance are re versibly modulated by hydrogen exposure5–11. Efficient hydrogen sensing based on magnetization dynamic s was also reported in similar materials5,12,13. Moreover, the inverse spin Hall effect (ISHE) induced by sp in pumping was successfully modulated by hydrogen exposure in Co/Pd bi layers14,15. Absorption of hydrogen gas at 3% concentration in the Pd layer reduces the ISHE volta ge by 20%15. This decrease in ISHE signals in the presence of hydrogen gas was attributed to the decrease in spin diffusion length due to enhanced scatterings to hydrogen atoms in the Pd layer. Af ter the hydrogen gas is flushed out of the setup, the ISHE voltage returns to the prehydrogenati on value; hence the observed effect is reversible. The studies of hydrogen effects on spintronic materials hav e been carried out for the combi- nation of Pd layers with itinerant magnetic films. For the mea surement of ISHE in ferromag- netic/nonmagnetic metallic bilayers, however, it is known that the precise estimation of ISHE voltage is difficult because of spurious spin rectification e ffects such as anisotropic magnetoresis- tance and anomalous Hall effect which couple the dynamic mag netization to microwave currents in the ferromagnetic layer16. Hence magnetic insulators such as Y 3Fe5O12(YIG) should be more suitable to investigate hydrogen effects on pure ISHE signa ls17. In this letter, we demonstrate the reversible manipulation of the spin Seebeck effect (SSE) by hydrogen exposure in Pd/YIG bilayers. The SSE is the generat ion of a spin current as a result of 2a temperature gradient applied across a junction consistin g of a magnet and a metal18,19. The spin current injected into a metal can be converted into a voltage by ISHE. Since the ISHE induced by SSE is a transverse thermoelectric effect, it can be employe d to realize transverse thermoelectric devices, which could potentially overcome the inherent lim itations of conventional thermoelectric devices20,21. Moreover, SSE is expected to be utilized for flexible heat-fl ow sensors22. The ma- nipulation of SSE by hydrogenation demonstrated below may o pen up new device potentials in spin caloritronics. We used epitaxial YIG films with 2 micron thickness grown by li quid phase epitaxy on Gd3Ga5O12(111) substrates. The YIG surfaces were mechanically polis hed, and then 5-nm-thick Pd films were sputtered at room temperature. The Pd layer is 5 m m long and 0.5 mm wide. For the Pd/YIG bilayers, the SSE measurements in the longitudin al configuration18were performed at room temperature using an electromagnet (3470 Electroma gnet System, GMW Associates). The bilayer sample was placed between sapphire and copper pl ates. A 1-k Ωresistive heater was attached to the upper sapphire plate and the lower copper pla te is a heat sink. To facilitate hydro- genation of the Pd layer, a breathable tape (TBAT-252, TRUSC O) was inserted between the upper plate and the sample [Fig. 1(a)]. The temperature gradient i s generated by applying an electric current to the heater. Two-pairs of leads were attached to th e Pd layer to measure not only the SSE but also the spin Hall magnetoresistance (SMR)23in the same setup. The distance between the voltage terminals is 2.5 mm. The thermoelectric voltage due to the ISHE induced by SSE was monitored with a Keithley 2182A nanovoltmeter. SMR was m easured by lockin detection using Anfatec USB Lockin Amplifier 250; the frequency and amp litude of ac electric current are 111 Hz and 0.8 mA. The sample was loaded into a small chamber to control the atmosphere. For hydrogenation measurements, the samples were first measure d in pure Ar gas ( >99.9999 vol.%) at atmospheric pressure followed by a 3%/97% H 2/Ar gas mixture. Before the measurements in Ar-H 2gas, we waited 20-40 minutes for the Pd layer to be completely hydrogenated15,24after the chamber was filled with 1 atm Ar-H 2gas. First, we measured SSE voltage VSSEof Pd/YIG in Ar atmosphere. Figure 1(b) shows the magnetic-field ( H) dependence of VSSEmeasured at several heater power levels. Here symmetric components of the output voltage with respect to Hare subtracted and antisymmetric components are plotted; note that the symmetric components which are no t to be attributed to the effect under study are almost independent of Hin our measurements. When the heater is off, VSSEis almost zero in the entire Hrange in Fig. 1(b). As the heater power Pincreases from zero, the clear 3FIG. 1. (a) Measurement setup of the SSE. A breathable tape wa s inserted between Pd/YIG and the heater part to facilitate hydrogen absorption and desorption in th e Pd film. (b) Magnetic field ( H) dependence of the SSE voltage ( VSSE) measured in 1 atm of Ar. The heater power ( P) was changed from 0 mW to 100 mW. (c) Heater power ( P) dependence of the SSE voltage ( VSSE) at 200 mT. The raw data is shown in (b). SSE signals appear and their magnitudes increase with P. The sign of VSSEis the same as that reported for Pt/YIG18. The saturated magnitude of VSSEis plotted against Pin Fig. 1(c). The VSSEmagnitude increases linearly with the heater power, which i ndicates that VSSEis proportional to temperature gradient generated across the Pd/YIG juncti on. The temperature difference ∆T generated at P=100 mW is estimated to be ∼1.5 K (see Fig. S1 in Supplementary Material). Next, the effect of exposing the 3% H 2mixture on the Pd/YIG sample is investigated in Fig. 2 (see also Fig. S2 in Supplementary Material for additional e xperimental results). Here the heater power Pis kept constant at 100 mW during the series of measurements. After the initial SSE measurement in pure Ar gas at atmospheric pressure already s hown in Figs. 1(b) and 1(c), the sample chamber was filled with H 23% H 2/Ar mixture and the SSE measurement was performed. As shown in Fig. 2, the magnitude of VSSEis found to be reduced by more than 50% in the presence of H 2gas. After completing the SSE measurement in Ar-H 2atmosphere, the sample was then remeasured in pure Ar. The VSSEmagnitude returned to the pristine value as shown in Fig. 2. Note that this data was taken 2.5 days after the cham ber was refilled with Ar. The observed decrease in the SSE signal is safely ascribed to the presence of hydrogen in Pd/YIG, and 4FIG. 2. Magnetic field ( H) dependence of the SSE voltage ( VSSE) measured before hydrogenation, during exposure to hydrogen gas, and after the hydrogen gas is flushe d out of the setup. The heater power is kept at 100 mW. importantly, the change is reversible. It is well known15,25that upon hydrogenation, Pd thin films undergo two stages of l attice ex- pansion depending on the hydrogen gas concentration. For li ght concentration levels up to 2-3 %, the lattice constant grows by approximately 1% in the out-of -plane direction only. This expansion is reversible. In the second stage, the lattice constant gro ws by up to 4% in both out-of-plane and in-plane directions. These changes are irreversible, caus ing structural changes to the Pd lattice. In our SSE measurements under 3% hydrogen gas, the sample sho uld undergo the first stage of lattice expansion and the SSE is thereby reversible. Note th at we confirmed by x-ray diffraction that the Pd films are (111) oriented as in the literature25(see Supplementary Material). Though the modulation of SSE by hydrogen absorption/desorp tion is reversible, the recovery 5FIG. 3. (a) Magnetic field ( H) dependence of the SSE voltage ( VSSE) measured 1.7-66 hours after Ar gas is refilled in the measurement chamber. The heater power is kept at 100 mW. (b) Time dependence of the SSE voltage ( VSSE) at 210 mT measured after Ar gas is refilled in the measurement chamber. The selected raw data is shown in (a). The black curve is a fit to the experimenta l data (see text). time of the SSE signal due to hydrogen desorption is as long as 2.5 days. It was reported that the hydrogen desorption takes a long time in contrast to the q uick hydrogen absorption9, and the response time depends significantly on materials. The time f or hydrogen desorption is typically at most several tens of minutes for Co/Pd5–11, while the completion of the entire desorption requires at least a few days at 10−3mbar for FePd alloys26. Our Pd/YIG also includes Fe and Pd, and the situation looks similar to FePd alloys. We then take a closer look on the dehydrogenation process by t he time dependent measurement of SSE in Fig. 3. Figure 3(a) shows VSSEcurves measured at different times after the measurement chamber is refilled with pure Ar gas. The VSSEmagnitude is approximately 0.5 µV just after the gas is replaced with Ar, and increases monotonically with ti me. After 50 hours, the VSSEmagnitude is almost saturated at ∼1µV. The time dependence of VSSEat 210 mT is plotted in Fig. 3(b). The VSSEmagnitude increases monotonically with time, as already shown in Fig. 3(a). We fit the experimental data by a standard relaxation function: VSSE∝1−e−t/τ, where tis the measurement time and τis a time constant of hydrogen desorption. The fitting curve matches the experi mental data very well, meaning that the hydrogen desorption follows an exponential function. T he same function was adopted for the hydrogenation effect on magneto-optical effects in Pd/Co/ Pd films9. The fit in Fig. 3(b) yields τ≈25 hour. Such a long time constant was not observed in the spin pumping measurement for 6FIG. 4. Magnetic field ( H) dependence of the magnetoresistance (MR) ratio ( ρ(H)/ρ(H=0)−1) measured before hydrogenation (a), during exposure to hydrogen gas ( b), and after the hydrogen gas is flushed out of the setup (c). Pd/Co bilayers15. In contrast to the spin pumping measurements, the attachmen t of the heater to the Pd surface is required in the SSE measurements, which may adversely affec t the absorption/desorption of hy- drogen because of small numbers of exposed surface atoms. To confirm that spin current signals in the Pd layer is indeed modulated by hydrogenation, we also perform the measurement of SMR (spin Hall magnetoresistance) for the same sample in the sam e setup. The SMR is a magnetoresis- tance effect related to a nonequilibrium proximity effect c aused by the simultaneous action of the SHE and ISHE23,27; the absorption/reflection of spin current at the ferromagn et/metal interface re- sults in magnetoresistance, since the spin-dependent scat tering at the metal/ferromagnet interface depends on the angle between the polarization of spin Hall cu rrent and the magnetization of the attached magnetic layer. The experimental setup is illustr ated in the inset to Fig. 4(a). Magnetic field is applied perpendicular to the electric-current dire ction in the film plane. Figure 4 shows the hydrogen effects on SMR in the Pd/YIG bilay er. Here, since the size of SMR is very small, the magnetoresistance measurements were repeated several times and aver- aged. The error bars stand for the standard errors. Before hy drogenation [Fig. 4(a)], a negative magnetoresistance effect is observed. The magnetic-field d ependence of resistance change follows the magnetization process of the YIG layer, consistent with the SMR23. The size of SMR is about 1×10−3%. This magnitude is about ten times smaller than that in Pt/Y IG23. A small SMR of about 10% compared to Pt/YIG was also reported in the literat ure28. During the exposure to 3% hydrogen gas, the SMR magnitude dec reases significantly as shown 7in Fig. 4(b). Although quantitative analysis is difficult be cause of the large error bars, the suppres- sion of SMR ratio by hydrogenation looks more than 50%, consi stent with the modulation in SSE voltages (Figs. 2 and 3). After the hydrogen gas is flushed out of the chamber and pure Ar gas is refilled, we confirmed that the size of SMR returns to the initi al value [Fig. 4(c)]. An important finding in the SMR measurement is that the SMR rat io has already returned to its original value 30 minutes after refilling Ar gas. Namely, the time constant of hydrogen desorption in the SMR measurement is much shorter than that in the SSE mea surement. Since both the measurements were performed for the same sample in the same s etup, the long time constant of hydrogen desorption in the SSE measurement cannot be attrib uted to impurities/defects in the Pd layer, surface oxidation, surface morphology9, or moisture which may trap hydrogen atoms and hinder the hydrogen desorption29,30. In our measurements of SSE and SMR, spin current signals are s ignificantly suppressed by hydrogen exposure as shown in Figs. 2-4. The decrease in the s pin Hall signals with hydrogen exposure is consistent with the previous spin pumping measu rements for Co/Pd14,15. Scatterings of conduction electrons to hydrogen atoms in the Pd layer dec rease the spin diffusion length due to the enhanced Elliot-Yafet relaxation mechanism, and res ult in the decrease in spin-pumping signals15. This mechanism should be also applicable to SSE and SMR. Sin ce the theory has shown that both effects depend on spin diffusion length and s pin Hall angle of the Pd layer19,27, the signal variation by hydrogenation can be attributed to the d ecrease in the spin diffusion length15. On the other hand, it is notable that magneto-optical Kerr si gnals are enhanced by hydrogenation in Co/Pd bilayers9,10, in contrast to the decrease in the spin current signals14,15. In transport measurements such as (inverse) spin Hall effects, enhanced electron scatterings due to interstitial hydrogen impurities are likely to play a dominant role in the hydrogenation effect. The significant scattering effect due to hydrogen atoms is also evidenced by the reduction of the anomalous Hall signal in hydrogenated Co xPd1−xfilms12. The decrease in the VSSEmagnitude ( >50%) by hydrogen exposure is noticeably greater than the change in ISHE signals reported in the spin pumping measu rements15; the decrease in the spin- pumping voltage in H 2/Ar mixture with 3% of hydrogen was only 20%. The larger signa l change in our results suggests that there may be other factors for the r eduction of VSSEbesides hydrogenation of the Pd layer. The first possibility is imperfect separatio n of ISHE signals from spin rectification effects in metallic Co/Pd bilayers14,15. Another possible origin is different interfacial stresse s to the Pd layer between YIG and Co. It is known that electrical re sistivity of single-layer Pd grown on 8Si substrates increases upon hydorogenation, while it tend s to decrease for bilayer cases15because of the interfacial compressive stress from the underlying l ayer. The interfacial stress can also affect the interface spin mixing conductance, modulating the inje ction efficiency of spin currents. Note that the resistivity of the Pd film on YIG decreases by hydroge nation, but the change in resistivity is as small as 1% (Fig. S3 in Supplementary Material), which c annot explain the large variation (>50%) of SSE voltage by hydrogen exposure. Moreover, since the SSE also depends on bulk spin transport i n the YIG layer31,32in contrast to the spin pumping, hydrogen effects on YIG may contribute t o the significant reduction in the VSSEmagnitude. Hydrogen diffusion in YIG was indeed reported fo r annealed samples in H 2 atmosphere33–35. The hydrogen diffusion in the YIG layer can suppress the mag non and phonon transport, which should reduce the VSSE. Also the interface spin-exchange coupling can be weak- ened by hydrogen around the interface, leading to the decrea se in the interface spin-injection efficiency. The presence of hydrogen effects on the YIG layer is also sugg ested by the different recovery time constants between SSE and SMR. Our measurements in Figs . 3 and 4 showed that the time constant for the signal recovery of SSE is much longer than th at of SMR. The different recovery time constants are attributable to different bulk sensitiv ity of these effects. In SMR, spin-dependent scattering at the Pd/YIG interface is essential. In contras t, bulk thermal spin current also plays an important role in the SSE voltage31,32in addition to the interfacial spin coupling. Bulk properti es of magnetic materials such as bulk magnetization, thermal c onductivity, and magnon transport coefficient contribute to the SSE signals, but not to SMR or sp in pumping. Also in the case of magneto-optical effects of Pd/Co frequently studied befor e for hydrogenation effects, the variation of perpendicular magnetic anisotropy originates from inte rface effects5,9. Hence the SSE is a rare spintronic phenomenon that depends not only on interface pr operties but also on bulk properties of magnons and phonons in the magnetic layer. Hydrogen effec ts on YIG could be related to the reduction of VSSEand also the long time constant for hydrogen desorption in SS E. In conclusion, we experimentally demonstrated the reversi ble modulation of SSE and SMR by hydrogenation in Pd/YIG bilayers. Absorption of hydrogen r esults in the decrease in both SSE and SMR signals by more than 50%. Enhanced scatterings of con duction electrons to hydrogen atoms in the Pd layer are partly responsible for the decrease in the spin-current signals, as reported in the previous spin pumping experiments. The modulation of SSE voltage is reversible, but the time constant for the signal recovery is longer than 2 days. T he long time constant for hydrogen 9desorption in the SSE measurement is in contrast to the case o f SMR, in which the SMR ratio already returned to the prehydrogenation value 30 minutes a fter the chamber was refilled with pure Ar. We speculate that the significant decrease in the SSE magnitude by hydrogen exposure and the long time constant for hydrogen desorption in SSE are related to the hydrogen modulation of bulk properties of the YIG layer, since the SSE depends not only on interfacial spin couplings but also on bulk properties of the magnetic layer. We hope tha t the present results will stimulate further research on hydrogen effects on Pd films grown on insu lating magnetic oxides. See the supplementary material for additional SSE data, x-r ay diffraction data, and resistivity change by hydrogen exposure. We thank Y . Miyazaki for the experimental help of sample prep aration and Dr. T. Yok- ouchi for the fruitful discussion. This research was suppor ted by JST CREST (JPMJCR20C1 and JPMJCR20T2), Institute for AI and Beyond of the Universi ty of Tokyo, and JSPS KAK- ENHI Grant Numbers JP20H05153, JP20H02599, JP20H04631, JP 21K18890, JP19H05600, and JP19H02424. The data that support the findings of this study are available from the corresponding author upon reasonable request. REFERENCES 1Etienne Rivard, Michel Trudeau, and Karim Zaghib. Hydrogen storage for mobility: A review. Materials , 12(12), 2019. 2William J. Buttner, Matthew B. Post, Robert Burgess, and Car l Rivkin. An overview of hydrogen safety sensors and requirements. International Journal of Hydrogen Energy , 36(3):2462–2470, 2011. The Third Annual International Conference on Hydroge n Safety. 3Steven G. Louie. Hydrogen on pd(111): Self-consistent elec tronic structure, chemical bonding, and photoemission spectra. Phys. Rev. Lett. , 42:476–479, Feb 1979. 4P. Jena, F. Y . Fradin, and D. E. Ellis. Models of electronic st ructure of hydrogen in metals: Pd-h. Phys. Rev. B , 20:3543–3551, Nov 1979. 5Crosby S. Chang, Mikhail Kostylev, and Eugene Ivanov. Metal lic spintronic thin film as a hydrogen sensor. Applied Physics Letters , 102(14):142405, 2013. 6Wen-Chin Lin, Cheng-Jui Tsai, Xin-Ming Liu, and Adekunle O. Adeyeye. Critical hydrogena- 10tion effect on magnetic coercivity of perpendicularly magn etized co/pd multilayer nanostruc- tures. Journal of Applied Physics , 116(7):073904, 2014. 7K. Munbodh, F. A. Perez, C. Keenan, D. Lederman, M. Zhernenko v, and M. R. Fitzsimmons. Effects of hydrogen/deuterium absorption on the magnetic p roperties of co/pd multilayers. Phys. Rev. B , 83:094432, Mar 2011. 8Satoshi Okamoto, Osamu Kitakami, and Yutaka Shimada. Enhan cement of magnetic anisotropy of hydrogenated pd/co/pd trilayers. Journal of Magnetism and Magnetic Materials , 239(1):313– 315, 2002. International Symposium on Physics of Magnetic M aterials/Internat ional Sympo- sium on Advanced Magnetic Technologies. 9Wen-Chin Lin, Cheng-Jui Tsai, Bo-Yao Wang, Chao-Hung Kao, a nd Way-Faung Pong. Hydro- genation induced reversible modulation of perpendicular m agnetic coercivity in pd/co/pd films. Applied Physics Letters , 102(25):252404, 2013. 10Wen-Chin Lin, Cheng-Jui Tsai, Han-Yuan Huang, Bo-Yao Wang, Venkata Ramana Mudinepalli, and Hsiang-Chih Chiu. Hydrogen-mediated long-range magne tic ordering in pd-rich alloy film. Applied Physics Letters , 106(1):012404, 2015. 11Satoshi Akamaru, Takashi Matsumoto, Mikako Murai, Katsuhi ko Nishimura, Masanori Hara, and Masao Matsuyama. Sensing hydrogen in the gas phase using ferromagnetic pd–co films. Journal of Alloys and Compounds , 645:S213–S216, 2015. Supplement Issue: Proceedings from the 14th International Symposium on Metal-Hydrogen System s: Fundamentals and Applica- tions, 2014 (MH2014). 12A. Gerber, G. Kopnov, and M. Karpovski. Hall effect spintron ics for gas detection. Applied Physics Letters , 111(14):143505, 2017. 13Jaw-Yeu Liang, Yun-Chieh Pai, Tu-Ngoc Lam, Wen-Chin Lin, Ti ng-Shan Chan, Chih-Huang Lai, and Yuan-Chieh Tseng. Using magnetic structure of co40 pd60/cu for the sensing of hydro- gen. Applied Physics Letters , 111(2):023503, 2017. 14Stuart Watt, Rong Cong, Chris Lueng, Manu Sushruth, Peter J. Metaxas, and Mikhail Kostylev. Impact of hydrogen gas on the inverse spin hall effect in pall adium/cobalt bilayer films. IEEE Magnetics Letters , 9:1–4, 2018. 15Stuart Watt and Mikhail Kostylev. Manipulation of the inver se spin hall effect in palladium by absorption of hydrogen gas. Phys. Rev. B , 101:174422, May 2020. 16Ryo Iguchi and Eiji Saitoh. Measurement of spin pumping volt age separated from extrinsic microwave effects. Journal of the Physical Society of Japan , 86(1):011003, 2017. 1117Y Kajiwara, K Harii, S Takahashi, Jun-ichiro Ohe, K Uchida, M Mizuguchi, H Umezawa, H Kawai, Kazuya Ando, K Takanashi, et al. Transmission of ele ctrical signals by spin-wave interconversion in a magnetic insulator. Nature , 464(7286):262–266, 2010. 18Ken-ichi Uchida, Hiroto Adachi, Takeru Ota, Hiroyasu Nakay ama, Sadamichi Maekawa, and Eiji Saitoh. Observation of longitudinal spin-seebeck eff ect in magnetic insulators. Applied Physics Letters , 97(17):172505, 2010. 19Ken-ichi Uchida, Hiroto Adachi, Takashi Kikkawa, Akihiro K irihara, Masahiko Ishida, Shinichi Yorozu, Sadamichi Maekawa, and Eiji Saitoh. Thermoelectri c generation based on spin seebeck effects. Proceedings of the IEEE , 104(10):1946–1973, 2016. 20Min Young Kim, Sang J. Park, Gi-Yeop Kim, Si-Young Choi, and H yungyu Jin. Designing efficient spin seebeck-based thermoelectric devices via si multaneous optimization of bulk and interface properties. Energy Environ. Sci. , 14:3480–3491, 2021. 21T. Kikkawa, D. Reitz, H. Ito, T. Makiuchi, T. Sugimoto, K. Tsu nekawa, S. Daimon, K. Oyanagi, R. Ramos, S. Takahashi, Y . Shiomi, Y . Tserkovnyak, and E. Sai toh. Observation of nuclear-spin Seebeck effect. NATURE COMMUNICATIONS , 12(1):4356, JUL 16 2021. 22Akihiro Kirihara, Koichi Kondo, Masahiko Ishida, Kazuki Ih ara, Yuma Iwasaki, Hiroko Someya, Asuka Matsuba, Ken-ichi Uchida, Eiji Saitoh, Naoha ru Yamamoto, et al. Flexible heat-flow sensing sheets based on the longitudinal spin seeb eck effect using one-dimensional spin-current conducting films. Scientific reports , 6(1):1–7, 2016. 23H. Nakayama, M. Althammer, Y .-T. Chen, K. Uchida, Y . Kajiwar a, D. Kikuchi, T. Ohtani, S. Geprägs, M. Opel, S. Takahashi, R. Gross, G. E. W. Bauer, S. T. B. Goennenwein, and E. Saitoh. Spin hall magnetoresistance induced by a nonequi librium proximity effect. Phys. Rev. Lett. , 110:206601, May 2013. 24D Lederman, Y Wang, EH Morales, RJ Matelon, GB Cabrera, UG V ol kmann, and AL Cabrera. Magnetooptic properties of fe/pd and co/pd bilayers under h ydrogen absorption. Applied physics letters , 85(4):615–617, 2004. 25Takashi Harumoto, Yusuke Ohnishi, Keishi Nishio, Takashi I shiguro, Ji Shi, and Yoshio Naka- mura. In-situ x-ray diffraction study of hydrogen absorpti on and desorption processes in pd thin films: Hydrogen composition dependent anisotropic expansi on and its quantitative description. AIP Advances , 7(6):065108, 2017. 26Po-Chun Chang, Tzu-Hung Chuang, Der-Hsin Wei, and Wen-Chin Lin. Thermally modulated hydrogenation in fexpd1-x alloy films: Temperature-driven peculiar variation of magnetism. 12Applied Physics Letters , 116(10):102407, 2020. 27Yan-Ting Chen, Saburo Takahashi, Hiroyasu Nakayama, Matth ias Althammer, Sebastian T. B. Goennenwein, Eiji Saitoh, and Gerrit E. W. Bauer. Theory of s pin hall magnetoresistance. Phys. Rev. B , 87:144411, Apr 2013. 28Li Ma, Lili Lang, Jeongwoo Kim, Zhe Yuan, Ruqian Wu, Shiming Z hou, and Xuepeng Qiu. Spin diffusion length and spin hall angle in pd1−xptx/YIG heterostructures: Examination of spin relaxation mechanism. Phys. Rev. B , 98:224424, Dec 2018. 29Deepa Jose and Balaji R. Jagirdar. Nature of hydrogen atom tr apped inside palladium lattice. International Journal of Hydrogen Energy , 35(13):6804–6811, 2010. ISMF-09. 30S. Kishore, J.A. Nelson, J.H. Adair, and P.C. Eklund. Hydrog en storage in spherical and platelet palladium nanoparticles. Journal of Alloys and Compounds , 389(1):234–242, 2005. 31Ryo Iguchi, Ken-ichi Uchida, Shunsuke Daimon, and Eiji Sait oh. Concomitant enhancement of the longitudinal spin seebeck effect and the thermal cond uctivity in a pt/yig/pt system at low temperatures. Phys. Rev. B , 95:174401, May 2017. 32Er-Jia Guo, Joel Cramer, Andreas Kehlberger, Ciaran A. Ferg uson, Donald A. MacLaren, Ger- hard Jakob, and Mathias Kläui. Influence of thickness and int erface on the low-temperature enhancement of the spin seebeck effect in yig films. Phys. Rev. X , 6:031012, Jul 2016. 33E. Milani and P. Paroli. Optical study of hydrogen diffusion in yttrium iron garnet. Journal of Applied Physics , 55(6):2173–2175, 1984. 34E. Milani, P. Paroli, and P. DeGasperis. Hydrogen diffusion in yttrium iron garnet films. Thin Solid Films , 126(1):73–76, 1985. 35A. Leiberich, E. Milani, P. Paroli, and R. Wolfe. Comment on “ hydrogen depth profiles in ion-implanted magnetic bubble garnets”. Journal of Applied Physics , 60(2):836–837, 1986. 13 | 2022-02-08 | We demonstrate the modulation of spin Seebeck effect (SSE) by hydrogenation
in Pd/YIG bilayers. In the presence of 3% hydrogen gas, SSE voltage decreases
by more than 50% from the magnitude observed in pure Ar gas. The modulation of
the SSE voltage is reversible, but the recovery of the SSE voltage to the
prehydrogenation value takes a few days because of a long time constant of
hydrogen desorption. We also demonstrate that the spin Hall magnetoresistance
of the identical sample reduces significantly with hydrogen exposure,
supporting that the observed modulation of spin current signals originates from
hydrogenation of Pd/YIG. | Modulation of Spin Seebeck Effect by Hydrogenation | 2202.03774v1 |
Eect of dipolar interactions on cavity magnon-polaritons Antoine Morin, Christian Lacroix, and David M enard Department of Engineering Physics, Polytechnique Montr eal, Montr eal, Qc (Dated: September 11, 2020) The strong photon-magnon coupling between an electromagnetic cavity and two yttrium iron garnet (YIG) spheres has been investigated in the context of a strong mutual dipolar interaction between the spheres. A decrease in the coupling strength between the YIG spheres and the elec- tromagnetic cavity is observed, along with an increase of the total magnetic losses, as the distance between the spheres is decreased. A model of inhomogeneous broadening of the ferromagnetic res- onance linewidth, partly mitigated by the dipolar narrowing eect, reproduces the reduction in the coupling strength observed experimentally. These ndings have important implications for the un- derstanding of strongly coupled photon-magnon system involving densely packed magnetic objects, such as ferromagnetic nanowires arrays, in which the total coupling strength with an electromagnetic cavity might become limited due to mutual dipolar interactions. I. INTRODUCTION Following recent works on the strong coupling between a magnonic mode of a ferromagnetic sample and a pho- tonic mode of a microwave cavity1,2, also called cavity magnon-polaritons3,4, much interest emerged in the sci- entic community to exploit the phenomenon as a mean to develop novel information transfer technologies5{10. Some interesting propositions involve multiple yttrium iron garnet (YIG) spheres placed inside an electromag- netic cavity, such as the magnon gradient memory11, the long distance modication of spin currents12, and the de- velopment of ultrahigh sensitivity magnetometers13. As these new ideas are being elaborated14, it is impor- tant to correctly predict the behavior of photon-magnon systems consisting of several ferromagnetic elements cou- pled to an electromagnetic resonator. In this context, the eect of dipolar interactions between the ferromagnetic objects on the strong photon-magnon coupling is crucial and remains relatively unexplored. The strong coupling of photon-magnon systems is well understood and has been recently reviewed15. Its ex- tension to multiple independent magnons is relatively straightforward16. For an ensemble of Nindependent and identical ferromagnetic objects, the ideal coupling is expected to be enhanced by a factor ofp Nas com- pared to the coupling strength of a single object to the cavity. However, due to dipolar interactions between the magnetic elements, some detuning along with inhomoge- neous broadening are expected for coupled magnon sys- tems. In this paper, we investigate the coupling strength of a simplied system consisting of two YIG spheres cou- pled to an adjustable microwave cavity. We show exper- imentally that the coupling constant gdecreases as the spheres are brought closer. The results are explained us- ing a model based on the Landau-Lifshitz equation and the Fourier expansion of the magnetization in order to include the coupling of the photons with the uniform fer- romagnetic mode as well as with the long wavelength spin wave modes, which are excited in presence of a non- uniform magnetic eld. Y I G s p h e r e s D e p t h c o n t r o l L e n g t h c o n t r o l M e t a l l i c r o d S h o r tH0hxYIG spheres Depth control Length control Metallic rodShortFIG. 1. Schematic representation of the tunable cavity used experimentally. The metallic rod allows the tuning of the resonance frequency and the losses of the cavity. The direction of the eld H0and the RF magnetic eld hxare also shown. II. EXPERIMENTAL PROCEDURE A tunable waveguide cavity consisting of a shorted X- band waveguide in which a metallic rod of 1 :36 mm of diameter is inserted in a slit located on one side of the waveguide17was used, as shown in Fig. 1. Varying the position of the rod along the slit and its length inside the waveguide allowed the tuning of the volume, resonance frequency and electromagnetic losses of the cavity. For the experiments, the TE 109mode with !c=2= 11:69 GHz (volume Vc= 32:37 cm3) was chosen18. The cavity and the YIG spheres were excited by a vector network analyzer, which was also used as a detector to obtain the resonance spectra for dierent applied elds through the S11re
ection coecient. In order to observe the strong coupling regime, the spheres were placed on the shorted end of the waveguide resonator where the amplitude of the RF magnetic eld is maximum. The two spheres, which will be called YIG 1and YIG 2hereafter, have a radiusR1= 0:620:01 mm and R2= 0:610:01 mm, respectively. They were placed so that the center line (or axis) generated by the two spheres was parallel to the direction of the external DC magnetic eld H0. The center-to-center distance dof the spheres was varied from 1:41 mm to 3 :58 mm. The strong coupling regime is observed when the cou- pling constant gexceeds both the cavity losses cand the magnetic losses m6. This is illustrated in Fig. 2arXiv:2009.04557v1 [cond-mat.mtrl-sci] 9 Sep 20202 for YIG 1, where the hybridization of the cavity photonic mode and the ferromagnetic uniform mode of resonance is observed. In this work, the rod insertion was adjusted to haveccomparable to min order to facilitate the observation of the coupling. The coupling gis obtained by subtracting the resonance frequency of both modes for the whole range of magnetic elds, whereas the min- imum value is equal to 2 g. The value of the magnetic eld corresponding to this minimum will be called Hc. III. RESULTS A coupling constant of g1=2= 29:2 MHz and g2=2= 28:5 MHz was independently extracted for YIG 1and YIG 2, respectively. This agrees well with the theoreti- cal value given by5 g=r Vs Vc!M!c 21=2 ; (1) whereVsis the volume of the sphere, !M=0j
jMs withMsthe saturation magnetization,
the gyromag- netic ratio, and represents the spatial overlap between the cavity photonic mode and the magnonic mode. The factoris given by19 =1 hmaxmmaxVsZ sphere(hm)dV(2) where his the dynamic magnetic eld of the cavity with hmaxbeing its maximum magnitude and mis the dy- namic magnetization of the sphere with mmaxbeing its maximum magnitude. The value of is usually equal to 1 when handmare both uniform, which is the case for a small sample placed at the maximum of the cavity eld. The input-output formalism6was used to extract the losses of each component. The losses of the cavity c=2 were8:65 MHz, similar to the losses m1=2= 8:44 MHz (YIG 1) andm2=2= 12:63 MHz (YIG 2) of the YIG spheres. The mean magnetic losses of both spheres, equal to 10:54 MHz, will be referred to as m1hereafter. With two spheres in the cavity, the hybridization of the modes is still exhibited, but accompanied with a shift in the value of Hcand a change in the coupling strength as the spheres are brought closer. This is shown in Fig. 3 for two values of d. The eld shift, due to the dipo- lar interaction, can be calculated by solving the coupled Landau-Lifshitz equations of motion of the two spheres treated as macrospins: @ @t M1 M2 = 0j
j M1 M2 H+N M2 M1 (3) where N=1 3R d32 4 1 0 0 0 1 0 0 0 23 5; (4) 0.41 0.415 0.4211.511.611.711.811.9 Magnetic Field (T)Frequency (GHz) Cavity FMR g/2π = 29.2 MHz Hc = 0.4137 TFIG. 2. Strong coupling spectra obtained for the sphere YIG 1with the setup described in Sect. II. The extracted cou- pling constant is g1=2= 29:2 MHz. The hybridization of the modes occurs at a eld Hc= 0:4137 T. When there is no coupling, the resonance frequency of the cavity and the YIG sphere is represented by the red dashed line and the blue dot- ted line, respectively. 0.41 0.415 0.4211.511.611.711.811.9 Magnetic Field (T)Frequency (GHz)g/2π = 40.5 MHz Hc = 0.4133 T(a) 0.395 0.4 0.40511.511.611.711.811.9 Magnetic Field (T)Frequency (GHz)g/2π = 20 MHz Hc = 0.3983 T(b) FIG. 3. Strong coupling between the microwave cavity and two YIG spheres placed at a mutual distance of (a) d= 3:58 mm and (b) d= 1:41 mm. The dipolar interaction between the spheres shifts the value of Hcand decreases the total coupling constant g. H=H0^z+h,Ris the mean radius of the spheres, considered identical, and dis the distance between the macrospins. Using a small signal approximation, the cou- pled equations yield the resonance condition !res=!0+R d3 !M (5) where!0=0j
jH0. Because the hybridization of the modes occurs for !res=!c, Eq. (5) shows that smaller distancesdlead to smaller values of Hc. This shift of Hcwas used to corroborate and correct the distances between the spheres, which were initially measured man- ually with a digital micrometer. A good agreement has been found between the two methods. The reduction of the coupling constant g, exhibited in Figure 3 as the spheres are brought closer, is reported in greater details in Fig. 4 (closed circles). For large dis- tances between the spheres, one expects from the input- output formalism20a total coupling strength of approxi- matelyp g2 1+g2 2(dotted line), which is indeed observed. However, for smaller distances d, the coupling constant3 11.522.533.5 d/2R2025303540g (MHz) ind. spin dip. narrowing(g12+g22) FIG. 4. Eect of dipolar interactions on the total coupling strengthgobtained experimentally (closed circles). Light gray curve: Expected decrease in the case of independent spins calculated with Eq. (8). Dark gray curve: Expected decrease in the case of dipolar narrowing calculated with Eq. (12). Dotted line: Coupling constant when d! 1 . 11.5 22.5 3 d/2R6810121416Mean losses (MHz)ind. spin dip. narrowingm FIG. 5. Magnetic losses mobtained experimentally as the spheres get closer (closed circles). Light gray curve: Expected increase in linewidth when considering independent spins ex- tracted from the susceptibility calculated with Eq. (8). Dark gray curve: m1+!where !is calculated using Eq. (11). Dotted line: Magnetic losses when d! 1 . is observed to decrease sharply from 40 :5 MHz down to 20 MHz. Considering the two YIG spheres as a whole, the usual expression of the S 11re
ection coecient, calculated from the input-output formalism, was used to extract the magnetic losses of the two spheres as a function of the distance between the spheres m(d) (closed circles in Fig. 5). In contrast with g, the magnetic losses increase sharply as the spheres get closer. For a distance d= 1:41 mm (d=2R= 1:13), the magnetic losses are just above 16 MHz, which is near the coupling strength of 20 MHz. For shorter distances, the magnetic losses would continue to increase while the coupling constant would decrease, causing the system to exit the strong coupling regime.IV. DISCUSSION In order to explain the reduction of the coupling con- stant, let us consider the impact of the dipolar interaction on. The dipolar eld can be separated into two com- ponents. A dominant non-uniform static component is added to the applied static eld and tend to spread the local eld on the spheres. A weaker non-uniform dynamic eld is further added to the cavity pumping eld, which could result in the excitation of non-uniform resonance modes. Assuming that the RF magnetic eld of the un- perturbed cavity is uniform, we can rewrite in terms of the uniform mode susceptibility using m=h. Since the real part of the susceptibility 00 near resonance, we keep only the imaginary part and rewrite Eq. (2) as =1 00maxVsZ sphere00dV=h00i 00max; (6) where the brackets hirepresents the mean value over the volume of the sphere. Further insights are provided by examining two lim- iting cases. Case 1 corresponds to the macrospin ap- proximation, in which all spins in a sphere are strongly coupled and locked parallel to each other's, which was as- sumed earlier in Eq. (3). Our calculations indicate that the dynamic part of both spheres will be in phase, re- sulting in a constant factor = 1 for any distance d. In Fig. 4, the macrospin approximation corresponds to the dotted line and a value of g=p g2 1+g2 2. Likewise, the macrospin approximation does not lead to an increase in the linewidth observed in Fig. 5 but rather gives a con- stant linewidth of m1(dotted line). In contrast, Case 2 assumes fully independent spins, that is, no long-range dynamic dipolar interaction and each spin constituent of the spheres is resonating at its own frequency depending on the value of its lo- cal static magnetic eld. This non-uniform magnetic eld, assumed to be along the ^ z-direction, is given by Hz=H0+Hdip., where Hdip.=R3(r2(3 cos2# 1) + 4drcos#+ 2d2) 3(r2+ 2drcos#+d2)5=2Ms:(7) Here,Hdip.is the static dipolar magnetic eld and the variablesrand#determine the position in a spherical coordinates system centered on a sphere placed at a dis- tancedfrom the source dipole. One can then numerically compute the probability density function f(Hdip.) over the volume of the sphere as a function of dto calculate the value of the mean susceptibility of the independent spins ensemble at resonance ( !=!c). Assuming no magnetic anisotropy, we have h00i=Z Hzm1!M (0j
jHz !c)2+ 2m1f(Hz)dHz;(8) which can be substituted in (6) and then (1) to calcu- late the coupling. In this limiting case, a strong decrease4 inis predicted, even for spheres separated by a rela- tively large distance d, as shown by the light gray curve in Fig. 4. Furthermore, the inhomogeneously broad- ened linewidth in the independent spins approximation is given by the light gray line in Fig. 5, which predicts a much broader linewidth than observed experimentally. In our two spheres experiment, we thus fall somewhere between these two limits: macrospin and independent spins. A more rigorous approach should include long- range dynamic dipolar interactions which are known to produce a phenomenon called \dipolar narrowing" in the literature21. We consider the original approach used by Clogston22in which the Landau-Lifshitz equation of mo- tion is solved for a non-uniform magnetic eld expanded in Fourier components as Hz=X kHkeikr: (9) Assuming the eld inhomogeneity is low with respect to the sample dimensions, we can neglect the terms related to the exchange interaction in the equation of motion, but consider the terms associated with dynamic dipolar elds. Further expanding the magnetization in Fourier series and by following a procedure similar to Ref. 22, we can derive an analytical expression for the imaginary part of the susceptibility of the uniform mode of resonance, ac- counting for the coupling between the uniform mode and the long wavelength spin wave modes, a process called two-magnon scattering23. With some simplications, it can be written in the form h00i=(m1+ !)!M (! !c)2+ (m1+ !)2(10) where != 2Var(!dip.) !M 1 +1 2!M 3!c !M2 2 3 !c 3!c !M 1=2 (11) is an additional loss term directly related to the variance of the static dipolar magnetic eld through the quantity !dip.=0j
jHdip., which can be calculated analytically (Appendix A). In the expression of !, the division by !Mrepresents the dipolar narrowing eect. This addi- tional loss term is added to m1, which yields a total loss term that can be compared with the measured mean losses of the magnetic system. As shown by the dark gray curve in Fig. 5, the general trend of the data is re- produced relatively well. Regarding the coupling constant g, the denition of in Eq. (6) is extended to account for the fact that spins, whose resonance frequency 0j
jHzis detuned from the resonance frequency of the cavity !c, can contribute to the coupling with the cavity. This can be achieved by introducing a weight function in the denition of sothat the spins whose resonance frequency is contained in- side the coupling range ( garound!c), have a stronger contribution (high energy exchange) to the total cou- pling than those whose resonance frequency falls outside the coupling range (low energy exchange). In contrast, in Eq. (6), only the spins resonating at frequency !c contribute, whereas the remaining spins (detuned from the cavity) do not contribute to the coupling. To in- clude this phenomenon, we use a weight function con- sisting in a Lorentz distribution L(!c;gmax) centered at !=!cand having a half-width at half maximum of gmax=p g2 1+g2 2. We thus have =Z1 0(m1+ !)!ML(!c;gmax) (! !c)2+ (m1+ !)2d! Z1 0m1!ML(!c;gmax) (! !c)2+ 2m1d!; (12) which equals unity if != 0, in absence of dipolar broadening. Equation (12) may be used with Eq. (1) to generate the dark gray curve in Fig. 4. The excellent agreement with the experimental data supports that the observed decrease in the coupling rate between the sys- tem of magnetic spheres and the cavity, as the spheres are brought closer together, originates from the increasingly non-uniform dipolar static magnetic eld on each sphere. It also shows that the long-range dynamic dipolar inter- action within each sphere, which gives rise to the dipolar narrowing eect, somewhat limits the adverse eect of the non-uniform eld distribution. Similarly, the expression of given in (12) implies that a larger coupling gmaxtends to smooth out the adverse eect of a given dipolar broadening !in reducing the total coupling strength. V. CONCLUSION We have demonstrated that the dipolar interaction be- tween two ferromagnetic objects can strongly aect their coupling with a microwave cavity. As the distance be- tween the spheres is gradually reduced, dipolar interac- tions force the spins to resonate at increasingly dierent frequencies. This results in increased magnetic losses and decreased coupling strength gof the system. A model based on inhomogeneous broadening with dipolar nar- rowing reproduces the main features observed on a sys- tem consisting of two YIG spheres in a tunable microwave cavity. While the reduction in the coupling strength can be linked with the variance of applied eld caused by the dipolar interaction, this eect is attenuated by dipo- lar narrowing and by strong coupling of each individual sphere with the cavity. Our results suggest that a number of Nindividual fer- romagnetic objects inserted in an electromagnetic cavity will eventually exhibit a reduced coupling as compared to the expected g/p Nbehavior as the density is in- creased. Yet the dipolar broadening will be mitigated5 by a compensating dipolar narrowing eect. A trade-o must be found to determine the optimal density of fer- romagnetic objects to be placed in the cavity to reach a maximum coupling strength while reducing the impact of dipolar interaction.Appendix A: Analytical expression for Var (!dip.). Integrating by parts Eq. (7), we have h!dip.i=a3 12!M (A1) wherea= 2R=d (0a1). The integration by parts also leads to an analytical expression for h!2 dip.i. The denition of the variance, Var( !dip.) =h!2 dip.i h!dip.i2, then gives Var(!dip.) !2 M=a3 4a 3(4 a2)3 5 +a2 2 1 +a2 8 +tanh 1(a=2) 24 a 43 32+a2 9 3(4 a2) 512ln2 +a 2 a :(A2) 1O. O. Soykal and M. Flatt e, Physical review letters 104, 077202 (2010). 2H. Huebl, C. W. Zollitsch, J. Lotze, F. Hocke, M. Greifen- stein, A. Marx, R. Gross, and S. T. Goennenwein, Physical review letters 111, 127003 (2013). 3Y. Cao, P. Yan, H. Huebl, S. T. Goennenwein, and G. E. Bauer, Physical Review B 91, 094423 (2015). 4B. Yao, Y. Gui, Y. Xiao, H. Guo, X. Chen, W. Lu, C. Chien, and C.-M. Hu, Physical Review B 92, 184407 (2015). 5Y. Tabuchi, S. Ishino, T. Ishikawa, R. Yamazaki, K. Usami, and Y. Nakamura, Physical review letters 113, 083603 (2014). 6X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, Physical review letters 113, 156401 (2014). 7P. Hyde, L. Bai, M. Harder, C. Dyck, and C.-M. Hu, Physical Review B 95, 094416 (2017). 8M. Goryachev, W. G. Farr, D. L. Creedon, Y. Fan, M. Kostylev, and M. E. Tobar, Physical Review Applied 2, 054002 (2014). 9Y. Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Ya- mazaki, K. Usami, and Y. Nakamura, Science 349, 405 (2015). 10N. Lambert, J. Haigh, S. Langenfeld, A. Doherty, and A. Ferguson, Physical Review A 93, 021803 (2016). 11X. Zhang, C.-L. Zou, N. Zhu, F. Marquardt, L. Jiang, andH. X. Tang, Nature communications 6(2015). 12L. Bai, M. Harder, P. Hyde, Z. Zhang, C.-M. Hu, Y. Chen, and J. Q. Xiao, Physical Review Letters 118, 217201 (2017). 13Y. Cao and P. Yan, Phys. Rev. B 99, 214415 (2019). 14J. T. Hou and L. Liu, Physical Review Letters 123, 107702 (2019). 15M. Harder and C.-M. Hu, in Solid State Physics , Vol. 69 (Elsevier, 2018) pp. 47{121. 16D. Zhang, X.-M. Wang, T.-F. Li, X.-Q. Luo, W. Wu, F. Nori, and J. You, npj Quantum Information 1, 1 (2015). 17A. Morin, C. Lacroix, and D. M enard, in 2016 17th Inter- national Symposium on Antenna Technology and Applied Electromagnetics (ANTEM) (IEEE, 2016) pp. 1{2. 18D. M. Pozar, Microwave engineering (John Wiley & Sons, 2009). 19N. Lambert, J. Haigh, and A. Ferguson, Journal of Applied Physics 117, 053910 (2015). 20D. Schuster, A. Sears, E. Ginossar, L. DiCarlo, L. Frunzio, J. Morton, H. Wu, G. Briggs, B. Buckley, D. Awschalom, et al. , Physical review letters 105, 140501 (2010). 21S. M. Rezende and A. Azevedo, Physical Review B 44, 7062 (1991). 22A. Clogston, Journal of Applied Physics 29, 334 (1958). 23R. D. McMichael, D. Twisselmann, and A. Kunz, Physical review letters 90, 227601 (2003). | 2020-09-09 | The strong photon-magnon coupling between an electromagnetic cavity and two
yttrium iron garnet (YIG) spheres has been investigated in the context of a
strong mutual dipolar interaction between the spheres. A decrease in the
coupling strength between the YIG spheres and the electromagnetic cavity is
observed, along with an increase of the total magnetic losses, as the distance
between the spheres is decreased. A model of inhomogeneous broadening of the
ferromagnetic resonance linewidth, partly mitigated by the dipolar narrowing
effect, reproduces the reduction in the coupling strength observed
experimentally. These findings have important implications for the
understanding of strongly coupled photon-magnon system involving densely packed
magnetic objects, such as ferromagnetic nanowires arrays, in which the total
coupling strength with an electromagnetic cavity might become limited due to
mutual dipolar interactions. | Effect of dipolar interactions on cavity magnon-polaritons | 2009.04557v1 |
arXiv:2102.12181v1 [quant-ph] 24 Feb 2021Phase-controlled pathway interferences and switchable fa st-slow light in a cavity-magnon polariton system Jie Zhao,1, 2, 3, 4, ∗Longhao Wu,1, 2, 3,∗Tiefu Li,5, 6Yu-xi Liu,5 Franco Nori,7, 8Yulong Liu,9, 10,†and Jiangfeng Du1, 2, 3,‡ 1Hefei National Laboratory for Physical Sciences at the Microscale and Department of Modern Physics, University of Science and Technology of China, Hefei 230026 , China 2CAS Key Laboratory of Microscale Magnetic Resonance, University of Science and Technology of China, Hefei 230026 , China 3Synergetic Innovation Center of Quantum Information and Qu antum Physics, University of Science and Technology of China, Hefei 230026 , China 4National Laboratory of Solid State Microstructures, School of Physics, Nanjing University, Nanjing 210093, Chi na 5Institute of Microelectronics, Tsinghua University, Beij ing 100084, China 6Quantum states of matter, Beijing Academy of Quantum Information Sciences, Beijing 100193, China 7Theoretical Quantum Physics Laboratory, RIKEN, Saitama, 3 51-0198, Japan 8Department of Physics, The University of Michigan, Ann Arbor, Michigan 48109-1040, USA 9Beijing Academy of Quantum Information Sciences, Beijing 1 00193, China 10Department of Applied Physics, Aalto University, P .O. Box 15100, FI-00076 Aalto, Finland (Dated: February 25, 2021) 1Abstract We study the phase controlled transmission properties in a c ompound system consisting of a 3D copper cavity and an yttrium iron garnet (YIG) sphere. By tuning the relative phase of the magnon pumping and cavity probe tones, constructive and destructive interfer ences occur periodically, which strongly modify both the cavity field transmission spectra and the group dela y of light. Moreover, the tunable amplitude ratio between pump-probe tones allows us to further improve the si gnal absorption or amplification, accompanied by either significantly enhanced optical advance or delay. B oth the phase and amplitude-ratio can be used to realize in-situ tunable and switchable fast-slow light. The tunable phase and amplitude-ratio lead to the zero reflection of the transmitted light and an abrupt fast-s low light transition. Our results confirm that direct magnon pumping through the coupling loops provides a versatile route to achieve controllable signal transmission, storage, and communication, which can be fur ther expanded to the quantum regime, realizing coherent-state processing or quantum-limited precise mea surements. I. INTRODUCTION Interference, due to superposed waves, plays a considerabl e role in explaining many classical and quantum physical phenomena. Based on the phase-differe nce-induced interference patterns, ultraprecise interferometers have been created, impactin g the development of modern physics and industry [1]. In addition to the phases, waves or particles p ropagating through different path- ways can also introduce interference patterns. Among vario us types of multiple-path-induced interference, the Fano resonance [2] and its typical manife stations, the electromagnetically in- duced transparency (EIT) and electromagnetically induced absorption (EIABS) [3, 4], are the most well-known ones. The Fano resonance and EIT-like (or EI ABS-like) line shapes are not only experimentally observed in quantum systems but also in vari ous classical harmonic-resonator sys- tems. Quantum examples include quantum dots [5], quantum we lls [6], superconducting qubits [7– 10], as well as Bose-Einstein condensates [11]. Classical e xamples [12] include coupled optical cavities [13–16], terahertz resonators [17, 18], microwav e resonators [19, 20], mechanical res- onators [21, 22], optomechanical systems [23]. However, wh ether in quantum or in classical systems, the Fano resonance, EIT- or EIABS-like spectra are normally experimentally realized ∗These authors contributed equally to this work †liuyl@baqis.ac.cn ‡djf@ustc.edu.cn 2separately. The switchable electromagnetically induced t ransparency and absorption, as well as fast and slow light, have been proposed using dressed superc onducting qubits [8], hybrid optome- chanical system [24, 25], dark-mode breaking [26–28], and s o on. Particularly, there appears growing interest to control the EIT and EIABS by introducing exceptional points [29–31]. Photon stops [32, 33], chiral EIT [34], and infinite slow light [33] h ave recently been realized around exceptional points. Motivated by their potential applicat ions in rapid transitions between fast and slow light, which facilitate coherent state storage and retrieval, it is highly desirable to have experimental realizations of in situ tunable and switchable absorption, transparency, and even am- plification. Meanwhile, cavity magnon polaritons in an yttrium-iron-ga rnet (YIG) sphere-cavity coupled system has attracted much attention due to its strong [35–42 ] and even ultrastrong couplings [43– 45]. The compatibility and scalability with microwave and o ptical light enable magnons to be a versatile interface for different quantum devices [46–51] . At low temperatures, strong coupling between magnons, superconducting resonators and qubits ha ve been demonstrated [52–56]. Sub- sequently, the EIT-like magnon-induced transparency (MIT ) or the EIABS-like magnon-induced absorption (MIABS) of the transmitted cavity field were obse rved for different external coupling conditions [57]. The underlying mechanism is attributed to interferences between two transition pathways, i.e., the direct cavity pathway and the cavity-ma gnon-cavity pathway, to transmit the probe field. In addition to the coupling strength [57] and frequency detu ning [58–60] between coupled modes, phases play a vital role in wave interference control . We thus focus on the controllabil- ity of pathway interferences through the phase difference b etween the cavity-probe tone and the magnon-pump tone, which is introduced by the coupling loops ’ technology [61–64]. The direct magnon pump is becoming useful in realizing the light-wave i nterface [46–48], enhancing the Kerr nonlinearity [65–67], and has also been adopted to observe t he magnetostriction-induced quantum entanglement [68–72], among other applications. Together with the cavity-probe tone, a magnon-pump tone int roduces a controllable relative phase to the system, and thus the path interference can be rea l-time controlled. Changing the two- tone phase difference, we can switch the cavity-probe spectra from the original magnon-i nduced transparency instantly to the magnon-induced absorption, or even the Fano line shape . Further- more, the tunable pump-probe amplitude ratio allows us to fu rther improve the signal absorption, transparency, or amplification, accompanied by a significan t enhancement by nearly 2 orders of 3magnitude of the optical advance or delay time compared to the case with out magnon pump [57]. In particular, the tunable phase and amplitude ratio also le ad to the zero reflection of the trans- mitted light, which is accompanied by an abrupt transition o f delay time. Our results confirm that direct magnon pumping provides a versatile route to con trol signal transmission, storage, and communication, and can be further expanded to coherent stat e processing in the quantum regime. -150 -75 0 75 150 p(MHz)-4-20S11(dB) -150 -75 0 75 150 p(MHz)-8-6-4-2S11(dB)Port 1 Port 2 1 32Q IL R , AB Splitter IQ Mixer CirculatorVNAAWG CavityMagnon gCavity gMagnon High Low(a) (b) (c) (d)xy z Destructive C t ti 4FIG. 1. Measurement setup and phase-induced interference m echanism diagrams. (a) The system consisting of a three-dimensional (3D) copper cavity and a YIG sphere, w hich is coherently pumped by the coupling loops shown as a black coil surrounding the YIG sphere. The re d arrows and colors indicate the magnetic field directions and amplitudes of the TE 101mode distribution, respectively. The YIG sphere is placed a t the area with maximum magnetic field distribution inside a 3D copper cavity box to obtain a strong cavity- magnon coupling. A small hole at the cavity sidewall is assem bled with a standard SubMiniature version A connector (SMA connector), allowing us to do the reflection m easurement S11of the probe field, i.e., such a SMA connector works as both the signal input and readout port . A beam of coherent microwave comes out from port 1 of the vector network analyzer (VNA) and splits in to two beams, working as the magnon-pump tone and the cavity-probe tone. Here, we use an in-phase and q uadrature mixer (I-Q mixer) and an arbitrary waveform generator (AWG) to control and tune the phase diffe renceϕand pump-probe amplitude ratio δ=εm/εcbetween the pump and probe tones. The interfering results ar e extracted by the circulator and finally transferred to port 2 of the VNA. (b) Diagram showing t he relative phase between the magnon pump and cavity probe in the cavity-magnon coupled system. (c) Th e corresponding energy-level diagram. Two transition pathways to the higher energy level: 1/circleco√yrtprobe-tone-induced direct excitation, and 2/circleco√yrtpump-tone excites magnons and then coherently transfers there to cavi ty photons. (d) Measurements of the reflection spectraS11versus the detuning ∆p=ωc−ωp=ωm−ωp. The relative phase difference between pump and probe tones can be developed to realize an in situ switchable constructive and destructive interference, presented as MIABS with ϕ= 0.35π,δ= 1.2and MIT with ϕ= 1.35π,δ= 1.2. II. EXPERIMENTAL SETUP As shown in Fig. 1(a), our system consists of a 3D copper (Cu) c avity with an inner dimension of40×20×8mm3and an YIG sphere with a 0.3 mm diameter. A static magnetic fiel dHstatic applied in the x-yplane tunes the magnon frequency. The simulated cavity-mod e magnetic field distribution is shown at the bottom of Fig. 1(a), where the ar rows and colors indicate the cavity mode magnetic field directions and amplitudes. The YIG spher e is placed near the magnetic field antinode of the cavity TE101mode. The magnetic components (along the zaxis) of the microwave field at this antinode is perpendicular to the static magneti c bias field. Here, we are only interested in the low excited states of the K ittel mode, in which all the spins precess in phase. Under the Holstein-Primakoff transforma tion, such collective spin mode can be 5simplified to a harmonic resonator, which introduces the mag non mode. In our setup, the cavity mode couples to the magnon mode with coupling strength g= 7.6 MHz , which is larger than the magnon decay rate κm= 1.2MHz , but smaller than the cavity decay rate κc= 113.9MHz . In our experiment, a beam of coherent microwave is emitted fr om port 1 of a VNA and then divided through a splitter into two beams, one of which is use d to probe the cavity (probe tone) and another beam is used to pump the magnon (pump tone) by inco rporating the coupling loop technique, which is schematically shown in the dashed recta ngle of Fig. 1(a). The probe tone is injected into the cavity through antenna 1, which induces th e cavity external decay rate κc1= 21.8 MHz . The pump tone is injected through antenna 2, which introduc es the magnon external decay rate κm1= 0.6 MHz . Note that the phase ϕc= 0 and amplitude εcof the probe tone are fixed (i.e., working as a reference), and the phase ϕand amplitude εmof the magnon-pump tone are tunable and controlled by an arbitrary wave generator wi th an in-phase and quadrature mixer (I-Q mixer). III. MODEL By considering the cavity-magnon coupling, as well as the pu mp and probe tones [model in Fig. 1(b)], the system Hamiltonian becomes H=ωca†a+ωmm†m+g(a†m+m†a) +i/radicalbig 2ηcκcεc/parenleftbig a†e−iωpt−aeiωpt/parenrightbig +i/radicalbig 2ηmκmεm/parenleftbig m†e−iωpt−iϕ−meiωpt+iϕ/parenrightbig . (1) Here,a†(a) andm†(m) are the creation (annihilation) operators for the microwa ve photon and the magnon at frequencies ωcandωm, respectively, and we choose units with /planckover2pi1= 1. The magnon frequency ωmlinearly depends on the static bias field Hstaticand is tunable within the range of a few hundred MHz to about 45 GHz; εc(εm) is the microwave amplitude applied to drive the cavity (magnon). Here, we introduce the coupling parameter ηc=κc1/κc, (2) ηm=κm1/κm (3) 6to classify the working regime of the cavity (the magnon). Th e parameter ηc(ηm) classifies three working regimes for the cavity (magnon) into three types: ov ercoupling regime for ηc(ηm)>1/2; critical-coupling regime for ηc(ηm) = 1/2; and undercoupling regime for ηc(ηm)<1/2. In our experiment, the cavity works in the undercoupling regime ( ηc<1/2) and the magnon works in the critical coupling regime ( ηm= 1/2). Experimentally, the reflection signal from the cavity is cir culated and then transferred to port 2 of the VNA to carry out the spectroscopic measurement, whic h corresponds to the steady-state solution of the Hamiltonian Eq. (1). The transmission coeffi cienttpof the probe field is defined as the ratio of the output-field amplitude εoutto the input-field amplitude εcat the probe frequency ωp:tp=εout/εc. With the input-output boundary condition, εout=εc−/radicalbig 2ηcκc/angbracketlefta/angbracketright, (4) we can solve the transmission coefficient tpof the probe field as [73] tp=tprobe+tpump, (5) with tprobe= 1−2ηcκc(i∆p+κm) (i∆p+κc)(i∆p+κm)+g2, (6) tpump=ig√2ηcκc√2ηmκmδe−iϕ (i∆p+κc)(i∆p+κm)+g2. (7) Here∆pis the detuning between the probe frequency ωpand either the cavity resonant frequency ωcor the magnon frequency ωm. In our experiment, the cavity is resonant with the cavity, i .e., ∆p=ωc−ωp=ωm−ωp; (8) and δ=εm/εc (9) is the pump-probe amplitude ratio. Equation (5) clearly sho ws that the transmission coefficient can be divided into two parts: 1.tprobe in Eq. (6), the contribution from the cavity-probe tone, rep resents the traditional pathway-induced interference; 2.tpump in Eq. (7), the contribution from the magnon-pump field, affe cts the interference and modifies the transmission of the probe field. 7As shown in Fig. 1(c), there exist two transition pathways fo r the cavity: the probe-tone-induced direct excitation, and the photons transferred from magnon excitations. When the cavity decay rate (analog to broadband of states) is much larger than the m agnon decay rate (analog to a nar- row discrete quantum state in other quantum systems), Fano i nterference happens and has been successfully used to explain the MIT and MIABS phenomenon in cavity magnon-polariton sys- tems [57]. Besides pathway-induced interference, the stee red phase ϕof the wave provides another useful way to generate and especially control the interfere nces, as shown in Fig. 1(d). We emphasize that in this paper we focus on how the phase difference ϕand pump-probe ratio δ=εm/εcaffect the interference, and we explore its potential appli cations, such as controllable field transmission and in situ switchable slow-fast light . TheS11spectrum and group-time delay measurement are carried out on the VNA and then fitted by T=|tp| (10) and τ=−∂[arg(tp)] ∂∆p, (11) respectively. IV . PHASE INDUCED INTERFERENCE AND CONTROLLABLE MICROWA VE FIELD TRANS- PORT We first study how the phase of the magnon-pump tone affects th e transmission of the cavity- probe field. In Fig. 2 (a), we present experimental results of the transmission, when the pump- probe ratio is δ=εm/εc= 1.7. In this setup, the phase ϕis continuously increased from 0 to 2 π using an I/Q mixer, and is shown in the xaxis of Fig. 2 (a). Then we conduct the S11measurements and the recorded spectra are plotted versus the detuning fre quencies ∆p. The colors represent the relative steady-state output amplitude (in dB units) at dif ferent frequency and pump-probe ratios. Figure 2(a) shows that the interference mainly happens arou nd∆p= 0 and can be controlled in situby changing the phase ϕ. As shown in Fig. 2(b), where ϕis set to0.35π, destructive interference happens and an obvious dip appears around ∆p= 0. This behavior can be regarded as MIABS. However, if we set ϕ= 8Theory Experiment -150 75 0 75 150 (a) (b) 0 -6.5 -130 -4.5 -9 2 -1.5 -51 -2.5 -6 -150 150-150-75 15075 FIG. 2.S11spectrum versus relative phase difference ϕ. (a) Measured transmission spectrum S11versus phaseϕand detuning ∆p. The colors indicate the transmitted amplitudes in dB units . (b) Measured output spectrum S11with phases: 1/circleco√yrtϕ= 0.35π,2/circleco√yrtϕ= 0.85π,3/circleco√yrtϕ= 1.35π, and 4/circleco√yrtϕ= 1.85π. Here, the pump-probe amplitude ratio is fixed at δ= 1.7. Red-solid lines are the corresponding theoretical result s. 1.35π, constructive interference happens and an obvious amplific ation window appears around ∆p= 0. This behavior can be described as magnon-induced amplifica tion (MIAMP). When ϕ is set to0.85πor1.85π, sharp and Fano-interference-like asymmetry spectra are o bserved even when the cavity and magnon are exactly resonant. Although the interference originates from the coherent cav ity-magnon coupling, Fig. 2 clearly shows that the phase ϕplays a key role in realizing an in situ tunable and controlla ble interfer- ence (e.g., constructive or destructive interference) , which can be further engineered to control the probe-field transmission. Note that in previous studies [57 ] MIABS was only observed in the cav- 9ity overcoupling regime (i.e., ηa>1/2) and MIT was only observed in the cavity undercoupling regime (i.e., ηa<1/2). In contrast to this, here we realize a phase-dependent and switchable MI- ABS and MIT, as well as MIAMP in a fixed undercoupling regime ( ηc= 0.19in our experiment). We emphasize that the destructive interference-induced MI ABS is a unique result of phase mod- ulation. The observed asymmetric Fano line shapes could be u seful to realize Fano-interference sensors or precise measurements, using the magnon-pump met hod realized in our work. V . AMPLITUDE RATIO OPTIMIZED MAGNON-INDUCED-ABSORPTION Recall the magnon-pump transmission coefficient tpump in Eq. (7). There, the phase ϕdeter- mines the type of interference, e.g., constructive or destr uctive. However, the pump-probe ratio δ=εm/εcalso affects the degree of interference, and thus can be used to control the probe-field transmissions tp. As shown in Fig. 3(a), a color map is used to present the exper iment results. Along the xaxis, the amplitude ratio δis continuously increased from 0 to 6.5, by changing the overall voltage amplitude applied to the I and Q ports of an I- Q mixer. Then we conduct the S11 measurements and the steady-state output-field amplitudes are plotted versus the frequency de- tuning∆p. The colors in Fig. 3(a) represent the relative strength of t he steady-state output field (in dB units) at a different frequency. Here, the chosen phas eϕ= 0.35πresults in MITs when δ <0.32, while MIABSs dominate the output response in the regime δ >0.32. We then study how the pump-probe ratio δaffects the central absorption window of the S11spectra. Figure 3(a) shows that interference occurs around ∆p= 0and is in situ controlled by changing the pump-probe ratio δ. The center blue-colored area represents an ideal absorpti on (transmission T <0.01) of the probe field. Figure 3 (b) shows the extreme values of the transmission coe fficients around ∆p= 0 versus the pump-probe ratio δ. In the yellow area, we find the local maximum values of the MIT s, and the local minimum values are found for MIABSs in the blue area . An obvious dip appears around δ= 3 and the minimum transmission value is less than 1% (voltage a mplitude ratio), which corresponds to an optimized and ideal probe-field absorptio n. Figure 3(c) shows the evolution process from MIT to MIABS by g radually increasing the pump-probe ratio δ. Whenδ= 0, corresponding to case 1/circleco√yrtof Fig. 3(c), our scheme recovers the traditional MIT case when no magnon pump is applied. When the magnon pump is introduced and its strength is continuously increased, the transparen cy window disappears and is replaced by 10-150 -75 0 75 150Theory -150 -75 0 75 150Experiment(a)( Extreme Amplitude (dB) -1.0 -2.5 -4.00 -20 -40 -60 0 2 4 6 -1.0 -2.5 -4.0 0 -20 -42 -150-75 1501.0 -4.0 -9.0 -150-75 150 FIG. 3. Measured transmission spectrum S11versus pump-probe amplitude ratio δwith phase fixed at ϕ= 0.35π. (a) Measured output spectrum versus amplitude ratio δand detuning ∆p. The colors indicate transmitted power in dBs. (b) The extreme values of the S11transmission spectra of the output field versus the amplitude ratio parameter δ. In the light-yellow (light-blue) regime, the extreme valu es represent the maximum (minimum) transmission amplitudes of the peaks (di ps) around ∆p= 0. (c) Measured transmis- sion spectrum S11with amplitude ratio: 1/circleco√yrtδ= 0,2/circleco√yrtδ= 0.3,3/circleco√yrtδ= 3.0, and 4/circleco√yrtδ= 5.7. Red-solid lines are the corresponding theoretical results. an obvious absorption dip, as shown in cases 2/circleco√yrtand 3/circleco√yrtof Fig. 3(c). With an even larger pump- probe ratio, the MIABS dips become asymmetry gradually, suc h as the spectrum in the case 4/circleco√yrtof Fig. 3(c). Comparing with other results in Fig. 3(c), we can fi nd that the experimental data do no fit so well with the theory in case 4/circleco√yrtof Fig. 3(c). This is induced by the additional cavity-anten na 112 coupling. Due to the existence of this tiny coupling, the ma gnon pump signal also pumps the cavity. With a modest magnon-pump strength, the additional cavity pump does not affect the sys- tem seriously, so that the theory fit the experiment data well . With a relatively strong magnon pump, the side effects of the additional cavity pump become l arger, though it does not change the line shape. Therefore, the experiment data and theory do not fit so well when the magnon pump is relatively strong [73]. Similar phenomena can also be obser ved in the case 4/circleco√yrtof Fig. 4(c). We emphasize one main result of this paper: the absorption dips appear with an under-coupling coefficient of ηa= 0.19in our experiment. However, absorptions only happen in the o vercoupling regime in traditional cases . Moreover, Figs. 3(a) and (c) show that δcan be used to switch the transmission behavior from the magnon-induced transparen cy to the magnon-induced absorption . Note that the type of interference, destructive interferen ce or constructive interference, depends on the value of the phase ϕ. However, the interference intensity is determined and opt imized by the pump-probe ratio δ. As shown in Fig. 3(c), the dip of S11is 42 dB lower than the baseline. The dip amplitude is quite close to zero, which indicates tha t a zero reflection is generated by the destructive interference. VI. AMPLITUDE RATIO OPTIMIZED MAGNON-INDUCED-AMPLIFICAT ION We now study how the amplitude ratio of δ=εm/εcaffects the MIAMP. In this case, the phase is fixed at ϕ= 1.35π, where constructive interference dominates the transmiss ion of the output field. As shown in Fig. 4(a), a color map is used to present the m easurement results. Along the xaxis, the pump-probe ratio δis continuously increased from 0 to 6.5. Then we conduct the S11 measurement, and the steady-state transmission spectra ar e plotted versus the frequency detuning parameter ∆p. The colors in Fig. 4(a) represent the transmission amplitu des of the steady-state output field (in dB units) at different frequencies. We then s tudy how the amplitude δaffects the center amplification window of the S11spectra. Figure 4(a) clearly shows that constructive interference h appens around ∆p= 0and are in situ controlled by changing the pump-probe ratio δ. Magnon-pump-induced constructive interference happens when the probe field is nearly resonant with the cavit y (also the magnon), and amplifi- cation windows appear. Around ∆p= 0, the color changes from light blue to orange when the pump-probe ratio δincreases from 0 to 6.5. This indicates that the higher ampli fication can be obtained with a larger pump-probe ratio δ. 120 1 2 3 4 5 6-30369Extreme Amplitude (dB) Theory Experiment (c)Experiment Theory (b)(a) -150 -75 0 75 150-4.0-2.00 -150 -75 0 75 150-4.004.0-150 -50 0 75 150-4-2.5-1 -150 -75 0 75 150-4.0-2.00 FIG. 4. Measured transmission spectrum S11versus pump-probe amplitude ratio δ=εm/εcwith phase fixed atϕ= 1.35π. (a) Measured output spectra S11versus amplitude ratio δand frequency detuning ∆p. The colors indicate the transmitted amplitude in dB units. (b) The extreme values of the S11trans- mission spectra of the output field versus the amplitude-rat io parameter δ. The extreme values represent the maximum transmission amplitude of the peaks around ∆p= 0. (c) Measured transmission spectra S11 with amplitude ratios: 1/circleco√yrtδ= 0,2/circleco√yrtδ= 0.9,3/circleco√yrtδ= 1.2, and 4/circleco√yrtδ= 4.5. The red-solid lines are the corresponding theoretical results. Figure 4(b) shows how the peak values in the amplification win dow change versus the ampli- tude ratio δ. The amplification coefficient is monotonously dependent on the increment of the pump-probe ratio δ. Although the maximum pump-probe ratio is δ= 6.5in our experiment, we emphasize that a higher transmission gain can be obtained us ing a larger pump power. 