Skip to main content
Log in

Wigner-Weyl calculus in Keldysh technique

  • Published:
Journal of Computational Electronics Aims and scope Submit manuscript

Abstract

We discuss the non-equilibrium dynamics of condensed matter/quantum field systems in the framework of Keldysh technique. In order to deal with the inhomogeneous systems we use the Wigner-Weyl formalism. Unification of the mentioned two approaches is demonstrated on the example of Hall conductivity. We express Hall conductivity through the Wigner transformed two-point Green’s functions. We demonstrate how this expression is reduced to the topological number in thermal equilibrium at zero temperature. At the same time both at finite temperature and out of equilibrium the topological invariance is lost. Moreover, Hall conductivity becomes sensitive to interaction corrections.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Keldysh, L.V.: Diagram technique for nonequilibrium processes. Zh. Eksp. Teor. Fiz. 47, 1515 (1964)

    MathSciNet  Google Scholar 

  2. Keldysh, L.V.: Diagram technique for nonequilibrium processes. Sov. Phys. JETP 20, 1018 (1965)

    MathSciNet  Google Scholar 

  3. Bonitz, M. (ed.): Progress in Nonequilibrium Green’s Functions. World Scientific, Singapore (2000)

  4. Bonitz, M., Semkat, D. (eds.): Progress in Nonequilibrium Green’s Functions II. World Scientific, Singapore (2003)

  5. Bonitz, M., Filinov, A. (Eds.): Progress in Nonequilibrium Green’s Functions III. In: Journal of Physics: Conference Series, vol. 35. (2006)

  6. Kadanoff, L.P., Baym, G.: Quantum Statistical Mechanics. Benjamin, New York (1962)

    MATH  Google Scholar 

  7. Baym, G.: Selfconsistent approximation in many body systems. Phys. Rev. 127, 1391 (1962)

    MathSciNet  MATH  Google Scholar 

  8. Schwinger, J.: Brownian motion of a quantum oscillator. J. Math. Phys. 2, 407 (1961)

    MathSciNet  MATH  Google Scholar 

  9. Matsubara, T.: A new approach to quantum statistical mechanics. Prog. Theor. Phys. 14, 351 (1955)

    MathSciNet  MATH  Google Scholar 

  10. Bloch, C., de Donimicis, C.: Un développement du potentiel de Gibbs d’un système quantique composé d’un grand nombre de particules III-La contribution des collisions binaires. Nucl. Phys. 10, 509 (1959)

  11. Gaudin, M.: Une démonstration simplifiée du théorème de wick en mécanique statistique. Nucl. Phys. 15, 89 (1960)

    MathSciNet  MATH  Google Scholar 

  12. Abrikosov, A.A., Gor’kov, L.P., Dzyaloshinski, I.E.: Methods of Quantum Field Theory in Statistical Physics. Prentice-Hall, Englewood Cliffs, N.J. (1963)

  13. Kamenev, A.: Field Theory of Non-Equilibrium Systems, by Alex Kamenev. Cambridge University Press, UK (2011)

    Google Scholar 

  14. Kamenev, A.: Many-body theory of non-equilibrium systems arXiv: cond-mat/0412296

  15. Langreth, D.C.: In: Devreese, J.T., van Doren, V.E. (eds.) Linear and Nonlinear Electron Transport in Solids, p. 3. Plenum Press, New York (1976)

  16. Danielewicz, P.: Quantum theory of nonequilibrium processes. Ann. of Phys. 1 152, 239 (1984)

    Google Scholar 

  17. Chou, K.C., Su, Z.B., Hao, B.L., Yu, L.: Equilibrium and nonequilibrium formalisms made unified. Phys. Rep. 118, 1 (1985)

    MathSciNet  Google Scholar 

  18. Rammer, J., Smith, H.: Quantum field-theoretical methods in transport theory of metals. Rev. Mod. Phys. 58, 323 (1986)

    Google Scholar 

  19. Berges, J.: Introduction to nonequilibrium quantum field theory. AIP Conf. Proc. 739(1), 3–62 (2004). arXiv:hep-ph/0409233

  20. Haug, H., Jauho, A.-P.: Quantum Kinetics in Transport and Optics of Semiconductors. Springer-Verlag, Berlin (1998)

    Google Scholar 

  21. Rammer, J.: Quantum Field Theory of Nonequilibrium States. Cambridge University Press, Cambridge (2007)

    MATH  Google Scholar 

  22. Lifshitz, E.M., Pitaevskii, L.P.: Physical Kinetics. Pergamon, New York (1981)

    Google Scholar 

  23. Mahan, G.D.: Many-Particle Physics. Kluwer Academic/Plenum, New York (2000)

    Google Scholar 

  24. Landau, L.D.: The theory of a fermi liquid. Zh. Eksp. Teor. Fiz. 30, 1058 (1956)

    Google Scholar 

  25. Landau, L.D.: The theory of a fermi liquid. Sov. Phys. JETP 3, 920 (1957)

    MathSciNet  MATH  Google Scholar 

  26. Landau, L.D.: Oscillations in a fermi liquid. Zh. Eksp. Teor. Fiz. 32, 59 (1957)

    MATH  Google Scholar 

  27. Landau, L.D.: Oscillations in a fermi liquid. Sov. Phys. JETP 5, 101 (1957)

    MathSciNet  MATH  Google Scholar 

  28. Luttinger, J.M., Ward, J.C.: Ground-State energy of a many-fermion system. II. Phys. Rev. 118, 1417 (1960)

    MathSciNet  MATH  Google Scholar 

  29. Luttinger, J.M.: Fermi surface and some simple equilibrium properties of a system of interacting fermions. Phys. Rev. 119, 1153 (1960)

    MathSciNet  MATH  Google Scholar 

  30. Cercignani, C.: The Boltzmann Equation and Its Applications. Springer-Verlag, New York (1988)

    MATH  Google Scholar 

  31. Caroli, C., Combescot, R., Nozières, P., Saint-James, D.: A direct calculation of the tunnelling current. II. Free electron description. J. Phys. C: Solid St. Phys. 4, 916 (1971)

  32. Aronov, A.G., Gurevich, V.L.: Finite-amplitude electron sound in superconductors. Fiz. Tverd. Tela 16, 2656 (1974)

    Google Scholar 

  33. Aronov, A.G., Gurevich, V.L.: Finite-amplitude electron sound in superconductors. Sov. Phys. Solid State 16, 1722 (1975)

    Google Scholar 

  34. Larkin, A.I., Ovchinnikov, Yu, B.: Nonlinear conductivity of superconductors in the mixed state. Zh. Eksp. Teor. Fiz. 68, 1915 (1975)

    Google Scholar 

  35. Larkin, A.I., Ovchinnikov, Yu, B.: Nonlinear conductivity of superconductors in the mixed state. Sov. Phys. JETP 41, 960 (1975)

    Google Scholar 

  36. Iancu, E., Leonidov, A., McLerran, L.D.: Nonlinear gluon evolution in the color glass condensate. 1. Nucl. Phys. A 692, 583 (2001)

    MATH  Google Scholar 

  37. Jensen, K., et al.: A panoply of schwinger-keldysh transport. Sci. Post Phys. 5.5 (2018)

  38. Akhmedov, Emil T., Godazgar, Hadi, Popov, Fedor K.: Hawking radiation and secularly growing loop corrections. Phys. Rev. D 93, 024029 (2016)

    MathSciNet  Google Scholar 

  39. Wigner, E.: On the quantum correction for thermodynamic equilibrium. Phys. Rev. 40, 749 (1932)

    MATH  Google Scholar 

  40. Groenewold, H.J.: On the principles of elementary quantum mechanics. Physica 12, 405 (1946)

    MathSciNet  MATH  Google Scholar 

  41. Moyal, J.E.: Quantum mechanics as a statistical theory. Proc. Cambridge Philos. Soc. 45, 99 (1949)

    MathSciNet  MATH  Google Scholar 

  42. Weyl, H.: Quantenmechanik und Gruppentheorie. Z. Phys. 46, 1 (1927)

    MATH  Google Scholar 

  43. Ali, S.T., Englis, M.: Quantization methods: a Guide for physicists and analysts. Rev. Math. Phys. 17, 391 (2005)

    MathSciNet  MATH  Google Scholar 

  44. Berezin, F.A., Shubin, M.A.: The Schrödinger Equation, p. 21. (1972)

  45. Curtright, T.L., Zachos, C.K.: Quantum mechanics in phase space. Asia Pac. Phys. Newsl. 1, 37 (2012). arXiv:1104.5269

  46. Zachos, C., Fairlie, D., Curtright, T.: Quantum Mechanics in Phase Space. World Scientic, Singapore (2005)

  47. Cohen, L.: Generalized phase-space distribution functions. J. Math. Phys. 7, 781 (1966)

    MathSciNet  Google Scholar 

  48. Agarwal, G.S., Wolf, E.: Calculus for functions of noncommuting operators and general phase-space methods in quantum mechanics. i. Mapping theorems and ordering of functions of noncommuting operators. Phys. Rev. D 2, 2161 (1970)

    MathSciNet  MATH  Google Scholar 

  49. Glauber, R.J.: Coherent and incoherent states of the radiation field. Phys. Rev 131, 2766 (1963)

    MathSciNet  MATH  Google Scholar 

  50. Husimi, K.: Some formal properties of the density matrix. Proc. Phys. Math. Soc. Jpn. 22, 264–314 (1940)

    MATH  Google Scholar 

  51. Cahill, K.E., R. J.: Causal signal transmission by quantum fields – V: Generalised Keldysh rotations and electromagnetic response of the Dirac sea. Phys. Rev. 177, 1882 (1969)

  52. Buot, F.A.: Nonequilibrium Quantum Transport Physics in Nanosystems. World Scientic, Singapore (2009)

    MATH  Google Scholar 

  53. Lorce, C., Pasquini, B.: Quark wigner distributions and orbital angular momentum. Phys. Rev. D 84, 014015 (2011)

    Google Scholar 

  54. Elze, H.T., Gyulassy, M., Vasak, D.: Transport equations for the QCD quark wigner operator. Nucl. Phys. B 706, 276 (1986)

    Google Scholar 

  55. Hebenstreit, F., Alkofer, R., Gies, H.: Schwinger pair production in space and time-dependent electric fields: relating the Wigner formalism to quantum kinetic theory. Phys. Rev. D 82, 105026 (2010). arXiv:1007.1099

    Google Scholar 

  56. Calzetta, E., Habib, S., Hu, B.L.: Quantum kinetic field theory in curved spacetime: covariant Wigner function and Liouville-Vlasov equations. Phys. Rev. D 37, 2901 (1988)

