Kubo formula for non-Hermitian systems and tachyon optical conductivity

Doru Sticlet,1, ∗ Bal´azsD´ora,2, 3 and C˘at˘alinPa¸scuMoca4, 5 1National Institute for R&D of Isotopic and Molecular Technologies, 67-103 Donat, 400293 Cluj-Napoca, Romania 2MTA-BME Lend¨ulet Topology and Correlation Research Group, Budapest University of Technology and Economics, 1521 Budapest, Hungary 3Department of Theoretical Physics, Budapest University of Technology and Economics, 1521 Budapest, Hungary 4MTA-BME Dynamics and Correlations Research Group, Institute of Physics, Budapest University of Technology and Economics, 1521 Budapest, Hungary 5Department of Physics, University of Oradea, 410087, Oradea, Romania Linear response theory plays a prominent role in various fields of physics and provides us with extensive information about the thermodynamics and dynamics of quantum and classical systems. Here we develop a general theory for the linear response in non-Hermitian systems with non-unitary dynamics and derive a modified Kubo formula for the generalized susceptibility for arbitrary (Her- mitian and non-Hermitian) system and perturbation. As an application, we evaluate the dynamical response of a non-Hermitian, one-dimensional Dirac model with imaginary and real masses, per- turbed by a time-dependent electric field. The model has a rich phase diagram, and in particular, features a tachyon phase, where excitations travel faster than an effective speed of light. Surprisingly, we find that the dc conductivity of tachyons is finite, and the optical sum rule is exactly satisfied for all masses. Our results highlight the peculiar properties of the Kubo formula for non-Hermitian systems and are applicable for a large variety of settings.

Introduction. Linear response theory is a fundamen- periments, e.g. in a quantum gas microscopy experi- tal framework in statistical physics that describes the way ment [35, 36], the measurement backaction can quali- in which a physical system responds to small external tatively alter the low-energy physics, so it is crucial to perturbations [1]. The response is expressed in term of understand and describe the effect of small perturba- a generalized susceptibility of the unperturbed system. tions in such non-Hermitian systems. In this regard, The change in the expectation value of an A only recently, a non-Hermitian perturbation of a Her- due to a time-dependent perturbation V (t) = Bf(t) is mitian Hamiltonian has been considered, and a modi- assumed to be linear in the perturbative force f(t), fied Kubo formula for the generalized susceptibility has Z emerged [37]. Still, at present, a complete non-Hermitian 0 0 0 δhA(t)i = dt χAB(t, t )f(t ) , (1) linear response theory, in which the Hamiltonian gov- erning the system and the perturbation are both non- 0 with χAB(t, t ) known as the response function. In gen- Hermitian is missing. eral, χ (t, t0) is real, and for systems invariant under AB In the present work we address this problem, and fill time translations χ (t, t0) = χ (t − t0). Traditionally, AB AB this gap, by developing a general formalism and pro- the explicit expression for the response function in Eq. (1) viding a generalized Kubo formula for the susceptibility is given by the Kubo formula [2], and has been for decades χ (t, t0) introduced in Eq. (1) which is suitable for non- a standard tool in the study of transport phenomena in AB Hermitian systems under very general conditions. We il- typical many-body [3–5], closed [6,7] or open systems [8– lustrate our approach by computing the correlation func- 10]. Through the fluctuation-dissipation theorem, the tion for the current operator, in response to an exter- linear response provides information about the equilib- nal electric field, and the associated optical conductiv- rium fluctuations of a system, and can reveal e.g. topo- ity for a non-Hermitian tachyon model. Although non- logical properties through the quantum Hall response [11] Hermiticity is in general associated with dissipation, we or demonstrate the presence of fractional charge carriers find remarkable features such as a finite dc conductivity for the fractional quantum Hall effect via noise measure- in the absence of scatterers and the conservation of the ments [12]. optical conductivity sum rule. Recently, a newly emerging field of non-Hermitian arXiv:2104.02428v2 [cond-mat.quant-gas] 11 Apr 2021 physics [13–19] has become quite active, mostly for its General theory. We consider a general setup in which promising applications [17] and for exhibiting unique fea- the system is modeled by a time-independent non- tures, which are absent in the Hermitian realm, such Hermitian Hamiltonian H0 with non-degenerate spec- as exceptional points [20–24], non-Hermitian skin ef- trum subjected to a time-dependent perturbation V (t) = fect [25–29], non-Bloch phase transitions [30], and uni- Bf(t), which starts to act at t = 0, where B is an op- directional invisibility [31] to mention a few. By now, erator which can be non-Hermitian and f(t), the pertur- quantum optics and cold atoms setups provide fertile bation force, assumed to be small and real-valued. We grounds where such non-Hermitian concepts are probed are interested in the linear response of a system operator experimentally [21, 32–34]. In such state-of-the-art ex- δhA(t)i = hA(t)i− hAi0. The expectation value hA(t)i at 2 time t is given by bation as