13Figure 4(c) clearly shows the evolution of the transmission spectrum from MIT to MIAMP when we gradually increase the pump-probe ratio δ. Whenδ <1.2, an obvious transparency window appears. When δ= 1.2, the peak value of the transparency window equals the value o f the baseline, showing the ideal MIT phenomenon. Further inc reasing the pump strength, we can observe MIAMP. When δ= 4.5, an obvious amplification window appears, producing MIAMP. Note that the phase is fixed at ϕ= 1.35πto produce constructive interference. When the amplitude ratio is set to δ= 0, i.e., no magnon pump, our scheme also recovers the traditio nal case without a magnon pump and only MIT is observed. This resu lt is, of course, the same as case 1/circleco√yrtin Fig. 3(c). We point out another main result that the pump-probe ratio δcan be used to realize and control the magnon-induced amplifications . Figures 4(a) and 4(c) show that δcan be used to switch the system response from MIT to MIAMP. Note t hat the interference type, such as constructive interference discussed here, depends on th e value of the phase ϕ; however, the interference intensity is determined and optimized by the p ump-probe ratio δ. VII. SWITCHABLE FAST- AND SLOW-LIGHT BASED ON THE PHASE AND A MPLITUDE RATIO The group delay or advance of light always accompanies EIT or EIABS. In this experiment, we show that the group delay (slow light) and group advance (f ast light) can also be realized in our cavity magnon-polariton system. Similar to the discuss ions above, the phase ϕis the key parameter that determines the interference type, e.g., des tructive or constructive. Therefore, the phaseϕprovides a tunable and in situ switched group advance or delay of the probe field. The extreme values of the delay time are measured and presented i n Fig. 5, choosing the same phases ϕ= 0.35πandϕ= 1.35π, which are also used in Figs. 3 and 4, respectively. In Fig. 5(a), the phase is set to ϕ= 0.35π. When we increase the pump-probe ratio δ, a longer advance time is achieved, but immediately changes to time de lay when δ >3.0. Further increasing δreduces the delay time. In Fig. 5(c), we present the phase of t ransmission signals at different probe frequencies with δ= 2.7(case 1/circleco√yrt) andδ= 3.3(case 2/circleco√yrt). The phase changes drastically around∆p= 0with opposite directions. The drastic changes of the phase r esult in a long advance or delay time, while the phase-change direction reversal re sults in the sharp transition from time advance to time delay. Accompanying the sharp transition in Fig. 5(a), we observe the longest either delay or advance times. Therefore, the pump-probe ra tioδallows to optimize and switch the 140 2 4 6-1000-50005001000 0 2 4 6205080Extreme Delay Time (ns) Phase(c) -50 500 -50 50 0-101 Phase (rad) -44 0 FIG. 5. Measured time delay versus pump-probe ratio δfor the phase ϕ= 0.35π(a); andϕ= 1.35π(b). Light-yellow area indicates the group-delay regime, and th e light-blue area indicates the group-advance regime. (c) Measured unwrapped phase versus frequency detu ning∆pwithδ= 2.7[point 1/circleco√yrtin (a)] and δ= 3.3[point 2/circleco√yrtin (a)] for ϕ= 0.35π. probe microwave from fast to slow light, or inversely . Comparing the abrupt transition in Fig. 5(a) with the zero reflection discussed in Sec. V , we find that the de lay time abrupt transition and the zero reflection occur at the same parameter setup. It is notab le that the discontinuity and abrupt transition are always accompanied by the zero reflection in c oupled resonator systems. In Fig. 5(b), we set the phase to ϕ= 1.35πand mainly observe constructive interference. In this case , the time delay monotonously increases with the pump-probe ratio δ. Note that the pump-probe ratio used in Fig. 5(b) is not its limitation, therefore longer delay ti mes can be achieved by further increasing δ. Figure 5 also shows that when the amplitude ratio δ≤3.0, the delay time is a negative number which corresponds to fast light with ϕ= 0.35π, and the positive delay time corresponds to slow 15TABLE I. Summary of MIT, MIABS, MIAMP and Fano resonance obse rved experimentally for different values in parameter space. Amplitude Ratio δ 0 - 0.3 0.3 0.3 - 1.2 1.2 1.2 - 3.0 >3.0 Phaseϕ0.35πMIT NULL MIABS MIABS MIABS Fano 1.35πMIT MIT MIT MIT (perfect) MIAMP MIAMP light with ϕ= 1.35π. Thus the phase parameter ϕcan also be used to switch fast and slow light. When δ= 0, i.e., no magnon pump, our scheme recovers the traditional M IT and only a 16-ns delay time is achieved. By applying the magnon pump and optimizing ϕandδ,the time delay, as well as advance, can be enhanced by nearly 2 orders o f magnitude compared with the case without magnon pump . For our scheme, the pump-probe amplitude ratio and phase di fference mediated path interference can result in the zero reflection , which is accompanied with a delay time abrupt transition. In our experiment, Fig. 5(a) clearly sho ws such an abrupt transition and greatly enhanced fast-slow light around this point. We can find that t he experimental data deviates from the theoretical result around the abrupt transition. This i s mainly induced by the imperfect system setups, such as limited output precision of AWG, imperfectn ess of the I-Q mixer and unstable magnon frequency [73]. VIII. CONCLUSION We experimentally study how the magnon pump affects the prob e-field transmission, and the observed results are summarized in Table. I. Two parameters , the relative phase ϕand the pump- probe ratio δbetween pump and probe tones, are studied in detail. The main results of this work are as follows: • the unconventional MIABS of the transmitted microwave fiel d is observed with the cavity in the undercoupling condition; • MIAMP phenomena is realized in our experiment; • asymmetric Fano-resonance-like spectra are observed eve n when the cavity is resonant with the magnon; 16• by tuning the phase of the magnon pump, we can easily switch b etween MIT, MIABS and MIAMP; • by tuning the pump and probe ratio, the MIABS and MIAMP can be further optimized, accompanied by greatly enhanced advanced or slow light by ne arly 2 orders of magnitude; • the tunable phase and amplitude ratio can lead to the zero re flection of the transmitted light and abrupt fast-slow light transitions.; • both the ϕandδcan be used to carry out the in situ switch of fast and slow light. Our results confirm that direct magnon pumping through the co upling loops provides a versa- tile route to achieve controllable signal transmission, st orage, and communication, which can be further expanded to coherent state processing in the quantu m regime. Furthermore, by exploiting multi-YIG spheres or multimagnon modes systems, the amplifi cation or absorption bandwidth can be increased, resulting in a broadband coherent signal stor e device. The sharp peak and asymmet- ric Fano line shape indicate that our platform has great pote ntial in the application of high-precision measurement of weak microwave fields [74, 75]. Our two-tone p ump scheme and phase-tunable interference can also be accomplished in other coupled-res onator systems, such as optomechanical resonators, which explores effects of mechanical pump on li ght transmission [76–84], and even in circuit-QED systems, in which photon transmission can be controlled through a circuit-QED system [85–88]. ACKNOWLEDGMENTS This work is supported by the National Key R&D Program of Chin a (Grant No. 2018YFA0306600), the CAS (Grants No. GJJSTD20170001 and No. QYZDY-SSW-SLH00 4), Anhui Initiative in Quantum Information Technologies (Grant No. AHY050000), a nd the Natural Science Foun- dation of China (NSFC) (Grant No. 12004044). F.N. is support ed in part by: NTT Research, Army Research Office (ARO) (Grant No. W911NF-18-1-0358), Ja pan Science and Technol- ogy Agency (JST) (via the CREST Grant No. JPMJCR1676), Japan Society for the Promotion of Science (JSPS) (via the KAKENHI Grant No. JP20H00134 and t he JSPS-RFBR Grant No. JPJSBP120194828), the Asian Office of Aerospace Research an d Development (AOARD), and the Foundational Questions Institute Fund (FQXi) via Grant No. FQXi-IAF19-06. 17Note added – Recently, we become aware of a study presenting a n infinite group delay and abrupt transition in a magnonic non-Hermitian system [33]. [1] M. Born, E. Wolf, A. B. Bhatia, P. C. Clemmow, D. Gabor, A. R. Stokes, A. M. Taylor, P. A. Wayman, and W. L. Wilcock, Principles of Optics: Electromagnetic Theory of Propagati on, Interference and Diffraction of Light , 7th ed. (Cambridge University Press, 1999). [2] A. E. Miroshnichenko, S. Flach, and Y . S. Kivshar, Fano re sonances in nanoscale structures, Reviews of Modern Physics 82, 2257 (2010). [3] M. Fleischhauer, A. Imamoglu, and J. P. Marangos, Electr omagnetically induced transparency: Optics in coherent media, Reviews of Modern Physics 77, 633 (2005). [4] P. M. Anisimov, J. P. Dowling, and B. C. Sanders, Objectiv ely discerning Autler-Townes splitting from electromagnetically induced transparency, Physical Revi ew Letters 107, 163604 (2011). [5] M. Kroner, A. O. Govorov, S. Remi, B. Biedermann, S. Seidl , A. Badolato, P. M. Petroff, W. Zhang, R. Barbour, B. Gerardot, et al. , The nonlinear Fano effect, Nature 451, 311 (2008). [6] D. Golde, M. Wagner, D. Stehr, H. Schneider, M. Helm, A. An drews, T. Roch, G. Strasser, M. Kira, and S. W. Koch, Fano signatures in the intersubband terahert z response of optically excited semicon- ductor quantum wells, Physical Review Letters 102, 127403 (2009). [7] L. Zhou, Z. Gong, Y .-x. Liu, C. P. Sun, and F. Nori, Control lable scattering of a single photon inside a one-dimensional resonator waveguide, Physical Review Let ters101, 100501 (2008). [8] H. Ian, Y .-x. Liu, and F. Nori, Tunable electromagnetica lly induced transparency and absorption with dressed superconducting qubits, Physical Review A 81, 063823 (2010). [9] Q. C. Liu, T. F. Li, X. Q. Luo, H. Zhao, W. Xiong, Y . S. Zhang, Z. Chen, J. S. Liu, W. Chen, and F. Nori, Method for identifying electromagnetically induc ed transparency in a tunable circuit quantum electrodynamics system, Physical Review A 93, 053838 (2016). [10] J. Long, H. S. Ku, X. Wu, X. Gu, R. E. Lake, M. Bal, Y .-x. Liu , and D. P. Pappas, Electromagneti- cally induced transparency in circuit quantum electrodyna mics with nested polariton states, Physical Review Letters 120, 083602 (2018). [11] M. J. Akram, F. Ghafoor, M. M. Khan, and F. Saif, Control o f Fano resonances and slow light using Bose-Einstein condensates in a nanocavity, Physical Revie w A 95, 023810 (2017). 18[12] C. Garrido Alzar, M. Martinez, and P. Nussenzveig, Clas sical analog of electromagnetically induced transparency, American Journal of Physics 70, 37 (2002). [13] M. F. Limonov, M. V . Rybin, A. N. Poddubny, and Y . S. Kivsh ar, Fano resonances in photonics, Nature Photonics 11, 543 (2017). [14] Y .-C. Liu, B.-B. Li, and Y .-F. Xiao, Electromagnetical ly induced transparency in optical microcavities, Nanophotonics 6, 789 (2017). [15] B. Peng, S ¸. K. ¨Ozdemir, W. Chen, F. Nori, and L. Yang, What is and what is not e lectromagnetically induced transparency in whispering-gallery microcavitie s, Nature Communications 5, 1 (2014). [16] S ¸. ¨Ozdemir, S. Rotter, F. Nori, and L. Yang, Parity–time symmet ry and exceptional points in photonics, Nature Materials 18, 783 (2019). [17] Z. Zhao, X. Zheng, W. Peng, J. Zhang, H. Zhao, and W. Shi, T erahertz electromagnetically-induced transparency of self-complementary meta-molecules on Cro atian checkerboard, Scientific Reports 9, 1 (2019). [18] R. Yahiaoui, J. A. Burrow, S. M. Mekonen, A. Sarangan, J. Mathews, I. Agha, and T. A. Searles, Electromagnetically induced transparency control in tera hertz metasurfaces based on bright-bright mode coupling, Physical Review B 97, 155403 (2018). [19] J. Johansson, G. Johansson, and F. Nori, Optomechanica l-like coupling between superconducting res- onators, Physical Review A 90, 053833 (2014). [20] C. Eichler and J. R. Petta, Realizing a circuit analog of an optomechanical system with longitudinally coupled superconducting resonators, Physical Review Lett ers120, 227702 (2018). [21] H. Xu, D. Mason, L. Jiang, and J. Harris, Topological ene rgy transfer in an optomechanical system with exceptional points, Nature 537, 80 (2016). [22] F. Liu, M. Ke, A. Zhang, W. Wen, J. Shi, Z. Liu, and P. Sheng , Acoustic analog of electromagnetically induced transparency in periodic arrays of square rods, Phy sical Review E 82, 026601 (2010). [23] M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, Cavi ty optomechanics, Reviews of Modern Physics 86, 1391 (2014). [24] M. J. Akram, M. M. Khan, and F. Saif, Tunable fast and slow light in a hybrid optomechanical system, Physical Review A 92, 023846 (2015). [25] L. Li, W. Nie, and A. Chen, Transparency and tunable slow and fast light in a nonlinear optomechanical cavity, Scientific Reports 6, 35090 (2016). [26] D.-G. Lai, X. Wang, W. Qin, B.-P. Hou, F. Nori, and J.-Q. L iao, Tunable optomechanically induced 19transparency by controlling the dark-mode effect, Physica l Review A 102, 023707 (2020). [27] D. P. Lake, M. Mitchell, B. C. Sanders, and P. E. Barclay, Two-colour interferometry and switching through optomechanical dark mode excitation, Nature Commu nications 11, 1 (2020). [28] M. C. Kuzyk and H. Wang, Controlling multimode optomech anical interactions via interference, Phys- ical Review A 96, 023860 (2017). [29] Y .-L. Liu, R. Wu, J. Zhang, S ¸. K. ¨Ozdemir, L. Yang, F. Nori, and Y .-x. Liu, Controllable optic al response by modifying the gain and loss of a mechanical reson ator and cavity mode in an optome- chanical system, Physical Review A 95, 013843 (2017). [30] B. Wang, Z.-X. Liu, C. Kong, H. Xiong, and Y . Wu, Mechanic al exceptional-point-induced trans- parency and slow light, Optics Express 27, 8069 (2019). [31] H. L¨ u, C. Wang, L. Yang, and H. Jing, Optomechanically i nduced transparency at exceptional points, Physical Review Applied 10, 014006 (2018). [32] T. Goldzak, A. A. Mailybaev, and N. Moiseyev, Light stop s at exceptional points, Physical Review Letters 120, 013901 (2018). [33] Y . Yang, Y .-P. Wang, J. Rao, Y . Gui, B. Yao, W. Lu, and C.-M . Hu, Unconventional singularity in anti-parity-time symmetric cavity magnonics, Physical Re view Letters 125, 147202 (2020). [34] C. Wang, X. Jiang, G. Zhao, M. Zhang, C. W. Hsu, B. Peng, A. D. Stone, L. Jiang, and L. Yang, Electromagnetically induced transparency at a chiral exce ptional point, Nature Physics 16, 334 (2020). [35] M. Goryachev, W. G. Farr, D. L. Creedon, Y . Fan, M. Kostyl ev, and M. E. Tobar, High-cooperativity cavity QED with magnons at microwave frequencies, Physical Review Applied 2, 054002 (2014). [36] Y . Tabuchi, S. Ishino, T. Ishikawa, R. Yamazaki, K. Usam i, and Y . Nakamura, Hybridizing ferromag- netic magnons and microwave photons in the quantum limit, Ph ysical Review Letters 113, 083603 (2014). [37] D. Zhang, X.-M. Wang, T.-F. Li, X.-Q. Luo, W. Wu, F. Nori, and J. Q. You, Cavity quantum electrody- namics with ferromagnetic magnons in a small yttrium-iron- garnet sphere, npj Quantum Information 1, 1 (2015). [38] D. Zhang, X.-Q. Luo, Y .-P. Wang, T.-F. Li, and J. Q. You, O bservation of the exceptional point in cavity magnon-polaritons, Nature Communications 8, 1 (2017). [39] B. Yao, Y . Gui, J. Rao, S. Kaur, X. Chen, W. Lu, Y . Xiao, H. G uo, K.-P. Marzlin, and C.-M. Hu, Cooperative polariton dynamics in feedback-coupled cavit ies, Nature Communications 8, 1 (2017). [40] L. Bai, M. Harder, P. Hyde, Z. Zhang, C.-M. Hu, Y . Chen, an d J. Q. Xiao, Cavity mediated manipu- 20lation of distant spin currents using a cavity-magnon-pola riton, Physical Review Letters 118, 217201 (2017). [41] S. Kaur, B. Yao, J. Rao, Y . Gui, and C.-M. Hu, V oltage cont rol of cavity magnon polariton, Applied Physics Letters 109, 032404 (2016). [42] I. Boventer, C. D¨ orflinger, T. Wolz, R. Macˆ edo, R. Lebr un, M. Kl¨ aui, and M. Weides, Control of the coupling strength and linewidth of a cavity magnon-polarit on, Physical Review Research 2, 013154 (2020). [43] N. Kostylev, M. Goryachev, and M. E. Tobar, Superstrong coupling of a microwave cavity to yttrium iron garnet magnons, Applied Physics Letters 108, 062402 (2016). [44] G. Flower, M. Goryachev, J. Bourhill, and M. E. Tobar, Ex perimental implementations of cavity- magnon systems: from ultra strong coupling to applications in precision measurement, New Journal of Physics 21, 095004 (2019). [45] J. Bourhill, N. Kostylev, M. Goryachev, D. Creedon, and M. Tobar, Ultrahigh cooperativity interac- tions between magnons and resonant photons in a YIG sphere, P hysical Review B 93, 144420 (2016). [46] D. Lachance-Quirion, Y . Tabuchi, A. Gloppe, K. Usami, a nd Y . Nakamura, Hybrid quantum systems based on magnonics, Applied Physics Express 12, 070101 (2019). [47] R. Hisatomi, A. Osada, Y . Tabuchi, T. Ishikawa, A. Noguc hi, R. Yamazaki, K. Usami, and Y . Naka- mura, Bidirectional conversion between microwave and ligh t via ferromagnetic magnons, Physical Review B 93, 174427 (2016). [48] X. Zhang, N. Zhu, C.-L. Zou, and H. X. Tang, Optomagnonic whispering gallery microresonators, Physical Review Letters 117, 123605 (2016). [49] S. V . Kusminskiy, H. X. Tang, and F. Marquardt, Coupled s pin-light dynamics in cavity optomagnon- ics, Physical Review A 94, 033821 (2016). [50] S. V . Kusminskiy, Cavity optomagnonics, preprint arXi v:1911.11104 (2019). [51] Z.-X. Liu, C. You, B. Wang, H. Xiong, and Y . Wu, Phase-med iated magnon chaos-order transition in cavity optomagnonics, Optics Letters 44, 507 (2019). [52] Y . Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Yamaz aki, K. Usami, and Y . Nakamura, Coherent coupling between a ferromagnetic magnon and a superconduct ing qubit, Science 349, 405 (2015). [53] Y . Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Yamaz aki, K. Usami, and Y . Nakamura, Quantum magnonics: The magnon meets the superconducting qubit, Com ptes Rendus Physique 17, 729 (2016). [54] R. Morris, A. Van Loo, S. Kosen, and A. Karenowska, Stron g coupling of magnons in a YIG sphere 21to photons in a planar superconducting resonator in the quan tum limit, Scientific Reports 7, 1 (2017). [55] Y . Li, T. Polakovic, Y .-L. Wang, J. Xu, S. Lendinez, Z. Zh ang, J. Ding, T. Khaire, H. Saglam, R. Divan, et al. , Strong coupling between magnons and microwave photons in o n-chip ferromagnet- superconductor thin-film devices, Physical Review Letters 123, 107701 (2019). [56] J. T. Hou and L. Liu, Strong coupling between microwave p hotons and nanomagnet magnons, Physical Review Letters 123, 107702 (2019). [57] X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, Strongly co upled magnons and cavity microwave photons, Physical Review Letters 113, 156401 (2014). [58] P.-C. Xu, J. Rao, Y . Gui, X. Jin, and C.-M. Hu, Cavity-med iated dissipative coupling of distant mag- netic moments: Theory and experiment, Physical Review B 100, 094415 (2019). [59] Y . Yang, J. Rao, Y . Gui, B. Yao, W. Lu, and C.-M. Hu, Contro l of the magnon-photon level attraction in a planar cavity, Physical Review Applied 11, 054023 (2019). [60] M. Harder, Y . Yang, B. Yao, C. Yu, J. Rao, Y . Gui, R. Stamps , and C.-M. Hu, Level attraction due to dissipative magnon-photon coupling, Physical Review Lett ers121, 137203 (2018). [61] Y .-P. Wang, G.-Q. Zhang, D. Zhang, X.-Q. Luo, W. Xiong, S .-P. Wang, T.-F. Li, C.-M. Hu, and J. Q. You, Magnon Kerr effect in a strongly coupled cavity-magnon system, Physical Review B 94, 224410 (2016). [62] Y .-P. Wang, G.-Q. Zhang, D. Zhang, T.-F. Li, C.-M. Hu, an d J. Q. You, Bistability of cavity magnon polaritons, Physical Review Letters 120, 057202 (2018). [63] J. Zhao, Y . Liu, L. Wu, C.-K. Duan, Y .-x. Liu, and J. Du, Ob servation of anti- PT-symmetry phase transition in the magnon-cavity-magnon coupled system, Ph ysical Review Applied 13, 014053 (2020). [64] I. Boventer, M. Kl¨ aui, R. Macˆ edo, and M. Weides, Steer ing between level repulsion and attraction: broad tunability of two-port driven cavity magnon-polarit ons, New Journal of Physics 21, 125001 (2019). [65] C. Kong, H. Xiong, and Y . Wu, Magnon-induced nonrecipro city based on the magnon Kerr effect, Physical Review Applied 12, 034001 (2019). [66] G. Zhang, Y . Wang, and J. Q. You, Theory of the magnon Kerr effect in cavity magnonics, Science China Physics, Mechanics & Astronomy 62, 987511 (2019). [67] S.-N. Huai, Y .-L. Liu, J. Zhang, L. Yang, and Y .-x. Liu, E nhanced sideband responses in a PT- symmetric-like cavity magnomechanical system, Physical R eview A 99, 043803 (2019). [68] X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, Cavity magn omechanics, Science Advances 2, 22e1501286 (2016). [69] J. Li, S.-Y . Zhu, and G. Agarwal, Magnon-photon-phonon entanglement in cavity magnomechanics, Physical Review Letters 121, 203601 (2018). [70] J. Li and S.-Y . Zhu, Entangling two magnon modes via magn etostrictive interaction, New Journal of Physics 21, 085001 (2019). [71] J. Li, S.-Y . Zhu, and G. Agarwal, Squeezed states of magn ons and phonons in cavity magnomechanics, Physical Review A 99, 021801 (2019). [72] M. Yu, H. Shen, and J. Li, Magnetostrictively induced st ationary entanglement between two mi- crowave fields, Physical Review Letters 124, 213604 (2020). [73] See supplemental material at [url will be inserted by pu blisher] for A. the derivation of the transmission spectra, B. the side-effect of the antenna 2, C. Error analys is. [74] J. A. Sedlacek, A. Schwettmann, H. K¨ ubler, R. L¨ ow, T. P fau, and J. P. Shaffer, Microwave electrometry with Rydberg atoms in a vapour cell using bright atomic reson ances, Nature Physics 8, 819 (2012). [75] K.-Y . Liao, H.-T. Tu, S.-Z. Yang, C.-J. Chen, X.-H. Liu, J. Liang, X.-D. Zhang, H. Yan, and S.-L. Zhu, Microwave electrometry via electromagnetically induced a bsorption in cold Rydberg atoms, arXiv preprint arXiv:2002.00855 (2020). [76] Y . Chang, T. Shi, Y .-x. Liu, C. Sun, and F. Nori, Multista bility of electromagnetically induced trans- parency in atom-assisted optomechanical cavities, Physic al Review A 83, 063826 (2011). [77] H. Suzuki, E. Brown, and R. Sterling, Nonlinear dynamic s of an optomechanical system with a coher- ent mechanical pump: Second-order sideband generation, Ph ysical Review A 92, 033823 (2015). [78] S.-C. Wu, L.-G. Qin, J. Jing, T.-M. Yan, J. Lu, and Z.-Y . W ang, Microwave-controlled optical dou- ble optomechanically induced transparency in a hybrid piez o-optomechanical cavity system, Physical Review A 98, 013807 (2018). [79] W. Jia, L. Wei, Y . Li, and Y .-x. Liu, Phase-dependent opt ical response properties in an optomechanical system by coherently driving the mechanical resonator, Phy sical Review A 91, 043843 (2015). [80] X.-W. Xu and Y . Li, Controllable optical output fields fr om an optomechanical system with mechanical driving, Physical Review A 92, 023855 (2015). [81] C. Jiang, Y . Cui, Z. Zhai, H. Yu, X. Li, and G. Chen, Phase- controlled amplification and slow light in a hybrid optomechanical system, Optics Express 27, 30473 (2019). [82] T.-X. Lu, Y .-F. Jiao, H.-L. Zhang, F. Saif, and H. Jing, S elective and switchable optical amplification with mechanical driven oscillators, Physical Review A 100, 013813 (2019). 23[83] H. Jing, S ¸. K. ¨Ozdemir, Z. Geng, J. Zhang, X.-Y . L¨ u, B. Peng, L. Yang, and F. Nori, Optomechanically- induced transparency in parity-time-symmetric microreso nators, Scientific Reports 5, 9663 (2015). [84] H. Wang, X. Gu, Y .-x. Liu, A. Miranowicz, and F. Nori, Opt omechanical analog of two-color elec- tromagnetically induced transparency: Photon transmissi on through an optomechanical device with a two-level system, Physical Review A 90, 023817 (2014). [85] Y .-x. Liu, X.-W. Xu, A. Miranowicz, and F. Nori, From blo ckade to transparency: Controllable photon transmission through a circuit-QED system, Physical Revie w A 89, 043818 (2014). [86] X. Gu, S.-N. Huai, F. Nori, and Y .-x. Liu, Polariton stat es in circuit QED for electromagnetically induced transparency, Physical Review A 93, 063827 (2016). [87] X. Wang, A. Miranowicz, H.-R. Li, F.-L. Li, and F. Nori, T wo-color electromagnetically induced transparency via modulated coupling between a mechanical r esonator and a qubit, Physical Review A 98, 023821 (2018). [88] H.-C. Sun, Y .-x. Liu, H. Ian, J. Q. You, E. Il’Ichev, and F . Nori, Electromagnetically induced trans- parency and Autler-Townes splitting in superconducting flu x quantum circuits, Physical Review A 89, 063822 (2014). 24 | 2021-02-24 | We study the phase controlled transmission properties in a compound system
consisting of a 3D copper cavity and an yttrium iron garnet (YIG) sphere. By
tuning the relative phase of the magnon pumping and cavity probe tones,
constructive and destructive interferences occur periodically, which strongly
modify both the cavity field transmission spectra and the group delay of light.
Moreover, the tunable amplitude ratio between pump-probe tones allows us to
further improve the signal absorption or amplification, accompanied by either
significantly enhanced optical advance or delay. Both the phase and
amplitude-ratio can be used to realize in-situ tunable and switchable fast-slow
light. The tunable phase and amplitude-ratio lead to the zero reflection of the
transmitted light and an abrupt fast-slow light transition. Our results confirm
that direct magnon pumping through the coupling loops provides a versatile
route to achieve controllable signal transmission, storage, and communication,
which can be further expanded to the quantum regime, realizing coherent-state
processing or quantum-limited precise measurements. | Phase-controlled pathway interferences and switchable fast-slow light in a cavity-magnon polariton system | 2102.12181v1 |
arXiv:1606.03469v1 [cond-mat.mes-hall] 10 Jun 2016Indirect Coupling between Two Cavity Photon Systems via Fer romagnetic Resonance Paul Hyde,a)Lihui Bai,b)Michael Harder, Christophe Match, and Can-Ming Hu Department of Physics and Astronomy, University of Manitob a, Winnipeg, Canada R3T 2N2 (Dated: 14 September 2018) We experimentally realize indirect coupling between two cavity modes v ia strong coupling with the ferromag- netic resonance in Yttrium Iron Garnet (YIG). We find that some ind irectly coupled modes of our system can have a higher microwave transmission than the individual uncoup led modes. Using a coupled harmonic oscillator model, the influence of the oscillation phase difference betw een the two cavity modes on the nature of the indirect coupling is revealed. These indirectly coupled microwav e modes can be controlled using an external magnetic field or by tuning the cavity height. This work has potential for use in controllable optical devices and information processing technologies. The indirect coupling of cavity modes via a waveguide has been studied theoretically and experimentally for use inopticalinformationprocessing1. Thisindirectcoupling dramatically modifies the transmission spectra, and is widely used for optical filtering, buffering, switching, and sensinginphotoniccrystalstructures2–5. Formicro/nano disk optical cavities, coupling properties are determined by the spatial distance between the disk and the waveg- uide during the fabrication process. Therefore, a tunable coupling between indirectly coupled cavity modes is re- quired for potential applications. Recently, strong coupling between a microwave cavity mode and ferromagnetic resonance (FMR) has been re- alized at room temperature6–17. Exchange interactions lock the high density of spins in YIG into a macro-spin state, leading to strong coupling with a cavity mode which can be adjusted using an external magnetic field. Potential applications of this form of strong coupling are currently being explored. For example, indirect coupling between the FMR in two YIG spheres has produced dark magnon modes with potential uses in information stor- age technologies18, and the FMR of YIG has been indi- rectly coupled with a qubit through a microwave cavity mode19. Instead of using a microwave cavity mode to build a bridge between two oscillators, we have used the FMR in YIG to produce indirect coupling between two cavity modes. In this work, we present two cavity modes which in- directly couple via their strong coupling with the FMR in YIG at room temperature. The two cavity modes are labelled hω1(ω) andhω2(ω) respectively, and are inde- pendent of each other when there is no direct coupling between them. Here ω1andω2are the uncoupled reso- nance frequencies of each cavity mode and ωis the in- put microwave frequency. The two cavity modes can be indirectly coupled with each other when they both in- dividually interact with the FMR in YIG and this indi- rect coupling can be controlled using an external mag- netic field. We found that the microwave transmission a)Electronic mail: umhydep@myumanitoba.ca b)Electronic mail: bai@physics.umanitoba.ca FIG.1. (Colour online) (a)Inanuncoupledsystemindividual elements do not interact with each other. In our experimenta l system a YIG sphere simultaneously couples to two separate cavity modes, indirectly coupling the modes together. (b) The frequencies of the two cavity modes are functions of the height of the cylindrical microwave cavity and cross near a height of 36 mm. The inset shows a sketch of the microwave cavity.(c)Transmission spectrum S21of our indirectly cou- pled system, as a function of the external magnetic field at a microwave cavity height of 36.5 mm [dashed line in (b)]. (d) Transmission spectrum S21ofourcavitysystem, withaheight of 36.5 mm at an external field µ0H= 0.412 T, in an uncou- pled state (NO YIG in cavity) and (e)an indirectly coupled state (YIG in cavity), showing the influence of coupling on the resonant modes. properties change dramatically for the coupled modes as the external field is tuned. Our experimental results,2 together with an extended coupled harmonic oscillator model, demonstrate the nature of indirect coupling and coherent information transfer. This tunable interaction between orthogonal cavity modes could potentially be used to build controllable optical and microwave devices. The microwave cavity used in our experiment was made of oxygen-free copper with a height tunable cylin- drical structure. The diameter of the cavity is 25 mm and the height is tunable in a range between 24 mm and 45 mm. Although multiple modes can exist inside of the cavity, the TM 012mode (with a cavity frequency ofω1) and the TE 211mode (with a cavity frequency of ω2) were chosen to demonstrate indirect coupling in this work. With no YIG inside of the cavity, the microwave transmission, S21, was measured using a Vector Network Analyser (VNA) as a function of frequency. The out- put microwave power of the VNA is 1 mW. The am- plitude of the transmission is proportional to the res- onance amplitudes of both cavity modes at a given mi- crowavefrequency, |S21(ω)|2∝ |hω1(ω)+hω2(ω)|2. Here, hω1=Γ1ω2 ω2−ω2 1+2iβ1ω1ωh0andhω2=Γ2ω2 ω2−ω2 2+2iβ2ω2ωh0are the response functions of each cavity mode near the res- onance conditions. ω1,ω2,β1, andβ2are the cavity mode resonance frequencies and damping. Γ 1and Γ2de- note the impedance matching parameters for each cavity mode.h0(ω) is the microwave field used to drive reso- nance in the cavity and is eliminated by normalization in the microwave transmission. The microwave trans- mission spectra with no YIG in the cavity allows the individual cavity mode frequencies and damping to be evaluated. Fig. 1(b) plots the resonant frequencies of ω1andω2as a function of the height of the microwave cavity, both agree well with the solutions for Maxwell’s equations (solid lines) in a cylindrical microwave cavity. Thatthetwocavitymodescrosseachotherindicatesthat there is no direct coupling between them. The different microwave magnetic field distributions of the two modes inside the cavity leads to them having different coupling strengthswiththe FMRinYIG.Foragivencavityheight of 36.5 mm, the parameters of the two cavity modes were determined to be: ω1/2π= 12.357GHz, β1= 1.9×10−4, Γ1= 6.1×10−5,ω2/2π= 12.382 GHz, β2= 0.91×10−4, and Γ2= 3.7×10−5. A YIG sphere20placed inside the cavity allows for in- direct coupling between the two cavity modes. The YIG sphere has a diameter of 1 mm, saturation magnetiza- tionµM0= 0.178 T, gyromagnetic ratio γ= 28×2πµ0 GHz/T, and Gilbert damping α= 1.15×10−4. The YIG sphere was placed at the bottom of the cavity near the wall as shown in the inset of Fig. 1(b). An external mag- netic field, H, was applied to the YIG as shown in the inset. This magnetic field allows us to tune the FMR fre- quency of the YIG, ωFMR, following an ω-H dispersion ωFMR=γ(H+HAni). Here, the anisotropy field of the sphere is µ0HAni= 0.0294 T. Transmission measurements of our coupled system are plotted in Fig. 1(c), which shows the amplitude |S21|2 as a function of the input microwave frequency ( ω) and FIG. 2. (Colour online) (a)and(b)display the ω-H disper- sion and damping evolution (symbols) of each of the Normal Modes in our system. They are compared to calculations from Eq. 1 (solid curves). (c)The amplitudes of the Nor- mal Modes, |S21|2, are dramatically enhanced or suppressed during coupling. (d)The relative phase between the two cav- ity modes, φ1−φ2, was calculated during indirect coupling. The in-phase point of Mode B corresponds to its maximum amplitude in (c). the external magnetic field H. By increasing the Hfield, the FMR frequency ( ωFMR) first increases to the lower cavity mode frequency ω1, then reaches the higher cavity mode frequency ω2as indicated by the dashed lines. By doing this, the two cavity modes are indirectly coupled together via their direct coupling with the FMR in YIG, producing three coupled modes. We observed a maxi- mum in the microwave transmission amplitude when the middle mode (later labelled Mode B) crosses the disper- sion of the YIG FMR (dashed line) due to the resonances ofthe twocavitymodesbeing in-phase. Fig. 1(d) and (e) show how the addition of the YIG sphere into the cavity affects the observed resonant modes at an external field of 0.412 T; with both the number and position of the ob- served modes changing once the sphere is placed in the cavity. To further understand the nature of this indirect cou- pling between the two cavity modes, an expanded cou- pled harmonic oscillator system is used to calculate cou- plingfeaturesincludingthe ω-Hdispersion,dampingevo- lution, and amplitudes. Coupled harmonic oscillators have previously been used to accurately model strong coupling between a cavity mode and FMR in YIG21. The coupling strengths between each cavity mode and the FMR in YIG, κ1= 0.070 and κ2= 0.043, were eval- uated using the two coupled harmonic oscillator model when the two cavity mode frequencies were well sepa- rated (not shown here). The local microwave magnetic field distribution of each mode, with respect to the exter- nal field orientation, lead to different coupling strengths between each cavity mode and the FMR in YIG22. A3 slight change of the cavity height does not change the coupling strength of each mode. However, the coupled system observed in this work can no longer be modelled by the two coupled harmonic oscillator model. To take into account the second cavity mode, a three oscillatorsystemis consideredratherthan the twoin Ref.[21]. Two of the oscillators describe the cavity modes with ampli- tudes of hω1andhω2, each separately coupled with the third representing the FMR in YIG with amplitude m. Therefore, the indirect coupling model can be written in the form; ω2−ω2 1+i2β1ω1ω 0 −κ2 1ω2 1 0 ω2−ω2 2+i2β2ω2ω κ2 2ω2 2 −κ2 1ω2 1 κ2 2ω2 2 ω2−ω2 FMR+i2αωFMRω hω1 hω2 m =ω2 Γ1 Γ2 0 h0(1) Here the diagonal terms are the uncoupled resonance conditions of the two cavity modes and the FMR in YIG. The off-diagonal terms are the coupling strengths. The two zeros indicate that there is no direct coupling be- tween the two cavity modes. To explain our experimen- talobservationswemustinclude a π-phasedelaybetween the resonance frequencies of neighbouring cavity modes, although the physical source of this phase shift is still an open question. This is the source of the additional minus sign in the κ1terms. Eq. 1 allows us to predict the char- acteristics of indirect coupling between cavity modes via FMR in YIG. By finding the complex eigen-frequencies ωn(n = A, B, C, denoting the Modes labelled in Figure 2) of the coupling matrix at a given Hfield, we can plot the cal- culated resonance frequency Re(ωn), and the normalized line width |Im(ωn)|/Re(ωn) in Fig. 2(a) and (b) using solid curves. This matches the observed ω-H dispersion and damping evolution seen in the measurements (sym- bols). Furthermore, we are able to calculate the amplitude and relative phase of the microwave transmission hω1, FIG. 3. (Colour online) (a),(b), and(c)show the transmis- sion spectrum of the three Normal Modes for different ω2−ω1 values.(d)Amplitude of the in-phase point of Mode B as a function of ω2−ω1.hω2, andmusing Eq. 1. The calculated transmission amplitude |S21|2was plotted in Fig. 2(c) (solid curves) and compared with that from our experimental results (symbols). An amplitude peak is seen in both the experi- mental results and the theoretical calculation. The phase difference( φ1−φ2)between hω1andhω2iscalculatedand plotted in Fig. 2(d). The applied field strength at the maximum amplitude corresponds to an in-phase point, highlighted in Fig. 2(d), where the phases of the two cavitymodes hω1andhω2areequal ( φ1−φ2= 0). Mean- while the amplitude decrease of the other Normal Modes is due to the relative phase difference between the two cavity modes approaching π. Hence, coherent phase con- trol between two indirectly coupled cavity modes is de- tected through amplitude enhancement of the microwave transmissionand explainedby our three oscillatormodel. The in-phase point observed occurs when Mode B crosses the uncoupled dispersion of the YIG FMR. The amplitude of this in-phase point also depends on the dif- ference between the resonant frequencies of the two cav- ity modes ( ω2−ω1). By tuning the cavity height, the in- phasepointcanbemeasuredfordifferentvaluesof ω2−ω1 as shown in Fig. 3(a), (b), and (c). The amplitude of these in-phase points, highlighted in red, increases when the two cavity mode frequencies are near to each other. As summarized in Fig. 3(d), the transmission amplitude |S21|2decreases as the two cavity mode frequencies are separated. Therefore, the microwave transmission of the in-phasepointcanalsobecontrolledbythe cavityheight. In summary, we experimentally demonstrate control- lable indirect coupling between two microwave cavity modes through a YIG sphere. The coupling features are analysed and explained using a three coupled harmonic oscillator model. Microwave modes produced due to in- direct coupling were observed to have a higher transmis- sion rate than the two uncoupled cavity modes. We also demonstrated that these indirectly coupled modes can be controlled with an external field and by changing the cavity’s height. Therefore, due to the controllable nature of our findings, our work can be useful for designing new optical devices for information processing. The authors would like to thank B. Yao for useful discussions. P.H. is supported by the UMGF program. M.H. is supported by an NSERC CGSD Scholarship.4 This work has been funded by NSERC, CFI, and NSFC (No. 11429401) grants (C.-M. Hu). 1S. Fan, ’Sharp asymmetric line shapes in side-coupled waveguide- cavity systems ’, Appl. Phys. Lett. 80, 908 (2002) 2P. Chak, S. Pereira, and J.E. Sipe, ’ Coupled-mode theory for pe- riodic side-coupled microcavity and photonic crystal stru ctures’, Phys. Rev. B 73, 035105 (2006) 3A.M. Armani, R.P. Kulkarni, S.E. Fraser, R.C. Flagan, and K. J. Vahala, ’ Label-free, single-molecule detection with optical micro - cavities’, Science 317, 783 (2007) 4J. Zhu, S.K. Ozdemir, Y.F. Xiao, L. Li, L. He, D.R. Chen, and L. Yang, ’On-chip single nanoparticle detection and sizing by mode splitting in an ultrahigh-Q microresonator ’, Nat. Photon. 4, 46 (2010) 5J. Li, R. Yu, C. Ding, and Y. Wu, ’ PT-symmetry-induced evo- lution of sharp asymmetric line shapes and high-sensitivit y re- fractive index sensors in a three-cavity array ’, Phys. Rev. A 93, 023814 (2016) 6H. Huebl, C.W. Zollitsch, J. Lotze, F. Hocke, M. Greifenstei n, A. Marx, R. Gross, and S.T.B. Goennenwein, ’ High Coopera- tivity in Coupled Microwave Resonator Ferrimagnetic Insul ator Hybrids’, Phys. Rev. Lett. 111, 127003 (2013) 7Y. Tabuchi, S. Ishino, T. Ishikawa, R. Yamazaki, K. Usami, and Y. Nakamura, ’ Hybridizing Ferromagnetic Magnons and Mi- crowave Photons in the Quantum Limit ’, Phys. Rev. Lett. 113, 083603 (2014) 8X. Zhang, C.-L. Zou, L. Jiang, and H.X. Tang, ’ Strongly coupled magnons and cavity microwave photons ’, Phys. Rev. Lett. 113, 156401 (2014) 9M. Goryachev, W.G. Farr, D.L. Creedon, Y. Fan, M. Kostylev, and M.E. Tobar, ’ High-Cooperativity Cavity QED with Magnons at Microwave Frequencies ’, Phys. Rev. Applied 2, 054002 (2014) 10B. Bhoi, T. Cliff, I.S. Maksymov, M. Kostylev, R. Aiyar, N. Venkataramani, S. Prasad, and R.L. Stamps, ’ Study of photon- magnon coupling in a YIG-film split-ring resonant system ’, J. Appl. Phys. 116, 243906 (2014) 11N.J. Lambert, J.A. Haigh, and A.J. Ferguson, ’ Identification ofspin wave modes in yttrium iron garnet strongly coupled to a co-axial cavity ’, J. Appl. Phys. 117, 053910 (2015) 12J.A. Haigh, N.J. Lambert, A.C. Doherty, and A.J. Ferguson, ’Dispersive readout of ferromagnetic resonance for strongl y cou- pled magnons and microwave photons ’, Phys. Rev. B 91, 104410 (2015) 13B.M. Yao, Y.S. Gui, M. Worden, T. Hegmann, M. Xing, X.S. Chen, W. Lu, Y. Wroczynskyj, J. van Lierop, and C.-M. Hu, ’Quantifying the complex permittivity and permeability of m ag- netic nanoparticles ’, Appl. Phys. Lett. 106, 142406 (2015) 14L. Bai, M. Harder, Y.P. Chen, X. Fan, J.Q. Xiao, and C.-M. Hu, ’Spin Pumping in Electrodynamically Coupled Magnon-Photon Systems ’, Phys. Rev. Lett. 114, 227201 (2015) 15L.V. Abdurakhimov, Y.M. Bunkov, and D. Konstantinov, ’Normal-Mode Splitting in the Coupled System of of Hybridize d Nuclear Magnons and Microwave Photons ’, Phys. Rev. Lett. 114, 226402 (2015) 16B.M. Yao, Y.S. Gui, Y. Xiao, H. Guo, X.S. Chen, W. Lu, C.L. Chien, and C.-M. Hu, ’ Theory and experiment on cavity magnon polariton in the 1D configuration ’, Phys. Rev. B 92, 184407 (2015) 17A Osada, R. Hisatomi, A. Noguchi, Y. Tabuchi, R. Yamazaki, K. Usami, M. Sadgrove, R. Yalla, M. Nomura, and Y. Naka- mura, ’Cavity optomagnonics with spin orbit coupled photons ’, arXiv:1510.01837 (2015) 18X. Zhang, C.-L. Zou, N. Zhu, F. Marquardt, L. Jiang, and H.X. Tang, ’ Magnon dark modes and gradient memory ’, Nature Comm.6, 8914 (2015) 19Y. Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Yamazaki, K. Usami, and Y. Nakamura, ’ Coherent coupling between a ferro- magnetic magnon and a superconducting qubit ’, Science 24, 405 (2015) 20http://www.ferrisphere.com 21M. Harder, L. Bai, C. Match, and C.-M. Hu, ’ Study of the cavity- magnon-polariton transmission line shape ’, arXiv:1601.06049 22L. Bai, K. Blanchette, M. Harder, Y. Chen, X. Fan, J. Xiao, and C.-M. Hu, Control of the Magnon-Photon Coupling ’, IEEE Transactions on Magnetics, PP, 2527691 (2016) | 2016-06-10 | We experimentally realize indirect coupling between two cavity modes via
strong coupling with the ferromagnetic resonance in Yttrium Iron Garnet (YIG).
We find that some indirectly coupled modes of our system can have a higher
microwave transmission than the individual uncoupled modes. Using a coupled
harmonic oscillator model, the influence of the oscillation phase difference
between the two cavity modes on the nature of the indirect coupling is
revealed. These indirectly coupled microwave modes can be controlled using an
external magnetic field or by tuning the cavity height. This work has potential
for use in controllable optical devices and information processing
technologies. | Indirect Coupling between Two Cavity Photon Systems via Ferromagnetic Resonance | 1606.03469v1 |
Thermally controlled connement of spin wave eld in a magnonic YIG waveguide Pablo Borysa,, O. Kolokoltseva, Iv an G omez-Aristab, V. Zavislyakc, G. A. Melkovc, N. Qureshia, Csar L. Ordez-Romerod aInstituto de Ciencias Aplicadas y Tecnologa, Universidad Nacional Autnoma de Mxico (UNAM), Ciudad Universitaria, 04510, Mxico bCtedras Conacyt Instituto Nacional de Astrofsica, ptica y Electrnica, 72840, Mxico cFaculty of Radiophysics, Electronics and Computer Systems, Taras Shevchenko National University of Kiev, Ukraine dInstituto de Fsica, Universidad Nacional Autnoma de Mxico, Ciudad Universitaria, 04510, Mxico. Abstract Methods for detecting spin waves rely on electrodynamical coupling between the spin wave dipolar eld and an inductive probe. While this coupling is usually treated as constant, in this work, we experimentally and theoretically show that it is indeed temperature dependent. By measuring the spin wave magnetic eld as a function of temperature of, and distance to the sample, we demonstrate that there is both a longitudinal and transversal connement of the eld near the YIG-Air interface. Our results are relevant for spin wave detection, in particular in the eld of spin wave caloritronics. Keywords: YIG, spin waves, thermal connement, spin wave- waveguide, electrodynamic coupling, inductive probe, dipolar eld. 1. Introduction It is expected that the emergence of thin lm logic elements based on spin waves in thin-lm ferromagnetic solids can lead to a new generation of Boolean and analogue processors [1, 2, 3]. One of the important points here is the tech- nique of spin wave excitation and modulation of their parameters. Traditionally, 5 spin waves have been excited and detected using the inductive coupling of micro- electrodes to the dipolar magnetic eld of the spin wave system. Usually, this electrodynamical coupling is considered to be constant, however, as shown in this work, it can suer signicant variations, depending on the temperature of the ferromagnetic material. The temperature of the sample can change due to 10 spin wave dissipation, from 1 to 10oC [4, 5] or up to 100-300oC because of Corresponding author Email address: pabloborys@ciencias.unam.mx (Pablo Borys) Preprint submitted to Journal of Magnetism and Magnetic Materials April 30, 2022arXiv:1910.04304v1 [cond-mat.mes-hall] 9 Oct 2019Figure 1: A schematic view of experimental set-up. The inductive probe is attached to YX motorized translation stages external heating used to control the spin wave propagation [6, 7]. Recently, the typical electrodynamical and magneto-optical methods for spin wave detec- tion/excitation were enriched with the spin transfer torque (STT) [8, 9, 10] in Pt/magnet thin lm structures caused by electrical or thermal spin currents 15 [11, 12, 13, 14, 15]. STT has been recognized to be a much promising tool to detect exchange SW, to control dipole SW, and to generate thermo-electricity on the basis of spin Seebeck eect. The discovery of the later has stimulated a number of ideas involving magnetocaloritronics [16]. For example, thermally assisted STT has been used for enhancement of spin oscillations in resonators, 20 spin wave amplication and spin auto-oscillations [17, 18, 19, 20]. The aim of this work is to reveal lateral eects of sample heating in experimental con- gurations on the inductive coupling between micro-antennas and the dipolar magnetic eld of a spin wave system [21, 22, 23, 24, 25, 26, 27, 28, 29]. 2. Experiment 25 A schematic diagram of the experimental set-up, designed to investigate the coupling between spin waves propagating on a YIG/GGG sample and an 2inductive micro-transducer, is shown in Fig. 1. The sample is 1 mm wide in the Zdirection and 28 mm long in the Ydirection. The thickness of the YIG lm is 7m. The sample was biased by a tangential magnetic eld ( H0) applied 30 along theZaxis to provide the propagation of Magneto-Static Surface Waves (MSSW) in the Ydirection. MSSW were excited at one end of the sample (at Y=Y0), in a pulse regime, by dc electric current pulse
owing through a 0.25 mm-wide microstrip line terminated to a 50 Ohm resistive load. This method provides very short spin wave packets, with duration of 10 ns. In the time 35 domain the shortest period of the magnetization precession in the wave packet is limited by the rise time of the electric current pulse, and in the k-space the largest wavenumber (k) is limited by the microstrip line width [3]. The MSSW pulse propagation characteristics were registered by an inductive frame-shaped probe [30] (Fig. 1) sensitive to the Ymagnetic component of microwave eld 40 (hy) induced by the spin wave in the vicinity of the YIG lm. The probe was scanned over the sample plane along the Ycoordinate (Fig. 1) by a motorized translation stage. The distance between the probe and the sample surface was also controlled by a motorized translation stage. It should be noted that we used a frame probe with reduced X-dimension to have high spatial resolution 45 of hy along the lm normal, as the probe is displaced in the X direction. The probe electrode was fabricated with a 50 m micro-wire. The sample was heated with a solid state green laser with variable output power ( Popt), from 40 to 300 mW. The laser spot on YIG was of 0.5 mm in diameter and was located at the distance of YLfrom the excitation port. 50 Fig 2 compares the time-space evolution of the amplitude of the hy pulse at room temperature (RT sample) in 2(a), with a sample heated at optical power Popt= 180 mW in 2(b). The pulse waveform was recorded by a real time Tektronix oscilloscope with 6 GHz- bandwidth, at dierent Y- positions of the probe, and at xed distance X= 50m between the probe and the YIG lm 55 plane. The measurements were done with a uniform bias eld H0= 120 Oe, and the laser spot at the position YL= 15 mm. As seen in Fig.2, the wave packet in the optically heated sample acquires an additional group delay, compared to the sample at room temperature. This phenomenon has been discussed in ref. [31], and is caused by a reduction on the saturation magnetization Msthat in 60 turn decreases the slope in the MSSW dispersion relation. Fig. 3 shows the signal detected by the probe, as the probe moves along the Y axis, at dierent Popt. The value of each point in the curves in Fig. 3 represents the energy of the pulse envelope. As clearly seen in the gure, the signal induced in the probe increases in the vicinity of the laser spot, and this 65 increment is proportional to temperature of the hot zone. On the other hand, in the sample at room temperature (Curve 1) the MSSW pulse propagates and attenuates exponentially, in the usual way. The data presented in Fig.3 are proportional to the overlap integral between a small eective area of the probe frame and an evanescent function hy(x) [30]. Hence, displacing the probe along 70 the lm normal one can obtain the prole jhy(x)j2, shown in Fig.4. In this experiment the probe was located in the center of the hot zone Y=YL Fig.4 presents the principal result of this study: the density of hynear 3Figure 2: Propagation of MSSW pulse along the magnonic waveguide YIG/GGG. The inset in Fig. 2a) shows details the pulse waveform, with a duration of 8 ns, at three adjacent positions along the Y axis. The data represent the pulse waveform (amplitude) a) in the sample at room temperature, and b) in the sample heated at 380 K in its center. 4Figure 3: Fig. 3 The energy of the MSSW pulse at dierent distances from excitation port. The curves were recorded at dierent Popt, which induce dierent temperatures in the region Y=YL: 1) T = T ROOM ; 2) T = T ROOM + 50 K ; 3) T = T ROOM + 70 K ; 4) T = TROOM + 90 K. 5Figure 4: Energy of hycomponent as a function of the distance between the probe and YIG lm surface, at the xed Y-position of the probe Y=YL. Red and black experimental points show the energy density in evanescent MSSW eld in the heated sample (at T = 380 K) and the RT sample, respectively. 6the lm interface increases as the sample temperature increases, i.e. the heat modies the eld connement. 75 3. Theoretical Background The eect of the thermally dependent eld connement is caused by the decrease ofMsin the ferrite lm, as its temperature increases. It can be analyzed analytically by a full set of Maxwell equations. In our case, considering that the sample is innite in YZ plane, the solutions for the magnetic and electric elds 80 of MSSW are h= (hx;hy;0) and e= (0;0;ez), respectively. Let us compare transversal proles of monochromatic magnetic eld components hx; hyin hot and RT samples, taking into account that the elds have to be normalized to transmit a given power
ow Pthrough the sample. It is clear that in both hot and RT samples a value of Pshould be the same, supposing equal excitation 85 eciency of MSSW. It can be shown that the Pointing vector for MSSW is calculated as: P=c 8 ezh yi+ezh xj (1) or P2=c 8k0? ke2 z+a ez@e z @x j (2) in the YIG lm, and P1;3=c 8k0ke2 zj (3) in air and substrate. 90 Here:kis the MSSW wavenumber, k0=!=c,cis the speed of light in the vacuum,!is the MSSW frequency, = (!2 !2 1)=(!2 !2 H),a=!!M=(!2 !2 H),!H=
H0, and!M= 4MS,!1=!H(!+!M), and
is the electron gyromagnetic ratio. 95 Then, taking into account that h,ein Eq. 1 are proportional to a certain constant,A, the value of Afor both hot and RT sample can be calculated using the conditionP i=1;2;3Pi(Hot sample ) =P i=1;2;3Pi(RT sample ) =Const . The explicit expressions for MSSW eld components in Eq. 1 are given in Appendix A. The calculated magnetic eld proles are shown in Fig. 5. 100 The results were obtained by using the experimental approximation for temperature dependence of the saturation magnetization in YIG: Ms= 140 T(G),0:3 G/K [31]. The eld proles in Fig.5 correlate well with the experimental proles in Fig.4. 4. Discussion and Conclusions 105 The peculiarity of the results for the pulse group delay shown in Fig. 2 is that the local heating increases the pulse delay, however, it does not change the group velocity dispersion. As seen in Fig.2, the pulse width (the pulse duration) in the 7Figure 5: Fig.5. The values of the tangential (hy) an the normal (hx) eld components calculated for hot (red curves) and RT (blue curves) samples. 8hot region remains unchanged, with respect to the pulse width in the RT sample. This means that spatial width of the pulse along the Y coordinate decreases, 110 i.e. there is spatial, longitudinal compression of the pulse along the propagation direction. This leads to the increase of a peak and average amplitude of the pulse envelope for pulse power to be conserved. The eect has been analyzed in [32], where we used a large diameter loop antenna that was not sensitive to the eect of the transversal connement of the evanescent eld shown in Fig.4, 115 and 5. On the other hand, the results presented in Fig.3, 4, and 5 indicate that increasing the sample temperature increases the coupling between MSSW eld and the micro-antenna. The experimental, Fig.4, and theoretical, Fig.5, data demonstrate that this eect takes place due to an increasing concentration of magnetic elds near YIG-Air interface, the so-called transversal connement. 120 In conclusion, it is shown that the increase of the sample temperature leads to the increase of both longitudinal and transversal connement of MSSW in the vecinity of YIG lm. This eect, in turn, is revealed as the increase of the signal induced in a micro-antenna, that has to be taken into account in the experiments on spin-wave caloritronics. 125 5. Appendix A Full system of Maxwell equations for electromagnetic waves in the sample saturated in the Z direction describes two kind of waves. The subsystem @hz @y=i"k0ex @hz @y= i"k0ey @ey @x @ex @y= ik0hz(4) describes fast waves, which neglects magnetism, and the subsystem @hy @x @hx @y=i"k0ez ik0(hx iahy) =@ez @y ik0(iahx+hy) =@ez @y(5) that is used to describe MSSW. It can be reduced to 130 @2ez @x2+@2ez @y2+"?k2 0ez= 0 (6) MSSW elds, where ?= (2 2 a)=, which satisfy Eq. 2b, and Eq.3 are ez=Aeax+i(!t ky);hx=k k0Aeax+i(!t ky);hy= ia k0Aeax+i(!t ky) | {z } Air(7) 9ez= (Bcosh (mx) +Csinh (mx))ei(!t ky) hx=1 k0(2 2a)[k(Bcosh (mx) +Csinh (mx)) +am(Bsinh (mx) +Ccosh (mx))]ei(!t ky) hy=1 k0(2 2a)[ak(Bcosh (mx) +Csinh (mx)) +m(Bsinh (mx) +Ccosh (mx))]ei(!t ky) | {z } YIG (8) ez=De ax+i(!t ky); hx=k k0De ax+i(!t ky); hy=ia k0De ax+i(!t ky); | {z } Substrate (9) witha=p k2 k2 0, andm=p k2 ?k2 0. The standard electrody- namic boundary conditions at the structure interfaces determine the following relations between the coecients A,B,C,D A=B; De as=Bcosh (ms) +Ccosh (ms); a 2 2 a A=akB+mC; a 2 2 a e asD=ak(Bcosh (ms) +Csinh (ms)) +m(Bsinh (ms) +Ccosh (ms)) (10) Then, the constant A is calculated from the condition 135 X i=1;2;3Pi(Hotsample ) =X i=1;2;3Pi(RT sample ) =Const: (11) 6. Acknowledgements This work was supported by the UNAM-DGAPA research grant IG100517, and by fellowship BECA UNAM Posdoctoral. Dr. O. Kolokoltsev is thankful to UNAM-DGAPA for sabbatical scholarship. References 140 [1] A. Khitun, M. Bao, K. L. Wang, Spin wave magnetic nanofabric: A new approach to spin-based logic circuitry, IEEE Transactions on Magnetics 44 (9) (2008) 2141{2152. doi:10.1109/TMAG.2008.2000812 . [2] A. Khitun, M. Bao, K. L. Wang, Magnonic logic circuits, Journal of Physics D: Applied Physics 43 (26) (2010) 264005. 145 [3] O. V. Kolokoltsev, C. L. Ord o~ nez-Romero, N. Qureshi, Synthesis and pro- cessing of pseudo noise signals by spin precession in y3 fe5o12 lms, Journal of Applied Physics 110 (2) (2011) 024504. 10[4] O. Kolokoltsev, C. Ordonez-Romero, N. Qureshi, R. Ortega-Martinez, V. Grimalsky, Optical characterization of thermal-stress induced by spin 150 waves in thin-lm ferrimagnetic structures, Solid State Communications 142 (3) (2007) 137{142. [5] Y. K. Fetisov, C. E. Patton, Thermal microwave foldover and bistability in ferromagnetic resonance, IEEE transactions on magnetics 40 (2) (2004) 473{482. 155 [6] O. Rousseau, B. Rana, R. Anami, M. Yamada, K. Miura, S. Ogawa, Y. Otani, Realization of a micrometre-scale spin-wave interferometer, Sci- entic reports 5 (2015) 9873. [7] O. Kolokoltsev, N. Qureshi, E. Mej a-Uriarte, C. L. Ord o~ nez-Romero, Hot spin-wave resonators and scatterers, Journal of Applied Physics 112 (1) 160 (2012) 013902. [8] J. C. Slonczewski, Current-driven excitation of magnetic multilayers, Jour- nal of Magnetism and Magnetic Materials 159 (1-2) (1996) L1{L7. [9] E. Saitoh, M. Ueda, H. Miyajima, G. Tatara, Conversion of spin current into charge current at room temperature: Inverse spin-hall eect, Applied 165 physics letters 88 (18) (2006) 182509. [10] Y. Kajiwara, K. Harii, S. Takahashi, J.-i. Ohe, K. Uchida, M. Mizuguchi, H. Umezawa, H. Kawai, K. Ando, K. Takanashi, et al., Transmission of electrical signals by spin-wave interconversion in a magnetic insulator, Na- ture 464 (7286) (2010) 262. 170 [11] Z. Wang, Y. Sun, Y.-Y. Song, M. Wu, H. Schulthei, J. E. Pearson, A. Ho- mann, Electric control of magnetization relaxation in thin lm magnetic insulators, Applied Physics Letters 99 (16) (2011) 162511. [12] Y. Kajiwara, K. Uchida, D. Kikuchi, T. An, Y. Fujikawa, E. Saitoh, Spin- relaxation modulation and spin-pumping control by transverse spin-wave 175 spin current in y3fe5o12, Applied Physics Letters 103 (5) (2013) 052404. [13] K. Uchida, S. Takahashi, K. Harii, J. Ieda, W. Koshibae, K. Ando, S. Maekawa, E. Saitoh, Observation of the spin seebeck eect, Nature 455 (7214) (2008) 778. [14] K.-i. Uchida, H. Adachi, T. Ota, H. Nakayama, S. Maekawa, E. Saitoh, Ob- 180 servation of longitudinal spin-seebeck eect in magnetic insulators, Applied Physics Letters 97 (17) (2010) 172505. [15] K. Uchida, J. Xiao, H. Adachi, J.-i. Ohe, S. Takahashi, J. Ieda, T. Ota, Y. Kajiwara, H. Umezawa, H. Kawai, et al., Spin seebeck insulator, Nature materials 9 (11) (2010) 894. 185 11[16] G. E. Bauer, E. Saitoh, B. J. Van Wees, Spin caloritronics, Nature materials 11 (5) (2012) 391. [17] C. Safranski, I. Barsukov, H. K. Lee, T. Schneider, A. Jara, A. Smith, H. Chang, K. Lenz, J. Lindner, Y. Tserkovnyak, et al., Spin caloritronic nano-oscillator, Nature communications 8 (1) (2017) 117. 190 [18] H. Chang, P. P. Janantha, J. Ding, T. Liu, K. Cline, J. N. Gelfand, W. Li, M. C. Marconi, M. Wu, Role of damping in spin seebeck eect in yttrium iron garnet thin lms, Science advances 3 (4) (2017) e1601614. [19] V. Demidov, S. Urazhdin, E. Edwards, S. Demokritov, Wide-range con- trol of ferromagnetic resonance by spin hall eect, Applied Physics Letters 195 99 (17) (2011) 172501. [20] L. Lu, Y. Sun, M. Jantz, M. Wu, Control of ferromagnetic relaxation in magnetic thin lms through thermally induced interfacial spin transfer, Physical review letters 108 (25) (2012) 257202. [21] T. An, V. Vasyuchka, K. Uchida, A. Chumak, K. Yamaguchi, K. Harii, 200 J. Ohe, M. Jung
eisch, Y. Kajiwara, H. Adachi, et al., Unidirectional spin- wave heat conveyer, Nature materials 12 (6) (2013) 549. [22] M. Jung
eisch, T. An, K. Ando, Y. Kajiwara, K. Uchida, V. Vasyuchka, A. Chumak, A. Serga, E. Saitoh, B. Hillebrands, Heat-induced damping modication in yttrium iron garnet/platinum hetero-structures, Applied 205 Physics Letters 102 (6) (2013) 062417. [23] B. Obry, V. I. Vasyuchka, A. V. Chumak, A. A. Serga, B. Hillebrands, Spin-wave propagation and transformation in a thermal gradient, Applied Physics Letters 101 (19) (2012) 192406. [24] K. An, D. R. Birt, C.-F. Pai, K. Olsson, D. C. Ralph, R. A. Buhrman, 210 X. Li, Control of propagating spin waves via spin transfer torque in a metallic bilayer waveguide, Physical Review B 89 (14) (2014) 140405. [25] E. Padr on-Hern andez, A. Azevedo, S. Rezende, Amplication of spin waves by the spin seebeck eect, Journal of Applied Physics 111 (7) (2012) 07D504. 215 [26] M. Evelt, V. Demidov, V. Bessonov, S. Demokritov, J. Prieto, M. Mu~ noz, J. Ben Youssef, V. Naletov, G. De Loubens, O. Klein, et al., High-eciency control of spin-wave propagation in ultra-thin yttrium iron garnet by the spin-orbit torque, Applied Physics Letters 108 (17) (2016) 172406. [27] O. Gladii, M. Collet, K. Garcia-Hernandez, C. Cheng, S. Xavier, P. Bor- 220 tolotti, V. Cros, Y. Henry, J.-V. Kim, A. Anane, et al., Spin wave ampli- cation using the spin hall eect in permalloy/platinum bilayers, Applied Physics Letters 108 (20) (2016) 202407. 12[28] M. Jung
eisch, A. Chumak, A. Kehlberger, V. Lauer, D. Kim, M. Onbasli, C. Ross, M. Kl aui, B. Hillebrands, Thickness and power dependence of the 225 spin-pumping eect in y 3 fe 5 o 12/pt heterostructures measured by the inverse spin hall eect, Physical Review B 91 (13) (2015) 134407. [29] T. Langner, D. A. Bozhko, S. A. Bunyaev, G. N. Kakazei, A. V. Chu- mak, A. A. Serga, B. Hillebrands, V. I. Vasyuchka, Spin-wave propagation through a magnonic crystal in a thermal gradient, Journal of Physics D: 230 Applied Physics 51 (34) (2018) 344002. [30] N. P. Vlannes, Optical probing of magnetostatic forward volume waves in thin lm yttrium-iron-garnet, Journal of applied physics 62 (3) (1987) 972{989. [31] H. Algra, P. Hansen, Temperature dependence of the saturation magneti- 235 zation of ion-implanted yig lms, Applied Physics A 29 (2) (1982) 83{86. [32] O. Kolokoltsev, I. G omez-Arista, N. Qureshi, A. Acevedo, C. L. Ord o~ nez- Romero, A. Grishin, Compression gain of spin wave signals in a magnonic yig waveguide with thermal non-uniformity, Journal of Magnetism and Magnetic Materials 377 (2015) 1{5. 240 13 | 2019-10-09 | Methods for detecting spin waves rely on electrodynamical coupling between
the spin wave dipolar field and an inductive probe. While this coupling is
usually treated as constant, in this work, we experimentally and theoretically
show that it is indeed temperature dependent. By measuring the spin wave
magnetic field as a function of temperature of, and distance to the sample, we
demonstrate that there is both a longitudinal and transversal confinement of
the field near the YIG-Air interface. Our results are relevant for spin wave
detection, in particular in the field of spin wave caloritronics | Thermally controlled confinement of spin wave field in a magnonic YIG waveguide | 1910.04304v1 |
