<<

Non-equilibrium field dynamics starting from arbitrary initial conditions: Absence of thermalization in one dimensional coupled to fermions

Md Mursalin Islam1, ∗ and Rajdeep Sensarma1 1Department of , Tata Institute of Fundamental Research, Mumbai 400005, India. (Dated: August 11, 2021) We propose a new method to study non-equilibrium dynamics of scalar fields starting from non-Gaussian initial conditions using Keldysh field theory. We use it to study dynamics of phonons coupled to bosonic and fermionic baths, starting from initial Fock states. We find that in one dimension long wavelength phonons coupled to fermionic baths do not thermalize both at low and high bath temperatures. At low temperature, constraints from energy-momentum conservation lead to a narrow bandwidth of particle-hole excitations and the phonons effectively do not feel the effects of this bath. On the other hand, the strong band edge divergence of particle-hole density of states leads to an undamped “polariton-like” mode of the dressed phonons above the band edge of the particle-hole excitations. These undamped modes contribute to the lack of thermalization of long wavelength phonons at high temperatures. In higher dimensions, these constraints and the divergence of density of states are weakened and leads to thermalization at all wavelengths.

I. INTRODUCTION this problem for obeying Schrodinger¨ equations. They found that arbitrary initial conditions can be incorporated in the field theory provided an additional source is coupled to Scalar fields are paradigmatic degrees of freedom in quan- the fields at the initial time. The correlators are calculated tum field theories, with wide ranging applications from par- within this theory with the additional source. The physical ticle physics to astro-particle physics and cosmology to con- correlation functions are then obtained by taking derivatives densed matter physics. In particle physics it played a big role of these correlators with respect to the initial sources, with the in the construction of the which leads to set of derivatives determined by the initial conditions. In this descriptions of massive particles1–3. In cosmology and astro- paper, we extend this formalism to describe non-equilibrium particle physics, scalar fields have been used to describe in- dynamics of scalar fields starting from arbitrary initial condi- flationary expansion of the early universe4,5; they have also tions. Our formalism also requires additional sources at initial been projected as possible candidates for dark matter in vari- times. However, unlike the case of Schrodinger¨ bosons, these ous models6–8. They are also ubiquitous in condensed matter sources couple to bilinears of both the initial fields and their physics, often rearing their heads as low energy fluctuations time derivatives, reflecting the fact that the classical equations around ordered ( broken) states of matter. From for these fields are now 2nd order in time derivatives and re- phonons in solids9 to magnons in certain types of magnets10 quire knowledge of both initial configurations and their time to phase fluctuations in a superfluid11,12, there are a large va- derivatives to be solved. The rest of the formalism is simi- riety of scalar fields which occur in descriptions of quantum lar to Ref. 51; i.e. we calculate the correlators in presence many body systems. More recently, with the advent of tech- of these sources and then take derivatives with respect to the nology to create strong light-matter interactions, photons in initial sources to get physical correlation functions. optical cavities13–16 or dressed excitations like polaritons17–21 can also be described by effective scalar fields. We use our formalism to study the thermalization of a sys- Compared to the well developed equilibrium and ground tem of scalar fields coupled to external baths, and initialized to state theory for scalar fields (including interacting field the- athermal Fock states. Although we consider phonons as a con- ories)22–24, the non-equilibrium dynamics of scalar fields are crete example in this case, similar considerations can be used relatively less studied. While the primary motivation for study for other scalar fields including complex and O(N) scalar of non-equilibrium dynamics of scalar fields were earlier re- fields. We first consider coupling the phonons to a Marko- 52 lated to studies of non-equilibrium processes in the early uni- vian Ohmic bath with a smooth ultra-violet cut-off . Here, verse and in high energy collisions25–28, more recent advances as expected, the phonons thermalize at a rate which becomes in creating non-equilibrium states, either by pumping solid momentum independent as we go to the broad bandwidth lim- 53 state systems with intense energy sources29–32, or by pulse its of pure white noise .