tr[δρI (t)AI (t)] tr[δρ(t)] δhA(t)i = − hA(t)i0 , (3) tr[ρ(t)A] tr[ρ (t)] tr[ρ (t)] hA(t)i = , (2) 0 0 tr[ρ(t)] † iH t/ −iH0t/ where AI (t) = e 0 ~Ae ~ is written in the interac- tion representation in terms of the unperturbed system, † iH0t/ −iH t/ where ρ(t) is the of the system in the and δρI (t) = e ~δρ(t)e 0 ~. The second, norm Schr¨odingerrepresentation obeying the non-Hermitian correction term, is specific to non-unitary dynamics and † von Neumann equation i~∂tρ(t) = Hρ(t) − ρ(t)H , with is absent in the regular unitary evolution, but it is impor- H = H0 + V (t), the full Hamiltonian. Its solution tant to reproduce correctly the dynamics of the expecta- is decomposed as ρ(t) = ρ0(t) + δρ(t) with ρ0(t) = tion values [39]. Keeping in mind that AI (t) is identical † −iH0t/ iH t/ e ~ρ(0)e 0 ~, the density matrix of the unper- to the time evolution in terms of H0, turbed system, and δρ(t) arising from the external per- i.e. A(t) = AI (t), we finally obtain a compact expression turbation. Rotating to the , the varia- for the generalized susceptibility [38] defined in Eq. (1) tion δhA(t)i is evaluated [38] to first order in the pertur- as

0 i  †  ρ (t )  0 iH τ/ −iH0τ/ 0 χAB(t, t ) = − θ(τ)tr [A(τ),B]∼ − hA(t)i0[e 0 ~e ~,B]∼ · , (4) ~ tr[ρ0(t)]

in terms of a modified commutator defined as [X,Y ]∼ = system Hamiltonian in momentum space is XY − Y †X, where τ = t − t0. The first term is the X 2 natural generalization of the Hermitian Kubo formula H0 = p c σx + ∆σy − imc σz, (5) for the non-Hermitian setting. The second term arises p solely from the non-unitary dynamics, and hA(t)i0 = which has two momentum-dependent energy bands p 2 2 2 2 4 tr[ρ0(t)A]/tr[ρ0(t)]. In contrast to the Hermitian Kubo E±(p) = ±ε(p), with ε(p) = p c + ∆ − m c [see formula, the non-Hermitian counterpart is not time- Fig.1(a)], with the Fermi velocity as the effective light 0 translation invariant and depends separately on t and t speed c. The Hamiltonian is invariant with respect to a due to the appearance of the system density matrix ρ0(t) PT -symmetry operator σxK, with K performing com- at various times, which may have a non-unitary evolution plex conjugation, [σxK,H0] = 0. The gapped non- in the absence of the perturbation. Hermitian Dirac Hamiltonian at |∆| > mc2 is in a PT - symmetric phase since the eigenvectors are invariant to When the system Hamiltonian H0 is Hermitian, the the PT -symmetry operator and the spectrum consists expression (4) simplifies considerably since ρ0(t) = ρ(0) of two real energy bands [13, 42, 43]. In this regime, is a time-independent equilibrium density matrix. For P the density of states, g(ω) = p δ(ω − E±(p)) van- example, if additionally, the perturbation proves Her- ishes below the gap√ edge and features a sharp thresh- mitian, the norm corrections drop out, and assum- old singularity at ∆2 − m2c4. Exactly at |∆| = mc2, ing that the system is invariant under time transla- the spectrum becomes gapless with a linear dispersion tion, one readily recovers the regular Kubo formula [1], ε(p) = c|p|. The p = 0 point marks an exceptional i χ(τ) = − θ(τ)h[A(τ),B]i0. If the perturbation is anti- point (EP), where not only the eigenvalues, but also the ~ Hermitian B† = −B, the correlation function recov- eigenvectors coalesce [20, 22]. Due to the linear disper- ers the central result of Ref. [37], where the modi- sion, the density of states is constant for all energies. fied commutator becomes an anticommutator, χ(τ) = For |∆| < mc2, a tachyon-like spectrum with a hyper- i − θ(τ)(h{A(τ),B}i0 − 2hAi0hBi0). If H0 is non- bolic gapless dispersion develops. Depending on p, the ~ Hermitian, the norm correction term is crucial and it system is in a PT -symmetric phase with real eigenval- cannot be discarded as it may encompass the dominant ues for p2c2 > m2c4 − ∆2 or in a PT -symmetry broken contribution to linear response. phase, with conjugate pairs of imaginary eigenvalues, at p2c2 < m2c4 −∆2. These two momentum regions are sep- Tachyon model. So far, our discussion was quite gen- arated by EPs at ε(p ) = 0. Close to the EP, the spec- EP √ eral, so in the following we illustrate the theory applied trum behaves as |ε(p)| ∼ p − pEP , and consequently on a genuine non-Hermitian Hamiltonian. In particu- g(ω) ∼ ω for small energies. This is reminiscent to the lar we investigate a one-dimensional (1d) non-Hermitian case of 2d graphene [44], but in 1d such a behavior is only Dirac model which contains a tachyon phase [40, 41]. The possible for non-Hermitian systems. Close to the EP, the 3

(a) (b) 2 ] 1.0 tion causes transitions between the bands, the response

/mc 2

E 2.0 e .8 / c

] is zero in the broken PT -symmetric phase for tachyons, 2 .9 m

4 0.5 e [ /

) c 1 i.e. when the spectrum consists of pairs of imaginary 0 ( m 1.5 1.1 4 0.0 [