All-optical cryogenic thermometry based on NV centers in nanodiamon ds M. Fukami1, C. G. Yale1,†, P. Andrich1,‡, X. Liu1, F. J. Heremans1,2, P. F. Nealey1,2, D. D. Awschalom1,2,* 1. Institute for Molecular Engineering, University of Chicago, Chicago, IL 60637 2. Institute for Molecular Engineering and Materials Science Division, Argonne National Lab, Argonne, IL 60439 †Present address: Sandia National Laboratories, Albuquerque, NM, 87185 ‡Present address: University of Cambridge, Cavendish Laboratory, JJ Thomson Ave, Camb ridge CB3 0HE *Email: awsch@uchicago.edu ABSTRACT The nitrogen -vacancy (NV) center in diamond has been recognized as a high -sensitivity nanometer -scale metrology platform . Thermometry has been a recent focus, with attention largely confined to room temperature applications. Temperature sensing at low temperatures , however, remains challenging as the sensitivity decreases for many commonly used technique s, which rely on a temperature dependent frequency shift of NV center’s spin resonance and its control with microwaves . Here w e use an alternative approach that does not require microwaves , ratiometric all -optical thermometry , and demonstrate that it may be utilized to liquid nitrogen temperatures without deterioration of the sensitivity . The use of an array of nanodiamonds embedded within a portable polydimethylsiloxane (PDMS) sheet provides a versatile temperature sensing platform that can probe a wide variety of systems without the configurational restrictions needed for applying microwaves . With this device, w e observe a temperature gradient over tens of microns in a ferromagnetic -insulator substrate (yttrium iron garnet, YIG) under local heating by a resistive heater . This thermometry technique provides a cryogenically compatible, microwave - free, minimally invasive approach capable of probing local temperatures with few restriction s on the substrate materials . I. INTRODUCTION Local temperature variation plays a central role in many -body physics governed by hydrodynamic description s [1,2] , in biomolecular science [3], as well as in thermal engineering of integrated circuit s. Among the existing high -sensitivity nanometer -scale thermometers, nitrogen vacancy (NV) centers in nanodiamond s (NDs) have emerged as promising temperature -sensitive fluorescent probes . The negatively -charged NV -center (NV-) consists of a ground state spin triplet manifold with a zero-field splitting 𝒟⋍2.87 GHz that sensitively responds to temperature s, where the shift can be measured by reading out the spin optically [3–6]. By vi rtue of diamond’s high thermal conductivity an d NV- centers’ long spin coherence time, ND-based thermometry has been demonstrated in a variety of systems , such as within a living cell at room temperature [3]. The temperature response of 𝒟 is significantly smaller at low temperatures, however, which reduces sensitivity and hinders the conventional thermometry technique [7,8] . Ratiometric all -optical thermometry has been proposed as an alternative to the convent ional microwave spin-resonance thermometry technique with compatible sensitivity at room temperature [9–12]. It also enables temperature sensing without the application of microwave s, which removes concern s of microwave heating . Interestingly, the temperature sensitivity of the all -optical thermometer is estimated to improve at lower temperatures (see Supplementary Material, Sec. A [13]), and indicates that this tech nique can offer a path forward towards ND-based cryogenic thermometry . The use of an array of NDs on a polydimethylsiloxane ( PDMS ) sheet [13] combined with all -optical thermometry completely removes configurational restrictions needed for microwave application s, offering a versatile device capable of probing a wide variety of solid -state systems over tens of microns with an adjustable spatial resolution on the order of a few microns. This makes all-optical thermome try suitable for probing and imaging a variety of condensed matter systems , and may have advantages over conventional NV-center thermometry technique s depending on the required thermal or spatial resolutions as well as the potential microwave response of the target system . Here we extend the all -optical thermometry technique based on the NV- centers in NDs from room temperature to liquid nitrogen temperatures , 85 K T 300 K , and demonstrate its application on a ferromagnetic insulato r (yttrium iron garnet, YIG) substrate . In particular, we focus on YIG as a platform to demonstrate our sensing approach both because the microwave s used to manipulate NV centers in conventional thermometry would impact the magnetic spins in the YIG [14–19], and the low temperature thermal response of YIG is of interest in the study of the spin -Seebec k effect [20–24]. We initially demonstrate that a laser -pulse sequence to control the NV centers’ charge states improve s the sensitivity of the all -optical thermometer by approximately a factor of 3 . Next , we systematically study the temperatu re dependence of the sensitivity , demonstrating that it improves at cryogenic temperatures . Finally , we apply this all-optical cryogenic thermometry technique at T 170 K to measure the surface temperature profile of a YIG slab in contact with a resistive heater, with the array of NDs embedded on the surface of a flexible PDMS sheet . The observed temperature gradient over a range of tens of micrometers confirms the applicability of the technique on the YIG substrate , indicating that it provides a tool for study ing local thermal properties of a wide variety of substrates over a broad range of temperatures. II. DEMONSTRATION OF CRYOGENIC ALL -OPTICAL THERMOMETRY We focus on the temperature dependence of the NV- centers’ zero phonon line (ZPL) amplitude ratio ( A), which is defined as the ratio of the ZPL intensity with respect to an average photoluminescence (PL) intensity in a spectral range around the ZPL. The ratio A strongly responds to temperature change due to the presence of a coupling between the orbital state of NV- and vibrational modes in diamond [25] (see Supplementary Material, Sec. B [13]), which leads to a high temperature sensitivity . The experiment was conducted on an array of NDs containing ensemble s of NV- centers measured with a confocal microscope using a high numerical aperture objective (NA =0.9) as shown in Fig. 1(a). An array of NDs embedded into the flexible PDMS sheet was placed on the surface of a 3.05- m-thick YIG film grown on a 500-m-thick gadolinium gallium garnet ( GGG ) substrate (MTI Corp.) . A Ti/Au (thickness: 8nm/200nm) resistive heater , for local heating , was patterned on the YIG film using a lithographic process . The bottom of the GGG substrate was affixed to a copper thermal sink within a flow cryostat . Both characterization (section II) and application (section III) of the thermometry were conducted on the same device with a YIG substrate for consistency (for data without a PDMS sheet on a quartz substrate, see Supplementary Material, Sec. I [13]). Figure 1(b) shows a two-dimensional PL scan of a n individual spot in the array of NDs under continuous 594-nm excitation measured by an avalanche photodiode (APD). The 594 -nm light does not excite the neutrally - charged NV -center (NV0) [26,27] and removes the noisy NV0 phonon -sideband spectral emission from the NV-‘s ZPL spectrum . The diameter of the spot is 1000 nm which is defined by our microfabrication technique [28], and contains tens of NDs, where each ND contains hundreds of NV centers [28]. Figure 1(c) shows a horizontal cut through the maximum of Fig. 1(b) . Interestingly, when we applied pulse sequence s of the 594-nm and 532-nm laser s as shown in F ig. 1(d) , which is in contrast to the previous studies with a continuous -wave excitation [11,12] , the PL count rate was enhanced by approxi mately a factor of three (see Supplementary Material, Sec. D [13]). The enhancement is due to the charge -state conversion between NV- and NV0 [13,29 –31]. While charge -state conversions of NV centers in NDs have not been comprehensively studied to our knowledge, we simply assume the results reported in bulk diamonds are applicable and attribute the PL enhancement to the charge -state convers ion. Since the sensitivity of the all-optical thermometer is limited by shot noise , improving the PL count rate by a factor of 3 increases sensitivity by a factor of 3 (see Supplementary Material, Sec. E [13]). In the following spectral measurements , we send the PL to a spectrometer and gate the intensifier of a single -photon sensitive CCD camera in the spectrometer (iStar 334T, Andor) triggered by the pulse sequence s. Every spectra l measurement was followed by a background measurement taken off the ND and the background counts were subtracted . (see Supplemental Material, Sec. F [13]). Figure 2(a) shows the PL spectra ()Lh of NV- centers in the temperature range 85 K T 100 K . Monotonic change in the spectra is observed except near T ≃230 K and T ≃150 K , which are due to the melting point and the glass transition point of the PDMS , respectively . We note that the presence of the PDMS sheet does not change the thermometry property of NV centers except PL count rate s, which is verified by the measurements done on NDs without a PDMS sheet (See Supplementary Material, Sec. I [13]). To maximize the PL count rat e, we widely opened the slit in the spectrometer, which results in a wavelength resolution =3.5 nm. For the temperature sensing, we focus on the ZPL emission peak at h≃1.94 eV (637 nm) . Importantly, the ZPL becomes sharper and more prominent at lower temperatures . In this experiment , we focused on the PL in the wavelength ranging from 605 nm to 660 nm , which we define as the spectral range (ℛ) (for the choice of this range, see Supplemental Materia l, Sec. G [13]). As shown in the inset of F ig. 2(b), we fit the relative spectrum LLR by a sum of a squared -Lorentzian function and an exponential function ZPL 2 2 2 B ZPL() 1( ) exp[ ( ) ]hL h L A Bk w h h R (1) where Bk is the Boltzmann constant , ℎ is the Plank constant, LR is the average PL intensity in the spectral window ℛ and ZPL { , , , , }A B w are fitting parameters. A squared -Lorentzian function instead of a Lorentzian function is used as suggested in Ref. [32] for better fit s at cryogenic temperatures. Temperature dependence of the ratio A is shown as solid marker s in Fig. 2(b) , where the solid and dot ted curves are derived from the two fits of the reduced Debye -Waller factor and the ZPL linewidth shown in Figs. 2(c) and 2(d). We note that the reduced Debye - Waller factor is defined in this work as the ratio of the integra ted ZPL emission , which corresponds to the area under the squared -Lorentzian fit , to the total PL in the range ℛ. Importantly , we find a maximum in the slope of the ratio 𝐴 around T 150 K, which coincidentally corresponds to the glass transition temperature of the PDMS , though does not appear to be related to it (See Supplementary Material, Sec. H and Sec. I [13]). While the stronger temperature response dA dT at lower temperatures observed in this study is desirable for the improved temperature sensing, t he presence of the maximum cannot be explained by a currently existing model , since it predicts a monotonic increase of the temperature response at lower temperatures . This can be resolved by taking into account a constant term ( a) in the linewidth 2w a bT , modifying t he analytical expression of the ZPL amplitude ratio to be (see Supplemental Material, Sec. J [13]) 2 22 exp( ) ()TAa bT R (2) where and are fitting parameters of the reduced Debye -Waller factor and R is the size of the spectral window ℛ. The constant contribution is due both to a resolution of the spectrometer and an inhomogeneous broadening. W avelength resolution can be improved by narrowing down the slit in the spectrometer with a trade -off of the PL count rate. The inhomogeneous broadening is not negligible at lower temperatures due to crystal strain variations both between different NDs and within the individual commercial NDs used in this study . These limitations could be overcome by introducing engineered nanoparticles [33,34] , leading to an enhanced temperature response at cryogenic temperatures . The temperature sensitivity of a thermometer, which is sometimes referred to as the noise floor, is not only quantified by the temperature response dA dT but also by the uncertainty A in the measurement of A . They are related by 1 At dA dT , where t is the measurement time . While the temperature response increases at lower temperatures , A grow s along with the temperature response. To fully characterize the sensitivity of the thermometry technique, we studied the uncertainty A as a function of temperature T . At each temperature, PL spectrum measurements with an integration time of t 2.5 s were repeated one hundred times (F ig. 3(a)). We then calculate the standard deviation A for each data set and show its temperature dependence in F ig. 3(b). Note that the standard deviation A is rescaled by a factor ZPLCt to quantitatively compare the results at different temperatures, where ZPLC is the ZPL count rate shown in the inse t of F ig. 3(c) that corresponds to the area under the squared -Lorentzian fit (see Supplemental Material, Sec. L [13]). The dashed curve shows the lower bound when the noise is coming only from photon shot noise, while the dotted curve shows the lower bound when the CCD camera’s dark-current shot noise also contributes to the noise in the measurement of the ratio A (see Supplementary Material, Sec. M [13]). The experimental observation is well explained by the dotted curve, demonstrating that the standard deviation A is limited both by the ZPL photon shot noise and the CCD’s dark current shot noise. Comb ining the temperature dependencies of A and dA dT as shown in F igs. 3(c) and 2(b), we plot the temperature dependence of the sensitivity in Fig. 3(d) . The lower bounds shown are derived from the same model s as in F ig. 3(c). Importantly, the sensitivity improves at cryogenic temperatures in contrast to the conventional thermometry technique based on the temperature dependent shift in the zero -field splitting . We note that the sensitivity calculated in this study at T 300 K does not reach the level of the sensitivity provided in the previous report on all -optical thermometry at room temperature [11]; however, taking into account detection efficiency differences, our result is found to be fully consistent with the one in Ref. [11]. This can be confirmed by introducing a projected sensitivity proj as shown in F ig. 3(d), wh ich assumes as high ZPL counts rate s as in Ref. [11] and shows an anticipated sensitivity compatible with their result (for detail, see Supplemental Material, Sec. O [13]). The highest temperature sensitivity is achieved near T 200 K, which can be understood through the simplified analytical model that only considers the temperature evolution of the DWF (for detail on the necessary assumptions , see Supplemental Material, Sec. P [13]) 12 tot 011exp2 2 (DWF)TTT C (3) resulting in a minimum at 1T =218 K, where totC is the total PL counts rate of NV- and 0(DWF)T is the (non-reduced) Debye -Waller factor at absolute zero (For the discussion of the effect of the PDMS sheet, see Supplemental Material, Sec. Q [13]). While there is a quantitative mismatch due to oversimplification in the model , this model captures the existence of the minimum well. To further improve the sensitivity at low temperature s, one could, for instance, increase the ZPL count rate by improving the detection efficiency and utilize brighter NDs that contain more NV- centers . III. SURFACE TEMPERATURE IMAGING OF A YIG FILM To demonstrate the applicability of the all -optical thermometer , we apply an 80-mA current to the resistive heater to generate a temperature gradient in the YIG and measure the spatial temperature variation of the YIG surface using an array of NDs, as illustrated in F ig. 1(a) . Since the YIG has spin -wave resonances at microwave frequencies near 𝒟 [14–19], this measurement confirms that the all -optical thermometry technique can be used independently of substrate materials where microwave control is problematic . In the se experiments , the base temperature of the copper heat sink is stabilized at T =170 K (see Supplemental Material , Sec. R [13]). Figure 4(a) shows a two -dimensional spatial scan of the PL from the array of NDs used in this study. To construct the temperature profile, we repeat temperature measurements at multiple spots in the array . The accuracy of the measured temperature is ensured by calibrating NDs individually (see Supplemental Material , Sec. S [13]) and the temperature dependencies of ZPL { , , , }Bw in addition to A are utilized for calculating the local temperature (see Supplemental Material , Sec. T [13]). For each measurement, the PL is collected in total for 500 s. Figure 4(b) shows the resulting temperature profile of the YIG surface, where we observe a temperature decay on the order of tens -of-microns from the heat source. The temperature of each spot as a function of the distance from the heater is shown in F ig. 4(c), where the error bars include both the uncertainty of the sensing and the err or in the calibration. The data is fit well by the Green’s function to the two -dimensional Poisson equation, showing that the temperature field in the YIG approximately follows the steady state diffusion equation with a single heat carrier. We note that the Poisson equation is not accurate in YIG because there are two kinds of heat carriers, phonons and magnons. A deviation from the Poisson equation is expected near the heat source within a length scale of a magnon -phonon thermalization, which is much small er than a few micrometers [35]. In our experiment, NDs directly measure temperatures of the YIG lattice, or phonon s, and we do not observe any perturbation to the qualitative feature of the steady -state phononic temperature profile by the presence of magnons in YIG , which is expected due to our thermal and spatial resolutions . (see Supplementary Material, Sec. U [13]) [20,24] . IV. CONCLUSION We demonstrate and characterize an all-optical thermometry technique based on NV- center ensembles in ND that can be deployed from room temperature to liquid nitrogen temperatures , with a sensitivity that increases with decreasing temperature . Furthermore , the PL intensity of NV- centers is enhanced by implem enting pulse sequences to convert NV0 into NV-, leading to a higher temperature sensitivity by approximately a factor of 3 . Systematic noise analysis reveal s that the sensitivity is limited by the shot noise and the inhomogeneous broadening of the ZPL linewidth , suggesting a pathway for further sensitivity improvements by optimizing the spectral resolution, improving the PL detection efficiency , and introducing engineered NDs with high brightness and homogeneous crystal strain s. Taking advantage of an array of NDs embedded i n a flexible PDMS sheet, w e show the utility of the all -optical thermometer at T =170 K by measuring the surface temperature profile of a YIG slab thermally driven by a resistive heater . This all-optical thermome try technique along with the versatility of the ND membrane array provides a microwave -free, minimally invasive , and cryogenic ally compatible way of measuring local temperatures within a variety of substrate materials . ACKNOWLEDGMENTS This work was supported by the Air Force Office of Scientific Research and the Army Research Office through the MURI program, grant no. W911NF -14-1-0016. The fabrication of the diamond nanoparticle arrays was supported by the US Department of Energy, Office of Science, Basic Energy Sciences, Materials Sciences and Engineering Division. FJH, PFN , and DDA were supported by the US Department of Energy, Office of Science, Basic Energy Sciences, Materials Sciences and Engineering Division. This work made use of shared facilities supported by the NSF MRSEC Program under grant no. DMR -0820054. The autho rs thank P. C. Jerger, B. B. Zhou, C. M. Anderson and J. C. Karsch for useful discussions. References : [1] S. A. Hartnoll, P. K. Kovtun, M. Müller, and S. Sachdev, Phys. Rev. B - Condens. Matter Mater. Phys. 76, 1 (2007). [2] B. I. Halperin and P. C. Hohenberg, Phys. Rev. 188, 898 (1969). [3] G. Kucsko, P. C. Maurer, N. Y. Yao, M. Kubo, H. J. Noh, P. K. Lo, H. Park, and M. D. Lukin, Nature 500, 54 (2013). [4] J. Wang, F. Feng, J. Zhang, J. Chen, Z. Zheng, L. Guo , W. Zhang, X. Song, G. Guo, L. Fan, C. Zou, L. Lou, W. Zhu, and G. Wang, Phys. Rev. B - Condens. Matter Mater. Phys. 91, 1 (2015). [5] P. Neumann, I. Jakobi, F. Dolde, C. Burk, R. Reuter, G. Waldherr, J. Honert, T. Wolf, A. Brunner, J. H. Shim, D. Suter, H. Sumiya, J. Isoya, and J. Wrachtrup, Nano Lett. 13, 2738 (2013). [6] D. M. Toyli, C. F. de las Casas, D. J. Christle, V. V. Dobrovitski, and D. D. Awschalom, Proc. Natl. Acad. Sci. 110, 8417 (2013). [7] D. M. Toyli, D. J. Christle, A. Alkauskas, B. B. Bu ckley, C. G. Van de Walle, and D. D. Awschalom, Phys. Rev. X 2, 1 (2012). [8] M. W. Doherty, V. M. Acosta, A. Jarmola, M. S. J. Barson, N. B. Manson, D. Budker, and L. C. L. Hollenberg, Phys. Rev. B - Condens. Matter Mater. Phys. 90, 1 (2014). [9] P. C. Ts ai, C. P. Epperla, J. S. Huang, O. Y. Chen, C. C. Wu, and H. C. Chang, Angew. Chemie - Int. Ed. 56, 3025 (2017). [10] T. Plakhotnik, Curr. Opin. Solid State Mater. Sci. 21, 25 (2017). [11] T. Plakhotnik, H. Aman, and H. C. Chang, Nanotechnology 26, (2015). [12] T. Plakhotnik, M. W. Doherty, J. H. Cole, R. Chapman, and N. B. Manson, Nano Lett. 14, 4989 (2014). [13] See Supplementary Information (n.d.). [14] P. Andrich, C. F. de las Casas, X. Liu, H. L. Bretscher, J. R. Berman, F. J. Heremans, P. F. Nealey, a nd D. D. Awschalom, Npj Quantum Inf. 1 (2017). [15] C. Du, T. van der Sar, T. X. Zhou, P. Upadhyaya, F. Casola, H. Zhang, M. C. Onbasli, C. A. Ross, R. L. Walsworth, Y. Tserkovnyak, and A. Yacoby, Science (80 -. ). 357, 195 (2017). [16] T. Van Der Sar, F. C asola, R. Walsworth, and A. Yacoby, Nat. Commun. 6, 1 (2015). [17] C. S. Wolfe, V. P. Bhallamudi, H. L. Wang, C. H. Du, S. Manuilov, R. M. Teeling -Smith, A. J. Berger, R. Adur, F. Y. Yang, and P. C. Hammel, Phys. Rev. B - Condens. Matter Mater. Phys. 89, 1 (2014). [18] M. R. Page, F. Guo, C. M. Purser, J. G. Schulze, T. M. Nakatani, C. S. Wolfe, J. R. Childress, P. C. Hammel, G. D. Fuchs, and V. P. Bhallamudi, ArXiv (2016). [19] C. S. Wolfe, S. A. Manuilov, C. M. Purser, R. Teeling -Smith, C. Dubs, P. C. Ham mel, and V. P. Bhallamudi, Appl. Phys. Lett. 108, 2 (2016). [20] M. Schreier, A. Kamra, M. Weiler, J. Xiao, G. E. W. Bauer, R. Gross, and S. T. B. Goennenwein, Phys. Rev. B - Condens. Matter Mater. Phys. 88, 1 (2013). [21] M. Hayashi, L. Thomas, and C. Rettner, Appl. Phys. … 92, 112510 (2008). [22] B. L. Giles, Z. Yang, J. S. Jamison, J. M. Gomez -Perez, S. Vélez, L. E. Hueso, F. Casanova, and R. C. Myers, Phys. Rev. B 96, 1 (2017). [23] C. M. Jaworski, J. Yang, S. MacK, D. D . Awschalom, R. C. Myers, and J. P. Heremans, Phys. Rev. Lett. 106, (2011). [24] K. An, K. S. Olsson, A. Weathers, S. Sullivan, X. Chen, X. Li, L. G. Marshall, X. Ma, N. Klimovich, J. Zhou, L. Shi, and X. Li, Phys. Rev. Lett. 117, 1 (2016). [25] M. F. H. G . Davies, Proc. R. Soc. London A 348, 285 (1976). [26] A. Gali, Phys. Rev. B 1 (2009). [27] S. Felton, A. M. Edmonds, and M. E. Newton, Phys. Rev. B 1 (2008). [28] P. Andrich, J. Li, X. Liu, F. J. Heremans, P. F. Nealey, and D. Awschalom, Nano Lett. 18, 4684 (2018). [29] X. D. Chen, S. Li, A. Shen, Y. Dong, C. H. Dong, G. C. Guo, and F. W. Sun, Phys. Rev. Appl. 7, 1 (2017). [30] X. D. Chen, L. M. Zhou, C. L. Zou, C. C. Li, Y. Dong, F. W. Sun, and G. C. Guo, Phys. Rev. B - Condens. Matter Mater. Phys. 92, 1 (2015). [31] N. Aslam, G. Waldherr, P. Neumann, F. Jelezko, and J. Wrachtrup, New J. Phys. 15, (2013). [32] G. Davies, Reports Prog. Phys. 44, (1981). [33] P. Maletinsky, S. Hong, M. S. Grinolds, B. Hausmann, M. D. Lukin, R. L. Walsworth, M. Loncar, and A. Yacoby, Nat. Nanotechnol. 7, 320 (2012). [34] and F. J. H. S. O. Hruszkewycz, W. Cha, P. Andrich, C. P. Anderson, A. Ulvestad, R. Harder, P. H. Fuoss, D. D. Awschalom, 026105 , (2017). [35] A. Prakash, B. Flebus, J. Brangham, F. Yang, Y. Tserkovnyak, and J. P. Heremans, Phys. Rev. B 020408 , 1 (2018). FIG. 1. (a) Schematic of an array of nanodiamonds (NDs) on a 3.05 -m YIG film grown on a GGG substrate. NDs are embedded on the surface of a flexible PDMS sheet and the YIG film was patterned with a resistive heater (central wire has a width of 5 m and a length of 200 m). (b) Two -dimensional photoluminescence (PL) image of NV centers in NDs collected under continuous 594 -nm excitation. PL intensity is measured by an ava lanche photodiode (APD). The measurement was conducted at T =170 K. Scale bar, 0.5 m. (c) Line cuts of PL intensity profiles of NV centers under two different excitation pulse sequences. (d) Schematic of the pulse sequences of a 532-nm laser ( NV- charge state initialization), a 594 -nm laser ( NV- detection) and a detector (APD/CCD camera). FIG. 2. (a) Evolution of NV centers’ PL spectrum ()Lh between temperatures T =85 K and T =300 K. The areas under the spectra are normalized to one. Discontinuities at T 230 K and T 150 K are associated with the PDMS’s phase transitions and not related to NV centers. Top (bottom) graph shows the spectrum at 300 K (85 K). (b) Temperature dependence of the ZPL amplitude ratio A (left axis) and its temperature respo nse dA dT (right axis). The solid blue curve is calculated from two fits: (i) temperature dependence of the reduced Debye -Waller factor (DWF) (shown in (c)), and (ii) temperature dependence of the ZPL linewidth (shown in (d)). The dotted red curve is the derivative of the solid (blue) curve with respect to temperature T . Inset shows the fit of the ZPL at T =170 K with a sum of an exponential function and a squared -Lorentzian function (black curve). The exponential - function part only is shown with a gray curve. LR is the mean PL intensity in the range ℛ from 605 nm to 660 nm. (c) Reduced DWF as a function of temperature T . A Gaussian -functional fit is shown. (d) ZPL linew idth as a function of temperature T . The solid blue fit is the second -order polynomial 2a bT and the dotted orange curve shows 2bT . FIG. 3. (a) ZPL amplitude ratio scanned over 100 times. A single scan consists of a total PL accumulation time of 2.5 s. The measurements were conducted at temperature T =170 K. (b) Histogram built from the measurements in (a). The standard deviation A is depicted. (c) Rescaled standard deviation ZPL AA Ct as a function of temperature T , where ZPLC is the PL counts rate under the squared -Lorentzian fit of ZPL and t is the total PL accumulation time. The dashed red curve shows the lower bound determined by photon shot noise and the dotted blue curve shows the lower bound dete rmined by photon and dark current shot noise. Inset shows ZPL counts rat e ZPLC as a function of temperature T . Solid black curve shows a one -parameter ( 1a ) fit of the ZPL counts rate () ZPL 1( ) DWFTC T aR , where ()DWFT R is the curve shown in F ig. 2 (c). (d) Temperature sensitivity as a function of temperature T . The dashed red and the dotted blue curves identify the lower bounds for the sensitivity as defined in (c). The spike near 160 K arises from the dip in the experimental data of ZPLC as shown in the inset of (c). Right axis shows a projected sensitivity proj under the assumption of a higher detection rate of the PL as explained in the main text. FIG. 4. (a) Spatial PL scan of NV centers in NDs in the array. (b) Two -dimensional temperature imaging of the YIG surface using NV centers in the array of NDs embedded on the surface of the PDMS sheet measured by the all - optical thermometry technique. An 8 0-mA current is applied to the resistive heater. The base temperature was set to T =170 K. (c) YIG surface temperature as a function of the distance from the resistive heater. Fit with a logarithmic function is shown. Supplemental Information for All-optical cryogenic thermometry based on NV centers in nanodiamonds M. Fukami1, C. G. Yale1,†, P. Andrich1,‡, X. Liu1, F. J. Heremans1,2, P. F. Nealey1,2, D. D. Awschalom1,2,* 1. Institute for Molecular Engineering, University of Chicago, Chicago, IL 60637 2. Institute for Molecular Engineering and Materials Science Division, Argonne National Lab, Argonne, IL 60439 †Present address: Sandia National Laboratories, Albuquerque, NM, 87185 ‡Present address: University of Cam bridge, Cavendish Laboratory, JJ Thomson Ave, Cambridge CB3 0HE *Email: awsch@uchicago.edu A. Temperature Dependence of the Sensitivity in a Range 𝟑𝟎𝟎 𝐊≲𝑻≲𝟒𝟎𝟎 𝐊 Reference [S1] provides a model that explains the temperature dependence of the zero-phonon line ( ZPL) amplitude ratio 𝐴 under temperature 𝑇 in the range 300 K≲𝑇≲400 K. The authors fit the ZPL with a sum of a Lorentzian function and an exponential function, with the coefficient s, 𝐴 and 𝐵, respectively . In the model, t he ratio 𝐴 is proportional to the Debye -Waller factor (DWF) divided by a ZPL linewidth 𝑤. Then t he temperature dependence of the ratio 𝐴 is given by 𝐴=𝛼𝑇−2exp(−𝛾𝑇2), (𝑆1) resulting in the temperature response |𝑑𝐴 𝑑𝑇|=2𝑇(𝑇−2+𝛾)𝐴, (𝑆2) where 𝛼 and 𝛾 are temperature independent con stants which are related to the electron -phonon coupling 𝑆, the Debye temperature 𝑇𝐷 and reference values . We n ote that the DWF is defined as the ratio of the integral ZPL intensity to the total PL. From this expression, the temperature response is expected to be larger at lower temperatures, which potentially give s rise to a higher temperature sensitivity at lower temperatures though it also depends on the uncertainty of the measurement of the ratio 𝐴. The uncertainty 𝜎𝐴 is given by [S2] 𝜎𝐴=𝑓(𝑟)𝐴 √𝐶ZPL 𝛥𝑡(𝑆3) 𝑓(𝑟)=√𝑐1+𝑐2𝑟+𝑐3√𝑟2+𝑟, 𝑤𝑖𝑡ℎ [𝑐1,𝑐2,𝑐3]=[3,3,1], (𝑆4) where 𝑟=𝐵/𝐴, 𝐶ZPL is the ZPL counts rate and 𝛥𝑡 is the measurement time. From the equation (S3) , the temperature sensitivity, or the noise floor, can be written as 𝜂≡𝜎𝐴√𝛥𝑡|𝑑𝑇 𝑑𝐴⁄ |=𝑇𝑓(𝑟) 2(1+𝛾𝑇2 )√𝐶ZPL. (𝑆5) Assuming , for simplicity , that the temperature dependence of the total PL is negligible, we can write 𝐶𝑍𝑃𝐿=𝐶𝑡𝑜𝑡(DWF )|𝑇=0exp(−𝛾𝑇2), (𝑆6) where 𝐶tot is the total PL counts rate and (DWF )|𝑇=0 is the DWF at absolute zero . From equations (S5) and (S6), we get 𝜂=𝑇𝑓(𝑟) 2(1+𝛾𝑇2 )√𝐶tot(DWF )|𝑇=0exp (1 2𝛾𝑇2). (𝑆7) As the temperature decreases , the factor 𝑟=𝐵/𝐴 decreases, which is not shown in the Ref. [S1] but is confirmed in a regime 85 K≤𝑇≤300 K as shown in the F ig. S6(b). Then we get 𝑑𝜂/𝑑𝑇>0, demonstrating a higher sensitivity at lower temperatures , at least in a regime 300 K≲𝑇≲400 K where the model is confirmed . B. Temperature response of the ratio 𝑨 The model described in Sec. A assumes that the temperature response of the ratio 𝐴 is dominated by that of the DWF and the ZPL linewidth. Another possible contribution is the temperature response of the amount of phonon -sideband emission in the range of i nterest (in our case, the spectral range ℛ) with regard to the total PL , though its temperature dependence is negligible as shown in Sec . I. C. Temperature stability of the flow cryostat In the experiment, the base temperature of the sample was stabilized with PID control. Temperature deviation was within ±0.3 K for all measurements. Though the thermocouple was positioned a few centimeters away from the sample position, temperature accuracy within ±0.5 K was ensured in a calibration of the setup which has a thermocouple right next to the sample position. D. Enhancement of the PL at Different Spot s in the Array Figure. 1(c) in the main text shows the enhancement of the PL at 𝑇=170 K with the pulse sequences shown in F ig. 1(d) in the main text. In the Fig. S1, we show the enhancement of the PL at different spots in the array. The figure S1(a) is identical to the Figure. 1(c) in the main text , while figures S1(b) and S1(c) show the PL scans at other spots. Each PL peak was fit by a sum of a Gaussian function and a constant, where the amplitude s of the Gaussian function s were extracted from the fits. The enhancement in the amplitudes due to the pulse sequences was observed, where the enhancement factors were approximately 3.1, 2.4, and 3.4 for F igs. S1(a), S1(b) and S1(c), respectively . Though the factor depends on the spots, enhancement s by approximately a factor of three were observed. The enhancement is due to the charge -state conversion between NV− and NV0. While the 594 -nm excitation preferentially converts NV− into NV0, the 532 -nm excitation preferentially converts NV0 back into NV− [S3–S5]. The time scale of the char ge-state conversion depends on the laser power [S4]. To m inimize the heating while keeping the PL counts high enough in our experiments, the powers of the two lasers were both set to 200 𝜇W, leading to an estimate that the relaxation time of the charge -state conversion is larger than 1 𝜇s. FIG. S1. Enhancement of the PL due to the pulse sequences as shown F ig. 1(d) in the main text under three different spots in the array. (a) shows the same figure as the F ig. 1(c) in the main text. E. Discussion of the practical sensitivity under the pulse sequence Since the sensitivity of the all -optical thermometer is limited by a shot noise, a higher PL count rate by roughly a factor of three results in a higher sensitivity of temperatures by approximately a factor of √3. Note that the pulse sequences also reduce the fraction of the measurement time in the total scanning time. While it improves the physical sensitivity 𝜂 of temperatures, it may result in worse practical sensitivity 𝜂practical if the enhancement factor is less than two. We observed, however , improved sensitivity with the pulse sequences not only because the enhancement is larger than two but also because it reduces the noise due to the CCD dark counts and the background counts . Here we note that physical sensitivity 𝜂 is defined as the minimum temperature difference that can be resolved by a given amount of NV -center -measurement time, while the practical sensitivity 𝜂practical is the minimum temperature difference that can be resolved by a given total time including the time necessary for charge state preparation, background measurement, control of the equipment, and feedback control to focus on the target spot. F. Background Measurement Each spectral measurement was followed by an off -spot background measurement with the same measurement duration. This not only deteriorates the practical sensitivity 𝜂practical , but also adds additional noise to the physical sensitivity 𝜂. The factor is considered in the calculation of the noise model where CCD camera’s dark - current s hot noise also contributes in addition to the ZPL photon shot noise, as explained in Sec. L . G. Choice of the Range 𝓡={𝒉𝝂| 𝒉𝒄(𝟔𝟔𝟎 𝐧𝐦)−𝟏≤𝒉𝝂≤𝒉𝒄(𝟔𝟎𝟓 𝐧𝐦)−𝟏 } The spectral range ℛ is chosen such that it is consistent with previous report s [S1,S6,S7]. With a choice of the grating in our spectrometer (Acton SP -2750, Princeton Instrument , 300 gr/mm with 750 nm blaze ; iStar 334T, Andor ), the range can be measured in the CCD with a single scan . This allowed us to take measurement s without stitching different spectral scans under different angles of the grating in the spectrometer. In contrast, the spectra shown in the F ig. 2(a) in the main text are stitched over multiple scans under different angles of the grating. H. Fitting the reduced Debye -Waller factor and the ZPL linewidth To get a curve for the ratio 𝐴 in the main text , we fit temperature dependencies of a reduced Debye -Waller factor (DWF )ℛ and a ZPL linewidth 𝑤. We fit the temperature dependence of (DWF )ℛ by a Gaussian function (DWF )ℛ=𝛼exp(−𝛾𝑇2), where 𝛼 and 𝛾 are fitting parameters but 𝛾 is related to the electron -phonon coupling 𝑆 and the Debye temperature 𝑇D by a relation 𝛾=2𝜋2𝑆3𝑇D2⁄ [S8]. In our experiment, we measured 𝛾=(218 K)−2 which corresponds to 𝑇D/√𝑆=560 K. The value of 𝛾 was consistent with a measurement conducted on NDs without the PDMS sheet as shown in Sec. I. The temperature dependence of the ZPL linewidth 𝑤 is fit by a second - order polynomial 𝑤=𝑎+𝑏𝑇2. As shown in Figure 2(d) in the main text , the constant contribution 𝑎 is not negligible at lower temperatures in our experiment due to the inhomogeneous broadening. Based on the two fits, we obtained the curve in Fig. 2(b). Based on the model under a simplifying assumption 𝑎≫𝑏𝑇2, one can easily find that the temperature response |𝑑𝐴 𝑑𝑇⁄ | takes maximum at 𝑇≃1√2𝛾 ⁄ =154 K, which is consistent with the exper imental observation. I. Temperature Dependencies of the Parameters without the PDMS Sheet The dependency of the ratio 𝐴 on the temperature shown in F ig. 2(b) in the main text is not largely affected by the presence of the PDMS sheet. To support this statement, we show the PL spectra of NV centers without the PDMS sheet in the F ig. S2 where the spectra were measured at 𝑇=85 K,110 K,150 K,200 K,250 K and 300 K. The measurement was conducted on NDs scattered on a quartz substrate, where hundreds of NDs existed under a laser -focused spot. The same analysis as in the main text is conducted. From the fit of the (DWF )ℛ as shown in F ig. S3(a) we got 𝛾=(283 K)−2, which stays within 25% from the value 𝛾=(218 𝐾)−2 of the NDs embedded in the PDMS array, showing that the presence of the PDMS sheet does not change the main result in this report. When we fit the reduced Debye -Waller factor with a Gaussian function in the main text , there was an implicit assumption that the temperature dependence of (DWF )ℛ is approximately that of (DWF ) since the temperature dependence of the DWF is known to be (DWF )=exp (−𝑆(1+2𝜋2𝑇2 3𝑇D2)). (𝑆8) Figure S3 (a) compares (DWF ) and (DWF )ℛ scanned on NDs without the PDMS sheet. Figure S3(b) shows almost constant ratio (DWF )/(DWF )ℛ over the temperature range 85 K≤𝑇≤300 K, confirming the assumption . FIG. S2. PL spectra and ZPL amplitude ratio of NV centers without PDMS sheet scanned under multiple temperatures 𝑇=85 K,110 K,150 K,200 K,250 K and 300 K. (b) inset shows the spectrum at 𝑇=150 K. FIG. S3. (a) Temperature dependencies of the Debye -Waller factor (DWF ) (left axis) and the reduced Debye -Waller factor (DWF )ℛ (right axis) measured on NDs without the PDMS sheet. (b) Temperature dependence of the ratio (DWF )(DWF )ℛ ⁄ , which equals to the fraction of the integrated PL in the range ℛ to the total PL. J. Temperature Dependence of the ZPL Amplitude Ratio In the fit of the spectrum as shown in the inset of the F ig. 2(b) in the main text, the PL intensity 𝐿(ℎ𝜈) was firstly divided by the mean PL intensity in the range ℛ={ℎ𝜈| ℎ𝑐(660 nm)−1≤ℎ𝜈≤ℎ𝑐(605 nm)−1 }. The mean PL intensity ⟨𝐿⟩ℛ can be explicitly written as ⟨𝐿⟩ℛ=1 𝛥ℛ∫𝐿(𝑛𝜈)ℎ𝑑𝜈𝜈f 𝜈i, (𝑆9) where 𝜈i=𝑐(660 nm)−1, 𝜈f=𝑐(605 nm)−1 and 𝛥ℛ=ℎ(𝜈f−𝜈i). Therefore, the reduced Debye -Waller factor (DWF )ℛ, the ZPL amplitude ratio 𝐴, and the linewidth 𝑤 are related by (DWF )ℛ=𝜋 2𝐴𝑤 𝛥ℛ. (𝑆10) We note that the reduced Debye -Waller factor is defined as the ratio of the integra ted ZPL emission intensity to the total PL in the range ℛ. With the use of the coefficients from the fits of (DWF )ℛ and 𝑤, the ZPL amplitude ratio 𝐴 can be written as 𝐴=2𝛼exp(−𝛾𝑇2)𝛥ℛ 𝜋(𝑎+𝑏𝑇2). (S11) From the equation (S11), we get |𝑑𝐴 𝑑𝑇|=2𝑇(𝑏 𝑎+𝑏𝑇2+𝛾)𝐴. (𝑆12) K. Discussion of the value 𝑻𝐃/√𝑺 From the fit of the reduced Debye -Waller factor in the main text, we obtained 𝑇𝐷/√𝑆=560 K. Though this is relatively small considering the bulk Debye temperature 𝑇Dbulk≃2200 K of diamond, Debye temperatures in nanodiamonds are know n to be around 30% smaller [S6]. Mismatch from the literature value of nanodiamonds 𝑇D/√𝑆|literature=1.0(1)×103 K given in Ref. [S1] would be due to the different ensemble of NVs used in our experiment. While tens of commercial NDs with 100 -nm diameter were used in this study, Ref. [S1] reports meas urements on a single ND with diameter smaller than 50 nm prepared from synthetic sub -micron diamond powder. In the measurement without the PDMS sheet shown in F ig. S3(a), we got similar value 𝛾=(283 K)−2 which corresponds to 𝑇D/√𝑆=725 K. This supports that the smaller value of 𝑇D/√𝑆 measured in our experiment compared to the value in Ref. [S1] is due to the different ensembles of NVs . L. Temperature Dependence of the ZPL Counts Rate There is a subtlety in modeling the temperature dependence of the ZPL counts rate 𝐶ZPL because it is not only determined by the temperature dependence of the NV center’s optical lifetime, but also affected by the temperature dependencies of the steady sta te population of NV− and the PDMS sheet’s optical transparency in our experiment. For simplicity, we conducted a one -parameter ( 𝑎1) fit of the ZPL counts rate 𝐶ZPL(𝑇)=𝑎1(DWF )ℛ(𝑇) where (DWF )ℛ(𝑇) is the curve we got in F ig. 2(c) in the main text. The underlying assumption is that the temperature dependence of 𝐶ZPL is dominated by that of the reduced DWF, which is valid when the integra ted PL intensity in the range ℛ is not significantly temperature dependent compared to th e temperature dependence of (DWF )ℛ. A detailed study of the temperature dependence of the ZPL counts rate is beyond the scope of this report. M. Two Models for the Rescaled Standard Deviation The dotted and dashed curves in the F ig. 3(c) in the main text show lower bounds for the rescaled standard deviation under different models. The dashed curve is the lower bound when the noise is coming only from the photon shot noise of the ZPL and the phonon sideband under the ZPL, while the dotted curve shows the lower bound when the CCD camera’s dark -current shot noise also contributes to the uncertainty 𝜎𝐴. In a model where photon counts under the ZPL peak add noise to the fit of the ZPL, the rescaled standard deviation 𝜎𝐴√𝐶ZPL𝛥𝑡 is given by 𝜎𝐴√𝐶ZPL𝛥𝑡=𝑔(𝑦)𝐴 (𝑆13) 𝑔(𝑦)=√𝑐1+𝑐2𝑦+𝑐3√𝑦2+𝑦, with [𝑐1,𝑐2,𝑐3]≃[2.00,1.98,0.763 ]. (𝑆14) The function 𝑔(𝑦) differs from the case when u sing a Lorentzian function [S2]. The dashed curve is drawn by setting 𝑦=𝐵/𝐴, while the dotted curve is drawn by setting 𝑦=𝐵/𝐴+2𝑐dark/⟨𝐿⟩ℛ̅̅̅̅̅̅𝐴ℎ𝛿𝜈, where the temperature dependence of 𝐵/𝐴 was fit by an exponential function as shown in F ig. S5, 𝛿𝜈 is the frequency range corresponds to one line of vertically binned pixel s in the CCD camera, ⟨𝐿⟩ℛ ̅̅̅̅̅̅≡(1𝑁⁄)∑ ⟨𝐿⟩ℛ(𝑇𝑖) 𝑁 𝑖=1 represents the average of ⟨𝐿⟩ℛ over temperatures 𝑇𝑖={85,90,⋯,300 K},, and 𝑐dark is the counts due to the CCD’s dark current whose average value is cancelled by the background measurement while it adds noise to the spectrum. The dotted curve explains the experimentally observed standard deviation 𝜎𝐴 and the residual would be associa ted with the background counts from the surroundings of NDs such as the PDMS sheet. The noise due to 𝑐dark is non -negligible because the PL is spread over thousands of pixels in the CCD camera in the spectrometer. N. Derivation of the Function 𝒈(𝒚) Applicatio n of the theory given in Ref. [S2] to the case with squared -Lorentzian function gives 𝑔(𝑦)=√𝑓2(𝑦) 𝑓1(𝑦)𝑓2(𝑦)−(𝑓3(𝑦))2√𝜋𝛤(𝛽−1 2) 𝛤(𝛽)(𝑆15) 𝑓1(𝑦)=∫𝑑𝑥∞ −∞((𝑥2+1)𝛽 𝑦(𝑥2+1)𝛽+1)(1 (𝑥2+1)𝛽)2 (𝑆16) 𝑓2(𝑦)=∫𝑑𝑥∞ −∞((𝑥2+1)𝛽 𝑦(𝑥2+1)𝛽+1)(𝑥2 (𝑥2+1)𝛽+1)2 (𝑆17) 𝑓3(𝑦)=∫𝑑𝑥∞ −∞((𝑥2+1)𝛽 𝑦(𝑥2+1)𝛽+1)(𝑥2 (𝑥2+1)𝛽+1), (𝑆18) where 𝛽=2 and 𝛤(𝑥) is the Gamma function. Instead of evaluating them analytically, we computed them numerically and fit the function 𝑔(𝑦) by a form √𝑐1+𝑐2𝑦+𝑐3√𝑦2+𝑦 as shown in the F ig. S4, where {𝑐1,𝑐2,𝑐3} are fitting parameters. The function was well fit by [𝑐1,𝑐2,𝑐3]≃[2.00,1.98,0.763 ]. FIG. S4. Numerical evaluation of the function 𝑔(𝑦) and the fit. Inset shows the residuals. O. Calculation of the projected sensitivity In our experiment, the ZPL counts rates were orders of magnitude smaller than those measured in the former study, where the ZPL counts rate from a single ND was observed to be 𝐶ZPL ,1(295 K)=900 kcps at 𝑇= 295 K [S1], in contrast to our measurement of 𝐶ZPL ,2(295 K)=760 kcps at 𝑇=295 K. High sensitivity all -optical thermometry with 𝜂=300 mK Hz−1/2 was demonstrated with this high ZPL detection rate 𝐶ZPL ,1(295 K). To compare our result with the previous study, we define a projected sensitivity 𝜂proj =√𝐶ZPL ,2(295 K)𝐶ZPL ,1(295 K)⁄ 𝜂 (𝑆19) and it is shown in the right axis of the F ig. 3(d) in the main text. Though the projected sensitivity only gives a rough estimate of a sensitivity given a higher detection efficiency of the PL, it shows our result is consistent with the previous report. P. Temperature Dependence of Sensitivity From the equations (S12 ) and (S13), we get the rescaled sensitivity 𝜂√𝐶ZPL≡𝜎𝐴√𝐶ZPL𝛥𝑡|𝑑𝑇 𝑑𝐴|=𝑔(𝑦) 2𝑇(𝑏 𝑎+𝑏𝑇2+𝛾). (𝑆19) The rescaled sensitivity represents the minimum temperature difference that can be resolved by a single ZPL photon detection. We show the temperature dependence of the rescaled sensitivity in F ig. S5. Two lower bounds due to the models ex plained in the previous section are shown. The equation (S19) gives a low temperature behavior 𝜂√𝐶ZPL∼1/𝑇, which is consistent with the experimental data shown in F ig. S5. FIG. S5. Temperature dependence of the rescaled sensitivity 𝜂√𝐶ZPL. Two curves showing the lower bound due to the two limitations as explained in the F ig. 3 in the main text are shown. The rescaled sensitivity represents the minimum possible temperature difference that can be measured by a single ZPL photon detection. Temperature dependence of the sensitivity can be derived from the equations (S4) and (S19), resulting in 𝜂=𝑔(𝑦) 2𝑇(𝑏 𝑎+𝑏𝑇2+𝛾)√𝐶tot(DWF )|𝑇=0exp (1 2𝛾𝑇2). (𝑆20) Under simplifying approximations 𝑎≫𝑇𝑏2, 𝛾≫𝑏/𝑎 and 𝑔(𝑦)≃𝑔(0)≃√2, we get 𝜂≃1 𝑇𝛾√2𝐶tot(DWF )|𝑇=0exp (1 2𝛾𝑇2), (𝑆21) which gives a minimum at 𝑇=1√𝛾⁄. Q. Discussion of the Effect of the PDMS Sheet on the Sensitivity Measurement Figure. 3(d) in the main text shows non -negligible effects of the PDMS sheet. This is mainly due to the temperature dependence of the absolute ZPL counts rate shown in the inset of F ig.3(c) , which is largely affected by the optical transparency of the PDMS sheet that modifies the PL collection efficiency of our setup. Temperature dependence of the rescaled sensitivity shown in F ig. S5 support s this statement, since there are no observable dips/peaks in the figure. While the inset of F ig. 3(c) and F ig. 3(d) are affected by the existence of the PDMS sheet, the general tendency of these figures are expected to be due to the NV-center’s intrinsic pr operties , since the temperature dependence of the total PL of NV centers below room temperatures are reported to be negligible [S9– S11], leading to the decrease of the ZPL counts rate with temperature increase, due to the Debye -Waller factor . R. Choice of the Base Temperature 𝑻=𝟏𝟕𝟎 𝐊 for the Temperature Imaging of YIG We chose the base temperature of 𝑇=170 K in the measurement of F ig. 4 in the main text . This is because there is a glass transition of PDMS at 𝑇≃150 K. Below the glass transition of PDMS, the proximity of the NDs on the YIG surface is not ensured and the local temperature measurement s become untrustworthy. Above the transition temperature, the NDs are in goo d contact with the YIG surface and they measure the local temperatures of the YIG. Note that the glass transitio n does not affect the main results in other parts of this report since the temperature gradient was not applied. S. Calibration of the temperature sensor For the calibration of the temperature sensors, we conducted multiple scans of the spectrum at 𝑇=170 K and 𝑇=180 K by changing the base temperatures of the copper thermal sink. The average value and the variance of the fitting coefficients {𝐴,𝐵,𝛩,𝑤,𝜈𝑍𝑃𝐿} were extracted. Then we calculated the linear dependence to convert the value of {𝐴,𝐵,𝛩,𝑤,𝜈𝑍𝑃𝐿} into temperatures. The calibration was conducted for each spot in the array. T. Temperature Imaging of YIG As mentioned in the main text, the temperature dependencies of the parameters {𝐵,𝛩,𝑤,𝜈𝑍𝑃𝐿} in addition to 𝐴 were used for temperature sensing by taking the weighted average of the temperatures measured by fitting coefficients . In the Figure S 6, the temperature dependencies of 𝛩,𝐵/𝐴 and 𝜈ZPL are shown, where 𝐵/𝐴 was fit by an exponential function which is empirical but the specific functional form does not matter in this report . The parameter 𝛩 represents the slope of the exponential function in the fit of the phonon sideband. The value is different from the true temperature by a factor of order one, which is called Urbach’s rule and similar dependencies are observ ed in many other materials [S12]. We note that t he temperature dependence of this exponential tail can potentially be used as a temperature sensor below the liquid nitrogen temperatures for future applications. FIG. S 6. Temperature dependencies of (a) 𝛩, (b) 𝐵/𝐴 and (c) 𝜈ZPL. The ratio 𝐵/𝐴 was fit by an exponential function in (b) with a solid (red) curve. Before the temperature measurements, YIG was magnetized to one direction by applying a DC magnetic field. After taking temper ature measurements on multiple spots in the ND array , the temperatures around the scanned spots are smoothly interpolated or extrapo lated and shown in the F ig. 4(b) in the main text . The spots used in the temperature measurement is shown in Fig. S7, where the white circles represent the spots that were used . FIG. S7. Two -dimensional scan of PL and temperatures as shown in the F ig. 4 in the main text with circles representing the spots in the array that were used for the temperature measurement. Temperatures around the scanned spots were smoothly interpolated or extrapolated in (b). U. Discussion of the temperature profile In the F ig. 4(c) in the main text, the YIG surface temperature 𝑇(𝑥) was fit by a logarithmic function 𝑇(𝑥)=−𝜉log(𝑥−𝜁), (𝑆19) where 𝑥 is the distance from the resistive heater and {𝜉,𝜁} are fitting parameters. A logarithmic function is used because the Green’s function to the steady state two -dimensional diffusion equat ion with a single hear carrier is logarithmic. Since the w ire has the length of 200 𝜇𝑚 and the center of the PDMS sheet was displaced to the left by approximately 45 𝜇m due to experimental imperfection , there would be a deviation from the logarithm ic function due to the imperfection of the two -dimensionality , i.e., the resistive heater is not infinitely long . A finite YIG thickness and the existence of an interface between YIG and GGG can also be a potential cause of the deviation from the logarithmic function. The deviation is, however, not clearly observed. In addition, both phonons and magnons are the heat carriers in YIG . According to the coupled magnon - phonon heat transport theory [S13], the steady state phonon temperature profile 𝑇𝑝(𝐱) does not obey a simple Poisson equation, but it obeys (𝜅𝑚𝜅𝑝𝛻4−𝑔(𝜅𝑚+𝜅𝑝)𝛻2)𝑇𝑝=−(𝜅𝑚𝛻2−𝑔)𝑄𝑝+𝑔𝑄𝑚. (𝑆20) Here we ignored , for simplicity, the spatial derivatives of the thermal conductivities, 𝜅𝑚 and 𝜅𝑝. Parameters 𝑄𝑝 and 𝑄𝑚 are the power densities of external heating absorbed by phonons and magnons, respectively [S14]. It is shown in the reference [S14], however, that the equation (S20) can be approximated to the Poisson equation in a regime where the phononic temperature gradient is dominant over the gradient of the magnon -phonon temperature difference . The observed logarithmic behavior of the phononic temperature profile supports this approximation and that the steady -state phononic temperature profile is not largely disturbed by magnons in YIG . For further study of the temperature profile of the YIG film, higher temperature sensitivity is required . References: [S1] T. Plakhotnik, H. Aman, and H. C. Chang, Nanotechnology 26, (2015). [S2] E. A. Donley and T. Plakhotnik, Single Mol. 2, 23 (2001). [S3] X. D. Chen, S. Li, A. Shen, Y. Dong, C. H. Dong, G. C. Guo, and F. W. Sun, Phys. Rev. Appl. 7, 1 (2017). [S4] X. D. Chen, L. M. Zhou, C. L. Zou, C. C. Li, Y. Dong, F. W. Sun, and G. C. Guo, Phys. Rev. B - Condens. Matter Mater. Phys. 92, 1 (2015). [S5] N. Aslam, G. Waldherr, P. Neumann, F. Jelezko, and J. Wrachtrup, New J. Phys. 15, (2013). [S6] T. Plakhotnik, M. W. Doherty, J. H. Cole, R. Chapman, and N. B. Manson, Nano Lett. 14, 4989 (2014). [S7] P. C. Tsai, C. P. Epperla, J. S. Huang, O. Y. Chen, C. C. Wu, and H. C. Chang, Angew. Chemie - Int. Ed. 56, 3025 (2017). [S8] D. B. Fitchen, R. H. Silsbee, T. A. Fulton, and E. L. Wolf, Phys. Rev. Lett. 11, 275 (1963). [S9] A. T. Collins, M. F. Thomaz, M. I. B. Jorge, A. T. Collins, A. T. Collins, A. T. Collins, P. M. Spear, A. T. Collins, A. T. Collins, M. Stanley, A. T. Collins, S. C. Lawson, and J. Walker, 2177 , (1983) . [S10] T. Plakhotnik and D. Gruber, Phys. Chem. Chem. Phys. 12, 9751 (2010). [S11] D. M. Toyli, D. J. Christle, A. Alkauskas, B. B. Buckley, C. G. Van de Walle, and D. D. Awschalom, Phys. Rev. X 2, 1 (2012). [S12] J. D. Dow and D. Redfield, Phys. Rev. B 5, 594 (1972). [S13] M. Schreier, A. Kamra, M. Weiler, J. Xiao, G. E. W. Bauer, R. Gross, and S. T. B. Goennenwein, Phys. Rev. B - Condens. Matter Mater. Phys. 88, 1 (2013). [S14] K. An, K. S. Olsson, A. Weathers, S. Sullivan, X. Chen, X. Li, L. G. Marshall , X. Ma, N. Klimovich, J. Zhou, L. Shi, and X. Li, Phys. Rev. Lett. 117, 1 (2016). | 2019-03-05 | The nitrogen-vacancy (NV) center in diamond has been recognized as a
high-sensitivity nanometer-scale metrology platform. Thermometry has been a
recent focus, with attention largely confined to room temperature applications.
Thermometry has been a recent focus, with attention largely confined to room
temperature applications. Temperature sensing at low temperatures, however,
remains challenging as the sensitivity decreases for many commonly used
techniques which rely on a temperature dependent frequency shift of the NV
centers spin resonance and its control with microwaves. Here we use an
alternative approach that does not require microwaves, ratiometric all-optical
thermometry, and demonstrate that it may be utilized to liquid nitrogen
temperatures without deterioration of the sensitivity. The use of an array of
nanodiamonds embedded within a portably polydimethylsiloxane (PDMS) sheet
provides a versatile temperature sensing platform that can probe a wide variety
of systems without the configurational restrictions needed for applying
microwaves. With this device, we observe a temperature gradient over tens of
microns in a ferromagnetic-insulator substrate (YIG) under local heating by a
resistive heater. This thermometry technique provides a cryogenically
compatible, microwave-free, minimally invasive approach capable of probing
local temperatures with few restrictions on the substrate materials. | All-optical cryogenic thermometry based on NV centers in nanodiamonds | 1903.01605v1 |
0018-9464 (c) 2020 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TMAG.2020.3021646, IEEE Transactions on Magnetics 1 Origin of Perpendicular Magnetic Anisotropy in Yttrium Iron Garnet Thin Films Grown on Si (100) Zurbiye Capku,1,2 Caner Deger,3 Perihan Aksu,4 Fikret Yildiz 1 1Department of Physics, Gebze Technical University, Gebze, Kocaeli, 41400, Turkey 2Department of Physics, Bo ğaziçi University, Beşiktaş, Istanbul, 34342, Turkey 3Department of Physics, Marmara University, Kadikoy, Istanbul, 34722, Turkey 4Institute of Nanotechnology, Gebze Technical University, Gebze, Kocaeli, 41400, Turkey We report the magnetic properties of yttrium iron garnet (YIG) thin films grown by pulsed laser deposition technique. The films were deposited on Si (100) substrates in the range of 15-50 nm thickness. Magnetic characterizations were investigated by ferromagnetic resonance spectra. Perpendicular magnetic easy axis was achieved up to 50 nm thickness. We observed that the perpendicular anisotropy values decreased by increasing the film thickness. The origin of the perpendicular magnetic anisotropy (PMA) was attributed to the texture and the lattice distortion in the YIG thin films. We anticipate that perpendicularly magnetized YIG thin films on Si substrates pave the way for a cheaper and compatible fabrication process. Index Terms —Ferromagnetic Resonance (FMR); Perpendicular Magnetic Anisotropy (PMA); Yttrium Iron Garnet (YIG). I. INTRODUCTION Magnetic garnet films have recently begun to take the place of conducting ferromagnetic materials in spintronic applications. The insulating features of the garnets eliminates the disadvantages of Eddy currents, which causes loss of information in relevant applications [1]. They have attracted great attention for high frequency and fast switching of magnetic properties [2]. In particular, perpendicular magnetization in garnet films is very crucial for the field of spintronics i.e. spin-orbit switching, spin transfer torque and, a reliable and rapid response [3, 4]. Yttrium iron garnet (YIG) is considered to be one of the most important magnetic insulators. Static and dynamic magnetic properties of bulk crystal or YIG films in the micrometer thickness range have been investigated in great detail and widely used in microwave applications (filtering, tunabling, isolators, phase shifters, etc.) [5, 6]. However, the process of thin/ultrathin YIG films plays a key role for spintronic [7-11] and magneto-optical applications [12- 14]. Many spintronic applications require a fine tuning of the orientation and magnitude of the magnetic anisotropy [15, 16] . Perpendicular magnetic anisotropy (PMA) has led to a revolutionary breakthrough in the technology such as the invention of high-density Magnetoresistive Random-Access Memory devices (MRAM). The effective control over the magnetic anisotropy leads to highly remarkable features such as increased data storage capacity in the magnetic recording media, magnon transistor [17] , and advancement in the logic devices [16]. Despite the fact that enhancement of PMA in metal thin films is a well-established phenomenon [18, 19] , generating PMA in insulating materials such as YIG remains a challenge. For such reasons, ferromagnetic insulators with PMA have been of particular importance for both fundamental scientific research and technological applications. Recent developments in the magnonic field have attracted great attention to ultrathin/thin YIG films perpendicularly magnetized. For example; YIG with the possess of PMA, has a unique feature in spin-orbit torque (SOT) applications [20] . The typical anisotropy in YIG films is in-plane anisotropy (IPA) which mainly originates from the strong shape anisotropy. When the magnetocrystalline anisotropy overcomes the shape anisotropy, the direction of the magnetic easy axis switches to the out of the film plane, resulting in PMA. In the literature, the control on magnetic anisotropy in YIG thin films has been studied by various substrate, temperature, and thicknesses [21, 22]. PMA in YIG films were achieved by using a buffer layer [23] and/or doping with rare earth elements [1, 24, 25]. In these studies, garnet substrates such as Yttrium aluminium garnet (YAG) [26] and Gadolinium gallium garnet (GGG) were used to grow epitaxial YIG thin films due to their similar crystalline structure [22, 27]. Lattice constants of YIG film and GGG substrate are aYIG= 12.376 Å and aGGG= 12.383 Å, respectively [28]. This lattice match between YIG and GGG provides high quality crystallized YIG films [29]. However, the use of the GGG substrate in certain areas is limited and also costs much for large area applications. The use of Silicon (Si) as a substrate has many advantages; cost-effectiveness and widespread use in electronic devices and integrated circuits. Si has an fcc diamond cubic crystal structure with a lattice constant of 5.43 Å. The nearest neighbor distance between two Si atoms is 2.35 Å [30]. On the other hand, YIG has a cubic structure consisting of Y3+ ions in dodecahedral (c) sites, Fe3+ ions in tetrahedral (d) and octahedral (a) sites in polyhedron of oxygen ions [31]. The nearest interionic distance in YIG is reported as (Y3+ - O2-) at 2.37 Å [31]. The atomic distances are comparable; thus, one can achieve crystalline / texture YIG on Si (100). In this study, we report the PMA enhancement in YIG films grown on Si substrates by pulsed laser deposition (PLD) technique. Several parameters such as oxygen pressure, substrate temperature, post-annealing treatment, and laser power play an important role in the stoichiometry and Authorized licensed use limited to: University of Edinburgh. Downloaded on September 05,2020 at 02:30:35 UTC from IEEE Xplore. Restrictions apply. 0018-9464 (c) 2020 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TMAG.2020.3021646, IEEE Transactions on Magnetics 2 crystallinity of the YIG films fabricated by PLD. In this report, YIG films with different thicknesses were grown on Si (100) substrates. A post-annealing process was carried out for all films to improve the crystallization and substrate-film lattice mismatch. The effect of the thickness on the magnetic anisotropy values was studied. In some reports, PMA in YIG films was obtained in the thickness range of 10-20 nm [22, 23, 32]. However, in this study, the lattice distortion/texture in YIG films gave rise to PMA in 15-50 nm thickness. We anticipate that our study not only offers a basis for fundamental understanding but also will inspire the integration of perpendicularly magnetized YIG thin films with technological applications. II. EXPERIMENTAL STUDIES Before the thin film deposition, the oxide layer was first etched from the surface of Si (100) substrates with diluted Hydrofluoric (HF) acid for a few minutes. The substrates were further cleaned in acetone, methanol and Isopropyl alcohol for 15 minutes by using an ultrasonic bath. Subsequently, the surfaces were spray dried with Nitrogen gas. Following the chemical cleaning, the substrates were introduced into high vacuum chamber and annealed at 500 oC for an hour. PLD with a KrF excimer laser, a Coherent COMPex Pro 205F operating at λ = 248 nm (20 ns pulse duration) was used to obtain the desired YIG film stoichiometry by adjusting the oxygen pressure and deposition temperature. The base pressure of the deposition chamber was 1.0 x 10-9 mbar. The commercial polycrystalline sintered YIG was used as the deposition target. The distance between the target and substrate was about 60 mm. The films were fabricated using laser energy of 220 mJ at a pulse repetition rate of 10 Hz in an oxygen atmosphere of 1.0 x 10-5 mbar. The substrate temperature was 400 °C during growth. The deposition rate was 0.96 nm/min. The films were cooled within a rate of 9.6 oC/min inside the chamber. Thereafter, the films were annealed at 850 °C for 2 hours in an air atmosphere and cooled down to room temperature by a ratio of 1.2 oC/min. The thicknesses of the annealed films were defined as 15 nm, 20 nm, 35 nm and 50 nm using X-ray reflectivity (XRR) method. Ⅲ. RESULTS Atomic Force Microscopy (AFM) was performed for the surface morphology and roughness of the films. A representative AFM image of the annealed YIG film at 850 o C with a root mean square (RMS) roughness value of 0.8 nm was given at the inset of Fig.1. Structural properties of the films were characterized by X-ray Diffraction (XRD) measurement using a Rigaku 2000 DMAX diffractometer with a Cu (alpha) wavelength of 1.54 nm. The θ -2θ scan XRD pattern was demonstrated in Fig.1. A typical (420) peak of YIG was observed for the annealed films. At each measurement, a signal from the sample plate was detected at 44o. The additional peaks around the (400) plane of the Si substrate correspond to the Kβ, Lα1, Lα2, Kα1 and Kα2 li nes of the incident x-ray. Fig. 1. XRD pattern of YIG thin film on a Si substrate. θ -2θ scan which shows the (420) characteristic peak of 20 nm YIG. (Inset: AFM image of a 20 nm YIG film.) The chemical analysis was performed by X-ray photoelectron spectroscopy (XPS) measurement. The survey scan XPS spectrum is represented in Fig. 2(a). The spectrum confirmed the presence of Y, Fe and O elements on the surface of YIG film grown on Si. The fittings of the spectral ranges related to the elements which are used to determine the composition ratio are given in Figs. 2(b)-2(d). XPS spectrum of the Fe 2p region in Fig. 2(c) shows the valance state of the Fe ions. Y/Fe compositional ratio was found to be 0.59 and Fe/O ratio was 0.44. These values are very close to the bulk YIG compositional ratios within the experimental error (Y/Fe = 0.6 and Fe/O = 0.42 in bulk YIG) [11]. Fig. 2. XPS spectrum of YIG thin film grown on Si substrate. (a) XPS survey scan. Fitted XPS spectrum of (b) O 1s. (c) Y 3p. (d) Fe 2p. Authorized licensed use limited to: University of Edinburgh. Downloaded on September 05,2020 at 02:30:35 UTC from IEEE Xplore. Restrictions apply. 0018-9464 (c) 2020 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TMAG.2020.3021646, IEEE Transactions on Magnetics 3 Ferromagnetic resonance (FMR) measurements were performed to the annealed YIG films by an X-Band (9.1 GHz) JEOL series ESR spectrometer at room temperature. FMR is a powerful technique, which the analysis of the spectra also provides the values of anisotropy constants [33- 35]. The resonance profıle is determined by the field (H) derivative of the absorbed RF power (P) dP/dH curve as a function of the applied magnetic field. The sample dimension for the FMR measurement was around 3 mm × 3 mm. The FMR spectra were registered by sweeping the applied field angle around the sample plane (SP) and sample normal (SN). In SP measurement, the magnetic component of the microwave field is al ways in the film plane, whereas the external magnetic field is rotated from the film plane towards the film normal. In SN measurement, the magnetic component of the microwave field is perpendicular to the film plane and the external magnetic field is rotat ed in the film plane for each spectrum. Representative FMR spectra of the YIG films in SP configuration were given in Fig. 3(c) for the applied field direction along with the film normal and in the film plane (Figs. 3(a) and 3(b)). When the applied field was parallel to the film normal, the spectrum was at low field (red spectra), whereas it shifted to higher field (black spectra) when the applied field was parallel to the film plane, for all samples. This behavior refers that the easy axis of the magnetic anisotropy is perpendicular to the film plane. In SN geometry measurements, there was no any anisotropic behavior, which is not surprising. Thin films having PMA do not represent any anisotropic behavior in the film plane [18, 23, 36] . Further analysis on intrinsic magnetic properties of the system is performed by angular FMR measurements and numerical calculations. To reveal the micromagnetic parameters of YIG/Si (100) structure, the energy Hamiltonian presented in Eq.1 is employed and numerically solved. (1) The Hamiltonian consists of two energy terms used to represent the magnetic behavior of the systems [33, 37]. Here, (θ, θH) and (φ, φH) are, respectively, the polar and azimuth angles for magnetization vector M and external DC magnetic field vector H with respect to the film plane. External DC magnetic field is represented by the first term of the Hamiltonian, i.e., Zeeman energy. Effective magnetic anisotropy energy consists of the demagnetization energy, the interface energy and the first-order term of magnetocrystalline energy of the system. And, the last term represents the second- order magnetocrystalline energy. In Eq. (1), Meff, Keff, and Keff_q are the effective magnetization, effective magnetic anisotropy energy density, second order term of magnetocrystalline energy density, respectively. We scan the DC magnetic field from 0 to 1 T to determine the field corresponding to the maximum value of the dynamic susceptibility, which is called as the resonance field (Hres). Dynamic susceptibility spectra are recorded by using the Soohoo formulation for ferromagnetic resonance in multilayer thin films [38- 40]. Fig. 3. FMR spectra of the YIG films in SP measurement geometry. (a) The applied field direction is along the film normal (H // [001] of Si substrate) and (b) along the film plane (H // [100] of Si substrate). (c) The black and red lines indicate the FMR spectrum when the magnetic field is parallel (H // [100]) and perpendicular (H // [001]) to the film plane. The resonance fields are extracted from the recorded spectrum for SP geometry. By performing the aforementione d procedure for different angles of the magnetic field with respect to the film plane, we are able to reproduce the experimental data. All calculations were performed at room temperature. Meff, Keff, and Keff_q were obtained by the simulation model for all samples. Here, the total energy was minimized with the values of the magnetic parameters given in Table Ⅰ. The angular dependence of the resonance field for different thicknesses in SP geometry is shown in Fig. 4. Table Ⅰ represents the result of the nu merical calculations. The positive effective anisotropy energy density confirms that the easy axis is perpendicular to the film plane. The effective magnetization is lower than the bulk YIG value, which may be Authorized licensed use limited to: University of Edinburgh. Downloaded on September 05,2020 at 02:30:35 UTC from IEEE Xplore. Restrictions apply. 0018-9464 (c) 2020 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TMAG.2020.3021646, IEEE Transactions on Magnetics 4 caused by possible crystal vacancies / deficiencies, inter diffusion between the substrate and the film, and Fe ion variation in the film [10]. In general, shape anisotropy is in the film plane . The increase in thickness strengthens the contribution of shape anisotropy to in -plane magnetic anisotropy, while the strain between th e substrate -film tends to relax and, therefore, t he effective perpendicular magnetic anisotropy reduces by increasing thickness as seen in Table Ⅰ. Fig. 4. A plot of angular variation of FMR resonance fields. Symbols and solid lines indicate the experimental and theoretical results; respectively. Table Ⅰ. Magnetic parameters obtained from the simulation for YIG thin films wit h PMA. Thickness(nm) Keff (J/m3) Keff_q(J/m3) Meff (kA/m) 15 nm 1773 320 105 20 nm 1456 250 105 35 nm 1260 150 105 50 nm 1180 120 105 Ⅳ. DISCUSSIONS In this section, the structural and magnetic characterization of YIG films grown on Si (100) will be discussed. The crystallization of the films was analyzed by XRD measurements. Since we did not get the characteristic XRD peaks in as-grown films, an annealing process was required to generate the YIG phase [28]. After annealing at the temperature of 850 °C for 2 hours, we were able to observe (420) peak of the YIG from θ -2θ scan of the XRD measurement as seen in Fig. 1, which indicates the formation of the YIG phase. Some studies report the polycrystalline YIG film grown on quartz with three characteristic peaks [27, 32] . However, it seems that there is a preferentia l crystalline ordering or texturing in our films. When the film was annealed, the lattice of YIG locates on Si by making an angle of 26.6o between (400) plane of Si and (420) plane. There are two different crystallographic orientations / domain of the film lattice repeating each other on the substrate with respect to the symmetry axis of c, as shown in Fig. 5. The lattice mismatch is , where “a” is the lattice constant. Since the lattice constant of YIG is larger than that of the substrate, a small compressive strain occurred at the interface, which can lead to a tetragonal distortion of the crystalline structure of the film. Fig. 5. A two-dimensional configuration of the lattice orientation of the film on Si substrate. Two preferential crystalline orderings/texturing were present in YIG after the annealing procedure. The composition and electronic state of the Y- Fe-O elements on the surface of the YIG film were investigated by XPS analysis. The stoichiometry of the YIG film was determined by the percentage of O 1s, Fe 2p and Y 3p XPS peak areas shown in Figs. 2(b)-2(d) using relative sensitivity factors in CasaXPS software. The stoichiometry of our samples was found to be Y: 3.06, Fe: 5.17, O: 11.7 which is close to the expected one for Y:Fe:O of 3:5:12. The Fe percentage in our YIG film stoichiometry is 20.6 % which is similar to the ratio of 20% in bulk YIG as known from the literature [11]. Fig. 2d shows the core level Fe 2p spectra. Both Fe3+ and Fe2+ are present in the films [41]. 711.1 eV and 724.4 eV are the binding energy values for the 2p3/2 and 2p1/2 peaks of Fe3+and Fe2+. Two peaks at 710.9 eV and 725.8 eV correspond to Fe3+ 2p3/2 and Fe3+ 2p1/2, and the binding energies at 708.86 eV and 724.16 eV refer to Fe2+ 2p3/2 and Fe2+ 2p1/2, respectively. The satellite structure of Fe 2p 3/2 was located at 718.8 eV, binding energy higher than 710.9 eV. This shows that the Fe ions are in +3 valance states in the spectrum and located at tetrahedral sites of YIG lattice [28, 42, 43]. Representative FMR spectra of the samples are given in Fig. 3 for the external magnetic field parallel to the film normal and film plane. Resonance field is determined by taking the minimum value when applied along the easy axis. Authorized licensed use limited to: University of Edinburgh. Downloaded on September 05,2020 at 02:30:35 UTC from IEEE Xplore. Restrictions apply. 0018-9464 (c) 2020 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TMAG.2020.3021646, IEEE Transactions on Magnetics 5 The spectra clearly show PMA for all thicknesses. The FMR spectrum for H // [001] orientation of Si shifts towards higher field values as the thickness increases. In contrast, for H // [100] orientation of Si, the FMR spectrum shifts to lower values while increasing the thickness. This behavior indicates that the uniaxial perpendicular magnetic anisotropy decreases as the thickness increases. Magnetic properties of the films can be affected by many factors such as thickness, substrate, interfacial energy, and strain. The strain is compressive by 1.9% and tensile by 0.65% when the film is grown on Si (100) and GGG, respectively. In our case, the strain in the YIG films grown on the Si substrate is much greater than that grown on GGG. Due to the lattice mismatch between Si and YIG, available thickness with PMA, the value of magnetic anisotropies and FMR linewidth were different from studies using lattice-compatible substrates in the YIG thin film fabrication process. In this study, all the YIG thin films had a single uniform ferromagnetic resonance peak. In Fig.3, t he shape and intensity of the FMR spectra vary depending on the thickness, crystalline quality, and magnetic homogeneity of films. Inhomogeneous broadening of the linewidth of the FMR spectra is due to the imperfections in film such as defects, roughness, symmetry breaking in surface and interface, oxygen vacancies, and inter ion diffusion between the layers [27]. The thinnest film has a wider and lower intensity FMR profile while the spectra gets clearer with the increase in thickness. Meanwhile, surface and interfacial strain effects show tendency to decrease and the increase in the amount of spins which interact with microwave field increases the FMR intensity. The reason for the FMR shape and intensity which do not vary in a systematic manner might be due to some uncontrollable parameters during deposition. However, it is observed that the out -of-plane magnetic anisotropy behavior of the films still exist in the pronounced thicknesses. The FMR linewidth was determined as the distance between the minimum and maximum point of the dP/dH curve, so called peak to peak lin ewidth. 20 nm YIG film has a linewidth of 230 Oe when the applied field is parallel to the sample plane and linewidth of 160 Oe when the applied field is perpendicular to the sample plane. The linewidths of the spectra are relatively larger compared to the reported value on GGG substrate [44]. However, there are similar linewidth values of that grown on quartz in the literature, as well [22, 27] . For example, 12 nm YIG film was reported to have an FMR linewidth of 250 Oe. For the film thickness range between 100 nm and 290 nm, linewidth values were between 340 Oe and 70 Oe. It is thought that defects due to the surface roughness and Fe iron deficiency may lead to magnon scatterings and increase of the FMR linewidth [11]. It is known that the magnetic easy axis in most of thin films are in the film plane due to the shape/dipolar anisotropy. Additional factor is necessary to overcome the shape anisotropy and switch the orientation of the easy axis from the film plane to the film normal. The crystalline or surface anisotropy or textured structure can trigger a perpendicular magnetic anisotropy [36]. Here, the lattice mismatch between Si and YIG thin film induced a compressive strain at the interface which led to a distortion of the lattice structure [4] . The compressive strain in the film plane results in an expansion along the c-axis, which switches the easy axis from the film plane to the film normal [36, 45]. In previous studies, PMA was realized in YIG films grown on different substrates in the thickness range of 10-20 nm [23, 32]. However, we achieved PMA up to 50 nm thickness as a result of texture and the lattice distortion of YIG. In the literature, PMA in YIG films was observed in those grown on the buffer layer except GGG [23, 46]. This study indicates that PMA was attained successfully in YIG films on a non-garnet substrate without using any additional buffer layer or doping. Ⅴ. CONCLUSION YIG thin films with perpendicular magnetic anisotropy can pave the way for cutting-edge magnonic and spin-related technologies, i.e. for the fast response in microwave devices, logic devices, spin-transfer torque and magneto-optical device applications. Existence of the PMA in an insulator material is a rare magnetic phenomenon. In this work, we have achieved perpendicular magnetization in YIG thin films grown on Si substrate which is a common and base material of the present- day electronic industry. The effect of post annealing- temperature on the crystal structure and magnetic anisotropy was explored. XRD analysis revealed that the crystallization of YIG films improved after annealing. The compressive strain due to the lattice mismatch between Si and YIG led to a distortion in the YIG films, resulting in PMA in the thickness range of 15-50 nm. As far as our best knowledge, we report PMA in pure YIG thin films grown on Si substrate for the first time. We anticipate that perpendicular magnetized YIG thin films will allow the YIG magnetic insulator to be widely used in many areas. ACKNOWLEDGEMENTS The authors are grateful to Dr. Ilhan Yavuz for his fruitful discussions on the results. REFERENCES [1] H. Wang, C. Du, P. C. Ham mel, and F. Yang, “Strain -tunable magnetocrystalline anisotropy in epitaxial Y 3 Fe 5 O 12 thin films,” Physical Review B, vol. 89, no. 13, pp. 134404, 2014. [2] A. Quindeau, C. O. Avci, W. Liu, C. Sun, M. Mann, A. S. Tang, M. C. Onbasli, D. Bono, P. M. Vo yles, and Y. Xu, “Tm3Fe5O12/Pt heterostructures with perpendicular magnetic anisotropy for spintronic applications,” Advanced Electronic Materials, vol. 3, no. 1, pp. 1600376, 2017. [3] K. Garello, C. O. Avci, I. M. Miron, M. Baumgartner, A. Ghosh, S. Auff ret, O. Boulle, G. Gaudin, and P. Gambardella, “Ultrafast magnetization switching by spin- orbit torques,” Applied Physics Letters, vol. 105, no. 21, pp. 212402, 2014. [4] S. M. Zanjani, and M. C. Onbasli, “Thin film rare earth iron garnets with perpendicul ar magnetic anisotropy for spintronic applications,” AIP Advances, vol. 9, no. 3, pp. 035024, 2019. [5] R. Linares, R. McGraw, and J. Schroeder, “Growth and Properties of Yttriu m Iron Garnet Single‐Crystal Films,” Journal o f Applied Physics, vol. 36, no. 9, pp. 2884-2886, 1965. [6] J. Robertson, “Liquid phase epitaxy of garnets,” Journal of Crystal Growth, vol. 45, pp. 233-242, 1978. Authorized licensed use limited to: University of Edinburgh. Downloaded on September 05,2020 at 02:30:35 UTC from IEEE Xplore. Restrictions apply. 0018-9464 (c) 2020 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TMAG.2020.3021646, IEEE Transactions on Magnetics 6 [7] L. Jin, D. Zhang, H. Zhang, X. Tang, F. Bai, Z. Zhong, X. Fan, and J. Q. Xiao, “Spin valve effect of the interfacial spin accumulation in yttrium iron garnet/platinum bilayers,” Applied Physics Letters, vol. 105, no. 13, pp. 132411, 2014. [8] M. Jungfleisch, V. Lauer, R. Neb, A. Chumak, and B. Hillebrands, “Improvement of the yttrium iron garnet/platinum interface for spin pumping - based applications,” Applied Physics Letters, vol. 103, no. 2, pp. 022411, 2013. [9] A. Kehlberger, K. Richter, M. C. Onbasli, G. Jakob, D. H. Kim, T. Goto, C. A. Ross, G . Götz, G. Reiss, and T. Kuschel, “Enhanced magneto - optic Kerr effect and magnetic properties of CeY 2 Fe 5 O 12 epitaxial thin films,” Physical Review Applied, vol. 4, no. 1, pp. 014008, 2015. [10] M. Onbasli, A. Kehlberger, D. Kim, G. Jakob, M. Kläui, A. Chumak, B. Hillebrands, and C. Ross, “Pulsed laser deposition of epitaxial yttrium iron garnet films with low Gilbert damping and bulk-like magnetization,” Apl Materials, vol. 2, no. 10, pp. 106102, 2014. [11] Y. Sun, Y.-Y. Song, H. Chang, M. Kabatek, M. Jantz, W. Schneider, M. Wu, H. Schultheiss, and A. Hoffmann, “Growth and ferromagnetic resonance properties of nanometer-thick yttrium iron garnet films,” Applied Physics Letters, vol. 101, no. 15, pp. 152405, 2012. [12] O. Galstyan, H. Lee, A. Babajanyan, A. Hakhoumian, B. Friedman, and K. Lee, “Magneto -optical visualization by Bi: YIG thin films prepared at low temperatures,” Journal of Applied Physics, vol. 117, no. 16, pp. 163914, 2015. [13] M. Nur-E-Alam, M. Vasiliev, V. A. Kotov, and K. Alameh, “Highl y bismuth-substituted, record-performance magneto-optic garnet materials with low coercivity for applications in integrated optics, photonic crystals, imaging and sensing,” Optical Materials Express, vol. 1, no. 3, pp. 413-427, 2011. [14] M. Shone, “The technology of YIG film growth,” Circuits, Systems and Signal Processing, vol. 4, no. 1-2, pp. 89-103, 1985. [15] D. Sander, “The magnetic anisotropy and spin reorientation of nanostructures and nanoscale films,” Journal of Physics: Condensed Matter, vol. 16, no. 20, pp. R603, 2004. [16] B. Tudu, and A. Tiwari, “Recent developments in perpendicular magnetic anisotropy thin films for data storage applications,” Vacuum, vol. 146, pp. 329-341, 2017. [17] A. V. Chumak, A. A. Serga, and B. Hillebrands, “Magnon transistor for all- magnon data processing,” Nature communications, vol. 5, pp. 4700, 2014. [18] F. Yildiz, M. Przybylski, and J. Kirschner, “Perpendicular anisotropy and oscillatory interlayer coupling in Fe 0.5 Co 0.5/Rh/Fe 0.5 Co 0.5 bilayers on Rh (001),” J ournal of Applied Physics, vol. 105, no. 7, pp. 07C312, 2009. [19] F. Yildiz, M. Przybylski, and J. Kirschner, “Direct Evidence of a Nonorthogonal Magnetization Configuration in Single Crystalline Fe 1− x Co x/Rh/Fe/Rh (001) System,” Physical review letter s, vol. 103, no. 14, pp. 147203, 2009. [20] C. Guo, C. Wan, M. Zhao, H. Wu, C. Fang, Z. Yan, J. Feng, H. Liu, and X. Han, “Spin -orbit torque switching in perpendicular Y3Fe5O12/Pt bilayer,” Applied Physics Letters, vol. 114, no. 19, pp. 192409, 2019. [21] H. Maier- Flaig, S. Geprägs, Z. Qiu, E. Saitoh, R. Gross, M. Weiler, H. Huebl, and S. Goennenwein, “Perpendicular magnetic anisotropy in insulating ferrimagnetic gadolinium iron garnet thin films,” arXiv preprint arXiv:1706.08488, 2017. [22] E. Popova, N. Keller, F. Gendron, L. Thomas, M.-C. Brianso, M. Guyot, M. Tessier, and S. Parkin, “Perpendicular magnetic anisotropy in ultrathin yttrium iron garnet films prepared by pulsed laser deposition technique,” Journal of Vacuum Science & Technology A: Vacuum, Su rfaces, and Films, vol. 19, no. 5, pp. 2567-2570, 2001. [23] J. Fu, M. Hua, X. Wen, M. Xue, S. Ding, M. Wang, P. Yu, S. Liu, J. Han, and C. Wang, “Epitaxial growth of Y3Fe5O12 thin films with perpendicular magnetic anisotropy,” Applied Physics Letters, vol . 110, no. 20, pp. 202403, 2017. [24] B. Calhoun, J. Overmeyer, and W. Smith, “Ferrimagnetic resonance in gadolinium iron garnet,” Physical Review, vol. 107, no. 4, pp. 993, 1957. [25] Y. Kigami, T. Namikawa, and Y. Yamazaki, “Stress Induced Magnetic Anisotropy of YBi 2 Fe 5- x Ga x O 12 Sputtered Films,” IEEE Translation Journal on Magnetics in Japan, vol. 5, no. 4, pp. 319-324, 1990. [26] M. Kubota, A. Tsukazaki, F. Kagawa, K. Shibuya, Y. Tokunaga, M. Kawasaki, and Y. Tokura, “Stress -induced perpendicular magnetization in epitaxial iron garnet thin films,” Applied Physics Express, vol. 5, no. 10, pp. 103002, 2012. [27] B. Bhoi, N. Venkataramani, R. Aiyar, and S. Prasad, “FMR and magnetic studies on polycrystalline YIG thin films deposited using pulsed laser,” IEEE Transactions on Magnetics, vol. 49, no. 3, pp. 990 -994, 2013. [28] Y. Zhang, J. Xie, L. Deng, and L. Bi, “Growth of phase pure yttrium iron garnet thin films on silicon: the effect of substrate and postdeposition annealing temperatures,” IEEE Trans actions on Magnetics, vol. 51, no. 11, pp. 1-4, 2015. [29] C. Tang, M. Aldosary, Z. Jiang, H. Chang, B. Madon, K. Chan, M. Wu, J. E. Garay, and J. Shi, “Exquisite growth control and magnetic properties of yttrium iron garnet thin films,” Applied Physics Le tters, vol. 108, no. 10, pp. 102403, 2016. [30] B. Schultrich, Tetrahedrally Bonded Amorphous Carbon Films I: Basics, Structure and Preparation: Springer, 2018. [31] S. Geller, and M. Gilleo, “The crystal structure and ferrimagnetism of yttrium-iron garnet , Y3Fe2 (FeO4) 3,” Journal of Physics and Chemistry of solids, vol. 3, no. 1-2, pp. 30-36, 1957. [32] E. Popova, N. Keller, F. Gendron, M. Guyot, M.-C. Brianso, Y. Dumond, and M. Tessier, “Structure and magnetic properties of yttrium– iron– garnet thin films prepared by laser deposition,” Journal of Applied Physics, vol. 90, no. 3, pp. 1422-1428, 2001. [33] C. Deger, P. Aksu, and F. Yildiz, “Effect of Interdot Distance on Magnetic Behavior of 2- D Ni Dot Arrays,” IEEE Transactions on Magnetics, vol. 52, no. 12, pp. 1-4, 2016. [34] M. Farle, “Ferromagnetic resonance of ultrathin metallic layers,” Reports on progress in physics, vol. 61, no. 7, pp. 755, 1998. [35] F. Yildiz, S. Kazan, B. Aktas, S. Tarapov, L. Tagirov, and B. Granovsky, “Ferromagnetic resonance studies on (Co40Fe40B20) x (SiO2) 1− x granular magnetic films,” Journal of magnetism and magnetic materials, vol. 305, no. 1, pp. 24-27, 2006. [36] F. Yildiz, M. Przybylski, X.- D. Ma, and J. Kirschner, “Strong perpendicular anisotropy in Fe 1− x Co x alloy films epitaxially grown on mismatching Pd (001), Ir (001), and Rh (001) substrates,” Physical Review B, vol. 80, no. 6, pp. 064415, 2009. [37] T. Kalaycı, C. Deger, S. Akbulut, and F. Yildiz, “Tuning magnetic properties of non-collinear magnetization configuration in pt/[pt/co] 6/pt/co/pt multilayer structure,” Journal of Magnetism and Magnetic Materials, vol. 436, pp. 11-16, 2017. [38] R. Soohoo, “General spin -wave dispersion relations,” Physical Review, vol. 120, no. 6, pp. 1978, 1960. [39] R. Soohoo, “Ex citation and boundary effects in spin-wave resonance,” Journal of Applied Physics, vol. 32, no. 3, pp. S148 -S150, 1961. [40] R. Soohoo, “Ferromagnetic and Spin -Wave Resonance in Multilayer Films,” Journal of Applied Physics, vol. 63, no. 8, pp. 3829 -3829, 1988. [41] G. Siegel, M. C. Prestgard, S. Teng, and A. Tiwari, “Robust longitudinal spin-Seebeck effect in Bi- YIG thin films,” Scientific reports, vol. 4, pp. 4429, 2014. [42] A. Abidov, B. Allabergenov, J. Lee, H.-W. Jeon, S.-W. Jeong, and S. Kim, “X-ray photoelectron spectroscopy characterization of Fe doped TiO2 photocatalyst,” Int. J. Mater. Mech. Manuf, vol. 1, no. 3, pp. 294, 2013. [43] L. Jin, K. Jia, Y. He, G. Wang, Z. Zhong, and H. Zhang, “Pulsed laser deposition grown yttrium-iron-garnet thin films: Effect of composition and iron ion valences on microstructure and magnetic properties,” Applied Surface Science, vol. 483, pp. 947-952, 2019. [44] C. Hauser, T. Richter, N. Homonnay, C. Eisenschmidt, M. Qaid, H. Deniz, D. Hesse, M. Sawicki, S. G. Ebbing haus, and G. Schmidt, “Yttrium iron garnet thin films with very low damping obtained by recrystallization of amorphous material,” Scientific reports, vol. 6, pp. 20827, 2016. [45] S. M. Zanjani, and M. C. Onbaşlı, “Predicting new iron garnet thin films with perpendicular magnetic anisotropy,” Journal of Magnetism and Magnetic Materials, vol. 499, pp. 166108, 2020. [46] P. Sellappan, C. Tang, J. Shi, and J. E. Garay, “An integra ted approach to doped thin films with strain-tunable magnetic anisotropy: powder synthesis, target preparation and pulsed laser deposition of Bi: YIG,” Materials Research Letters, vol. 5, no. 1, pp. 41-47, 2017. Authorized licensed use limited to: University of Edinburgh. Downloaded on September 05,2020 at 02:30:35 UTC from IEEE Xplore. Restrictions apply. | 2021-07-12 | We report the magnetic properties of yttrium iron garnet (YIG) thin films