    MathSciNet  Google Scholar 

  57. Bastos, C., Bertolami, O., Dias, N.C., Prata, J.N.: Weyl-Wigner formulation of noncommutative quantum mechanics. J. Math. Phys. 49, 072101 (2008). arXiv:hep-th/0611257]

    MathSciNet  MATH  Google Scholar 

  58. Dayi, O.F., Kelleyane, L.T.: Wigner functions for the Landau problem in noncommutative spaces. Mod. Phys. Lett. A 17, 1937 (2002). arXiv:hep-th/0202062

    MathSciNet  MATH  Google Scholar 

  59. Habib, S., Laamme, R.: Wigner function and decoherence in quantum cosmology. Phys. Rev. D 42, 4056 (1990)

    MathSciNet  Google Scholar 

  60. Chapman, S., Heinz, U.W.: HBT correlators: current formalism versus wigner function formulation. Phys. Lett. B 340, 250 (1994). arXiv:hep-ph/9407405

    Google Scholar 

  61. Berry, M.V.: Semi-classical mechanics in phase space: A study of wigner’s function. Phil. Trans. R. Soc. Lond. A 287, 0145 (1977)

  62. Zhang, C.X., Zubkov, M.A.: Feynman rules in terms of the wigner transformed green’s functions. Phys. Lett. B 802, 135197 (2020)

  63. Zubkov, M.A., Wu, X.: Topological invariant in terms of the green functions for the quantum hall effect in the presence of varying magnetic field. Ann. Phys. 418, 168179 (2020). Annals Phys. 430 (2021) 168510 (erratum); arXiv:1901.06661

  64. Zubkov, M.A.: Wigner transformation, momentum space topology, and anomalous transport. Ann. Phys. 373, 298 (2016)

    MathSciNet  MATH  Google Scholar 

  65. Thouless, D.J., Kohmoto, M., Nightingale, M.P., den Nijs, M.: Classification of Hamiltonians in neighborhoods of band crossings in terms of the theory of singularities. Phys. Rev. Lett. 49, 405 (1982)

    Google Scholar 

  66. Suleymanov, M., Zubkov, M.A.: p. 07178. Chiral separation effect in non-homogeneous systems. Phys. Rev. D 102(7), 076019 (2007)

  67. Zhang, C.X., Zubkov, M.A.: Influence of interactions on the anomalous quantum Hall effect. J. Phys. A Math. Theor. 53(19), 195002

  68. Zhang, C.X., Zubkov, M.A.: Hall conductivity as the topological invariant in the phase space in the presence of interactions and a nonuniform magnetic field. Jetp Lett. 110, 487 (2019)

  69. Shitade, A.: Anomalous thermal hall effect in a disordered weyl ferromagnet. J. Phys. Soc. Jpn. 86(5), 054601 (2017)

    Google Scholar 

  70. Onoda, S., Sugimoto, N., Nagaosa, N.: Theory of non-equilibirum states driven by constant electromagnetic fields: non-commutative quantum mechanics in the keldysh formalism. Prog. Theor. Phys. 116, 61 (2006)

    MATH  Google Scholar 

  71. Sugimoto, N., Onoda, S., Nagaosa, N.: Gauge covariant formulation of the wigner representation through deformation quantization: application to keldysh formalism with an electromagnetic Field. Prog. Theor. Phys. 117, 415 (2007)

    MATH  Google Scholar 

  72. Onoda, S., Sugimoto, N., Nagaosa, N.: Intrinsic versus extrinsic anomalous hall effect in ferromagnets. Phys. Rev. Lett. 97, 126602 (2006)

    Google Scholar 

  73. Onoda, S., Sugimoto, N., Nagaosa, N.: Quantum transport theory of anomalous electric, thermoelectric, and thermal Hall effects in ferromagnets. Phys. Rev. B 77, 165103 (2008)

    Google Scholar 

  74. Polkovnikov, A.: Phase space representation of quantum dynamics. Ann. Phys. 325, 1790–1852 (2010). https://doi.org/10.1016/j.aop.2010.02.006

    Article  MathSciNet  MATH  Google Scholar 

  75. Haldane, F.D.M.: Model for a quantum hall effect without landau levels: condensed-matter realization of the parity anomaly. Phys. Rev. Lett. 02, 076019 (1988)

  76. Vilenkin, A.: Equilibrium parity-violating current in a magnetic field. Phys. Rev. D 22, 3080–3084 (1980)

    Google Scholar 

  77. Kharzeev, D.E.: Chern-Simons current and local parity violation in hot QCD matter. Nucl. Phys. A 830, 543C (2009). https://doi.org/10.1016/j.nuclphysa.2009.10.049

    Article  Google Scholar 

  78. Kharzeev, D.E., Warringa, H.J.: Chiral magnetic conductivity. Phys. Rev. D 80, 034028 (2009). https://doi.org/10.1103/PhysRevD.80.034028

    Article  Google Scholar 

  79. Kharzeev, D.E.: The Chiral magnetic effect and anomaly-induced transport. Prog. Part. Nucl. Phys. 75, 133 (2014). https://doi.org/10.1016/j.ppnp.2014.01.002

    Article  Google Scholar 

  80. Kharzeev, D.E., Liao, J., Voloshin, S.A., Wang, G.: Chiral magnetic effect in high-energy nuclear collisions — a status report. arXiv:1511.04050

  81. Metlitski Max, A., Zhitnitsky Ariel, R.: Anomalous axion interactions and topological currents in dense matter. Phys. Rev. D 72, 045011 (2005). https://doi.org/10.1103/PhysRevD.72.045011

    Article  Google Scholar 

  82. Khaidukov, Z.V., Zubkov, M.A.: Chiral separation effect in lattice regularization. Phys. Rev. D 95, 074502 (2017). https://doi.org/10.1103/PhysRevD.95.074502

    Article  MathSciNet  Google Scholar 

  83. Volovik, G.E.: The Universe in a Helium Droplet. Clarendon Press, Oxford, UK (2003)

    MATH  Google Scholar 

  84. Abramchuk, R., Khaidukov, Z.V., Zubkov, M.A.: Anatomy of the chiral vortical effect. Phys. Rev. D 2018, 98, https://doi.org/10.1103/PhysRevD.98.076013

  85. Golkar, S., Son Dam, T.: (Non)-renormalization of the chiral vortical effect coefficient. JHEP 2015, 169, https://doi.org/10.1007/JHEP02(2015)169

  86. Hou, D., Liu, H., Ren, H.: A possible higher order correction to the chiral vortical conductivity in a gauge field plasma. Phys. Rev. D 86, 121703(R) (2012). https://doi.org/10.1103/PhysRevD.86.121703

    Article  Google Scholar 

  87. Khaidukov, Z.V., Zubkov, M.A.: Chiral torsional effect. JETP Lett. 108(10), 670–674 (2018). https://doi.org/10.1134/S0021364018220046

    Article  Google Scholar 

  88. Vazifeh, M.M., Franz, M.: Electromagnetic response of weyl semimetals. Phys. Rev. Lett. 111, 027201 (2013)

    Google Scholar 

  89. Yamamoto, N.: Generalized bloch theorem and chiral transport phenomena. Phys. Rev. D 92, 085011 (2015)

    Google Scholar 

  90. Yan, W., Hou, D.F., Ren, H.C.: Field theoretic perspectives of the Wigner function formulation of the chiral magnetic effect. Phys. Rev. D 96, 096015 (2017)

  91. Li, Q., Kharzeev, D.E., Zhang, C., Huang, Y., Pletikosic, I., Fedorov, A.V., Zhong, R.D., Schneeloch, J.A., Gu, G.D., Valla, T.: Observation of the chiral magnetic effect in ZrTe5. Nature Phys. 12, 550–554 (2016)

    Google Scholar 

  92. Fialkovsky, I.V., Zubkov, M.A.: Wigner-Weyl, precise, calculus for lattice models, Nucl. Phys. B 954, 114999 (2020). https://doi.org/10.1016/j.nuclphysb.2020.114999

  93. Aoki, H.: Aharonov-Bohm effect for the quantum hall conductivity on a disordered lattice. Phys. Rev. Lett. 55, 1136 (1985)

    Google Scholar 

  94. Ortmann, F., Leconte, N., Roche, S.: Efficient linear scaling approach for computing the Kubo Hall conductivity. Phys. Rev. B 91, 165117 (2015)

    Google Scholar 

  95. Aoki, H., Ando, T.: Effect of localization on the hall conductivity in the two-dimensional system in strong magnetic fields. Solid State Commun. 38, 1079 (1981)

    Google Scholar 

  96. Lux, F.R., Freimuth, F., Blügel, S., Mokrousov, Y.: Chiral hall effect in noncollinear magnets from a cyclic cohomology approach. Phys. Rev. Lett. 124(9), 096602 (2020)

    MathSciNet  Google Scholar 

  97. Lux, F.R., Prass, P., Blügel, S., Mokrousov, Y.: Effective seiberg-witten gauge theory of noncollinear magnetism, arXiv preprint arXiv:2005.12629

Download references

Funding

Not applicable

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to M. A. Zubkov.

Ethics declarations

Conflict of interest

Not applicable

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Lesser component of a product of operators

Here we consider the static non-interacting system with the one-particle Hamiltonian that does not depend on time. We will show that if the derivative of distribution function may be neglected

$$\begin{aligned} H &\equiv \int d^{{D+1}}\pi \, \mathop {\mathrm{tr}}\nolimits \left( K_1 \star K_2\star \ldots \star K_n\right) ^< \nonumber \\ &=\int d^{{D+1}}\pi \, f(\pi )\left[ \mathop {\mathrm{tr}}\nolimits \left( K_1 \star K_2\star \ldots \star K_n\right) ^\mathrm{A}\nonumber \right. \\&\quad \left. - \mathop {\mathrm{tr}}\nolimits \left( K_1 \star K_2\star \ldots \star K_n\right) ^\mathrm{R}\right] , \end{aligned}$$
(A1)

where each of the \(K_i\) is a Green’s function G, its inverse Q or a derivative of one of those. \(f(\pi )\) stands for the initial distribution of the system, not necessarily the equilibrium one. See (14), (19) and (20) for the notions of A/R and “lesser” components.

To prove the above, we first note that

$$\begin{aligned}&\left( K_1 \star K_2\star \ldots K_n\right) ^< = \sum _{i=0}^n \left[ K_1^\mathrm{R}\star \ldots \star K_i^\mathrm{R}\star K^<_{i+1} \right. \left. \star K_{i+2}^\mathrm{A}\star \ldots \star K_n^\mathrm{A}\right] , \end{aligned}$$
(A2)

so that in every term of this sum there only one factor in the “lesser” form.