arXiv:2108.04264v1 [cond-mat.mes-hall] 9 Aug 2021 33–37 38–41 shaping and cavity engineering , has led to a renewed We then consider the coupling of phonons with the particle- focus on this topic. In many of these cases, the system evolves hole excitations of a non-interacting fermionic bath. We note starting from an athermal state. A typical example is a cav- that the dynamics of electrons coupled to phonons has been 39,42–45 ity where the photons are initiated to a squeezed state . extensively studied in the context of thermalization of “hot Thus it is very important to obtain a description of dynamics electrons”54–62. Relatively less attention has been paid to the of scalar fields starting from athermal initial conditions. dynamics of the phonons coupled to fermions 63–67. We note The standard Schwinger-Keldysh (SK) field theory used that the evolution of the one-particle distribution functions to describe non-equilibrium dynamics of many body sys- have been treated before using Kadanoff-Baym equations27. tems, unfortunately requires a thermal initial condition to In one dimension, at low temperatures, the density of particle- work24,46–50. Recently Chakraborty et. al.51 had looked at hole excitations with a fixed momentum transfer is finite over 2 a narrow region of energy in the low momentum limit. In II. DYNAMICS WITH INITIAL CONDITIONS the long wavelength limit this almost reduces to a linearly dis- persing mode with a width which scales quadratically with the Several problems in non-equilibrium dynamics relevant to momentum. As a result the long wavelength modes different branches of physics require the description of dy- effectively do not see the fermionic bath as it is not possi- namics of scalar fields starting from athermal initial condi- ble to satisfy both energy and momentum conservation during tions. Such initial conditions can result from sophisticated exchange with the bath. Hence long wavelength modes un- pulse-shaping techniques creating squeezed lights in optical dergo quantum oscillations and remember their initial occu- cavities39,44,45. They can also be created when large amount pation numbers even at very long times. These modes thus do of energy is dumped into a system with broken symmetry as not thermalize at low bath temperature. This lack of thermal- is done in the case of pump-probe experiments29–32,70–72. This ization should show up in generic one-dimensional systems. can also be relevant for high energy particles which can be At high temperatures, the particle-hole density of states ex- created in non-equilibrium distributions due to particle colli- 73 tend from zero energy to a band maximum (due to a finite sion/decay . bandwidth of the fermions). An interesting feature of one di- The standard textbook formalism of Schwinger-Keldysh mensional fermions is a singularity in the particle-hole den- field theory used to describe non-equilibrium dynamics how- sity of states at the upper edge of the corresponding band. ever works when the initial condition is described by a thermal 24,48–50 This leads to a strong divergence in the particle-hole density initial density matrix . In a recent paper, Chakraborty 51 of states near their band edge. The divergence in the bath den- et. al had constructed a formalism to treat open system dy- sity of states result in the formation of new dressed modes namics of Bosons and Fermions starting from arbitrary ini- of the phonons just outside the upper band edge for particle- tial conditions. The Bosons/Fermions were described by a hole excitations. This new mode, which we call a ‘polarinon’ Schrodinger¨ equation, which was suitably modified for open due to its similarity to ‘polariton’s formed by dressing of light system dynamics. Here we will extend this formalism to the with excitons68,69, remains undamped leading to absence of case of scalar fields whose equation of motion has 2nd order thermalization of phonons at long wavelengths, even at high time-derivative which brings additional complications. Be- bath temperatures. Due to matrix element effects, the spectral fore we describe the formalism for scalars, we first review the weight of this mode increases with the bath temperature, and structure of the formalism put forward in Ref. 51. It will serve hence the deviation from thermalization increases with tem- as a paradigm for comparing and contrasting the formalism perature. This new mode can be studied in 1-d systems with developed here. strong electron-phonon coupling. In Ref. 51 it was proposed that the athermal initial density matrix can be exponentiated by using a source term which To contrast with the singular case of one dimensional allows it to be included in the . In this formalism the particle-hole bath, we also consider a bath of two dimensional SK action is modified by adding a source dependent action non-interacting electrons. In this case there is no lower bound δS(u) where the source u couples to the bilinears of the fields for existence of particle-hole excitations and the phonons ther- only at the initial time. The source only couple to the Keldysh malize unless the Debye frequency of the phonons is larger (q − q) part of the action which carries information about the than the bandwidth of the electrons. Further the singularities initial state. For example, in the case of a single bosonic mode in the density of states of the particle-hole bath are weakened 1+u ∗ δS(u) = i φq (0)φq(0), when the system starts with an ar- and hence there is no additional mode formation in this case. 1−u P bitrary initial density matrix ρ0 = n cn|nihn| in the Fock The scalar fields thermalize unless they are massive, which basis. The correlation functions are calculated from the ac- describes optical phonons in the present context. In that case tion with this extra source dependent part, which makes them coupling to the Fermionic bath fails to thermalize the long dependent on the source u in general. The physical correla- wavelength phonons as energy-momentum constraints cannot tion functions Gρ are then calculated by taking a particular set be satisfied simultaneously. of derivatives on the source dependent correlation functions We now provide a brief route maps of the different sections G(u). This set of derivatives depends on the initial condition of our paper. In sectionII, we set up the Keldysh field the- the system starts with. For the simple case of a single bosonic P cn n 1 ory based formalism to describe non-equilibrium dynamics of mode, Gρ = n n! ∂u 1−u G(u)|u=0. We will now extend scalar field starting from Fock states. We initially work out the Schwinger-Keldysh field theory for scalar fields to include the case of single harmonic oscillator in sectionIIA and ex- athermal initial conditions. We will first consider the case of a tend it to the case of scalar fields in sectionIIB. In section III single Harmonic oscillator in position basis which represents we first set up the problem of dynamics of phonons coupled a single mode of scalar fields. Then we will generalize it for to external baths, where the phonons are initialized to a non- the scalar fields containing multiple modes. thermal Fock state. In section III A, we focus on the effects of a Bosonic Ohmic bath, while section III B 1 illustrates the example of a one dimensional Fermionic bath, where the long A. The Case of a single Harmonic Oscillator wavelength phonons do not thermalize. We finish this section with the case of two dimensional Fermionic bath in section It is instructive to first consider the case of a single har- III B 2. Finally we summarized the results in the concluding monic oscillator in the position basis. This will provide us sectionIV. with the basic structure of the theory, which can then be ex- 3 tended to real scalar fields. We consider the dynamics of a non-negative integer eigenvalues. Unless ω = Ω, these will closed system of one dimensional single harmonic oscillator not be the stationary states of our harmonic oscillator; never- with frequency Ω, which is governed by the action theless they can provide a complete basis to write the initial density matrix. This will allow us to consider density matri- Z m 1 S = dt x˙ 2 − mΩ2x2. (1) ces which do not commute with the Hamiltonian of the sys- 2 2 tem. In a closed system, this allows us to consider quench problems with non-trivial dynamics74–76. We will consider an We want to consider the dynamics of this system starting from initial density matrix which is diagonal in the number basis the number or Fock states. However, in a harmonic oscilla- corresponding to aω operators, i.e. tor (and similarly in scalar field theory), construction of the X Fock states involves an additional energy scale, in contrast ρˆ = c |n ihn |, (2) to Schrodinger¨ theories, where one can define a Fock basis 0 n ω ω n independent of any external scales. To see this, note that † p mω creation-annihilation operators defined by a = [ xˆ − where P c = 1 for probability conservation. Shifting √ ω√ ~ n n pˆ p mω pˆ i √ ]/ 2 and aω = [ xˆ + i √ ]/ 2 satisfies the now to the Keldysh theory with doubled time contours (+ for ~mω ~ ~mω † forward propagation and − for backward propagation), the standard bosonic algebra [aω, aω] = 1 for any value of ω, and can be used to construct a complete number basis with Keldysh partition function can be written as