grown by pulsed laser deposition technique. The films were deposited on Si
(100) substrates in the range of 15-50 nm thickness. Magnetic characterizations
were investigated by ferromagnetic resonance spectra. Perpendicular magnetic
easy axis was achieved up to 50 nm thickness. We observed that the
perpendicular anisotropy values decreased by increasing the film thickness. The
origin of the perpendicular magnetic anisotropy (PMA) was attributed to the
texture and the lattice distortion in the YIG thin films. We anticipate that
perpendicularly magnetized YIG thin films on Si substrates pave the way for a
cheaper and compatible fabrication process. | Origin of Perpendicular Magnetic Anisotropy in Yttrium Iron Garnet Thin Films Grown on Si (100) | 2107.05591v1 |
Spin wave based tunable switch between superconducting flux qubits Shaojie Yuan1*, Chuanpu Liu2*, Jilei Chen2*,Song Liu1, Jin Lan5, Haiming Yu2†, Jiansheng Wu†1, Fei Yan1, Man-Hong Yung1, Jiang Xiao3†, Liang Jiang4†, Dapeng Yu1† 1Institute for Quantum Science and Engineering, and Department of Physics, Southern University of Science and Technology, Shenzhen 518055, China 2Fert Beijing Research Institute, School of Electronic and Information Engineering, BDBC, Beihang University, 100191 Beijing, China 3Department of Physics and State Key Laboratory of Surface Physics, Fudan University, Shanghai 200433, China 4Pritzker School of Molecular Engineering, The University of Chicago, Chicago, Illinois 60637, USA 5Center of Joint Quantum Studies and Department of Physics, School of Science, Tianjin University, Tianjin 300350, China * Equally contributed authors. † Corresponding authors Quantum computing hardware has received world-wide attention and made considerable progress recently. YIG thin film have spin wave (magnon) modes with low dissipation and reliable control for quantum information processing. However, the coherent coupling between a quantum device and YIG thin film has yet been demonstrated. Here, we propose a scheme to achieve strong coupling between superconducting (SC) flux qubits and magnon modes in YIG thin film. Unlike the direct √𝑵 enhancement factor in coupling to the Kittel mode or other spin ensembles, with N the total number of spins, an additional spatial-dependent phase factor needs to be considered when the qubits are magnetically coupled with the magnon modes of finite-wavelength. To avoid undesirable cancelation of coupling caused by the symmetrical boundary condition, a CoFeB thin layer is added to one side of the YIG thin film to break the symmetry. Our numerical simulation demonstrates avoided crossing and coherent transfer of quantum information between the flux qubit and the standing spin waves in YIG thin films. We show that the YIG thin film can be used as a tunable switch between two flux qubits, which have modified shape with small direct inductive coupling between them. Our results manifest that it is possible to couple flux qubits while suppressing undesirable cross-talk. Quantum computing and simulation based on superconducting qubits have achieved significant progress in recent years (1-3). Many efforts were devoted to hybridizing the solid-state qubits with other physical systems, such as mechanical or magnetic systems (4-9). For instance, the Kittel mode of a macroscopic YIG sphere was coherently coupled to a transmon qubit in a 3D cavity with the microwave photons manipulated inside the cavity (8). Besides, the superconducting flux qubit was successfully hybridized with spin ensembles, i.e., nitrogen-vacancy (NV) centers in diamond via magnetic interaction (4-6). On the other hand, because of the zero Joule heating, the wave nature with microwave working frequency, spin wave (whose quanta is called magnon) has become a promising candidate for conventional information transmission and processing and acquired the potential to establish a spin-wave based computing technology, far beyond its CMOS counterpart (10-16). Due to its favorably low damping, ferrimagnetic insulator yttrium iron garnet (YIG) is particularly promising for these applications (17-19). In this work, we propose a novel hybrid system, consisting of superconducting flux qubits and the standing spin waves (20) in ferrimagnetic YIG thin film. The latter system has been widely used in spintronics and magnonics (17-19), while, its magnetic coupling to superconducting qubits and the corresponding application in quantum information processing has not been extensively investigated. As shown in the following, unlike the coupling to spin ensembles or Kittel mode of spin waves (4, 5, 7), the enhancement factor for the coupling strength does not follow the √𝑁 law, but carries a modulation associated with the finite spin-wave wavelength. In our proposal, an additional thin pinning layer of CoFeB is deposited on one side of the YIG thin film to break the symmetry at the boundary conditions (21-30). Avoided crossing of the energy spectrum can be numerically simulated by solving Heisenberg equation based on the full Hamiltonian of the flux qubit, the spin waves in the YIG thin film and their coupling. We find that it is possible to transfer quantum information coherently between the flux qubit and the spin wave mode in the YIG thin film. Moreover, we propose an experimentally feasible design to switch “on” and “off” the coupling between two shape-modified flux qubits or to entangle them via the perpendicular standing spin waves (PSSWS) of the YIG thin film. Hybridizing one flux qubit with and further “tuning” the inductive coupling, which causes cross-talk (31), between multiple flux qubits or entangling them through PSSWs highlights the application of spin wave bus in quantum computing, further expanding the application of spin wave-based computation technology (32-35). A superconducting (SC) loop with three Josephson junctions compose of a flux qubit with the superposition of the clockwise and counter clockwise persistent currents state as the qubit ground state: |g>=|↺>−|↻> and first excited state, |e>=|↺>+|↻> (36, 37), respectively. The net currents and the resulting the magnetic field threading the loop for the |g> and |e> states are distinct. Consequently, the Rabi oscillation between the two states of the flux qubit generates an alternating magnetic field perpendicular to the SC loop, which can be used to excite spin waves in YIG system. The basic setup of the hybrid system is shown schematically in Fig. 1, which consists of a 5x 5 µm! superconducting loop and a 3x 0.08x 3 µm" YIG thin film above. A much thinner CoFeB capping layer ~ 10 nm in thickness is deposited on the top side of the YIG thin film to pin the magnetization in YIG at the interface. The magnetization follows Dirichlet boundary condition at the pinned surface, and Neumann boundary condition at the other free surface (21-27). The resonant frequencies of the perpendicular standing spin wave (PSSW) modes are (20), 𝑓#$$%=&'!!(23𝐻)*++!,"#'!-$5.(/6!73𝐻)*++!,"#'!-$5.(/6!+𝑀07 (1) with gyromagnetic ratio &!(=28 GHz/T, vacuum permeability 𝜇1=1.256 ∗1023 𝑁/𝐴!, saturation magnetization 𝑀$=192 kA/m for YIG, thicknessδ=80 nm, the exchange constant 𝐴)*=3.1 pJ/m, external field 𝐻456 and mode number 𝑛=1,2,3,…, . Values of 𝑀$ and 𝐴45 are obtained by fitting the resonance of a 295 nm YIG thin film from Ref (19) using equation (1) with mode number n=1,2,3,4,5,6. Experiment has measured the resonance value for the PSSW mode of 80 nm YIG thin film at near zero external field to be 4.57 GHz, which is different from theoretical prediction 3.39 GHz. The discrepancy may be due to choosing of order parameter to be integer for unsymmetrical pinning in the fitting process, as actually there are ¾ wavelength in thickness direction for n=1 as illustrated in Fig. 1 b. For our quantum control schemes, we will use the experimental resonance values and design a flux qubit with transition frequency close to 𝑓#$$%(.89) and sufficiently detuned from the CoFeB resonance. Using the geometric confinement, the proper boundary conditions and the suitable coupling strength (see Eq. 4 with later discussion), the PSSW of wavelength of λ=;"δ=;* =1" nm can be excited. An external field of 10 Gauss is applied to align spins in YIG and the field created by the YIG thin film and the CoFeB capping layer on the flux qubit is of the same order (see Fig 1. c), assuming the spin density 𝑛>?@4A=1.61x 10!B 𝑚2" for CoFeB and 𝑛CDE=2.14x10!= 𝑚2" for YIG. The distance between the flux qubit and the YIG thin film is chosen to be around 1-1.5 µm for later simulation in Fig.3. At these distances, the total magnetic field on the qubit is between 21.5 to 37 Gauss, which is less than the critical field of the aluminum superconductor (around 100 Gauss) and guarantees superconductivity of the flux qubit. In addition, other superconducting material such as Niobium can be used to fabricate the loop and junctions of the flux qubit, which has a much higher critical magnetic field for superconductivity, i.e., above 1000 Gauss. From Ref (19), the decay rates for YIG thin film and CoFeB pinning layer are estimated as ΓFGH,J8!~40 MHz and ΓKLM)N~300 MHz, where n is the PSSW mode number. Since the decay rate is proportional to the frequency and the frequency is approximately proportional to the square of mode number, intrinsic decay rate for n=1 PSSW mode is ~ 10 MHz. The resonance frequency for n=1 PSSW mode in YIG thin film and CoFeB pinning layer are 𝑓CDE~4.6 GHz and𝑓>?@4A~1.35 GHz and the exchange coupling strength is 𝑔KLM)N,FGH~500 MHz, which makes converted decay rate of CoFeB on n=1 PSSW mode as Γ>?@4A→CDE,.89=(P%&'()*+,-Q+,-2Q%&'())!∗300~7 MHz and total decay rate for n=1 PSSW mode being 17 to 20 MHz. In our proposal, we replace the microwave antenna in Ref (19) a flux qubit loop, which has a lower decay rate and a much smaller inductive coupling with the sample, and we expect the magnon decay rates will be further reduced. Therefore, it is reasonable to assume that the decay rate for n=1 PSSW mode is about 20-30 MHz. Fig. 1 Hybrid structure coupling a flux qubit and a YIG thin film (spin wave, or a magnon). (a) A YIG thin film with the dimension of 3 x 0.08 x 3 µm" is placed in the center of the 5 x 5 µm! flux qubit loop separated by a distance d. An external field of 10 Gauss is applied along the x axis to align the spins in the YIG thin film. The thickness of YIG thin film is δFGH=80 nm. A perpendicular standing spin wave (n = 1) with wavelength λ=;"δ is excited. The frequency of the flux qubit is chosen to match that of the spin wave, which is experimentally around 4.7 GHz. The alternating magnetic field in the flux qubit loop excites the spin wave in the YIG thin film. (b) A cartoon depicts the PSSW of mode number n=0,1,2,3 with bottom spins being pinned and top spins unpinned. In our proposal, n=1 mode is selected. (c) The magnetic field on the flux qubit created by the YIG thin film and CoFeB thin layer. The spins in both YIG and CoFeB are fully aligned along x. In the following, we consider the coupling strength between the flux qubit and YIG thin film. Hamiltonian for a ferromagnetic or ferrimagnetic material in a magnetic field takes the following general form (38) 𝐻o=−∑𝑆r𝐦𝐦,𝐧.𝐽𝐦,𝐧.𝑆r𝐧+ ∑𝑆r𝐦𝐦.𝐵 (2) with coupling matrix J𝐦,𝐧 between the spins, with the assumptions that deviations from the group state are small, we can perform the Holstein–Primakoff approximation and transforming into the momentum space, we obtain 𝐻o=−𝐽𝑁𝑆!+𝑁𝑆𝜇N𝐵T+∑∑(ℏ𝜔U."V89+𝜇N.𝐵T)𝑎𝐤†𝑎𝐤A.Y.𝐤+∑𝑎𝐤†A.Y.𝐤∑(2𝑆)/0𝜇Nx9√[𝑒\𝐤.𝐦5A#]\A1!6z+𝑎𝐤∑(2𝑆)/0𝜇Nx9√[𝑒2\𝐤.𝐦5A#2\A1!6z𝐦𝐦 (3) Where i=1,2,3, ℏωU.=2JS(1−coskV)=4JSsin!5U.!6, 𝑆 is total spin at each lattice site, 𝑘⃗ is the wavevector of the spin wave and 𝑁 is the number of lattice sites in each direction. From Equation (3), by replacing the summation over each site with integration over space and insert the spin density, the integral form of the coupling strength between the flux qubit and YIG thin film is obtained as following: 𝑔4QQ^_⃗~(!a)/0∫c').42344⃗.744⃗d)892):0ef;g(∫cf;g)/0 (4) where B(x,y,z) is the microwave excitation field created by the flux qubit and 𝜌 is the spin density in YIG. Unlike the simple √𝑁 enhancement associated with coupling to Kittel mode, there is an extra spatial-dependent phase factor 𝑒\^_⃗.g⃗ in Eq. (4). For long wavelength spin wave, |𝒌|≪1 𝜇𝑚29 and 𝑒\𝐤.𝐫 ~1and if 𝐵* or 𝐵i only vary slowly compared to 9|𝒌| in real space, 𝑔4QQ𝐤 will be proportional to √𝑁. However, for the short wavelength spin wave, 𝑔4QQ𝐤 is not necessarily proportional to √𝑁 and can even be zero if the integration region covers exactly integer times of the wavelength along wavevector direction. This is also the reason, to excite PSSW mode in YIG thin film by an almost homogenous field, an asymmetric boundary condition is required to avoid zero coupling strength caused by the phase factor. Given the dimension of the flux qubit square loop 5x 5 µm! and the persistent current I ~ 500 nA (39, 40), the magnetic field produced by the flux qubit can be evaluated using Ampere’s law 𝐵(r)='!;(∮𝐼fl____⃗×gn___⃗|gn___⃗|; and 𝐵o dominates while, 𝐵5, 𝐵p is close to zero in Fig .1. Given a net spin density 𝜌=2.14x 10!= 𝑚2" in YIG, we obtain the absolute value 𝑔4QQ^ as a function of the separation distance d between of the coupling strength between the flux qubits and the YIG thin film as in Fig. 2: Fig. 2, Coupling strength |𝒈𝐞𝐟𝐟𝐤| as a function of the separating spacing d. For small distance (d<2 µm), |𝒈𝒆𝒇𝒇𝒌| for 𝑘⃗o=!(u , where y is the direction in Fig. 1, decreases slowly with the distance and is above 30 MHz, which is larger than decay rate of magnon in YIG thin film. For large d, |𝒈𝐞𝐟𝐟𝐤| decreases as 𝑑2" indicated by the red curve. With the coupling strength estimated above, the full Hamiltonian with the flux qubit and YIG thin film can be written as 𝐻o=−𝐽𝑁𝑆!+𝑁𝑆𝜇N𝐵T+∑(ℏ𝜔𝐤+𝜇A.𝐵)𝑎𝐤]𝑎𝐤A.Y.𝐤+v!(∆𝜎5+𝜀𝜎p)−ℎ𝑔4QQU𝑎𝐤†+𝑔4QQ^∗𝑎𝐤𝜎p+ℎ𝜆cos𝜔𝑡.𝜎p (5) where the first and second terms are the ferromagnetic and Zeeman terms, the third term describes the spin wave excitation, the fourth term is the flux qubit with ∆ the tunneling energy splitting and 𝜀 being the energy bias between the two qubit states, the fifth term characterizes interaction between the two devices, and the last term is the external driving of the flux qubit. Here, 𝜎5,p are the Pauli matrices. The first two terms can be neglected for the reason that spin wave energy is a small perturbation compared to these two energies. By changing the basis of the flux-qubit, neglecting the Zeeman splitting and performing the rotating wave approximation, the Hamiltonian becomes, 𝐻o=v!(√∆!+𝜀!−𝜔)𝜎p+(ℏ𝜔^+𝜇A.𝐵)𝑎^]𝑎^+ℎ∆√∆0]y0𝑔4QQ^𝑎^]𝜎2+𝑔4QQ^∗𝑎^𝜎]+ℎ∆√∆0]y0u!(𝜎]+𝜎2) (6) Where 𝜎],𝜎2are rasing and lowering operator. 𝜎p=2∗(𝜎]𝜎2−9!) . Approximating the flux qubit as a harmonic oscillator and let 𝜎]→𝑐̂] and 𝜎2→𝑐̂, the Hamiltonian can be written in a different form. Employing the Heisenberg relation fẑf6=[𝑐̂,𝐻], sloving in Fourier space and transforming back to the lab frame, we obtain simulation of the energy spectrum 𝜎2,|~ 9|2√∆0]y0]\}<=2~P(<<3∆?∆09@0~0/(|2|AB]\}AB) (7) with 𝜔 being the driving pulse frequency and 𝜔$% being the resonance frequency of the standing spin wave of the YIG thin film. The expression of Eq. (7) describes the spectroscopic measurement of the flux qubit hybridized with spin waves in YIG thin film. Chossing parematers as ∆!(=4.52 GHz, }'C!(=2 MHz, |AB!(=4.57 GHz and }AB!(=20 MHz, which is a rsonable number since the decay rate for Kittel spin wave in a perfect sphere is around 1 MHz (7) and for finite wavelength spin wave in the YIG thin film is 6.8 MHz at 20 mk with GGG substrate and 1.4 MHz without substrate (41), let |P(<<3|!(=0 MHz and 30 MHz, we obtain a simulated spectrum for a bare qubit and a hybridized qubit-spin wave system, as shown in Fig.3. The avoided cross or gap shows the strong coupling between flux qubit and standing spin wave of YIG thin film with vacuum Rabi splitting 2g = 60 MHz, which supports coherent energy or information exchange between them. Before preceding further, let us have a brief discussion about the influence of the CoFeB thin layer on the flux qubit. Using Eq. 4, with long wavelength approximation (𝑘~0) and spin density of CoFeB being 1.61∗10!B 𝑚2" (Co), and 𝑑=1.2 𝜇𝑚 as the parameter chosen in Fig. 3, a rough coupling strength between flux qubit and Co thin layer is 200 MHz. Decay rate for CoFeB is Γ>?@4A ~ 300 MHz and the converted influence on the flux qubit from Co electrons would be Γ>? *$"∆%$~ 1.2 MHz, where 𝑔 is the coupling strength and ∆ is the off resonance between the flux qubit and CoFeB. We may introduce the the damping constant 𝛼=}Q , where Γ is the decay rate and 𝑓 is the resonace frequency. For YIG, 𝛼 is on the order of 102 to 102;, which makes decay rate as small as 3.3 MHz at a resonace of 4.57 GHz, most possibly by improving the thin film growing quality. In addition, a low ferromagntic alloy Co25Fe75 with damping constant as low as 5∗102; is reoprted. This material could substitute the CoFeB capping layer, which would have the decay rate ΓKL!M)<1 MHz instead of ΓKLM)𝐁~300 MHz and decrease the total decay rate of YIG-pinning layer to below 5 MHz. These further ensure the possibilities to implement thicknees mode of YIG thin film in quantum information processing. Fig. 3 Simulation of the energy spectrum of a flux qubit coupled to standing spin waves in the YIG thin film. (a) Spectrum of a bare flux qubit with ∆=4.52 GHz, Γ@=2 MHz and 𝑔4QQ^=0 in Eq. (7). (b) Spectrum of a flux qubit coupled to the standing spin wave of the flux qubit with |P(<<3|!(=30 MHz, |AB!(=4.57 GHz, }AB!(=20 MHz. Next, we propose a scheme to entangle and further switch the coupling “on” and “off” between two shape-modified flux qubits through PSSW mode in YIG thin film. Fig. 4 shows the schematic: two modified flux qubits with center-to-center distance of 20√2 𝜇𝑚, are placed on top of a YIG thin film with a vertical separation d. The left/right arc of a flux qubit is a quarter of a 10 𝜇𝑚 radius circle and the top/down arc is a quarter of a 13.2 𝜇𝑚 radius circle. Mutual inductance of the two loops is given by the Neumann formula 𝐿,.='!;(∮∮f𝑿Df𝑿E|𝑿D2𝑿E|. The designed orientations of those arcs are to decrease the mutual inductance between the two flux qubit loops from several tens MHz for comparable size square loops to 3.97 MHz for the current design with circulating current as much as 500 nA. YIG thin film is ~ 80 nm in thickness with left/right sides being a quarter of a 10 𝜇𝑚 circle and top/down sides having the length of 10√2 𝜇𝑚, which is also deposited with 10 𝑛𝑚 CoFeB on one side. As oscillation occurs between the two states of a flux qubit, alternating magnetic fields are created outside the loop and Fig. 4 (d) shows the coupling strength between each flux qubit and the YIG thin film as a function of the distance d in between. As shown in Fig. 4, stray magnetic field created by the YIG-CoFeB thin film is below the superconducting critical field of material of Niobium, i.e., 1000 Gauss, that is used to fabricate the flux qubit. Readout of a flux qubit can be realized via another shaped-modified squid loop as in Fig. 4 c. Mutual inductance between the squid loop and flux qubit is 3.8∗10299 𝐻, while the one between the squid loop and the neighboring flux qubit is 5.6∗1029; 𝐻. This guarantees that reading-out flux qubit will not be influenced much by the state of neighboring qubit, even operating simultaneously. Microwave line which is not shown, can quickly tune flux quit resonance frequency to the frequency of the (PSSW) spin wave mode of 4.57 GHz. At distance 𝑑=0.5 𝜇m, the absolute value of coupling strength is about 50 MHz. If both flux qubits are detuned simultaneously to 630 MHz below 4.57 GHz, effective coupling strength J between the two flux qubits can be J~"!""%!∆!&!∆"'$≈−3.97 MHz (8) This will cancel the mutual inductive coupling between (+ 3.97 MHz) the two flux qubits loops, thus switching off the coupling. On the other hand, if detuning both flux qubit to 400 MHz above 4.57 GHz, J would be 6.25 MHz, and plus additional mutual inductive 3.97 MHz, the total coupling strength would be about 10 MHz. Since the intrinsic life time for flux qubit can be about 1 µs, coupling strength of 10 MHz is strong enough to entangle the two qubits. In this way the coupling between two flux qubits is switched “on” and “off‘. In addition, the intrinsic decay rate of thickness mode spin wave in YIG thin film is about ΓCDE =10 MHz, which will introduce an extra broadening of 10*51;116!=0.15 MHz on the flux qubit. Similarly, the CoFeB thin layer gives rise to another 300*5!1""116!~0.01 MHz broadening on flux qubit. Fig. 4 Proposed setup for a tunable switch between two shape-modified flux qubits utilizing (with) YIG thin film. (a) two shape modified flux qubits are placed at a distance d above the 80 nm thick YIG thin film, which is capped with 10 nm CoFeB layer on one side. Special geometry of flux qubits is to decrease mutual inductance and detail dimensions of both flux qubits and YIG thin film are given in the context. (a) the sideview (b) the top view. (c) a special designed squid loop used for reading out the state of flux qubit. Mutual inductance between flux qubit and squid loop is given in the context and reading out one flux qubit will not be influenced much by the neighboring qubit. (d) the absolute value of effective coupling strength (left axis) between one flux qubit and YIG thin film and the total magnetic field (right axis) at point p as in (a) created by YIG thin film as a distance of d. As demonstrated above, different from coupling to spin ensembles or Kittel mode of spin waves, the coupling of the flux qubit with finite-wavelength (fundamental) spin wave mode has an extra phase term, which enables us to obtain the coupling strength and proposed a scheme to hybridize flux qubit with a perpendicularly standing spin wave in the YIG thin film. We further show the PSSW spin wave mode in an YIG thin film can switch “on” and “off” the coupling between two flux qubits and generate entanglement. Our results manifest that it is possible to couple flux qubits while suppressing cross-talk. This opens a possibility of utilizing YIG thin film for quantum information processing. The authors thank Huaiyang Yuan, Peihao Huang, Xiuhao Deng for fruitful discussions. This work is supported by Key-Area Research and Development Program of GuangDong Province (No. 2018B030326001), the National Key Research and Development Program of China (2016YFA0300802), the National Natural Science Foundation of China (Grants No. 11704022, No. U1801661), the Guangdong Innovative and Entrepreneurial Research Team Program (Grant No. 2016ZT06D348), the Science, Technology and Innovation Commission of Shenzhen Municipality (Grant No. KYTDPT20181011104202253). 1. A. D. Córcoles et al., Nature Communications 6, 6979 (2015). 2. R. Barends et al., Nature 508, 500 (2014). 3. M. Mirrahimi et al., New Journal of Physics 16, 045014 (2014). 4. D. Marcos et al.,Physical Review Letters 105, 210501 (2010). 5. X. Zhu et al., Nature 478, 221 (2011). 6. X. Zhu et al., Nature Communications 5, 3424 (2014). 7. Y. Tabuchi et al., Physical Review Letters 113, 083603 (2014). 8. Y. Tabuchi et al., Science 349, 405 (2015). 9. D. Lachance-Quirion et al., Science Advances 3, e1603150 (2017). 10. I. L. Markov, Nature 512, 147 (2014). 11. Y. Kajiwara et al., Nature 464, 262 (2010). 12. V. V. Kruglyak, S. O. Demokritov, D. Grundler, Journal of Physics D: Applied Physics 43, 264001 (2010). 13. B. Lenk, H. Ulrichs, F. Garbs, M. Münzenberg, Physics Reports 507, 107-136 (2011). 14. A. V. Chumak, V. I. Vasyuchka, A. A. Serga, B. Hillebrands, Nature Physics 11, 453 (2015). 15. S. Dutta et al., Scientific Reports 5, 9861 (2015). 16. W. Yu, J. Lan, J. Xiao, Physical Review Applied 13, 024055 (2020). 17. T. Schneider et al., Applied Physics Letters 92, 022505 (2008). 18. A. V. Chumak, A. A. Serga, B. Hillebrands, Nature Communications 5, 4700 (2014). 19. C. S. Davies et al., IEEE Transactions on Magnetics 51, 1-4 (2015). 20. H. Qin, S. J. Hämäläinen, S. van Dijken, Scientific Reports 8, 5755 (2018). 21. C. Kittel, Physical Review 110, 1295-1297 (1958). 22. P. E. Wigen, C. F. Kooi, M. R. Shanabarger, T. D. Rossing, Physical Review Letters 9, 206-208 (1962). 23. R. F. Soohoo, Physical Review 131, 594-601 (1963). 24. Y. S. Gui, N. Mecking, C. M. Hu, Physical Review Letters 98, 217603 (2007). 25. R. Magaraggia et al., Physical Review B 83, 054405 (2011). 26. S. Klingler et al., Journal of Physics D: Applied Physics 48, 015001 (2014). 27. M. A. W. Schoen, J. M. Shaw, H. T. Nembach, M. Weiler, T. J. Silva, Physical Review B 92, 184417 (2015). 28. S. Klingler et al., Physical Review Letters 120, 127201 (2018). 29. J. Chen et al., Physical Review Letters 120, 217202 (2018). 30. S. J. H. H. Qin, and S. van Dijken, Sci. Rep. 8, (2018). 31. D. Abraham et al., Quantum Computing Architecture. 27003 (2013). 32. A. Khitun, M. Bao, K. L. Wang, Journal of Physics D: Applied Physics 43, 264005 (2010). 33. A. V. Chumak, V. I. Vasyuchka, A. A. Serga, B. Hillebrands, Nature Physics 11, 453-461 (2015). 34. G. Csaba, Á. Papp, W. Porod, Physics Letters A 381, 1471-1476 (2017). 35. J. Lan, W. Yu, R. Wu, J. Xiao, Physical Review X 5, 041049 (2015). 36. T. P. Orlando et al., Physical Review B 60, 15398-15413 (1999). 37. J. E. Mooij et al., Science 285, 1036-1039 (1999). 38. A. Altland, Condensed Matter Field Theory 2nd Edition. (Cambridge University Press, ed. 2 2010). 39. J. Bylander et al., Nature Physics 7, 565 (2011). 40. F. Yan et al., Nature Communications 7, 12964 (2016). 41. S. Kosen, A. F. van Loo, D. A. Bozhko, L. Mihalceanu, A. D. Karenowska, APL Materials 7, 101120 (2019). | 2020-04-05 | Quantum computing hardware has received world-wide attention and made
considerable progress recently. YIG thin film have spin wave (magnon) modes
with low dissipation and reliable control for quantum information processing.
However, the coherent coupling between a quantum device and YIG thin film has
yet been demonstrated. Here, we propose a scheme to achieve strong coupling
between superconducting flux qubits and magnon modes in YIG thin film. Unlike
the direct $\sqrt{N}$ enhancement factor in coupling to the Kittel mode or
other spin ensembles, with N the total number of spins, an additional spatial
dependent phase factor needs to be considered when the qubits are magnetically
coupled with the magnon modes of finite wavelength. To avoid undesirable
cancelation of coupling caused by the symmetrical boundary condition, a CoFeB
thin layer is added to one side of the YIG thin film to break the symmetry. Our
numerical simulation demonstrates avoided crossing and coherent transfer of
quantum information between the flux qubits and the standing spin waves in YIG
thin films. We show that the YIG thin film can be used as a tunable switch
between two flux qubits, which have modified shape with small direct inductive
coupling between them. Our results manifest that it is possible to couple flux
qubits while suppressing undesirable cross-talk. | Spin wave based tunable switch between superconducting flux qubits | 2004.02156v1 |
Engineering Entangled Coherent States of Magnons and Phonons via a Transmon Qubit Marios Kounalakis,1, 2,∗Silvia Viola Kusminskiy,2and Yaroslav M. Blanter1 1Kavli Institute of Nanoscience, Delft University of Technology, 2628 CJ Delft, The Netherlands 2Institute for Theoretical Solid State Physics, RWTH Aachen University, 52074 Aachen, Germany (Dated: September 29, 2023) We propose a scheme for generating and controlling entangled coherent states (ECS) of magnons, i.e. the quanta of the collective spin excitations in magnetic systems, or phonons in mechani- cal resonators. The proposed hybrid circuit architecture comprises a superconducting transmon qubit coupled to a pair of magnonic Yttrium Iron Garnet (YIG) spherical resonators or mechanical beam resonators via flux-mediated interactions. Specifically, the coupling results from the mag- netic/mechanical quantum fluctuations modulating the qubit inductor, formed by a superconducting quantum interference device (SQUID). We show that the resulting radiation-pressure interaction of the qubit with each mode, can be employed to generate maximally-entangled states of magnons or phonons. In addition, we numerically demonstrate a protocol for the preparation of magnonic and mechanical Bell states with high fidelity including realistic dissipation mechanisms. Furthermore, we have devised a scheme for reading out the prepared states using standard qubit control and res- onator field displacements. Our work demonstrates an alternative platform for quantum information using ECS in hybrid magnonic and mechanical quantum networks. I. INTRODUCTION The development of quantum technologies aims to- wards disruptive practical applications in several fields such as computing, communication and sensing by ex- ploiting the effects of quantum mechanics [1, 2]. The success of this venture largely relies on the evolution of hybrid quantum systems that incorporate the advantages of different physical platforms in a constructive way [3, 4]. For example, circuit quantum electrodynamics (QED), where light-matter interactions in superconducting cir- cuits are used to manipulate quantum information, is one of the leading platforms in quantum computing, combin- ing strong nonlinearities with advanced quantum control and readout as well as high coherence times relative to qubit operations [5, 6]. However, superconducting cir- cuits do not directly couple to optical photons, hindering their integration with optical networks [3]. In this direc- tion, the development of hybrid circuit QED platforms based on mechanical and magnetic systems is an essen- tial requirement towards networked quantum computa- tion [4]. In addition, the evolution of high-quality me- chanical systems operating in the quantum regime pro- vides unique opportunities not only in transduction but also in building quantum memories and sensors [3, 4, 7]. Moreover, hybrid quantum systems based on magnons, i.e., the quanta of the collective spin excitations in mag- netic materials, offer distinctive advantages, such as uni- directional propagation and chiral coupling to phonons and photons [8, 9], making them prime candidates for technological applications in quantum information sci- ences [10, 11]. The ability to generate entanglement is at the heart of most protocols in quantum information. For macro- scopic mechanical and magnonic resonators, which carry bosonic degrees of freedom and typically operate in thelinear regime, the special class of entangled coherent states (ECS) [12, 13] is of particular interest. Such states exhibit continuous-variable entanglement between different bosonic modes and provide a valuable resource for quantum teleportation [14, 15], quantum computa- tion [16–18] and communication [19, 20]. In addition, ECS are useful for fundamental studies of quantum me- chanics with applications in quantum metrology [21, 22] and tests of collapse models [23, 24]. Macroscopic entanglement between mechanical modes has recently been achieved on aluminum drum res- onators [25, 26] and micromechanical photonic/phononic crystal cavities [27, 28], however, an experimental demon- stration of entanglement between metallic nanobeams such as the ones studied in Refs. [29–31] is currently lack- ing. Furthermore, while entanglement between atomic ensembles has been experimentally realised in an optical setup [32], entangling magnons in two distant magnets still remains a challenge. Recent theoretical proposals have investigated the possibility of entangling magnons in two Yttrium Iron Garnet (YIG) spheres interacting via photons in a microwave cavity. More specifically, in Ref. [33] the emerging Kerr nonlinearity in strongly driven magnons is used, relying on driving the magnon modes far from equilibrium in order to create entangle- ment. In Ref. [34] the nonlinearity stemming from the parametric magnetorestrictive interaction is employed to create magnon-magnon entanglement, although requir- ing a much larger magnetorestrictive coupling strength than experimentally attainable [35]. Alternatively, in Refs. [36, 37] it is shown that two YIG spheres can be en- tangled by driving the magnon-cavity system with strong squeezing fields. However, while the above schemes show promise for creating magnon-magnon entanglement in distant YIG spheres, the absence of a highly controllable nonlinear element, such as a qubit, hinders the gener- ation and control of more complex states and ECS, inarXiv:2309.16514v1 [quant-ph] 28 Sep 20232 particular. Here we propose a scheme for generating ECS of magnons/phonons in a hybrid circuit QED architecture comprising a superconducting transmon qubit and two magnonic/mechanical modes. Concerning magnonic sys- tems, without loss of generality, we consider two YIG sphere modes in a hybrid qubit-magnon setup similar to Ref. [38], where the qubit-magnon coupling is me- diated via a superconducting quantum interference de- vice (SQUID). We showcase a protocol for generating maximally-entangled states such as Bell and NOON states with high fidelity, by exploiting the parametric na- ture of the qubit-magnon radiation-pressure interaction and the transmon quantum control toolbox. Further- more, we analyze a readout scheme for verifying the en- tanglement in the system based on qubit measurements and displacements of the magnon field. Contrary to pre- vious proposals for generating magnon-magnon entangle- ment, there is no need for placing the YIG spheres inside a cavity, therefore, increasing scalability and modularity. Furthermore, we numerically demonstrate the validity of our proposal for entangling SQUID-embedded mechan- ical beam resonators [29–31, 39, 40], thereby extending the possibilities for quantum control using mechanical ECS. II. HYBRID SYSTEM DESCRIPTION The fundamental element in the proposed circuit ar- chitecture is a dc SQUID, i.e., a superconducting loop interrupted by two Josephson junctions, as schematically depicted in Fig. 1. When shunted by a capacitance C, with charging energy EC= 2e2/C, this nonlinear induc- tor can realize a flux-tunable transmon qubit described by the Hamiltonian, ˆHT= 4ECˆN2−EJcosˆδ, (1) where ˆN,ˆδare conjugate operators describing the tunnel- ing Cooper-pairs and the superconducting phase across the SQUID, respectively [41, 42]. In the case where the two junctions are the same (symmetric SQUID), an external flux bias Φ btunes the Josephson energy EJ=Emax J|cosϕb|, where ϕb˙ =πΦb/Φ0and Φ 0is the flux quantum. For magnetic systems, without loss of generality we focus our description on micro-sized YIG spheres sim- ilar to Refs. [38, 43]. Upon application of an in-plane magnetic field Bz, a YIG sphere acquires a magne- tization Msand its excitations can be approximated as a set of independent quantum harmonic oscillators with Hamiltonian ˆHM=ℏP mωmˆa† mˆam, where a(†) m are bosonic operators describing the annihilation (cre- ation) of single magnons [44, 45]. Note that this de- scription is valid in the limit ⟨ˆm†ˆm⟩ ≪ NS, where NSis the total number of spins in the sphere [44, 45]. The fundamental excitation, or Kittel mode, is a uni- EJC 1 EJ2Φb δm^ δm^(a) Bz (b) Bzδx δx^ ^EJ1 EJ2C ΦbFIG. 1. Proposed hybrid circuit architecture. A flux- tunable transmon qubit, formed by a C-shunted SQUID loop, is coupled to (a) two nearby YIG spheres or (b) two SQUID- embedded mechanical beams. The magnetization of both spheres in (a) is oriented by an in-plane field Bz. The mag- netic quantum fluctuations ˆδmmodulate the SQUID flux as well as the transmon inductive energy, thereby giving rise to a qubit-magnon coupling. In (b) the coupling stems from the mechanical quantum fluctuations ˆδxinducing a modulating flux in the SQUID in the presence of the in-plane field Bz. An additional flux bias Φ bcan be externally applied to tune the qubit frequency and modulate the coupling. formly polarized state of all the spins acting as a sin- gle “macrospin” precessing around z, with ferromagnetic resonance (FMR) frequency ω0=γ0(Bz+Bani), where Baniis the anisotropy field [46]. Higher mode frequencies are given by ωm=ω0+γ0Msl−1 3(2l+1)depending on the magnon angular momentum quantum number l[47]. The mechanical systems of interest in this work consist of SQUID-embedded aluminum beams [29, 30, 39, 48]. Such mechanical beams are realised by suspending part of the SQUID loop such that it can freely oscillate out of plane [29, 30]. Similar to the YIG sphere, its ex- citations can also be described by a set of indepen- dent quantum harmonic oscillators, with Hamiltonian ˆHX=ℏP xωxˆa† xˆax, where a(†) xare bosonic operators that annihilate (create) a phonon. The fundamental mode, which is the one considered in this work, oscil- lates with frequency ω0=ℏ/(2mx2 zpf), where mis the beam mass and xzpfthe magnitude of its zero-point mo- tion [29]. Upon application of an in-plane magnetic field Bz, the quantum fluctuations in the out-of-plane displace- ment of the beam ˆδx=xzpf(ˆax+ ˆa† x) induce a flux Φ(ˆδx) =β0Bzlˆδxthrough the loop, where lis the beam length and β0is a geometric factor that depends on the mode shape [29]. Similarly, quantum fluctuations of the magnetic moment in the magnetized YIG sphere, ˆδm=µzpf(ˆam+ ˆa† m), result in an additional flux Φ( ˆδm) through the SQUID loop. Let us assume that the sphere is placed at an in-plane and out-of-plane distance dfrom the closest point in the loop. Then in the far-field limit Φ(ˆδm) =µ0ˆδm/(4√ 2πd) [38]. The additional flux from each source of quantum fluc-3 tuation, Φ( ˆδj), modulates the SQUID flux and conse- quently its Josephson energy, E′ J(ϕb,ˆδj)≃EJ 1−tanϕbX jϕ(ˆδj) , (2) where we assume ϕ(ˆδj) ˙ =πΦ(ˆδj)/Φ0≪1 and a symmetric SQUID; for a full treatment including fi- nite junction asymmetry see Refs. [38, 39]. Re- placing EJwith E′ Jin Eq. (1) and expressing the transmon operators in terms of annihilation (cre- ation) operators ˆ c(†), i.e., ˆN=i[EJ/(32EC)]1/4(ˆc†−ˆc), ˆδ= [2EC/EJ]1/4(ˆc+ ˆc†) [42], yields the total system Hamiltonian ˆH=ˆHq+ℏX jh ωjˆa† jˆaj−gjˆc†ˆc(ˆaj+ ˆa† j)i ,(3) where ˆHq=ℏωqˆc†ˆc−EC 2ˆc†ˆc†ˆcˆc, is the bare transmon Hamiltonian (valid for EJ≫EC), with qubit frequency ωq= (√8EJEC−EC)/ℏ[41]. The last term in Eq. (3) describes the radiation- pressure interaction between the qubit and each bosonic mode, with coupling strength gj=∂ωq ∂ϕjϕzpf j, (4) where ϕzpf jis the magnitude of the flux fluctuations induced by either the beam or the magnet, given by ϕzpf x=πβ0Bzlxzpf/Φ0andϕzpf m=µ0µzpf/(4√ 2dΦ0), re- spectively. In the case of a symmetric SQUID, the trans- mon frequency sensitivity to flux changes is ∂ωq ∂ϕj=ωp 2sinϕb√cosϕb, (5) where ωp=p8Emax JEC/ℏis the Josephson plasma fre- quency at ϕb= 2πk(k∈Z). The behavior of the cou- pling strength as a function of the SQUID asymmetry andϕbis studied in detail in Ref. [38]. III. ENTANGLED COHERENT STATE GENERATION The system Hamiltonian in Eq. (3) describes a qubit interacting with a set of bosonic modes via bipartite radiation-pressure interactions. However, in the ab- sence of additional driving, these radiation-pressure cou- plings lead to interesting dynamics only in the ultra- strong coupling regime, gj≳ωj[39, 49, 50]. Typi- cally, mechanical beam resonators have frequencies of a few MHz [29, 30] and operating magnon frequencies lie above 100 MHz [44, 45], whereas gj≲10 MHz [38, 39]. Therefore, while the ultrastrong coupling condition seems promising for optomechanical setups [39], it is far fromrealistic for magnonic devices. On the other hand, when external driving is introduced to the system, the radiation-pressure interaction can be “activated” even for gj< ωj, e.g., by a stroboscopic application of short π qubit pulses [51] or by modulating the coupling [38, 52]. Here, without loss of generality, we consider the case gj≪ωjand assume that the radiation-pressure inter- action is activated by applying a weak flux modula- tion through the SQUID loop as in Ref. [38]. In this scheme the qubit operates around the transmon “sweetspot”, i.e., ϕb≃0, and an applied ac flux with amplitude ϕacat frequency ωac, modulates the flux, ϕb=ϕaccos (ωact−θ)≪1, resulting in a modulated coupling strength gj(t) =ωp 2ϕzpf jcos (ωact−θ), where θ is a constant phase. In the frame rotating at ωacthe transformed Hamiltonian reads, ˆeH=ˆHq+ℏX jh ∆jˆa† jˆaj−egjˆc†ˆc(ˆajeiθ+ ˆa† je−iθ)i ,(6) where egj=ωp 4ϕzpf j, ∆j=ωj−ωacand we have omit- ted fast-rotating terms ˆ c†ˆcˆa(†) je±i(ωj+ωac)twhich do not contribute to the dynamics since egi≪(ωj+ωac). We now describe a simple protocol for generating ECS that are maximally entangled using the Hamiltonian in Eq. (6). Let us assume there are Nbosonic modes, in- teracting with the qubit via bipartite radiation-pressure couplings. First, a microwave pulse, prepares the qubit in a superposition state |χ⟩q˙ =(|0q⟩+eiχ|1q⟩)/√ 2. The next step is to activate the bipartite interaction of the qubit with each mode. In the simple case where all the modes we want to entangle have the same frequency, ωj, then by turning on the flux modulation, i.e., setting ωac=ωj, for a variable duration, τj, the system evolves into a hybrid generalized Greenberger–Horne–Zeilinger state |ψ⟩GHZ=1 N |0q01···0N⟩+eiχ|1qα1···αN⟩ ,(7) where |αj⟩denotes a coherent state with complex phase space amplitude αj=−iegjτj. For |αj|≳4 the normal- ization factor is N ≃√ 2 [53]. Note that if there are M modes with different frequencies, then the flux modula- tion should be activated Mtimes in order to prepare the state in Eq. (7). Applying a qubit pulse Rˆy,π 2followed by a strong pro- jective measurement collapses the qubit in its ground or excited state and projects the bosonic system into 1 N± |0102···0N⟩ ±eiχ|α1α2···αN⟩ , where the “+” or “−” state results from measuring the qubit in |0q⟩or |1q⟩, respectively. For the case of two bosonic modes with eg1,2, τ1,2chosen such that α1=α2=αandχ= 0 the prepared state corresponds to the maximally-entangled Bell state, |±ΨBell⟩=1 N±(|00⟩ ± |αα⟩), (8)4 where N±=p 2(1±e−|α|2)≃√ 2 for |α|≳4 [54]. Alternatively, in the case of different frequency modes, ω1̸=ω2, a maximally-entangled NOON state of the form |±ΦNOON⟩=1 N±(|0α⟩ ± |α0⟩), (9) can be obtained by performing a πpulse to flip the qubit state right after turning on the first interaction and be- fore the second one. The protocol would then require the following steps: (a) start modulating at ωac=ω1, (b) turn off the interaction after time τ1, (c) apply πqubit pulse, and (d) switch on the second flux modulation with ωac=ω2for time τ2=τ1eg1/eg2. Additionally, more general ECS of the form, |Ψ⟩ij=c00|0i0j⟩+c1α|0iαj⟩+cα0|αi0j⟩+cαα|αiαj⟩, (10) with cα0, c0α̸= 0, may also be generated using appro- priately adjusted protocols. For example, starting from |ψ⟩qij=|(0q+ 1q)0i0j⟩, then turning on the interaction with mode ifor time τisuch that |α| ≡ |egiτi|≳4, and applying a Rˆyπ 2qubit pulse, results in the state |ψ⟩qij=1 2[|0q0i0j⟩+|0qαi0j⟩+|1q0i0j⟩ − |1qαi0j⟩]. If we subsequently turn on the interaction with mode j(for time τj=α/egj) and apply another Rˆyπ 2qubit pulse, the resulting state is, |ψ⟩qij=1√ 2 |0⟩q|Ψ⟩+ ij+|1⟩q|Ψ⟩− ij , where |Ψ⟩± ij=1 2(|0i0j⟩+|0iαj⟩ ± |αi0j⟩ ∓ |αiαj⟩).(11) Finally, a strong measurement collapses the qubit in |0⟩q or|1⟩q, projecting the system in the maximally-entangled two-mode state |Ψ⟩+ ijor|Ψ⟩− ij, respectively. IV. NUMERICAL MODELING & BENCHMARKING We benchmark the protocol described above for gener- ating the Bell state |+ΨBell⟩against realistic experimen- tal conditions including dissipation using the quantum statistical Lindblad master equation [55] ˙ρ=i ℏ[ρ,ˆeH] +X jωj Qj nth jL[ˆa† j]ρ+ (nth+ 1)L[ˆaj]ρ +1 T1L[ˆc]ρ+1 T2L[ˆc†ˆc]ρ, (12) where Qjis the quality factor of each resonator, L[ˆo]ρ= (2ˆ oρˆo†−ˆo†ˆoρ−ρˆo†ˆo)/2 are superoper- ators describing each bare dissipation channel and nth j= 1/[exp(ℏωj/(kBT))−1] is the number of thermally excited magnons/phonons at temperature T.T1and T2are the qubit relaxation and dephasing times, re- spectively, for which we pick a realistic value of 50 µs throughout our simulations [6]. Of note, the in-plane (a) (b) (c) (d)FIG. 2. Bell state benchmarking for the case of two Kittel modes in two identical YIG spheres, as schematically shown in Fig. 1(a). (a) Magnon number in each magnonic mode, as a function of time during the protocol, shown for different resonator quality factors. (b) Wigner function of the individ- ual magnonic state in one mode, after tracing out the other mode, at the end of the protocol for Qm= 105. The fidelity of the prepared state to the ideal Bell state |+ΨBell⟩is shown as a function of time in (c) and as a function of the magnon number in (d). System parameters: ω1,2/(2π) = 1 GHz, eg1,2/(2π) = 2 MHz, T1=T2= 50 µs,T= 10 mK. magnetic field that is required to enable the qubit cou- pling to the magnonic or the mechanical resonator, Bz∼10−50 mT [38, 39], is not expected to limit the qubit performance [56]. In addition, while the transmon is effectively a qubit, it is more accurately described as a three-level system with negative anharmonicity given by ∼ −EC. We therefore model it as such choosing a typical value of EC/h= 300 MHz [6, 41]. We first study the case, schematically depicted in Fig. 1(a), of two YIG spheres placed diametrically op- posite with respect to the center of the SQUID. For simplicity, we assume two identical spheres and Kittel modes with the same frequency, ω1,2/(2π) = 1 GHz, as well as coupling to the qubit, eg1,2/(2π) = 2 MHz, and study the performance of the protocol proposed above as a function of the resonator quality factor, Qm, at T= 10 mK ( nth 1,2≃0.01). For typical val- ues of the Gilbert damping constant αGwe expect Qm= 1/αG∼103−105[45, 57, 58]. In Fig. 2(a) we plot the evolution of the magnon num- ber in either mode jand compare it to the ideal case, i.e., without dissipation, where ⟨ˆa† jˆaj⟩(t) =|egmt|2/2. In ad- dition, in Fig. 2(b) we plot the Wigner quasi-probability5 (a) (b) FIG. 3. (a) Logarithmic negativity and (b) conditional quan- tum entropy as a function of the magnon number for the magnon-magnon system described in Fig. 2. In the absence of dissipation (dashed-dotted curves) an ideal Bell state is created for magnon numbers ⟨ˆa† jˆaj⟩>2 with EN→1 and S(m1|m2)→ − log 2. distribution at t= 0.24µs for Qm= 105, which is de- fined as W(αj) = 2 /πTrn D†(αj)ρjD(αj)eiπˆa† jˆajo , where ρj≡Tri[ρij] is the reduced density matrix of mode j andD(αj) =eαˆa† j−α∗ˆajis the displacement operator act- ing on this mode. The two-mode density matrix, ρij, is obtained after projecting on |+q⟩, and tracing out the qubit, i.e., ρij≡Trq[ρ|+q⟩⟨+q|]. We note that since we have two identical modes, the magnon number evo- lution as well as the reduced-state Wigner functions are exactly the same for both. Furthermore, Figs. 2(c) and 2(d) show the fidelity F=p ⟨+ΨBell|ρ12|+ΨBell⟩[55, 59] of the prepared two-mode state to the ideal Bell state, as a function of time and magnon number, respectively. Evidently, for realistic values of the magnonic quality fac- torsQm≳104[40, 57, 58], the desired Bell state can be prepared with high fidelity F≲90%. To showcase the evolution of the bipartite entangle- ment during the protocol, in Fig. 3(a) we plot the loga- rithmic negativity EN= log2(2N(ρ12)+1), where N(ρ12) is the sum of negative eigenvalues of the partial transpose of the two-mode density matrix ρ12[60]. The dashed- dotted curve shows the logarithmic negativity evolution in the ideal case, EN(t) = log2h 2/(e−|egjt|2+ 1)i [61]. For |α| ≡ |egjt|≳2 it approaches the ideal value of Emax N= 1, where the two modes are maximally entangled, before magnon dissipation eventually takes over and the entan- glement gets lost. Furthermore, in Fig. 3(b) we plot the conditional quantum entropy S(m1|m2) =S(ρ12)−S(ρ2) [62, 63], where S(ρij) and S(ρj) are the Von Neumann en- tropies of the joint and reduced state, respectively, with S(ρ) =−Tr[ρlnρ]. Negative conditional quantum en- tropy serves as a sufficient criterion for the quantum state to be entangled and provides a measure of the degree of coherent quantum communication between the two en- tangled modes [62, 63]. For maximally-entangled Bell (a) (b) (c)FIG. 4. (a) Bell state fidelity, (b) logarithmic negativity and (c) conditional quantum entropy as a function of the phonon number for the case of two SQUID-embedded mechanical nanobeams interacting via the transmon. System parameters: ω1,2/(2π) = 10 MHz, eg1,2/(2π) = 100 kHz, T1=T2= 50 µs, T= 10 mK, initial nth 1,2= 0.1. states we have S(ρij) = 0 and S(ρj) = ln 2. There- fore, in the limit of large magnon numbers, we ex- pect S(m1|m2)→ − ln 2, as illustrated by the dashed- dotted curve plotting the ideal (dissipationless) case. However, as the entanglement starts decreasing due to magnon dissipation, the joint entropy of the system be- comes positive and both S(ρij) and S(ρi) start increas- ing. Therefore, as expected, the positive value threshold forS(m1|m2) is surpassed faster and at lower magnon numbers as the quality factors get smaller. Note that initially S(m1|m2)>0 due to the fact that the modes start in a thermal state with nth≃0.01. The protocol described above can also be applied to entangle mechanical beam resonators embedded in the SQUID loop, as depicted in Fig. 1(b). These can be realized using carbon nanotubes [40] or aluminum-based mechanical beams [29–31, 39] interacting via radiation- pressure couplings with the transmon. The former have operating frequencies and quality factors similar to the magnonic case studied above, therefore, the results in Figs. 2 and 3 are applicable as well. On the other hand, mechanical beam resonators made of aluminum typically operate in the range 1 −10 MHz, with quality factors Qx≳105[29–31]. Therefore, in conjunction with the magnonic case, we numerically test the same protocol for creating me- chanical Bell states between two SQUID-embedded alu- minum beam resonator modes [30], with the same fre- quency ω1,2/(2π) = 10 MHz and coupling to the qubiteg1,2/(2π) = 100 kHz. Typical temperatures of T∼10 mK, correspond to high thermal population at these frequencies, however, cooling schemes can reduce the number of thermal phonons to ≲0.1 [39, 40]. We therefore assume an attainable initial thermal population nth 1,2= 0.1 and an operating temperature of T= 10 mK. In Figs. 4(a) and 4(b) we plot the Bell-state fidelity and the logarithmic negativity, respectively, as a function of6 the phonon number during the protocol for quality fac- tors in the range Qx= 105−107. Note that initially the fidelity is less than 1, due to the finite thermal popula- tion in both resonators, however, as the protocol evolves it starts increasing before phonon dissipation takes over. We find that, for realistic quality factors Qx≳106, high phonon number Bell-states can be prepared with high fi- delity and sufficiently high entanglement as quantified by EN. However, as shown in Fig. 4(c), the effects of the initial thermal population seem to be detrimental to the conditional quantum entropy S(x1|x2) which remains far from the ideal limit during the whole protocol and only reaches negative values for Qj∼106. Experimental verification of the prepared states can be obtained by performing state tomography. For example, in the case of mechanical resonators, by sideband driving on the qubit one may engineer beam-splitter and two- mode squeezing interactions that can be used to detect correlations of the entangled state similar to Ref. [26]. This method may also be applied to the magnonic res- onators, for which independent state tomography tech- niques exist as well [64]. However, strong driving may severely impact the qubit state [65] limiting the suc- cess of such protocols. For this reason we have also analyzed an alternative scheme for reading out the en- tangled states, presented in the Appendix, which relies solely on switching on/off the interaction and performing magnon/phonon displacements and qubit measurements. V. CONCLUSION In summary, we have proposed a scheme for generat- ing ECS of magnons/phonons in a hybrid circuit QED architecture comprising a superconducting transmon qubit coupled to different magnonic/mechanical modes via bipartite flux-mediated interactions. In particu- lar, we have highlighted several schemes for creating maximally-entangled states and, as a proof-of-principle demonstration, we have numerically tested a simple protocol for generating magnonic and mechanical Bell states under realistic experimental conditions. We show that high-fidelity Bell states can be prepared in the presence of typical dissipation mechanisms in the system. Furthermore, in the Appendix we have analyzed a readout scheme, using standard circuit operations, that can be used as an alternative to existing tomography methods for verifying the prepared states. Our results pave the way towards creating controllable quantum networks of entangled magnons in a flexible and scalable platform without relying on microwave 3D cavities or strong driving. Although for simplicitywe have considered identical YIG spheres, our results are also applicable to nonidentical modes and other geometries such as micro-disk resonators [66]. Finally, as we demonstrate numerically, the proposed scheme for creating and controlling ECS is also applicable to SQUID-embedded mechanical beam resonators, opening up new opportunities for quantum information tasks in this platform and potentially giving rise to novel magnonic-mechanical hybrid devices. ACKNOWLEDGMENTS We thank Sanchar Sharma and Victor Bittencourt for helpful discussions. This research was supported by the Dutch Foundation for Scientific Research (NWO). M.K. and S.V.K. would like to acknowledge financial support by the German Federal Ministry of Education and Research (BMBF) project QECHQS (Grant No. 16KIS1590K). APPENDIX: READOUT SCHEME We now describe a method for reading out the two- mode ECS discussed in the main text, using only qubit measurements and displacement operations on the bosonic modes. We start with the assumption that the most general state one can prepare with the system Hamiltonian in Eq. (6) is of the following form, |Ψ⟩ij=c0eiθ0|0i0j⟩+c1eiθ1|0iαj⟩ +c2eiθ2|αi0j⟩+c3eiθ3|αiαj⟩, (A13) where cjare real positive numbers andP3 j=0c2 j= 1. Our assumption is based on the fact that the engineered radiation-pressure interaction in Eq. (6) can only lead to magnon/phonon displacements when the qubit is in the excited state, therefore, for the protocols described in the main text, where the interaction is activated at least once for each bosonic mode, Eq. (A13) describes the mode general state one can prepare. In addition, single-photon losses acting on coherent states result in a coherent state of smaller amplitude, therefore this decay channel does not alter the form of the state described in Eq. (A13). Assuming the state in Eq. (A13) has been prepared, we start the readout protocol by preparing the qubit in a general superposition state |ϕ⟩q= (|0⟩q+eiϕ|1⟩q)/√ 2. After switching on both interactions, the system wave- function evolves as7 U(i) intU(j) int|ϕ⟩q|Ψ⟩ij=1√ 2h |0⟩q c0eiθ0|0i0j⟩+c1eiθ1|0iαj⟩+c2eiθ2|αi0j⟩+c3eiθ3|αiαj⟩ + |1⟩qei(ϕ+¯ϕ) c0eiθ0|βiβj⟩+c1eiθ1+γj|βi(α+β)j⟩+c2eiθ2+γi|(α+β)iβj⟩+c3eiθ3+γi+γj|(α+β)i(α+β)j⟩i , (A14) where U(j) int= expn iegjˆc†ˆc(ˆajeiθ+ ˆa† je−iθ)to . The displacement amplitudes and corresponding ge- ometric phases, which arise from the radiation pres- sure interactions, are given by βi,j˙ =β(ti,j) = (gi,j/ωi,j) (eiωi,jti,j− 1) and ¯ϕ˙ =¯ϕ(ti,j) = (gi,j/ωi,j)2(ωi,jti,j−sin (ωi,jti,j)) [38, 67]. For simpli- fication purposes we have assumed that the latter are equal and, since ϕis arbitrarily determined at the qubit preparation stage, they can be absorbed into a redefini- tion of ϕ→¯ϕ+ϕ. The phases γi,j= Im( α∗βi,j) arise fromthe fact that in general two consecutive displacements do not commute. The above state can also be written as |ψ⟩qij=1 2 |+⟩qΨ+ ij+|−⟩qΨ− ij , (A15) where |±⟩= (|0⟩ ± |1⟩)/√ 2 are the eigenstates of the Pauli ˆ σxoperator and Ψ± ij=c0eiθ0|0i0j⟩+c1eiθ1|0iαj⟩+c2eiθ2|αi0j⟩+c3eiθ3|αiαj⟩ ± c0eiθ0+ϕ|βiβj⟩+c1eiθ1+ϕ+γ|βi(α+β)j⟩+c2eiθ2+ϕ+γ|(α+β)iβj⟩+c3eiθ3+ϕ+2γ|(α+β)i(α+β)j⟩ . (A16) The expectation value of the qubit in the |±⟩basis is then given by ⟨ˆσx⟩β,β=1 4 |⟨Ψ+ ij|Ψ+ ij⟩|2− |⟨Ψ− ij|Ψ− ij⟩|2 .(A17) We now consider several cases for each displacement: (I) First, assuming the coupling strength and inter- action times for both resonators are chosen such that βi,j=αi,j, we have ( γi,j= 0): Ψ± ij=c0eiθ0|0i0j⟩+c1eiθ1|0iαj⟩+c2eiθ2|αi0j⟩+(c3eiθ3±c0eiθ0+ϕ)|αiαj⟩ ±c1eiθ1+ϕ|αi(2α)j⟩ ±c2eiθ2+ϕ|(2α)iαj⟩ ±c3eiθ3+ϕ|(2α)i(2α)j⟩ (A18) From Eq. (A17) we obtain ⟨ˆσx⟩α,α=|c3eiθ3+c0eiθ0+ϕ|2− |c3eiθ3−c0eiθ0+ϕ|2 =c0c3cos (ϕ+θ0−θ3). (A19) Additionally, for βi,j=−αi,jit can be shown that ⟨ˆσx⟩−α,−α=c0c3cos (ϕ+θ3−θ0). (A20) (II) For the case βi=αi,βj=−αj, using Eq. (A16)and Eq. (A17), it follows that ⟨ˆσx⟩α,−α=c1c2cos (ϕ+θ1−θ2). (A21) Similarly for βi=−αi,βj=αjwe obtain ⟨ˆσx⟩−α,α=c1c2cos (ϕ+θ2−θ1) (A22) (III) For the cases βi=αi,βj= 0 and βi=−αi, βj= 0 we have ⟨ˆσx⟩α,0=c0c2cos (ϕ+θ0−θ2) +c1c3cos (ϕ+θ1−θ3) (A23)8 and ⟨ˆσx⟩−α,0=c0c2cos (ϕ+θ2−θ0)+c1c3cos (ϕ+θ3−θ1) (A24) respectively. (IV) For βi= 0,βj=αiandβi= 0,βj=−αiwe find two more equations, ⟨ˆσx⟩0,α=c0c1cos (ϕ+θ0−θ1) +c2c3cos (ϕ+θ2−θ3). (A25) and ⟨ˆσx⟩0,−α=c0c1cos (ϕ+θ1−θ0)+c2c3cos (ϕ+θ3−θ2) (A26) Finally for βi,j= 0 we obtain the following relation ⟨ˆσx⟩0,0= c2 0+c2 1+c2 2+c2 3 cosϕ, (A27) which is equivalent to the normalisation condition for |Ψ⟩ijwith the additional degree of freedom ϕ. The above equations are not yet in a form where they can be used to obtain all pairs of ci, θistraightforwardly. However, they can be combined and further simplified using basic trigonometric relations as shown below: (i) First, adding and subtracting equations (A19) and (A20) we obtain ⟨ˆσx⟩α,α+⟨ˆσx⟩−α,−α= 2c0c3cosϕcos (θ3−θ0),(A28) and ⟨ˆσx⟩α,α− ⟨ˆσx⟩−α,−α= 2c0c3sinϕsin (θ3−θ0).(A29) If the qubit is prepared such that ϕ=π/4 then by com- bining the above two equations we obtain a relation for c0, c3that does not depend on θ0, θ3: c0c3=q |⟨ˆσx⟩α,α|2+|⟨ˆσx⟩−α,−α|2. (A30)Ifc0c3̸= 0 we can also determine the phases. First, eiθ0 in Eq. (A13) can be absorbed into a global phase factor multiplying |Ψ⟩ijfollowed by a redefinition of θ1,2,3→ θ1,2,3/θ0(equivalent to defining θ0= 0 or 2 π). Then for ϕ=π/4 we have θ3= arctan⟨ˆσx⟩α,α− ⟨ˆσx⟩−α,−α ⟨ˆσx⟩α,α+⟨ˆσx⟩−α,−α . (A31) (ii) Following the same recipe we can obtain similar relations for c1, c2andθ1, θ2. In this case, by combining equations (A21) and (A22) for ϕ=π/4 we obtain the following equations c1c2=q |⟨ˆσx⟩α,−α|2+|⟨ˆσx⟩−α,α|2, (A32) and (assuming c1c2̸= 0) θ2−θ1= arctan⟨ˆσx⟩α,−α− ⟨ˆσx⟩−α,α ⟨ˆσx⟩α,−α+⟨ˆσx⟩−α,α . (A33) (iii) Furthermore, from equations (A23) and (A24) we obtain (for ϕ=π/4) (⟨ˆσx⟩α,0+⟨ˆσx⟩−α,0)2±(⟨ˆσx⟩α,0− ⟨ˆσx⟩−α,0)2 = 2 (c0c2)2+ (c1c3)2+ 2c0c1c2c3cos (θ2±θ1∓θ3) . (A34) Using equations (A30), (A31), (A32) and (A33) we can obtain a relation for c0, c1, c2, c3with no dependence on the phases: (c0c2)2+ (c1c3)2=f(⟨ˆσx⟩α,0,⟨ˆσx⟩−α,0,⟨ˆσx⟩α,α,⟨ˆσx⟩−α,−α,⟨ˆσx⟩α,−α,⟨ˆσx⟩−α,α) = 2⟨ˆσx⟩α,0⟨ˆσx⟩−α,0−2"q (|⟨ˆσx⟩α,α|2+|⟨ˆσx⟩−α,−α|2) (|⟨ˆσx⟩α,−α|2+|⟨ˆσx⟩−α,α|2) ×cos arctan⟨ˆσx⟩α,α− ⟨ˆσx⟩−α,−α ⟨ˆσx⟩α,α+⟨ˆσx⟩−α,−α + arctan⟨ˆσx⟩α,−α− ⟨ˆσx⟩−α,α ⟨ˆσx⟩α,−α+⟨ˆσx⟩−α,α# . (A35) (iv) Similarly, from equations (A25) and (A26) we ob- tain (for ϕ=π/4) (⟨ˆσx⟩0,α+⟨ˆσx⟩0,−α)2±(⟨ˆσx⟩0,α− ⟨ˆσx⟩0,−α)2 = 2 (c0c1)2+ (c2c3)2+ 2c0c1c2c3cos (θ1±θ2∓θ3) . (A36) Again, using equations (A30), (A31), (A32) and (A33) we can obtain another relation for c0, c1, c2, c3with no dependence on the phases:9 (c0c1)2+ (c2c3)2=g(⟨ˆσx⟩0,α,⟨ˆσx⟩0,−α,⟨ˆσx⟩α,α,⟨ˆσx⟩−α,−α,⟨ˆσx⟩α,−α,⟨ˆσx⟩−α,α) = 2⟨ˆσx⟩0,α⟨ˆσx⟩0,−α−2"q (|⟨ˆσx⟩α,α|2+|⟨ˆσx⟩−α,−α|2) (|⟨ˆσx⟩α,−α|2+|⟨ˆσx⟩−α,α|2) ×cos arctan⟨ˆσx⟩α,α− ⟨ˆσx⟩−α,−α ⟨ˆσx⟩α,α+⟨ˆσx⟩−α,−α −arctan⟨ˆσx⟩α,−α− ⟨ˆσx⟩−α,α ⟨ˆσx⟩α,−α+⟨ˆσx⟩−α,α# . (A37) In our case we are interested in reading out the Bell state |Ψ⟩ij=1√ N |0i0j⟩+eiθ|αiαj⟩ , (A38) i.e. the state in Eq. (A13) with θ3=θ,c0=c3=1√ Nand c1=c2= 0. Let us assume that we have prepared the general state in Eq. (A13). First, we can measure ⟨ˆσx⟩α,α and⟨ˆσx⟩−α,−αand from Eq. (A30) determine c0c3. If we have indeed prepared the target state shown in Eq. (A38) then this product should be nonzero. Then we proceedby measuring ⟨ˆσx⟩α,−αand⟨ˆσx⟩−α,αwhich should both be zero indicating that either c1= 0 or c2= 0 according to Eq. (A32). Additionally equations (A35) and (A37) should also equate to zero indicating c1=c2= 0. Finally, combining equations (A27) and (A30) we have (c0−c3)2=√ 2⟨ˆσx⟩0,0−2q |⟨ˆσx⟩α,α|2+|⟨ˆσx⟩−α,−α|2. (A39) If indeed the state in Eq. (A38) is prepared then we should find that ⟨ˆσx⟩0,0=p 2 (|⟨ˆσx⟩α,α|2+|⟨ˆσx⟩−α,−α|2), and therefore c0=c3= |⟨ˆσx⟩α,α|2+|⟨ˆσx⟩−α,−α|21/4= 2−1/4⟨ˆσx⟩0,0. ∗marios.kounalakis@gmail.com 1A. G. J. MacFarlane, J. P. Dowling, and G. J. Milburn, Philosophical Transactions of the Royal Society of London. Series A: Mathematical, Physical and Engineering Sciences 361, 1655 (2003). 2A. Ac´ ın, I. Bloch, H. Buhrman, T. Calarco, C. Eichler, J. Eisert, D. Esteve, N. Gisin, S. J. Glaser, F. Jelezko, S. Kuhr, M. Lewenstein, M. F. Riedel, P. O. Schmidt, R. Thew, A. Wallraff, I. Walmsley, and F. K. Wilhelm, New Journal of Physics 20, 080201 (2018). 3G. Kurizki, P. Bertet, Y. Kubo, K. Mølmer, D. Petrosyan, P. Rabl, and J. Schmiedmayer, Proceedings of the Na- tional Academy of Sciences 112, 3866 (2015). 4A. Clerk, K. Lehnert, P. Bertet, J. Petta, and Y. Naka- mura, Nature Physics 16, 257 (2020). 5M. Devoret and R. Schoelkopf, Science 339, 1169 (2013). 6M. Kjaergaard, M. E. Schwartz, J. Braum¨ uller, P. Krantz, J. I.-J. Wang, S. Gustavsson, and W. D. Oliver, Annual Review of Condensed Matter Physics 11, 369 (2020). 7M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, Rev. Mod. Phys. 86, 1391 (2014). 8T. Yu, Y.-X. Zhang, S. Sharma, X. Zhang, Y. M. Blanter, and G. E. W. Bauer, Phys. Rev. Lett. 124, 107202 (2020). 9I. Bertelli, J. J. Carmiggelt, T. Yu, B. G. Simon, C. C. Pothoven, G. E. W. Bauer, Y. M. Blanter, J. Aarts, and T. van der Sar, Science Advances 6, eabd3556 (2020). 10D. Lachance-Quirion, Y. Tabuchi, A. Gloppe, K. Usami, and Y. Nakamura, Applied Physics Express 12, 070101 (2019). 11H. Yuan, Y. Cao, A. Kamra, R. A. Duine, and P. Yan, Physics Reports 965, 1 (2022), quantum magnonics: When magnon spintronics meets quantum information science. 12B. Yurke and D. Stoler, Phys. Rev. Lett. 57, 13 (1986).13B. C. Sanders, Journal of Physics A: Mathematical and Theoretical 45, 244002 (2012). 14S. J. van Enk and O. Hirota, Phys. Rev. A 64, 022313 (2001). 15X. Wang, Phys. Rev. A 64, 022302 (2001). 16P. T. Cochrane, G. J. Milburn, and W. J. Munro, Phys. Rev. A 59, 2631 (1999). 17M. C. de Oliveira and W. J. Munro, Phys. Rev. A 61, 042309 (2000). 18H. Jeong and M. S. Kim, Phys. Rev. A 65, 042305 (2002). 19P. van Loock, W. J. Munro, K. Nemoto, T. P. Spiller, T. D. Ladd, S. L. Braunstein, and G. J. Milburn, Phys. Rev. A 78, 022303 (2008). 20A. E. Allati, Y. Hassouni, and N. Metwally, Physica Scripta 83, 065002 (2011). 21W. J. Munro, K. Nemoto, G. J. Milburn, and S. L. Braun- stein, Phys. Rev. A 66, 023819 (2002). 22J. Joo, W. J. Munro, and T. P. Spiller, Phys. Rev. Lett. 107, 083601 (2011). 23L. Di´ osi, Journal of Physics: Conference Series 306, 012006 (2011). 24D. Kafri, J. M. Taylor, and G. J. Milburn, New Journal of Physics 16, 065020 (2014). 25C. Ockeloen-Korppi, E. Damsk¨ agg, J.-M. Pirkkalainen, M. Asjad, A. Clerk, F. Massel, M. Woolley, and M. Sil- lanp¨ a¨ a, Nature 556, 478 (2018). 26S. Kotler, G. A. Peterson, E. Shojaee, F. Lecocq, K. Ci- cak, A. Kwiatkowski, S. Geller, S. Glancy, E. Knill, R. W. Simmonds, J. Aumentado, and J. D. Teufel, Science 372, 622 (2021). 27R. Riedinger, A. Wallucks, I. Marinkovi´ c, C. L¨ oschnauer, M. Aspelmeyer, S. Hong, and S. Gr¨ oblacher, Nature 556, 473 (2018).10 28E. A. Wollack, A. Y. Cleland, R. G. Gruenke, Z. Wang, P. Arrangoiz-Arriola, and A. H. Safavi-Naeini, Nature 604, 463 (2022). 29I. Rodrigues, D. Bothner, and G. Steele, Nature commu- nications 10, 5359 (2019). 30P. Schmidt, M. T Amawi, S. Pogorzalek, F. Deppe, A. Marx, R. Gross, and H. Huebl, Communications Physics 3, 1 (2020). 31T. Bera, S. Majumder, S. K. Sahu, and V. Singh, Com- munications Physics 4, 1 (2021). 32J. Simon, H. Tanji, S. Ghosh, and V. Vuleti´ c, Nature Physics 3, 765 (2007). 33Z. Zhang, M. O. Scully, and G. S. Agarwal, Phys. Rev. Res.1, 023021 (2019). 34J. Li and S.-Y. Zhu, New Journal of Physics 21, 085001 (2019). 35X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, Science Advances 2, e1501286 (2016). 36J. M. P. Nair and G. S. Agarwal, Applied Physics Letters 117, 084001 (2020). 37M. Yu, S.-Y. Zhu, and J. Li, Journal of Physics B: Atomic, Molecular and Optical Physics 53, 065402 (2020). 38M. Kounalakis, G. E. W. Bauer, and Y. M. Blanter, Phys. Rev. Lett. 129, 037205 (2022). 39M. Kounalakis, Y. M. Blanter, and G. A. Steele, Phys. Rev. Research 2, 023335 (2020). 40K. E. Khosla, M. R. Vanner, N. Ares, and E. A. Laird, Phys. Rev. X 8, 021052 (2018). 41J. Koch, T. M. Yu, J. Gambetta, A. A. Houck, D. I. Schus- ter, J. Majer, A. Blais, M. H. Devoret, S. M. Girvin, and R. J. Schoelkopf, Phys. Rev. A 76, 042319 (2007). 42U. Vool and M. Devoret, International Journal of Circuit Theory and Applications 45, 897 (2017). 43H. Yuan, J. Xie, and R. A. Duine, arXiv preprint arXiv:2301.09095 (2023). 44Y. Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Ya- mazaki, K. Usami, and Y. Nakamura, Comptes Rendus Physique 17, 729 (2016). 45B. Zare Rameshti, S. Viola Kusminskiy, J. A. Haigh, K. Us- ami, D. Lachance-Quirion, Y. Nakamura, C.-M. Hu, H. X. Tang, G. E. Bauer, and Y. M. Blanter, Physics Reports 979, 1 (2022). 46D. D. Stancil and A. Prabhakar, Spin waves , Vol. 5 (Springer, 2009). 47S. Sharma, Cavity optomagnonics: Manipulating mag- netism by light , Ph.D. thesis, Delft University of Technol- ogy (2019).48M. Kounalakis, Y. M. Blanter, and G. A. Steele, npj Quan- tum Information 5, 100 (2019). 49A. Nunnenkamp, K. Børkje, and S. M. Girvin, Phys. Rev. Lett.107, 063602 (2011). 50P. D. Nation, J. Suh, and M. P. Blencowe, Phys. Rev. A 93, 022510 (2016). 51L. Tian, Phys. Rev. B 72, 195411 (2005). 52D. Kielpinski, D. Kafri, M. J. Woolley, G. J. Milburn, and J. M. Taylor, Phys. Rev. Lett. 108, 130504 (2012). 53M. Mirrahimi, Z. Leghtas, V. V. Albert, S. Touzard, R. J. Schoelkopf, L. Jiang, and M. H. Devoret, New Journal of Physics 16, 045014 (2014). 54T. C. Ralph, A. Gilchrist, G. J. Milburn, W. J. Munro, and S. Glancy, Phys. Rev. A 68, 042319 (2003). 55J. Johansson, P. Nation, and F. Nori, Computer Physics Communications 183, 1760 (2012). 56J. Krause, C. Dickel, E. Vaal, M. Vielmetter, J. Feng, R. Bounds, G. Catelani, J. M. Fink, and Y. Ando, Phys. Rev. Applied 17, 034032 (2022). 57Y. Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Ya- mazaki, K. Usami, and Y. Nakamura, Science 349, 405 (2015). 58S. Klingler, H. Maier-Flaig, C. Dubs, O. Surzhenko, R. Gross, H. Huebl, S. T. B. Goennenwein, and M. Weiler, Applied Physics Letters 110, 092409 (2017). 59M. A. Nielsen and I. L. Chuang, Quantum computation and quantum information (Cambridge university press, 2010). 60G. Vidal and R. F. Werner, Phys. Rev. A 65, 032314 (2002). 61P. Liu, X.-M. Feng, and G.-R. Jin, Chinese Physics B 23, 030310 (2014). 62N. J. Cerf and C. Adami, Phys. Rev. Lett. 79, 5194 (1997). 63M. Horodecki, J. Oppenheim, and A. Winter, Nature 436, 673 (2005). 64T. Hioki, H. Shimizu, T. Makiuchi, and E. Saitoh, Phys. Rev. B 104, L100419 (2021). 65R. Lescanne, L. Verney, Q. Ficheux, M. H. Devoret, B. Huard, M. Mirrahimi, and Z. Leghtas, Phys. Rev. Ap- plied11, 014030 (2019). 66T. Srivastava, H. Merbouche, I. Ngouagnia Yemeli, N. Beaulieu, J. Ben Youssef, M. Mu˜ noz, P. Che, P. Bor- tolotti, V. Cros, O. Klein, S. Sangiao, J. De Teresa, S. Demokritov, V. Demidov, A. Anane, C. Serpico, M. d’Aquino, and G. de Loubens, Phys. Rev. Appl. 19, 064078 (2023). 67A. Asadian, C. Brukner, and P. Rabl, Phys. Rev. Lett. 112, 190402 (2014). | 2023-09-28 | We propose a scheme for generating and controlling entangled coherent states