In the case of a static system for any \(K_i\) that does not contain derivatives of G/Q, we have (see Eqs. 19 and 20)

$$\begin{aligned} K^<_i = (K^\mathrm{A}_i-K^\mathrm{R}_i) f(\pi ). \end{aligned}$$
(A3)

Then H becomes

$$\begin{aligned} \begin{aligned} H&= \sum _{i=0}^n \int d^{{D+1}}\pi \, \mathop {\mathrm{tr}}\nolimits \left[ K_1^\mathrm{R}\star \ldots \star K_i^\mathrm{R}\right. \\&\left. \quad \star \left[ (K^\mathrm{A}_{i+1}-K^\mathrm{R}_{i+1}) f(\pi )\right] \star K_{i+2}^\mathrm{A}\star \ldots \star K_n^\mathrm{A}\right] \\&= \sum _{i=0}^n \int d^{{D+1}}\pi \, \mathop {\mathrm{tr}}\nolimits \left[ K_1^\mathrm{R}\star \ldots \star K_i^\mathrm{R}\star \left[ K^\mathrm{A}_{i+1} f(\pi )\right] \right. \\&\left. \quad \star K_{i+2}^\mathrm{A}\star \ldots \star K_n^\mathrm{A}\right] \\&\quad - \mathop {\mathrm{tr}}\nolimits \left[ K_1^\mathrm{R}\star \ldots \star K_i^\mathrm{R}\star \left[ K^\mathrm{R}_{i+1} f(\pi )\right] \right. \\&\quad \left. \star K_{i+2}^\mathrm{A}\star \ldots \star K_n^\mathrm{A}\right] . \end{aligned} \end{aligned}$$
(A4)

This expression can be simplified in the case when position of \(f(\pi )\) is irrelevant, i.e. when \(K_i\star f(\pi ) = K_i f(\pi )\). It is possible in two cases: a) when derivatives of \(f(\pi )\) can be assumed being small; b) when all \(K_i\) do not depend on time, while \(f(\pi )\) depends only on the energy. In the former case, the following expression is only approximate, while in the latter it is exact:

$$\begin{aligned} \begin{aligned} H&= \sum _{i=0}^n \int d^{{D+1}}\pi \, f(\pi )\, \\&\quad \mathop {\mathrm{tr}}\nolimits \left( K_1^\mathrm{R}\star \ldots \star K_i^\mathrm{R}\star K^\mathrm{A}_{i+1} \star K_{i+2}^\mathrm{A}\star \ldots \star K_n^\mathrm{A}\right) \\&\quad - f(\pi )\, \mathop {\mathrm{tr}}\nolimits \left( K_1^\mathrm{R}\star \ldots \star K_i^\mathrm{R}\star K^\mathrm{R}_{i+1} \right. \\&\quad \left. \star K_{i+2}^\mathrm{A}\star \ldots \star K_n^\mathrm{A}\right) . \end{aligned}\end{aligned}$$
(A5)

We can see now that in the i-th contribution to the sum the first term in the parentheses cancels the second one at \(i-1\). Thus, we will be left only with the first and the last terms of the summation:

$$\begin{aligned} \begin{aligned} H&= \int d^{{D+1}}\pi \, f(\pi )\, \left[ \mathop {\mathrm{tr}}\nolimits \left( K_1^\mathrm{A}\star \ldots \star K_n^\mathrm{A}\right) \right. \\&\quad \left. - \mathop {\mathrm{tr}}\nolimits \left( K_1^\mathrm{R}\star \ldots \star K_n^\mathrm{R}\right) \right] . \end{aligned} \end{aligned}$$
(A6)

Very similar consideration is valid when

$$\begin{aligned} K_i = \partial _\pi G \mathrm{\ or\ } K_i = \partial _\pi Q. \end{aligned}$$
(A7)

To show it we recall, that due to the mutually inverse nature of G and Q, we can always write a derivative of one as a sandwiched derivative of the other,

$$\begin{aligned} \partial _\pi Q = -Q\star \partial _\pi G \star Q,\qquad \partial _\pi G = -G\star \partial _\pi Q \star G . \end{aligned}$$

Then, it is sufficient to analyse expressions of the following type entering (A2) in place of one of the \(K_i\)-s:

$$\begin{aligned} \begin{aligned} G^\mathrm{R}\star \partial _\pi Q^< \star G^\mathrm{A}&= G^\mathrm{R}\star \partial _\pi \left[ (Q^\mathrm{A}-Q^\mathrm{R}) f(\pi )\right] \star G^\mathrm{A}\\&= G^\mathrm{R}\star \left[ \partial _\pi (Q^\mathrm{A}-Q^\mathrm{R})\right] f(\pi )\star G^\mathrm{A}\\&\quad + G^\mathrm{R}\star \left[ \partial _\pi f(\pi )\right] (Q^\mathrm{A}-Q^\mathrm{R})\star G^\mathrm{A}\end{aligned}\end{aligned}$$
(A8)

Provided that the derivative of f may be neglected, the derivatives of G and Q behave similarly to G (Q) themselves in considered construction, and (A1) holds valid. In Appendix B we give the more detailed derivation in an important particular case of Hall conductivity, when last term of the above is important.

Finally, we note that disregarding the derivatives of \(f(\pi )\) is justified, in particular, when considering the low temperature limit of the initially equilibrium distribution. Moreover, if \(f=f(\pi _0)\) is thermal equilibrium distribution indeed, we can further rewrite (A2) as a sum over the Matsubara frequencies.

Indeed, by using the analytic properties of \(K^{R/A}\) (as inherited from those of G/Q) we have

$$\begin{aligned} H= & {} \int _{{\mathrm{Im}\pi _0 = {-} 0 }} d^{{D+1}}\pi \, f(\pi _0) \mathop {\mathrm{tr}}\nolimits \left( K_1 \star K_2\star \ldots \star K_n\right) \nonumber \\&- \int _{{\mathrm{Im}\pi _0 ={+}0 }} d^{{D+1}}\pi \, f(\pi _0) \mathop {\mathrm{tr}}\nolimits \left( K_1 \star K_2\star \ldots \star K_n\right) , \end{aligned}$$
(A9)

Closing now the integration contour over \(\pi ^0\) into upper (lower) half-plane in the first (second) terms of the above, we transform the original integrals to the two integrals surrounding poles of \(f(\pi _0)\). Those are given by imaginary unit times Matsubara frequency \(\omega _j = (2j+1)\pi /{\beta }\). Calculating the integral using residue theorem we obtain:

$$\begin{aligned} H= & {} 2\pi {i T} \int d^{{D}} \pi \sum _{\omega _j} \mathop {\mathrm{tr}}\nolimits \left( K_1^\mathrm{M}\star K_2^\mathrm{M}\star \ldots \star K_n^\mathrm{M}\right) \end{aligned}$$
(A10)

here the superscript M indicates that corresponding function is taken at Matsubara frequency: \(K^\mathrm{M}(\omega _j,\pi _1,...,\pi _D) \equiv K(i \omega _j,\pi _1,...,\pi _D)\). In the limit of small temperatures the sum over the Matsubara frequencies is reduced to an integral and we arrive at:

$$\begin{aligned} H ={i} \int d^{{D+1}} {\Pi } \mathop {\mathrm{tr}}\nolimits \left( K_1^\mathrm{M}\star K_2^\mathrm{M}\star \ldots \star K_n^\mathrm{M}\right) . \end{aligned}$$
(A11)

Here \(\Pi\) is “Euclidean” \(D+1\) - momentum, i.e. \(\Pi ^{D+1} = \omega\) is continuous Matsubara frequency, \(\Pi ^i = \pi ^i\) for \(i=1,...,D\).

Appendix B: Conductivity in a static system without interactions

Here we derive expressions for Hall and symmetric conductivities through Wigner transformed Green’s functions for the case of a noninteracting static system, in which initial non-thermal distribution depends on energy only in \(2+1 D\).

We start from tensor \({{\mathcal {K}}}^{\mu \nu \rho }\) defined in Eq. (74):

$$\begin{aligned}&{ {\mathcal {K}}}^{ijk} = \frac{1}{4 \, {{\mathcal {V}}}} \int \frac{d^3\pi d^2 x}{(2\pi )^3 } \mathop {\mathrm{tr}}\nolimits \left( \partial _{\pi _{i}}\hat{Q}_0 \right. \left. \star \hat{G}_0\star \partial _{\pi _{{j}}}\hat{Q}_0 \star \partial _{\pi _{{k}}}\hat{G}_0 \right) ^< +\mathrm{c.c.} \end{aligned}$$
(B1)

Hall conductivity is given by \({{\mathcal {K}}}^{i0j} - {{\mathcal {K}}}^{ij0}\), so we study both terms separately. We have