Z R m 2 2 1 2 2 2 i dt (x˙+ −x˙− )− mΩ (x −x ) Z = D[x+]D[x−]e 2 2 + − hx+(0)|ρˆ0|x−(0)i. (3)

1 1 2 √ mω  4 1 − mωx Using the Harmonic oscillator energy eigenfunctions, ψ (x) = hx|n i = √ e 2 H ( mωx), where H n ω π 2nn! n n are the Hermite polynomials (we have set ~ = 1), the matrix element of the initial density matrix is given by r mω − 1 mω(x2 +x2 ) X cn √ √ hx |ρˆ |x i = e 2 + − H ( mωx )H ( mωx ) + 0 − π 2nn! n + n − n r n mω X cn ∂ 1 mωf(u)x x − 1 mωg(u)(x2 +x2 ) = √ e + − 2 + − , (4) π n! ∂un 2 n 1 − u u=0

2u 1+u2 77 where f(u) = 1−u2 , g(u) = 1−u2 , and we have used the Mehler formula for the generating function of products of Hermite polynomials to obtain the last expression. One can now write the partition function describing the dynamics starting from this initial condition as

Z = L˜(∂u, ρ0)Z(u) , (5) u=0 where the differential operator L˜ = P cn (∂ )n √ 1 depends on the initial density matrix. It is useful to write L˜ = LM(u), n n! u 1−u2 where L = P cn (∂ )n and M(u) = √ 1 . Here, the partition function in presence of source terms at the initial time is n n! u 1−u2 p R iS(u) Z(u) = mω/π D[x+]D[x−]e , where

Z Z  2 2 0 0   0  m 0 ∂t + Ω + δ(t)δ(t )[∂t0 − iωg(u)] iωf(u)δ(t)δ(t ) x+(t ) S(u) = − dt dt (x+(t), x−(t)) 0 2 2 0 0 . 2 iωf(u)δ(t)δ(t ) −∂t − Ω − δ(t)δ(t )[∂t0 + iωg(u)] x−(t ) (6)