1.2 eigenvalues. This is due to exact cancellation between

) 0.0 0.5 1.0 2

( /mc 0 the regular commutator contribution to the response,

y 1.0

p t i [jp(τ), jp(0)], and the norm corrections. We have also v i t c

u checked this numerically by solving Eq. (5) in the pres- d

n 0.5

o ence of a weak electromagnetic field. Therefore, only C the PT -symmetric part of the model with pairs of real 0.0 eigenvalues contributes to linear response. The initial 0 1 2 3 4 [mc2/ ] conditions fix ρ0(t) = ρ(0), and Eq. (4) becomes X0 χ(p, τ) iθ(τ)X0 FIG. 1. (a) Dispersion relation for the non-Hermitian Hamil- χ(τ) = = h[j (τ), j (0)]i L L p p 0 tonian (5) with (green) parabolic spectrum of gapped Dirac p ~ p 2 Hamiltonians at |∆| > mc , (red) hyperbolic tachyon dis- † 2 iH τ/ −iH0τ/  persion for |∆| < mc , and (black) linear Dirac spectrum at −hjp(0)i0h[e 0 ~e ~, jp(0)]i0 , (6) |∆| = mc2. The dispersion is imaginary (dashed-dotted line) 2 2 2 4 2 P0 only in the tachyon phase for p c < m c − ∆ . The wig- where L is the system size and p restricts the sum to gly lines denote resonant optical transitions under the electric momenta in the PT -symmetric region p2c2 > m2c4 − 0 field application. (b) The real part of the conductivity σ (ω) ∆2. For our particular model, at half-filling, χ(τ) is as a function of frequency for several values of the ratio ∆/mc2 in the case of (green) parabolic, (black) linear, or (red) hyper- time-translation invariant, unlike the general formula in bolic (tachyon) dispersion. (Inset) Tachyon dc conductivity Eq. (4). as a function of ∆/mc2. Let us first consider the case ∆ = 0, where the sys- tem has a purely imaginary mass. Noticing that the Hamiltonian H0(∆ = 0) is self-conjugate with respect √ group velocity diverges as ∂pε(p) ∼ 1/ p − pEP . to σx, it implies [jp,H0(∆ = 0)]∼ = 0, and therefore Such a non-Hermitian Dirac-like Hamiltonian has been the current operator is time independent. Hence the first recently realized experimentally on waveguide lattices contribution to χ(τ) in Eq. (6), containing the commu- on a photonic chip [45, 46] as well as proposed in ion tator [jp(τ), jp(0)] = 0, drops out. However, the norm traps [41]. We can also give a physical meaning to Hamil- correction term remains finite and contributes to the lin- tonian (5) by envisioning it as an effective Hamiltonian ear response. Such a term is unique to non-Hermitian associated with a Lindbladian evolution [47, 48] in which systems, without any analogue in the conventional Her- the quantum jumps or decay events are absent [38]. mitian Kubo formula. For generic real gap ∆, both the regular and the norm As a perturbation, we consider a time-dependent elec- correction terms contribute, and χ(p, τ) reads tric field E that couples to the current operator and in- duces optical transitions between the two bands. The 2e2c2 (ε(p)∆ − pmc3)2 2ε(p)τ electric field effect is taken into account through a time- χ(p, τ) = θ(τ) sin . (7) (p2c2 + ∆2)2 dependent vector potential A(t) as p → p + eA(t), with ~ ~ e, the absolute value of electric charge, and E = −∂A/∂t. The result recovers correctly the Hermitian limit m → P Then the perturbation becomes V (t) = p ecσxA(t). 0 [49, 50]. The correlation function (7) captures an inter- By construction, the model is translation invariant and esting behavior for the Dirac-point dispersion |∆| = mc2: P the current operator j = −δH/δA = p jp, with jp = only half of the momenta contribute to linear response. −ecσx, is momentum independent. Indeed, for p > 0 and ∆ = mc2, χ(p > 0, τ) = 0 2 Optical conductivity. Our main task is to determine and hjp>0(t)i = ec, while for p < 0 and ∆ = −mc , 0 0 the current-current correlator χ(t, t ) ≡ χj,−j(t, t ) us- χ(p < 0, τ) = 0 and hjp<0(t)i = −ec. This is a re- ing Eq. (4) and the corresponding optical conductivity, sult of destructive quantum interference, due to the two σ(ω) = χ(ω)/iω that describes the optical response to the masses of equal amplitude, that inhibits half of the mo- external electric field. As initial conditions we assume mentum states to transition from the ground to the ex- the half filling case. In the gapped and PT -symmetric cited states [38]. This is analogous to unidirectional in- tachyon phase, the system starts in the visibility in non-Hermitian systems [17, 31]. The time with all the −ε(p) levels initially occupied, while in the dependent χ(τ) is obtained by summation over all mo- PT -broken phase the levels with a larger weight, and mentum states. We find that in the gapped phase χ(τ) corresponding to the slowest decay rate, associated with exhibits a damped oscillating behavior with a slow decay +i|ε(p)| states dominate the dynamics at t < 0 [38] are ∼ τ −1/2 which transforms into an exponential decay for assumed to be filled. the linear dispersion and in the tachyon phase [38]. The First of all, we find that while in general the perturba- asymptotic behavior of χ(τ) in all the three regimes is 4