$$\begin{aligned} \begin{aligned} { {\mathcal {K}}}^{i0j}&= \frac{1}{4 \, {{\mathcal {V}}}} \int \frac{d^3\pi d^2 x}{(2\pi )^3 } \\&\quad \mathop {\mathrm{tr}}\nolimits \left( \partial _{\pi _{i}}\hat{Q}_0\star \partial _{\pi _{0}}\hat{G}_0 \star \partial _{\pi _{j}}\hat{Q}_0 \star \hat{G}_0 \right) ^< +\mathrm{c.c.}\\&= \frac{1}{4 \, {{\mathcal {V}}}} \int \frac{d^3\pi d^2 x}{(2\pi )^3 } \\&\quad \mathop {\mathrm{tr}}\nolimits \left( \partial _{\pi _{i}} {Q}^\mathrm{R}_0 \star \partial _{\pi _{0}} {G}^\mathrm{R}_0\star \partial _{\pi _{j}} {Q}^\mathrm{R}_0 \star \Big ((G^\mathrm{A}_0-G^\mathrm{R}_0) f(\pi _0)\Big ) \right) \\&\quad + \frac{1}{4 \, {{\mathcal {V}}}} \int \frac{d^3\pi d^2 x}{(2\pi )^3 } \\&\quad \mathop {\mathrm{tr}}\nolimits \left( \partial _{\pi _{i}} {Q}^\mathrm{R}_0 \star \partial _{\pi _{0}} {G}^\mathrm{R}_0\star \partial _{\pi _{j}}\Big ((Q^\mathrm{A}_0-Q^\mathrm{R}_0) f(\pi _0)\Big ) \star {G}^\mathrm{A}_0 \right) \\&\quad + \frac{1}{4 \, {{\mathcal {V}}}} \int \frac{d^3\pi d^2 x}{(2\pi )^3 } \\&\quad \mathop {\mathrm{tr}}\nolimits \left( \partial _{\pi _{i}} {Q}^\mathrm{R}_0 \star \partial _{\pi _{0}} \Big ((G^\mathrm{A}_0-G^\mathrm{R}_0) f(\pi _0)\Big )\star \partial _{\pi _{j}} {Q}^\mathrm{A}_0 \star {G}^\mathrm{A}_0 \right) \\&\quad + \frac{1}{4 \, {{\mathcal {V}}}} \int \frac{d^3\pi d^2 x}{(2\pi )^3 } \\&\quad \mathop {\mathrm{tr}}\nolimits \left( \partial _{\pi _{i}}\Big ((Q^\mathrm{A}_0-Q^\mathrm{R}_0) f(\pi _0)\Big ) \star \partial _{\pi _{0}} {G}^\mathrm{A}_0\star \partial _{\pi _{j}} {Q}^\mathrm{A}_0 \star {G}^\mathrm{A}_0 \right) +\mathrm{c.c.} \\&= -\frac{1}{4 \, {{\mathcal {V}}}} \int \frac{d^3\pi d^2 x}{(2\pi )^3 } \\&\quad \mathop {\mathrm{tr}}\nolimits \left( \partial _{\pi _{i}} {Q}^\mathrm{R}_0 \star \partial _{\pi _{0}} {G}^\mathrm{R}_0\star \partial _{\pi _{j}} {Q}^\mathrm{R}_0 \star G^\mathrm{R}_0 \right) f(\pi _0)\\&\quad + \frac{1}{4 \, {{\mathcal {V}}}} \int \frac{d^3\pi d^2 x}{(2\pi )^3 } \\&\quad \mathop {\mathrm{tr}}\nolimits \left( \partial _{\pi _{i}}Q^\mathrm{A}_0 \star \partial _{\pi _{0}} {G}^\mathrm{A}_0\star \partial _{\pi _{j}} {Q}^\mathrm{A}_0 \star {G}^\mathrm{A}_0 \right) f(\pi _0)\\&\quad + \frac{1}{4 \, {{\mathcal {V}}}} \int \frac{d^3\pi d^2 x}{(2\pi )^3 } \\&\quad \mathop {\mathrm{tr}}\nolimits \left( \partial _{\pi _{i}}Q^\mathrm{R}_0\star ( {G}^\mathrm{A}_0- {G}^\mathrm{R}_0) \star \partial _{\pi _{j}} {Q}^\mathrm{A}_0 \star {G}^\mathrm{A}_0 \right) \partial _{\pi _{0}} f(\pi _0) +\mathrm{c.c.} \end{aligned}\end{aligned}$$
(B2)

In the last equality we used that position of \(f(\pi _0)\) is irrelevant, see App. A. Very similarly, for \({{\mathcal {K}}}^{ij0}\) we have

$$\begin{aligned} \begin{aligned} { {\mathcal {K}}}^{ij0}&= -\frac{1}{4 \, {{\mathcal {V}}}} \int \frac{d^3\pi d^2 x}{(2\pi )^3 } \\&\quad \mathop {\mathrm{tr}}\nolimits \left( \partial _{\pi _{i}} {Q}^\mathrm{R}_0 \star {G}^\mathrm{R}_0\star \partial _{\pi _{j}} {Q}^\mathrm{R}_0 \star \partial _{\pi _{0}}G^\mathrm{R}_0 \right) f(\pi _0)\\&\quad + \frac{1}{4 \, {{\mathcal {V}}}} \int \frac{d^3\pi d^2 x}{(2\pi )^3 }\\&\quad \mathop {\mathrm{tr}}\nolimits \left( \partial _{\pi _{i}}Q^\mathrm{A}_0 \star {G}^\mathrm{A}_0\star \partial _{\pi _{j}} {Q}^\mathrm{A}_0 \star \partial _{\pi _{0}} {G}^\mathrm{A}_0 \right) f(\pi _0)\\&\quad + \frac{1}{4 \, {{\mathcal {V}}}} \int \frac{d^3\pi d^2 x}{(2\pi )^3 }\\&\quad \mathop {\mathrm{tr}}\nolimits \left( \partial _{\pi _{i}}Q^\mathrm{R}_0\star {G}^\mathrm{R}_0\star \partial _{\pi _{j}} {Q}^\mathrm{R}_0 \star ( {G}^\mathrm{A}_0- {G}^\mathrm{R}_0) \right) \partial _{\pi _{0}} f(\pi _0)+\mathrm{c.c.} \end{aligned}\end{aligned}$$
(B3)

Both in (B2) and (B3) one might be tempted to put in the limit \(\epsilon \rightarrow 0\)

$$\begin{aligned}&\partial _{\pi _{j}} {Q}^\mathrm{R}_0 \star ( {G}^\mathrm{A}_0- {G}^\mathrm{R}_0) = - {Q}^\mathrm{R}_0\star \partial _{\pi _{j}} {G}^\mathrm{R}_0\star {Q}^\mathrm{R}_0 \star ( {G}^\mathrm{A}_0- {G}^\mathrm{R}_0)=0, \end{aligned}$$

alleging that \({Q}^\mathrm{R}_0 \star ( {G}^\mathrm{A}_0- {G}^\mathrm{R}_0) = {{\mathcal {O}}}(\epsilon )\). While the latter is true, the former is not legitimate since near the pole \(\partial _{\pi _{j}} {G}^\mathrm{R}_0 \sim 1/\epsilon ^2\) and the whole expression is not vanishing.

Thus, for the Hall conductivity we obtain

$$\begin{aligned} \begin{aligned} {-}\sigma _H&= \frac{1}{2\pi }\times \frac{1}{48\pi ^2 \, {{\mathcal {V}}}} \epsilon ^{\mu \nu \rho } \int {d^3\pi d^2 x}\\&\quad \mathop {\mathrm{tr}}\nolimits \left( \partial _{\pi ^{\mu }} {Q}^\mathrm{R}_0 \star \partial _{\pi ^{\nu }} {G}^\mathrm{R}_0\star \partial _{\pi ^{\rho }} {Q}^\mathrm{R}_0 \star G^\mathrm{R}_0 \right) f(\pi ^0)\\&\quad - \frac{1}{2\pi }\times \frac{1}{48\pi ^2 \, {{\mathcal {V}}}} \epsilon ^{\mu \nu \rho }\int {d^3\pi d^2 x} \\&\quad \mathop {\mathrm{tr}}\nolimits \left( \partial _{\pi ^{\mu }}Q^\mathrm{A}_0 \star \partial _{\pi ^{\nu }} {G}^\mathrm{A}_0\star \partial _{\pi ^{\rho }} {Q}^\mathrm{A}_0 \star {G}^\mathrm{A}_0 \right) f(\pi ^0)\\&\quad + \frac{1}{8 \, {{\mathcal {V}}}} \epsilon ^{ij}\int \frac{d^3\pi d^2 x}{(2\pi )^3 } \\&\quad \mathop {\mathrm{tr}}\nolimits \left( \partial _{\pi ^{i}}Q^\mathrm{R}_0\star ( {G}^\mathrm{A}_0- {G}^\mathrm{R}_0) \star \partial _{\pi ^{j}} {Q}^\mathrm{A}_0 \star {G}^\mathrm{A}_0 \right) \partial _{\pi ^{0}} f(\pi ^0)\\&\quad - \frac{1}{8 \, {{\mathcal {V}}}}\epsilon ^{ij} \int \frac{d^3\pi d^2 x}{(2\pi )^3 } \\&\quad \mathop {\mathrm{tr}}\nolimits \left( \partial _{\pi ^{i}}Q^\mathrm{R}_0\star {G}^\mathrm{R}_0\star \partial _{\pi ^{j}} {Q}^\mathrm{R}_0 \star ( {G}^\mathrm{A}_0- {G}^\mathrm{R}_0) \right) \partial _{\pi ^{0}} f(\pi ^0)+\mathrm{c.c.}\\&= \frac{1}{2\pi } {{\mathcal {N}}}_f + \epsilon ^{ij} {{\mathcal {A}}}_{ij}. \end{aligned} \end{aligned}$$
(B4)

Here

$$\begin{aligned} {{\mathcal {A}}}_{ij}= & {} \frac{1}{8 \, {{\mathcal {V}}}} \int \frac{d^3\pi \, d^2 x}{(2\pi )^3 } \mathop {\mathrm{tr}}\nolimits \left( \left( \partial _{\pi ^{{j}}}Q^\mathrm{A}_0\star {G}^\mathrm{A}_0 \star \partial _{\pi ^{{i}}} {Q}^\mathrm{R}_0\nonumber \right. \right. \\&\left. \left. -\partial _{\pi ^{{i}}}Q^\mathrm{R}_0\star {G}^\mathrm{R}_0 \star \partial _{\pi ^{{j}}} {Q}^\mathrm{R}_0 \right) \nonumber \right. \\&\left. \star ( {G}^\mathrm{A}_0- {G}^\mathrm{R}_0) \right) \partial _{\pi _{0}} f(\pi _0) +\mathrm{c.c.}, \end{aligned}$$
(B5)

and

$$\begin{aligned} {{\mathcal {N}}}_f= & {} -\frac{1}{48\pi ^2 \, {{\mathcal {V}}}} \epsilon ^{{ijk}} \oint d\pi _0 \int {d^2 \mathbf {\pi } d^2 x} \nonumber \\&\mathop {\mathrm{tr}}\nolimits \left( G_0\star \partial _{\pi ^{{i}}} {Q}_0 \star \partial _{\pi ^{{j}}} {G}_0\star \partial _{\pi ^{{k}}} {Q}_0 \right) f(\pi _0) +\mathrm{c.c.}, \end{aligned}$$
(B6)

where \(\oint\) is an integral over the contour encompassing the whole real axis in positive direction, while

$$\begin{aligned} \begin{aligned} Q_0(x,\pi )&= \pi _0- {H}_W(x,\pi ), \\ G_0(x,\pi )&\star Q_0(x,\pi ) =1. \end{aligned}\end{aligned}$$
(B7)

In a similar way for the longitudinal conductivity we obtain

$$\begin{aligned} \sigma _\parallel ^{ij} = {{\mathcal {A}}}_{\{ij\}} + \mathrm{c.c.} \end{aligned}$$
(B8)