(ECS) of magnons, i.e. the quanta of the collective spin excitations in
magnetic systems, or phonons in mechanical resonators. The proposed hybrid
circuit architecture comprises a superconducting transmon qubit coupled to a
pair of magnonic Yttrium Iron Garnet (YIG) spherical resonators or mechanical
beam resonators via flux-mediated interactions. Specifically, the coupling
results from the magnetic/mechanical quantum fluctuations modulating the qubit
inductor, formed by a superconducting quantum interference device (SQUID). We
show that the resulting radiation-pressure interaction of the qubit with each
mode, can be employed to generate maximally-entangled states of magnons or
phonons. In addition, we numerically demonstrate a protocol for the preparation
of magnonic and mechanical Bell states with high fidelity including realistic
dissipation mechanisms. Furthermore, we have devised a scheme for reading out
the prepared states using standard qubit control and resonator field
displacements. Our work demonstrates an alternative platform for quantum
information using ECS in hybrid magnonic and mechanical quantum networks. | Engineering Entangled Coherent States of Magnons and Phonons via a Transmon Qubit | 2309.16514v1 |
arXiv:1407.4957v2 [cond-mat.mes-hall] 22 Jul 2014Microwave-induced spin currents in ferromagnetic-insula tor|normal-metal bilayer system Milan Agrawal,1,2,a)Alexander A. Serga,1Viktor Lauer,1Evangelos Th. Papaioannou,1Burkard Hillebrands,1 and Vitaliy I. Vasyuchka1 1)Fachbereich Physik and Landesforschungszentrum OPTIMAS, Technische Universit¨ at Kaiserslautern, 67663 Kaiserslautern, Germany 2)Graduate School Materials Science in Mainz, Gottlieb-Daim ler-Strasse 47, 67663 Kaiserslautern, Germany (Dated: 1 September 2018) A microwave technique is employed to simultaneously examine the spin p umping and the spin Seebeck effect processes in a YIG |Pt bilayer system. The experimental results show that for these t wo processes, the spin current flows in opposite directions. The temporal dynamics of the longitudinal spin Seebeck effect exhibits that the effect depends on the diffusion ofbulk thermal-magnonsin t he thermal gradientin the ferromagnetic- insulator |normal-metal system. Since its discovery in 2008, the spin Seebeck effect1—a route to generate a spin current by applying a heat cur- rent to ferromagnets—has given a new dimension to the field of spin-caloritronics2. In particular, the longitudi- nal spin Seebeck effect (LSSE)3, where the spin current flows along the thermal gradient in the magnetic mate- rial, drives the field due to its technologically promising applications in energy harvesting4, and in temperature, temperature gradient, and position sensing5. Having conceptual understanding and future applica- tions in the centre of attention, a comparative study of the spin current direction and its temporal evolution for different spin-current-generation processes like spin pumping (SP) and spin Seebeck effect (SSE) is very im- portant. In previous experiments6,7, these issues have not been explicitly addressed. In this letter, we demon- strate microwaves as a simple-and-controlled tool to in- vestigate both, SP and SSE, processes simultaneously in a single experiment. Such investigations are not pos- sible with other techniques including laser heating8or direct-current heating9employed to study the SSE. For example, in Ref. 10, the direction of the spin current in the SP and SSE processes has been determined by combining the FMR technique with additional Peltier or dc/ac based heating techniques. Our results reveal that inaferromagnet |normalmetal(paramagnet)system, the spin current flows from the ferromagnet (FM) to the nor- mal metal (NM) in the case of SP process, while the flow reverses for the LSSE provided that the NM is hot- ter than the FM. The time-resolved measurements show that the spin current dynamics of the LSSE is on sub- microsecond timescale compared to nanosecond fast spin pumping process11. The experiment was realized using a bilayer of a mag- netic insulator, Yttrium Iron Garnet (YIG), and a nor- mal metal, Pt. The sample structure consists a 6.7- µm- thick YIG film of dimensions 14 mm ×3 mm, grown by a)Electronic mail: magrawal@physik.uni-kl.deYIG Microwave Generator HF-Diode Pt12 3 OscilloscopeLow-pass FilterY CirculatorAmplifier50 Ω load Voltage Amplifierzxy Temperature (°C) 16.5 19.0 21.5 FIG. 1. A schematic sketch of the experimental setup. Mi- crowaves were employed to heat a 10-nm-thick Pt strip grown on a 6.7- µm-thick YIG film placed on top of the micro- stripline. The reflected microwaves were monitored on an oscilloscope. The inverse spin Hall voltage generated in th e Pt strip was amplified and measured by the oscilloscope. The inset is an infrared thermal image of the sample obtained by continuous microwave heating of the Pt strip situated in the middle of the film. liquid phase epitaxy on a 500- µm-thickGallium Gadolin- ium Garnet (GGG) substrate, and a10-nm-thick Ptstrip (3 mm×100µm), structured by photolithography and deposited by molecular beam epitaxy at a growth rate of 0.05˚A/s under a pressure of 5 ×10−11mbar. A 0.6-mm- wide and 17- µm-thick copper (Cu) microstrip antenna wasdesignedona dieletricsubstratetoapplymicrowaves to the sample. In order to obtain a maximum microwave heating efficiency by eddy currents in the metal (Pt), the Pt-covered surface of the sample was placed on top of the micro-stripline. An insulating layer was inserted in between the sample and the micro-stripline to avoid any direct electric contact. Gold wires were glued to the Pt strip with silver paste to connect with the external cir- cuit. A schematic diagram of the experimental setup is shown in Fig. 1. Microwaves from an Anritsu MG3692 C generator were amplified (+30 dB) by an amplifier and guided to the sample structure. The microwaves are2 absorbed by the thin Pt metal strip and start to heat it up. As a result, a thermal gradient along the + z- direction was established. The reflected microwavesfrom the micro-stripline were received by connecting a Y- circulator to the microwave circuit-line. The reflected microwaves were rectified using a high-frequency (HF) diode and monitored on an oscilloscope. This signal pro- vides the information about the shape of the microwave signal envelope and its duration. The YIG film was mag- netized in plane by an applied magnetic field µ0Halong thex-axis. The perpendicular thermal gradient in the YIG |Pt bi- layer generates a spin current in the system due to the LSSE. The generated spin current flows along the ther- mal gradient ( z-axis). In the normal metal Pt, the spin current converts into a charge current by the inverse spin Hall effect (ISHE)12asJISHE∝Js×σ,whereJsis the spin current, and σthe spin polarization. The gener- ated charge current along the y-axis in the Pt strip was passed through a low-pass filter to block alternating cur- rents generated directly by the electric field components of the microwaves. The filtered signal was amplified and observed on the oscilloscope. In the first experiment, continuous microwave mea- surements at a fixed frequency of 6 .8 GHz were carried out by varying the magnetic field. In Fig. 2, the in- verse spin Hall voltage VISHEis plotted versus the ap- plied magnetic field µ0H. Clearly, three features can be noticed here: (i) two peaks with opposite polarities at the magnetic fields of +168 .6 mT and −168.6 mT, (ii) their unequal magnitude, and (iii) an offset for all non- resonancemagnetic fields, which hasan opposite polarity to the peaks. The first feature originates from the spin pumping processby spin wavesexcited close to ferromag- netic resonance (FMR) in the YIG film13–15. The FMR conditions were achieved for both positive and negative magnetic fields. Corresponding to these fields, a spin current is injected into Pt by the spin pumping process. The inverse spin Hall voltage generated in Pt is given by VISHE∝θSHE(Js×σ), where θSHEdenotes the spin Hall angle. The direction of σdepends on the direction of the magnetic field. Therefore, on inverting the magnetic field direction, the polarity of VISHEreverses. In order to understand the second feature of the spec- trum that the signals have unequal amplitude, it is im- portant to discuss the spin-wave modes excited in the YIG film for our experimental geometry. It is clear from the sample orientation, shown in Fig. 1 and inset to Fig. 2, that the spin waves excited by the Oersted field of the micro-stripline propagate along the y-axis, perpen- dicular to the magnetic field applied along the x-axis. These kinds of spin waves with wave vector k⊥Hare known as magnetostatic surface spin waves (MSSW) or Damon-Eshbach (DE) spin-waves16. The DE spin-waves are nonreciprocal spin waves and travel along a direction given by k=H×n, wherenis the normal to the film surface. Therefore, the propagation of these spin waves on the surface of a film can be reversed by inverting the-150 -100 -50 0 50 100 150-6-4-20246 VISHV(µV) Magnetic field (mT) µ H0Spin pumping by DE spin-waves Longitudinal SSE signal H kn PtMW stripline σE JSPt-H-k nMW stripline -σ -EJS zxy FIG. 2. The inverse spin Hall voltage ( VISHE) generated in the Pt strip as a function of the applied magnetic field µ0H. An asymmetry in the amplitude of VISHEat FMR-magnetic field appears due to unequal efficiency of Damon-Eshbach spin- waves excitation (shaded area) for two opposite directions of the applied magnet field, shown in the insets. direction of the magnetic field. In our experiment, when a magnetic field is applied along the + x-direction, The DE-spin waves, in the YIG film surface close to the micro-stripline17, can only be ex- cited along the −y-direction ( ˆk=x×z) with respect to the micro-stripline as shown in the inset to Fig 2. On the other hand, when the magnetic field is applied along the −x-direction, the spin waves can propagate only along the +y-axis. If the YIG film is not positioned symmet- rically around the micro-stripline, as in our case, the ef- fective YIG film area, where spin waves can be excited, will be unequal for two opposite fields as shown in the inset to Fig 2. Since the strength of VISHEsignal is pro- portional to the spin-wave intensity in the system7, we observed an unequal amplitude of VISHEin our experi- ment. We performed alike measurements with displacing the YIG film and find that the amplitude of the spin pumping signals can be altered by varying the relative positions of the film with respect to the micro-stripline. Therefore, we conclude that the unidirectional nature of the DE spin-waves regulates the asymmetry of the ISHE signal18. The third feature seen in Fig. 2, i.e., an offset for non- resonant magnetic fields; is attributed to the LSSE. A similar signal could also be produced by the anomalous Nernst effect in Pt, magnetized due to the proximity ef- fect. However, recent observations19,20discard any such possibility in YIG |Pt systems. The polarity of the LSSE signal changes with the direction of the magnetic field; however, it is important to notice that the LSSE signal has an opposite polarity than that of the signal at FMR for a same direction of the magnetic field. This evidence excludes the possibility of non-resonant spin pumping in the system. When the Pt strip is heated by microwave absorption, a thermal gradient ( ∇Tz) from YIG to Pt develops normal to the interface. The thermal gradient generates a spin current flowing along the z-axis. Since the Pt strip is hot, the spin currentgeneratedvia the lon- gitudinal SSE ( Js∝ −∇T) flows from Pt to YIG21,22, in3 (a) VLSSE(µV) 0 0.4 0.8 1.204812 Microwave power (W) μ0H(mT)(b) VLSSE(µV) -20 -10 0 10 20-40418.6 mW 117.5 mW 295.1 mW 468 mW 741 mW 933 mW 1175 mW8 FIG. 3. (a) Plotted is VLSSEas a function of the applied magnetic field for various applied microwave powers. (b) The peak-to-peak amplitude of VLSSEis plotted versus the ap- plied microwave power. The peak-to-peak amplitude of VLSSE scales linearly with the microwave power. contrast to spin pumping where the spin current flows from YIG to Pt23. This argument explains the oppo- site polarities of the resonant (spin pumping) and the non-resonant (longitudinal SSE) inverse-spin-Hall volt- ages (VISHE) observed in Fig. 2. These results are con- sistent with previous experimental studies6,7,10. As dis- cussed above, the non-resonant VISHEis attributed to the longitudinalSSE;henceforth,wedenotethenon-resonant VISHEvalues as VLSSE. Magnetic field scans for various microwave input pow- ers were carried out. In Fig. 3, the VLSSEversusthe mag- netic field data is plotted for various applied microwave powers. With increasing microwave power, the tempera- ture increases in the Pt strip which enlarges the thermal gradient close to the YIG |Pt interface and, hence, in- jects a larger spin current ( Js∝ −∇T) into the YIG film3,21,22. Impact of the large spin current appears as a higherVLSSEsignal highlighted in Fig. 3(b). The peak- to-peak amplitude of VLSSEscales linearly with the ap- plied microwave power. The signature that the VLSSE signal scales linearly with the microwave power verifies that the signal originates from the heating produced in Pt shown in the inset to Fig. 1. The above experiment demonstrates that microwaves can be utilized to create a thermal gradient in ferromagnetic-insulator |normal-metal system, thereby, to study the LSSE along with the SP. The experimental setup shown in Fig. 1 can also be employed to investi- gate the temporal dynamics of the longitudinal SSE and to compare it with the SP dynamics11. In the second ex- periment, instead of continuous microwaves, 10- µs-long microwave pulses with rise-fall times of less than 10 ns were used to perform the time-resolved measurements of the longitudinal SSE. The frequency of the microwave pulses (6 .8 GHz) was chosen such that the magnetic sys- tem stayed at non-resonance condition of the magnetic field in the range of interest ( ±25 mT). The experiment was executed at various microwave powers. The mea- surementswere recordedfor both positive (+25mT) and negative (-25 mT) magnetic fields, and an average valueof the non-resonant VISHE, i.e.,VLSSEwas considered. In Fig. 4, VLSSEis plotted versus time. The longitu- dinal SSE signal takes around 1 µs to reach to the sat- uration level. The 10% −90% rise time of the signal is found to be around 530 ns. The longitudinal SSE sig- nal (VLSSE) shows similar features as reported in Ref. 8, where a pulsed laser is employed to create the vertical thermal gradient in the YIG |Pt system. The main dif- ference observed here is that the VLSSEsignal appears as soon as the microwave current runs; contrarily, in the laser heating experiment8a time lag (200 ns) exists due to the laser switching time. The model of thermal magnon diffusion8,22,24is em- ployed here to understand the timescale of the longitudi- nal SSE. The model states that the spin current from a FM injected into a NM depends on the diffusion of ther- malmagnonsinthe FM. Thedensityofthermalmagnons is proportional to the local phonon temperature25,26. Due to the thermal gradient in the FM, magnons diffuse from hotter regions (higher population) to colder regions (lower population) of the FM and create a magnon den- sity inequilibrium at the FM |NM interface which leads to the injection of a spin current into the NM21. The timescale of the effect depends on the temporal devel- opment of the magnon density inequilibrium, i.e., the thermal gradient in the system. According to the model, VLSSEis given by8 VLSSE(t)∝l/integraldisplay interface∇Tz(z,t)exp(−|z| L)dz,(1) where∇Tzis the phonon thermal gradient in the FM, perpendicular to the interface, lis the magnetic film thickness, and Lis the effective magnon diffusion length. Wefitted ourexperimentaldatashownin Fig4(a)with Eq. 1 using ∇Tz(z,t) calculated numerically by solving the heat equation for our system in accordance with a model described in Ref. 8. In Fig. 4(b), the normalized experimental VLSSE-signal was plotted together with the FIG. 4. (a) Plotted is the temporal evolution of VLSSEon the application of a 10 µs long microwave pulse which creates a vertical temperature gradient in the YIG |Pt structure by heatingthePtstrip. (b)Acomparison ofexperimentallymea - suredVLSSEdata with calculated values using Eq. (1) for vari- ous effective magnon diffusion lengths L= 300,500,700 nm.4 calculated ones for various magnon diffusion lengths of 300 nm, 500 nm, and 700 nm. The model resembles the experimental data well. The fitting shows that a typi- cal magnon diffusion length for thermal magnons in the YIG|Pt system is around 500 nm. An identical value for the magnon diffusion length was obtained in the laser heating experimental performed on the same sample, re- ported in Ref. 8. In summary, we presented microwavesas a perspective heating technique to generate a thermal gradient in fer- romagnetic insulator |normal metal systems to study the static and temporal dynamics of the longitudinal spin Seebeck effect. The static measurements provide cru- cial information about the direction of the spin current flow in the spin pumping and longitudinal SSE processes. The experiment demonstrates that in the longitudinal SSE a spin current flows from the normal metal (hot) towards the ferromagnet (cold) while in the spin pump- ing case, the flow is opposite. The temporal dynamics of the longitudinal SSE experiment manifests the sub- microsecond timescale of the effect which is slower than the spin pumpingprocess. Thethermal magnondiffusion model can explain the outcomes of the experiment and leadstoconcludethatthetimescaleoftheeffect reliesthe evolution of the vertical thermal gradient in the vicinity of the ferromagnet |normal metal interface. From our ex- periment, a typical magnon diffusion length of 500 nm is estimated for the YIG |Pt system. The authors thank A. V. Chumak, M. B. Jungfleisch, and P. Pirro for valuable discussions. M.A. was sup- ported by a fellowship of the Graduate School Material Sciences in Mainz (MAINZ) through DFG funding of the Excellence Initiative (GSC-266). We acknowledge financial support by Deutsche Forschungsgemeinschaft (SE 1771/4) within Priority Program 1538 “Spin Caloric Transport”, and technical support from the Nano Struc- turing Center, TU Kaiserslautern. 1K. Uchida, S. Takahashi, K. Harii, J. Ieda, W. Koshibae, K. Ando, S. Maekawa, and E. Saitoh, Nature 455, 778 (2008). 2G. E. W. Bauer, E. Saitoh, and B. J. van Wees, Nature Materials 11, 391 (2012). 3K. Uchida, H. Adachi, T. Ota, H. Nakayama, S. Maekawa, and E. Saitoh, Applied Physics Letters 97, 172505 (2010). 4A. Kirihara, K. Uchida, Y. Kajiwara, M. Ishida, Y. Nakamura, T. Manako, E. Saitoh, and S. Yorozu, Nature Materials 11, 686 (2012).5K. Uchida, A. Kirihara, M. Ishida, R. Takahashi, and E. Saito h, Japanese Journal of Applied Physics 50, 120211 (2011). 6M. B. Jungfleisch, T. An, K. Ando, Y. Kajiwara, K. Uchida, V. I. Vasyuchka, A. V. Chumak, A. A. Serga, E. Saitoh, and B. Hillebrands, Applied Physics Letters 102, 062417 (2013). 7C. Sandweg, Y. Kajiwara, A. V. Chumak, A. A. Serga, V. Vasyuchka, M. Jungfleisch, E. Saitoh, and B. Hillebrands, Physical Review Letters 106, 216601 (2011). 8M. Agrawal, V. I. Vasyuchka, A. A. Serga, A. Kirihara, P. Pirr o, T. Langner, M. B. Jungfleisch, A. V. Chumak, E. T. Papaioan- nou, and B. Hillebrands, Physical Review B 89, 224414 (2014). 9M. Schreier, N. Roschewsky, E. Dobler, S. Meyer, H. Huebl, R. Gross, and S. T. B. Goennenwein, Applied Physics Letters 103, 242404 (2013). 10M. Schreier, G. E. W. Bauer, V. Vasyuchka, J. Flipse, K.-i. Uchida, J. Lotze, V. Lauer, A. Chumak, A. Serga, S. Daimon, T. Kikkawa, E. Saitoh, B. J. van Wees, B. Hillebrands, R. Gros s, and S. T. B. Goennenwein, arXiv:1404.3490. 11M. B. Jungfleisch, A. V. Chumak, V. I. Vasyuchka, A. A. Serga, B.Obry, H.Schultheiss, P.A.Beck, A.D.Karenowska, E.Sait oh, and B. Hillebrands, Applied Physics Letters 99, 182512 (2011). 12E. Saitoh, M. Ueda, H. Miyajima, and G. Tatara, Applied Physics Letters 88, 182509 (2006). 13Y. Kajiwara, K. Harii, S. Takahashi, J. Ohe, K. Uchida, M. Mizuguchi, H. Umezawa, H. Kawai, K. Ando, K. Takanashi, S. Maekawa, and E. Saitoh, Nature 464, 262 (2010). 14M. Costache, M. Sladkov, S. Watts, C. van der Wal, and B. van Wees, Physical Review Letters 97, 1 (2006). 15K. Ando, J. Ieda, K. Sasage, S. Takahashi, S. Maekawa, and E. Saitoh, Applied Physics Letters 94, 262505 (2009). 16R. Damon and J. Eshbach, Journal of Physics and Chemistry of Solids19, 308 (1961). 17The efficiency of spin wave excitation at the back surface of th e YIG film is very poor. 18R. Iguchi, K. Ando, Z. Qiu, T. An, E. Saitoh, and T. Sato, Applied Physics Letters 102, 022406 (2013). 19S. Gepraegs, S. Meyer, S. Altmannshofer, M. Opel, F. Wilhelm , A. Rogalev, R. Gross, and S. T. B. Goennenwein, Applied Physics Letters 101, 262407 (2012). 20T. Kikkawa, K. Uchida, Y. Shiomi, Z. Qiu, D. Hou, D. Tian, H. Nakayama, X.-F. Jin, and E. Saitoh, Physical Review Lette rs 110, 067207 (2013). 21J. Xiao, G. E. W. Bauer, K. Uchida, E. Saitoh, and S. Maekawa, Physical Review B 81, 214418 (2010). 22S. M. Rezende, R. L. Rodr´ ıguez-Su´ arez, R. O. Cunha, A. R. Rodrigues, F. L. A. Machado, G. A. Fonseca Guerra, J. C. Lopez Ortiz, and A. Azevedo, Physical Review B 89, 014416 (2014). 23Y. Tserkovnyak, A. Brataas, and G. E. W. Bauer, Physical Review B 66, 1 (2002). 24U. Ritzmann, D. Hinzke, and U. Nowak, Physical Review B 89, 024409 (2014). 25M. Agrawal, V. I. Vasyuchka, A. A. Serga, A. D. Karenowska, G. A. Melkov, and B. Hillebrands, Physical Review Letters 111, 107204 (2013). 26M. Schreier, A. Kamra, M. Weiler, J. Xiao, G. E. W. Bauer, R. Gross, and S. T. B. Goennenwein, Physical Review B 88, 094410 (2013). | 2014-07-18 | A microwave technique is employed to simultaneously examine the spin pumping
and the spin Seebeck effect processes in a YIG|Pt bilayer system. The
experimental results show that for these two processes, the spin current flows
in opposite directions. The temporal dynamics of the longitudinal spin Seebeck
effect exhibits that the effect depends on the diffusion of bulk
thermal-magnons in the thermal gradient in the
ferromagnetic-insulator|normal-metal system. | Microwave-induced spin currents in ferromagnetic-insulator|normal-metal bilayer system | 1407.4957v2 |
arXiv:2301.05820v1 [quant-ph] 14 Jan 2023Quantum entanglement generation on magnons assisted with m icrowave cavities coupled to a superconducting qubit Jiu-Ming Li and Shao-Ming Fei∗ School of Mathematical Sciences, Capital Normal Universit y, Beijing 100048, China We present protocols to generate quantum entanglement on no nlocal magnons in hybrid systems composed of yttrium iron garnet (YIG) spheres, microwave ca vities and a superconducting (SC) qubit. In the schemes, the YIGs are coupled to respective mic rowave cavities in resonant way, and the SC qubit is placed at the center of the cavities, which int eracts with the cavities simultaneously. By exchanging the virtual photon, the cavities can indirect ly interact in the far-detuning regime. Detailed protocols are presented to establish entanglemen t for two, three and arbitrary Nmagnons with reasonable fidelities. Keywords: magnon, superconducting qubit, quantum electro dynamics, quantum entanglement, indirect in- teraction I. INTRODUCTION Quantum entanglement is one of the most important features in quantum mechanics. The quantum entan- gled states [1–4] are significant ingredients in quantum information processing. Over past decades, various the- oretical and experimental proposals have been presented for processing quantum information by using varioussys- tems such as atoms [5–14], spins [15–21], ions [22–29], photons [5, 30–39], phonons [40–42], and so on. With the development of technologies, the quantum entanglement hasbeenestablishednotonlyin microscopicsystems, but also in the macroscopic systems such as superconducting circuits [43–48] and magnons system [49–54]. Hybrid systems exploit the advantages of different quantum systems in achieving certain quantum tasks, such as creating quantum entanglement and carryingout quantum logic gates. Many works have been presented so far for quantum information processing in the hybrid systems [55–58]. For instance, as an important quan- tum technology [59], the hybrid quantum circuits com- bine superconducting systems with other physical sys- tems which can be fabricated on a chip. The supercon- ducting (SC) qubit circuits [60, 61], based on the Joseph- son junctions, can exhibit quantum behaviors even at macroscopic scale. Generally, the interaction between the SC qubits and the environment, e.g., systems in strong or even ultrastrong coupling regime via quantized electromagnetic fields, would result in short coherence time. Thus many researches on circuit quantum electro- dynamics (QED) [62] have been presented with respect to the SC qubits, superconducting coplanar waveguide resonators, LCresonators and so on. This circuit QED focuses on studies of the light-matter interaction by us- ing the microwave photons, and has become a relative independent research field originated from cavity QED. The hybrid systems composed of collective spins (magnons) in ferrimagnetic systems and other systems ∗Electronic address: feishm@cnu.edu.cnare able to constitute the magnon-photon [63, 64], magnon-phonon[65–67], magnon-photon-phonon[49, 50, 68] systems and so on, giving rise to new interesting ap- plications. Ferrimagnetic systems such as yttrium iron garnet (YIG) sphere have attracted considerable atten- tion in recent years, which provide new platforms for in- vestigating the macroscopic quantum phenomena partic- ularly. Such systems are able to achieve strong and even ultrastrong couplings [69] between the magnons and the microwave photons, as a result of the high density of the collective spins in YIG and the lower dissipation. The YIG has the unique dielectric microwave properties with very lower microwave magnetic loss parameter. Mean- while, some important works have been presented on magnon Kerr effect [70, 71], quantum transduction [72], magnon squeezing [73, 74], magnon Fock state [75] and entanglement of magnons. For example, In 2018 Li et al. [49] proposed a system consisted of magnons, microwave photonsandphononsforestablishingtripartiteentangled states based on the magnetostrictiveinteraction and that the entangled state in magnon-photon-phonon system is robust. In 2019 Li et al.[50] constructed the entangled state of two magnon modes in a cavity magnomechani- cal system by applying a strong red-detuned microwave field on a magnon mode to activate the nonlinear mag- netostrictive interaction. In 2021 Kong et al.[52] used the indirect coherent interaction for accomplishing two magnonsentanglementand squeezingvia virtualphotons in the ferromagnetic-superconducting system. Inthiswork,wefirstpresentahybridsystemcomposed of two YIG spheres, two identical microwave cavities and a SC qubit to establish quantum entanglement on two nonlocal magnons. In this system, two YIGs are coupled to respective microwave cavities that cross each other. And a SC qubit is placed at the center of the crossing of twoidenticalcavities,namely,theSCqubitinteractswith the two cavities simultaneously. The magnons in YIGs can be coupled to the microwave cavities in the resonant way, owing to that the frequencies of two magnons can be tuned by biased magnetic fields, respectively. Com- pared with other works, the SC qubit is coupled to the two microwavecavities in the far-detuning regime, mean-2 FIG. 1: (Color online) Schematic of the hybrid system com- posed oftwoyttriumiron garnet spheres coupledtorespecti ve microwave cavities. Two cavities cross each other, and a su- perconducting qubit (black spot) is placed at the center of t he crossing. ing that the two identical cavities indirectly interact with each other by exchanging virtual photons. Then, we give the effective Hamiltonian of the subsystem composed of the SC qubit and two cavities, and present the protocol of entanglement establishment. In Sec. III, we consider the caseofthreemagnons. Inthe hybridsystemshownin Fig.3, the three identical microwave cavities could indi- rectly interact via the virtual photons, and each magnon is resonant with the respective cavity by tuning the fre- quency of the magnon. At last, we get the isoprobability entanglement on three nonlocal magnons. Moreover, the hybrid system composed of Nmagnons, Nidentical mi- crowave cavities and a SC qubit is derived in Sec. IV. We summarize in Sec. V. II. QUANTUM ENTANGLEMENT ON TWO NONLOCAL MAGNONS A. Hamiltonian of the hybrid system We consider a hybrid system, see Fig.1, in which two microwavecavitiescrosseachother,twoyttriumirongar- net (YIG) spheres are coupled to the microwave cavities, respectively. A superconducting (SC) qubit, represented by black spot in the Fig.1, is placed at the center of the crossing in order to interact with the two microwave cav- ities simultaneously. The YIG spheres are placed at the antinode of two microwave magnetic fields, respectively, and a static magnetic field is locally biased in each YIG sphere. In our model, the SC qubit is a two-level system with ground state |g/an}bracketri}htqand excited state |e/an}bracketri}htq. The magnetostatic modes in YIG can be excited when the magnetic component of the microwave cavity field is perpendicular to the biased magnetic field. We only con- sider the Kittel mode [76] in the hybrid system, namely,the magnon modes can be excited in YIG. The fre- quency of the magnon is in the gigahertz range. Thus the magnon generally interacts with the microwave pho- ton via the magnetic dipole interaction. The frequency of the magnon is given by ωm=γH, whereHis the biased magnetic field and γ/2π= 28 GHz/T is the gyro- magnetic ratio. In recent years, some experiments have already real- izedthestrongandultrastrongmagnon-magnoncoupling [77–79] as well as the magnon-qubit interaction [80, 81], which means that in the hybrid system shown in Fig.1 the magnon is both coupled to the SC qubit and an- other magnon. However, we mainly consider that the magnons which frequencies are tuned by the locally bi- ased static magnetic fields can be resonant with the cav- ities. In the meantime, the two cavities modes interact indirectly in the far-detuning regime for exchanging pho- tons. The entanglement of two nonlocal magnons can be constructed by using two cavities and the SC qubit. Given that there are magnon-magnon and magnon-qubit interactions, the magnon can be detuned with the qubit and another magnon in order to neglect their interac- tions. In the rotating wave approximation the Hamilto- nian of the hybrid system is ( /planckover2pi1= 1 hereafter) [82] H(S)=H0+Hint H0=ωm1m† 1m1+ωm2m† 2m2+1 2ωqσz +ωa1a† 1a1+ωa2a† 2a2 Hint=λm1(a1m† 1+a† 1m1)+λm2(a2m† 2+a† 2m2) +λq1(a1σ++a† 1σ)+λq2(a2σ++a† 2σ).(1) Here,H0is the free Hamiltonian of the two cavities, two magnonsandtheSCqubit. Hintisthe interactionHamil- tonian among the cavities, magnons and SC qubit. ωm1 andωm2are the frequencies of the two magnons, which are tunable under biased magnetic fields, respectively. ωa1andωa2are the frequencies of two cavities, and ωq is the state transition frequency between |g/an}bracketri}htq↔ |e/an}bracketri}htqof the SC qubit. In the Kittel mode, the collective spins in YIGs can be expressed by the boson operators. m1(m2) andm† 1(m† 2) are the annihilation and creation opera- tors of magnon mode 1 (2). a1(a2) anda† 1(a† 2) denote the annihilation and creation operators of cavity mode 1 (2), respectively. They satisfy commutation relations [O,O†] = 1 forO=a1,a2,m1,m2.σz=|e/an}bracketri}htq/an}bracketle{te|−|g/an}bracketri}htq/an}bracketle{tg|. σ=|g/an}bracketri}htq/an}bracketle{te|andσ+=|e/an}bracketri}htq/an}bracketle{tg|are the lowing and rais- ing operators of the SC qubit. λq1(λq2) is the coupling strengthbetweenthe SCqubit andthe cavitymode1(2). λm1(λm2) is the coupling between the magnon mode 1 (2) and the cavity mode 1 (2). As mentioned above, the two microwave cavities are identical ones with the same frequency ωa1=ωa2=ωa. Meanwhile, one can assume that λq1=λq2=λq. In the3 interaction picture with respect to e−iH0t, the Hamilto- nian is expressed as H(I)=λm1a1m† 1eiδ1t+λm2a2m† 2eiδ2t+λqa1σ+ei∆1t +λqa2σ+ei∆2t+H.c., (2) whereδ1=ωm1−ωa,δ2=ωm2−ωa, ∆1=ωq−ωa and ∆ 2=ωq−ωa. The SC qubit is coupled to the two cavities simultaneously. Owing to ∆ 1= ∆2= ∆0/ne}ationslash= 0 and ∆ 0≫λq, the two identical microwave cavities indi- rectlyinteractwitheachotherinthefar-detuningregime. Therefore, the effective Hamiltonian of the subsystem composedofthe twomicrowavecavitiesand the SC qubit in the far-detuning regime is given by [83] Heff=/tildewideλq/bracketleftBig σz(a† 1a1+a† 2a2+a† 1a2+a1a† 2)+2|e/an}bracketri}htq/an}bracketle{te|/bracketrightBig ,(3) where/tildewideλq=λ2 q/∆0. B. Entangled state generation on two nonlocal magnons We now give the protocol of quantum entanglement generation on two nonlocal magnons. Generally, the magnon can be excited by a drive magnetic field. For convenience the state of magnon 1 is prepared as |1/an}bracketri}htm1 via the magnetic field. The initial state of the hybrid system is |ϕ/an}bracketri}ht0=|1/an}bracketri}htm1|0/an}bracketri}htm2|0/an}bracketri}hta1|0/an}bracketri}hta2|g/an}bracketri}htq, in which the two cavities are all in the vacuum state, magnon 2 is in the state|0/an}bracketri}htm2, and the SC qubit is in state |g/an}bracketri}htqwhich is unal- tered all the time due to the indirect interaction between the two cavities. step 1: The frequency of magnon 1 is tuned to be ωm1=ωa1so that the cavity 1 could be resonated with it. Therefore, the magnon 1 and cavity 1 are in a super- posed state after time T1=π/4λm1. The local evolution is|1/an}bracketri}htm1|0/an}bracketri}hta1→1√ 2(|1/an}bracketri}htm1|0/an}bracketri}hta1−i|0/an}bracketri}htm1|1/an}bracketri}hta1), which means that the states of SC qubit, magnon 2 and cavity 2 are unchanged due to decoupling between the SC and two cavities, and the magnon 2 is far-detuned with cavity 2. The state evolves to |ϕ/an}bracketri}ht1=1√ 2(|1/an}bracketri}htm1|0/an}bracketri}hta1−i|0/an}bracketri}htm1|1/an}bracketri}hta1) ⊗|0/an}bracketri}htm2⊗|0/an}bracketri}hta2⊗|g/an}bracketri}htq. (4) step 2: The magnons are tuned to far detune with respective cavities. From Eq. (3), the evolution of sub- system composedof twomicrowavecavities and SC qubit is given by |χ(t)/an}bracketri}htsub=ei/tildewideλqt/bracketleftbig cos(/tildewideλqt)|1/an}bracketri}hta1|0/an}bracketri}hta2+isin(/tildewideλqt)|0/an}bracketri}hta1|1/an}bracketri}hta2/bracketrightbig ⊗|g/an}bracketri}htq (5) under the condition ∆ 0≫λq. After timeT2=π/2/tildewideλq, the evolution between two cav- ities is|1/an}bracketri}hta1|0/an}bracketri}hta2→ −|0/an}bracketri}hta1|1/an}bracketri}hta2, which indicates that thephoton can be indirectly transmitted between the two cavities, with the state of SC qubit unchanged. There- fore, the state after this step changes to |ϕ/an}bracketri}ht2=1√ 2(|1/an}bracketri}htm1|0/an}bracketri}hta1|0/an}bracketri}hta2+i|0/an}bracketri}htm1|0/an}bracketri}hta1|1/an}bracketri}hta2) ⊗|0/an}bracketri}htm2⊗|g/an}bracketri}htq. (6) step 3: The frequency of magnon 2 is tuned with ωm2=ωa2to resonate with the cavity 2. In the mean- time the cavities are decoupled to the SC qubit and the magnon 1 is far detuned with the cavity 1. After time T3=π/2λm2, the local evolution |0/an}bracketri}htm2|1/an}bracketri}hta2→ −i|1/an}bracketri}htm2|0/an}bracketri}hta2 is attained. The final state is |ϕ/an}bracketri}ht3=1√ 2(|1/an}bracketri}htm1|0/an}bracketri}htm2+|0/an}bracketri}htm1|1/an}bracketri}htm2) ⊗|0/an}bracketri}hta1⊗|0/an}bracketri}hta2⊗|g/an}bracketri}htq, (7) which is just the single-excitation Bell state on two non- local magnons. In the whole process, we mainly consider the interac- tions between the magnonsand the cavities, and between the cavities and the SC qubit. However, the SC qubit can be coupled to the magnons. In terms of Ref.[80], the interactions between the magnons and the SC qubit are described as Hqm,1=λqm,1(σ+m1+H.c.) and Hqm,2=λqm,2(σ+m2+H.c.) whereλqm,1=λqλm1/∆0 andλqm,2=λqλm2/∆0, while the conditions ωq=ωm1 andωq=ωm2are attained. In the meantime, the two magnons are interacts each other by using the SC qubit. Generally, the frequencies of two magnon modes are tuned by the locally biased magnetic fields. There- fore, the magnon can be detuned with the SC qubit and another magnon in order to neglect the interactions be- tween the magnons and the SC qubit. C. Numerical result We here simulate [84] the fidelity of the Bell state on two nonlocal magnons by considering the dissipations of all constituents of the hybrid system. The realistic evo- lution of the hybrid system composed of magnons, mi- crowave cavities and SC qubit is governed by the master equation ˙ρ=−i[H(I),ρ]+κm1D[m1]ρ+κm2D[m2]ρ +κa1D[a1]ρ+κa2D[a2]ρ+γqD[σ]ρ.(8) Here,ρis the density operator of the hybrid system, κm1 andκm2are the dissipation rates of magnon 1 and 2, κa1andκa2denote the dissipation rates for the two mi- crowave cavities 1 and 2, γqis the dissipation rate of the SC qubit, D[X]ρ=(2XρX†−X†Xρ−ρX†X)/2 for X=m1,m2,a1,a2,σ. The fidelity of the entangled state of two nonlocal magnons is defined by F=3/an}bracketle{tϕ|ρ|ϕ/an}bracketri}ht3. The related parameters are chosen as ωq/2π= 7.92 GHz,ωa/2π= 6.98 GHz,λq/2π= 83.2 MHz,λm1/2π=4 74 78 82 86 90 9488909294 λq/2π [MHz] Fidelity (%)(a) 0.6 0.8 1.0 1.2909294 (κa)−1 [µs] Fidelity (%)(b) 0.8 1.0 1.2 1.4 1.6909294 (κm)−1 [µs]Fidelity (%)(c) 0.6 0.8 1.0 1.29091929394 (γq)−1 [µs] Fidelity (%)(d) FIG. 2: (a) The fidelity of the Bell state of two nonlocal magno ns with respect to the coupling strength λq. Since/tildewideλq=λ2 q/∆0 in Eq.(3), the fidelity is similar to parabola. (b)-(d) The fid elity of the Bell state versus the dissipations of cavities, magnons, and SC qubit, respectively. 15.3 MHz,λm2/2π= 15.3 MHz [81], κm1/2π=κm2/2π= κm/2π= 1.06 MHz,κa1/2π=κa2/2π=κa/2π= 1.35 MHz [69], γq/2π= 1.2 MHz [80]. The fidelity of the entanglement between two nonlocal magnons can reach 92.9%. The influences of the imperfect relationship among pa- rameters is discussed next. The Fig.2(a) shows the fi- delity influenced by the coupling strength between the microwave cavities and the SC qubit. Since /tildewideλq=λ2 q/∆0 in Eq.(3), the fidelity is similar to parabola. In Fig.2(b)- (d), we give the fidelity varied by the dissipations of cav- ities, magnons, and SC qubit. As a result of the virtual photon, the fidelity is almost unaffected by the SC qubit, shown in Fig.2(d). III. ENTANGLEMENT GENERATION FOR THREE NONLOCAL MAGNONS A. Entangled state of three nonlocal magnons Similar to the protocol of entangled state generation for two nonlocal magnons in two microwave cavities, we consider the protocol for entanglement of three nonlocalmagnons. As shown in Fig.3, similar to the hybrid sys- tem composed of two magnons coupled to the respective microwave cavities and a SC qubit in Fig.1, there are three magnons in three YIGs coupled to respective mi- crowave cavities and a SC qubit placed at the center of the three identical cavities ( ωa1=ωa2=ωa3=ωa). Each magnon is in biased static magnetic field and is located at the antinode of the microwave magnetic field. In the interaction picture, the Hamiltonian of the hy- brid system depicted in Fig.3 is H(I) 3=λm1a1m† 1eiδ1t+λm2a2m† 2eiδ2t +λm3a3m† 3eiδ3t+λqa1σ+ei∆1t +λqa2σ+ei∆2t+λqa3σ+ei∆3t+H.c.,(9) whereλm3is the coupling strength between magnon 3 and microwavecavity3, a3andm† 3areannihilation oper- ator of the cavity 3 and creation operator of the magnon 3, respectively. λqis the coupling between the SC qubit and three cavities, δ3=ωm3−ωa. The frequency ωm3 can be tuned by the biased magnetic field in microwave cavity 3. ∆ 3=ωq−ωa= ∆0. At the beginning we have the initial state |ψ/an}bracketri}ht(3) 0= |ψ/an}bracketri}ht(3) m⊗|ψ/an}bracketri}ht(3) a⊗|g/an}bracketri}htqwith|ψ/an}bracketri}ht(3) m=|1/an}bracketri}htm1|0/an}bracketri}htm2|0/an}bracketri}htm3=|100/an}bracketri}htmand5 FIG. 3: (Color online) Schematic of the hybrid system com- posed of three yttrium iron garnet spheres coupled to respec - tivemicrowave cavities. Asuperconductingqubit(blacksp ot) is placed at the center of the three cavities. |ψ/an}bracketri}ht(3) a=|0/an}bracketri}hta1|0/an}bracketri}hta2|0/an}bracketri}hta3=|000/an}bracketri}hta. Thesingle-excitationissetin the magnon 1. The magnon 1 is resonant with the cavity 1 by tuning the frequency of magnon 1, and the SC qubit is decoupled to the cavities. After time T(3) 1=π/2λm1, the local evolution |1/an}bracketri}htm1|0/an}bracketri}hta1→ −i|0/an}bracketri}htm1|1/an}bracketri}hta1is attained. The state is evolved to |ψ/an}bracketri}ht(3) 1=−i|000/an}bracketri}htm|100/an}bracketri}hta|g/an}bracketri}htq. (10) The SC qubit is coupled to the three identical mi- crowave cavities at the same time in far-detuning regime ∆0≫λq. Therefore, the effective Hamiltonian of the subsystem composed of the SC qubit and the three iden- tical cavities is of the form [83] H(3) eff=/tildewideλq/bracketleftbigg σz(a† 1a1+a† 2a2+a† 3a3)+3|e/an}bracketri}htq/an}bracketle{te| +σz(a1a† 2+a1a† 3+a2a† 3+H.c.)/bracketrightbigg .(11) The magnons are then all detuned with the cavities. The local evolution e−iH(3) efft|100/an}bracketri}hta|g/an}bracketri}htof the subsystem is given by |χ(t)/an}bracketri}ht(3) sub=/bracketleftbigg C(3) 1,t|100/an}bracketri}hta+C(3) 2,t|010/an}bracketri}hta+C(3) 3,t|001/an}bracketri}hta/bracketrightbigg ⊗|g/an}bracketri}htq, (12) whereC(3) 1,t=ei3/tildewideλqt+2 3andC(3) 2,t=C(3) 3,t=ei3/tildewideλqt−1 3. It is easy to derive that |C(3) 1,t|2+|C(3) 2,t|2+|C(3) 3,t|2= 1. (13) Fig.4 shows the probability related to the states |100/an}bracketri}hta|000/an}bracketri}htm|g/an}bracketri}htq,|010/an}bracketri}hta|000/an}bracketri}htm|g/an}bracketri}htqand|001/an}bracketri}hta|000/an}bracketri}htm|g/an}bracketri}htq. In0 1 2 3 4 5 6 70.00.51.0 3λ2 qt/∆0Pn |C1,t(3)|2 |C2,t(3)|2=|C3,t(3)|2 FIG. 4: (Color online) Evolution probabilities of the state s: P1=|C(3) 1,t|2for|100/angbracketrighta|000/angbracketrightm|g/angbracketrightq(red),P2=|C(3) 2,t|2for |010/angbracketrighta|000/angbracketrightm|g/angbracketrightq,P3=|C(3) 3,t|2for|001/angbracketrighta|000/angbracketrightm|g/angbracketrightq, andP2= P3(blue). particular, one has |C(3) 1,t|2=|C(3) 2,t|2=|C(3) 3,t|2=1 3, with C(3) 1=√ 3+i 2√ 3,C(3) 2=C(3) 3=−√ 3+i 2√ 3(14) at timeT(3) 2= 2π/9/tildewideλq. Correspondingly, the state evolves to |ψ/an}bracketri}ht(3) 2=/bracketleftbigg√ 3+i 2√ 3|100/an}bracketri}hta+−√ 3+i 2√ 3|010/an}bracketri}hta+−√ 3+i 2√ 3|001/an}bracketri}hta/bracketrightbigg ⊗(−i)|000/an}bracketri}htm⊗|g/an}bracketri}htq. (15) Finally, the magnonscan be resonatedwith the respec- tive cavities under the condition {δ1,δ2,δ3}= 0. The local evolution and the time are |0/an}bracketri}htmk|1/an}bracketri}htak→ −i|1/an}bracketri}htmk|0/an}bracketri}htak andT(3) 3k=π/2λmk(k= 1,2,3), respectively. Thus the final state is |ψ/an}bracketri}ht(3) 3=−/bracketleftbigg√ 3+i 2√ 3|100/an}bracketri}htm+−√ 3+i 2√ 3|010/an}bracketri}htm+−√ 3+i 2√ 3|001/an}bracketri}htm/bracketrightbigg ⊗|000/an}bracketri}hta⊗|g/an}bracketri}htq. (16) In the whole process, the state of the SC qubit is kept unchanged. B. Numerical result The entanglement fidelity of three nonlocal magnons is given here by taking into account the dissipations of hy- brid system. Firstly, the master equation which governs the realistic evolution of the hybrid system composed of three magnons, three microwave cavities and a SC qubit can be expressed as ˙ρ(3)=−i[H(I) 3,ρ(3)]+κm1D[m1]ρ(3)+κm2D[m2]ρ(3) +κm3D[m3]ρ(3)+κa1D[a1]ρ(3)+κa2D[a2]ρ(3) +κa3D[a3]ρ(3)+γqD[σ]ρ(3), (17)6 0.6 0.8 1.0 1.281838587 (κa)−1 [µs]Fidelity (%)(a) 0.8 1.0 1.2 1.4 1.6838485 (κm)−1 [µs]Fidelity (%)(b) 0.6 0.8 1.0 1.28283848586 (γq)−1 [µs]Fidelity (%)(c) FIG. 5: (a)-(c) The fidelity of the entanglement on three nonlocal magnons versus the dissipations of cavities, magn ons and SC qubit. whereρ(3)is the density operator of realistic evolu- tion of the hybrid system, κm3is the dissipation rate of magnon 3 with κm3/2π=κm/2π= 1.06 MHz [69],κa3denotes the dissipation rate for the microwave cavities 3 with κa3/2π=κa/2π= 1.35 MHz [69], D[X]ρ(3)=(2Xρ(3)X†−X†Xρ(3)−ρ(3)X†X)/2 for any X=m1,m2,m3,a1,a2,a3,σ. The entanglement fidelity for three nonlocal magnons is defined by F(3)=(3) 3/an}bracketle{tψ|ρ(3)|ψ/an}bracketri}ht(3) 3, which can reach 84.9%. The fidelity with respect to the parameters is FIG. 6: (Color online) Schematic of the hybrid system com- posed of Nyttrium iron garnet spheres coupled to respective microwave cavities. A superconducting qubit is placed at th e center of the Nidentical microwave cavities. shown in Fig.5. IV.NMAGNONS SITUATION In Sec. II and Sec. III, the entanglement of two and three nonlocal magnons have been established. In this section we consider the case of Nmagnons. In the hy- brid system shown in Fig.6, the SC qubit is coupled to Ncavity modes that have the same frequencies ωa. A magnon is coupled to the cavity mode in each cavity. Each magnon is placed at the antinode of microwave magnetic field of the respective cavity and biased static magnetic field. In the interaction picture the Hamiltonian of whole system shown in Fig.6 can be expressed as H(I) N=/summationdisplay n/bracketleftbigg λmn(anm† neiδnt+H.c.) +λq(anσ+ei∆nt+H.c.)/bracketrightbigg ,(18) whereanandm† n(n= 1,2,3,···,N) are the annihi- lation operator of the nth cavity mode and the creation operatorofthe nth magnon, λmnis the couplingbetween thenth magnon and the nth cavity mode, λqdenotes the coupling strength between the SC qubit and the nth cav- ity mode,δn=ωmn−ωa,ωmnis the frequency of the nth magnon, ∆ n= ∆0=ωq−ωa. The initial state is prepared as |ψ/an}bracketri}ht(N) 0=|ψ/an}bracketri}ht(N) m⊗|ψ/an}bracketri}ht(N) a⊗|g/an}bracketri}htq, (19) |ψ/an}bracketri}ht(N) m=|1/an}bracketri}htm1|0/an}bracketri}htm2|0/an}bracketri}htm3···|0/an}bracketri}htmN=|100···0/an}bracketri}htm, |ψ/an}bracketri}ht(N) a=|0/an}bracketri}hta1|0/an}bracketri}hta2|0/an}bracketri}hta3···|0/an}bracketri}htaN=|000···0/an}bracketri}hta. At first, we tune the frequency of magnon 1 under the conditionδ1= 0. The magnon 1 is resonant with the7 cavity 1, which means that the single photon is trans- mitted to cavity 1, and the SC qubit is decoupled to all the cavities. The state evolves to |ψ/an}bracketri}ht(N) 1=−i|000···0/an}bracketri}htm|100···0/an}bracketri}hta|g/an}bracketri}htq(20) after timeT(N) 1=π/2λm1. Next the magnons are tuned to detune with respec- tive cavities. The SC qubit is coupled to the Nmi- crowave cavities at the same time in far-detuning regime ∆0≫λq. Under the condition ∆ n= ∆0, the effective Hamiltonian of the subsystem composed of the SC qubit andNmicrowave cavities is of the form [83] H(N) eff=/summationdisplay n/tildewideλq/bracketleftbigg σza† nan+|e/an}bracketri}htq/an}bracketle{te|/bracketrightbigg +/summationdisplay l<n/tildewideλq/bracketleftbigg σz(ala† n+H.c.)/bracketrightbigg .(21) Consequently, the evolution of the hybrid system is given by |ψ/an}bracketri}ht(N) 2=/bracketleftbigg C(N) 1,t|100···0/an}bracketri}hta+C(N) 2,t|010···0/an}bracketri}hta +C(N) 3,t|001···0/an}bracketri}hta+···+C(N) N,t|000···1/an}bracketri}hta/bracketrightbigg ⊗(−i)|000···0/an}bracketri}htm⊗|g/an}bracketri}htq, (22) where C(N) 1,t=eiN/tildewideλqt+(N−1) N, C(N) 2,t=C(N) 3,t=···=C(N) N,t=eiN/tildewideλqt−1 N.(23) In addition, we have the following relation /summationdisplay n|C(N) n,t|2=|C(N) 1,t|2+|C(N) 2,t|2+|C(N) 3,t|2+···+|C(N) N,t|2 = 1 (24) by straightforward calculation. At last, the SC qubit is decoupled to the cavities, and the magnons are resonant with the cavities, respectively. Thus, after the time T(N) 3n=π/2λmn, the final state is given by |ψ/an}bracketri}ht(N) 3=−/bracketleftbigg C(N) 1,t|100···0/an}bracketri}htm+C(N) 2,t|010···0/an}bracketri}htm +C(N) 3,t|001···0/an}bracketri}htm+···+C(N) N,t|000···1/an}bracketri}htm/bracketrightbigg ⊗|000···0/an}bracketri}hta⊗|g/an}bracketri}htq. (25) In the whole process, the state of SC qubit is unchanged all the time. [Remark] Concerning the coefficients Eq. (23), the probabilities with respect to the states|100···0/an}bracketri}htm|000···0/an}bracketri}hta|g/an}bracketri}htq,|010···0/an}bracketri}htm|000···0/an}bracketri}hta|g/an}bracketri}htq, |001···0/an}bracketri}htm|000···0/an}bracketri}hta|g/an}bracketri}htq,···,|000···1/an}bracketri}htm|000···0/an}bracketri}hta|g/an}bracketri}htq arep(N) 1=|C(N) 1,t|2,p(N) 2=|C(N) 2,t|2,p(N) 3=|C(N) 3,t|2,···, p(N) N=|C(N) N,t|2, respectively, and p(N) 2=p(N) 3=···= p(N) N. If the condition p(N) 1=p(N) 2can be attained, the isoprobability entanglement can be obtained. For instance, for N= 4, the entangled state of the four nonlocal magnons is given by |ψ/an}bracketri}ht(4) 3=−1 2/bracketleftbigg |1000/an}bracketri}htm−|0100/an}bracketri}htm−|0010/an}bracketri}htm−|0001/an}bracketri}htm/bracketrightbigg ⊗|0000/an}bracketri}hta⊗|g/an}bracketri}htq. (26) However, if N/greaterorequalslant5, the isoprobability entanglement does not exist as a result of p(N) 1/ne}ationslash=p(N) 2, see illustration in Fig.7(b)(c). V. SUMMARY AND DISCUSSION We have presented protocols of establishing entan- glement on magnons in hybrid systems composed of YIGs, microwave cavities and a SC qubit. By exploit- ing the virtual photon, the microwave cavities can indi- rectly interact in far-detuning regime, and the frequen- cies of magnons can be tuned by the biased magnetic field, which leads to the resonant interaction between the magnons and the respective microwavecavities. We have constructed single-excitation entangled state on two and three nonlocal magnons, respectively, and the entangle- ment for Nmagnons has been also derived in term of the protocol for three magnons. By analyzing the coefficients in Eq. (23), the isoprob- ability entanglement has been also constructed for cases N= 2,N= 3 andN= 4. In particular, such isoproba- bility entanglement no longer exists for N/greaterorequalslant5. In the protocol for the case of two magnons discussed in Sec. II, we have firstly constructed the superposition of magnon 1 and microwave cavity 1. Then the photon could be transmitted |1/an}bracketri}hta1|0/an}bracketri}hta2→ −|0/an}bracketri}hta1|1/an}bracketri}hta2between two cavities. At last, the single-excitation Bell state is finally constructedinresonantway. Asfor N/greaterorequalslant3,however,such method is no longer applicable because of |100···0/an}bracketri}hta/notarrowright α2|010···0/an}bracketri}hta+α3|001···0/an}bracketri}hta+···+αN|000···1/an}bracketri}hta, namely, p(N) 1/ne}ationslash= 0. Acknowledgements This work is supported by the National Natural Sci- ence Foundation of China (NSFC) under Grant Nos. 12075159and 12171044,Beijing Natural Science Founda- tion (Grant No. Z190005), the Academician Innovation Platform of Hainan Province.8 0 2 4 6 8 10 12 1400.51 4λq2t/∆0pn(4)(a) p1(4) p2(4) 0 2 4 6 8 10 12 1400.51 5λq2t/∆0pn(5)(b) p1(5) p2(5) 0 2 4 6 8 10 12 1400.51 6λq2t/∆0pn(6)(c) p1(6) p2(6) FIG. 7: (a)-(c) Evolution probabilities for N= 4 (left), N= 5 (middle), and N= 6 (right). If N/greaterorequalslant5,p(N) 1/negationslash=p(N) 2implies that the isoprobability entanglement does not exist. [1] A. Einstein, B. Podolsky, and N. Rosen, Can quantum- mechanical description of physical reality be considered complete, Phys. Rev. 47, 777 (1935) [2] D. Bouwmeester, J.-W. Pan, M. Daniell, H. Wein- furter, and A. Zeilinger, Observation of three-photon Greenberger-Horne-Zeilinger entanglement, Phys. Rev. Lett.82, 1345 (1999) [3] W. D¨ ur, G. Vidal, and J. I. Cirac, Three qubits can be entangled in two inequivalent ways, Phys. Rev. A 62, 062314 (2000) [4] R. Horodecki, P. Horodecki, M. Horodecki, and K. Horodecki, Quantum entanglement, Rev. Mod. Phys. 81, 865 (2009) [5] J. M. Raimond, M. Brune, and S. Haroche, Manipulat- ing quantum entanglement with atoms and photons in a cavity,Rev. Mod. Phys. 73, 565 (2001) [6] K. G. H. Vollbrecht and J. I. Cirac, Delocalized entan- glement of atoms in optical lattices, Phys. Rev. Lett. 98, 190502 (2007) [7] M. Bina, F. Casagrande, A. Lulli, and E. Solano, Mon- itoring atom-atom entanglement and decoherence in a solvabletripartiteopensystemincavityQED, Phys. Rev. A77, 033839 (2008) [8] J. Le´ on and C. Sab´ ın, Photon exchange and correlation transfer in atom-atom entanglement dynamics, Phys. Rev. A79, 012301 (2009) [9] W.-A. Li and G.-Y. Huang, Deterministic generation of a three-dimensional entangled state via quantum Zeno dynamics, Phys. Rev. A 83, 022322 (2011) [10] M. Roghani, H. Helm, and H.-P. Breuer, Entanglement dynamics of a strongly driven trapped atom, Phys. Rev. Lett.106, 040502 (2011) [11] R. Rogers, N. Cummings, L. M. Pedrotti, and P. Rice, Atom-fieldentanglementincavityQED:nonlinearityand saturation, Phys. Rev. A 96, 052311 (2017) [12] H. Jo, Y. Song, M. Kim, and J. Ahn, Rydberg atom entanglements in the weak coupling regime, Phys. Rev. Lett.124, 033603 (2020) [13] X.-Y. Luo, Y. Yu, J.-L. Liu, M.-Y. Zheng, C.-Y. Wang, B. Wang, J. Li, X. Jiang, X.-P. Xie, Q. Zhang, X.-H. Bao, and J.-W. Pan, Postselected entanglement between two atomic ensembles separated by 12.5 km, Phys. Rev. Lett.129, 050503 (2022) [14] Z. Jin, S.-L. Su, A.-D. Zhu, H.-F. Wang, and S. Zhang, Engineering multipartite steady entanglement of distant atoms via dissipation, Front. Phys. 13, 134209 (2018)[15] S. Michalakis and B. Nachtergaele, Entanglement in finitely correlated spin states, Phys. Rev. Lett. 97, 140601 (2006) [16] G. T´ oth, C. Knapp, O. G¨ uhne, and H. J. Briegel, Spin squeezing and entanglement, Phys. Rev. A 79, 042334 (2009) [17] H, Zheng, H. T. Dung, and M. Hillery, Application of entanglement conditions to spin systems, Phys. Rev. A 81, 062311 (2010) [18] F. Troiani, S. Carretta, and P. Santini, Detection of en - tanglement between collective spins, Phys. Rev. B 88, 195421 (2013) [19] A. Str¨ om, H. Johannesson, and P. Recher, Controllable spin entanglement production in a quantum spin Hall ring,Phys. Rev. B 91, 245406 (2015) [20] J. C. Szabo and N. Trivedi, Entanglement dynamics be- tween Ising spins and a central ancilla, Phys. Rev. A 105, 052431 (2022) [21] V. Azimi-Mousolou, A. Bergman, A. Delin, O. Eriks- son, M. Pereiro, D. Thonig, and E. Sj¨ oqvist, Entangle- ment duality in spin-spin interactions, Phys. Rev. A 106, 032407 (2022) [22] A. Retzker, J. I. Cirac, and B. Reznik, Detecting vacuum entanglement in a linear ion trap, Phys. Rev. Lett. 94, 050504 (2005) [23] G.-X. Li, S.-P. Wu, and G.-M. Huang, Generation of entanglement and squeezing in the system of two ions trapped in a cavity, Phys. Rev. A 71, 063817 (2005) [24] F. L. Semi˜ ao and K. Furuya, Entanglement in the dis- persive interaction of trapped ions with a quantized field, Phys. Rev. A 75, 042315 (2007) [25] F. Nicacio, K. Furuya, and F. L. Semi˜ ao, Motional en- tanglement with trapped ions and a nanomechanical res- onator,Phys. Rev. A 88, 022330 (2013) [26] C. D. B. Bentley, A.R. R.Carvalho, D.Kielpinski, andJ. J. Hope, Detection-enhanced steady state entanglement with ions, Phys. Rev. Lett. 113, 040501 (2014) [27] K. Lake, S. Weidt, J. Randall, E. D.Standing, S. C. Web- ster, and W. K. Hensinger, Generation of spin-motion entanglement in a trapped ion using long-wavelength ra- diation, Phys. Rev. A 91, 012319 (2015) [28] J. Hannegan, J. D. Siverns, J. Cassell, and Q. Quraishi, Improving entanglement generation rates in trapped-ion quantumnetworks usingnondestructivephotonmeasure- ment and storage, Phys. Rev. A 103, 052433 (2021) [29] D. C. Cole, S. D. Erickson, G. Zarantonello, K. P. Horn,9 P.-Y. Hou, J. J. Wu, D. H. Slichter, F. Reiter, C. P. Koch, and D. Leibfried, Resource-efficient dissipative en- tanglement of two trapped-ion qubits, Phys. Rev. Lett. 128, 080502 (2022) [30] H. S. Eisenberg, G. Khoury, G. A. Durkin, C. Simon, and D. Bouwmeester, Quantum entanglement of a large number of photons, Phys. Rev. Lett. 93, 193901 (2004) [31] J. O. S. Yin and S. J. van Enk, Entanglement and purity of one- and two-photon states, Phys. Rev. A 77, 062333 (2008) [32] D. Salart, O. Landry, N. Sangouard, N. Gisin, H. Her- rmann, B. Sanguinetti, C. Simon, W. Sohler, R. T. Thew, A. Thomas, and H. Zbinden, Purification of single-photon entanglement, Phys. Rev. Lett. 104, 180504 (2010) [33] H. Wang, M. Mariantoni, R. C. Bialczak, M. Lenander, E. Lucero, M. Neeley, A. D. O’Connell, D. Sank, M. Wei- des, J. Wenner, T. Yamamoto, Y. Yin, J. Zhao, John M. Martinis, and A. N. Cleland, Deterministic entanglement of photons in two superconducting microwave resonators, Phys. Rev. Lett. 106, 060401 (2011) [34] X.-L. Wang, L.-K. Chen, W. Li, H.-L. Huang, C. Liu, C. Chen, Y.-H. Luo, Z.-E. Su, D. Wu, Z.-D. Li, H. Lu, Y. Hu, X. Jiang, C.-Z. Peng, L. Li, N.-L. Liu, Y.-A. Chen, C.-Y. Lu, and J.-W. Pan, Experimental ten-photon en- tanglement, Phys. Rev. Lett. 117, 210502 (2016) [35] O.Kfir, Entanglementsof electrons andcavityphotonsi n the strong-coupling regime, Phys. Rev. Lett. 123, 103602 (2019) [36] A. C. Dada, J. Kaniewski, C. Gawith, M. Lavery, R. H. Hadfield, D. Faccio, and M. Clerici, Near-maximal two-photon entanglement for optical quantum communi- cation at 2.1 µm,Phys. Rev. Applied 16, L051005 (2021) [37] J. V. Rakonjac, D. Lago-Rivera, A. Seri, M. Mazzera, S. Grandi, and Hugues de Riedmatten, Entanglement between a telecom photon and an on-demand multi- mode solid-state quantum memory, Phys. Rev. Lett. 127, 210502 (2021) [38] H. Wang, B.-C. Ren, A. H. Wang, A. Alsaedi, T. Hayat, and F.-G. Deng, General hyperentanglement concentra- tion for polarizationspatial- time-bin multi-photon sys- tems with linear optics, Front. Phys. 13, 130315 (2018) [39] T. Liu, Z.-F. Zheng, Y. Zhang, Y.-L. Fang, and C.-P. Yang, Transferring entangled states of photonic cat-state qubits in circuit QED, Front. Phys. 15, 21603 (2020) [40] U. Akram, W. Munro, K. Nemoto, and G. J. Milburn, Photon-phonon entanglement in coupled optomechanical arrays,Phys. Rev. A 86, 042306 (2012) [41] S.Finazzi andI. Carusotto, Entangled phononsinatomi c Bose-Einstein condensates, Phys. Rev. A 90, 033607 (2014) [42] S.-S. Chen, H. Zhang, Q. Ai, and G.-J. Yang, Phononic entanglement concentration via optomechanical interac- tions,Phys. Rev. A 100, 052306 (2019) [43] M. Di Tullio, N. Gigena, and R. Rossignoli, Fermionic entanglement in superconducting systems, Phys. Rev. A 97, 062109 (2018) [44] M. Gong, M.-C. Chen, Y. Zheng, S. Wang, C. Zha, H. Deng, Z. Yan, H. Rong, Y. Wu, S. Li, F. Chen, Y. Zhao, F. Liang, J. Lin, Y. Xu, C. Guo, L. Sun, A. D. Castellano, H. Wang, C. Peng, C.-Y. Lu, X. Zhu, and J.- W. Pan, Genuine 12-qubit entanglement on a supercon- ducting quantum processor, Phys. Rev. Lett. 122, 110501 (2019)[45] W. Ning, X.-J. Huang, P.-R. Han, H. Li, H. Deng, Z.-B. Yang, Z.-R. Zhong, Y. Xia, K. Xu, D. Zheng, and S.-B. Zheng, Deterministic entanglement swapping in a super- conducting circuit, Phys. Rev. Lett. 123, 060502 (2019) [46] Y. Ma, X. Pan, W. Cai, X. Mu, Y. Xu, L. Hu, W. Wang, H. Wang, Y. P. Song, Z.-B. Yang, S.-B. Zheng, and L. Sun, Manipulating complex hybrid entanglement and testing multipartite bell inequalities in a supercon- ducting circuit, Phys. Rev. Lett. 125, 180503 (2020) [47] S. Krastanov, H. Raniwala, J. Holzgrafe, K. Jacobs, M. Lonˇ car, M. J. Reagor, and D. R. Englund, Optically her- alded entanglement of superconducting systems in quan- tum networks, Phys. Rev. Lett. 127, 040503 (2021) [48] A. Cervera-Lierta, M. Krenn, A. Aspuru-Guzik, and A. Galda, Experimental high-dimensional Greenberger- Horne-Zeilinger entanglement with superconducting transmon qutrits, Phys. Rev. Applied 17, 024062 (2022) [49] J. Li, S.-Y. Zhu, and G. S. Agarwal, Magnon-photon- phonon entanglement in cavity magnomechanics, Phys. Rev. Lett. 121, 203601 (2018) [50] J. Li and S.-Y. Zhu, Entangling two magnon modes via magnetostrictive interaction, New J. Phys. 21, 085001 (2019) [51] H. Y. Yuan, S. Zheng, Z. Ficek, Q. Y. He, and M.- H. Yung, Enhancementof magnon-magnon entanglement inside a cavity, Phys. Rev. B 101, 014419 (2020) [52] D. Kong, X. Hu, L. Hu, and J. Xu, Magnon-atom in- teraction via dispersive cavities: magnon entanglement, Phys. Rev. B 103, 224416 (2021) [53] V. Azimi Mousolou, Y. Liu, A. Bergman, A. Delin, O. Eriksson, M. Pereiro, D. Thonig, and Erik Sj¨ oqvist, Magnon-magnon entanglement and its quantification via a microwave cavity, Phys. Rev. B 104, 224302 (2021) [54] F. Wang, C. Gou, J. Xu, and C. Gong, Hybrid magnon- atom entanglement and magnon blockade via quantum interference, Phys. Rev. A 106, 013705 (2022) [55] B. Rogers, M. Paternostro, G. M. Palma, and G. De Chiara, Entanglement control in hybrid optomechanical systems, Phys. Rev. A 86, 042323 (2012) [56] M. Borrelli, M. Rossi, C. Macchiavello, and S. Manis- calco, Witnessing entanglement in hybrid systems, Phys. Rev. A90, 020301(R) (2014) [57] C. Joshi, J. Larson, and T. P. Spiller, Quantum state engineering in hybrid open quantum systems, Phys. Rev. A93, 043818 (2016) [58] Q.-Z. Ye, Z.-T. Liang, W.-X. Zhang, D.-J. Pan, Z.-Y. Xue, and H. Yan, Speedup of entanglement generation in hybrid quantum systems through linear driving, Phys. Rev. A106, 012407 (2022) [59] Z.-L. Xiang, S. Ashhab, J. Q. You, and F. Nori, Hybrid quantum circuits: superconducting circuits interacting with other quantum systems, Rev. Mod. Phys. 85, 623 (2013) [60] A. Blais, R.-S. Huang, A. Wallraff, S. M. Girvin, and R. J. Schoelkopf, Cavity quantum electrodynamics for superconducting electrical circuits: an architecture for quantum computation, Phys. Rev. A 69, 062320 (2004) [61] J. Q. You and F. Nori, Superconducting circuits and quantum information, Phys. Today 58, 42 (2005) [62] A. Blais, A. L. Grimsmo, S. M. Girvin, and A. Wallraff, Circuit quantum electrodynamics, Rev. Mod. Phys. 93, 025005 (2021) [63] H. Y. Yuan, P. Yan, S Zheng, Q. Y. He, K. Xia, and M.-H. Yung, Steady Bell state generation via magnon-10 photon coupling, Phys. Rev. Lett. 124, 053602 (2020) [64] Y. Zhao, L. Wang, J. Xue, Q. Zhang, Y. Tian, S. Yan, L. Bai, M. Harder, Q. Guo, and Y. Zhai, Measuring the magnon phase in a hybrid magnon-photon system, Phys. Rev. B105, 174405 (2022) [65] X. Zhang, C.-L. Zou, L. Jiang, and H. Tang, Cavity mag- nomechanics, Sci. Adv. 2, e1501286 (2016) [66] S. Streib, N. Vidal-Silva, K. Shen, and G. E. W. Bauer, Magnon-phonon interactions in magnetic insula- tors,Phys. Rev. B 99, 184442 (2019) [67] M. Yu, H. Shen, and J. Li, Magnetostrictively induced stationary entanglement between two microwave fields, Phys. Rev. Lett. 124, 213604 (2020) [68] S.-F. Qi and J. Jing, Magnon-assisted photon-phonon conversion in the presence of structured environments, Phys. Rev. A 103, 043704 (2021) [69] X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, Strongly coupled magnons and cavity microwave photons, Phys. Rev. Lett. 113, 156401 (2014) [70] Y.-P. Wang, G.-Q. Zhang, D. Zhang, X.-Q. Luo, W. Xiong, S.-P. Wang, T.-F. Li, C.-M. Hu, and J. Q. You, Magnon Kerr effect in a strongly coupled cavity-magnon system,Phys. Rev. B 94, 224410 (2016) [71] Y.-P. Wang, G.-Q. Zhang, D. Zhang, T.-F. Li, C.-M. Hu, and J. Q. You, Bistability of cavity magnon polaritons, Phys. Rev. Lett. 120, 057202 (2018) [72] R. Hisatomi, A. Osada, Y. Tabuchi, T. Ishikawa, A. Noguchi, R. Yamazaki, K. Usami, and Y. Nakamura, Bidirectional conversion between microwave and light via ferromagnetic magnons, Phys. Rev. B 93, 174427 (2016) [73] J. Li, S.-Y. Zhu, and G. S. Agarwal, Squeezed states of magnons and phonons in cavity magnomechanics, Phys. Rev. A99, 021801(R) (2019) [74] J. Zou, S. K. Kim, and Y. Tserkovnyak, Tuning entan- glement by squeezing magnons in anisotropic magnets, Phys. Rev. B 101, 014416 (2020) [75] V. A. S. V. Bittencourt, V. Feulner, and S. V. Kusmin-skiy, Magnon heralding in cavity optomagnonics, Phys. Rev. A100, 013810 (2019) [76] C. Kittel, On the theory of ferromagnetic resonance ab- sorption, Phys. Rev. 73, 155 (1948) [77] J. Chen, C. Liu, T. Liu, Y. Xiao, K. Xia, G. E.W.Bauer, M. Wu, and H. Yu, Strong interlayer magnon-magnon coupling in magnetic metal-insulator hybrid nanostruc- tures,Phys. Rev. Lett. 120, 217202 (2018) [78] L. Liensberger, A. Kamra, H. Maier-Flaig, S. Gepr¨ ags, A. Erb, S. T.B.Goennenwein, R. Gross, W. Belzig, H. Huebl, and M. Weiler, Exchange-enhanced ultrastrong magnon-magnon coupling in a compensated ferrimagnet, Phys. Rev. Lett. 123, 117204 (2019) [79] Y. Li, V. G. Yefremenko, M. Lisovenko, C. Trevillian, T. Polakovic, T. W. Cecil, P. S. Barry, J. Pearson, R. Divan, V. Tyberkevych, C. L. Chang, U. Welp, W.-K. Kwok, and V. Novosad, Coherent coupling of two re- mote magnonic resonators mediated by superconducting circuits, Phys. Rev. Lett. 128, 047701 (2022) [80] Y. Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Ya- mazaki, K. Usami, and Y. Nakamura, Coherent coupling between a ferromagnetic magnon and a superconducting qubit,Science349, 405 (2015) [81] D. Lachance-Quirion, S. P. Wolski, Y. Tabuchi, S. Kono, K. Usami, andY. Nakamura, Entanglement-basedsingle- shot detection of asingle magnon with asuperconducting qubit,Science367, 425 (2020) [82] J.-K. Xie , S.-L Ma, and F.-L. Li, Quantum-interference - enhanced magnon blockade in an yttrium-iron-garnet sphere coupled to superconducting circuits, Phys. Rev. A101, 042331 (2020) [83] D. F. James and J. Jerke, Effective Hamiltonian the- ory and its applications in quantum infromation, Can. J. Phys. 85, 625 (2007) [84] I. Buluta and F. Nori, Quantum Simulators, Science326, 108 (2009) | 2023-01-14 | We present protocols to generate quantum entanglement on nonlocal magnons in
hybrid systems composed of yttrium iron garnet (YIG) spheres, microwave
cavities and a superconducting (SC) qubit. In the schemes, the YIGs are coupled
to respective microwave cavities in resonant way, and the SC qubit is placed at
the center of the cavities, which interacts with the cavities simultaneously.