settings [18], where additional external perturbations re- veal various aspects of the underlying model. For exam-  −2mc2τ/ 2 τe ~, |∆| = mc ple, the role of topology in non-Hermitian systems, com-  5/2 −1/2 −2|∆ |τ/ 2 χ(τ) ∼ |∆eff | τ e eff ~, |∆| < mc (8) bined with many-body physics, as well as various sources 5/2 ∆ τ −1/2 sin( 2∆eff τ + π ), |∆| > mc2, of dissipation can be probed using our theory. eff ~ 4 √ 2 2 4 with ∆eff = ∆ − m c , the effective band gap. The re- In terms of the tachyon physics, imaginary mass sponse in frequency space χ(ω) follows by Fourier trans- particle dynamics has been demonstrated recently us- forming Eq. (7) and summing over momenta. The real ing a waveguide lattice, realizing an effective non- (absorptive) part of the optical conductivity is given Hermitian Hamiltonian (5) experimentally [45, 46]. An- by the imaginary part of the susceptibility, σ0(ω) = other promising avenue involves ion trap physics, which χ00(ω)/ω. allows to simulate the 1d in a minimal The Fig.1(b) displays the evolution of σ0(ω) in all setup consisting of two atomic levels and a motional de- three regimes. For |∆| = mc2, the dispersion is linear, gree of freedom [41, 54, 55]. The momentum term in the similarly to the gapless 1d Dirac equation, but, in con- Hamiltonian results from coupling between the atomic trast to the expectations from a 1d Hamiltonian ∼ pcσx, levels and vibrational degrees of freedom. The real mass where σ0(ω) = 0 [50], now σ0(ω > 0) > 0. This is a man- is generated by a laser field with frequency proportional ifestation of Zitterbewegung, and follows from the non- to ∆, while the imaginary mass is modeled by the finite commutativity of current operator and Hamiltonian, due lifetime of the excited state, and it is proportional to to the additional real and imaginary mass terms [51, 52]. the decay rate [41]. In such setups, the density matrix of The conductivity in the gapped phase |∆| > mc2, be- the system evolves according to a Lindblad equation, but haves qualitatively similar to the one in the gapped Her- postselecting outcomes with no photon emission events mitian model [50]. There is a threshold for excitations singles out the dynamics driven by the non-Hermitian given by the band gap ~ω ' 2 ∆eff , above which a thresh- Hamiltonian in Eq. (5). Finally, the optical conductivity old singularity appears, similarly to the density of states. can be measured by in situ fluorescence spectroscopy [56]. Finally, in the tachyon phase |∆| < mc2, σ0(ω) re- mains always finite, with no anomalous behavior when Conclusions. We developed a unified linear response the system approaches the EP, although the group veloc- theory for non-Hermitian systems and perturbations, and ity ∂pε(p) diverges at the EPs. The low-frequency con- provided an expression for the Kubo formula suitable for stant optical conductivity parallels closely to graphene, non-Hermitian models. It contains (a) a generalized com- where the linear density of states produces similar be- mutator, (b) norm correction terms due to non-unitary havior. Furthering this analogy with graphene, we also dynamics, and (c) is not generally time-translation in- identify a non-zero dc conductivity σdc = σ(ω = 0) for variant. tachyons which depends on the ratio of the effective gap and the imaginary mass, As a physically motivated illustration, we applied it to investigate the optical conductivity of a one dimensional 2 e |∆eff | Dirac model, with both real and imaginary masses. This σdc = , (9) 4mc mc2 model features a tachyon phase where excitations travel faster than an effective speed of light. We find that the while σ = 0 for |∆| ≥ mc2. A finite dc conductiv- dc low-energy physics in this phase is graphene-like with ity without scatterers was only thought to be possible constant optical conductivity and finite minimal conduc- for graphene [44, 53] in 2d, and the present 1d non- tivity. For all masses, the optical sum rule is satisfied. Hermitian Dirac system represents another occurrence. We argue that these results can be tested experimentally. However, with increasing frequency, the analogy with Our Kubo formula represents an ideal starting point to graphene stops, as the EPs do not play a major role and gain information about non-Hermitian many-body sys- the optical conductivity decays with frequency as ∼ ω−3. tems as well, where the available methods are limited. Finally, we also observe that the optical sum rule [1] is satisfied for all phases of the non-Hermitian model, The authors thank A. Mostafazadeh for clarifying Z ∞ 2 discussions. This research is supported by the Na- 0 e c σ (ω)dω = . (10) tional Research, Development and Innovation Office - 0 2~ NKFIH within the Quantum Technology National Ex- It would be interesting to investigate this and other sum cellence Program (Project No. 2017-1.2.1-NKP-2017- rules in non-Hermitian, higher dimensional systems as 00001), K119442, K134437 and by the Romanian Na- well. tional Authority for Scientific Research and Innovation, Connection to experiments. Our general theory is ex- UEFISCDI, under projects no. PN-III-P4-ID-PCE-2020- pected to find applications in a variety of non-Hermitian 0277 and PN-III-P1-1.1-TE-2019-0423. 5