Note that in addition to the source terms related to impos- by adding sources coupling linearly to x at arbitrary times and ing the correct initial condition on the dynamics, there is taking derivatives with respect to these sources. The deriva- an additional time derivative term at t = 0. This bound- tives with respect to these linear sources commute with the ary term comes from an integration by parts, which converts derivatives with respect to the the bilinear initial source u. x˙ 2 → xx¨ in the action and is ignored in theories where the Hence one can write time co-ordinate extends up-to −∞ . As a simple check, one can easily show that without any additional sources present, D(ρ0) = L(∂u, ρ0)N (u)D(u)|u=0 (7) q 1+u P Z(u) = , and hence Z = cn = 1, as one expects 1−u n where D(ρ0) is the correlator corresponding to the physical for the Keldysh partition function. dynamics with the initial condition, while D(u) is the cor- The correlation functions in this theory can be calculated 4 relation function in the theory with the initial sources. The single particle Green’s function, we can invert the matrix in 1 normalization N (u) = M(u)Z(u) = 1−u . Focussing on the Eq. (6) to write

1 i 1 + u  ω2  D+−(t, t0, u) = sin Ω(t − t0) − cos Ωt cos Ωt0 + sin Ωt sin Ωt0 2mΩ 2mω 1 − u Ω2 −1 i 1 + u  ω2  D−+(t, t0, u) = sin Ω(t − t0) − cos Ωt cos Ωt0 + sin Ωt sin Ωt0 2mΩ 2mω 1 − u Ω2 D++(t, t0, u) = Θ(t − t0)D−+(t, t0, u) + Θ(t0 − t)D+−(t, t0, u) and D−−(t, t0, u) = Θ(t − t0)D+−(t, t0, u) + Θ(t0 − t)D−+(t, t0, u). (8)

The details of this calculation is presented in AppendixA.

It is useful to work with the Keldysh rotated classical xcl = 0 1 1 R 0 0 sin Ω(t − t ) 2 (x+ + x−) and the quantum xq = 2 (x+ − x−) degrees of D (t, t , u) = −Θ(t − t ) and (10) freedom. In this basis, the one particle Green’s functions take 2mΩ i 1 + u  ω2  the form DK (t, t0, u) = − cos Ωt cos Ωt0 + sin Ωt sin Ωt0 2mω 1 − u Ω2  DK DR  Dˆ = . (9) 1 + u  1 ←−−→ DA 0 R A 0 = −2imω D (t, ) 1 + 2 ∂ ∂ D (, t ) . 1 − u ω =0 with

We find that the retarded (and advanced) Green’s functions are independent of the initial conditions/ initial sources, while the Keldysh Green’s function depends on them. It is then easy to show that the retarded Green’s function for the actual dynamics starting from the initial condition has the same form as in Eq. (10), while the physical Keldysh Green’s function is given by

X  1 ←−−→ DK (t, t0, ρ ) = −2imω c (1 + 2n)DR(t, ) 1 + ∂ ∂ DA(, t0) . (11) 0 n ω2   n =0 One can now write the action in the Keldysh rotated basis including the source terms, which will reproduce the correlators

Z Z 0 2(∂2 + Ω2) !  0  m 0 t xcl(t ) S = − dt dt (xcl(t), xq(t)) 2 2 1+u h 1 ←− 0 −→i 0 . (12) 2 2(∂t + Ω ) −2iω 1−u 1 + ω2 ∂t δ(t)δ(t )∂t0 xq(t )