By exchanging the virtual photon, the cavities can indirectly interact in the
far-detuning regime. Detailed protocols are presented to establish entanglement
for two, three and arbitrary $N$ magnons with reasonable fidelities. | Quantum entanglement generation on magnons assisted with microwave cavities coupled to a superconducting qubit | 2301.05820v1 |
1 Observation of nonlinear planar Hall effect in magnetic insulator/topological insulator heterostructures Yang Wang1, Sivakumar V. Mambakkam2, Yue-Xin Huang3, Yong Wang2, Yi Ji1, Cong Xiao4,5, Shengyuan A. Yang3, Stephanie A. Law1,2, and John Q. Xiao1,* 1 Department of Physics and Astronomy, University of Delaware, Newark, Delaware 19716, USA 2 Department of Materials Science and Engineering, University of Delaware, Newark, Delaware, 19716, USA 3 Research Laboratory for Quantum Materials, Singapore University of Technology and Design, Singapore 487372, Singapore 4 Department of Physics, The University of Hong Kong, Hong Kong, China 5 HKU -UCAS Joint Institute of Theoretical and Computational Physics, Hong Kong, China *jqx@udel.edu Abstract Interfacing topological insulators (TIs) with magnetic insulators (MIs) has been widely used to study the interaction between topological surface states and magnetism. Previous transport studies typic ally interpret the suppression of weak antilocalization or appearance of the anomalous Hall effect as signatures of magnetic proximity effect (MPE) imposed to TIs. Here, we report the observation of nonlinear planar Hall effect (NPHE) in Bi 2Se3 films grown on MI thulium and yttrium iron garnet (TmIG and YIG) substrates, which is an order of magnitude larger than that in Bi 2Se3 grown on nonmagnetic gadolinium gallium garnet (GGG) substrate. The nonlinear Hall resistance in TmIG/Bi 2Se3 depends linearly on the external magnetic field, while that in YIG/Bi 2Se3 exhibits an extra hysteresis loop around zero field. The magnitude of the NPHE is found to scale inversely with carrier density. We speculate the observed NPHE is related to the MPE -induced exchange gap opening and out -of-plane spin textures in the TI surface states, which may be used as an alternative transport signature of the MPE in MI/TI heterostructures. I. INTRODUCTION Magnetic topological insulators [1] have been intensively stu died in the past decade, because scientifically fundamental as well as technologically promising phenomena like quantum anomalous Hall effect (QAHE) [2] and topological magnetoelectric effect [ 3] can arise in the surface states of magnetic TIs. So far doping magnetic elements like Cr or V [4,5] into the TI (Bi,Sb) 2Te3 has been the most developed and robust method to observe QAHE [ 6,7]. Another approach to magnetize the topological surface states (TSS) is to proximately couple the TI to a magnetic material [8 -11], preferentially an insulator, so there is no doping or current shunting effect. Although robust anomalous Hall effect (AHE) has been reported in (Bi,Sb) 2Te3 coupled to MIs like TmIG [12] o r Cr 2Ge2Te6 [13], weak or absence of AHE is more commonly seen in Bi 2Se3 or Bi 2Te3 when grown or transferred on top of MI substrates, and the suppression of weak antilocalization (WAL) is often taken a s a qualitative signature of the MPE [14 -2 17]. Conventional surface -sensitive spectroscopic methods like a ngle-resolved photoemission spectroscopy (ARPES) cannot be applied to probe the band structure at the buried MI/TI interface, and x - ray- or neutron -based measurements have not been successful in detecting any induced magnetic moments in TIs grown on M Is [18 -20]. Therefore, searching an alternative and more sensitive transport probe of the MPE will be helpful to clarify and understand the interaction between magnetism and TSS in MI/TI heterostructures . In recent years, planar Hall effect (PHE) in both linear [21] and nonlinear [22] re gimes have been reported in nonmagnetic TIs. The linear PHE was interpreted as a result of magnetic -field-induced anisotropic backscattering [21] or magnetic -field-induced tilting of the Dirac cone with particle -hole (p - h) asymmetry [23,24]. The nonlinear PHE was attributed to the distortion or tilting of the Dirac dispersion with higher -order k terms like p -h asymmetry ( k2) or hexagonal warping ( k3) by the external magnetic field [22,24] . In this work, we report the observation of NPHE in Bi 2Se3 (BS) films grown on magnetic Tm 3Fe5O12 (TmIG) and Y 3Fe5O12 (YIG) substrates, which is an order of magnitude larger than that in Bi 2Se3 film grown on nonmagnetic Gd 3Ga5O12 (GGG) substrate. While the NPHE in TmIG/BS with out -of-plane (OP) easy axis shows a linear dependence on the in -plane (IP) magnetic field, the NPHE in YIG/BS with IP easy axis takes an extra hysteretic jump around zero field, reflecting the reversal of the IP YIG magnetization. The same temperature dependence of the linear -in-B and hysteretic components suggests they share the same origin. The carrier density dependence of the NPHE excludes hexagonal warping as the dominant mechanism. The enhancement of the NPHE in Bi 2Se3 grown on magnetic substrates as well as the sharp increase o f it below 30 K indicate MPE plays a critical role. Our results suggest the NPHE may work as a convenient and sensitive transport probe of the MPE in MI/TI heterostructures. II. METHODS The 8 quintuple -layer (QL) Bi 2Se3 films used in this study were grown on Tm 3Fe5O12, Y3Fe5O12, and Gd3Ga5O12 substrates with very close lattice constants and smooth surfaces in a molecular beam epitaxy (MBE) system with a base pressure of 1 ×10-9 Torr, following the two -step Se -buffer layer method reported in Ref. [25]. The TmIG (30 nm) and two YIG (2.5 𝜇m and 100 nm) films with OP and IP magnetic anisotropy respectively were deposited on GGG(111) substrates by magnetron sputtering or liquid phase epitaxy method . Prior Bi 2Se3 growth, the Tm IG and YIG substrates were soaked in Piranha solution (H2SO 4:H2O2=3:1) for 5 min to clean the surface [26]. After annealing the substrates at 650 °C for 30 min and cooling down to 50 °C in the growth chamber, ~2 nm thick amorphous Se and 1 nm Bi xSe1-x were deposited. Then the substrate temperature was slowly ramped to 325 °C at 10 °C/min to evaporate extra Se and crystalize the first QL Bi 2Se3. And the remaining 7 QL Bi 2Se3 was subsequently deposited by co - evaporating Bi and Se. After cooling to room t emperature, the Bi 2Se3 films were capped with 5 nm SiO 2 layer for protection in another magnetron sputtering chamber. As shown in Fig. 1(a), streaky reflection high-energy electron diffraction (RHEED) patterns were observed from the very first QL and x-ray diffraction (XRD) results of all the Bi 2Se3 films exhibit clear (0,0,3 n) peaks, indicating the c-axis growth orientation and the high and similar crystalline quality. 3 After the growth, the Bi 2Se3 films were fabricated into 200×100 𝜇m Hall bar devices by standard photolithography method and contacted with Ti/Au electrodes. The magneto -transport measurements were carried out in a home -built cryogenic system with base temperature 4.5 K. Low -frequency lock -in technique was used to detect the first - and second -harmonic longitudinal and transverse voltages. III. RESULTS AND DISCUSSION A. Linear transport properties Fig. 2(a) displays the temperature dependence of the longitudinal resistance of four Bi 2Se3 devices made on TmIG, YIG and GGG substrates. All of them show metallic behavior above 50 K . Below 50 K, the resistance upturn in TmIG and YIG/BS samples is more pronounced compared to that of the GGG/BS sample, especially in the YIG2/BS sample with a lower carrier density. Such insulating behavior may result from the suppression of WAL due to MPE [27] and is consistent with previous reports on iron garnet/Bi 2Se3 bilayers [14,16,17] . As shown in Fig. 2(b), from ordinary Hall effect (OHE) measurements we extracted the sheet carrier densities for these devices to be 2.45 -3.46 ×1013 cm-2. This means these Bi2Se3 samples are n-doped with a Fermi level of ~0.3 eV, so significant amount of the current is carried by the bulk states. We did not observe hysteretic or nonlinear AHE in the TmIG or YIG/BS samples after subtracting the linear OHE background. This suggests that the garnet/Bi 2Se3 interfaces formed here may not be as good as those in Ref. [17], where weak AHE was observed. Thi s does not rule out the existence of a small exchange -interaction -induced gap in the TSS in our samples, because when the gap ∆ is small, e.g. ~1 meV , and it is much smaller than the Fermi energy 𝜀F, the AH conductivity 𝜎𝑦𝑥AH∝8𝑒2 ℏ(∆ 𝜀F)3 [28] will give rise to a n AH resistance in the order of 0.1 mΩ, which can not be discerned from the large OHE background. Although AHE was not detected, we observed suppression of WAL in Bi 2Se3 films grown on TmIG and YIG substrates as compared with that on GGG [Fig. 2(c)]. By fitting to the Hikami -Larkin - Nagaoka equation, the electron phase coherence of length of the three Bi 2Se3 samples on GGG, TmIG, and YIG are 203, 119, and 90 nm, respectively. Given the similar crystalline quality [Fig. 1] and carrier densities [Fig. 2(b)], the suppressed WAL in TmIG and YIG/BS is most likely due to MPE -induced OP spin textures and correspondingly, the reduced Berry phase of TI surface electrons [30]. B. Observat ion of the NPHE The linear PHE has sin𝜙𝐵cos𝜙𝐵 dependence on the IP magnetic field direction [21,23] where 𝜙𝐵 is the angle between the current and magnetic field directions. Thus, the first -order Hall voltage is zero at 𝜙𝐵=𝑛90° with n being an integer. Differently, the NPHE depends on the IP magnetic field as 𝐵cos𝜙𝐵 [22,24], so it can be detected by sweepin g the magnetic field between 0 and 180 °. As illustrated in Fig. 3(a) inset, in our experiments we sent a sinusoidal a.c. current to the Hall channel in the x-direction, and measured the second harmonic Hall voltage 𝑉𝑦2𝜔 while sweeping the external magne tic field also in the x- direction. As shown in Fig. 3, the second harmonic Hall resistance 𝑅𝑦𝑥2𝜔=𝑉𝑦2𝜔/𝐼 as a function of B for three Bi 2Se3 devices fabricated on GGG, TmIG, and YIG substrates have dramatically different behaviors. Compared with the TmIG and YIG/BS samples, the magnitude of the NPHE in GGG/BS measured by the slope 𝑑𝑅 𝑦𝑥2𝜔/𝑑𝐵 is an order of magnitude smaller. In the TmIG/BS bilayer with OP magnetic anisotropy, 4 the NPHE is enhanced and exhibits a linear dependence on the IP magne tic field [Fig. 3(b)]. When Bi2Se3 is grown on YIG substrate with IP anisotropy, 𝑅𝑦𝑥2𝜔 not only exhibits a linear dependence on B, but also takes a hysteretic shape around zero field, corresponding to the switching of the IP magnetization of YIG. Moreover, the NPHE in both TmIG and YIG/BS samples is only observable below ~30 K and shows sharp increase when temperature is decreased from 30 K to 4.5 K. Such enhance d NPHE with similar characteristics was observed in different YIG/Bi 2Se3 samples grown at different times and in different TmIG/Bi 2Se3 devices , confirming the reproducibility of the results . The NPHE previously reported in Al 2O3/Bi2Se3 was attributed to the nonlinear, i.e., the hexagonal warping k3 and the p -h asymmetry k2 terms in the topological surface dispersion [22]. Although this contribution should exist in all three samples shown here, it cannot account the enhancement of the NP HE in TmIG and YIG/BS. When scaled by the coefficient 𝛾𝑦≡𝑅𝑦𝑥2𝜔 𝑅𝑥𝑥𝐼𝐵, the magnitude of the NPHE in TmIG and YIG/BS at 4.5 K is 0.025 and 0.031 respectively, which is one and two orders of magnitude larger than that in GGG/BS ( 𝛾𝑦=3×10-3) and Al2O3/BS (𝛾𝑦=1×10-4) [23] , respectively. Similarly, if Nernst effect was responsible for the measured second harmonic voltage, it should appear on the same order of magnitude in all three samples. The large difference shows that it should not be th e dominant contribution. Therefore, the observed NPHE is likely to have a magnetic origin. Ref. [31] reported a large hysteretic NPHE in magnetic TI Cr x(Bi,Sb) 2-xTe3/(Bi,Sb) 2Te3 heterostructures and explains it as a result of asymmetric scattering of surface electrons by magnons. This magnon scattering mechanism cannot be applied to the TmIG/BS sample, because with OP magnetization, the magnons are polarized in the z-direction an d cannot participate the scattering of surface electrons with IP - polarized spins. To see whether it is responsible for the hysteretic loop of the 𝑅𝑦𝑥2𝜔 𝑣𝑠 𝐵 curve in the YIG/BS sample, we parsed the NPHE into the linear -in-B (𝑑𝑅 𝑦𝑥2𝜔/𝑑𝐵) and hysteretic ( ∆𝑅𝑦𝑥2𝜔) part, as displayed in Fig. 4(a). Fig. 4(b) shows that these two components have almost the same temperature dependence, suggesting that they share the same origin. Therefore, in the YIG/BS sample, it is likely that the IP exchange field experienced by the Dirac electrons from the IP magnetic moments of YIG plays the same role with the external magnetic field, which gives rise to the hysteretic loop around zero field. We note that although the easy axis of YIG is IP, the Dirac -electron -mediated exchange interaction [8,32] tilts the magnetic moments to the OP direction at the interface [8,32 -34], which can also open an exchange gap for the TSS, just like that in TmIG/BS. As will be discussed in Section III.C, this gap o pening may play a decisive role in generating the large NPHE in MI/TI heterostructures. The other characteristics of the NPHE are shown in Fig. 5. First, the second -harmonic Hall resistance divided by field 𝑑𝑅𝑦𝑠2𝜔/𝑑𝐵 depends linear on the current de nsity [Fig. 5(a) top inset], demonstrating its nonlinear nature. The deviation from linear relationship at higher current densities is due to Joule heating. Second, as shown in Fig. 5(a), when scanned under a y-direction field, 𝑅𝑦𝑥2𝜔 becomes much small er, consistent with the cos𝜙𝐵 dependence of NPHE. Third, we also measured the longitudinal second - harmonic resistance 𝑅𝑥𝑥2𝜔 under a y-direction magnetic field scan and observed the longitudinal counterpart response , the so-called bilinear magnetoresistance (BMR) [35]. Lastly, as summarized in Fig. 5(b), the magnitude of the NPHE measure by 𝛾𝑦 is enhanced at low carrier densities. This also points out the insignificant role of the hexagonal warping effect in the NPHE observed here, because hexagonal warping is negligible at low Fermi levels and is enhanced when carrier density increases . 5 C. Discussion The nonlinear charge current under IP electric and magnetic fields can be expressed as 𝑗𝑎(2)= 𝜒𝑎𝑏𝑐𝑑 𝐸𝑏𝐸𝑐𝐵𝑑, where 𝜒𝑎𝑏𝑐𝑑 is the nonlinear conductivity tensor and 𝑎,𝑏,𝑐,𝑑=𝑥 or 𝑦. For an ideal linear Dirac dispersion, an IP magnetic field merely shift s the Dirac cone in k-space, leaving no effect on transport properties [22 -24]. However, in real TIs, higher -order k terms exist like the quadratic p -h asymmetry term and the cubic warping term. Here, for simplicity, we consider a linear TI dispersion with a parabolic 𝐷𝑘2 term. As sketched in Fig. 6(a) , when there is no IP magnetic field, under a driving electric field, there are equal number of electrons moving to the left and right with opposite spin polarizations. This generates a second -order spin current 𝑗s(2) but there is no net charge current. When an IP magnetic field 𝐵𝑥 is applied [Fig. 6(b)], it not only shifts the Dirac cone, but also tilts it [23,24] in the y-direction due to the existence of the 𝑘2 term. As a result, the transverse currents carried by the left - and right -moving electrons with opposite spin polarizations no longer cancel each other, and the nonlinear spin current is partially converted into a nonlinear charge current [22,35] , giving rise to a nonlinear Hall conductivity 𝜒𝑦𝑥𝑥𝑥. When the TI is further coupled to a magnetic material , the exchange interaction with OP moments opens a gap and introduces OP spin textures to the TSS [30]. Expanding around the shifted Dirac point, the Hamiltonian of TSS can be put in the form of 𝐻=ℏ𝑣F(𝑘𝑥𝜎𝑦−𝑘𝑦𝜎𝑥)+𝛼𝐵𝑥𝑘𝑦+∆𝜎𝑧 [23,24] , where ℏ is the reduced Planck constant, 𝑣F the Fermi velocity, and 𝝈 the Pauli matrices. The term with coefficient 𝛼𝐵𝑥 describes the magnetic -field-induced tilting in the y-direction , and 2∆ is the OP exchange - interaction -induced gap. The IP magnetic field Bx ~0.1 T used in this study is presumably smaller than the perpendicular magnetic anisotropy fields, so the size of the exchange gap is not affected by the small Bx. For similar tilted mas sive Dirac models, previous theoretical studies show that both intrinsic Berry -phase - related [36 -39] and extrinsic skew -scattering or side -jump mechanisms [40-42] can contribute to the NPHE with different 𝜏 scaling . In our experiment, the one to two orders of magnitude increase of 𝜒𝑦𝑥𝑥𝑥 (∝𝑑𝑅𝑦𝑥2𝜔/𝐼𝑑𝐵) [Fig. 3] from 30 K to 4.5 K and the relatively small change of the linear conductivity in this regime [Fig. 2(a)] suggest that some other factor other than t he relaxation time, governs the temperature dependence of the NPHE. As a result, we are not able to use 𝜏-scaling to narrow down the candidate NPHE mechanisms. More systematic future study is needed to clarify the dominant mechanism in the observed effect . Our experimental result s also suggest the importance of the OP spin textures formed in the TSS (due to MPE) for the NPHE , which was not considered in previous studies. As sketched in Fig. 6(c), with broken time-reversal symmetry, the spins of the electro ns with ±𝑘 momenta are no longer orthogonal with each other, and backscattering between these states by nonmagnetic impurities is allowed. This is reflected as the suppression of WAL in the linear transport regime [Fig. 2(c)] . Ref. [23] shows that the tilting -induced PHE on the surface of TIs can be enhanced by scattering off nonmagnetic impurities. Here, we speculate that similar effect also occur s in the nonlinear regime. When backscattering is allowed due to OP spin texture formation, the nonlinear s pin to charge current conversion may be increased , resulting in the large NPHE in MI/TI heterostructures. We note that in the above analysis, we only considered the TI surface states and did not consider the contribution from the bulk states . Because of th e short -range nature of the 6 MPE, the inversion symmetry of the bulk is presumably preserved. As a result, neither second -order spin nor charge current can be generated from the bulk [4 3]. IV. CONCLUSION In summary, we observed enhanced nonlinear planar Hall effect in Bi 2Se3 films grown on magnetic TmIG and YIG substrates as compared to that on nonmagnetic GGG substrate. This NPHE is only observable below 30 K and scales inversely with carrier density. Compared with the previously reported NPHE in Al 2O3/Bi2Se3 [22] arising from hexagonal warping or p -h asymmetry, the NPHE in MI/TI heterostructures is orders of magnitude larger, indicating a different origin. In YIG/Bi 2Se3 we find the IP exchange field plays the same role as the external magnetic field, giv ing rise to an extra hysteresis loop in the 𝑅𝑦𝑥2𝜔 𝑣𝑠 𝐵 scans. Actually, a large hysteretic NPHE was previously observed in EuS/(Bi,Sb) 2Te3 bilayers in Ref. [4 4], and a mechanism based on tilting of the p -h asymmetric Dirac cone by the IP exchange field was speculated. Our control experiments suggest the necessary role of the OP exchange - interaction -induced gap opening or modified spin textures in generating the large NPHE. Further experimental and theoretical work is needed to reveal the u nderlying mechanism and establish the NPHE as a convenient and sensitive transport probe of the MPE in MI/TI heterostructures. ACKNOWLEDGMENTS This work was supported by the U.S. DOE, Office of Basic Energy Sciences under Contract No. DE- SC0016380 and by NSF DMR Grant No. 1904076. Y.-X. H and S. A. Yang are supported by Singapore NRF CRP22 -2019 -0061. C. X. is supported by the UGC/ RGC of Hong Kon g SAR (AoE/P -701/20). The authors acknowledge the use of the Materials Growth Facility (MGF) at the University of Delaware, which is partially supported by the National Science Foundation Major Research Instrumentation under Grant No. 1828141 and UD -CHARM, a National Science Foundation MRSEC, under Award No. DMR - 2011824. 7 References [1] Y. Tokura, K. Yasuda, and A. Tsukazaki, Magnetic Topological Insulators , Nat. Rev. Phys. 1, 126 (2019). [2] R. Yu, W. Zhang, H. J. Zhang, S. C. Zhang, X. Dai, and Z. Fang, Quantized anomalous Hall effect in magnetic topological insulators , Science 329, 61 (2010). [3] X. L. Qi., T. L. Hughes, and S. C. Zhang, Topological field theory of time -reversal invariant insula tors, Phys. Rev. B 78, 195424 (2008). [4] C. Z. Chang et al., Thin Films of Magnetically Doped Topological Insulator with Carrier - Independent Long -Range Ferromagnetic Order , Adv. Mater. 25, 1065 (2013). [5] M. Mogi, R. Yoshimi, A. Tsukazaki , K. Yasuda, Y. Kozuka, K. S. Takahashi, M. Kawasaki, and Y. Tokura, Magnetic modulation doping in topological insulators toward higher -temperature quantum anomalous Hall effect , Appl. Phys. Lett. 107, 182401 (2015). [6] C.-Z. Chang et al., Experimental observ ation of the quantum anomalous Hall effect in a magnetic topological insulator , Science 340, 167 (2013). [7] C.-Z. Chang, W. Zhao, D. Y. Kim, H. Zhang, B. A. Assaf, D. Heiman, S. -C. Zhang, C. Liu, M. H. W. Chan, and J. S. Moodera, High -Precision Realization of Robust Quantum Anomalous Hall State in a Hard Ferromagnetic Topological Insulator , Nat. Mater. 14, 473 ( 2015). [8] Q. Liu, C. -X. Liu, C. Xu, X. -L. Qi, and S. -C. Zhang, Magnetic impurities on the surface of a topological insulator , Phys. Rev. Lett. 102, 156603 (2009). [9] L. A. Wray, S. -Y. Xu, Y. Xia, D. Hsieh, A. V. Fedorov, Y. S. Hor, R. J. Cava, A. Bansil, H. Lin, and M. Z. Hasan, A Topological Insulator Surface under Strong Coulomb, Magnetic and Disorder Perturbations , Nat. Phys. 7, 32 (2011). [10] S. Bhattacharyya, G. Akhgar, M. Gebert, J. Karel, M. T. Edmonds, and M. S. Fuhrer , Recent progress in proximity coupling of magnetism to topological insulators , Adv. Mater. 33, 2007795 (2021 ). [11] A. J. Grutter and Q. L. He, Magnetic proximity effects in topological insulator heterostructures: Implementation and characterization , Phys. Rev. Mater. 5, 090301 (2021) . [12] C. Tang, C. -Z. Chang, G. Zhao, Y. Liu, Z. Jiang, C. -X. Liu, M. R. McCartney, D. J. Smith, T. Chen, J. S. Moodera, and J. Shi, Above 400 -K robust perpendicular ferromagnetic phase in a topological insulator , Sci. Adv. 3, e1700307 (2017). [13] M. Mogi et al., Large Anomalous Hall Effect in Topological Insulators with Proximitized Ferromagnetic Insulators , Phys. Rev. Lett. 123, 016804 (2019) . [14] Z. Jiang, F. Katmis, C. Tang, P. Wei, J. S. Moodera, and J. Shi, A comparative transport study of Bi 2Se3 and Bi 2Se3/yttrium iron garnet , Appl. Phys. Lett. 104, 222409 (2014). [15] Z. Jiang, C. -Z. Chang, C. Tang, J. -G. Zheng, J. S. Moodera, and J. Shi, Structural and proximity -induced ferromagnetic properties of topological insulator -magnetic insulator heterostructures , AIP Adv. 6, 055809 (2016). [16] X. Che et al., Proximity -induced magnetic order in a transferred topological insulator thin film on a magnetic insulator , ACS Nano 12, 5042 (2018). [17] S. R. Yang, Y. T. Fanchiang , C. C. Chen, C. C. Tseng, Y. C. Liu, M. X. Guo, M. Hong, S. F. Lee, and J. Kwo, Evidence for exchange Dirac gap in magnetotransport of topological insulator–magnetic insulator heterostructures , Phys. Rev. B 100, 045138 (2019). 8 [18] A. I. Figueroa, F. Bonell, M . G. Cuxart, M. Valvidares, P. Gargiani, G. van der Laan, A. Mugarza, and S. O. Valenzuela, Absence of Magnetic Proximity Effect at the Interface of Bi 2Se3 and (Bi ,Sb) 2Te3 with EuS , Phys. Rev. Lett. 125, 226801 (2020). [19] K. Prokeš et al., Search for enhanced magnetism at the interface between Bi 2Se3 and EuSe , Phys. Rev. B 103, 115438 (2021). [20] L. J. Riddiford et al ., Understanding signatures of emergent magnetism in topological insulator/ferrite bilayers , Phys. Rev. Lett. 128, 126802 (2022). [21] A. A. Taskin, H. F. Legg, F. Yang, S. Sasaki, Y. Kanai, K. Matsumoto, A. Rosch, and Y. Ando, Planar Hall effect from the surface of topological insulators , Nat. Commun. 8, 1340 (2017). [22] P. He, S. S. -L. Zhang, D. Zhu, S. Shi, O. G. Heinonen, G. Vignale, and H. Yang, Nonlinear planar Hall effect , Phys. Rev. Lett. 123, 016801 (2019). [23] S.-H. Zheng, H. -J. Duan, J. -K. Wang, J. -Y. Li, M. -X. Deng, and R. -Q. Wang, Origin of planar Hall effect on the surface of topological insulators: Tilt of Dirac cone by an in -plane magnetic field , Phys. Rev. B 101, 041408(R) (2020). [24] W. Rao, Y. -L. Zhou, Y. -j. Wu, H. -J. Duan, M. -X. Deng, and R. -Q. Wang, Theory for linear and nonlinear planar Hall effect in topological insulator thin films , Phys. Rev. B 103, 155415 (2021). [25] C. C. Chen et al., Topological insulator Bi 2Se3 films on rare earth iron garnets and their high-quality interfaces , Appl. Phys. Lett. 114, 031601 (2019). [26] M. Li, L . Jin, Z . Zhong, X . Tang, Q . Yang, L . Zhang, and H . Zhang , Impact of interfacial chemical state on spin pumping and inverse spin Hall effect in YIG/Pt hybrids , Phys. Rev. B 102, 174435 (2020). [27] M. Liu et al ., Crossover between Weak Antilocalization and Weak Localization in a Magnetically Doped Topological Insulator , Phys. Rev. Lett. 108, 036805 (2012). [28] I. A. Ado, I. A. Dmitriev, P. M. Ostrovsky, and M. Titov, Anomalous Hall effect with massive Dirac fermions , Europhys. Lett. 111, 37004 (2015). [29] S. Hikami, A. I. Larkin, and N. Nagaoka, Spin-orbit interaction and magnetoresistance in the two dimensional random system , Prog. Theor. Phys. 63, 707 (1980). [30] S.-Y. Xu et al., Hedgehog spin texture and Berry's phase tuning in a magnetic topological insulator , Nat. Phys. 8, 616 (2012). [31] K. Yasuda, A. Tsukazaki, R. Yoshimi, K. Kondou , K. S. Takahashi, Y. Otani, M. Kawasaki, and Y. Tokura, Current -nonlinear Hall effect and spin -orbit torque magnetization switching in a magnetic topological insulator , Phys. Rev. Lett. 119, 137204 (2017). [32] F. Katmis et al ., A high -temperature ferromagneti c topological insulating phase by proximity coupling , Nature (London) 533, 513 (2016) . [33] M. Li et al ., Dirac -electron -mediated magnetic proximity effect in topological insulator/magnetic insulator heterostructures , Phys. Rev. B 96, 201301(R) (2017). [34] T. Liu et al ., Changes of Magnetism in a Magnetic Insulator due to Proximity to a Topological Insulator , Phys. Rev. Lett. 125, 017204 (2020). [35] P. He, S. S. L. Zhang, D. Zhu, Y. Liu, Y. Wang, J. Yu, G. Vignale, and H. Yang, Bilinear magnetoelectric resistanc e as a probe of three -dimensional spin texture in topological surface states , Nat. Phys. 14, 495 (2018). 9 [36] I. Sodemann and L. Fu, Quantum nonlinear Hall effect induced by Berry curvature dipole in time -reversal invariant materials , Phys. Rev. Lett. 115, 216806 (2015). [37] Y. Gao, S. A. Yang, and Q. Niu, Field induced positional shift of Bloch electrons and its dynamical implications , Phys. Rev. Lett. 112, 166601 (2014). [38] C. Wang, Y. Gao, and D. Xiao, Intrinsic Nonlinear Hall Effect in Ant iferromagnetic Tetragonal CuMnAs , Phys. Rev. Lett. 127, 277201 (2021). [39] H. Liu, J. Zhao, Y. -X. Huang, W. Wu, X. -L. Sheng, C. Xiao, and S. A. Yang, Intrinsic Second -Order Anomalous Hall Effect and Its Application in Compensated Antiferromagnets , Phys. Rev. Lett. 127, 277202 (2021) . [40] Z. Z. Du, C. M. Wang, H. -Z. Lu, and X. C. Xie, Disorder -induced nonlinear Hall effect with time -reversal symmetry , Nat. Commun. 10, 3047 (2019). [41] S. Nandy and I. Sodemann, Symmetry and quantum kinetics of the nonlinear Hall effect , Phys. Rev. B 100, 195117 (2019). [42] C. Xiao, Z. Z. Du, and Q. Niu, Theory of nonlinear Hall effects: Modified semiclassics from quantum kinetics , Phys. Rev. B 100, 165422 (2019) . [43] K. Hamamoto, M. Ezawa , K. W. Kim, T. Morimoto, and N. Nagaosa, Nonlinear spin current generation in noncentrosymmetric spin -orbit coupled systems , Phys. Rev. B 95, 224430 (2017). [44] D. Rakhmilevich, F. Wang, W. Zhao, M. H. W. Chan, J. S. Moodera, C. Liu, and C. -Z. Chang, Unconventional planar Hall effect in exchange -coupled topological insulator – ferromagnetic insulator heterostructures , Phys. Rev. B 98, 094404 (2018). 10 FIG. 1 XRD and RHEED results of Bi 2Se3(8 QL) films grown on TmIG, YIG and GGG substrates. FIG. 2 (a) Temperature dependence of the longitudinal resistance 𝑅𝑥𝑥 of four Garnet/Bi 2Se3(8 QL) samples. YIG and YIG2 are two films with thicknesses 2.5 𝜇m and 100 nm respectively. The result of YIG2/BS is plotted in the inset due to its large resistance. (b) Hall resistance as a function of OP magnetic field and extracted sheet carrier densities. Inset: Schematic of the measurement setup. (c) Change of sheet conductance under an OP magnetic field for three Garnet/Bi 2Se3(8 QL) devices at 4.5 K. 10 20 30 40 50 60 70Intensity (a.u.) 2q (degree)(006) (009)(012)(015) (018)(021) GGG/Bi2Se3 YIG/Bi2Se3 TmIG/Bi2Se3Garnet (444)Bi2Se3 1QL 8QL 0 100 200 3002.533.54 0 100 200 30010.511Rxx (kW) T (K)YIG2/Bi2Se3Rxx (kW) T (K)Bi2Se3(8 QL) on TmIG YIG GGG(a) (b) -0.5 0 0.5-10-50 DGs (mW) B (T)Bi2Se3 (8 QL) on TmIG YIG GGG 4.5 K (c) -0.1 0 0.1-4-2024Ryx (W) B (T)Garnet/Bi2Se3(8 QL) ns (´1013 cm-2) TmIG 3.46 YIG 3.02 YIG2 2.45 GGG 3.16 11 FIG. 3 The nonlinear Hall resistance 𝑅𝑦𝑥2𝜔=𝑉𝑦2𝜔/𝐼 as a function of x-direction magnetic field for three Bi 2Se3 samples grown on (a) GGG, (b) TmIG , and (c) YIG substrates at various temperatures. The bottom inset in (a) is an optical image of a device with illustrated measurement setup. Scale bar, 100 𝜇m. The top insets in (a) -(c) display the temperature dependence of the slope −𝑑𝑅 𝑦𝑥2𝜔/𝑑𝐵. (a) (b) (c) -2-1012 -2-1012 -0.1 0 0.1-505 4.5 K 6 K 7.5 K 10 K 15 K 20 K 30 K 50 K 150 KR2w yx (mW) 25 K 30 K 50 K 100 K 200 K 4.5 K 6 K 7.5 K 10 K 12.5 K 15 K 20 K 0 100 200010-dR2w yx/dB (mW/T) T (K) 0 100 200 30001020-dR2w yx/dB (mW/T) T (K) B (T) 25 K 30 K 50 K 100 K 200 K 4.5 K 6 K 7.5 K 10 K 12.5 K 15 K 20 K 0.2 mA 0 100 200010-dR2w yx/dB (mW/T) T (K) 12 FIG. 4 (a) The 𝑅𝑦𝑥2𝜔 𝑣𝑠 𝐵 curve (black) of the YIG/BS sample consists of a linear -in-B component (red) with slope 𝑑𝑅 𝑦𝑥2𝜔/𝑑𝐵 and a hysteretic component (blue) with magnitude ∆𝑅𝑦𝑥2𝜔. (b) Temperature dependence of −𝑑𝑅 𝑦𝑥2𝜔/𝑑𝐵 and −∆𝑅𝑦𝑥2𝜔. Inset shows the linear relationship between them. (a) -0.1 0.0 0.1-505R2w yx (mW) B (T) Total Linear-in- B HystereticdR2w yx/dB DR2w yx 0.2 mA (b) 10 10001020 0 10 20012-DR2w yx (mW) -dR2w yx/dB (mW/T) -dR2w yx/dB -DR2w yx T (K)-dR2w yx/dB (mW/T) 012 -DR2w yx (mW)13 FIG. 5 (a) Longitudinal and transverse second -harmonic resistance as a function of IP x- or y-direction magnetic field for the TmIG/Bi 2Se3 sample. Top inset: Current density dependence of the nonlinear Hall resistance −𝑑𝑅 𝑦𝑥2𝜔/𝑑𝐵. Bottom inset: Schematic of the measurement setup. (b) Carrier density dependence of the coefficient 𝛾𝑦=𝑅𝑦𝑥2𝜔 𝑅𝑥𝑥𝐼𝐵 in several devices. The solid l ine fit to 𝑛𝑠−3 is guide for the eye. FIG. 6 Illustration of the hypothetical backscattering enhanced NPHE mechanism. (a) For a TI surface dispersion containing a parabolic term, when there is no IP magnetic field, under a driving electric field, the number of electrons traveling to the right carrying up spins is equal to that of the electrons traveling to the left with down spins. This genera tes a second -order spin current 𝑗s(2) but there is no charge current. (b) An IP magnetic field tilts the upright Dirac cone and causes an imbalance between the left - and right -moving electrons, resulting in a second -order charge current, i.e., the NPHE. (c) When the TI is further coupled with a magnetic material, the exchange interaction opens an exchange gap 2∆ and introduces OP spin polarization components to the surface electrons. The restriction on backscattering is lifted, which may enhance the nonlinear spin to charge conversion efficien cy, giving rise to the large NPHE in MI/TI heterostructures. -0.1 0 0.1-4-2024 0 0.1 0.2 0.3010-dR2w yx/dBx (mW/T) I (mA)R2w y(x)x (mW) B (T)R2w xx vs By R2w yx vs Bx R2w yx vs By TmIG/Bi2Se3 4.5 K 2.5 3 3.5 40.020.030.040.05YIG/BS Device1 Device2 Device3 YIG2/BS gy (A-1T-1) ns (´1013 cm-2)TmIG/BS Device1 Device2 Device3(a) (b) (a) (b) (c) B=0 | 2022-03-12 | Interfacing topological insulators (TIs) with magnetic insulators (MIs) has
been widely used to study the interaction between topological surface states
and magnetism. Previous transport studies typically interpret the suppression
of weak antilocalization or appearance of the anomalous Hall effect as
signatures of magnetic proximity effect (MPE) imposed to TIs. Here, we report
the observation of nonlinear planar Hall effect (NPHE) in Bi2Se3 films grown on
MI thulium and yttrium iron garnet (TmIG and YIG) substrates, which is an order
of magnitude larger than that in Bi2Se3 grown on nonmagnetic gadolinium gallium
garnet (GGG) substrate. The nonlinear Hall resistance in TmIG/Bi2Se3 depends
linearly on the external magnetic field, while that in YIG/Bi2Se3 exhibits an
extra hysteresis loop around zero field. The magnitude of the NPHE is found to
scale inversely with carrier density. We speculate the observed NPHE is related
to the MPE-induced exchange gap opening and out-of-plane spin textures in the
TI surface states, which may be used as an alternative transport signature of
the MPE in MI/TI heterostructures. | Observation of nonlinear planar Hall effect in magnetic insulator/topological insulator heterostructures | 2203.06293v2 |
arXiv:1506.02902v1 [cond-mat.mtrl-sci] 9 Jun 2015Identification of spin wave modes strongly coupled to a co-ax ial cavity. N. J. Lambert,1J. A. Haigh,2and A. J. Ferguson1 1)Microelectronics Group, Cavendish Laboratory, Universit y of Cambridge, Cambridge, CB3 0HE, UK 2)Hitachi Cambridge Laboratory, Cavendish Laboratory, Univ ersity of Cambridge, Cambridge, CB3 0HE, UK (Dated: 16 October 2018) We demonstrate, at room temperature, the strong coupling of th e fundamental and non-uniform magnetostatic modes of an yttrium iron garnet (YIG) ferrimagnetic sphere to theelectromagnetic modesof aco-axial cavity. The well- defined field profile within the cavity yields a specific coupling strength for each magneto static mode. We experimentally measure the coupling strength for the different mag netostatic modes and, by calculating the expected coupling strengths, are able to ide ntify the modes themselves. 1A magnet may be excited in a uniform mode1,2, where all the constituent moments are precessing in phase, or in non-uniform modes3,4where there is a spatially varying phase difference between the moments. The uniform oscillating field that us ually drives ferromag- netic resonance excites only the uniform mode or higher order mode s with a net dynamic magnetisation. In contrast, if the oscillating field is spatially depende nt, perhaps due to the skin depthinthecaseofa metalferromagnet5,6or bydesign inanelectromagnetic waveguide or cavity7,8, then the modes are excited according to the spatial symmetry of the drive field. Such modes are the standing spin waves, and their propagating cou nterparts are central to the research field of magnonics which introduces the possibility to tr ansfer information over millimeter length scales9,10and perform specific information processing tasks11. Recently there has been a surge of interest in the coupling of magne ts to high quality fac- tor electromagnetic cavities12,13, motivated by the possibility of performing experiments in quantummagnonicswhichmightallowsinglelocalisedmagnonstatestob ecreatedandmea- sured. Sofar, thestrongcouplingregimeofquantumelectrodyna mics hasbeenreached8,14–16 along with demonstrations of magnetically induced transparency14. The strong coupling has been enabled by the high moment density and low magnetic damping17in YIG. Both uni- form and non-uniform modes have shown strong coupling8. The work reported in this Letter has been performed in such a context. We fabricate an easily made cavity (Fig. 1a) with a well-defined non-un iform field specif- ically so that we can couple into the non-uniform excited modes. It is m ade from a short (L= 28 mm) length of 3.5 mm diameter copper semi-rigid coaxial cable cut fl at at each end. These ends are brought into proximity with similarly flat ends in connec torised leads, with a small air gap forming the coupling capacitance. SMA screw connect ors provide mechan- ical stability and allow the size of the air gap, and hence the coupling ca pacitance, to be varied in a controlled way. At one extreme, the coaxial cables can be brought into contact with each other, transforming the cavity back into a transmission lin e. We find that the internal quality factor ( Q) of our cavity is 515, in close agreement with the theoretical value ofQ= 517 calculated from the specified attenuation in the co-axial cable . For the cavity experiments described in this Letter, we tuned the coupling streng ths to be κc/2π= 3.3 MHz, giving a loaded Qof 261, a fundamental frequency of ω0/2π= 3.535 GHz and a total cavity linewidth of (2 κc+κint)/2π= 13.5 MHz. A YIG sphere18of diameter 1 mm is inserted into the cable dielectric at the midpoint 23.50 3.55 3.60-6-4-20S22,S11 (dB) f (GHz)-40-30-20-10(a) (b) (c)xy zBx,yYIG sphere S21S11 S22SMA female thread SMA male nut Coupling capacitance gaps FIG. 1. The cavity and YIG sphere. (a) Diagram and longitudin al cross-section of the cavity. It is made from 3.5 mm diameter (UT141) semirigid coaxial cable, a nd the gap capacitances controlled with SMA coupling threads. (b) |S21|,|S11|and|S12|for the cavity configuration used in this experiment. (c) Non-uniform magnetic field around the YIG sp here due to the alternating cavity drive. The global field is applied in the zdirection. of the cavity (Fig. 1c). A key feature of our cavity is the well define d and non-uniform magnetic field profile in the dielectric gap, which has a 1 /rform in the radial direction. This non-uniform field allows the cavity to couple to both uniform and n on-uniform spin- wave modes. We measure the transmission, S21, of the system using a vector ne twork analyser. The incident power on the cavity is -10 dBm; the driven FMR in this regime is lin ear, as ob- served by the independence of S21 on power. We sweep the freque ncy from 2 GHz to 8 GHz, encompassing both the fundamental mode and the second ha rmonic of the cavity. A magnetic field is applied parallel to the cavity, and is varied between 50 and 330 mT. In this field range the magnetization of the YIG is fully saturated. The transmission of the system is shown in Fig. 2. In Fig. 2a we show d|S21|/dHfor the case in which the coupling capacitors are shorted; this is theref ore simply transmission line FMR. The magnetostatic band can be clearly seen, comprising a mu ltitude of modes. Unambiguous identification of each one is not trivial; the intensity of e ach line depends on 3|S21| (dB) 78 6 5 4 3 78 6 5 4 3 2Frequency (GHz) 50 100 150 200 250 300-20 -40 -60 -80 -100 -120 -140 d|S21| dH 01 0.75 0.5 0.25 Applied field (mT)(a) (b)(a.u.) (2,1) (1,1) FIG. 2. Transmission of the system. (a) Derivative of cavity transmission amplitude, d|S21|/dH, with both coupling capacitors shorted; it acts as a 50 Ω trans mission line. Many magnetostatic modes are visible. (b) Transmission amplitude |S21|of the cavity with both coupling capacitances setto≈28fF.Anticrossingsbetweencavity modesandmagnetostati cmodesareseen. Thecoupling depends strongly upon which magnetostatic mode is being exc ited. The anticrossing between the (2,1) mode and the second cavity harmonic is labelled. both the coupling of the magnetostatic mode to the transmission line , and the damping of that mode19, and the linewidth is also dependent on the measurement method20. In Fig 2b we revert to the gap coupled cavity as earlier described. An ticrossings between magnetostatic modes and the cavity resonances at both 3 .53 GHz and 7 .12 GHz are seen, with a maximum coupling strength of 130 MHz for the uniform FMR mode and the funda- mental cavity frequency. Coupling to the second harmonic of the c avity is in general much weaker, as the sphere is positioned at a magnetic field node of this ca vity mode. The spatial form and resonant frequencies of modes in magnetized spheres is well known3,4. Following Walker3we label them with indices nandm21. The radial form of the mode is characterized by n, andmdetermines the number of lobes in the mode pattern. 4The coupling of the ( n,m) mode to the cavity is given by16 gj=ηn,m 2γ/radicalbigg /planckover2pi1ωcµ0ǫr Vc√ 2Ns. Hereωris the resonance frequency, Vcis the volume of the cavity mode, Nis the total number of spins in the YIG sphere, s= 5/2 is the spin per site, µ0is the permeability of free space and ǫris the relative permitivity of the dielectric within the co-axial cable. Th e overlap between the cavity mode and the sphere mode ( n,m) is described by ηn,m, which is given by ηn,m=/vextendsingle/vextendsingle/vextendsingle/vextendsingle1 HmaxMmaxVs×/integraldisplay sphere(H·M)dV/vextendsingle/vextendsingle/vextendsingle/vextendsingle. His the r.f. driving field, and Mis the complex time-dependent off zaxis sphere mag- netization for mode ( n,m).HmaxandMmaxare the maximum magnitudes of these, and Vs is the sphere volume. The coupling strength is independent of magne tostatic damping. The coupling to a particular FMR mode is dependent on the relative sym metries of the mode and the r.f. drive field. It is forced to zero if the mode is antisym metric with respect to thedrive. Inparticular, forthecoupling tothefundamental cavit y modetobesignificant the FMR mode must be symmetric and low-order in z(as the cavity mode is also symmetric). This condition is only met by modes for which n=m. In contrast, in order to couple to the second harmonic cavity mode, the mode must be antisymmetric abou tz= 0. We tabulate calculated coupling constants larger than 1 MHz in Table I. In order to compare these values to our measurement we model th e transmission of strongly coupled cavity using the input-output formalism14–16,22. Close to the fundamental mode of the cavity S21 = κc i(ω−ωc)−1 2(2κc+κint)+/summationtext j|gj|2 −1 2γj+i(ω−ωj), wherejruns over the magnetostatic modes and γjare the FMR linewidths. In Fig. 3 we examine the region around the uniform mode’s anticrossing with th e cavity fundamental 5110 120 130 140 Applied field (m T)(a) (b) -20 -40 -60 -80 -100 -120 -140|S21| (dB)Frequency (GHz)3.23.43.63.84.0 3.23.43.63.84.0(1,1)(2,2) (3,3) FIG. 3. Strong coupling betwen cavity and FMR modes. (a) The r egion around the anticrossing of the uniform mode and the fundamental mode of the cavity. Th e most strongly coupled modes are labelled. (b) Simulation of the same region using the inp ut-output formalism. more closely. In Fig. 3a we show the measured transmission, and in Fig . 3b show the calcu- lated transmission over the same range. For m=nmodes the two are in good agreement. We attribute the appearance of additional weakly coupled modes to the YIG sphere being slightly off-center in the cavity, which lifts the symmetry conditions d escribed above. This also accounts for the weak coupling of the uniform mode to the seco nd harmonic of the cavity. In conclusion, we have described a simple tunable cavity-spin ensemb le system which can nevertheless achieve the strong coupling limit due to the high spin den sity in ferrimagnetic YIG. We show that the coupling to the uniform mode is 130 MHz, giving a cooperativity ofC=g2/κγ≈200. Furthermore, the asymmetric but well defined field profile in th e cavity permitsaquantitative understanding ofthecoupling tohighe r orderspinwave modes. Coupling between microwave cavities andhighlytunablemagnonicexcit ations isa candidate building block for hybrid quantum systems, and the ability to selective ly excite specific spin wave modes offers intriguing possibilities in the emerging field of quantu m magnonics. 6TABLE I. Calculated coupling strengths of selected FMR mode s to the fundamental and second harmonic cavity resonances. g/2π(MHz) n m Fundamental Second harmonic 1 1 130 0 2 1 0 2.9 2 2 27.1 0 3 3 8.1 0 4 4 2.8 0 5 5 1.1 0 We would like to acknowledge support from Hitachi Cambridge Labora tory, and EPSRC Grant No. EP/K027018/1. A.J.F. is supported by a Hitachi Researc h fellowship. REFERENCES 1C. Kittel, “Interpretation of Anomalous Larmor Frequencies in Fer romagnetic Resonance Experiment,” Physical Review 71, 270–271 (1947). 2C. Kittel, “On the Theory of Ferromagnetic Resonance Absorption ,” Physical Review 73, 155–161 (1948). 3L. R. Walker, “Magnetostatic Modes in Ferromagnetic Resonance,” Physical Review 105, 390–399 (1957). 4P. C. Fletcher and R. O. Bell, “Ferrimagnetic Resonance Modes in Sph eres,” Journal of Applied Physics 30, 687 (1959). 5C. Kittel, “Excitation of Spin Waves in a Ferromagnet by a Uniform rf F ield,” Physical Review110, 1295–1297 (1958). 6M. H. Seavey andP. E. Tannenwald, “Direct observationof spin-wa ve resonance,” Physical Review1, 168–169 (1958). 7Y. V. Khivintsev, L. Reisman, J. Lovejoy, R. Adam, C. M. Schneider , R. E. Camley, and Z. J. Celinski, “Spin wave resonance excitation in ferromagnetic films using planar waveguide structures,” Journal of Applied Physics 108, 023907 (2010). 78M. Goryachev, W. G. Farr, D. L. Creedon, Y. Fan, M. Kostylev, an d M. E. Tobar, “High-Cooperativity Cavity QED with Magnons at Microwave F requencies,” Physical Review Applied 2, 054002 (2014). 9J. R. Eshbach, “Spin-wave propogationandthe magnetoelastic int eraction inYttriumIron Garnet,” Physical Review Letters 8, 357–359 (1962). 10M. Tsoi, A. G. M. Jansen, J. Bass, W. Chiang, V. Tsoi, and P. Wyder, “Generation and detection of phase-coherent current-driven magnons in magetic multilayers,” Nature 406, 46–48 (2000). 11A. V. Chumak, A. A. Serga, and B. Hillebrands, “Magnon transistor for all-magnon data processing.” Nature Communications 5, 4700 (2014). 12O. O. Soykal and M. E. Flatt´ e, “Strong Field Interactions betwee n a Nanomagnet and a Photonic Cavity,” Physical Review Letters 104, 077202 (2010). 13O. O. Soykal and M. E. Flatt´ e, “Size dependence of strong couplin g between nanomagnets and photonic cavities,” Physical Review B 82, 104413 (2010). 14X. Zhang, C.-L. Zou, L. Jiang, and H. X. Tang, “Strongly Coupled Ma gnons and Cavity Microwave Photons,” Physical Review Letters 113, 156401 (2014). 15H. Huebl, C. W. Zollitsch, J. Lotze, F. Hocke, M. Greifenstein, A. Ma rx, R. Gross, and S. T. B. Goennenwein, “High Cooperativity in Coupled Microwave Reso nator Ferrimag- netic Insulator Hybrids,” Physical Review Letters 111, 127003 (2013). 16Y. Tabuchi, S. Ishino, T. Ishikawa, R. Yamazaki, K. Usami, and Y. Na kamura, “Hybridizing Ferromagnetic Magnons and Microwave Photons in the Q uantum Limit,” Physical Review Letters 113, 083603 (2014). 17E. G.Spencer, R.C. Lecraw, andA.M. Clogston, “Low-temperat ureline-width maximum in Yttrium Iron Garnet,” Physical Review Letters 3, 32–33 (1959). 18Ferrisphere, Inc. 19C. Bilzer, T. Devolder, P. Crozat, C. Chappert, S. Cardoso, and P . P. Freitas, “Vector net- work analyzer ferromagnetic resonance of thin films on coplanar wa veguides: Comparison of different evaluation methods,” Journal of Applied Physics 101, 074505 (2007). 20S. S. Kalarickal, P. Krivosik, M. Wu, C. E. Patton, M. L. Schneider, P . Kabos, T. J. Silva, and J. P. Nibarger, “Ferromagnetic resonance linewidth in metallic th in films: Comparison of measurement methods,” Journal of Applied Physics 99, 093909 (2006). 21For some modes, the resonance equation admits more than one solu tion, which is generally 8labelled with a third index. As none of the modes we explicitly discuss her e have multiple resonances, for simplicity we omit this index. 22A. A. Clerk, M. H. Devoret, S. M. Girvin, F. Marquardt, and R. J. Schoelkopf, “Introduction to quantum noise, measurement, and amplification,” Reviews of Modern Physics 82, 1155–1208 (2010). 9 | 2015-06-09 | We demonstrate, at room temperature, the strong coupling of the fundamental
and non-uniform magnetostatic modes of an yttrium iron garnet (YIG)
ferrimagnetic sphere to the electromagnetic modes of a co-axial cavity. The
well-defined field profile within the cavity yields a specific coupling
strength for each magnetostatic mode. We experimentally measure the coupling
strength for the different magnetostatic modes and, by calculating the expected
coupling strengths, are able to identify the modes themselves. | Identification of spin wave modes strongly coupled to a co-axial cavity | 1506.02902v1 |
Micro magnet location using spin waves Michael Balinskiy and Alexander Khitun* Electrical Engineering Department, University of California - Riverside, Riverside, CA, USA, 92521 Abstract. In this work, we present experimental data demonstrating the feasibility of magnetic object location using spin wave s. The test structure includes a Y3Fe2(FeO 4)3) (YIG) film with four micro -antennas placed on the edges . A constant in -plane bias magnetic field is provided by NdFeB permanent magnet . Two antennas are used for spin wave excitation while the other two are used for the inductive voltage measurement. There are nine selected places for the magnet on the film. The magnet was subsequentl y placed in all nine positions and spin wave transmission and reflection were measured. The obtained experimental data show the difference in the output signal amplitude depending on the magnet position. All nine locations can be identified by the frequenc y and the amplitude of the absolute minimum in the output power. All experiments are accomplished at room temperature. Potentially, spin waves can be utilized for remote magnetic bit read -out. The disadvantages and physical constraints of this approach are also discussed. I. Introduction The ability to locate objects using transmitted or reflected waves is widely used in d ifferent technologies 1-3. The position is related to the amplitude/frequency of the reflected/ transmitted wave. It provides a convenient tool for non -contact object location. It may be possible to apply s imilar techniques to locate magnetic objects using spin waves. Spin wave is a collective oscillation of spins in a spin lattice around the direct ion of magnetization. Spin waves appear in magnetically ordered structures, and a quantum of spin wave is called a “magnon”. The collective nature of spin wave phenomena manifests itself in relatively long coherence length, which may be order of the tens o f micrometers in conducting ferromagnetic materials (e.g. Ni 81Fe19) and exceed millimeters in non -conducting ferrites (e.g. YIG) at room temperature 4. The latter makes it possible t o build magnonic interferometers exploiting spin waves within the coherence length. For example, a Mach –Zehnder -type spin wave interferometer based on YIG structure was demonstrated in 2005 by M. Kostylev et al. 5. The phase difference among the interfering spin waves was controlled by the magnetic field produced by an electric current flowing through a conducting wire under one of the arms. In general, spin wave dispersion depends on the strength and direction of the bias magnetic field 6. Even a relatively weak (e.g. tens of Oersted) magnetic field produced by a micro -scale magnet placed on the top of magnonic waveguide may result in a prominent phase shift /amplitude change 7. This phenomenon was utilized for magnonic holographic imaging of magnetic microstructures 8. Here, we consider the feasibility of magnetic object location using spin waves. The rest of the paper is organized as follows. In the next Section II, we describe the experimental setup and present experimental data. The Discussion and Conclusions are given in Sections III and IV, respectively. II. Experimental data The schematics of the experimental setup are shown in Fig.1. The cross -section of the device under study is shown in Fig. 1(a). It consists of a permanent magnet made of NdFeB, a Printed Circuit Board (PCB) substrate with four short -circuited antennas, a ferrite film made of YIG, and a micro magnet that can be placed on different parts on top of the film. The permanent magnet is aimed to create a constant bias magnetic field. This bias magnetic field defines the frequency window as well as the type of spin waves that can propagate in the ferrite film. The strength of the field depends on the type of perman ent magnet and the substrate thickness. The bias field is about 195 Oe and directed in -plane on the film surface. The photo of the PCB substrate with four antennas is shown in Fig. 1(b). Each antenna is 6 mm long and 150 μm wide. The antennas are marked as 1,2,3, and 4 in the figure. The four antennas are placed on the side of a virtual square with a length of 11.4 mm. These antennas are aimed to excite and detect spin waves. The antennas are connected to the Programmable Network Analyzer (Keysight N5241A). The details of the spin -wave measurements with micro antennas can be found elsewhere 9,10. The ferrite film is made of YIG. YIG was chosen due to the low spin wave damping. The thickness of the film is 31.5 μm. The saturation magnetization is close to 1750 G, the dissipation parameter ΔH = 0.6 Oe. The planar dimensions of YIG -film are significantly larger than a virtual s quare of antennas providing the total coverage of all microstrip antennas. The schematics in Fig.1(c) show the top view of the YIF film. It is shown a frame of reference , where one division corresponds to 0.75 mm. The red circles depict the nine possible p ositions of the magnet. Hereafter, the position of the magnet will be referred to this reference frame. The magnet has a disk shape whose diameter is about 1 mm and the thickness is 0.3 mm . The magnet is made of magnetic steel . The first set of experiments is aimed to confirm spin -wave propagation through the film. These experiments are accomplished without a magnet on the top of the film. The collection of data showing S21 parameter measured with PNA is presented in Fig.2. In Fi g.2(a), there are shown data for the case when the signal is excited by antenna #1 and detected by antenna #3 (see Fig.1(c)). The data are taken in the frequency range from 1.3 GHz to 2 .9 GHz. The spectrum reveals Magnetostatic Surface Spin Wave (MSSW) propagation perpendicular to the direction of the bias magnetic field. In Fig. 2(b), there is shown S21 parameter measured at PNA for the frequency range from 1. 4 GHz to 2. 2 GHz. The signal is excited by antenna # 2 and detected by antenna # 4. The spectrum reveals Backward Volume Magnetostatic Spin Wave ( BVMSW) propagation directed along the bias magnetic field. In Fig.1(c), there are shown experimental data when two input antennas #1 and #3 operate simul taneously. The output signal is detected at antennas # 2 and #4. The detected inductive voltage is a superposition of the two signal s transmitted by MSSW and BVMSW. This configuration with two working input antennas is the most preferable for magnet location as it allows us to see the difference in spin wave p ropagation along the X and Y axes at a same time. Next, the experiments with two operating input antennas (i.e., #1 and #2) and two output antennas (i.e., #3 and #4) were repeated for the different position s of the magnet. The magnet was sequentially placed in the nine positions marked with the red circles in Fig.1(c). The measurements were accomplished in the frequency range from 1. 5 GHz to 3. 0 GHz. In Fig.3, there are shown data obtained after the subtraction (i.e., signal without magnet) and filtering. The graph shows the change of the output transmitted power ΔP (i.e., measured by the two output antennas) as a function of frequency. There are nine curves of different color s that correspond to the nine magnet locations. The data reveal a prominent variation in the output power depending on the magnet position. In Fig.4., there are shown data on spin wave signal transmission for the four selected cases corresponding to four selected positions of the magnet. The position of the magnet is shown in the inset . Fig.4(a) shows the normalized output power (i.e., after the subtraction) for the case when the magnet is placed in the center of the film. Figs. 4(b) and 4(c) show the output power for the magnet located on the corners of the film. In Fig.4(d), there are shown data for magnet shifted from the center towards the excitation antenna #1. The minimum of the signal transmission appears at different frequencies and reach es different amplitudes . That is the key result that allows us to conclude on the magnet position by the results of spin wave transport measurements. The summary of the experimental data obtained for different magnet locations is shown In Table I . The first two column s contain data on the magnet location while the last two columns show the frequency of the signal minimum and the normalized minimum amplitude. As one can see, there is a unique combination of frequency/minimum amplitude for each of the nine positions. The accuracy of the frequency measurements is 1 MHz . The accuracy of the output power measurements is 4.5 pW. All measurements are done at room temperature. It is also interesting to investigate the change in the signal reflection (i.e. , the S11 parameter) depending on the position of the magnet on the film. In Fig. 5, there are shown data on signal reflection. The reflected signal is measured by antenna #1 and antenna #2. The change in the reflected power is shown after the subtracti on (i.e., signal reflection without magnet) and filtering. The graph shows the change of the reflected power ΔP (i.e., measured by the two antennas) as a function of frequency. There are nine curves of different colors that correspond to the nine magnet lo cations. The re is a lso a difference in the reflection where the minimum of the reflected power occurs at different frequencies and reaches different minimum values. There is a less difference in the frequency compared to the transmission signal. The diffe rence in the amplitude of the reflected signal is prominent for the different locations of the magnet. In Fig. 6., there are shown data on spin wave signal reflection for the four selected cases corresponding to four selected positions of the magnet. The position of the magnet is shown in the inset. Fig. 6(a) shows the normalized reflected power (i.e., after the subtraction) for the case when the magnet is placed in the center of the film. Figs. 4(b) and 4(c) show the reflected power for the magnet located on the corners of the film. In Fig.4(d), there are shown data for magnet shifted from the center towards the excitation antenna #1. The summary of the experimental dat a obtained for the reflected signal is shown In Table II. The first two columns contain data on the magnet location while the last two columns show the frequency of the signal minimum and the normalized minimum amplitude. There are two prominent minima in the reflected power that occur for several magnet positions (e.g., (2,0), (2,1), (1,2), and (2,2). The minima in the reflected power appear on the same frequencies for the different magnet locations. Overall, the reflected spectra are less informative for the magnet location compared to the ones obtained for the transmitted signal. III. Discussion There several observations we want to make based on the obtained experimental data. (i) The output power spectra for transmitted signal are not symmetric for sym metric position of the magnet. For instance, one can see in Table I that the minimum of the output power differs in frequency and amplitude for magnet placed in different corners. The movement of the magnet along the X or Y axes results in the different ou tput power depending the direction of motion. This fact ca n be attributed to the asymmetry of the spin wave diffraction on the magnet and different dispersion of MSSW and BVMSW. The using of the same types of spin waves (e.g., only MSSW or only BVMSW) woul d smash the difference in the output characteristics. The calculation of the spin wave intensity profile over the film is a quite complicated computational task that goes beyond the scope of this work. The main focus of this work is the feasibility of magn et location via spin waves. (ii) The difference in the output power is quite prominent in the range of tens or a hundred of pW. It may be possible to recognize hundreds of magnet locations only by the amplitude of the output signal. The difference in the f requency of the output is also prominent, which provides an additional degree of freedom for magnet location. (iii) The exper iments were accomplished on a relatively large template (e.g., the area of the film with four antennas is about 1 cm2, the size of the magnet is about 1mm2). These millimeter -scale dimensions are mainly defined by the wavelength of the spin waves. It is estimated that the wavelength of MSSW is about 0.5 mm . The wavelength of BVMSW is about 0.5 mm as well . These large dim ensions are possible due to the long coherence length of spin waves in YIG. There is a lot of room for scaling down and increasing the number of possible magnet positions. The scaling down will require the reduction of the spin wave wavelength to microme ter range. In this work , the wavelength of the spin waves is mainly defined by the thickness of the YIG -film and the size of the micro -antennas. There should be a different mechanism for micrometer wavelength spin wave generation. For example, spin waves c an be excited and detected by synthetic multiferroics 11. However, this technique remains mainly unexplored. The ability to search fo r a number of possible magnet positions is the most appealing property of the described approach to magnet location using spin waves. In contrast to the existing technologies based on magnetoresistance measurements, i t does not require any physical contact between the magnet and the sensing element. Overall, it may provide a fundamental advantage over the existing practices for magnetic bit addressing and read -out. The spin wave location technique may be further extended by increasing the number of input/o utput ports 12 or/and exploiting spin wave interference 13. It would be of great interest to validate the possibility to identify multiple magnet configurations (i.e., configuration of several magnets on selected locations) . That would significantly enhance the read -out information capacity and lead to a new clas s of magnetic memory. There are several physical limitations and constraints inherent in the spin wave approach. The recognition of the magnet position requires the scan over a frequency range to find the location of the absolute minimum. It complicates th e whole search procedure and requires additional resources for input frequency modulation. The accuracy of output power measurements is another physical constraint that limits the number of possible magnet locations or magnets configuration to be recognize d. The physical origin of the prominent change in the signal transmission/reflection depending on the position of the magnet is not well understood. There are multiple factors that affect spin wave propagation (e.g., non - uniformity of the bias magnetic fi eld, non -uniformity of the magnetic field produced by the magnet, etc). The position of the magnet may also affect the generation of spin waves by the input antennas. One of the critical concerns is related to the scalability of the proposed approach. On t he one hand, quite a large propagation length of spin waves (i.e., up to 1 cm) at room temperature in YIG serves as the base for further device scaling. On the other hand, it is not clear if the difference in the signal transmission will be still recogniza ble for nanometer scale magnets. IV. Conclusions We present experimental data showing the change of spin wave transmitted and reflected signal depending on the magnet position on the film. Overall, the data show a prominent variation in the frequency an d the amplitude of the signal depending on the magnet position. It is possible to conclude on the location of the magnet (i.e., one of the nine pre -selected positions) based on the spin wave measurements. All experiments are accomplished at room temperatur e. It demonstrates the practical feasibility of using spin wave for magnetic object location. It may be utilized for magnetic bit addressing and read -out. The physical origin of the prominent signal modulation is not clear. There are multiple factors affe cting spin wave propagation/generation/detection which need further investigation. The experiments are accomplished on a relatively large template with millimeter -sized antennas. The main practical challenge toward nanometer magnet location is associated w ith a short -wavelength spin wave generation and detection. Author Contributions M.B. carried out the experiments. A.K. conceived the idea of magnet location using spin wave and wrote the manuscript . All authors discussed the data and the results and co ntributed to the manuscript preparation. Competing financial interests The authors declare no competing financial interests. Data availability All data generated or analyzed during this study are included in this published article . Acknowledgment This work was supported by the National Science Foundation (NSF) under Award # 2006290. Figure Captions Figure 1. (a) Schematics of the test structure. It consists of a permanent magnet made of NdFeB , a Printed Circuit Board (PCB) substrate with four short -circuited antennas, a ferrite film made of YIG, and a micro magnet that can be placed on different parts on top of the film. (b) The photo of the PCB substrate with four antennas. Each antenna is 6 mm long and 150 μm wide. The antennas are marked as 1,2,3, and 4 in the figure. The antennas are connected to the Programmable Network Analyzer (Keysight N5241A . (c) The top view of the YIF film. It is shown a frame of reference, where one division corres ponds to 0.75 mm. The red circles depict the nine possible positions of the magnet. The magnet has a disc oidal shape whose diameter is about 1 mm and the thickness is 0.3 mm. The magnet is made of magnetic steel. Figure 2. The collection of data showing S 21 parameter measured with PNA without a magnet. (a) Experimental data for the case when the signal was is excited by antenna #1 and detected by antenna #3. The data are taken in the frequency range from 1.3 GHz to 2.4 GHz. (b) Data obtained for the case w hen the signal is excited by antenna #2 and detected by antenna #4. Data are collected in the frequency range from 1.4 GHz to 2.2 GHz. (c) Experimental data for the case when two input antennas #1 and #3 operate simultaneously. The output signal is detecte d at antennas #2 and #4. Figure 3. Experimental data on spin wave transmission collected for nine positions of the magnet on the top of YIG film. There are nine curves of different colors that correspond to the nine magnet locations. The measurements were accomplished in the frequency range from 1.5 GHz to 3.0 GHz. The data are after the subtraction (i.e., signal without magnet) and filtering. Figure 4. Experimental data on spin wave signal transmission for the four selected positions of the magnet. The position of the magnet is shown in the inset. (a) data for magnet placed in the center of the film ; (b) data for magnet placed in the left top corner; (c) magnet is placed in the right down corner; (d) magnet is shifted from the center towards antenna #1. Table I. Summary of the experimental data showing the frequency and the amplitude of the absolute minimum of the transmitted signal for nine magnet locations. The first two columns contain data on the magnet location while the last two columns show the frequency of the signal minimum and the normalized minimum amplitude. Figure 5. Experimental data on spin wave reflection collected for nine positions of the magnet on the top of YIG film. There are nine curves of different colors that correspond to t he nine magnet locations. The measurements were accomplished in the frequency range from 1.5 GHz to 3.0 GHz. The data are after the subtraction (i.e., signal without magnet) and filtering. Figure 6. Experimental data on spin wave signal reflection for th e four selected positions of the magnet. The position of the magnet is shown in the inset. (a) data for magnet placed in the center of the film; (b) data for magnet placed in the left top corner; (c) magnet is placed in the right down corner; (d) magnet is shifted from the center towards antenna #1. Table I I. Summary of the experimental data showing the frequency and the amplitude of the absolute minimum of the reflected signal for nine magnet locations . The first two columns contain data on the magnet lo cation while the last two columns show the frequency of the signal minimum and the normalized minimum amplitude. Figure 1 Figure 2 Figure 3 1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0-0.1896-0.1580-0.1264-0.0948-0.0632-0.03160.00000.0316 DP [nW] (2,2) (1,2) (0,2) (2,1) (1,1) (0,1) (2,0) (1,0) (0,0) Frequency [GHz] Figure 4 Table I Figure 5 1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0-0.72-0.60-0.48-0.36-0.24-0.120.000.12 (2,2) (1,2) (0,2) (2,1) (1,1) (0,1) (2,0) (1,0) (0,0) DP [nW] Frequency [GHz] Figure 6 Table II References 1 Born, W. T. A REVIEW OF GEOPHYSICAL INSTRUMENTATION. GEOPHYSICS 25, 77-91, doi:10.1190/1.1438705 (1960). 2 Sarvazyan, A. P., Urban, M. W. & Greenleaf, J. F. Acoustic Waves in Medical Imaging and Diagnostics. Ultrasound in Med icine & Biology 39, 1133 -1146, doi:https://doi.org/10.1016/j.ultrasmedbio.2013.02.006 (2013). 3 Zhukov, V. Y. & Shchukin, G. G. Current problems of meteorological radiolocation. Journal of Communications Technology and Electronics 61, 1069 -1080, doi:10.1134/S1064226916100235 (2016). 4 Serga, A. A., Chumak, A. V. & Hillebrands, B. YIG MAgnonics. Journal of Physics D: Applied Physics 43 (2010). 5 Kostylev, M. P., Serga, A. A., Schneider, T., L even, B. & Hillebrands, B. Spin -wave logical gates. Applied Physics Letters 87, 153501 -153501 -153503 (2005). 6 Gurevich, A. G. & Melkov, G. A. (CRC press, 1996). 7 Gertz, F., Kozhevnikov, A., Filimonov Y., Nikonov, D. E. & Khitun, A. Magnonic Holograph ic Memory: From Proposal to Device. IEEE Journal on Exploratory Solid -State Computational Devices and Circuits 1, 67-75 (2015). 8 Gutierrez, D. et al. Magnonic holographic imaging of magnetic microstructures. Journal of Magnetism and Magnetic Materials 428, 348 -356, doi:10.1016/j.jmmm.2016.12.022 (2017). 9 Khivintsev, Y. et al. Prime factorization using magnonic holographic devices. Journal of Applied Physics 120, doi:10.1063/1.4962740 (2016). 10 Balinskiy, M. et al. Spin wave interference in YIG cross junc tion. Aip Advances 7, doi:10.1063/1.4974526 (2017). 11 Cherepov, S. et al. Electric -field -induced spin wave generation using multiferroic magnetoelectric cells. Applied Physics Letters 104, doi:10.1063/1.4865916 (2014). 12 Balynsky, M. et al. Quantum compu ting without quantum computers: Database search and data processing using classical wave superposition. Journal of Applied Physics 130, 164903, doi:10.1063/5.0068316 (2021). 13 Balinskiy, M., Chiang, H., Gutierrez, D. & Khitun, A. Spin wave interference detection via inverse spin Hall effect. Applied Physics Letters 118, 242402, doi:10.1063/5.0055402 (2021). | 2022-04-14 | In this work, we present experimental data demonstrating the feasibility of