Rev. Lett. 124, 056802 (2020). [30] S. Longhi, Phys. Rev. Research 1, 023013 (2019). [31] Z. Lin, H. Ramezani, T. Eichelkraut, T. Kottos, H. Cao, ∗ [email protected] and D. N. Christodoulides, Phys. Rev. Lett. 106, 213901 [1] G. D. Mahan, Many-Particle Physics (3rd ed. , Kluwer, (2011). Boston, 2000). [32] R. El-Ganainy, M. Khajavikhan, D. N. Christodoulides, [2] R. Kubo, J. Phys. Soc. Jpn. 12, 570 (1957). and S. K. Ozdemir, Commun. Phys. 2, 37 (2019). [3] H. U. Baranger and A. D. Stone, Phys. Rev. B 40, 8169 [33] L. Li, C. H. Lee, S. Mu, and J. Gong, Nat Commun 11, (1989). 5491 (2020). [4] A. Cr´epieuxand P. Bruno, Phys. Rev. B 64, 014416 [34] Y. Takasu, T. Yagami, Y. Ashida, R. Hamazaki, (2001). Y. Kuno, and Y. Takahashi, Prog. Theor. Exp. Phys. [5] N. Nagaosa, J. Sinova, S. Onoda, A. H. MacDonald, and 2020, 12A110 (2020). N. P. Ong, Rev. Mod. Phys. 82, 1539 (2010). [35] W. S. Bakr, J. I. Gillen, A. Peng, S. F¨olling, and [6] D. Cohen, Phys. Rev. B 68, 155303 (2003). M. Greiner, Nature 462, 74 (2009). [7] D. Bohr, P. Schmitteckert, and P. W¨olfle, EPL 73, 246 [36] S. Kuhr, Natl. Sci. Rev. 3, 170 (2016). (2006). [37] L. Pan, X. Chen, Y. Chen, and H. Zhai, Nat. Phys. 16, [8] M. Michel, J. Gemmer, and G. Mahler, Eur. Phys. J. B 767 (2020). 42, 555 (2004). [38] D. Sticlet, B. Dora, and C. P. Moca, Supplementary Ma- [9] T. Fujii, J. Phys. Soc. Jpn. 76, 044709 (2007). terial. [10] A. Kundu, A. Dhar, and O. Narayan, J. Stat. Mech.: [39] The expectation value dynamics of an operator A ac- Theory Exp. 2009 (03), L03001. quires additional terms, due to the fact that H 6= H† [11] D. J. Thouless, M. Kohmoto, M. P. Nightingale, and and a the norm is not conserved. M. den Nijs, Phys. Rev. Lett. 49, 405 (1982). [40] G. Feinberg, Phys. Rev. 159, 1089 (1967). [12] L. Saminadayar, D. C. Glattli, Y. Jin, and B. Etienne, [41] T. E. Lee, U. Alvarez-Rodriguez, X.-H. Cheng, Phys. Rev. Lett. 79, 2526 (1997). L. Lamata, and E. Solano, Phys. Rev. A 92, 032129 [13] C. M. Bender, Rep. Progr. Phys. 70, 947 (2007). (2015). [14] I. Rotter, J. Phys. A: Math. Theor. 42, 153001 (2009). [42] C. M. Bender and S. Boettcher, Phys. Rev. Lett. 80, 5243 [15] N. Moiseyev, Non-Hermitian (Cam- (1998). bridge University Press, Cambridge New York, 2011). [43] A. Mostafazadeh, J. Math. Phys. 43, 205 (2002). [16] H. Cao and J. Wiersig, Rev. Mod. Phys. 87, 61 (2015). [44] A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. [17] R. El-Ganainy, K. G. Makris, M. Khajavikhan, Z. H. Novoselov, and A. K. Geim, Rev. Mod. Phys. 81, 109 Musslimani, S. Rotter, and D. N. Christodoulides, Nat. (2009). Phys. 14, 11 (2018). [45] W. Song, S. Gao, H. Li, C. Chen, S. Wu, S. Zhu, and [18] Y. Ashida, Z. Gong, and M. Ueda, (2020), T. Li, (2020), arXiv:2011.08496 [physics.optics]. arXiv:2006.01837 [cond-mat.mes-hall]. [46] P. Xue, L. Xiao, D. Qu, K. Wang, H.-W. Li, J.-Y. Dai, [19] E. J. Bergholtz, J. C. Budich, and F. K. Kunst, Rev. B. Dora, M. Heyl, R. Moessner, and W. Yi, (2020), Mod. Phys. 93, 015005 (2021). arXiv:2004.05928 [quant-ph]. [20] M. V. Berry, Czech. J. Phys. 54, 1039 (2004). [47] G. Lindblad, Comm. Math. Phys. 48, 119 (1976). [21] K. G. Makris, R. El-Ganainy, D. N. Christodoulides, and [48] V. Gorini, A. Kossakowski, and E. C. G. Sudarshan,J. Z. H. Musslimani, Phys. Rev. Lett. 100, 103904 (2008). Math. Phys. 17, 821 (1976). [22] W. D. Heiss, J. Phys. A: Math. Theor. 45, 444016 (2012). [49] B. D´oraand F. Simon, Sci. Rep. 5, 14844 (2015). [23] B. Zhen, C. W. Hsu, Y. Igarashi, L. Lu, I. Kaminer, [50] P. Lee, T. Rice, and P. Anderson, Solid State Commun. A. Pick, S.-L. Chua, J. D. Joannopoulos, and M. Soljaˇci´c, 14, 703 (1974). Nature 525, 354 (2015). [51] J. Cserti and G. D´avid, Phys. Rev. B 74, 172305 (2006). [24] B. D´ora,M. Heyl, and R. Moessner, Nat Commun 10, [52] B. Thaller, The Dirac equation (Springer, 2011). 2254 (2019). [53] M. I. Katsnelson, Eur. Phys. J. B 51, 157 (2006). [25] S. Yao and Z. Wang, Phys. Rev. Lett. 121, 086803 (2018). [54] L. Lamata, J. Le´on,T. Sch¨atz,and E. Solano, Phys. Rev. [26] F. K. Kunst, E. Edvardsson, J. C. Budich, and E. J. Lett. 98, 253005 (2007). Bergholtz, Phys. Rev. Lett. 121, 026808 (2018). [55] R. Gerritsma, G. Kirchmair, F. Z¨ahringer,E. Solano, [27] F. Song, S. Yao, and Z. Wang, Phys. Rev. Lett. 123, R. Blatt, and C. F. Roos, Nature 463, 68 (2010). 246801 (2019). [56] R. Anderson, F. Wang, P. Xu, V. Venu, S. Trotzky, [28] L. Li, C. H. Lee, and J. Gong, Phys. Rev. Lett. 124, F. Chevy, and J. H. Thywissen, Phys. Rev. Lett. 122, 250402 (2020). 153602 (2019). [29] D. S. Borgnia, A. J. Kruchkov, and R.-J. Slager, Phys. [57] E. M. Graefe, H. J. Korsch, and A. E. Niederle, Phys. Rev. Lett. 101, 150408 (2008). 6

