Spin texture in a bilayer high-temperature cuprate superconductor
Abstract
We investigate the possibility of spin texture in the bilayer cuprate superconductor using cluster dynamical mean field theory (CDMFT). The one-band Hubbard model with a small interlayer hopping and a Rashba spin-orbit coupling is used to describe the material. The -wave order parameter is not much affected by the presence of the Rashba coupling, but a small triplet component appears. We find a spin texture circulating in the same direction around and and stable against the superconducting phase. The amplitude of the spin structure, however, is strongly affected by the pseudogap phenomenon, more so than the spectral function itself.
While electron-electron interactions are a key ingredient in the study of quantum materials, the presence of a spin-orbit coupling (SOC) is the source of new emergent phenomena, especially in heavy transition metal compounds Witczak-Krempa et al. (2014). The SOC is a key ingredient of the topological states of matter Hasan and Kane (2010); Qi and Zhang (2011). The interplay or competition between SOC and electron correlations is relevant in systems like the heavy fermion superlattices Shimozawa et al. (2016), iridium oxides Rau et al. (2016), and optical lattices Manchon et al. (2015), in which exotic phases are expected. Within the Rashba-Hubbard model, a mixed singlet-triplet superconducting state Frigeri et al. (2004); Yokoyama et al. (2007); Yanase and Sigrist (2008); Tada et al. (2008); Greco and Schnyder (2018); Lu and Sénéchal (2018); Ghadimi et al. (2019); Nogaki and Yanase (2020); Wolf and Rachel (2020), novel magnetism Zhang et al. (2015); Greco et al. (2020), and nontrivial topological properties Farrell and Pereg-Barnea (2014); Laubach et al. (2014); Lu and Sénéchal (2018); Marcelino (2017) are theoretically predicted.
The spin-orbit coupling also manifests itself in a globally centrosymmetric crystal, which contains subunits in which the inversion symmetry is broken locally Fischer et al. (2011); Maruyama et al. (2012); Sigrist et al. (2014); Zhang et al. (2014), i.e., a locally non-centrosymmetric crystal. The bilayer SOC system is a typical example, in which the SOC on two non-equivalent layers have opposite signs, such as the hybrid structure Goh et al. (2012), Nishikubo et al. (2011); Sigrist et al. (2014), and bilayer transition metal dichalcogenides (TMDs) Jones et al. (2014); Liu (2017). The absence of local inversion symmetry in the bilayer system can lead to a “hidden” spin polarization Zhang et al. (2014); Riley et al. (2014), nontrivial topological states Nakosai et al. (2012); Das and Balatsky (2013); Dong et al. (2015), and unconventional superconductivity Sigrist et al. (2014); Liu (2017); Ishizuka and Yanase (2018).
Recent spin- and angle-resolved photoemission spectroscopy (SARPES) experiments have shown that, in one of the most studied cuprate superconductors (Bi2212), a striking spin texture develops in the Brillouin zone with spin-momentum locking Gotlieb et al. (2018). The observed spin texture is consistent with the prediction of a bilayer model with opposite Rashba SOC on the two CuO layers of the unit cell. This has motivated new studies focusing on the hidden SOC in high- cuprates Raines et al. (2019); Hitomi and Yanase (2019); Atkinson (2020). However, correlation effects and a possible competition between the spin texture and -wave superconductivity have not been fully considered so far. In this paper, we will address these important issues in a fully dynamical study employing cluster dynamical mean field theory (CDMFT).
Model —
To describe the locally non-centrosymmetric bilayer high- cuprate, we use the following tight-binding bilayer Rashba model Gotlieb et al. (2018); Raines et al. (2019); Hitomi and Yanase (2019); Harrison et al. (2015):
(1) |
The non-interacting part includes three terms,
(2) | |||||
(3) | |||||
(4) |
where is the annihilation (creation) operator of an electron on the th layer () with spin and wave vector . The dispersion relation on a square lattice is , in which the hopping terms up to third nearest neighbor (, , and ) and the chemical potential are included. defines the antisymmetric SOC of Rashba type and is the vector of Pauli matrices. Due to the global inversion symmetry, the SOC in layers 1 ad 2 are opposite in sign: . The interlayer hoping is , which causes the bilayer splitting in high- cuprate. For Bi2212, the nearest neighbor hopping is meV Markiewicz et al. (2005); Drozdov et al. (2018). In the remainder of this paper, we set as the energy unit, and choose the other tight-binding parameters to be , and Drozdov et al. (2018). Finally, we set .