magnetic object location using spin waves. The test structure includes a
Y$_3$Fe$_2$(FeO$_4$)$_3$) (YIG) film with four micro-antennas placed on the
edges. A constant in-plane bias magnetic field is provided by NdFeB permanent
magnet. Two antennas are used for spin wave excitation while the other two are
used for the inductive voltage measurement. There are nine selected places for
the magnet on the film. The magnet was subsequently placed in all nine
positions and spin wave transmission and reflection were measured. The obtained
experimental data show the difference in the output signal amplitude depending
on the magnet position. All nine locations can be identified by the frequency
and the amplitude of the absolute minimum in the output power. All experiments
are accomplished at room temperature. Potentially, spin waves can be utilized
for remote magnetic bit read-out. The disadvantages and physical constraints of
this approach are also discussed. | Micro magnet location using spin waves | 2204.07238v1 |
arXiv:1502.05244v1 [cond-mat.mtrl-sci] 13 Feb 2015Applied Physics Letters Spin-current injection and detection in strongly correlat ed organic conductor Z. Qiu∗,1,2M. Uruichi,3D. Hou,1,2K. Uchida,2,4,5H. M. Yamamoto,6,7and E. Saitoh1,2,4,8 1WPI Advanced Institute for Materials Research, Tohoku University, Sendai 980-8577, Japan 2Spin Quantum Rectification Project, ERATO, Japan Science and Technology Agency, Aoba-ku, Sendai 980-8 577, Japan 3Institute for Molecular Science, Okazaki 444-8585, Japan. 4Institute for Materials Research, Tohoku University, Send ai 980-8577, Japan 5PRESTO, Japan Science and Technology Agency, Saitama 332-0 012, Japan 6Research Center of Integrative Molecular Systems (CIMoS), Institute for Molecular Science, 38 Nishigounaka, Myodaiji, Okazaki 444-8585, Japan. 7RIKEN , 2-1 Hirosawa, Wako 351-0198, Japan. 8Advanced Science Research Center, Japan Atomic Energy Agency, Tokai 319-1195, Japan ∗Author to whom correspondence should be addressed; electronic mail: qiuzy@imr.tohoku.ac.jp 1Abstract Spin-current injection into an organic semiconductor κ-(BEDT-TTF) 2Cu[N(CN) 2]Br film in- duced by the spin pumping from an yttrium iron garnet (YIG) fil m. When magnetiza- tion dynamics in the YIG film is excited by ferromagnetic or sp in-wave resonance, a volt- age signal was found to appear in the κ-(BEDT-TTF) 2Cu[N(CN) 2]Br film. Magnetic-field- angle dependence measurements indicate that the voltage si gnal is governed by the inverse spin Hall effect in κ-(BEDT-TTF) 2Cu[N(CN) 2]Br. We found that the voltage signal in the κ-(BEDT-TTF) 2Cu[N(CN) 2]Br/YIG system is critically suppressed around 80 K, around which magnetic and/or glass transitions occur, implying that the efficiency of the spin-current injection is suppressed by fluctuations which critically enhanced nea r the transitions. PACS numbers: 72.80.Le, 85.75.-d, 72.25.Pn Keywords: Organic semiconductor, spintronics, inverse spin Hall e ffect, spin pumping 2Thefieldofspintronics hasattractedgreatinterest inthelastdec adebecauseofanimpact on the next generation magnetic memories and computing devices, w here the carrier spins play a key role in transmitting, processing, and storing information [ 1]. Here, a method for direct conversion of a spin current into an electric signal is indispens able. The spin-charge conversion has mainly relied on the spin-orbit interaction, which caus es a spin current to induce an electric field EISHEperpendicular to both the spin polarization σand the flow directionofthespin current Js:EISHE/ba∇dblJs×σ. This phenomenon isknown astheinverse spin Hall effect (ISHE) [2, 3]. To investigate spin-current physics and r ealize large spin-charge conversion, the ISHE has been measured in various materials rangin g from metals and semiconductors to an organic conjugated polymer [4–11]. A recent study has suggested that conjugated polymers can work as a spin-charge converter [10, 11 ], and further investigation of the ISHE in different organic materials is now necessary. In the present study, we report the observation of the ISHE in an organic molecular semi- conductor κ-(BEDT-TTF) 2Cu[N(CN) 2]Br (called κ-Br). The κ-Br consists of alternating layers of conducting sheets (composed of BEDT-TTF dimers) and in sulating sheets (com- posed of Cu[N(CN) 2]Br anions), which is recognized as an ideal system with anisotropic strongly correlated electrons (Fig. 1(a)). The ground state of b ulkκ-Br is known to be superconducting with the transition temperature Tc≈12 K[12], which becomes antiferro- magnetic and insulating by replacing Cu[N(CN) 2]Br with Cu[N(CN) 2]Cl (Fig. 1(d))[13]. To inject a spin current into the κ-Br, we employed a spin-pumping method by using κ-Br/yttrium iron garnet (YIG) bilayer devices. In the κ-Br/YIG bilayer devices, mag- netization precession motions driven by ferromagnetic resonance (FMR) and/or spin-wave resonance (SWR) in the YIG layer inject a spin current across the in terface into the con- ductingκ-Br layer in the direction perpendicular to the interface[14]. This inje cted spin 3current is converted into an electric field along the κ-Br film plane if κ-Br exhibits ISHE. The preparation process for the κ-Br/YIG bilayer devices is as follows. A single- crystalline YIG film with the thickness of 5 µm was put on a gadolinium gallium garnet wafer by a liquid phase epitaxy method. The YIG film on the substrate was cut into a rect- angular shape with the size of 3 ×1 mm2. Two separated Cu electrodes with the thickness of 50 nm were then deposited near the ends of the YIG film. The dista nce between the two electrodes was 0.4 mm. Here, to completely avoid the spin injection int o the Cu electrodes, 20-nm-thick SiO 2films were inserted between the electrodes and the YIG film, as show n in Fig. 1(b). Finally, a laminated κ-Br single crystal, grown by an electrochemical method [15], was placed on the top of the YIG film between the two Cu electrod es (Fig. 1(b)). We prepared two κ-Br/YIG samples A and B to check reproducibility. The thicknesses o f theκ-Br films for the samples A and B are around 100 nm but a little different from each other, resulting in the difference of the resistance of the κ-Br film (Fig. 1(b)). To observe the ISHE in κ-Br induced by the spin pumping, we measured the Hdependence of the mi- crowave absorption and DC electric voltage between the electrode s at various temperatures with applying a static magnetic field Hand a microwave magnetic field with the frequency of 5 GHz to the device. Figure 1(c) shows the temperature dependence of resistance of theκ-Br films on YIG for the samples A and B. The κ-Br films in the present system exhibit no superconducting transition [12, 16], but do insulator-like behavior similar to a bulk κ-Cl [13, 16]. This result can be ascribed to tensile strain induced by the substrate du e to the different thermal expansion coefficients of κ-Br and YIG. The similar phenomenon was reported in κ-Br on a SrTiO 3substrate[15], of which the thermal expansion coefficient is close to that of YIG (∼10 ppm/K at room temperature[17, 18]). Thus, the ground state o f theκ-Br film on 4YIG is expected to be slightly on the insulator side of the Mott transit ion. The red arrow in Fig. 1(d) schematically indicates a state trajectory of our κ-Br films with decreasing the temperature [15, 16, 19–25]. Figure 2(a) shows the FMR/SWR spectrum dI/dHfor theκ-Br/YIG sample A at 300 K. Here, Idenotes the microwave absorption intensity. The spectrum shows that the mag- netization in the YIG film resonates with the applied microwave around the FMR field HFMR≈1110 Oe. As shown in Fig. 2(b), under the FMR/SWR condition, electr ic voltage with peak structure was observed between the ends of the κ-Br film at θ=±90◦, where θdenotes the angle between the Hdirection and the direction across the electrodes (Fig. 2(b)). The voltage signal disappears when θ= 0◦. Thisθdependence of the peak voltage is consistent with the characteristic of the ISHE induced by the spin pumping. Because the SiO 2film between the Cu electrode and the YIG film blocks the spin-curren t injection across the Cu/YIG interfaces, the observed voltage signal is irre levant to the ISHE in the Cu electrodes. The magnitude of the electric voltage is one or two or ders of magnitude smaller than that in conventional Pt/YIG devices [26–29], but is close to that observed in polymer/YIG devices [10]. To establish the ISHE in the κ-Br/YIG sample exclusively, it is important to separate the spin-pumping-induced signal from thermoelectric voltage induced b y temperature gradients generated by nonreciprocal surface-spin-wave excitation [30], s ince thermoelectric voltage in conductors whose carrier density is low, such as κ-Br, may not be negligibly small. In order to estimate temperature gradient under the FMR/SWR condition, w e excited surface spin waves ina 3-mm-lengthYIG sample by using a microwave of which the po wer is much higher than that used in the present voltage measurements, and measur ed temperature images of the YIG surface with an infrared camera (Figs. 3(a) and (b)). We f ound that a temperature 5gradient is created in the direction perpendicular to the Hdirection around the FMR field and its direction is reversed by reversing H, consistent with the behavior of the spin-wave heat conveyer effect (Fig. 3(c) and (d)) [30]. Figure 3(e) shows th at the magnitude of the temperature gradient is proportional to the absorbed microwave power. This temperature gradient might induce an electric voltage due to the Seebeck effect in κ-Br with the similar symmetry as the ISHE voltage. However, the thermoelectric volta ge is expected to be much smaller than the signal shown in Fig. 2(b); since all the measurement s in this work were carried out with a low microwave-absorption power (marked with a gr een line in Fig. 3(e)), the magnitude of the temperature gradient in the κ-Br/YIG film is less than 0.015 K/mm. Even when we use the Seebeck coefficient of κ-Brat the maximum valuereported inprevious literatures [31–33], the electric voltage due to the Seebeck effect in theκ-Br film is estimated to be less than 0 .01µVat 300 K, where the effective length of κ-Br is 0.4 mm. This is at least one order of magnitude less than the signals observed in our κ-Br/YIG sample. Therefore, we can conclude that the observed electric voltage with the peak st ructure is governed by ISHE. Figure 4 shows the electric voltage spectra in the κ-Br/YIG sample A for various values of temperature, T. Clear voltage signals were observed to appear at the FMR fields whe n T >80 K. We found that the sign of the voltage signals is also reversed wh enHis reversed, which is consistent with the ISHE as discussed above. Surprisingly, h owever, the peak voltage signals at the FMR fields decrease steeply with decreasing Tand merge into noise around 80 K. This anomalous suppression of the voltage signals cann ot be explained by the resistance Rchange of the κ-Br film because no remarkable Rchange was observed in the same temperatures (Fig. 1(c)). At temperatures lower than 60 K, large voltage signals appear around the FMR fields as shown in Fig. 4, but its origin is not con firmed because 6of the big noise and poor reproducibility. Therefore, hereafter we focus on the temperature dependence of the voltage signals above 80 K. In Fig. 5, we plot the Tdependence of V∗/Rfor theκ-Br/YIG samples A and B, where V∗=/parenleftbig VFMR(−H)−VFMR(+H)/parenrightbig /2 withVFMR(±H)being the electric voltage at the FMR fields, to take into account the resistance difference of the κ-Br films. V∗/Rfor both the samples exhibit almost same Tdependence, indicating that the observed voltage suppression is a n intrinsic phenomenon in the κ-Br/YIG samples. Here we discuss a possible origin of the observed temperature depe ndence of the voltage in theκ-Br/YIG systems. ISHE voltage is determined by two factors. One is spin-to- charge conversion efficiency, i.e. the spin-Hall angle, in the κ-Br film. The mechanism of ISHE consists of intrinsic contribution due to spin-orbit coupling in the band structure and extrinsic contribution due to the impurity scattering [34]. In or ganic systems such asκ-Br, the extrinsic contribution seem to govern the ISHE since intrin sic contribution is expected to be weak because of their carbon-based light-element composition. Judging from the predicted rather weak temperature dependence of impurity s cattering, the temperature dependenceofthespin-Hallanglecannotbetheoriginofthesharp suppressionofthevoltage signal in the κ-Br/YIG systems (Fig. 5). The other factor is the spin-current in jection efficiency across the κ-Br/YIG interface, which can be affected by spin susceptibility [35] in κ-Br. Importantly, the temperature dependence of the spin susc eptibility for the κ-X family was shown to exhibit a minimum at temperatures similar to those at whic h the anomalous suppression of the spin-pumping-induced ISHE voltage was observ ed [16], suggesting an importance of the temperature dependence of the spin-current injection efficiency in the κ- Br/YIG systems. We also mention that the temperature at which th e ISHE suppression was observed coincides with a glass transition temperature of κ-Br films [36, 37]. However, at the 7present stage, thereisnoframeworktodiscusstherelationbetw eenthespin-current injection efficiency and such lattice fluctuations. To obtain the full understa nding of the temperature dependence of the spin-pumping-induced ISHE voltage in the κ-Br/YIG systems, more detailed experimental and theoretical studies are necessary. In summary, we have investigated the spin pumping into organic semic onductor κ-(BEDT-TTF) 2Cu[N(CN) 2]Br (κ-Br) films from adjacent yttrium iron garnet (YIG) films. The experimental results show that an electric voltage is generate d in the κ-Br film when ferromagnetic or spin-wave resonance is excited in the YIG film. Sinc e this voltage signal was confirmed to be irrelevant to extrinsic temperature gradients generated by spin-wave excitation and the resultant thermoelectric effects, we attribute it to the inverse spin Hall effect in the κ-Br film. The temperature-dependent measurements reveal tha t the voltage signal in the κ-Br/YIG systems is critically suppressed around 80 K, implying that t his suppression relates with the spin and/or lattice fluctuations in κ-Br. This work was supported by PRESTO “Phase Interfaces for Highly E fficient Energy Utilization”, Strategic International Cooperative Program ASPIM ATT from JST, Japan, Grant-in-Aid for Young Scientists (A) (25707029), Grant-in-Aid f or Young Scientists (B) (26790038), Grant-in-Aid for Challenging Exploratory Research ( 26600067), Grant-in-Aid for Scientific Research (A) (24244051), Grant-in-Aid for Scientifi c Research on Innovative Areas “Nano Spin Conversion Science” (26103005) from MEXT, Jap an, NEC Corporation, and NSFC. 8[1] I. Zutic, J. Fabian, and S. Das Sarma, Reviews of Modern Ph ysics76, 323 (2004). [2] E. Saitoh, M. Ueda, H. Miyajima, and G. Tatara, Applied Ph ysics Letters 88, 182509 (2006). [3] A. Azevedo, L.H. V.Leao, R.L.Rodriguez-Suarez, A. B. Ol iveira, andS.M.Rezende, Journal of Applied Physics 97, 10C715 (2005). [4] Z. Qiu, Y. Kajiwara, K. Ando, Y. Fujikawa, K. Uchida, T. Ta shiro, K. Harii, T. Yoshino, and E. Saitoh, Applied Physics Letters 100, 022402 (2012). [5] K. Ando, M. Morikawa, T. Trypiniotis, Y. Fujikawa, C. H. W . Barnes, and E. Saitoh, Journal of Applied Physics 107, 113902 (2010). [6] Z. Qiu, K. Ando, K. Uchida, Y. Kajiwara, R. Takahashi, H. N akayama, T. An, Y. Fujikawa, and E. Saitoh, Appl. Phys. Lett. 103, 4 (2013). [7] K. Ando, S. Takahashi, J. Ieda, H. Kurebayashi, T. Trypin iotis, C. H. W. Barnes, S. Maekawa, and E. Saitoh, Nature Materials 10, 655 (2011). [8] Z. Qiu, T. An, K. Uchida, D. Hou, Y. Shiomi, Y. Fujikawa, an d E. Saitoh, Appl. Phys. Lett. 103, 3 (2013). [9] Y.Kajiwara, K.Harii, S.Takahashi, J.Ohe, K.Uchida, M. Mizuguchi, H.Umezawa, H.Kawai, K. Ando, K. Takanashi, et al., Nature 464, 262 (2010). [10] K. Ando, S. Watanabe, S. Mooser, E. Saitoh, and H. Sirrin ghaus, Nature Materials 12, 622 (2013). [11] S. Watanabe, K. Ando, K. Kang, S. Mooser, Y. Vaynzof, H. K urebayashi, E. Saitoh, and H. Sirringhaus, Nature Physics 10, 308 (2014). [12] A. M. Kini, U. Geiser, H. H. Wang, K. D. Carlson, J. M. Will iams, W. K. Kwok, K. G. 9Vandervoort, J. E. Thompson, and D. L. Stupka, Inorganic Che mistry29, 2555 (1990). [13] J. M. Williams, A. M. Kini, H. H. Wang, K. D. Carlson, U. Ge iser, L. K. Montgomery, G. J. Pyrka, D. M. Watkins, and J. M. Kommers, Inorganic Chemistry 29, 3272 (1990). [14] C. W. Sandweg, Y. Kajiwara, K. Ando, E. Saitoh, andB. Hil lebrands, AppliedPhysics Letters 97, 252504 (2010). [15] H. M. Yamamoto, M. Nakano, M. Suda, Y. Iwasa, M. Kawasaki , and R. Kato, Nature Com- munications 4, 2379 (2013). [16] K. Kanoda, Journal of the Physical Society of Japan 75, 051007 (2006). [17] S. H. Ling, Y. S. Tang, W. S. Au, and H. K. Wong, Applied Phy sics Letters 62, 1757 (1993). [18] T.Boudiar, B.Payet-Gervy, M.-F. Blanc-Mignon, J.-J. Rousseau, M.LeBerre, andH.Joisten, Journal of Magnetism and Magnetic Materials 284, 77 (2004). [19] K. Miyagawa, K. Kanoda, and A. Kawamoto, Chemical Revie ws104, 5635 (2004). [20] D. Faltermeier, J. Barz, M. Dumm, M. Dressel, N. Drichko , B. Petrov, V. Semkin, R. Vlasova, C. Me´ zi` ere, and P. Batail, Physical Review B 76, 165113 (2007). [21] F. Kagawa, K. Miyagawa, and K. Kanoda, Nature Physics 5, 880 (2009). [22] F. Kagawa, K. Miyagawa, and K. Kanoda, Nature 436, 534 (2005). [23] P. Limelette, P. Wzietek, S. Florens, a. Georges, T. Cos ti, C. Pasquier, D. J´ erome, C. M´ ezi` ere, and P. Batail, Physical Review Letters 91, 016401 (2003). [24] D. Fournier, M. Poirier, M. Castonguay, and K. Truong, P hysical Review Letters 90, 127002 (2003). [25] Y. Kuwata, M. Itaya, and A. Kawamoto, Physical Review B 83, 144505 (2011). [26] R. Iguchi, K. Ando, R. Takahashi, T. An, E. Saitoh, and T. Sato, Japanese Journal of Applied Physics51, 103004 (2012). 10[27] R. Iguchil, K. Ando, T. An, E. Saitoh, and T. Sato, Ieee Tr ansactions on Magnetics 48, 3051 (2012). [28] S. M. Rezende, R. L. Rodr´ ıguez-Su´ arez, M. M. Soares, L . H. Vilela-Le˜ ao, D. L. Dom´ ınguez, and A. Azevedo, Applied Physics Letters 102, 012402 (2013). [29] Z. H. Wang, Y. Y. Sun, Y. Y. Song, M. Z.Wu, H. Schultheiss, J.E. Pearson, andA. Hoffmann, Applied Physics Letters 99, 162511 (2011). [30] T. An, V. I. Vasyuchka, K. Uchida, a. V. Chumak, K. Yamagu chi, K. Harii, J. Ohe, M. B. Jungfleisch, Y. Kajiwara, H. Adachi, et al., Nature material s12, 549 (2013). [31] R. Yu, J. Williams, H. Wang, J. Thompson, A. Kini, K. Carl son, J. Ren, M.-H. Whangbo, and P. Chaikin, Physical Review B 44, 6932 (1991). [32] G. Y. LOGVENOV, H. Ito, T. Ishiguro, G. Saito, J.-i. Yama da, and H. Anzai, Journal of the Physical Society of Japan 64, 3622 (1995). [33] V.Yefanov, S.Kagoshima, M.Tanatar, T.Ishiguro, V.Bo ndarenko, N.Kushch, andE.Yagub- skii, Physica C: Superconductivity 388-389 , 589 (2003). [34] G. Vignale, Journal of Superconductivity and Novel Mag netism23, 3 (2009). [35] Y. Ohnuma, H. Adachi, E. Saitoh, and S. Maekawa, Physica l Review B 89, 174417 (2014). [36] C. Strack, C. Akinci, V. Pashchenko, B. Wolf, E. Uhrig, W . Assmus, M. Lang, J. Schreuer, L. Wiehl, J. A. Schlueter, et al., Physical Review B 72, 54511 (2005). [37] J. M¨ uller, M. Lang, F. Steglich, J. Schlueter, a. Kini, and T. Sasaki, Physical Review B 65, 144521 (2002). 11T (K) Cu[N(CN)2]Br-1 BEDT-TTF+0.5 Cu[N(CN)2]Br-1 BEDT-TTF+0.5 Cu[N(CN)2]Br-1 0.4 mm Tensile Compressive StrainSC PM PI AFIMixed phase10 100κ-Clκ-BrR/R300 K 10 2κ-Cl (bulk) κ-Br (bulk)sample Bsample A 10 100 T (K)SiO2 (c) (d)Cu YIGV κ-Br(a) (b) SS SSS S SSBEDT-TTF H microwave FIG. 1: (a) Structural formula of the BEDT-TTF molecule (upp er panel) and schematic cross- section of the (BEDT-TTF) 2Cu[N(CN) 2]Br (κ-Br) crystal, where cationic BEDT-TTF and anionic Cu[N(CN) 2]Br layers alternate each other (lower panel). (b) Schemati c illustration of the sample structure and experimental setup. Hdenotes the static external maganetic field applied along th e film plane. (c) Temperature dependence of R/R300Kof the two κ-Br/YIG samples A and B, a bulk κ-Br crystal, and a bulk κ-Cl crystal. Here, R(R300K) denotes the resistance between the ends of theκ-Br film at each temperature (at 300 K). (d) Conceptual phase d iagram of κ-X systems. PI, PM, AFI, and SC denote paramagnetic insulator, paramagneti c metal, antiferromagnetic insulator, and superconductor, respectively. The red arrow indicates the trajectory that the κ-Br crystal on the YIG substrate experiences upon cooling. 12800 1200 1600-1 01 -90°θ=90° 0° H (Oe)V (µV) 0dI/dH (a.u.)(a) (b)V H θH FIG. 2: (a) The FMR/SWR spectrum dI/dHof theκ-Br/YIG sample A at 300 K. Here, Iand denotes the microwave absorption intensity. The dashed lin e shows the magnetic field HFMRat which the FMR is excited. (b) The electric voltage Vbetween the ends of the κ-Br film as a function of H. 13-1 0 1-0.20.00.2 Position from center (mm)T-Tcenter (K) -1 0 1 0.050.100.15 0 5 10T (K/mm) Pab (mW)exp. fitting H H 1.5 mm 1.5 mm -0.4 0.0 (K) 0.4 (a) (b) (c) (d) (e) FIG. 3: (a),(b) Temperature distributions of the YIG surfac e near the FMR fields (5 GHz) for the opposite orientations of H, measured with an infrared camera. (c),(d) Temperature pro files of the YIG surface. (e) The microwave-power absorption Pabdependence of the temperature gradient ∇Tof the YIG surface. The voltage measurements were carried ou t with a low Pabvalue (marked with a green line). 141 μV FIG. 4:Hdependence of Vin theκ-Br/YIG sample A for various values of the temperature T. The scales of the longitudinal axis for the data at T≤70 K are shrinked by a factor of 0.1. 100 200 3000.010.101 T (K)Normalized V*/Rsample A sample B FIG. 5: Tdependence of V∗/Rfor the κ-Br/YIG samples A and B. Here, V∗=/parenleftbig VFMR(−H)−VFMR(+H)/parenrightbig /2 withVFMR(±H)being the electric voltage at HFMR. 15 | 2015-02-13 | Spin-current injection into an organic semiconductor
$\rm{\kappa\text{-}(BEDT\text{-}TTF)_2Cu[N(CN)_2]Br}$ film induced by the spin
pumping from an yttrium iron garnet (YIG) film. When magnetization dynamics in
the YIG film is excited by ferromagnetic or spin-wave resonance, a voltage
signal was found to appear in the
$\rm{\kappa\text{-}(BEDT\text{-}TTF)_2Cu[N(CN)_2]Br}$ film.
Magnetic-field-angle dependence measurements indicate that the voltage signal
is governed by the inverse spin Hall effect in
$\rm{\kappa\text{-}(BEDT\text{-}TTF)_2Cu[N(CN)_2]Br}$. We found that the
voltage signal in the $\rm{\kappa\text{-}(BEDT\text{-}TTF)_2Cu[N(CN)_2]Br}$/YIG
system is critically suppressed around 80 K, around which magnetic and/or glass
transitions occur, implying that the efficiency of the spin-current injection
is suppressed by fluctuations which critically enhanced near the transitions. | Spin-current injection and detection in strongly correlated organic conductor | 1502.05244v1 |
Spatial Control of Hybridization-Induced Spin-Wave Transmission Stop Band Franz Vilsmeier∗1, Christian Riedel1, and Christian H. Back1 1Fakult¨ at f¨ ur Physik, Technische Universit¨ at M¨ unchen, Garching, Germany March 23, 2024 Abstract Spin-wave (SW) propagation close to the hybridization-induced transmission stop band is investigated within a trapezoid-shaped 200 nm thick yttrium iron garnet (YIG) film using time-resolved magneto-optic Kerr effect (TR-MOKE) microscopy and broadband spin wave spectroscopy, supported by micromagnetic simulations. The gradual reduc- tion of the effective field within the structure leads to local variations of the SW dispersion relation and results in a SW hybridization at a fixed position in the trapezoid where the propagation vanishes since the SW group velocity approaches zero. By tuning external field or frequency, spatial control of the spatial stop band position and spin-wave propagation is demonstrated and utilized to gain transmission control over several microstrip lines. I Introduction Driven by potential spin-wave-based applications in computing and data processing, the field of magnonics has garnered growing interest in recent years [1–11]. To perform logic operations encoded within magnon currents various approaches were suggested and realized, such as interference-based logic gates [1,10,12,13] or magnonic crystals that ex- ploit the periodicity-induced formation of bandgaps in the spin-wave spectrum [14–21]. These devices rely on precise control and manipulation of spin- waves with wave vector kwithin a material with magnetization M. Recently, a hybridization-induced ∗franz.vilsmeier@tum.despin-wave-transmission stop band was demonstrated in 200 nm yttrium iron garnet (YIG) [22], adding to the list of options for engineering spin-wave propa- gation. It was shown that the hybridization of two different Damon Eshbach-like (DE) ( k⊥M) SW modes causes a frequency- and field-dependent sup- pression of SW propagation in a film with in-plane magnetization. Furthermore, it is well known that at the edges of thin magnetic films, depending on the magnetization direction, the effective field is locally reduced in order to avoid the generation of magnetic surface charges [16, 23, 24]. This allows for shape- modulated local variations of SW propagation. Com- bining this effect with the transmission stop band may provide enhanced control over spin-wave prop- agation dynamics and facilitate the implementation of magnonic devices. In this report, we investigate the effect of the geometry-induced variation of the effective field in a 200 nm YIG film on the hybridization-induced stop band. We demonstrate that the spin-wave propa- gation distance can be actively controlled within a trapezoid-shaped magnetic film as the reduced ef- fective field locally enforces the hybridization con- dition. Experimental dispersion measurements and micromagnetic simulations using TetraX [25] are conducted to determine the full film stop band con- dition. From further micromagnetic simulations us- ingMuMax3 [26], the effective internal field of a trapezoid geometry is determined. An inhomoge- neous field distribution with a gradual decrease along the trapezoid’s length is observed. We experimen- tally investigate the corresponding spin-wave prop- agation within the trapezoid in a DE-like geometry using time-resolved magneto-optic Kerr effect (TR- MOKE) microscopy [27–31] and broadband spin- wave spectroscopy. Bending of wavefronts, the for- 1arXiv:2403.15840v1 [physics.app-ph] 23 Mar 2024(a) (b)Figure 1 Sketch of the experimental setup. (a) Schematic for TR-MOKE measurement. Spin-waves are excited in the dipolar regime by the CPW and propagate through the trapezoid structure. The trapezoid was chosen to have a maximum width of 30 µm, a minimum width of 5 µm and a length of 80 µm. A static external field along the x-direction was applied throughout the experiment. (b) Schematic of all-electrical VNA spin-wave spectroscopy measurement. Spin-waves are excited from the first microstrip and detected via two more microstrips at different positions along the trapezoid geometry. Each microstrip is connected to a separate port of a four-port VNA. mation of edge channels, and a gradual decrease of wavelength along the propagation direction are ob- served. At a distinct position in the trapezoid, the propagation ceases. We show that this stop posi- tion is locally induced by the reduced effective field, which grants access to the spin-wave transmission stop band. Based on these findings, we demonstrate spatial control of spin-wave propagation within a trapezoid-shaped device by tuning the static external field close to the stop band. We utilize this effect for the active transmission control between microstrip lines. II Experimental Results The first set of experiments was carried out us- ing time-resolved magneto-optic Kerr effect (TR- MOKE) microscopy. Here, the dynamic out-of-plane magnetization component δmzis spatially mapped in the xy-plane, and a direct observation of spin- wave propagation in the sample is obtained. Simul- taneously, the reflectivity is detected, providing a to- pographic map of the sample. The measurements were conducted on a 200 nm thick yttrium iron gar- net (YIG) film grown by liquid phase epitaxy on a gadolinium gallium garnet (GGG) substrate. Thetrapezoid shape, with a gradual continuation back to the full film, was patterned by means of opti- cal lithography and subsequent Argon sputtering of the YIG film. For the excitation of spin-waves, a coplanar waveguide (CPW) was fabricated on top of the YIG film by optical lithography and electron beam evaporation of Ti(5 nm)/Au(210 nm). During the measurements, the external bias field was fixed along the CPW, so spin-waves in a DE-like geometry were excited [32]. A schematic of the measurement geometry can be found in Fig. 1(a). As a preliminary step, the spin-wave stop band in the unpatterned plane YIG film was identified by examining SW propagation far away from the patterned trapezoid structure. In this context, line scans of the Kerr signal along the y-direction were recorded as a function of the applied external field at a constant microwave frequency of f= 2.8 GHz. The result is depicted in Fig. 2(a). Here, a clear suppres- sion of spin-wave propagation around 32 mT can be observed. Previous work [22,33] has shown that hy- bridization between the DE-mode and the first-order perpendicular standing spin-wave (PSSW) mode can create a spin-wave stop band in 200 nm YIG. This is further illustrated in Fig. 2(b) by micromag- netic simulations with TetraX [25], an open-source 2Figure 2 (a) Measurement of SW propagation ex- cited by the CPW (gold) in full film YIG as a func- tion of the external field. A suppression of propaga- tion is visible around 32 mT. The grey-scale repre- sents the measured Kerr amplitude. (b) Micromag- netic simulations with TetraX forf= 2.8 GHz. The DE-mode (blue dash-dotted line) and the n=1 mode (blue dashed line) hybridize and form an anti- crossing in the micromagnetic simulations (red line). This results in an attenuation of SW propagation since the group velocity approaches zero. For all the simulations, the following material parameters were used: saturation magnetization Ms= 1.4·105A m, ex- change stiffness Aex= 3.7·10−12J m, gyromagnetic ratio γ= 176GHz T, film thickness L= 200 nm. Python package for finite-element-method micro- magnetic modelling [25]. In zeroth-order pertur- bation theory, according to Kalinikos and Slavin (KS) [34], the n=0 mode (blue dash-dotted line) and n=1 mode (blue dashed line) cross each other. This degeneracy is lifted by the formation of an avoided crossing in the micromagnetic simulations (red line). This leads to a flattening of the disper- sion relation and, in turn, to a decrease in group ve- locity [22]. For the given experimental parameters, the stop band is predicted at approximately 32 mT, consistent with the observed suppression of propa- (a) 313233µ0Hx,eff(mT) 0 20 40 60 80 y (µm)313233µ0Hx,eff (mT)(c)03060 µ0Hx,eff (mT)010203040 x (µm)(b)Figure 3 Micromagnetic simulations of the x- component of the effective field inside the trapezoid structure. The effective magnetic field varies locally across the geometry in (a). In the vicinity of edges, it is significantly reduced. The grey contour depicts the spatial boundaries of the simulated YIG struc- ture. Across the width of the structure shown in (b), a strong dip of the field at the edges of the ge- ometry is visible. Here, the grey-shaded rectangles indicate the areas outside of the magnetic structure. The effective field along the length of the trapezoid gradually decreases, as shown in (c). The full film hybridization condition at 2.8 GHz is marked with a green dot. gation in the full film line scans. We thus conclude that the pronounced attenuation can be attributed to the hybridization-induced stop band. To understand the influence of the trapezoid ge- ometry on SW propagation, and in turn, on the hy- bridization condition, further micromagnetic simu- lations were performed [26] to determine the effec- tive field of the tapered SW waveguide. Fig. 3(a) shows the spatial distribution of the x-component µ0Hx,effof the effective field at an externally ap- plied field µ0Hx= 33 .5 mT. The simulations re- vealed that the effective field varies locally inside the trapezoid and is strongly reduced at the YIG edges. The iso-field lines (black lines), which dis- play rounded triangular-like features, further illus- trate the inhomogeneous spatial distribution. Along the width of the trapezoid (Fig. 3(b)), we observe sharp edge pockets of low internal field, and a grad- ual decrease of field along the axis of spin-wave prop- agation (Fig. 3(c)). The origin of this inhomogeneity of the effective field lies in the geometry-induced de- magnetizing field, which aims to avoid the formation of magnetic surface charges [23,24]. Another important consideration regarding the ef- 3fect of the modified trapezoid waveguide is the emer- gence of additional width modes in the SW disper- sion relation due to the finite waveguide width [35– 37]. However, this width quantization does not af- fect the PSSWs, and the intersection of modes is still present. Thus, we argue that the key influence of the trapezoid geometry on the stop band is the reduction in effective field which results in a local variation of the dispersion relation. A more detailed discussion concerning the width quantization can be found in the supplementary material. Next, we experimentally investigate the effect of the geometry-induced field distribution on spin-wave propagation. Fig. 4(a) displays TR-MOKE measure- ments in which plane spin-waves are launched from the CPW into the trapezoid in the y-direction, with an excitation frequency f= 2.8 GHz and a static ex- ternal field µ0Hx= 33.5 mT. Changes in the prop- agation characteristics are observed upon entering the trapezoid. Apart from a prominent mode with slightly bent wavefronts in the trapezoid center, a localized mode with strongly bent wavefronts close to the edges appears. We also note that a magnetic contrast right at the edges of the patterned struc- ture was observed in some Kerr images. We argue that this artifact is due to imperfections in the fab- rication process and discuss it in more detail in the supplementary material. The observed bending of wavefronts can be at- tributed to the inhomogeneous internal field pro- file [24], where a local reduction in the effective field causes a shift towards lower fields and lower wave- lengths in the spin-wave dispersion relation. Addi- tionally, the edge localization of modes is a direct consequence of the low-field pockets in the effec- tive field distribution (Fig. 3(b)) as reported previ- ously [30,38,39]. Furthermore, the center mode changes wavelength as it travels through the trapezoid structure. No- tably, it comes to a halt at a specific position in space, beyond which the spin-wave propagation is almost entirely suppressed. This behaviour is illus- trated in more detail in Fig. 4(b), where a y-line pro- file (red curve) along the red dashed line in Fig. 4(a) is plotted. A line scan on plane YIG (blue curve), far away from any patterned structure, is also presented for comparison. As the spin-wave enters the trape- zoid, its wavelength gradually decreases up to more than 50%, consistent with the simulated decrease in the effective field (Fig. 3(c)). However, the propa- gation abruptly ceases at a specific position in space Figure 4 Kerr images at 2.8 GHz. (a) Propaga- tion of the main mode stops at a distinct position in space. The golden area depicts the excitation source. Light red and blue indicate saturation of the grey-scale. (b) Gradual decrease in wavelength and suppression of propagation along the trapezoid (red line in (a)) is observed. Propagation in the non- patterned plane YIG (blue curve) is also shown. (c) Below the hybridization field, propagation along the full trapezoid is observed. (d)-(e) By slightly tuning the field above the stop band, spin-wave propagation vanishes at different positions in space. 4(y≈74µm). From Fig. 3(c), we observe that the stop position of the center mode within the wedge corresponds to an estimated effective field of about 32 mT which aligns well with the measured full film hybridization condition (Fig. 2). Thus, we conclude that the reduction in effective field at different po- sitions in space leads to the local dispersion enter- ing the hybridization regime at a specific position in space, resulting in a sharp local attenuation of spin- wave propagation. Now, we aim to apply our findings towards the active manipulation of spin-wave propagation. To this end, additional Kerr images as a function of the external field were taken and are depicted in Figs. 4(c)-(e). Below the hybridization field, at 31.5 mT (Fig. 4(c)), propagation along the full length of the trapezoid without any sharp attenuation is present. No spatial suppression of propagation is observable since the effective field is only further reduced inside the trapezoid, and thus, the stop band regime is never reached. We also note a com- plex spatial beating profile with a prominent node aty≈45µm, and several less prominent ones. This self-focusing effect results from interference of the width modes induced by the tapered waveguide geometry and has been reported in magnonic mi- crostripes before [35, 36]. Furthermore, caustic-like beams induced by the corners where the full film transitions into the trapezoid may emerge [22, 40]. These caustic-like beams are reflected back and forth at the edges, resulting in non-equidistant areas of higher and lower amplitude. On tuning the external field slightly above 32 mT (Figs. 4(d)-(e)), however, the spin-wave propagation ceases at different positions in space. Furthermore, the boundaries of the spin-wave pattern display a shape reminiscent of the iso-field lines in the effective field. As the external field increases, the positions where the dispersion relation locally gains access to the transmission stop band also shift further outward along the y-direction. As a result, the geometry- induced hybridization allows to actively control the spin-wave propagation distance merely by tuning the external field within a reasonable range. Further all-electrical Vector Network Analyzer (VNA) spin-wave spectroscopy measurements were conducted with the intention to demonstrate the control of spin-wave propagation within a potential magnonic device. For this purpose, three 800 nm wide Au microstrips were patterned at different posi- tions along the trapezoid structure. One microstripserved as a source of spin-wave excitation, while the other two served for detection. The microstrips were connected to separate ports of a four-port VNA, and broadband spin-wave spectroscopy was performed. Note that the choice to employ microstrips instead of CPWs was made in order to obtain a more contin- uous range of wavenumbers for both the excitation and detection processes. A sketch of the measure- ment geometry is depicted in Fig. 1(b). Fig. 5(a) displays the detected spin-wave transmis- sion spectra showcasing the amplitudes of the scat- tering parameters S21,S31in terms of |∆S21|and |∆S31|. We point out that, in the following discus- sion,|Sij|denotes the absolute values of the detected scattering parameters whereas |∆Sij|refers to ab- solute values where a high-field subtraction method was applied. Also note that for better visibility, only data close to the stop band condition is de- picted. Full transmission spectra, along with more detailed information about the data processing pro- cedure, can be found in the supplementary material. Both spectra exhibit amplitude oscillations with the |∆S31|spectrum displaying shorter spacing between these oscillations. This is due to the change of the lateral spin wave profile, where the positions of high- amplitude and caustic-like nodes shift due to changes in the external magnetic field and applied frequency. This effect leads to a smaller node spacing in the field domain at the location of the third microstrip due to the gradual decrease in trapezoid width [35]. Moreover, distinct wide regions with low to no transmission in the spectra (highlighted by red dot- ted lines) occur at conditions in accordance with the spin wave stop band. For the transmission |∆S31|, this band is noticeably broader compared to the |∆S21|spectrum and is reached at lower frequencies at a given field (compare white dashed lines). This is a direct consequence of the spatially varying oc- currence of the hybridization condition suppressing the propagation of spin waves over a broader range of fields and frequencies the further they advance along the trapezoid. To put it differently, distinct external field and frequency conditions exist where transmission is absent in both |∆S21|and|∆S31|, transmission is observed only in |∆S21|, and trans- mission occurs in both |∆S21|and|∆S31|. As a re- sult, selective control over the transmission between the microstrips can be achieved by slightly tuning the frequency or the applied bias field. This is further illustrated in the continuous wave (CW) mode measurements at fixed frequen- 515 25 35 45 field (mT)2.22.32.42.52.62.72.82.9f (GHz)|∆S21| 15 25 35 45 field (mT)|∆S31| min max Magnitude (arb. u.) 1001012.45 GHz(a) (b) 18 19 20 21 22 23 24 25 field (mT)100101|S21||S31| 1001012.7 GHz 24 25 26 27 28 29 30 31 field (mT)100101Magnitude (arb. u.)Figure 5 Selective control of transmission between microstrips along trapezoid-like structure. (a) Spin-wave transmission spectra (amplitudes |∆S21|and|∆S31|). A region of low to no transmission consistent with the expected hybridization conditions occurs in both spectra (red dotted lines serve as guides to the eye). This region appears broader in the transmission spectrum from port 1 to port 3. Moreover, at a given external field, the stop band starts at lower frequencies in |∆S31|compared to |∆S21|as highlighted by white dashed lines. (b) Transmission signals in CW mode at fixed frequencies. The hybridization-induced stop band (roughly marked by gray-shaded areas) spans to higher applied fields in the |S31|transmission trace. cies shown in Fig. 5(b). The regions of suppressed transmission shift with applied frequency and span over a broader field range in the |S31|parameter. The hybridization-induced stop band (highlighted by gray-shaded regions) extends to higher fields due to localized effective field reduction. For instance, at 2.45 GHz and with a field of 23 mT, we observe trans- mission in the |S21|channel but minimal transmis- sion in the |S31|trace, similar to 2.7 GHz at 29 mT. Interestingly, the transmission in |S31|also appears to be suppressed for fields slightly below the hy- bridization field. Additional TR-MOKE data in the supplementary material reveals that this behavior can be attributed to the formation of caustic-like beams that are significantly attenuated upon propa- gation along the geometry. To conclude this section, we suggest employing multiple microstrips along the trapezoid geometry for potential logic operation. Moreover, in the sup- plementary material, we provide further discussion on properties of the hybridization, such as its thick- ness dependence.III Conclusion In conclusion, we demonstrated the feasibility of ac- tively controlling the spin-wave propagation distance by combining the hybridization-induced stop band and a geometry-induced variation of the effective field in 200 nm YIG within a trapezoid-shaped mag- netic film. Experiments and micromagnetic simula- tions were performed to gain insight into the effect of the trapezoid geometry on the effective field. The re- sults show that the spin-wave transmission stop band locally induces the stop position, allowing for spatial control of spin-wave propagation within a trapezoid- shaped device by tuning the static external field close to the stop band. Using multiple microstrips along the trapezoid, we further demonstrated the feasibility of active transmission control between mi- crostrips by external field and frequency. The pro- posed method offers a promising approach for fur- ther advancing spin-wave-based computing and data processing applications. 6References [1] K.-S. Lee and S.-K. Kim, “Conceptual design of spin wave logic gates based on a mach–zehnder- type spin wave interferometer for universal logic functions,” J. Appl. Phys. , vol. 104, p. 053909, Sept. 2008. [2] T. Schneider, A. A. Serga, B. Leven, B. Hille- brands, R. L. Stamps, and M. P. Kostylev, “Re- alization of spin-wave logic gates,” Appl. Phys. Lett., vol. 92, p. 022505, Jan. 2008. [3] V. V. Kruglyak, S. O. Demokritov, and D. Grundler, “Magnonics,” J. Phys. D: Appl. Phys. , vol. 43, p. 260301, July 2010. [4] A. Khitun and K. L. Wang, “Non-volatile magnonic logic circuits engineering,” J. Appl. Phys. , vol. 110, p. 034306, Aug. 2011. [5] A. V. Sadovnikov, C. S. Davies, S. V. Gr- ishin, V. V. Kruglyak, D. V. Romanenko, Y. P. Sharaevskii, and S. A. Nikitov, “Magnonic beam splitter: The building block of parallel magnonic circuitry,” Appl. Phys. Lett. , vol. 106, p. 192406, May 2015. [6] A. V. Chumak, V. I. Vasyuchka, A. A. Serga, and B. Hillebrands, “Magnon spintronics,” Nat. Phys. , vol. 11, pp. 453–461, June 2015. [7] Q. Wang, P. Pirro, R. Verba, A. Slavin, B. Hille- brands, and A. V. Chumak, “Reconfigurable nanoscale spin-wave directional coupler,” Sci. Adv., vol. 4, Jan. 2018. [8] Q. Wang, A. Hamadeh, R. Verba, V. Lomakin, M. Mohseni, B. Hillebrands, A. V. Chumak, and P. Pirro, “A nonlinear magnonic nano-ring res- onator,” npj Comput. Mater. , vol. 6, Dec. 2020. [9] A. Barman, G. Gubbiotti, S. Ladak, A. O. Adeyeye, M. Krawczyk, J. Gr¨ afe, C. Adel- mann, S. Cotofana, A. Naeemi, V. I. Vasyuchka, B. Hillebrands, S. A. Nikitov, H. Yu, D. Grundler, A. V. Sadovnikov, A. A. Grachev, S. E. Sheshukova, J.-Y. Duquesne, M. Marangolo, G. Csaba, W. Porod, V. E. Demidov, S. Urazhdin, S. O. Demokritov, E. Al- bisetti, D. Petti, R. Bertacco, H. Schultheiss, V. V. Kruglyak, V. D. Poimanov, S. Sa- hoo, J. Sinha, H. Yang, M. M¨ unzenberg, T. Moriyama, S. Mizukami, P. Landeros, R. A.Gallardo, G. Carlotti, J.-V. Kim, R. L. Stamps, R. E. Camley, B. Rana, Y. Otani, W. Yu, T. Yu, G. E. W. Bauer, C. Back, G. S. Uhrig, O. V. Do- brovolskiy, B. Budinska, H. Qin, S. van Dijken, A. V. Chumak, A. Khitun, D. E. Nikonov, I. A. Young, B. W. Zingsem, and M. Winklhofer, “The 2021 magnonics roadmap,” J. Phys.: Con- dens. Matter , vol. 33, p. 413001, Aug. 2021. [10] ´A. Papp, W. Porod, and G. Csaba, “Nanoscale neural network using non-linear spin-wave inter- ference,” Nat Commun , vol. 12, Nov. 2021. [11] A. V. Chumak, P. Kabos, M. Wu, C. Abert, C. Adelmann, A. O. Adeyeye, J. Aker- man, F. G. Aliev, A. Anane, A. Awad, C. H. Back, A. Barman, G. E. W. Bauer, M. Becherer, E. N. Beginin, V. A. S. V. Bitten- court, Y. M. Blanter, P. Bortolotti, I. Boven- ter, D. A. Bozhko, S. A. Bunyaev, J. J. Carmiggelt, R. R. Cheenikundil, F. Ciubotaru, S. Cotofana, G. Csaba, O. V. Dobrovolskiy, C. Dubs, M. Elyasi, K. G. Fripp, H. Fulara, I. A. Golovchanskiy, C. Gonzalez-Ballestero, P. Graczyk, D. Grundler, P. Gruszecki, G. Gub- biotti, K. Guslienko, A. Haldar, S. Ham- dioui, R. Hertel, B. Hillebrands, T. Hioki, A. Houshang, C.-M. Hu, H. Huebl, M. Huth, E. Iacocca, M. B. Jungfleisch, G. N. Kakazei, A. Khitun, R. Khymyn, T. Kikkawa, M. Klaui, O. Klein, J. W. Klos, S. Knauer, S. Koraltan, M. Kostylev, M. Krawczyk, I. N. Krivorotov, V. V. Kruglyak, D. Lachance-Quirion, S. Ladak, R. Lebrun, Y. Li, M. Lindner, R. Macedo, S. Mayr, G. A. Melkov, S. Mieszczak, Y. Naka- mura, H. T. Nembach, A. A. Nikitin, S. A. Nikitov, V. Novosad, J. A. Otalora, Y. Otani, A. Papp, B. Pigeau, P. Pirro, W. Porod, F. Porrati, H. Qin, B. Rana, T. Reimann, F. Riente, O. Romero-Isart, A. Ross, A. V. Sadovnikov, A. R. Safin, E. Saitoh, G. Schmidt, H. Schultheiss, K. Schultheiss, A. A. Serga, S. Sharma, J. M. Shaw, D. Suess, O. Surzhenko, K. Szulc, T. Taniguchi, M. Urbanek, K. Usami, A. B. Ustinov, T. van der Sar, S. van Dijken, V. I. Vasyuchka, R. Verba, S. V. Kusminskiy, Q. Wang, M. Weides, M. Weiler, S. Wintz, S. P. Wolski, and X. Zhang, “Advances in magnet- ics roadmap on spin-wave computing,” IEEE Trans. Magn. , vol. 58, pp. 1–72, June 2022. [12] T. Goto, T. Yoshimoto, B. Iwamoto, K. Shi- mada, C. A. Ross, K. Sekiguchi, A. B. Gra- 7novsky, Y. Nakamura, H. Uchida, and M. In- oue, “Three port logic gate using forward vol- ume spin wave interference in a thin yttrium iron garnet film,” Sci. Rep. , vol. 9, Nov. 2019. [13] H. Qin, R. B. Holl¨ ander, L. Flajˇ sman, F. Her- mann, R. Dreyer, G. Woltersdorf, and S. van Dijken, “Nanoscale magnonic fabry-p´ erot res- onator for low-loss spin-wave manipulation,” Nat Commun , vol. 12, Apr. 2021. [14] G. Gubbiotti, S. Tacchi, M. Madami, G. Car- lotti, A. O. Adeyeye, and M. Kostylev, “Bril- louin light scattering studies of planar metal- lic magnonic crystals,” J. Phys. D: Appl. Phys. , vol. 43, p. 264003, June 2010. [15] A. A. Serga, A. V. Chumak, and B. Hillebrands, “YIG magnonics,” J. Phys. D: Appl. Phys. , vol. 43, p. 264002, June 2010. [16] B. Lenk, H. Ulrichs, F. Garbs, and M. M¨ unzenberg, “The building blocks of magnonics,” Phys. Rep. , vol. 507, pp. 107–136, Oct. 2011. [17] Z. K. Wang, V. L. Zhang, H. S. Lim, S. C. Ng, M. H. Kuok, S. Jain, and A. O. Adeyeye, “Nanostructured magnonic crystals with size- tunable bandgaps,” ACS Nano , vol. 4, pp. 643– 648, Jan. 2010. [18] M. Krawczyk and D. Grundler, “Review and prospects of magnonic crystals and devices with reprogrammable band structure,” J. Phys.: Condens. Matter , vol. 26, p. 123202, Mar. 2014. [19] M. A. Morozova, S. V. Grishin, A. V. Sadovnikov, D. V. Romanenko, Y. P. Sharaevskii, and S. A. Nikitov, “Band gap control in a line-defect magnonic crystal waveg- uide,” Appl. Phys. Lett. , vol. 107, p. 242402, Dec. 2015. [20] A. J. E. Kreil, H. Y. Musiienko-Shmarova, S. Eggert, A. A. Serga, B. Hillebrands, D. A. Bozhko, A. Pomyalov, and V. S. L 'vov, “Tun- able space-time crystal in room-temperature magnetodielectrics,” Phys. Rev. B , vol. 100, July 2019. [21] F. Lisiecki, J. Rych/suppress ly, P. Ku´ swik, H. G/suppress lowi´ nski, J. W. K/suppress los, F. Groß, I. Bykova, M. Weigand, M. Zelent, E. J. Goering, G. Sch¨ utz, G. Gub- biotti, M. Krawczyk, F. Stobiecki, J. Dubowik,and J. Gr¨ afe, “Reprogrammability and scalabil- ity of magnonic fibonacci quasicrystals,” Phys. Rev. Applied , vol. 11, May 2019. [22] C. Riedel, T. Taniguchi, L. K¨ orber, A. K´ akay, and C. H. Back, “Hybridization-induced spin- wave transmission stop band within a 1d diffraction grating,” Advanced Physics Re- search , vol. 2, Feb. 2023. [23] J. M. D. Coey, Magnetism and Magnetic Mate- rials. Cambridge University Press, Jan. 2001. [24] P. Gruszecki, J. Romero-Vivas, Y. S. Da- doenkova, N. N. Dadoenkova, I. L. Lyubchan- skii, and M. Krawczyk, “Goos-h¨ anchen ef- fect and bending of spin wave beams in thin magnetic films,” Appl. Phys. Lett. , vol. 105, p. 242406, Dec. 2014. [25] L. K¨ orber, A. Hempel, A. Otto, R. A. Gallardo, Y. Henry, J. Lindner, and A. K´ akay, “Finite- element dynamic-matrix approach for propagat- ing spin waves: Extension to mono- and multi- layers of arbitrary spacing and thickness,” AIP Adv., vol. 12, p. 115206, Nov. 2022. [26] A. Vansteenkiste, J. Leliaert, M. Dvornik, M. Helsen, F. Garcia-Sanchez, and B. V. Waeyenberge, “The design and verification of MuMax3,” AIP Adv. , vol. 4, p. 107133, Oct. 2014. [27] K. Perzlmaier, G. Woltersdorf, and C. H. Back, “Observation of the propagation and interfer- ence of spin waves in ferromagnetic thin films,” Phys. Rev. B , vol. 77, Feb. 2008. [28] M. Farle, T. Silva, and G. Woltersdorf, “Spin dynamics in the time and frequency domain,” inSpringer Tracts in Modern Physics , pp. 37– 83, Springer Berlin Heidelberg, Sept. 2012. [29] Y. Au, T. Davison, E. Ahmad, P. S. Keatley, R. J. Hicken, and V. V. Kruglyak, “Excitation of propagating spin waves with global uniform microwave fields,” Appl. Phys. Lett. , vol. 98, p. 122506, Mar. 2011. [30] H. G. Bauer, J.-Y. Chauleau, G. Woltersdorf, and C. H. Back, “Coupling of spinwave modes in wire structures,” Appl. Phys. Lett. , vol. 104, p. 102404, Mar. 2014. 8[31] J. Stigloher, T. Taniguchi, H. K¨ orner, M. Decker, T. Moriyama, T. Ono, and C. Back, “Observation of a goos-h¨ anchen-like phase shift for magnetostatic spin waves,” Phys. Rev. Lett. , vol. 121, Sept. 2018. [32] R. Damon and J. Eshbach, “Magnetostatic modes of a ferromagnet slab,” J. Phys. Chem. Solids , vol. 19, pp. 308–320, May 1961. [33] M. Vaˇ natka, K. Szulc, O. Wojewoda, C. Dubs, A. V. Chumak, M. Krawczyk, O. V. Dobrovol- skiy, J. W. K/suppress los, and M. Urb´ anek, “Spin-wave dispersion measurement by variable-gap propa- gating spin-wave spectroscopy,” Phys. Rev. Ap- plied, vol. 16, Nov. 2021. [34] B. A. Kalinikos and A. N. Slavin, “Theory of dipole-exchange spin wave spectrum for ferro- magnetic films with mixed exchange bound- ary conditions,” J. Phys. C: Solid State Phys. , vol. 19, pp. 7013–7033, Dec. 1986. [35] V. E. Demidov, S. O. Demokritov, K. Rott, P. Krzysteczko, and G. Reiss, “Mode interfer- ence and periodic self-focusing of spin waves in permalloy microstripes,” Physical Review B , vol. 77, Feb. 2008. [36] X. Xing, Y. Yu, S. Li, and X. Huang, “How do spin waves pass through a bend?,” Scientific Reports , vol. 3, Oct. 2013. [37] A. V. Chumak, “Fundamentals of magnon- based computing,” 2019. [38] J. Topp, J. Podbielski, D. Heitmann, and D. Grundler, “Internal spin-wave confinement in magnetic nanowires due to zig-zag shaped mag- netization,” Phys. Rev. B , vol. 78, July 2008. [39] J. Jorzick, S. O. Demokritov, B. Hillebrands, M. Bailleul, C. Fermon, K. Y. Guslienko, A. N. Slavin, D. V. Berkov, and N. L. Gorn, “Spin wave wells in nonellipsoidal micrometer size magnetic elements,” Phys. Rev. Lett. , vol. 88, Jan. 2002. [40] A. Wartelle, F. Vilsmeier, T. Taniguchi, and C. H. Back, “Caustic spin wave beams in soft thin films: Properties and classification,” Phys. Rev. B , vol. 107, Apr. 2023. 9Supplementary Material: Spatial Control of Hybridization Induced Spin-Wave Transmission Stop Band Franz Vilsmeier∗1, Christian Riedel1, and Christian H. Back1 1Fakult¨ at f¨ ur Physik, Technische Universit¨ at M¨ unchen, Garching, Germany March 23, 2024 I Time-Resolved Magneto-Optic Kerr Effect Microscopy As a source of illumination, a mode-locked Ti:Sa laser with a centre wavelength of 800 nm and a pulse width of around 150 fs is used. The pulse trains are applied at a fixed repetition rate of 80 MHz. Subsequently, we fix the polar- ization plane of the laser and focus it onto the sample through an objective lens with a numerical aperture of 0.7, giving a maximum resolution of ∼0.6µm. Upon reflection at a magnetic surface, the polarization changes due to the po- lar magneto-optical Kerr effect. Here, the change of polarization rotation is directly proportional to the change in the dynamic out-of-plane magnetization component. A Wollaston prism splits the reflected signal into two beams with orthogonal polarization components, which are detected by two photodiodes. The difference between the two photodiode signals then gives a direct represen- tation of the change in magnetization - the Kerr signal -, and the sum of the two is proportional to the sample’s reflectivity. Since the sample is mounted onto a piezo stage, the relative laser focus position can be spatially scanned in the sample plane. Hence, Kerr image and topography are obtained. In addition, during the acquisition of a Kerr image, the relative phase relation between the applied rf-frequency in the GHz regime and the laser repetition rate is fixed. This requires the driving field frequency to always be a multiple of the laser repetition rate. As a result of the constant phase and the short laser pulses (much shorter than one period of the excitation), we can directly access the dynamic out-of-plane magnetization component and observe the propagation of spin-waves excited by our antenna structure. Fig. 1 shows some spatial Kerr maps approaching the full film hybridization condition from lower fields. Caustic-like beams emerge, which are damped along the trapezoid with increasing field. ∗franz.vilsmeier@tum.de 1arXiv:2403.15840v1 [physics.app-ph] 23 Mar 2024020406080 y (µm)02040x (µm)32.2 mT 020406080 y (µm)32.3 mT 020406080 y (µm)32.4 mT 020406080 y (µm)32.5 mT min maxδmzFigure 1 Kerr images recorded at f= 2.8 GHz at different fields coming from below the stop band. Caustic-like beams are reflected back and forth at the edges and fade away along the trapezoid with increasing field. II Effect of Waveguide Width on Hybridiza- tion Following the analytical model by Kalinikos and Slavin [1], the full film disper- sion relation in the presence of an in-plane external field with totally unpinned surface states is given by ω2 n=/parenleftbig ωH+l2 exk2 nωM/parenrightbig/parenleftbig ωH+l2 exk2 nωM+ωMFnn/parenrightbig , (1) where ωH=γµ0H, (2) ωM=γµ0MS, (3) Fnn=Pnn+/parenleftbigg 1−Pnn/parenleftbig 1 + cos2φ/parenrightbig +ωMPnn(1−Pnn) sin2φ ωH+l2exk2nωM/parenrightbigg , (4) and Pnn=k2 k2n−k4 k4nFn1 1 +δ0n, Fn=2 kL/parenleftbig 1−(−1)ne−kL/parenrightbig .(5) Furthermore, n= 0,1,2, ...denote the eigenmode orders across the film thick- nessL,kn=/radicalig k2+/parenleftbignπ L/parenrightbig2, and φdescribes the angle between kandM(so for k⊥M,φ=π 2). Considering a spin wave waveguide of finite width w, an additional quanti- zation across the waveguide width is introduced and the dispersion relation can be represented using equ. (1) by letting k→/radicalig k2+/parenleftbigmπ w/parenrightbig2andφ→ φ−arctan/parenleftbigmπ kw/parenrightbig [2–4]. Here, m= 0,1,2, ...denote the eigenmode orders across the width, and kdenotes the wavenumber along the waveguide. In the specific case of a tangentially magnetized waveguide in the DE-geometry ( k⊥M), de- magnetization also has to be taken into account and the non-uniform effective field µ0Heffneeds to be considered in the dispersion relation in place of the externally applied field. In the case of spin wave propagation in the center, a uniform field is assumed, but an effective waveguide width weffis introduced to account for the strong reduction in the effective field at the edges. The effective width can be defined in different ways. Here, we follow the definition by Chu- mak [4], where weffis given by the distance of points across the width where the effective field is reduced by 10%. i.e., to the value 0 .9·µ0Hmax eff. 2From micromagnetic simulations [5], the effective field within the trapezoid geometry at an externally applied field of 32 mT (close to the full film hybridiza- tion field at 2.8 GHz) was determined. From the field distribution, the effective field and effective width for the center mode were extracted as a function of trapezoid width w. The respective results are depicted in Figs 2(a)-(b). At the smallest width, the effective field is reduced by almost 3 mT with respect to the applied field. The effective width is maximally reduced to about 65 % of the actual waveguide width, allowing for a rather wide region of uniform field and mode propagation across the width. 29303132µ0Hx,eff(mT)(a) 5 10 15 20 25 30 w(µm)0.70.80.9weff/w(b)2.12.42.73.03.3f (GHz)(c) wg, m=0 wg, m=1 wg, m=2wg, n=1 ff, n=0 ff, n=1 0 1 2 3 k (µm−1)2.72.9f (GHz)(d) n=1, m=0 n=1, m=1 Figure 2 Effect of width modes on dispersion relation. (a) Effective field from micromagnetic simulations for the center mode as a function of trapezoid width w. (b) Ratio of extracted effective width weffand actual trapezoid width w. (c) Dispersion relations of the waveguide (wg) modes considering effective field and effective width for w= 6µm. Modes with n=0 and m=0 (blue dash-dotted line), n=0 and m=1 (red dash-dotted line), n=0 and m=2 (purple dash-dotted line), and n=1 and m=0 (red dashed line) are shown. The full film (ff) modes with n=0 and n=1 are also depicted (grey lines) for comparison. (d) Waveguide modes with n=1 and m=0, and n=1 and m=1. The width modification does not significantly affect the first-order PSSW. Forw= 6µm, the resulting dispersion relations for a waveguide with several width modes (m=0, m=1, m=2) and thickness modes (n=0, n=1) are displayed in Fig. 2(c). Note that for the waveguide case, only the m=0 mode of the first- order (n=1) PSSW mode is shown, as the width quantization doesn’t notably affect the higher-order PSSWs (see Fig. 2(d)). The reduced effective field gener- ally shifts the dispersion relation towards lower frequencies compared to the full film case (grey lines). The higher-order width modes (m=1, m=2) also display lower frequencies in the dipolar regime than the m=0 mode. More importantly, however, the n=0 and n=1 thickness modes still intersect in the dipolar regime, facilitating a hybridization and corresponding stop band. From this, we con- clude that the main effect of the width modulation on the hybridization is the reduced effective field and the resulting shift in the hybridization condition. This is especially the case for the m=0 width mode, which should be dominant in the trapezoid structure due to the transmission of spin waves from the full film into the tapered waveguide. 3III Broadband Spin-Wave Spectroscopy A four-port vector network analyzer (Agilent N5222A) was used for the broad- band spin-wave spectroscopy. All measurements were conducted at a microwave power of 3 dBm, and the real and imaginary parts of the complex scattering pa- rameters S21andS31were recorded. A frequency sweep method was applied at different external magnetic field values for the transmission spectra. The mag- netic field’s strength was changed stepwise (5 mT steps) from high to low field. To improve contrast, a high-field subtraction method was applied. Reference data S21,refandS31,reftaken at 200 mT was recorded. The presented spectra were then obtained by subtracting the absolute value of the reference data from the absolute value of the scattering parameters, i.e., |∆S21|=|S21|−|S21,ref|and |∆S31|=|S31|−|S31,ref|. Exemplary recorded spectra are shown in Fig. 3 where modes close or at the FMR are prominent. In the CW mode measurements, no reference was taken. −60−40−200204060 field (mT)0.51.01.52.02.53.03.5f (GHz)|∆S21| minmax Magnitude (arb. u.) −60−40−200204060 field (mT)|∆S31| Figure 3 Broadband spin-wave spectroscopy spectra. Modes close to FMR are very prominent in the transmission spectra |∆S21|and|∆S31|. IV Edge Mode Due to Imperfect Fabrication Fig. 4 shows TR-MOKE measurements at different frequencies and relative phases between the microwave excitation and laser pulses at an applied field of 33.5 mT. Apart from the propagation inside the trapezoid, Kerr images where intense caustic-like beams are detected (2.4 GHz, 2.48 GHz, 2.56 GHz) also ex- hibit a magnetic contrast right at the edges of the patterned YIG. In close vicinity to the points where the beams are reflected from the trapezoid edges, a localized mode profile outside the previous propagation region in x-direction is visible. However, no such distinct feature seems to occur when DE-like modes become dominant in the profile (2.72 GHz). Fig. 5(a) depicts an AFM image of the patterned YIG trapezoid. Linescans across the edges (Fig. 5(b)) reveal that the transition from YIG to GGG sub- strate is not sharp, but rather gradual along a distance of about 1-2 µm. This is attributed to imperfections in the fabrication process. The boundary regions, from which the caustic-like beams scatter, serve as secondary point-like excitation sources with a finite size of the order of the 402040x (µm)2.4 GHz 0 deg 2.48 GHz 0 deg 2.56 GHz 0 deg 2.72 GHz 0 deg 020 40 60 y (µm)02040x (µm)2.4 GHz 90 deg 020 40 60 y (µm)2.48 GHz 90 deg 020 40 60 y (µm)2.56 GHz 90 deg 020 40 60 y (µm)2.72 GHz 90 deg min maxδmzFigure 4 TR-MOKE measurements for different frequencies and phases be- tween microwave excitation and laser pulses at an external field of 33.5 mT. In the vicinity where the caustic-like beams scatter from the edges, an additional mode profile is observed. We note that the antisymmetric beam directions stem from a slight mismatch of the external field angle with respect to the DE- geometry. 051015202530 y (µm)05101520253035x (µm)(a) 0 100 200profile (nm) −2−1012 ∆x (µm)050100150200250profile (nm)(b) upper lower −6−4−20246 ky(µm−1)−8−6−4−202468kx(µm−1)(c) 50 nm 100 nm150 nm 200 nm Figure 5 (a) AFM profile of patterned YIG structure. (b) Linescans taken across the edges highlighted in blue and red in (a). A gradual transition from the YIG to the GGG is observed. (c) Iso-frequency curves for several film thicknesses at 2.48 GHz and 33.5 mT. beam’s width, as noted in previous works [4,6]. Consequently, spin wave modes may be excited within the transitional region where the thickness of YIG de- creases. Examining the iso-frequency curves at 2.48 GHz and 33.5 mT (see Fig. 5 (c)) reveals that the modes potentially excited fall within our resolution limits across a considerable range of film thicknesses. 5V Some Properties of Hybridization-Induced Stop Band This section offers a brief overview of some characteristics of the anticrossing in full YIG films. All micromagnetic simulations were executed utilizing the TetraX [7] software package. To provide a qualitative understanding of the hybridization’s coupling strength, we introduce the quantity ∆ fas the minimal gap between the upper and lower band determined by micromagnetic simulations. We note that here, we only consider the strength of hybridization in the frequency domain, as this needed significantly less computing time. Furthermore, we inferred the wave vector of hybridization khybby the intercept of the n=0 and n=1 modes according to the model by Kalinikos and Slavin [1]. 150 250 350 450 YIG thickness (nm)01234khyb(µm−1)(a)khyb ∆f 0.000.020.040.060.080.10 field (mT)12345khyb(µm−1)(b)170 nm 200 nm 230 nm 020406080100120 ∆f(MHz) Figure 6 Some properties of hybridization. (a) Trend of hybridization wave number khyband hybridization strength with increasing film thickness at ex- ternal field of 32 mT. Both parameters decrease with increasing thickness. (b) Dependence of khybon the external field for different film thicknesses. khybcan be increased to some extent by the external field strength. In Fig. 6(a), the relationship between film thickness and both ∆ fandkhyb is illustrated at an external field strength of 32 mT. As film thickness increases, both these parameters exhibit a decreasing trend. Notably, below a thickness of 170 nm, the formation of an anticrossing appears to be absent. Here, the increased separation between n=0 and n=1 prevents hybridization from occur- ring. We also note that potential crossing is only possible in the dipolar regime as modes follow k2-dependence in the exchange regime. Fig. 6(b) depicts the influence of the external field on khyb. Higher external field strengths result in higher wave numbers. Furthermore, the external field strength also affects the existence of frequency degeneracy. At higher fields, the n=0 mode becomes flatter and no longer intersects with the n=1 mode. In summary, the manipu- lation of external field strength and sample thickness allows for the adjustment of hybridization properties, facilitating higher wave number values and stronger coupling within a specific range. 6References [1] B. A. Kalinikos and A. N. Slavin, “Theory of dipole-exchange spin wave spec- trum for ferromagnetic films with mixed exchange boundary conditions,” J. Phys. C: Solid State Phys. , vol. 19, pp. 7013–7033, Dec. 1986. [2] V. E. Demidov and S. O. Demokritov, “Magnonic waveguides studied by microfocus brillouin light scattering,” IEEE Transactions on Magnetics , vol. 51, p. 1–15, Apr. 2015. [3] T. Br¨ acher, O. Boulle, G. Gaudin, and P. Pirro, “Creation of unidirectional spin-wave emitters by utilizing interfacial dzyaloshinskii-moriya interaction,” Physical Review B , vol. 95, Feb. 2017. [4] A. V. Chumak, “Fundamentals of magnon-based computing,” 2019. [5] A. Vansteenkiste, J. Leliaert, M. Dvornik, M. Helsen, F. Garcia-Sanchez, and B. V. Waeyenberge, “The design and verification of MuMax3,” AIP Adv., vol. 4, p. 107133, Oct. 2014. [6] T. Schneider, A. A. Serga, A. V. Chumak, C. W. Sandweg, S. Trudel, S. Wolff, M. P. Kostylev, V. S. Tiberkevich, A. N. Slavin, and B. Hille- brands, “Nondiffractive subwavelength wave beams in a medium with exter- nally controlled anisotropy,” Phys. Rev. Lett. , vol. 104, May 2010. [7] L. K¨ orber, A. Hempel, A. Otto, R. A. Gallardo, Y. Henry, J. Lindner, and A. K´ akay, “Finite-element dynamic-matrix approach for propagating spin waves: Extension to mono- and multi-layers of arbitrary spacing and thickness,” AIP Adv. , vol. 12, p. 115206, Nov. 2022. 7 | 2024-03-23 | Spin-wave (SW) propagation close to the hybridization-induced transmission