Appendix C: Conductivity in terms of the velocity operator

Let us now rewrite (B4) in terms of the matrix elements of the velocity operator, similar to the derivation given in [63]. For this end we shall use that the trace of the Weyl symbols over the phase space is equal to the functional trace of a product of corresponding operators, for instance, given by the trace of their matrix elements over momentum space

$$\begin{aligned} \begin{aligned} \mathop {\mathrm{Tr}}\nolimits (A_W *B_W)&\equiv \int d^3 X \int \frac{d^3 P}{(2\pi )^3}\mathop {\mathrm{tr}}\nolimits (A_W *B_W) \\&= {\mathop {\mathrm{Tr}}\nolimits } \hat{A} \hat{B} = \int d^3 P d^3 Q A(P,Q) B(Q,P), \end{aligned} \end{aligned}$$
(C1)

where \(P^i = (p_0,p_1,p_2) = (\omega ,p)\) and \(X^i = (t,x_1,x_2) = (t,x)\). Applying this formula to Eq. (B6) we come to

$$\begin{aligned} \begin{aligned} {{\mathcal {N}}}_f&= -\frac{\epsilon ^{\mu \nu \rho }}{48\pi ^2 \, {{\mathcal {V}}}} \oint d \omega ^{(1)} \, f(\omega ^{(1)}) \int d^2 p^{(1)} \prod _{i=2}^4 d^3 P^{(i)} \\&\quad \mathop {\mathrm{tr}}\nolimits \Bigg [ {G}(P^{(1)},P^{(2)}) \left[ \partial _{P^{(2)}_\mu } + \partial _{P^{(3)}_\mu }\right] {Q}(P^{(2)},P^{(3)})\\&\quad \left( \left[ \partial _{P^{(3)}_\nu } + \partial _{P^{(4)}_\nu }\right] {G}(P^{(3)},P^{(4)}) \right) \\&\quad \left[ \partial _{P^{(4)}_\rho } + \partial _{P^{(1)}_\rho }\right] {Q}(P^{(4)},P^{(1)}) \Bigg ] +\mathrm{c.c.} \end{aligned} \end{aligned}$$
(C2)

For the non-interacting fermions described by Hamiltonian \({{\mathcal {H}}}\) with energy eigenstates \(|n\rangle\): \({{\mathcal {H}}}|n\rangle = {{\mathcal {E}}}_n |n\rangle\), the matrix elements in the above are given by

$$\begin{gathered} Q\left( {P^{{(1)}} ,P^{{(2)}} } \right) \equiv \left\langle {P^{{(1)}} |\hat{Q}|P^{{(2)}} } \right\rangle \hfill \\ \quad \quad \quad \quad \quad \,\, = \left( {\delta ^{{(2)}} (p^{{(1)}} - p^{{(2)}} )\omega ^{{(1)}} - \langle p^{{(1)}} |{\mathcal{H}}|p^{{(2)}} \rangle } \right)\delta \left( {\omega ^{{(1)}} - \omega ^{{(2)}} } \right) \hfill \\ G\left( {P^{{(1)}} ,P^{{(2)}} } \right) = \delta \left( {\omega ^{{(1)}} - \omega ^{{(2)}} } \right)\sum\limits_{n} {\frac{{\left\langle {p^{{(1)}} |n} \right\rangle \left\langle {n|p^{{(2)}} } \right\rangle }}{{\omega ^{{(1)}} - {\mathcal{E}}_{n} }}} . \hfill \\ \end{gathered}$$
(C3)

here \(\sum _n\) may stand both for discrete spectrum summation, and integration \(\int dn\) in the case of continuum one.

To perform further simplifications, we note that

$$\begin{aligned} \partial _{p_i} G = - G \partial _{p_i} Q G , \end{aligned}$$
(C4)

and more importantly,

$$\begin{aligned} \left[ \partial _{p^{(4)}_j} + \partial _{p^{(1)}_j} \right] \langle p ^{(4)}| {{\mathcal {H}}} | p ^{(1)} \rangle= & {} \mathrm {i}\langle p ^{(4)}| {{\mathcal {H}}} {{\hat{x}} }_j \\&-{{\hat{x}} }_j{{\mathcal {H}}} | p ^{(1)}\rangle \equiv \langle p ^{(4)}| {\hat{v}}_j | p ^{(1)}\rangle , \end{aligned}$$

where we introduced the velocity operator \({{\hat{v}}}_i = \mathrm {i}[{{\mathcal {H}}}, {{\hat{x}} }_i]\). By operator \({\hat{x}}_i\) we understand \(\mathrm {i}\partial _{p_i}\) acting on the wavefunction written in momentum representation:

$$\begin{aligned} \hat{x}_j \Psi (p) = \langle p|\hat{x}_j |\Psi \rangle = \mathrm {i}\partial _{p_j} \langle p|\Psi \rangle = \mathrm {i}\partial _{p_j} \Psi (p). \end{aligned}$$

Then, for example,

$$\begin{aligned} \hat{x}_j \delta ^{(2)}(q-p)= & {} \langle p|\hat{x}_j |q\rangle = \mathrm {i}\partial _{p_j} \langle p|q \rangle \\= & {} \mathrm {i}\partial _{p_j} \delta ^{(2)}(p-q) = -\mathrm {i}\partial _{p_j}\langle q|p \rangle . \end{aligned}$$

Therefore, we can write

$$\begin{aligned} \hat{x}_j |p\rangle = -\mathrm {i}\partial _{p_j} |p \rangle . \end{aligned}$$

Using the above formulae, we derive that

$$\begin{aligned} \begin{aligned} {{\mathcal {N}}}_f&= {+}\frac{\epsilon ^{ij} }{4\, {{\mathcal {V}}}}\,\sum _{n,k} \int \prod _{l=1}^4 d^2 p ^{(l)} \\&\quad \oint \frac{ f(\omega ) d \omega }{(\omega ^{}-{{\mathcal {E}}}_n)^2 (\omega ^{}-{{\mathcal {E}}}_k)} \langle p ^{(1)}| n \rangle \langle n p ^{(2)}\rangle \langle p ^{(2)}| {\hat{v}}_i | p ^{(3)}\rangle \\&\quad \langle p ^{(3)}| k \rangle \langle k | p ^{(4)}\rangle \langle p ^{(4)}| {\hat{v}}_j | p ^{(1)}\rangle + \mathrm{c.c.} \\&= {+}\frac{2\pi \mathrm {i}\, \epsilon ^{ij} }{4 \, {{\mathcal {V}}}} \sum _{n,k } \frac{ f({{\mathcal {E}}}_k)-f({{\mathcal {E}}}_n) + ({{\mathcal {E}}}_n-{{\mathcal {E}}}_k) f'({{\mathcal {E}}}_n)}{({{\mathcal {E}}}_k-{{\mathcal {E}}}_n)^2}\\&\quad \langle n | {\hat{v}}_i | k \rangle \langle k | {\hat{v}}_j | n \rangle + \mathrm{c.c.} \end{aligned}\end{aligned}$$
(C5)

We used here that the momentum eigenvalues compose a full set, \(\int d^2p\, |p\rangle \langle p|=1\). Note, that in the case of a discreet spectrum, the term \(n=k\) should be understood as a limit \({{\mathcal {E}}}_n\rightarrow {{\mathcal {E}}}_k\), which gives a finite result.

For the non-topological contribution to \(\sigma _H\) and for \(\sigma _\Vert\) we shall similarly analyze \({{\mathcal {A}}}\) given by (B5). Advanced and retarded components needed for its calculation can be obtained from (C3) as

$$\begin{aligned} Q^{\mathrm{A}/\mathrm{R}}(p^{(1)},p^{(2)})= & {} \left( \delta ^{(2)} (p^{(1)}-p^{(2)})\, \left[ \omega ^{(1)}\pm \mathrm {i}\epsilon \right] \nonumber \right. \\&\left. - \langle p^{(1)}| {{\mathcal {H}}} | p^{(2)}\rangle \right) \delta \left( \omega ^{(1)}-\omega ^{(2)}\right) \nonumber \\&G^{\mathrm{A}/\mathrm{R}}(P^{(1)},P^{(2)}) = \delta \left( \omega ^{(1)}-\omega ^{(2)}\right) \nonumber \\&\sum _{n} \frac{\langle p ^{(1)}| n \rangle \langle n | p ^{(2)}\rangle }{\omega ^{(1)}-{{\mathcal {E}}}_n\pm \mathrm {i}\epsilon } . \end{aligned}$$
(C6)

and thus,

$$\begin{aligned}&\left( {G}^\mathrm{A}_0-{G}^\mathrm{R}_0\right) \left( P^{(1)},P^{(2)}\right) =2\pi \mathrm {i}\, \delta (\omega ^{(1)}-\omega ^{(2)}) \nonumber \\&\quad \sum _{n} \delta (\omega ^{(1)}-{{\mathcal {E}}}_n) \langle p ^{(1)} | n \rangle \langle n | p ^{(2)}\rangle . \end{aligned}$$
(C7)

Then

$$\begin{aligned}&\left( \partial _{\pi _{j}}Q^\mathrm{A}_0 \hat{G}^\mathrm{A}_0 \partial _{\pi _{i}}\hat{Q}^\mathrm{R}_0 \right) \left( P^{(1)},P^{(4)}\right) \nonumber \\&\quad = \delta \left( \omega ^{(1)}-\omega ^{(4)}\right) \int dp^{(2)}dp^{(3)} \langle p ^{(1)}| {\hat{v}}_j | p ^{(2)}\rangle \frac{\left\langle p ^{(2)}| n \right\rangle \left\langle n | p^{(3)}\right\rangle }{\omega ^{(1)}-{{\mathcal {E}}}_n + \mathrm {i}\epsilon } \nonumber \\&\qquad \left\langle p ^{(3)}| {\hat{v}}_i | p ^{(4)}\right\rangle , \end{aligned}$$
(C8)

and

$$\begin{aligned}&(\partial _{\pi _{i}}Q^\mathrm{R}_0 \hat{G}^\mathrm{R}_0 \partial _{\pi _{j}}\hat{Q}^\mathrm{R}_0 )(P^{(1)},P^{(4)}) \nonumber \\&\quad = \delta (\omega ^{(1)}-\omega ^{(4)}) \int dp^{(2)}dp^{(3)} \langle p ^{(1)}| {\hat{v}}_i | p ^{(2)}\rangle \nonumber \\&\quad \frac{\langle p ^{(2)}| n \rangle \langle n | p^{(3)}\rangle }{\omega ^{(1)}-{{\mathcal {E}}}_n - \mathrm {i}\epsilon } \langle p ^{(3)}| {\hat{v}}_j | p ^{(4)}\rangle . \end{aligned}$$
(C9)

All together it gives

$$\begin{aligned} {{\mathcal {A}}}_{ij}= & {} {-} \frac{ \mathrm {i}}{8 \, {{\mathcal {V}}}} \sum _{n,k} f' ({{\mathcal {E}}}_k) \nonumber \\&\left[ \frac{\langle k| {\hat{v}}_j | n \rangle \langle n | {\hat{v}}_i | k \rangle }{{{\mathcal {E}}}_k-{{\mathcal {E}}}_n + \mathrm {i}\epsilon }\nonumber \right. \\&\left. -\frac{\langle k| {\hat{v}}_i | n \rangle \langle n | {\hat{v}}_j | k \rangle }{{{\mathcal {E}}}_k-{{\mathcal {E}}}_n - \mathrm {i}\epsilon }\right] +\mathrm{c.c.}, \end{aligned}$$
(C10)