The interacting part of Hamiltonian reads
(5) |
in which is the number of electrons of spin at lattice site of layer . Only on-site interactions are considered.
CDMFT —
In order to reveal the spin texture arising in model (1), we use cluster dynamical mean-field theory (CDMFT) Lichtenstein and Katsnelson (2000); Kotliar et al. (2001); Liebsch et al. (2008); Sénéchal (2015) with an exact diagonalization solver at zero temperature (or ED-CDMFT). In this approach, the infinite lattice is tiled into identical units (or supercells) defining a superlattice. The supercell is made of one or more clusters, each of which coupled to a bath of uncorrelated, auxiliary orbitals. The parameters describing this bath (energy levels, hybridization, etc.) are then found by imposing a self-consistency condition.
In this work the supercell is made of two superimposed, four-site plaquettes (one per layer), each of which coupled to a bath of eight uncorrelated orbitals. The cluster-bath system, or impurity model, is illustrated on Fig. 1 and defined by the following Anderson impurity model (AIM):
(6) |
where is the Hamiltonian (1), but restricted to a single cluster, and and destroy electrons on the cluster sites and the bath orbitals, respectively. Probing superconductivity forces us to use the Nambu formalism, in which each degree of freedom is occurring in particle and hole form in a multiplet. Thus, the index is a composite index comprising cluster site , spin and Nambu indices: . This index takes values in the particular AIM that we use. Likewise, the index comprises bath orbital index , spin and Nambu indices and takes values: . is a complex-valued, hybridization matrix between cluster and bath orbitals, whereas is a matrix of one-body terms within the bath, including possible superconducting pairing. In principle, the matrix could be diagonalized (this would change the values of the hybridizations ), but we find it convenient and intuitive to allow pairing operators between bath orbitals.
The bath parameters are assumed to be spin independent, since we are not looking for magnetic ordering. In order to probe superconductivity, we include singlet and triplet pairing operators within the bath. Given two bath orbitals labeled by and , the following pairing operators are defined:
(singlet) | (7) | |||
(triplet) | (8) |
In terms of the bath orbital numbering scheme defined on Fig. 1b, the pairing terms in are
(9) |
These terms are included in the one-body matrix .
The AIM is characterized by 10 variational parameters, all illustrated on Fig. 1b: bath energy levels (diagonal elements of ), hybridization amplitudes , singlet pairing amplitudes and triplet pairing amplitudes , , , . It turns out that, owing to the rather small value of in Eq. (3), the triplet bath parameters are too small to have an observable effect and can be neglected. This reduces the number of independent bath parameters to six. The two clusters forming the supercell also happen to have the same bath parameter values in the converged solutions, which is expected from symmetry.
For a given set of bath parameters, the AIM (6) may be solved and the electron Green function computed. The latter may be expressed as
(10) |
where is the one-body matrix in the cluster part of the impurity Hamiltonian , is the associated self-energy, and is the bath hybridization matrix:
(11) |
where is the matrix with components and the matrix with components . Equation (10) allows us to extract the cluster self-energy from computed quantities. The fundamental approximation of CDMFT is to replace the full self-energy of the problem with the local self-energy . More precisely, when the supercell contains more than one cluster, as is the case here, the supercell self-energy is the direct sum of the self-energies of the different clusters: . The lattice Green function is then approximated as
(12) |
where is a wave vector of the reduced Brillouin zone (associated with the superlattice) and is the one-body Hamiltonian (2)-(4) expressed in that mixed basis of reduced wave vectors and supercell orbitals. In our system, the matrix has dimension , because of the two clusters forming the supercell.

Let us finally summarize the self-consistent procedure used to set the bath parameters, as proposed initially in Caffarel and Krauth (1994): (i) trial values of the bath parameters are chosen on the first iteration. (ii) For each iteration, the AIM (6) is solved, i.e., the cluster Green functions are computed using the Lanczos method, for each cluster. (iii) The bath parameters are updated, by minimizing the distance function:
(13) |
where is the restriction to cluster of the projected Green function , defined as
(14) |
(iv) We go back to step (ii) and iterate until the bath parameters or the bath hybridization functions stop varying within some preset tolerance.