Supplementary Material

Linear response theory

Let us assume a time-independent Hamiltonian H0 with non-degenerate energy eigenvalues perturbed by a time- dependent potential V (t) which starts to act at t = 0. Unless stated otherwise, H0 is considered non-Hermitian † H0 6= H0 . The time-dependent potential is V (t) = Bf(t), where operator B is a time-independent operator in the Schr¨odingerpicture, not necessarily Hermitian. The external force f(t) is considered without loss of generality as a real-valued quantity. The change in the expectation value hA(t)i of an operator A due to the perturbation is encoded 0 R 0 0 0 in the correlation function χ(t, t ), δhA(t)i = dt χAB(t, t )f(t ) Expectation values (2) are normalized by a time-dependent norm. Their dynamics is obtained using the equation † of motion for the density matrix i~∂tρ(t) = Hρ(t) − ρ(t)H , where H = H0 + V (t), which generates additional terms which occur only for non-Hermitian systems [57],

† † i~∂thAi = hAH − H Ai − hAihH − H i. (11)

In order to treat the effect of V (t), it is convenient to switch to the interaction picture, where operators evolve with H0. The regular expression relating the interaction and the Schr¨odingerpicture for the ket states remains the same, † † −iH0 t/~ but since H0 6= H0 , the Hermitian conjugate expression for the bra changes, hψI (t)| = hψ(t)|e . The evolution of an operator in the interaction representation is constructed in the regular way, AI (t) = iH†t/ −iH t/ e 0 ~Ae 0 ~. Since the evolution is not unitary, the density matrix in the interaction picture

† iH0t/ −iH t/ ρI (t) = e ~ρ(t)e 0 ~, (12) is not properly normalized, and therefore hA(t)i also reads

tr[ρ (t)A (t)] hA(t)i = I I , (13) tr[ρ(t)] where the denominator is included for proper normalization. Under the effect of the external perturbation, following the general strategy [1], we look at the changes in the density matrices in the Schr¨odingerand interaction pictures induced by the external perturbation. For that we write ρ(t) = ρ0(t) + δρ(t) and ρI (t) = ρ0,I (t) + δρI (t) where the density matrices ρ0 evolve solely due to H0, † −iH0t/ iH t/ ρ0(t) = e ~ρ(0)e 0 ~. Consequently, ρ0,I (t) is identified with the initial density matrix ρ0,I (t) = ρ(0). The linear response requires estimating δρ(I) only to first order in the perturbation. To this order, the operator expectation value (13) reads

tr[δρI (t)AI (t)] tr[δρ(t)] hAi ' hAi0 + δhAi, δhA(t)i = − hA(t)i0 . (14) tr[ρ0(t)] tr[ρ0(t)] where hA(t)i0 represents the expectation value in the absence of the perturbation hA(t)i0 = tr[ρ0(t)A]/tr[ρ0(t)] = tr[ρ(0)AI (t)]/tr[ρ0(t)]. Note that since H0 is generally non-Hermitian, the norm of the density matrix under the evolution with H0 may change in time, which requires an explicit time dependence in ρ0(t), even in the absence of V (t). To find an explicit expression for δρI , we use the equation of motion for the density matrix in interaction picture

dρI (t) iH t/ −iH†t/ † iH t/ −iH†t/ i = e 0 ~e 0 ~V (t)ρ (t) − ρ (t)V (t)e 0 ~e 0 ~, (15) ~ dt I I I I which gives the first-order correction to ρ(0),

Z t i 0 † 0 † 0 † 0 0 iH0t /~ −iH0 t /~ 0 0 iH0t /~ −iH0 t /~ δρI (t) ' − dt e e VI (t )ρ(0) − ρ(0)VI (t )e e . (16) ~ 0 The effect of the perturbation on the expectation hA(t)i is made explicit by using Eqs. (12) and (16) into Eq. (14). Then comparison with Eq. (1) yields the generalized susceptibility in Eq. (4) from the main text after the conventional † iH t/ −iH0t/ identification, that the operator time evolution in the interaction picture, AI (t) = e 0 ~Ae ~ is identical to the Heisenberg picture time evolution in terms of H0, i.e. A(t) = AI (t). 7

The rest of the section discusses some particular cases. For example, if the initial Hamiltonian H0 is Hermitian, then the system density matrix remains time-independent, ρ0(t) = ρ(0) is usually just the equilibrium density matrix, and hence hA(t)i0 = hA(0)i0. If the perturbation B is Hermitian as well, the classic Kubo formula follows immediately, i χAB(τ) = − θ(τ)h[A(τ),B]i0. (17) ~