So, that

$$\begin{aligned} \begin{aligned} {{\mathcal {A}}}_{\{ij\}}&= {-} \frac{\mathrm {i}}{8 \, {{\mathcal {V}}}} \sum _{n,k} f' ({{\mathcal {E}}}_k) \left[ \frac{1}{{{\mathcal {E}}}_k-{{\mathcal {E}}}_n + \mathrm {i}\epsilon }\right. \\&\left. \quad -\frac{1}{{{\mathcal {E}}}_k-{{\mathcal {E}}}_n - \mathrm {i}\epsilon }\right] \left( \langle k| {\hat{v}}_j | n \rangle \langle n | {\hat{v}}_i | k \rangle \right. \\&\quad \left. +\langle k| {\hat{v}}_i | n \rangle \langle n | {\hat{v}}_j | k \rangle \right) +\mathrm{c.c.}, \\ {{\mathcal {A}}}_{[ij]}&= {-} \frac{ \mathrm {i}}{8 \, {{\mathcal {V}}}} \sum _{n,k} f' ({{\mathcal {E}}}_k)\\&\quad \left[ \frac{1}{{{\mathcal {E}}}_k-{{\mathcal {E}}}_n + \mathrm {i}\epsilon } + \frac{1}{{{\mathcal {E}}}_k-{{\mathcal {E}}}_n - \mathrm {i}\epsilon }\right] \\&\quad \left( \langle k| {\hat{v}}_j | n \rangle \langle n | {\hat{v}}_i | k \rangle - \langle k| {\hat{v}}_i | n \rangle \langle n | {\hat{v}}_j | k \rangle \right) +\mathrm{c.c.}, \end{aligned}\end{aligned}$$
(C11)

The limit \(\epsilon \rightarrow 0\) of these expressions depends on the nature of the spectrum. In the continuum case, the Sokhotski-Plemelj formula gives

$$\begin{aligned}&\frac{1}{{{\mathcal {E}}}_k-{{\mathcal {E}}}_n + \mathrm {i}\epsilon } - \frac{1}{{{\mathcal {E}}}_k-{{\mathcal {E}}}_n - \mathrm {i}\epsilon } = - 2 \pi \mathrm {i}\, \delta ({{\mathcal {E}}}_k-{{\mathcal {E}}}_n), \qquad \nonumber \\&\frac{1}{{{\mathcal {E}}}_k-{{\mathcal {E}}}_n + \mathrm {i}\epsilon } + \frac{1}{{{\mathcal {E}}}_k-{{\mathcal {E}}}_n - \mathrm {i}\epsilon } = 2 {{\mathcal {P}}}\frac{1}{{{\mathcal {E}}}_k-{{\mathcal {E}}}_n}, \end{aligned}$$
(C12)

while in the discrete case, the expression for \({{\mathcal {A}}}_{\{ij\}}\) (and thus, for symmetric conductivity) will be divergent in \(\epsilon \rightarrow 0\)

$$\begin{aligned}&\frac{1}{{{\mathcal {E}}}_k-{{\mathcal {E}}}_n + \mathrm {i}\epsilon } - \frac{1}{{{\mathcal {E}}}_k-{{\mathcal {E}}}_n - \mathrm {i}\epsilon } = \left\{ \begin{array}{ll} 0, &{} n\ne k\\ \frac{2}{\mathrm {i}\epsilon }, &{} n = k \end{array}\right. , \nonumber \\&\frac{1}{{{\mathcal {E}}}_k-{{\mathcal {E}}}_n + \mathrm {i}\epsilon } + \frac{1}{{{\mathcal {E}}}_k-{{\mathcal {E}}}_n - \mathrm {i}\epsilon } = \left\{ \begin{array}{ll} \frac{2}{{{\mathcal {E}}}_k-{{\mathcal {E}}}_n}, &{} n\ne k\\ 0, &{} n = k \end{array}\right. \end{aligned}$$
(C13)

Summarizing, we have

$$\begin{aligned} \sigma _H= {- \sigma _{xy} } &= - \frac{\mathrm {i}\, \epsilon ^{ij} }{4 \, {{\mathcal {V}}}}\,\sum _{n,k} \left( \frac{ f({{\mathcal {E}}}_k)-f({{\mathcal {E}}}_n) + ({{\mathcal {E}}}_n-{{\mathcal {E}}}_k) f'({{\mathcal {E}}}_n)}{({{\mathcal {E}}}_k-{{\mathcal {E}}}_n)^2} \nonumber \right. \\& \quad \left. + f'({{\mathcal {E}}}_k) {{\mathcal {P}}}\frac{1}{{{\mathcal {E}}}_k-{{\mathcal {E}}}_n}\right) \langle n | {\hat{v}}_i | k \rangle \langle k | {\hat{v}}_j | n \rangle + \mathrm{c.c.} \end{aligned}$$
(C14)

We can use this expression both for continuum and discreet spectrum if in the latter case we put \({{\mathcal {P}}}\frac{1}{{{\mathcal {E}}}_k-{{\mathcal {E}}}_n}=0\), \({{\mathcal {E}}}_n={{\mathcal {E}}}_k\). One can see, that in the absence of the singularities at \({{\mathcal {E}}}_n = {{\mathcal {E}}}_k\) the term with \(f^\prime\) is cancelled. In Eq. (C14) the singularity is isolated in the second term in the brackets while the first term remains finite at \({{\mathcal {E}}}_n = {{\mathcal {E}}}_k\) (it is reduced to \(f^{\prime \prime }({{\mathcal {E}}}_n)/2\)). It is worth mentioning that one can rewrite the whole expression in the following alternative form:

$$\begin{aligned} \sigma _H= & {} {-}\frac{\mathrm {i}\, \epsilon ^{ij} }{2 \, {{\mathcal {V}}}}\,\sum _{n,k} \frac{ f({{\mathcal {E}}}_k)}{({{\mathcal {E}}}_k-{{\mathcal {E}}}_n-i\epsilon )({{\mathcal {E}}}_k-{{\mathcal {E}}}_n+i\epsilon )} \nonumber \\&\langle n | {\hat{v}}_i | k \rangle \langle k | {\hat{v}}_j | n \rangle + \mathrm{c.c.} \end{aligned}$$
(C15)

Written in this form it coincides with expression proposed in [93] (see also [94]). In order to show equivalence of Eqs. (C15) and (C14) let us represent the quotient from the former as follows:

$$\begin{aligned} \begin{aligned} \frac{ f({{\mathcal {E}}}_k)-f({{\mathcal {E}}}_n)}{({{\mathcal {E}}}_k-{{\mathcal {E}}}_n-i\epsilon )({{\mathcal {E}}}_k-{{\mathcal {E}}}_n+i\epsilon )}&= \frac{ f({{\mathcal {E}}}_k)-f({{\mathcal {E}}}_n) + ({{\mathcal {E}}}_n-{{\mathcal {E}}}_k) f^{\prime }({{\mathcal {E}}}_n)}{({{\mathcal {E}}}_k-{{\mathcal {E}}}_n-i\epsilon )({{\mathcal {E}}}_k-{{\mathcal {E}}}_n+i\epsilon )} \\&\quad - \frac{ ({{\mathcal {E}}}_n-{{\mathcal {E}}}_k -i\epsilon ) f^{\prime }({{\mathcal {E}}}_n)}{2({{\mathcal {E}}}_k-{{\mathcal {E}}}_n-i\epsilon )({{\mathcal {E}}}_k-{{\mathcal {E}}}_n+i\epsilon )}\\&\quad -\frac{ ({{\mathcal {E}}}_n-{{\mathcal {E}}}_k+i\epsilon ) f^{\prime }({{\mathcal {E}}}_n)}{2({{\mathcal {E}}}_k-{{\mathcal {E}}}_n-i\epsilon )({{\mathcal {E}}}_k-{{\mathcal {E}}}_n+i\epsilon )} \\&= \frac{ f({{\mathcal {E}}}_k)-f({{\mathcal {E}}}_n) + ({{\mathcal {E}}}_n-{{\mathcal {E}}}_k) f^{\prime }({{\mathcal {E}}}_n)}{({{\mathcal {E}}}_k-{{\mathcal {E}}}_n-i\epsilon )({{\mathcal {E}}}_k-{{\mathcal {E}}}_n+i\epsilon )} \\&\quad + \frac{ f^{\prime }({{\mathcal {E}}}_n)}{2({{\mathcal {E}}}_k-{{\mathcal {E}}}_n+i\epsilon )}+\frac{ f'({{\mathcal {E}}}_n)}{2({{\mathcal {E}}}_k-{{\mathcal {E}}}_n-i\epsilon )} \\&= \frac{ f({{\mathcal {E}}}_k)-f({{\mathcal {E}}}_n) + ({{\mathcal {E}}}_n-{{\mathcal {E}}}_k) f^{\prime }({{\mathcal {E}}}_n)}{({{\mathcal {E}}}_k-{{\mathcal {E}}}_n-i\epsilon )({{\mathcal {E}}}_k-{{\mathcal {E}}}_n+i\epsilon )} \\&\quad + f'({{\mathcal {E}}}_n) {{\mathcal {P}}}\frac{1 }{({{\mathcal {E}}}_k-{{\mathcal {E}}}_n)} \end{aligned}\end{aligned}$$
(C16)

Appendix D: Hall conductivity for the noninteracting 2D system in the presence of constant magnetic field

Here we demonstrate how the derived expressions allow to obtain final expressions for the conductivity. We take as an example the simplest system of free non-relativistic electrons in the presence of constant magnetic field. The one-particle Hamiltonian is taken in its simplest form

$$\begin{aligned} {{\mathcal {H}}} = \frac{1}{2m}\left( \pi _1^2 + \pi _2^2\right) -\mu \end{aligned}$$

with \(\pi _1 = \hat{p}_1\) and \(\pi _2 =\hat{p}_2-{{\mathcal {B}}}x_1\). We have the following property specific for this Hamiltonian to be used further:

$$\begin{aligned} \epsilon ^{ij}\pi _i {{\mathcal {H}}} \pi _j = 3i{{\mathcal {B}}} {{\mathcal {H}}} \end{aligned}$$