Ideally, should coincide with the impurity Green function , but the finite number of bath parameters does not allow for this correspondence at all frequencies. This is why a distance function is defined, with emphasis on low frequencies along the imaginary axis. The weight function is where the method has some arbitrariness; in this work is taken to be a constant for all Matsubara frequencies lower than a cutoff , with a fictitious temperature .
The lattice Green function (12) can be used to compute the average of any one-body operator defined on the lattice. In addition, we can go back to a fully wave vector-dependent representation , where now belongs to the original Brillouin zone and is a smaller, matrix, by a procedure called periodization. The simplest periodization scheme is to Fourier transform directly from the supercell to the original Brillouin zone, as follows Sénéchal et al. (2000):
(15) |
where are composite spin, Nambu and layer indices, and the difference between and is an element of the reciprocal superlattice: . Note that since is by construction a periodic function of the reduced Brillouin zone.


Results and discussion —
Figure 2 shows the -wave order parameter, computed from the Green function (12), as a function of hole doping, for . This is the ground state average of the following operator:
(16) |
where and denote the nearest-neighbor vectors on the square lattice. The blue curve is obtained in a single-layer model, without spin orbit coupling. The black squares are obtained in the current bilayer model, and differ very little from the single layer values, because of the small value of both and . The green diamonds are the average of the following triplet operator:
(17) |
Note the factor of 100 in the scale. A similar operator defined along the axis with the component of the triplet -vector has equal expectation values.
The periodized Green function (15) give us access to quantities observable by SARPES, such as the spectral function:
(18) |
where means a trace over spin and layer indices that excludes the Nambu sector. One can also extract spin information by projecting the Green function on various spin directions. We thus define the following spin spectral functions:
(19) |
One can also define the corresponding layer-resolved quantities.
Figure 3 shows the spin texture measured in three of the superconducting solutions obtained, for three different values of density (underdoped, optimally doped and overdoped), on the first layer. The left panel shows a color plot of the spectral function (18) at the Fermi level (the non-interacting Fermi surface is indicated by a dashed line). The right panel shows the magnitude of the projection on the - plane of the spin spectral function (19), on which we superimposed a (diluted) vector plot indicating the direction of the spin in the - plane, as given by the vector . On this plot the arrows only provide a direction, not a magnitude. The color scale is the same for all three values of electron density in the figure. The structure of the spin texture is similar for all three cases illustrated: it is made of a clockwise rotating pattern around and around , on the first layer. The corresponding quantities on the second layer are obtained simply by reversing the arrows.
As expected, the amplitude of the spin texture is maximal in the vicinity of the non-interacting Fermi surface, with the important proviso that the pseudogap phenomenon suppresses the spectral weight (and the amplitude of the spin texture) away from the diagonals as we go in the underdoped regime. Exactly on the Fermi surface, at least near the diagonals, the spin texture vanishes as it must change direction, leading to split maxima along the diagonals on the right panels of Fig. 3. Figure 4 shows the drop of the spectral weight (indeed its maximum value over the Brillouin zone at the Fermi level) as doping decreases (blue circles). This suppression is even more pronounced for the maximum value of (red squares).
Overall, the known physics of the two-dimensional, one-band Hubbard model is not affected by the presence of the spin-orbit coupling between the two layers or by the inter-layer coupling. The -wave order parameter (Fig. 2) is not affected in any visible way and the triplet component of the order parameter is two orders of magnitude smaller than the singlet component. The spin-orbit coupling and the associated spin texture do not interfere with the pseudogap physics illustrated by the concentration of the spectral weight along the diagonal as one gets closer to half-filling. The spin-orbit coupling is too small to make the system topological Nakosai et al. (2012), and we checked that the Chern number, computed by adding the real part of the zero-frequency self-energy to the non-interacting Hamiltonian Wang and Zhang (2012a, b), is indeed zero in all the cases studied. The amplitude of the spin texture, however, is more suppressed in the underdoped region than the spectral function is (roughly twice as much).
Acknowledgements.