Another case discussed in the literature corresponds to a Hermitian Hamiltonian H0 perturbed with an anti- Hermitian potential B† = −B. Then general expression (4) recovers the result derived in Ref. [37], with additional norm corrections,

i  χAB(τ) = − θ(τ) h{A(τ),B}i0 − 2hAi0hBi0 . (18) ~

If H0 is some generic non-Hermitian Hamiltonian, but the perturbation B is Hermitian, then one has to use Eq. (4), with the only simplification that the two modified commutators [... ]∼ are replaced by regular commutators. Evidently, if the perturbation B is anti-Hermitian, then [... ]∼ are replaced by anti-commutators. Finally, note that the trace in Eq. (4) must be taken over an orthonormal basis, and that eigenstates of H0 are generally not orthogonal. Other simplifications occur for a typical PT -symmetric model with real eigenspectrum, where the initial density matrix is constructed from the ground-state wavefunction H0|0i = E0|0i. In this case ρ(0) = |0ih0| remains time independent,

† −iH0t/ iH t/ ρ0(t) = e ~|0ih0|e 0 ~ = ρ(0), (19) since E0 is real. The correlation function is explicitly time-translation invariant and reduces to

i † iH τ/ −iH0τ/ χAB(τ) = − θ(τ)h[A(τ),B]∼ − hA(0)i0[e 0 ~e ~,B]∼i0. (20) ~ The first term in the correlation function bears resemblance to the commutator in the Hermitian case. However, the second term has no obvious analogue, but its presence crucial to obtain the correct linear response expression, as we discuss in the main paper in the study of tachyons.

Tachyon model and the optical conductivity

The non-Hermitian tachyon model was introduced in Eq. (5) in the main paper, but here we derive it as an effective Hamiltonian for a Dirac particle with decay [41]. A relativistic particle is described by the Hamiltonian Hp = pcσx + ∆σy where p is the momentum, and σx,y are the regular Pauli matrices, and ∆ is the energy gap. If the excited levels of Hp have a finite lifetime, modeled by a decay rate γ, the time evolution of the model is described by P the Lindblad equation for the density matrix, ρ = p ρp, with

dρp i 1 † † = − [Hp, ρp] − {L L, ρp} + LρpL , (21) dt ~ 2 √ with Lindblad operators L = γσ−, σ± = σx ± iσy. The non-Hermitian evolution is conditioned by the absence of photon emission events, i.e by neglecting the recycling (quantum jump) term ∼ LρL†. Then the system is modeled by P i~γ 2 a non-Hermitian Hamiltonian H0 = p Hp − 2 σ+σ− which describes the dynamics. Denoting mc = ~γ/4, apart from a term proportional to the identity matrix σ0 we recover the Hamiltonian introduced in Eq. (5),

X 2 H0 = p c σx + ∆σy − imc (σ0 + σz). (22) p

2 2 The extra term, δH = −imc σ0, shifts the eigenvalues in the complex plane by a constant E± = −imc ± ε, with ε = pp2c2 + ∆2 − m2c4, but does not chage the eigenvectors. For a given momentum p the eigenvectors are

1  pc − i∆  1  imc2 + ε  |p+i = √ , |p−i = √ , (23) N imc2 + ε N −pc − i∆ where |p+i and |p−i correspond to E+ and E− bands, respectively. The normalization factor depends on the phase in which the system is investigated: N = 2(ε2 + m2c4) if ε is real, and N = 2mc2(|ε|2 + mc2) if ε is imaginary. 8

(a) (b) (c)

0.92 1 ]

c 0.9998 e [

) t ( p j 0.90

t n e

r 0.9999 0 r u

C 2 2 2 2 p/mc, /mc , eA0/mc, 0/mc p/mc, /mc , eA0/mc, 0/mc 2 2 p/mc, /mc , eA0/mc, 0/mc ( 2.2, 1, 0.1, 0.8) (1.05, 0.05, 0.1, 0.8) (2, 0.1, 0.1, 1) ( 2.2, 1.05, 0.1, 0.8) (1.05, 0.05, 0.05, 0.8) 0.88 (2.1, 0.2, 0.05, 1.7) ( 2.2, 0.95, 0.1, 0.8) (1.05, 0.05, 0.03, 0.8) (1.1, 1.6, 0.1, 0.8) ( 3.2, 0.95, 0.1, 0.8) (1.05, 0.05, 0.01, 0.8) 1.0000 0 5 10 15 20 0 5 10 15 20 0 5 10 15 20 Time t [ /mc2]

FIG. 2. Comparison between the time-evolution of the current operator hjp(t)i calculated by using the numerical solution to the Schr¨odingerequation (dashed lines) and the linear response approach (solid lines). The legend shows the parameters sets at different momenta p, masses ∆, vector potential amplitudes A0, and frequencies ω0. Expectation values at (a) some generic points in parameter space, (b) the critical point ∆ = −mc2 (similar to the ∆ = mc2) and (c) the exceptional point ε(p) = 0.