The average Hall conductivity may be represented as

$$\begin{aligned} \bar{\sigma }_H= & {} -\frac{\mathrm {i}}{4 {{\mathcal {V}}}}\, \epsilon ^{ij} \, \Big (\sum _{n,k|{{\mathcal {E}}}_n\ne {{\mathcal {E}}}_k} \frac{ f({{\mathcal {E}}}_k)-f({{\mathcal {E}}}_n) }{({{\mathcal {E}}}_k-{{\mathcal {E}}}_n)^2} \nonumber \\&+ \frac{1}{2}\sum _{n,k|{{\mathcal {E}}}_n={{\mathcal {E}}}_k}f^{\prime \prime }({{\mathcal {E}}}_n)\Big ) \langle n|{{\hat{v}}}_i| k \rangle \langle k | {{\hat{v}}}_j | n \rangle + \mathrm{c.c.} \end{aligned}$$
(D1)

In order to calculate the value of \(\bar{\sigma }_{H}\) we decompose the coordinates \(x_1, x_2\) as follows:

$$\begin{aligned} \hat{x}_1= & {} -\frac{\hat{p}_2-{{\mathcal {B}}}x_1}{{\mathcal {B}}} + \hat{{\mathcal {X}}}_1 = \hat{\xi }_1 + \hat{{\mathcal {X}}}_1,\\ \hat{x}_2= & {} \frac{\hat{p}_1}{{\mathcal {B}}} + \hat{{\mathcal {X}}}_2= \hat{\xi }_2 + \hat{{\mathcal {X}}}_2\,. \end{aligned}$$

The commutation relations follow:

$$\begin{aligned}&[\hat{\xi }_1,\hat{\xi }_2] = \frac{i}{{\mathcal {B}}}\,, \\&[\hat{{\mathcal {X}}}_1,\hat{{\mathcal {X}}}_2] =-\frac{i}{{\mathcal {B}}}\,,\\&[{{\mathcal {H}}}, \xi _1] = -i \frac{\partial }{\partial p_1} {{\mathcal {H}}}\,, \quad \\&[{{\mathcal {H}}}, \xi _2] = -i \frac{\partial }{\partial p_2} {{\mathcal {H}}}\,,\\&[{{\mathcal {H}}}, \hat{{\mathcal {X}}}_1] = [{{\mathcal {H}}}, \hat{{\mathcal {X}}}_2] = 0\,. \end{aligned}$$

Here we use that the Hamiltonian contains the following dependence on x:

$$\begin{aligned} {{\mathcal {H}}}(\hat{p}_1, \hat{p}_2-{{\mathcal {B}}} x_1)\, \end{aligned}$$

and \(\frac{\partial ^2}{\partial p_1 \partial p_2}{{\mathcal {H}}}=0\). Thus we obtain:

$$\begin{aligned} \bar{\sigma }_H= & {} -\frac{\mathrm {i}}{2 {{\mathcal {V}}}}\, \epsilon ^{ij} \, \Big (\sum _{n,k|{{\mathcal {E}}}_n\ne {{\mathcal {E}}}_k} \frac{ f({{\mathcal {E}}}_k)-f({{\mathcal {E}}}_n) + ({{\mathcal {E}}}_n-{{\mathcal {E}}}_k) f'({{\mathcal {E}}}_n)}{({{\mathcal {E}}}_k-{{\mathcal {E}}}_n)^2} \nonumber \\&+ \frac{1}{2}\sum _{n,k|{{\mathcal {E}}}_n={{\mathcal {E}}}_k}f^{\prime \prime }({{\mathcal {E}}}_n)\Big )\langle n| [{{\mathcal {H}}}, {{\hat{\xi }}}_i] | k \rangle \langle k | [{{\mathcal {H}}}, {{\hat{\xi }}}_j] | n \rangle \nonumber \\= & {} \frac{\mathrm {i}}{2 {{\mathcal {V}}}}\, \epsilon ^{ij} \, \Bigl (- \sum _{n,k|{{\mathcal {E}}}_n\ne {{\mathcal {E}}}_k} 2f({{\mathcal {E}}}_n)\nonumber \\&+ \frac{1}{2}\sum _{n,k|{{\mathcal {E}}}_n={{\mathcal {E}}}_k}f^{\prime \prime }({{\mathcal {E}}}_n)({{\mathcal {E}}}_k-{{\mathcal {E}}}_n)^2 \Bigr )\,\nonumber \\&\Big [ \langle n| {{\hat{\xi }}}_i | k \rangle \langle k | {{\hat{\xi }}}_j | n \rangle \Big ] \nonumber \\= & {} \frac{\mathrm {i}}{2 {{\mathcal {V}}}}\, \epsilon ^{ij} \, \Bigl (-\sum _{n,k} 2f({{\mathcal {E}}}_n) \nonumber \\&+ \sum _{n,k|{{\mathcal {E}}}_n={{\mathcal {E}}}_k} 2f({{\mathcal {E}}}_n)\Bigr ) \Big [ \langle n| {{\hat{\xi }}}_i | k \rangle \langle k | {{\hat{\xi }}}_j | n \rangle \Big ] \nonumber \\= & {} \frac{\mathrm {i}}{2 {{\mathcal {V}}}}\, \sum _{n} \nonumber \\&\Bigl ( -2f({{\mathcal {E}}}_n)\Bigr )\,\Big [ \langle n| [{{\hat{\xi }}}_1, {{\hat{\xi }}}_2] | n \rangle \Big ] \nonumber \\= & {} \frac{1 }{{{\mathcal {B}}} {{\mathcal {V}}}}\, \sum _{n} f({{\mathcal {E}}}_n)\, \langle n| n \rangle = \frac{\rho }{{\mathcal {B}}}. \end{aligned}$$
(D2)

Here \(\rho\) is density of electrons. We used here that momentum \(p_2\) is a good quantum number, and it enumerates the eigenstates of the Hamiltonian:

$$\begin{aligned}&{{\mathcal {H}}} |n\rangle = {{\mathcal {H}}}(\hat{p}_1, p_y-{{\mathcal {B}}} x_1)|p_2, q\rangle = {{\mathcal {E}}}_{q}|p_2, q\rangle , \quad q\in \mathbb {N}\,. \end{aligned}$$

We assume that the size of the system is \(L\times L\). Properties of the eigenstates of Hamiltonian guarantee that \(\langle p^\prime _2, q| \hat{p}_1|p_2, q\rangle =0\) for \(p^\prime _2 \ne p_2\). This gives

$$\begin{aligned} \bar{\sigma }_H= & {} \,\sum _{q}\int \frac{dp_2 L}{2\pi {{\mathcal {V}}}} \, \frac{ f({{\mathcal {E}}}_q)}{{\mathcal {B}}}. \end{aligned}$$
(D3)

\(\langle x_1 \rangle = p_2/{{\mathcal {B}}}\) plays the role of the center of orbit, and this center should belong to the interval \((-L/2, L/2)\) while \({{\mathcal {E}}}_q\) does not depend on momentum. This gives

$$\begin{aligned} \bar{\sigma }_H= & {} \frac{1}{2\pi } \,\sum _{q = 0,1,...} \, f({{\mathcal {E}}}_q). \end{aligned}$$
(D4)

In case of thermal equilibrium this expression receives the form:

$$\begin{aligned} \bar{\sigma }_H= & {} \frac{1}{2\pi } \,\sum _{q=0,1,...} \, \frac{1}{e^{{{\mathcal {E}}}_q/T}+1}. \end{aligned}$$
(D5)

Here

$$\begin{aligned} {{\mathcal {E}}}_q = \frac{{\mathcal {B}}}{2m} (2q+1)-\mu \end{aligned}$$

where \(\mu\) is chemical potential. One can see, that at \(T \ll \frac{{\mathcal {B}}}{m}\) this expression is reduced to the zero temperature expression \(\bar{\sigma }_H = \frac{N}{2\pi }\), where N is the number of occupied Landau Levels.

Appendix E: Robustness of \({{\mathcal {N}}}_f\) with respect to modification of one-particle Hamiltonian

Let us consider the following quantity for the 2+1 dimensional system

$$\begin{aligned}&{{\mathcal {N}}} = \left ( \frac{\epsilon _{ijk}}{48 \pi ^2 \, {{\mathcal {V}}}} \int {d^3\pi d^2 x} \mathop {\mathrm{tr}}\nolimits \left( \hat{G}_0 \star \partial _{\pi _{i}}\hat{Q}_0 \star \hat{G}_0\star \partial _{\pi _{j}}\hat{Q}_0 \star \hat{G}_0\star \partial _{\pi _{k}}\hat{Q}_0 \right) ^< +\mathrm{c.c.}\right ) \end{aligned}$$
(E1)

We consider the static system with distribution function that depends only on \(\pi _0\). Variation of \({\mathcal {N}}\) caused by a variation of \(\hat{Q}\) (such that \(\delta f(\pi _0)=0\)) gives:

$$\begin{aligned} {\delta } {{\mathcal {N}}}= & {} \Big ( \frac{3\epsilon _{ijk}}{48 \pi ^2 \, {{\mathcal {V}}}} \int {d^3\pi d^2 x} \, \nonumber \\&\mathop {\mathrm{tr}}\nolimits \left( \delta \hat{G}_0 \star \partial _{\pi _{i}}\hat{Q}_0 \star \hat{G}_0\star \partial _{\pi _{j}}\hat{Q}_0 \star \hat{G}_0\star \partial _{\pi _{k}}\hat{Q}_0 \right) ^< +\mathrm{c.c.}\Big )\nonumber \\&+ \Big ( \frac{3\epsilon _{ijk}}{48 \pi ^2 \, {{\mathcal {V}}}} \int {d^3\pi d^2 x} \, \nonumber \\&\mathop {\mathrm{tr}}\nolimits \left( \hat{G}_0 \star \partial _{\pi _{i}}\delta \hat{Q}_0 \star \hat{G}_0\star \partial _{\pi _{j}}\hat{Q}_0 \star \hat{G}_0\star \partial _{\pi _{k}}\hat{Q}_0 \right) ^< +\mathrm{c.c.}\Big )\nonumber \\= & {} \Big ( -\frac{3\epsilon _{ijk}}{48 \pi ^2 \, {{\mathcal {V}}}} \int {d^3\pi d^2 x} \, \nonumber \\&\mathop {\mathrm{tr}}\nolimits \left( \delta \hat{G}_0 \star \partial _{\pi _{i}}\hat{Q}_0 \star \partial _{\pi _{j}}\hat{G}_0 \star \partial _{\pi _{k}}\hat{Q}_0 \right) ^< +\mathrm{c.c.}\Big )\nonumber \\&+ \Big ( -\frac{3\epsilon _{ijk}}{48 \pi ^2 \, {{\mathcal {V}}}} \int {d^3\pi d^2 x} \, \nonumber \\&\mathop {\mathrm{tr}}\nolimits \left( \hat{G}_0 \star \partial _{\pi _{i}}\delta \hat{Q}_0 \star \partial _{\pi _{j}}\hat{G}_0 \star \partial _{\pi _{k}}\hat{Q}_0 \right) ^< +\mathrm{c.c.}\Big )\nonumber \\= & {} \Big (- \frac{3\epsilon _{ijk}}{48 \pi ^2 \, {{\mathcal {V}}}} \int {d^3\pi d^2 x} \, \nonumber \\&\mathop {\mathrm{tr}}\nolimits \left( \hat{G}_0 \star \delta \hat{Q}_0 \star \hat{G}_0 \star \partial _{\pi _{i}}\hat{Q}_0 \star \hat{G}_0 \star \partial _{\pi _{j}}\hat{Q}_0 \star \hat{G}_0 \star \partial _{\pi _{k}}\hat{Q}_0 \right) ^< +\mathrm{c.c.}\Big )\nonumber \\&+ \Big ( \frac{3\epsilon _{ijk}}{48 \pi ^2 \, {{\mathcal {V}}}} \int {d^3\pi d^2 x} \, \nonumber \\&\mathop {\mathrm{tr}}\nolimits \left( \hat{G}_0 \star \partial _{\pi _{i}}\hat{Q}_0 \star \hat{G}_0 \star \delta \hat{Q}_0 \star \hat{G}_0 \star \partial _{\pi _{j}}\hat{Q}_0 \star \hat{G}_0 \star \partial _{\pi _{k}}\hat{Q}_0 \right) ^< +\mathrm{c.c.}\Big ) \end{aligned}$$
(E2)