Discussions with Yuehua Su and A.-M. Tremblay are gratefully acknowledged. Computing resources were provided by Compute Canada and Calcul Québec. X.L. is supported by the National Natural Science Foundation of China (Grant No. 11974293) and the Fundamental Research Funds for Central Universities (Grant No. 20720180015).References
- Witczak-Krempa et al. (2014) W. Witczak-Krempa, G. Chen, Y. B. Kim, and L. Balents, Annual Review of Condensed Matter Physics 5, 57 (2014).
- Hasan and Kane (2010) M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010).
- Qi and Zhang (2011) X.-L. Qi and S.-C. Zhang, Rev. Mod. Phys. 83, 1057 (2011).
- Shimozawa et al. (2016) M. Shimozawa, S. K. Goh, T. Shibauchi, and Y. Matsuda, Reports on Progress in Physics 79, 074503 (2016).
- Rau et al. (2016) J. G. Rau, E. K.-H. Lee, and H.-Y. Kee, Annual Review of Condensed Matter Physics 7, 195 (2016).
- Manchon et al. (2015) A. Manchon, H. C. Koo, J. Nitta, S. M. Frolov, and R. A. Duine, Nature Materials 14, 871 (2015).
- Frigeri et al. (2004) P. A. Frigeri, D. F. Agterberg, A. Koga, and M. Sigrist, Phys. Rev. Lett. 92, 097001 (2004).
- Yokoyama et al. (2007) T. Yokoyama, S. Onari, and Y. Tanaka, Phys. Rev. B 75, 172511 (2007).
- Yanase and Sigrist (2008) Y. Yanase and M. Sigrist, Journal of the Physical Society of Japan 77, 124711 (2008).
- Tada et al. (2008) Y. Tada, N. Kawakami, and S. Fujimoto, Journal of the Physical Society of Japan 77, 054707 (2008).
- Greco and Schnyder (2018) A. Greco and A. P. Schnyder, Phys. Rev. Lett. 120, 177002 (2018).
- Lu and Sénéchal (2018) X. Lu and D. Sénéchal, Phys. Rev. B 98, 245118 (2018).
- Ghadimi et al. (2019) R. Ghadimi, M. Kargarian, and S. A. Jafari, Phys. Rev. B 99, 115122 (2019).
- Nogaki and Yanase (2020) K. Nogaki and Y. Yanase, Phys. Rev. B 102, 165114 (2020).
- Wolf and Rachel (2020) S. Wolf and S. Rachel, Phys. Rev. B 102, 174512 (2020).
- Zhang et al. (2015) X. Zhang, W. Wu, G. Li, L. Wen, Q. Sun, and A.-C. Ji, New Journal of Physics 17, 073036 (2015).
- Greco et al. (2020) A. Greco, M. Bejas, and A. P. Schnyder, Phys. Rev. B 101, 174420 (2020).
- Farrell and Pereg-Barnea (2014) A. Farrell and T. Pereg-Barnea, Phys. Rev. B 89, 035112 (2014).
- Laubach et al. (2014) M. Laubach, J. Reuther, R. Thomale, and S. Rachel, Phys. Rev. B 90, 165136 (2014).
- Marcelino (2017) E. Marcelino, Phys. Rev. B 95, 195112 (2017).
- Fischer et al. (2011) M. H. Fischer, F. Loder, and M. Sigrist, Phys. Rev. B 84, 184533 (2011).
- Maruyama et al. (2012) D. Maruyama, M. Sigrist, and Y. Yanase, Journal of the Physical Society of Japan 81, 034702 (2012).
- Sigrist et al. (2014) M. Sigrist, D. F. Agterberg, M. H. Fischer, J. Goryo, F. Loder, S.-H. Rhim, D. Maruyama, Y. Yanase, T. Yoshida, and S. J. Youn, Journal of the Physical Society of Japan 83, 061014 (2014).
- Zhang et al. (2014) X. Zhang, Q. Liu, J.-W. Luo, A. J. Freeman, and A. Zunger, Nature Physics 10, 387 (2014).
- Goh et al. (2012) S. K. Goh, Y. Mizukami, H. Shishido, D. Watanabe, S. Yasumoto, M. Shimozawa, M. Yamashita, T. Terashima, Y. Yanase, T. Shibauchi, A. I. Buzdin, and Y. Matsuda, Phys. Rev. Lett. 109, 157006 (2012).
- Nishikubo et al. (2011) Y. Nishikubo, K. Kudo, and M. Nohara, Journal of the Physical Society of Japan 80, 055002 (2011).