stop band is investigated within a trapezoid-shaped 200\,nm thick yttrium iron
garnet (YIG) film using time-resolved magneto-optic Kerr effect (TR-MOKE)
microscopy and broadband spin wave spectroscopy, supported by micromagnetic
simulations. The gradual reduction of the effective field within the structure
leads to local variations of the SW dispersion relation and results in a SW
hybridization at a fixed position in the trapezoid where the propagation
vanishes since the SW group velocity approaches zero. By tuning external field
or frequency, spatial control of the spatial stop band position and spin-wave
propagation is demonstrated and utilized to gain transmission control over
several microstrip lines. | Spatial Control of Hybridization-Induced Spin-Wave Transmission Stop Band | 2403.15840v1 |
Detection sensitivity enhancement of magnon Kerr nonlinearity in cavity magnonics induced by coherent perfect absorption Guo-Qiang Zhang,1,Yimin Wang,2,yand Wei Xiong3,z 1School of Physics, Hangzhou Normal University, Hangzhou, Zhejiang 311121, China 2Communications Engineering College, Army Engineering University of PLA, Nanjing 210007, China 3Department of Physics, Wenzhou University, Zhejiang 325035, China (Dated: February 23, 2023) We show how to enhance the detection sensitivity of magnon Kerr nonlinearity (MKN) in cavity magnonics. The considered cavity-magnon system consists of a three-dimensional microwave cavity containing two yttrium iron garnet (YIG) spheres, where the two magnon modes (one has the MKN, while the other is linear) in YIG spheres are simultaneously coupled to microwave photons. To obtain the e ective gain of the cavity mode, we feed two input fields into the cavity. By choosing appropriate parameters, the coherent perfect absorption of the two input fields occurs, and the cavity-magnon system can be described by an e ective non-Hermitian Hamil- tonian. Under the pseudo-Hermitian conditions, the e ective Hamiltonian can host the third-order exceptional point (EP3), where the three eigenvalues of the Hamiltonian coalesce into one. When the magnon frequency shiftKinduced by the MKN is much smaller than the linewidths of the peaks in the transmission spectrum of the cavity (i.e., K ), the magnon frequency shift can be amplified by the EP3, which can be probed via the output spectrum of the cavity. The scheme we present provides an alternative approach to measure the MKN in the region K and has potential applications in designing low-power nonlinear devices based on the MKN. I. INTRODUCTION In the past decade, the progress in cavity-magnon systems has been impressive, where magnons (i.e., collective spin ex- citations) in ferrimagnetic materials are strongly coupled to photons in microwave cavities via the collective magnetic- dipole interaction [1–3]. Experimentally, the most widely used cavity-magnon system is composed of the millimeter- scale yttrium iron garnet (Y 3Fe5O12or YIG) crystal and the three-dimensional (3D) microwave cavity [4–7]. Up to now, various exotic phenomena have been extensively in- vestigated in cavity-magnon systems, such as magnon dark modes [8], manipulating spin currents [9, 10], steady-state magnon-photon entanglement [11], magnon blockade [12– 14], non-Hermitian physics [15–17], cooperative polariton dynamics [18], enhancing spin-photon coupling [19], quan- tum states of magnons [20–23], microwave-to-optical trans- duction [24, 25], and dissipative coupling [26, 27]. Based on the coherent perfect absorption (CPA), the second-order exceptional point (EP2) was observed [28] and the third-order EP (EP3) was subsequently predicted [29] in cavity-magnon systems. The CPA refers to a phenomenon that when two (or more) coherent electromagnetic waves are fed into a medium, the waves are completely absorbed by the medium due to both destructive interference between them and medium dissipation, and there are no output waves from the medium [30, 31]. Intriguing applications of CPA include, e.g., engineering EPs [28, 29, 32, 33], antilasing [34, 35], optical switches [36, 37], and coherent polarization con- trol [38, 39]. The nth-order EP (EP n) refers to the degen- erate point in non-Hermitian systems, where neigenvalues zhangguoqiang@hznu.edu.cn yvivhappyrom@163.com zxiongweiphys@wzu.edu.cnas well as corresponding neigenvectors coalesce simultane- ously [40]. Owing to its fundamental importance and po- tential applications, the EPs have been explored in various physical systems (see, e.g., Refs. [41–49]). Contrary to the degenerate point in Hermitian systems, the EPs have some unique features. For example, the energy splitting follows a1=ndependence around the EP nwhen the non-Hermitian systems are subjected to a weak perturbation with strength (1) [50, 51], which makes it possible to enhance the detec- tion sensitivity [52–54]. It is worth noting that the cavity-magnon system also has reached the nonlinear regime [55], where the magnon Kerr nonlinearity (MKN) stems from the magnetocrystalline anisotropy in the YIG [56]. The MKN not only results in cavity-magnon bistability [57–59] and tristability [60–62], nonreciprocal microwave transmission [63], and strong long- distance spin-spin coupling [64], but it also leads to magnon- photon entanglement [65, 66] as well as dynamical quantum phase transition [67, 68]. In experiments, many phenomena induced by MKN can be detected by measuring the transmis- sion spectrum of the microwave cavity, where the MKN is equivalent to the magnon frequency shift Kdependent on the magnon population [55–63]. This probe method works well only when the magnon frequency shift Kis comparable to (or larger than) the linewidths of the peaks in the transmis- sion spectrum of the cavity (i.e., K ), while it is not valid in the region K [18, 69]. In this paper, we propose a scheme to enhance the detection sensitivity of MKN around an EP in cavity magnonics when K . Here, the considered hybrid system consists of a 3D microwave cavity with two YIG spheres (YIG 1 and YIG 2) embedded (cf. Fig. 1), where the magnon mode in YIG 1 has the MKN, while the auxiliary magnon mode in YIG 2 is linear. By feeding two input fields with the same frequency into the 3D microwave cavity via its two ports, an e ective pseudo- Hermitian Hamiltonian of the cavity-magnon system can be obtained, where the e ective gain of the cavity mode resultsarXiv:2211.08922v2 [quant-ph] 22 Feb 20232 from the CPA of the two input fields. In the absence of the MKN (corresponding to K=0), we analyze the eigenvalues of the pseudo-Hermitian Hamiltonian and find the EP3 in the parameter space. Further, we show that the magnon frequency shiftK( ) induced by the MKN can be amplified by the EP3. Finally, we derive the output spectrum of the 3D cavity and display how the amplification e ect can be probed via the output spectrum. Recently, Ref. [70] has proposed to enhance the sensitivity of the magnon-population response to the coe cient of MKN via the anti-parity-time-symmetric phase transition, where the strength of the drive field on the system is fixed. In contrast to Ref. [70], we show that the EP3 can enhance the sensitivity of the eigenvalue response to the small magnon frequency shift induced by MKN in the present work. Our study provides a possibility to detect the MKN in the region K , which is a complement to the existing approach (i.e., measuring the transmission spectrum of the microwave cavity) [55–63] and may find promising applications in designing low-power non- linear devices in cavity magnonics. In addition to MKN, other weak signals (such as a weak magnetic field), which can re- sult in the changes of system parameters, can also be detected using our scheme. II. THE MODEL As shown in Fig. 1, the considered cavity-magnon system consists of two YIG spheres (YIG 1 and YIG 2) and a 3D mi- crowave cavity, where YIG 1 and YIG 2 are uniformly mag- netized to saturation by the bias magnetic fields B1andB2, re- spectively. Here, to enhance the detection sensitivity of MKN in YIG 1, the YIG 2 provides a magnon mode serving as an ancilla. Now the entire cavity-magnon system is described by the Hamiltonian [56, 57] H=!caya+X j=1;2h !jby jbj+Kjby jbjby jbj+gj(aybj+aby j)i + d(by 1e i!dt+b1ei!dt); (1) where aanday(bjandby jwith j=1;2) are the annihilation and creation operators of the cavity mode (magnon mode in YIG j) at frequency !c(!j),gjis the coupling strength be- tween the cavity mode aand the magnon mode bj, and d (!d) is the strength (frequency) of the drive field on YIG 1. In the two YIG spheres, the magnetocrystalline anisotropy re- sults in the MKN term Kjby jbjby jbj, where the nonlinear coef- ficient Kjcan be continuously tuned from negative values to positive values by adjusting the angle between the crystallo- graphic axis of YIG jand the bias magnetic field Bj[71, 72]. Without loss of generality, we assume K1>0 and K2=0 in our scheme. When macroscopic magnons are excited in YIG 1 (i.e.,hby 1b1i1), the system Hamiltonian in Eq. (1) can be linearized as H=!caya+X j=1;2h !jby jbj+gj(aybj+aby j)i + Kby 1b1 + d(by 1e i!dt+b1ei!dt); (2) FIG. 1. Schematic of the proposed setup for enhancing the detection sensitivity of MKN in YIG 1. The cavity magnonic system is com- posed of two YIG spheres coupled to a 3D microwave cavity, where YIG 1 (YIG 2) is magnetized by a static magnetic field B1(B2). To measure the weak MKN in YIG 1, one microwave field with Rabi frequency dis used to drive YIG 1. In addition, two input fields a(in) 1anda(in) 2are fed into the microwave cavity via ports 1 and 2, re- spectively, and a(out) 1anda(out) 2denote the corresponding output fields. with the frequency shift K=2K1hby 1b1iof the magnon mode b1, where the mean-field approximation by 1b1by 1b1 2hby 1b1iby 1b1has been used [56, 57]. When the magnon frequency shift Kis comparable to (or larger than) the linewidths of the peaks in the transmission spectrum of the cavity (i.e., K ), the MKN can be probed by measuring the transmission spectrum of the cavity [55– 63], where the linewidths are comparable to the decay rates of cavity mode and magnon modes. However, in the case of K , it is di cult to probe the MKN in this way [18, 69]. For measuring the magnon frequency shift Kin this circumstance, we feed two weak input fields a(in) 1anda(in) 2with same frequency !pinto the microwave cavity via ports 1 and 2, respectively. Using the input-output formalism [73], we get the equations of motion of the cavity-magnon system as follows: ˙a= i(!c ic)a X j=1;2 igjbj q 2ja(in) je i!pt +p 2cf(in) a; ˙b1= i(!1+ K i
1)b1 ig1a i de i!dt+p 2
1f(in) b1; ˙b2= i(!2 i
2)b2 ig2a+p 2
2f(in) b2; (3) where
1(
2) is the decay rate of the magnon mode b1(b2), the total decay rate c=int+1+2of the cavity mode is composed of the intrinsic decay rate intand the decay rates 1 and2induced by the ports 1 and 2, and f(in) a(f(in) bj) with zero mean valuehf(in) ai=0 (hf(in) bji=0) describes the quantum3 noise from the environment related to the cavity mode (the magnon mode bj). Following the above equations of motion, the expected values haiandhbjisatisfy h˙ai= i(!c ic)hai X j=1;2 igjhbji q 2jha(in) jie i!pt ; h˙b1i= i(!1+ K i
1)hb1i ig1hai i de i!dt; h˙b2i= i(!2 i
2)hb2i ig2hai: (4) In the absence of the two input fields (corresponding to ha(in) 1i=ha(in) 2i=0), we denotehai=Ae i!dtandhbji= Bje i!dt. When the input fields are considered, we assume that the changes of haiandhbjican be expressed as Ae i!pt andBje i!pt, i.e., hai=Ae i!dt+Ae i!pt; hbji=Bje i!dt+Bje i!pt; (5) wherejAj j AjandjBjj j Bjj[56]. This assumption is reasonable, because compared with the drive field, the input fields are very weak and can be treated as a perturbation. Now the magnon frequency shift becomes K=2K1jB1j2. Substi- tuting Eq. (5) into Eq. (4), we have ˙A= i(cd ic)A ig1B1 ig2B2; ˙B1= i(1d+ K i
1)B1 ig1A i d; ˙B2= i(2d i
2)B2 ig2A; (6) and ˙A= i(cp ic)A X j=1;2 igjBj q 2jha(in) ji ; ˙B1= i(1p+ K i
1)B1 ig1A; ˙B2= i(2p i
2)B2 ig2A; (7) wherecd=!c !d(jd=!j !d) is the frequency detuning between the cavity mode (magnon mode j) and the drive field, andcp=!c !p(jp=!j !p) is the frequency detuning between the cavity mode (magnon mode j) and the two in- put fields. Eq. (6) determines the magnon frequency shift K, while Eq. (7) determines the output spectrum of the cavity. According to the input-output theory [73], the output field ha(out) jifrom the port jof the cavity is given by ha(out) ji=q 2jA ha(in) ji: (8) Under the pseudo-Hermitian conditions [cf. Eq. (12) in Sec. III], the CPA may occur by carefully choosing appropri- ate parameters of the two input fields [cf. Eqs. (16) and (17) in Sec. III] [29]. The CPA means that the two input fields are nonzero but there are no output fields, i.e., ha(in) 1i,0 and ha(in) 2i,0 butha(out) 1i=ha(out) 2i=0 [28, 32, 33]. When ha(out) 1i=ha(out) 2i=0, ha(in) ji=q 2jA: (9)Inserting the above relation into Eq. (7) to eliminate ha(in) ji, Eq. (7) can be rewritten as 0BBBBBBB@˙A ˙B1 ˙B21CCCCCCCA= iHe0BBBBBBB@A B1 B21CCCCCCCA; (10) where He=0BBBBBBB@cp+ig g1 g2 g11p+ K i
1 0 g2 02p i
21CCCCCCCA(11) is the e ective non-Hermitian Hamiltonian of the cavity- magnon system. Due to the occurrence of CPA, the cavity mode has an e ective gaing=1+2 int(>0) [28, 29]. III. ENHANCING THE DETECTION SENSITIVITY OF MKN A. The EP3 in the cavity-magnon system In this section, we study the EP3 in the cavity-magnon system when K=0. Usually, the eigenvalues of a non- Hermitian Hamiltonian are complex. However, when the sys- tem parameters satisfy the pseudo-Hermitian conditions [29], g=(1+)
2; 2= 1; 2 1=1+k2 (1+)g2 1
2 2;g1gmin; (12) the eective non-Hermitian Hamiltonian Hein Eq. (11) has the pseudo-Hermiticity and thus can also own either three real eigenvalues or one real and two complex-conjugate eigenval- ues [74–76]. The parameter =
1=
2(k=g2=g1) de- notes the ratio between the decay rates
1and
2(coupling strengths g1andg2),j=!j !cis the frequency detun- ing of the magnon mode jrelative to the cavity mode, and gmin=[(1+)=(1+k2)]1=2
2is the allowed minimal value of the coupling strength g1for ensuring 2 10. For engineering the EP3 under the pseudo-Hermitian con- ditions in Eq. (12), the parameters andkmust satisfy the following constraint [29]: k= 1+2 2+2!3=2 : (13) In the symmetric case of =k=1, the non-Hermitian Hamiltonian Hehas three eigenvalues, 0=cpand = cpq 3g2 1 4
2 2[29]. Obviously, 0is real and indepen- dent of the coupling strength g1and the decay rate
2, while are functions of g1and
2. To have three real eigenvalues, the coupling strength g1should be in the region g1>gEP3, where gEP3=2
2=p 3. For g1=gEP3in particular, the three eigenvalues and 0coalesce to = 0= EP3=cp, and the corresponding three eigenvectors of Healso coalesce4 toji=ji0=jiEP3=1p 3 1; 1+p 3i 2;1 p 3i 2T . This co- alescent point at g1=gEP3is referred to as the EP3. While gming1<gEP3, become complex. For the asymmet- ric case with ,1 and k,1, the expressions of and 0are cumbersome and not shown here, and we only give the coalesced eigenvalues = 0= EP3atg1=gEP3= [2(2+2)1=2=(1+2)]
2, where [29] EP3=cp p 3( 1) 22+5+2
2: (14) At the EP3, the three eigenvectors of Hecoalesce to jiEP3=1p N0BBBBB@1; 2p 2+2p 3 i(1+2);2p 2+1p 3+i(2+)1CCCCCAT ;(15) with the normalization factor N=(22+5+2)=(2++1), i.e.,ji=ji0=jiEP3. Note that the results in Eqs. (14) and (15) are also valid for the symmetric case of =k=1. As stated in Sec. II, the e ective non-Hermitian Hamil- tonian Hein Eq. (11) is obtained in the presence of CPA. For engineering the CPA in the pseudo-Hermitian conditions in Eq. (12), the strengths of the two input fields should sat- isfy [29] ha(in) 2i ha(in) 1i=r2 1: (16) In addition, the same frequency of the two input fields need to be equal to the real eigenvalues of He[29], i.e., !(CPA) p= ;0when Im[ ;0]=0: (17) Therefore, the eigenvalues and the EP3 of the pseudo- Hermitian cavity-magnon system can be probed by measur- ing the CPA via the output spectrum of the cavity in experi- ments [28, 32, 33]. B. Eigenvalue response to the MKN near the EP3 Here we investigate the eigenvalue response to the MKN in YIG 1 near the EP3. Considering the magnon frequency shift K(,0), the three eigenvalues of the cavity-magnon system can be obtained by solving the corresponding characteristic equation jHe Ij=0; (18) with an identity matrix I. Because the magnon frequency shiftKis much smaller than other parameters of the cavity- magnon system, we can perturbatively expand the eigenvalue near the EP3 as = EP3+11=3
2+22=3
2 (19) using a Newton-Puiseux series [77–79], where only the first two terms are considered, and EP3is given in Eq. (14). The 0.0 0.1 0.2 0.3-1.0-0.50.00.51.01.5 0.0 0.1 0.2 0.3-1.0-0.50.00.51.01.50.0 0.1 0.2 0.3-1.0-0.50.00.51.01.5 0.0 0.1 0.2 0.3-1.0-0.50.00.51.01.5(a) (b)(c) (d) FIG. 2. The changes of the real and imaginary parts of the eigen- values and 0, (Re[ ;0] EP3)=
2and Im[ ;0]=
2, versus the magnon frequency shift K=
2near the EP3, where =1 in (a,b), while=2 in (c,d). In (a)–(d), the thick curves correspond to the numerical results obtained by numerically solving the characteristic equation in Eq. (18), and the thin curves correspond to the analytical results in Eq. (23). Note that the thick curves almost overlap the thin curves in (b,d). coecients1and2are complex, while = K=
2(1) is real. With Eq. (19), the characteristic equation of the cavity- magnon system in Eq. (18) can be expressed as f1+f4=34=3+f5=35=3+f22+f7=37=3=0; (20) where the coe cients are f1=3 1 42(1 p 3i) 1+2; f4=3=32 12 2[p 3 i(1+2)] 1+21; f5=3=312 2 22 1 2[p 3 i(1+2)] 1+22; f2=3 2 412; f7=3= 22 2: (21) Since4=35=327=3, we can ignore the contri- butions from the last three terms in Eq. (20), and Eq. (20) is reduced to f1+f4=34=3=0. To ensure the relation f1+f4=34=3=0 is valid for any , the coe cients f1and f4=3must be zero, i.e., f1=f4=3=0. Solving f1=f4=3=0, we obtain three sets of solutions for the coe cients1and2, (l) 1= 82 1+2!1=3 eil; (l) 2=2[p 3 i(1+2)] 3(1+2)(l) 1; (22) with l=;0, where+=17=9, =11=9, and0=5=9. Now the three complex eigenvalues of the cavity-magnon sys-5 tem read += EP3+(+) 11=3
2+(+) 22=3
2; 0= EP3+(0) 11=3
2+(0) 22=3
2; = EP3+( ) 11=3
2+( ) 22=3
2: (23) Clearly, the changes of the eigenvalues, ;0 EP3, are pro- portional to 1=3in the case of 1, i.e., ;0 EP3 (;0) 11=3
2. By numerically solving the characteristic equation in Eq. (18), we further study the eigenvalue response to the MKN near the EP3 when K=
2<0:3. In the symmetric case of =1, we plot the changes of the real and imaginary parts of and 0, (Re[ ;0] EP3)=
2and Im[ ;0]=
2, as func- tions of magnon frequency shift K=
2(i.e.,) in Figs. 2(a) and 2(b), where the thick curves correspond to the numerical results, and the thin curves correspond to the analytical results in Eq. (23). The analytical results and the numerical results are almost consistent for K=
2<0:1, while the analytical results deviate from the numerical results when K=
2>0:1 because the condition K=
21 has been used in deriving Eq. (23). Obviously, (Re[ ;0] EP3)=
2and Im[ ;0]=
2 versus K=
2sharply change. This is because the small fre- quency shift Kis amplified by the EP3 [50, 51]. In the re- gion1, (Re[ ;0] EP3)=
2and Im[ ;0]=
2follow the cube-root of , i.e., (Re[ ;0] EP3)=
2Re[(;0) 1]1=3 and Im[ ;0]=
2Im[(;0) 1]1=3. It is very di erent from the existing approach of measuring MKN, where the energy split- ting follows a dependence [55–57]. Further, we find that the amplification e ect is more significant for a larger value of [cf. Figs. 2(a) and 2(c); Figs. 2(b) and 2(d)], which results from the monotonous increase of j(l) 1j=[82=(1+2)]1=3ver- sus. Considering the experimentally accessible parameters, we choose 13 in our study [1, 28, 29]. This amplifica- tion e ect of the EP3 can be used to measure the MKN in the case of K=
2<1 (cf. Sec. IV). IV . MEASURING THE MKN VIA THE OUTPUT SPECTRUM OF THE CA VITY In the cavity-magnon system, we can measure the eigen- value response to the MKN via the output spectrum of the cavity [28, 29]. In the theory, the output spectrum can be de- rived using Eqs. (7) and (8). At the steady state, we solve Eq. (7) with ˙A=˙B1=˙B2=0 and obtain the change Aof the cavity fieldhaidue to the two input fields, A=p21ha(in) 1i+p22ha(in) 2i c+icp+P(!p); (24) where X (!p)=g2 1
1+i(1p+ K)+g2 2
2+i2p(25) -2 -1 0 1 2-200-150-100-500 (a) -2 -1 0 1 2-50-40-30-20-10 (b)(dB) (dB)FIG. 3. (a) The output spectrum jS(!p)j2of the cavity at the EP3, where K=0. (b) The output spectrum jS(!p)j2of the cavity near the EP3 when K,0 (e.g., K=
2=0:01). The (red) dashed ver- tical lines in (b) highlight the locations of the two dips in the output spectrum. Other parameters are chosen to be
1=
2=1,int=
2=1, and1=
2=2=
2=1:5. is the self-energy. Correspondingly, the two output fields ha(out) 1iandha(out) 2iin Eq. (8) can be expressed as ha(out) 1i=21ha(in) 1i+2p12ha(in) 2i c+icp+P(!p) ha(in) 1i; ha(out) 2i=2p12ha(in) 1i+22ha(in) 2i c+icp+P(!p) ha(in) 2i: (26) It follows from Eq. (26) that ha(out) 1i=S(!p)ha(in) 1iand ha(out) 2i=S(!p)ha(in) 2iunder the constraint in Eq. (16), where S(!p)=21+22 c+icp+P(!p) 1 (27) is the output spectrum of the microwave cavity. It can be easily verified that in the case of K=0, the output spec- trum S(!p) is zero [i.e., S(!p)=0] when the system pa- rameters satisfy the pseudo-Hermitian conditions in Eq. (12) and the same frequency of the two input fields is given in Eq. (17) [29]. At the EP3, the three eigenvalues and 0of the cavity- magnon system coalesce to EP3, and the CPA occurs at !(CPA) p= EP3, i.e., there is only one CPA point with jS(!p)j= 0 in the output spectrum [see Fig. 3(a)]. In the presence of the MKN (i.e., K,0), the CPA disappears, and there are6 0.00 0.05 0.10 0.15 0.20 0.25 0.300.00.51.01.52.0(a) 0.00 0.05 0.10 0.15 0.20 0.25 0.301101001000 (b) 10-510-410-310-210-110-210-1100 FIG. 4. (a) The distance !p=
2between the two dips in the output spectrum of the cavity versus the magnon frequency shift K=
2for dierent. The inset displays the logarithmic relationship between !p=
2andK=
2for di erent, where the three (violet) thin curves with a same slope of 1 =3 serve as guides to the eyes. (b) Detection sensitivity enhancement factor !p=Kversus the magnon frequency shift K=
2for di erent. Here=1 for the (black) solid curve, =2 for the (red) dashed curve, and =3 for the (blue) dotted curve. Other parameters are chosen to be
1=
2=,int=
2=1, and 1=
2=2=
2=1+0:5. two dips in the output spectrum highlighted by the two (red) dashed vertical lines in Fig. 3(b). The locations and linewidths of the dips in the output spectrum are determined by the real and imaginary parts of the complex eigenvalues of the cavity- magnon system given in Eq. (23). The left dip at !(dip1) p Re[ ] (right dip at !(dip2) pRe[ +]) corresponds to the eigenvalue ( +). Note that because jIm[ 0]j>jIm[ ]j [cf. Figs. 2(b) and 2(d)], there is no dip in the output spec- trum corresponding to the eigenvalue 0. Therefore, we can measure the MKN by the output spectrum of the cavity. To characterize the detection sensitivity enhancement of MKN near the EP3, we introduce an experimentally measur- able quantity !p=!(dip2) p !(dip1) p; (28) which presents the distance between the two dips in the out- put spectrum of the cavity. By numerically solving the output spectrum S(!p) in Eq. (27), we plot the frequency di erence !p=
2as a function of the magnon frequency shift K=
2 for di erent values of in Fig. 4(a), where !p=
2increases monotonically with K=
2. Obviously, for a given value of K=
2, the corresponding frequency di erence!p=
2be-tween the two dips is far larger than the magnon frequency shift K=
2, i.e.,!pK. In contrast, the frequency dif- ference induced by Kis approximately equal to Kin the existing approach of measuring MKN [55–57]. This means that the magnon frequency shift Kis amplified by the EP3. For su ciently small K=
2,!pfollows a ( K=
2)1=3depen- dence [see the inset in Fig. 4(a)]. Especially, for a larger value of, the amplification e ect of the EP3 is more significant. Moreover, we also display the detection sensitivity enhance- ment factor !p=Kversus the magnon frequency shift K=
2 in Fig. 4(b), where !p=Kmonotonically decreases for dif- ferent. In the region K=
21,!p=Kis proportional to (K=
2) 2=3. When K=
2tends to 0, the sensitivity enhance- ment factor !p=Ktends to infinity, i.e., !p=Kdiverges at K=
2=0. V . DISCUSSIONS AND CONCLUSIONS In our study, all results are based on the equations of motion in Eq. (4), which describes the average behavior of the cavity- magnon system in the mean-field approximation by neglecting the impacts of noises [including classical noise related to fluc- tuations of system parameters and quantum noise related to termsp2cf(in) aandp2
jf(in) bjin Eq. (3)] and quantum fluc- tuations (related to a=a haiandbj=bj hbji). Using Eq. (4), we investigate the detection sensitivity enhancement of MKN by deriving the e ective non-Hermitian Hamiltonian Heof the cavity-magnon system in Eq. (11) and the output spectrum S(!p) of the microwave cavity in Eq. (27). This pro- cedure is widely applied in studying EP-based sensors [50– 54], and the related theoretical predictions have been demon- strated experimentally in various physical systems [80]. For example, the detection sensitivity enhancement factor of 23 has been realized experimentally in a ternary micro-ring sys- tem [77]. However, in the region with the signal being comparable to the noises and quantum fluctuations, the impacts of noises and quantum fluctuations on the EP-based sensor should be considered [80]. The classical noise caused by the techni- cal limitation can reduce the resolvability of frequency di er- ence!pby broadening the linewidth of the output spectrum S(!p) [81, 82]. In principle, the classical noise can be made arbitrarily small in the cavity-magnon system. Di erent from the classical noise, the quantum noise cannot be made arbi- trarily small owing to the vacuum noise. Due to the quantum noise and quantum fluctuations, the diverging sensitivity en- hancement factor [cf. Fig. 4(b) and related discussions] does not necessarily lead to arbitrary high measurement precision, where the measurement precision refers to the smallest mea- surable change of signal [83–86]. This is because the EP- based sensor is sensitive to not only the signal but also the quantum noise, and thus the quantum-limited signal-to-noise ratio cannot be improved [80]. Following the procedures in Refs. [84–86], one can derive the upper bound of the signal- to-noise ratio by calculating the quantum Fisher information based on Heisenberg-Langevin equations in Eq. (3). For the MKN term K1by 1b1by 1b1, the corresponding e ective Hamil-7 tonian for quantum fluctuations can be expressed as Hflu= 2Kby 1b1+by 1by 1+b1b1with=K1hb1i2[64–66]. The two-magnon terms by 1by 1andb1b1can squeeze the quantum fluctuations of magnon mode b1, which can be transferred to cavity mode aand magnon mode b2via their interactions and leads to the squeezing of cavity mode aand magnon mode b2[66]. The squeezing of quantum fluctuations induced by MKN may be helpful for improving the measure- ment precision [87, 88]. Before concluding, we briefly analyze the experimental fea- sibility of the present scheme. In cavity magnonics, both the intrinsic decay rate of the 3D microwave cavity as well as the decay rate of the magnon mode are of the order 1 MHz (i.e., int=21 MHz and
1;2=21 MHz) [1], while the decay rates1;2due to the two ports of the cavity can be tuned from 0 to 8 MHz [28]. Since the frequency of the magnon mode in the YIG is proportional to the bias magnetic field, the fre- quencies!1;2can be easily controlled [8, 60]. In Ref. [28], the EP2 based on CPA has been observed, where the cavity- magnon coupling can be adjusted (ranging from 0 to 9 MHz) via moving the YIG sphere, and the relative amplitudes (rela- tive phases) of the two input fields, ha(in) 1iandha(in) 2i, are also tunable via a variable attenuator (a phase shifter). In addi- tion, the magnon frequency shift Kcaused by the MKN isdependent on the strength of the drive field on the magnon mode [55, 57, 58]. These available conditions ensure that our scheme in the present work is experimentally accessible. In conclusion, we have presented a feasible scheme to en- hance the detection sensitivity of MKN via the CPA around an EP3. In the proposed scheme, the cavity-magnon system con- sists of a 3D microwave cavity and two YIG spheres. With the assistance of the CPA, an e ective pseudo-Hermitian Hamil- tonian of the cavity-magnon system can be obtained, which makes it possible to engineer the EP3 in the parameter space. Considering the magnon frequency shift caused by the MKN, we find that it can be amplified by the EP3. Moreover, we show that this amplification e ect can be measured using the output spectrum of the 3D cavity. Our proposal paves a way to measure the MKN in the case of K . ACKNOWLEDGMENTS This work is supported by the National Natural Science Foundation of China (Grant No. 12205069) and the key pro- gram of the Natural Science Foundation of Anhui (Grant No. KJ2021A1301). [1] D. Lachance-Quirion, Y . Tabuchi, A. Gloppe, K. Usami, and Y . Nakamura, Hybrid quantum systems based on magnonics, Appl. Phys. Express 12, 070101 (2019). [2] H. Y . Yuan, Y . Cao, A. Kamra, R. A. Duine, and P. Yan, Quan- tum magnonics: When magnon spintronics meets quantum in- formation science, Phys. Rep. 965, 1 (2022). [3] B. Z. Rameshti, S. V . Kusminskiy, J. A. Haigh, K. Usami, D. Lachance-Quirion, Y . Nakamura, C. M. Hu, H. X. Tang, G. E. W. Bauer, and Y . M. Blanter, Cavity magnonics, Phys. Rep. 979, 1 (2022). [4] Y . Tabuchi, S. Ishino, T. Ishikawa, R. Yamazaki, K. Usami, and Y . Nakamura, Hybridizing Ferromagnetic Magnons and Mi- crowave Photons in the Quantum Limit, Phys. Rev. Lett. 113, 083603 (2014). [5] X. Zhang, C. L. Zou, L. Jiang, and H. X. Tang, Strongly Cou- pled Magnons and Cavity Microwave Photons, Phys. Rev. Lett. 113, 156401 (2014). [6] M. Goryachev, W. G. Farr, D. L. Creedon, Y . Fan, M. Kostylev, and M. E. Tobar, High-Cooperativity Cavity QED with Magnons at Microwave Frequencies, Phys. Rev. Applied 2, 054002 (2014). [7] D. Zhang, X. M. Wang, T. F. Li, X. Q. Luo, W. Wu, F. Nori, and J. Q. You, Cavity quantum electrodynamics with ferromagnetic magnons in a small yttrium-iron-garnet sphere, npj Quantum Inf.1, 15014 (2015). [8] X. Zhang, C. L. Zou, N. Zhu, F. Marquardt, L. Jiang, and H. X. Tang, Magnon dark modes and gradient memory, Nat. Com- mun. 6, 8914 (2015) [9] L. Bai, M. Harder, Y . P. Chen, X. Fan, J. Q. Xiao, and C. M. Hu, Spin Pumping in Electrodynamically Coupled Magnon-Photon Systems, Phys. Rev. Lett. 114, 227201 (2015). [10] L. Bai, M. Harder, P. Hyde, Z. Zhang, C. M. Hu, Y . P. Chen, and J. Q. Xiao, Cavity Mediated Manipulation of Distant Spin Cur-rents Using a Cavity-Magnon-Polariton, Phys. Rev. Lett. 118, 217201 (2017). [11] J. Li, S. Y . Zhu, and G. S. Agarwal, Magnon-Photon-Phonon Entanglement in Cavity Magnomechanics, Phys. Rev. Lett. 121, 203601 (2018). [12] Z. X. Liu, H. Xiong, and Y . Wu, Magnon blockade in a hy- brid ferromagnet-superconductor quantum system, Phys. Rev. B100, 134421 (2019). [13] J. K. Xie, S. L. Ma, and F. L. Li, Quantum-interference- enhanced magnon blockade in an yttrium-iron-garnet sphere coupled to superconducting circuits, Phys. Rev. A 101, 042331 (2020). [14] Y . Wang, W. Xiong, Z. Xu, G. Q. Zhang, and J. Q. You, Dissipation-induced nonreciprocal magnon blockade in a magnon-based hybrid system, Sci. China-Phys. Mech. Astron. 65, 260314 (2022). [15] M. Harder, L. Bai, P. Hyde, and C. M. Hu, Topological prop- erties of a coupled spin-photon system induced by damping, Phys. Rev. B 95, 214411 (2017). [16] Y . Cao and P. Yan, Exceptional magnetic sensitivity of PT- symmetric cavity magnon polaritons, Phys. Rev. B 99, 214415 (2019). [17] J. Zhao, Y . Liu, L. Wu, C. K. Duan, Y . Liu, and J. Du, Obser- vation of Anti-PT-Symmetry Phase Transition in the Magnon- Cavity-Magnon Coupled System, Phys. Rev. Appl. 13, 014053 (2020). [18] B. Yao, Y . S. Gui, J. W. Rao, S. Kaur, X. S. Chen, W. Lu, Y . Xiao, H. Guo, K. P. Marzlin, and C. M. Hu, Cooperative po- lariton dynamics in feedback-coupled cavities, Nat. Commun. 8, 1437 (2017). [19] X. L. Hei, X. L. Dong, J. Q. Chen, C. P. Shen, Y . F. Qiao, and P. B. Li, Enhancing spin-photon coupling with a micromagnet, Phys. Rev. A 103, 043706 (2021).8 [20] H. Y . Yuan, P. Yan, S. Zheng, Q. Y . He, K. Xia, and M.-H. Yung, Steady Bell State Generation via Magnon-Photon Cou- pling, Phys. Rev. Lett. 124, 053602 (2020). [21] F. X. Sun, S. S. Zheng, Y . Xiao, Q. Gong, Q. He, and K. Xia, Remote generation of magnon Schr ¨odinger cat state via magnon-photon entanglement, Phys. Rev. Lett. 127, 087203 (2021). [22] G. Q. Zhang, W. Feng, W. Xiong, Q. P. Su, and C. P. Yang, Gen- eration of long-lived Wstates via reservoir engineering in dis- sipatively coupled systems, Phys. Rev. A 107, 012410 (2023). [23] S. F. Qi and J. Jing, Generation of Bell and Greenberger-Horne- Zeilinger states from a hybrid qubit-photon-magnon system, Phys. Rev. A 105, 022624 (2022). [24] R. Hisatomi, A. Osada, Y . Tabuchi, T. Ishikawa, A. Noguchi, R. Yamazaki, K. Usami, and Y . Nakamura, Bidirectional conver- sion between microwave and light via ferromagnetic magnons, Phys. Rev. B 93, 174427 (2016). [25] N. Zhu, X. Zhang, X. Han, C. L. Zou, C. Zhong, C. H. Wang, L. Jiang, and H. X. Tang, Waveguide cavity optomagnonics for broadband multimode microwave-to-optics conversion, Optica 7, 1291 (2020). [26] V . L. Grigoryan, K. Shen, and K. Xia, Synchronized spin- photon coupling in a microwave cavity, Phys. Rev. B. 98, 024406 (2018). [27] M. Harder, Y . Yang, B. M. Yao, C. H. Yu, J. W. Rao, Y . S. Gui, R. L. Stamps, and C. M. Hu, Level Attraction Due to Dissi- pative Magnon-Photon Coupling, Phys. Rev. Lett. 121, 137203 (2018). [28] D. Zhang, X. Q. Luo, Y . P. Wang, T. F. Li, and J. Q. You, Ob- servation of the exceptional point in cavity magnon-polaritons, Nat. Commun. 8, 1368 (2017). [29] G. Q. Zhang and J. Q. You, Higher-order exceptional point in a cavity magnonics system, Phys. Rev. B 99, 054404 (2019). [30] Y . D. Chong, L. Ge, H. Cao, and A. D. Stone, Coherent Per- fect Absorbers: Time-Reversed Lasers, Phys. Rev. Lett. 105, 053901 (2010). [31] W. Wan, Y . Chong, L. Ge, H. Noh, A. D. Stone, and H. Cao, Time-reversed lasing and interferometric control of absorption, Science 331, 889 (2011). [32] Y . Sun, W. Tan, H. Q. Li, J. Li, and H. Chen, Experimental Demonstration of a Coherent Perfect Absorber with PT Phase Transition, Phys. Rev. Lett. 112, 143903 (2014). [33] C. Wang, W. R. Sweeney, A. D. Stone, and L. Yang, Coherent perfect absorption at an exceptional point, Science 373, 1261 (2021). [34] Z. J. Wong, Y . L. Xu, J. Kim, K. O’Brien, Y . Wang, L. Feng, and X. Zhang, Lasing and anti-lasing in a single cavity, Nat. Photonics 10, 796 (2016). [35] K. Pichler, M. K ¨uhmayer, J. B ¨ohm, A. Brandst ¨otter, P. Am- bichl, U. Kuhl, and S. Rotter, Random anti-lasing through co- herent perfect absorption in a disordered medium, Nature (Lon- don) 567, 351 (2019). [36] X. Fang, M. L. Tseng, J. Y . Ou, K. F. MacDonald, D. P. Tsai, and N. I. Zheludev, Ultrafast all-optical switching via coherent modulation of metamaterial absorption, Appl. Phys. Lett. 104, 141102 (2014). [37] W. Xiong, J. Chen, B. Fang, C. H. Lam, and J. Q. You, Coher- ent perfect absorption in a weakly coupled atom-cavity system, Phys. Rev. A 101, 063822 (2020). [38] M. Kang and Y . D. Chong, Coherent optical control of polar- ization with a critical metasurface, Phys. Rev. A 92, 043826 (2015). [39] Y . Ye, D. Hay, and Z. Shi, Coherent perfect absorption in chiral metamaterials, Opt. Lett. 41, 3359 (2016).[40] W. D. Heiss, The physics of exceptional points, J. Phys. A Math. Theor. 45, 444016 (2012). [41] T. Gao, E. Estrecho, K. Y . Bliokh, T. C. H. Liew, M. D. Fraser, S. Brodbeck, M. Kamp, C. Schneider, S. H ¨ofling, Y . Yamamoto, F. Nori, Y . S. Kivshar, A. G. Truscott, R. G. Dall, and E. A. Ostrovskaya, Observation of non-Hermitian degeneracies in a chaotic exciton-polariton billiard, Nature (London) 526, 554 (2015). [42] X. Y . L ¨u, H. Jing, J. Y . Ma, and Y . Wu, PT-Symmetry- Breaking Chaos in Optomechanics, Phys. Rev. Lett. 114, 253601 (2015). [43] G. Q. Zhang, Z. Chen, D. Xu, N. Shammah, M. Liao, T. F. Li, L. Tong, S. Y . Zhu, F. Nori, and J. Q. You, Exceptional point and cross-relaxation e ect in a hybrid quantum system, PRX Quantum 2, 020307 (2021). [44] H. Jing, S ¸. K. ¨Ozdemir, H. L ¨u, and F. Nori, High-order excep- tional points in optomechanics, Sci. Rep. 7, 3386 (2017). [45] T. X. Lu, H. Zhang, Q. Zhang, and H. Jing, Exceptional-point- engineered cavity magnomechanics, Phys. Rev. A 103, 063708 (2021). [46] W. Xiong, Z. Li, G. Q. Zhang, M. Wang, H. C. Li, X. Q. Luo, and J. Chen, Higher-order exceptional point in a blue- detuned non-Hermitian cavity optomechanical system, Phys. Rev. A 106, 033518 (2022). [47] J. Doppler, A. A. Mailybaev, J. B ¨ohm, U. Kuhl, A. Girschik, F. Libisch, T. J. Milburn, P. Rabl, N. Moiseyev, and S. Rotter, Dy- namically encircling an exceptional point for asymmetric mode switching, Nature (London) 537, 76 (2016). [48] J. Q. Zhang, J. X. Liu, H. L. Zhang, Z. R. Gong, S. Zhang, L. L. Yan, S. L. Su, H. Jing, and M. Feng, Topological optome- chanical amplifier in synthetic PT-symmetry, Nanophotonics 11, 1149 (2022). [49] Y . Zhiyenbayev, Y . Kominis, C. Valagiannopoulos, V . Kovanis, and A. Bountis, Enhanced stability, bistability, and exceptional points in saturable active photonic couplers, Phys. Rev. A 100, 043834 (2019). [50] J. Wiersig, Enhancing the Sensitivity of Frequency and Energy Splitting Detection by Using Exceptional Points: Application to Microcavity Sensors for Single-Particle Detection, Phys. Rev. Lett. 112, 203901 (2014). [51] W. Chen, S ¸. K. ¨Ozdemir, G. Zhao, J. Wiersig, and L. Yang, Exceptional points enhance sensing in an optical microcavity, Nature (London) 548, 192 (2017). [52] Z. P. Liu, J. Zhang, S ¸. K. ¨Ozdemir, B. Peng, H. Jing, X. Y . L ¨u, C. W. Li, L. Yang, F. Nori, and Y . X. Liu, Metrology with PT- Symmetric Cavities: Enhanced Sensitivity near the PT-Phase Transition, Phys. Rev. Lett. 117, 110802 (2016). [53] G. Q. Zhang, Y . P. Wang, and J. Q. You, Dispersive readout of a weakly coupled qubit via the parity-time-symmetric phase transition, Phys. Rev. A 99, 052341 (2019). [54] X. G. Wang, G. H. Guo, and J. Berakdar, Enhanced Sensitiv- ity at Magnetic High-Order Exceptional Points and Topological Energy Transfer in Magnonic Planar Waveguides, Phys. Rev. Appl. 15, 034050 (2021). [55] Y . P. Wang, G. Q. Zhang, D. Zhang, X. Q. Luo, W. Xiong, S. P. Wang, T. F. Li, C. M. Hu, and J. Q. You, Magnon Kerr e ect in a strongly coupled cavity-magnon system, Phys. Rev. B 94, 224410 (2016). [56] G. Q. Zhang, Y . P. Wang, and J. Q. You, Theory of the magnon Kerr e ect in cavity magnonics, Sci. China-Phys. Mech. As- tron. 62, 987511 (2019). [57] Y . P. Wang, G. Q. Zhang, D. Zhang, T. F. Li, C. M. Hu, and J. Q. You, Bistability of Cavity Magnon-Polaritons, Phys. Rev. Lett. 120, 057202 (2018).9 [58] R. C. Shen, J. Li, Z. Y . Fan, Y . P. Wang, and J. Q. You, Me- chanical Bistability in Kerr-modified Cavity Magnomechanics, Phys. Rev. Lett. 129, 123601 (2022). [59] J. M. P. Nair, D. Mukhopadhyay, and G. S. Agarwal, Ultralow threshold bistability and generation of long-lived mode in a dis- sipatively coupled nonlinear system: Application to magnonics, Phys. Rev. B 103, 224401 (2021). [60] R. C. Shen, Y . P. Wang, J. Li, S. Y . Zhu, G. S. Agarwal, and J. Q. You, Long-Time Memory and Ternary Logic Gate Using a Multistable Cavity Magnonic System, Phys. Rev. Lett. 127, 183202 (2021). [61] J. M. P. Nair, Z. Zhang, M. O. Scully, and G. S. Agarwal, Non- linear spin currents, Phys. Rev. B 102, 104415 (2020). [62] M. X. Bi, X. H. Yan, Y . Zhang, and Y . Xiao, Tristability of cavity magnon polaritons, Phys. Rev. B 103, 104411 (2021). [63] C. Kong, H. Xiong, and Y . Wu, Magnon-Induced Nonreciproc- ity Based on the Magnon Kerr E ect, Phys. Rev. Appl. 12, 034001 (2019). [64] W. Xiong, M. Tian, G. Q. Zhang, and J. Q. You, Strong long- range spin-spin coupling via a Kerr magnon interface, Phys. Rev. B 105, 245310 (2022). [65] Z. Zhang, M. O. Scully, and G. S. Agarwal, Quantum entangle- ment between two magnon modes via Kerr nonlinearity driven far from equilibrium, Phys. Rev. Research 1, 023021 (2019). [66] Z. B. Yang, H. Jin, J. W. Jin, J. Y . Liu, H. Y . Liu, and R. C. Yang, Bistability of squeezing and entanglement in cavity magnonics, Phys. Rev. Research 3, 023126 (2021). [67] G. Q. Zhang, Z. Chen, W. Xiong, C. H. Lam, and J. Q. You, Parity-symmetry-breaking quantum phase transition via para- metric drive in a cavity magnonic system, Phys. Rev. B 104, 064423 (2021). [68] Y . Qin, S. C. Li, K. Li, and J. J. Song, Controllable quantum phase transition in a double-cavity magnonic system, Phys. Rev. B106, 054419 (2022). [69] J. A. Haigh, N. J. Lambert, A. C. Doherty, and A. J. Fergu- son, Dispersive readout of ferromagnetic resonance for strongly coupled magnons and microwave photons, Phys. Rev. B 91, 104410 (2015). [70] J. M. P. Nair, D. Mukhopadhyay, and G. S. Agarwal, Enhanced Sensing of Weak Anharmonicities through Coherences in Dissi- patively Coupled Anti-PT Symmetric Systems, Phys. Rev. Lett. 126, 180401 (2021). [71] A. G. Gurevich and G. A. Melkov, Magnetization Oscillations and Waves (CRC Press, Boca Raton, 1996). [72] D. D. Stancil and A. Prabhakar, Spin Waves (Springer, Berlin, 2009). [73] D. F. Walls and G. J. Milburn, Quantum Optics (Springer, Berlin, 1994).[74] A. Mostafazadeh, Pseudo-Hermiticity versus PT-symmetry: The necessary condition for the reality of the spectrum of a non- Hermitian Hamiltonian, J. Math. Phys. 43, 205 (2002). [75] A. Mostafazadeh, Pseudo-Hermiticity versus PT-symmetry II: A complete characterization of non-Hermitian Hamiltonians with a real spectrum, J. Math. Phys. 43, 2814 (2002). [76] A. Mostafazadeh, Pseudo-Hermiticity versus PT-symmetry III: Equivalence of pseudo-Hermiticity and the presence of an- tilinear symmetries, J. Math. Phys. 43, 3944 (2002). [77] H. Hodaei, A. U. Hassan, S. Wittek, H. Garcia-Gracia, R. El-Ganainy, D. N. Christodoulides, and M. Khajavikhan, En- hanced sensitivity at higher-order exceptional points, Nature (London) 548, 187 (2017). [78] W. Xiong, Z. Li, Y . Song, J. Chen, G. Q. Zhang, and M. Wang, Higher-order exceptional point in a pseudo-Hermitian cavity optomechanical system, Phys. Rev. A 104, 063508 (2021). [79] C. Zeng, K. Zhu, Y . Sun, G. Li, Z. Guo, J. Jiang, Y . Li, H. Jiang, Y . Yang, and H. Chen, Ultra-sensitive passive wireless sensor exploiting high-order exceptional point for weakly cou- pling detection, New J. Phys. 23, 063008 (2021). [80] J. Wiersig, Review of exceptional point-based sensors, Photon- ics Res. 8, 1457 (2020). [81] C. Wol , C. Tserkezis, and N. A. Mortensen, On the time evo- lution at a fluctuating exceptional point, Nanophotonics 8, 1319 (2019). [82] J. Wiersig, Robustness of exceptional-point-based sensors against parametric noise: the role of Hamiltonian and Liouvil- lian degeneracies, Phys. Rev. A 101, 053846 (2020). [83] W. Langbein, No exceptional precision of exceptional-point sensors, Phys. Rev. A 98, 023805 (2018). [84] H. K. Lau and A. A. Clerk, Fundamental limits and non- reciprocal approaches in non-Hermitian quantum sensing, Nat. Commun. 9, 4320 (2018). [85] C. Chen, L. Jin, and R. B. Liu, Sensitivity of parameter estima- tion near the exceptional point of a non-Hermitian system, New J. Phys. 21, 083002 (2019). [86] M. Zhang, W. Sweeney, C. W. Hsu, L. Yang, A. D. Stone, and L. Jiang, Quantum noise theory of exceptional point amplifying sensors, Phys. Rev. Lett. 123, 180501 (2019). [87] I. Kruse, K. Lange, J. Peise, B. L ¨ucke, L. Pezz `e, J. Arlt, W. Ertmer, C. Lisdat, L. Santos, A. Smerzi, and C. Klempt, Im- provement of an Atomic Clock using Squeezed Vacuum, Phys. Rev. Lett. 117, 143004 (2016). [88] M. Malnou, D. A. Palken, B. M. Brubaker, L. R. Vale, G. C. Hilton, and K. W. Lehnert, Squeezed Vacuum Used to Accel- erate the Search for a Weak Classical Signal, Phys. Rev. X 9, 021023 (2019). | 2022-11-16 | We show how to enhance the detection sensitivity of magnon Kerr nonlinearity
(MKN) in cavity magnonics. The considered cavity-magnon system consists of a
three-dimensional microwave cavity containing two yttrium iron garnet (YIG)
spheres, where the two magnon modes (one has the MKN, while the other is
linear) in YIG spheres are simultaneously coupled to microwave photons. To
obtain the effective gain of the cavity mode, we feed two input fields into the
cavity. By choosing appropriate parameters, the coherent perfect absorption of
the two input fields occurs, and the cavity-magnon system can be described by
an effective non-Hermitian Hamiltonian. Under the pseudo-Hermitian conditions,
the effective Hamiltonian can host the third-order exceptional point (EP3),
where the three eigenvalues of the Hamiltonian coalesce into one. When the
magnon frequency shift $\Delta_K$ induced by the MKN is much smaller than the
linewidths $\Gamma$ of the peaks in the transmission spectrum of the cavity
(i.e., $\Delta_K\ll \Gamma$), the magnon frequency shift can be amplified by
the EP3, which can be probed via the output spectrum of the cavity. The scheme
we present provides an alternative approach to measure the MKN in the region
$\Delta_K\ll \Gamma$ and has potential applications in designing low-power
nonlinear devices based on the MKN. | Detection sensitivity enhancement of magnon Kerr nonlinearity in cavity magnonics induced by coherent perfect absorption | 2211.08922v2 |
Electronic control of the spin-wave damping in a magnetic insulator A. Hamadeh,1O. d'Allivy Kelly,2C. Hahn,1H. Meley,1R. Bernard,2A.H. Molpeceres,2V. V. Naletov,1, 2, 3M. Viret,1A. Anane,2V. Cros,2S. O. Demokritov,4J. L. Prieto,5M. Mu~ noz,6G. de Loubens,1and O. Klein1, 1Service de Physique de l' Etat Condens e (CNRS URA 2464), CEA Saclay, 91191 Gif-sur-Yvette, France 2Unit e Mixte de Physique CNRS/Thales and Universit e Paris Sud 11, 1 av. Fresnel, 91767 Palaiseau, France 3Institute of Physics, Kazan Federal University, Kazan 420008, Russian Federation 4Department of Physics, University of Muenster, 48149 Muenster, Germany 5Instituto de Sistemas Optoelectr onicos y Microtecnolog a (UPM), Madrid 28040, Spain 6Instituto de Microelectr onica de Madrid (CNM, CSIC), Madrid 28760, Spain (Dated: November 28, 2021) It is demonstrated that the decay time of spin-wave modes existing in a magnetic insulator can be reduced or enhanced by injecting an in-plane dc current, Idc, in an adjacent normal metal with strong spin-orbit interaction. The demonstration rests upon the measurement of the ferromagnetic resonance linewidth as a function of Idcin a 5m diameter YIG(20nm) jPt(7nm) disk using a magnetic resonance force microscope (MRFM). Complete compensation of the damping of the fun- damental mode is obtained for a current density of 31011A.m 2, in agreement with theoretical predictions. At this critical threshold the MRFM detects a small change of static magnetization, a behavior consistent with the onset of an auto-oscillation regime. The spin-orbit interaction (SOI) [1{3] has been re- cently shown to be an interesting and useful addition in the eld of spintronics. This subject capitalizes on adjoining a strong SOI normal metal next to a thin mag- netic layer [4]. The SOI converts a charge current, Jc, to a spin current, Js, with an eciency parametrized by SH, the spin Hall angle [5, 6]. Recently, it was demon- strated experimentally that the spin current produced in this way can switch the magnetization in a dot [7, 8] or can partially compensate the damping [9{11], allow- ing the lifetime of propagating spin-waves [12] to be in- creased beyond their natural decay time, . These two eects open potential applications in storage devices and in microwave signal processing. The eect is based on the fact that the spin current Js exerts a torque on the magnetization, corresponding to an eective damping s=
Js=(tFMMs), wheretFMis the thickness of the magnetic layer, Msits spontaneous magnetization, and
the gyromagnetic ratio. In the case of metallic ferromagnets [13{15], it was established that scan fully compensate the natural damping 1 =at a critical spin current J s, which determines the onset of auto-oscillation of the magnetization: J s= 1 tFMMs
: (1) An important benet of the SOI is that JcandJsare linked through a cross-product, allowing a charge current
owing in-plane to produce a spin current
owing out- of-plane. Hence it enables the transfer of spin angular momentum to non-metallic materials and in particular to insulating oxides, which oer improved performance compared to their metallic counterparts. Among all ox- ides, Yttrium Iron Garnet (YIG) holds a special place for having the lowest known spin-wave (SW) damping factor. In 2010, Kajiwara et al. reported on the e-cient transmission of spin current through the YIG jPt interface [16]. It was shown that Jsproduced by the excitation of ferromagnetic resonance (FMR) in YIG can cross the YIGjPt interface and be converted into Jcin Pt through the inverse spin Hall eect (ISHE). This nding was reproduced in numerous experimental works [17{23]. In the same paper, the reciprocal eect was also reported asJsproduced in Pt by the direct spin Hall eect (SHE) could be transferred to the 1.3 m thick YIG, resulting in damping compensation. However, attempts to directly measure the expected change of the resonance linewidth of YIG as a function of the dc current have so far failed [21, 22] [24]. This is raising fundamental questions about the reciprocity of the spin transparency, T, of the in- terface between a metal and a magnetic insulator. This coecient enters in the ratio between Jcin Pt andJsin YIG through: Js=TSHh 2eJc; (2) whereeis the electron charge and hthe reduced Planck constant.Tdepends on the transport characteristics of the normal metal as well as on the spin-mixing conduc- tanceG"#, which parametrizes the scattering of the spin angular momentum at the YIG jPt interface [25]. At the heart of this debate lies the exact value of the threshold current. The lack of visible eects reported in Refs.[21, 22], although inconsistent with [16], is co- herent with the estimation of the threshold current of 1011 12A.m 2using Eqs.(1) and (2) and typical pa- rameters for the materials [26]. This theoretical cur- rent density is at least one order of magnitude larger than the maximum Jcthat could be injected in the Pt so far. Importantly, the previous reported experiments were performed on large (millimeter sized) structures, where many nearly degenerate SW modes compete forarXiv:1405.7415v1 [cond-mat.mes-hall] 28 May 20142 TABLE I. Transport and magnetic properties of the Pt and bare YIG layers, respectively from Ref.[31] and Ref.[22]. PttPt(nm)( 1.m 1)SD(nm) SH 7 5:81063.5 0.056 YIGtYIG(nm) 4Ms(G)
(rad.s 1.G 1)0 20 2 :11031:791072:310 4 feeding from the same dc source of angular momentum, a phenomenon that could become self-limiting and pre- vent the onset of auto-oscillations [11]. To isolate a single candidate mode, we have recently reduced the lateral di- mensions of the YIG pattern, as quantization results in increased frequency gaps between the dynamical modes [27]. This requires to grow very thin lms of high qual- ity YIG [23, 28{30]. Beneting from our progress in the epitaxial growth of YIG lms by pulsed laser deposition (PLD) [22], we propose to study the FMR linewidth as a function of the dc current in a micron-size YIG jPt disk. FIG.1 shows a schematic of the experimental setup. A YIGjPt disk of 5 m in diameter is connected to two Au contact electrodes (see the microscopy image) across which a positive voltage generates a current
ow Jcalong the +^x-direction. The microdisk is patterned out of a 20 nm thick epitaxial YIG lm with a 7 nm thick Pt layer sputtered on top. The YIG and Pt layers have been fully characterized in previous studies [22, 31]. Their characteristics are reported in Table I. The sample is mounted inside a room temperature magnetic resonance force microscope (MRFM) which de- tects the SW absorption spectrum mechanically [32{34]. The excitation is provided by a stripline (not shown in the sketches of FIG.1) generating a linearly polarized microwave eld h1along the ^x-direction. The detec- tion is based on monitoring the de
ection of a mechan- ical cantilever with a magnetic Fe particle axed to its tip, coupled dipolarly to the sample. The FMR spec- trum is obtained by recording the vibration amplitude of the cantilever while scanning the external bias magnetic eld,H0, at constant microwave excitation frequency, f=!=(2) [35]. The MRFM is placed between the poles of an electromagnet, generating a uniform magnetic eld, H0, which can be set along ^ yor ^z(i.e., perpendicularly to bothh1andJc). We start by measuring the eect of a dc current, Idc, on the FMR spectra when the disk is magnetized in- plane by a magnetic eld along the +^ y-direction (positive eld). The spectra recorded at f= 6:33 GHz are shown in FIG.1a in red tones. The middle row shows the ab- sorption at zero current. The MRFM signal corresponds to a variation of the static magnetization of about 2 G, i.e., a precession cone of 2.5. As the the electrical cur- rent is varied, we observe very clearly a change of the linewidth. At negative current, the linewidth decreases, FIG. 1. (Color online) MRFM spectra of the YIG jPt mi- crodisk as a function of current for dierent eld orientations: a)H0k+^yatf= 6:33 GHz (red tone); b) H0k+^zat f= 10:33 GHz (black); c) H0k ^yatf= 6:33 GHz (blue tone). The highest amplitude mode is used for linewidth anal- ysis (shaded area). Field axes are shifted so as to align the peaks vertically. In-plane and out-of-plane eld orientations are sketched above. The top right frame is a microscopy im- age of the sample. to reach about half the initial value at Idc= 8 mA. This decrease is strong enough so that the individual modes can be resolved spectroscopically within the main peak. Concomitantly the amplitude of the MRFM sig- nal increases. The opposite behavior is observed when the current polarity is reversed. At positive current, the linewidth increases to reach about twice the initial value atIdc= +8 mA, and the amplitude of the signal de- creases. Idc=12 mA is the maximum current that we have injected in our sample to avoid irreversible eects. We estimate from the Pt resistance, the sample temperature to be 90C at the maximum current. This Joule heating reduces 4M sat a rate of 4 :8 G/K, which results in an even shift of the resonance eld towards higher eld [36]. In FIG.1b, we show the FMR spectra at f= 10:33 GHz in the perpendicular geometry, i.e.,H0is along ^z. In con- trast to the previous case, the linewidth does not change with current. This is expected as no net spin transfer torque is exerted by the spin current on the precessing magnetization in this conguration. Note that due to Joule heating, the spectrum now shifts towards lower eld due to the decrease of Msas the current increases. We now come back to the in-plane geometry, but this time, the magnetic eld is reversed compared to FIG.1a,3 FIG. 2. (Color online) Variation of the full linewidth Hk measured at 6.33 GHz as a function of IdcforH0k+^y(red) andH0k ^y(blue). Inset: detection of VISHE as a function ofH0atf= 6:33 GHz and Idc= 0. i.e., applied along ^y(negative eld). The correspond- ing spectra are presented in FIG.1c using blue tones. As expected for the symmetry of the SHE, the observed be- havior is inverted with respect to FIG.1a: a positive (neg- ative) current now reduces (broadens) the linewidth. We report in FIG.2 the values of Hk, the full linewidth measured in the in-plane geometry, as a func- tion of current. The data points follow approximately a straight line, whose slope 0:5 Oe/mA reverses with the direction of H0along^yand whose intercept with the abscissa axis occurs at I 6.33 GHz =14 mA. More- over, we emphasize that the variation of linewidth covers about a factor ve on the full range of current explored. The inset of FIG.2 shows the inverse spin Hall voltage VISHE measured at Idc= 0 mA and f= 6:33 GHz. This voltage results from the spin current produced by spin pumping from YIG to Pt and its subsequent conversion into charge current by ISHE [16]. Its sign changes with the direction of the bias magnetic eld, as shown by the blue and red VISHE spectra. This observation conrms that a spin current can
ow from YIG to Pt and that damping reduction occurs for a current polarity corre- sponding to a negative product of VISHE andIdc. To gain more insight into these results, we now an- alyze the frequency dependence of the full linewidth at half maximum for three values of dc current (0, 6 mA) for both the out-of-plane and in-plane geometries. We start with the out-of-plane data, plotted in FIG.3a. The dispersion relation displayed in the inset follows the Kit- tel law,!=
(H0 4NeMs), whereNeis an eective demagnetizing factor close to 1 [37, 38]. The linewidth H?increases linearly with frequency along a line that intercepts the origin, a signature that the resonance is homogeneously broadened [27]. In this geometry, the Gilbert damping coecient is simply =
H?=(2!) = 1:110 3and the reaxation time = 1=(!). We also report on this gure the fact that at 10.33 GHz, H?= 7 Oe is independent of the current (see FIG.1b). FIG. 3. (Color online) Frequency dependence of the linewidth for three values of the dc current (0, 6 mA) a) in the per- pendicular geometry and b) in the parallel geometry. Insets show the corresponding dispersion relations f(H0). The damping found in our YIG jPt microdisk is signif- icantly larger than the one measured in the bare YIG lm0= 2:310 4(cf. Table I). This dierence is due to the spin pumping eect, and enables to determine the spin-mixing conductance of our YIG jPt interface through [39, 40]: =0+
h 4M stYIGG"# G0; (3) whereG0= 2e2=his the quantum of conductance. The measured increase of almost 9 10 4for the damping corresponds to G"#= 1:51014 1m 2, in agree- ment with a previous determination made on similar YIGjPt nanodisks [27]. This value allows us to esti- mate the spin transparency of our interface [25], T= G"#=(G"#coth (tPt=sd) +=(2sd))'0:15, whereis the Pt conductivity and sdits spin-diusion length. Moreover, the spin-mixing conductance can be used to analyze quantitatively the dc ISHE voltage produced at resonance [21, 41, 42]. Using the parameters of Table I and the value of G"#, we nd that the 50 nV voltage measured in the inset of FIG.2 is produced by an angle of precession '3:5, which lies in the expected range. We now turn to the in-plane data, presented in FIG.3b. The dispersion relation plotted in the inset follows the Kittel law !=
p H0(H0+ 4NeMs). In this case, 1==(@!=@H 0) (!=
). ForIdc= 0 mA the slope of the linewidth vs. frequency is exactly the same as that in the perpendicular direction = 1:110 3. For this geometry, however, the line does not intercept the origin, indicating a nite amount of inhomogeneous broadening H0= 2:5 Oe, i.e, the presence of several modes within the resonance line. Setting Idcto6 mA shifts Hkby 3 Oe independently of the frequency, which is consistent with the rate of 0.5 Oe/mA reported at 6.33 GHz in FIG.2. In fact, in the presence of the eective damping4 FIG. 4. (Color online) a) Density plot of the MRFM spectra at 4.33 GHz vs. eld and current Idc2[ 12;+12] mA. The color scale represents 4 Mz(white: 0 G, black: 1.5 G). b) Evolution of integrated power vs. Idc. c) Dependence of linewidth on Idc. d) Dierential measurements of Mz(Idc modulated by 0.15 mA pp, no rf excitation) vs. Idcat six dierent values of the in-plane magnetic eld. s, the linewidth of the resonance line varies as Hk= H0+ 2!
+ 2Js MstYIG: (4) This expression is valid when ( @!=@H 0)'
,i.e., at large enough eld or frequency (see inset of FIG.3b). It de- scribes appropriately the experimental data on the whole frequency range measured. In order to investigate the autonomous dynamics of the YIG layer and exceed the compensation current, I, we now perform measurements at lower excitation fre- quency, where the threshold current is estimated below 12 mA. In FIG.4a, we present a density plot of the MRFM spectra acquired at 4.33 GHz as a function of the in-plane magnetic eld and Idcthrough the Pt. The measured signal is clearly asymmetric in Idc. At positive current, it broadens and its amplitude decreases, almost disappearing above +8 mA, whereas at negative current, it becomes narrower and the amplitude is maximal at Idc< |