The constant shift ensures that all eigenvalues have negative imaginary part, and the wave function norm has an exponential decay, in contrast to the gain-loss mechanism in the absence of the shift. Moreover, the additional identity matrix has no impact on the current-current correlations and is henceforth omitted in the Hamiltonian (5) in the main paper. By construction, the Hamiltonian (5) is invariant under a PT -type symmetry represented by the antiunitary oper- ator σxK, where K performs complex conjugation, [σxK,H0] = 0. The model is in a PT -symmetric phase when also the eigenvectors are invariant to the symmetry σxK|p±i = |p±i and consequently the eigenvalues are real. In the PT -symmetry broken phase σxK|p±i = |p∓i and the eigenvalues are imaginary. The effect of an electric field perturbation is incorporated using the minimal coupling scheme p → p + eA(t), with A(t), the associated vector potential and e, the absolute value of the electric charge. Consequently the full Hamiltonian P P reads H = H0 +V (t), with V (t) = p ecA(t)σx. Correspondingly, the current operator is j = −δH/δA = − p ecσx. The optical conductivity gives the current response to the external perturbation V (t) and requires the calculation of the current-current correlation function. In order to compare the analytical results with the numerical data we consider the case of a monochromatic perturbation, with an oscillating behavior for the vector potential of the form A(t) = A0 sin(ω0t). Fig.2 shows a comparison between the linear response results and the numerical ones. The latter are obtained by integrating the time-dependent Schr¨odingerequation and computing the averaged current hjp(t)i for different sets of parameters, as indicated in the legend. The linear response theory assumes that the perturbation is much smaller compared to the energy separation between the eigenvalues, and within this approximation it reproduces almost perfectly the numerical results in Fig.2(a) and (b). The approximation only starts to fail very close to the exceptional point, but is improved by decreasing the perturbation amplitude [see Fig.2(c)]. For the Dirac-point dispersion, when ∆ = mc2, the correlator χ(p, τ) vanishes for positive momenta, while at 2 ∆ = −mc , for negative momenta. In these cases the current expectations value is determined entirely by hjp(0)i0 = ec, 2 2 for ∆ = mc , and hjp(0)i0 = −ec, for ∆ = −mc . The latter case is illustrated both in the numerical and analytical calculation in Fig.2(b) (blue line). To understand this effect it is relevant to study the transition amplitudes between 2 the current eigenstates for the√ Hamiltonian at ∆ = mc , H0(p) = pcσx +∆σy −i∆σz. Under a unitary transformation U = exp(iπ(σx + σy + σz)/3 3) the Hamiltonian reads

 pc 0  UH (p)U † = . (24) 0 2∆ −pc

For p > 0, |p−i = (0, 1)T belongs to the ground state and |p+i to the excited spectrum. The system cannot be excited, but can only decay with a decay rate ∼ 2∆. Then in the T = 0 limit, when all |p−i states are filled, the system does not respond to the electric field at p > 0 and the current remains independent in time, i.e. hjp>0(t)i = −echp−|σz|p−i = ec. At p < 0, the lowest energy state is |p−i = (1, 0)T , and now the system can be excited with a 2∆ rate. In conclusion, 9

0.6 /mc2 0.5 0.75 0.4 1

] 1.25 3

c 1.5 m

2 0.2 e / 2 [

)

( 0.0

0.2

0 5 10 15 20 Time [ /mc2]

FIG. 3. Numerical results for current-current correlations as a function of time match the asymptotic behavior from Eq. (25). There is an exponential decay for the gapless spectrum |∆| ≤ mc2 and a damped oscillatory behavior for |∆| > mc2. only a half of all momentum states in the system respond to the electric field. Similar considerations work for 2 ∆ = −mc , where H0 = pcσx + ∆σy + i∆σz, and now hjp<0(t)i = −ec, with transitions at p < 0 inhibited. R The correlation function integrated over momenta gives the time-dependent susceptibility, χ(τ) = dpχ(p, τ)/2π~. An asymptotic analysis, using the saddle-point approximation for the integrals, determines the time-dependence of correlations presented in the main text. The leading terms, including the overall constants are

2 2  8e c 2 4 −2mc τ/~ 2 − 3 m c τe , |∆| = mc  3π~ √  2 |∆ |5/2 2 4 2 √e c eff −2 m c −∆ τ/~ 2 χ(τ) ∼ − 3 ∆2 e , |∆| < mc (25) π~ τ  2 ∆5/2  √e c eff 2 3 ∆2 sin(2∆eff τ/~ + π/4), |∆| > mc , π~ τ √ 2 2 4 with ∆eff = ∆ − m c the effective band gap. Fig.3 displays the numerical results for χ(τ) in all these phases, and confirms the above asymptotic results. A Fourier transform of Eq. (7) readily gives χ(ω, p) and integration over R momentum yields the total correlation function χ(ω) = dpχ(ω, p)/2π~. The absorptive part of the spectrum is related to the imaginary part of the current-current correlation function, σ0(ω) = χ00(ω)/ω. The dc conductivity is 0 00 00 σdc ≡ σ(0) = σ (0), since σ (−ω) = −σ (ω). In the gapped phase, |∆| > mc2, the dc conductivity is trivially zero and χ00(ω) reads

e2c ω∆¯ 2 ω¯2(∆2 + m2c4) − m2c4 ω χ00(ω) = √ eff θ(|ω¯| − 1), ω¯ = ~ . (26) 2 2 2 2 4 2 2~ ω¯ − 1 (¯ω ∆eff + m c ) 2∆eff For the linear dispersion |∆| = mc2,

e2c ω¯2 ω e2 |ω| χ00(ω) = sgn(¯ω), ω¯ = ~ , lim σ0(ω) = lim ~ , (27) ~ (1 +ω ¯2)2 2mc2 ω→0 ω→0 4m2c3 and the dc conductivity vanishes. In the tachyon phase |∆| < mc2, χ00(ω) and the finite dc conductivity are

e2c ω¯|∆ |2 ω¯2(∆2 + m2c4) + m2c4 ω e2 |∆ | 00 √ eff ~ eff χ (ω) = 2 2 2 4 2 , ω¯ = , σdc = 2 . (28) 2~ ω¯2 + 1 (¯ω |∆eff | + m c ) 2|∆eff | 4mc mc

R ∞ 0 The following optical sum rule indicates the conservation of the total spectral weight for conductivity 0 dωσ (ω) = e2c/2~ in all the phases of the model.