- Jones et al. (2014) A. M. Jones, H. Yu, J. S. Ross, P. Klement, N. J. Ghimire, J. Yan, D. G. Mandrus, W. Yao, and X. Xu, Nature Physics 10, 130 (2014).
- Liu (2017) C.-X. Liu, Phys. Rev. Lett. 118, 087001 (2017).
- Riley et al. (2014) J. M. Riley, F. Mazzola, M. Dendzik, M. Michiardi, T. Takayama, L. Bawden, C. Granerød, M. Leandersson, T. Balasubramanian, M. Hoesch, T. K. Kim, H. Takagi, W. Meevasana, P. Hofmann, M. S. Bahramy, J. W. Wells, and P. D. C. King, Nature Physics 10, 835 (2014).
- Nakosai et al. (2012) S. Nakosai, Y. Tanaka, and N. Nagaosa, Phys. Rev. Lett. 108, 147003 (2012).
- Das and Balatsky (2013) T. Das and A. V. Balatsky, Nature Communications 4, 1972 (2013).
- Dong et al. (2015) X.-Y. Dong, J.-F. Wang, R.-X. Zhang, W.-H. Duan, B.-F. Zhu, J. O. Sofo, and C.-X. Liu, Nature Communications 6, 8517 (2015).
- Ishizuka and Yanase (2018) J. Ishizuka and Y. Yanase, Phys. Rev. B 98, 224510 (2018).
- Gotlieb et al. (2018) K. Gotlieb, C.-Y. Lin, M. Serbyn, W. Zhang, C. L. Smallwood, C. Jozwiak, H. Eisaki, Z. Hussain, A. Vishwanath, and A. Lanzara, Science 362, 1271 (2018).
- Raines et al. (2019) Z. M. Raines, A. A. Allocca, and V. M. Galitski, Phys. Rev. B 100, 224512 (2019).
- Hitomi and Yanase (2019) T. Hitomi and Y. Yanase, Journal of the Physical Society of Japan 88, 054712 (2019).
- Atkinson (2020) W. A. Atkinson, Phys. Rev. B 101, 024513 (2020).
- Harrison et al. (2015) N. Harrison, B. J. Ramshaw, and A. Shekhter, Scientific Reports 5, 10914 (2015).
- Markiewicz et al. (2005) R. S. Markiewicz, S. Sahrakorpi, M. Lindroos, H. Lin, and A. Bansil, Phys. Rev. B 72, 054519 (2005).
- Drozdov et al. (2018) I. K. Drozdov, I. Pletikosić, C. K. Kim, K. Fujita, G. D. Gu, J. C. S. Davis, P. D. Johnson, I. Božović, and T. Valla, Nature Communications 9, 5210 (2018).
- Lichtenstein and Katsnelson (2000) A. I. Lichtenstein and M. I. Katsnelson, Phys. Rev. B 62, R9283 (2000).
- Kotliar et al. (2001) G. Kotliar, S. Y. Savrasov, G. Pálsson, and G. Biroli, Phys. Rev. Lett. 87, 186401 (2001).
- Liebsch et al. (2008) A. Liebsch, H. Ishida, and J. Merino, Phys. Rev. B 78, 165123 (2008).
- Sénéchal (2015) D. Sénéchal, in Many-Body Physics: From Kondo to Hubbard, Vol. 5, edited by E. Pavarini, E. Koch, and P. Coleman (Forschungszentrum Jülich, 2015) pp. 13.1–13.22.
- Caffarel and Krauth (1994) M. Caffarel and W. Krauth, Phys. Rev. Lett. 72, 1545 (1994).
- Sénéchal et al. (2000) D. Sénéchal, D. Perez, and M. Pioro-Ladrière, Phys. Rev. Lett. 84, 522 (2000).
- Verret et al. (2019) S. Verret, J. Roy, A. Foley, M. Charlebois, D. Sénéchal, and A.-M. S. Tremblay, Phys. Rev. B 100, 224520 (2019).
- Wang and Zhang (2012a) Z. Wang and S.-C. Zhang, Phys. Rev. X 2, 031008 (2012a).
- Wang and Zhang (2012b) Z. Wang and S.-C. Zhang, Phys. Rev. B 86, 165116 (2012b).