Here similar to the case of quantities considered in Appendix A and Appendix 1 the above expressions obey the cyclic property provided that function \(f(\pi _0)\) remains unchanged and the contribution proportional to its derivative may be neglected. Further we simplify these expressions and obtain:

$$\begin{aligned} {\delta } {{\mathcal {N}}} &= \left(- \frac{3\epsilon _{ijk}}{48 \pi ^2 \, {{\mathcal {V}}}} \int {d^3\pi d^2 x} \, \mathop {\mathrm{tr}}\nolimits \left( \hat{G}_0 \star \delta \hat{Q}_0 \star \hat{G}_0 \star \partial _{\pi _{i}}\hat{Q}_0 \star \hat{G}_0 \star \partial _{\pi _{j}}\hat{Q}_0 \star \hat{G}_0 \star \partial _{\pi _{k}}\hat{Q}_0 \right) ^< +\mathrm{c.c.}\right)\nonumber \\& \quad + \left ( \frac{3\epsilon _{ijk}}{48 \pi ^2 \, {{\mathcal {V}}}} \int {d^3\pi d^2 x} \, \mathop {\mathrm{tr}}\nolimits \left( \hat{G}_0 \star \delta \hat{Q}_0 \star \hat{G}_0 \star \partial _{\pi _{j}}\hat{Q}_0 \star \hat{G}_0 \star \partial _{\pi _{k}}\hat{Q}_0 \star \hat{G}_0 \star \partial _{\pi _{i}}\hat{Q}_0 \right) ^< +\mathrm{c.c.}\right )=0 \end{aligned}$$
(E3)

Notice that the property proven here is not the complete topological invariance unlike the case of equilibrium at \(T=0\). The value of \({\mathcal {N}}\) may be changed smoothly under the change of distribution function \(f(\pi _0)\). Moreover, its value is modified also when \(f(\pi _0)\) remains unchanged but the terms in \({\mathcal {N}}\) proportional to the derivative of f gives valuable contribution.

Appendix F: Hall conductivity for the system of massive 2D Dirac fermions

The system of massive 2D Dirac fermions in equilibrium at zero temperature has been concerned in Sect. 5.2. It corresponds to

$$\begin{aligned} Q = {\varvec{1}} \omega -v_F (\sigma ^1 \pi _1 + \sigma ^2 \pi _2 + \sigma ^3 m). \end{aligned}$$

Here m is a mass-type parameter, \(\sigma _i\) are Pauli matrices, and \({\varvec{1}}\) is a unit \(2\times 2\) matrix. In equilibrium at \(T=0\) Hall conductivity is given by

$$\begin{aligned} \sigma _H = {-}\frac{{{\mathcal {N}}}}{2\pi }. \end{aligned}$$

with [83]

$$\begin{aligned} {{\mathcal {N}}}^{(0)} = \frac{1}{2} \mathrm{sign} \, m. \end{aligned}$$

Recall that for purely two-dimensional systems these fermions always come in pairs, and the total value of \({\mathcal {N}}\) is integer rather than half-integer.

Now let us calculate corrections to \(\sigma _H\) at finite temperatures using the developed formalism. For simplicity we consider the case when Fermi energy is set to zero. Our starting point is Eq. (C14) for Hall conductivity. The Hamiltonian can be written as follows:

$$\begin{aligned} H = v_F \left( \begin{array}{cc}m &{} p_1 - \mathrm {i}p_2 \\ p_1 +\mathrm {i}p_2 &{} - m \end{array}\right) \end{aligned}$$

In the following for simplicity we will consider the case \(v_F = 1\) (the nontrivial value of Fermi velocity may be easily restored in the final answer). The eigenvalues of this Hamiltonian are \(E_\pm ({p}) =\pm \sqrt{|{p}|^2+m^2}\). The corresponding eigenvectors are

$$\begin{aligned} |n \rangle\equiv & {} |a {p}\rangle = \frac{|{p}|}{\sqrt{2\sqrt{|{ p}|^2+m^2} \left(\sqrt{|{ p}|^2+m^2}+ a m\right)}} &\left( \begin{array}{c} 1 \\ - \frac{m -a \sqrt{|{ p}|^2+m^2}}{p_x +\mathrm {i}p_y} \end{array}\right) |{ p}\rangle \end{aligned}$$

here \(a= \pm 1\), with \(+1\) corresponding to the conductance band with positive energy while \(-1\) marks valence band with negative energy. Momenta eigenstates are normalized to 1 in discrete space, \(\langle {q}|{p}\rangle = (2\pi )^2 \delta ^{(2)}({q}-{p})/{{\mathcal {V}}}\).

Since velocity operator \(\hat{v}_k = \sigma _k\) does not contain momentum, its matrix elements between the states with definite momenta p and q contain a delta-function \(\delta ^{(2)}({p-q})\). In Eq. (C14) each of two sums over the quantum states is to be substituted by \({{\mathcal {V}}}\sum _{a \in \{c,v\}} \int \frac{d^2 p}{(2\pi )^2}\). We also denote

$$\begin{aligned}&\langle a,{p}|\hat{v}_i|b,{ p}^\prime \rangle \langle b,{ p}^\prime |\hat{v}_j|a,{p}\rangle \\&\quad = \langle a,b;i,j;{p} \rangle \frac{(2\pi )^2}{{\mathcal {V}}} \delta ^{(2)}({p}-{p}^\prime ) \frac{(2\pi )^2}{{\mathcal {V}}} \delta ^{(2)}({p}-{p}^\prime ) \\&\quad = \langle a,b;i,j;{p} \rangle \frac{(2\pi )^2}{{\mathcal {V}}} \delta ^{(2)}({p}'-{p}). \end{aligned}$$

In the last equality we used the fact that the factor \(\frac{(2\pi )^2}{{\mathcal {V}}} \delta ^{(2)}(0)\) is to be replaced by unity, which becomes clear if we consider the system inside a large but finite rectangular box with periodic boundary conditions and replace the integral over momentum by the sum over its discrete values. Furthermore, it is easy to see that \(\langle a,b;i,j; {p}\rangle =\langle a,b;j,i; {p}\rangle\). Using these notations we may rewrite Eq. (C14) as follows

$$\begin{aligned} \sigma _H= & {} {+} \frac{\mathrm {i}}{4}\int \frac{d^2p}{(2\pi )^2} \nonumber \\&\sum _{a,b = \pm 1} \Big [ \frac{f(E_a({p}))-f(E_b({p}))}{(E_a({p})-E_b({p}))^2} \Big ] \nonumber \\&\epsilon _{ij} \langle a,b;i,j;{p}\rangle +c.c., \end{aligned}$$
(F1)

where f is Fermi distribution.

Using the given above explicit expressions for the 2D Dirac spinors, we obtain \(\epsilon _{ij} \langle +,-;i,j;{p}\rangle =2\mathrm {i}\mathrm{Im} \langle +,-;1,2;{p}\rangle =2\mathrm {i}m/\sqrt{{p}^2+m^2}\). We represent \(\sigma _H = I/(8\pi ^2)\). Here

$$\begin{aligned} I= & {} -\mathrm {i}\int d^2p \sum _{a,b} \frac{f(E_a({p}))-f(E_b({p}))}{(E_a({p})-E_b({p}))^2} \epsilon _{ij} \langle a,b;i,j; {p}\rangle \nonumber \\= & {} -2\mathrm {i}\int d^2p \frac{f(E_+({p}))-f(E_-({p}))}{(E_+({p})-E_-({p}))^2} \epsilon _{ij} \langle c,v;i,j; {p}\rangle \nonumber \\= & {} -2\mathrm {i}\int d^2p \Big (\frac{1}{e^{-\beta \sqrt{{p}^2+m^2}}+1}-\frac{1}{e^{\beta \sqrt{{p}^2+m^2}}+1} \Big )\nonumber \\&\frac{1}{4({p}^2+m^2)}\frac{2\mathrm {i}m}{\sqrt{{p}^2+m^2}} \nonumber \\= & {} \int d^2p \frac{m}{({p}^2+m^2)^{3/2}}\, \mathrm{th}\left( \frac{\beta \sqrt{{p}^2+m^2}}{2} \right) \end{aligned}$$
(F2)

Then, after changing variables we obtain the following expression for Hall conductivity:

$$\begin{aligned} \sigma _H= & {} {-} \frac{\alpha }{4\pi } \int _{|\alpha |}^{+\infty } \frac{du}{u^2 \, \mathrm{th}(u/2)}. \end{aligned}$$
(F3)

where \(\alpha =v_F\beta m\equiv v_F m/T\) is dimensionless (we restored here Fermi velocity). This expression tends to \({-}\frac{1}{4\pi }\mathrm{sign}\,m\) at \(T\rightarrow 0\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Banerjee, C., Fialkovsky, I.V., Lewkowicz, M. et al. Wigner-Weyl calculus in Keldysh technique. J Comput Electron 20, 2255–2283 (2021). https://doi.org/10.1007/s10825-021-01775-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10825-021-01775-8

Keywords

Navigation