We note that the structure of the action is similar to that ob- herent state fields in terms of position and momentum degrees tained for a Schrodinger¨ theory of Bosons, i.e. we add a bi- of freedom in a path integral on the phase space. Integrating linear source u which couples only to the quantum fields at out the momentum degrees of freedom, one can arrive at the the initial time. This anti-Hermitian term effectively acts as effective action derived above. In this case, it is easy to see a Keldysh self energy for the fields. However, there is one that the initial sources couple to both position and momen- key difference: unlike the Schrodinger¨ bosons, where the u tum, and hence, on integrating out the momentum degrees of terms only coupled to bilinears of the fields at t = 0, here freedom, they couple to the time derivative of the position at the additional terms couple both to the position and its time- initial times. We note that this derivation works when ω and derivatives at the initial time. This is a reflection of the fact Ω are same, while the detailed derivation provided in this pa- that the saddle point equation is second order in time in this per also works even when the Fock states of the initial density case and both the initial position and velocity of the particle matrix are not the eigenstates of the closed system dynamics. are required to solve this equation. Before we move to the case of the scalar fields, we need to There is an alternate way of deriving the effective action discuss how the dynamics changes if the system is coupled to Eq. (12). One can start with the coherent state representa- an external bath starting at t = 0, leading to open quantum tion of the harmonic oscillator and using the theory derived in system dynamics. The addition of a bath ( at least the sim- Ref. 51, derive the effective action in the Keldysh rotated basis plest baths where the position of the oscillator couples to the of complex Schrodinger¨ bosons. One can then write the co- bath degrees of freedom), leads to a quadratic theory where 5 the effect of the bath can be incorporated through a retarded The Keldysh self energy simply adds to the initial self-energy self-energy ΣR and a Keldysh self-energy ΣK , corresponding from the bilinear sources, so that the correlators in presence respectively to dissipative and stochastic effects of the bath. of the external bath are given by

Z t Z t1 R 0 R 0 R R R 0 D (t − t ) = D0 (t − t ) + dt1 dt2 D0 (t − t1)Σ (t1 − t2)D (t2 − t ) (13) t0 t0 0 X  1 ←−−→ Z t Z t DK (t, t0, ρ ) = −2imω c (1 + 2n)DR(t, ) 1 + ∂ ∂ DA(, t0) + dt dt DR(t − t )ΣK (t , t )DA(t − t0). 0 n 2   1 2 1 1 2 2 ω 0 0 n =0

Finally, we note that one can introduce effects of interaction B. Scalar Fields by adding an interaction term in the Keldysh action. In this case, the dynamics is not exactly solvable, and one needs to We would now like to extend our formalism to the dynam- make approximations to treat effects of interaction on the open ics of scalar fields. In this paper, we will discuss the case of system dynamics. It is important to note that even in this case, real scalar fields for the sake of brevity, but the formalism can one can use standard diagrammatic techniques for calculating be extended to complex scalar fields, or to O(N) fields in a D(u), and then obtain D(ρ0) by taking appropriate deriva- straightforward way. We consider real scalar fields φ(x, t), tives. whose dynamics is governed by the action

Z Z  2 2 2 2    d 0 −∂t + c ∇ − m φcl(x, t) S = dt d x (φcl(x, t) , φq(x, t)) 2 2 2 2 . (14) −∂t + c ∇ − m 0 φq(x, t)

Working with momenta instead of real space, this gives

Z Z  2 2    d 0 −∂t − Ωk φcl(k, t) S = dt d k (φcl(−k, t) , φq(−k, t)) 2 2 , (15) −∂t − Ωk 0 φq(k, t) √ 2 2 2 where Ωk = c k + m is the dispersion of a generic massive scalar field. One can of course set m = 0 to describe massless scalar fields. We note that since the theory would be written in terms of Ωk, one can also put the system on a lattice and consider lattice dispersions with associated finite range of lattice momenta (Brillouin zones).

We assume that the system is initialized to a density matrix which generates the dynamics of the system. which is diagonal in the Fock basis of the number operators q In this case, we can generalize the answers we obtained for † p ωk 1 ˙ nˆk = a ak, where ak = φ(k) + φ(−k). Here k 2 2ωk the simple harmonic oscillator in a straightforward way. The ωk is a dispersion which can be different from Ωk, and hence source u is generalized to a source field uk for each mode k the initial state may not be an eigenstate of the Hamiltonian and the source dependent action in the Keldysh basis is

Z Z 0 −∂2 − Ω2 !   d h t ←− k −→i φcl(k, t) S = dt d k (φcl(−k, t) , φq(−k, t)) 2 2 1+uk 1 0 . (16) −∂t − Ωk iωk 1 + 2 ∂t δ(t)δ(t )∂t0 φq(k, t) 1−uk ωk

We consider a system starting from ρ0 = this dynamics is given by P {n} c{n}|{n}ih{n}|. The physical correlation functions for 6

Z t Z t1 R 0 R 0 R R R 0 D (k; t − t ) = D0 (k; t − t ) + dt1 dt2 D0 (k; t − t1)Σ (k; t1 − t2)D (k; t2 − t ) (17) t0 t0

  K 0 X R 1 ←−−→ A 0 D (k; t, t , ρ0) = −2iωk c{n}(1 + 2nk)D (k; t, ) 1 + ∂ ∂ D (k; , t ) ω2 {n} k =0 Z t Z t0 R K A 0 + dt1 dt2 D (k; t − t1)Σ (k; t1, t2)D (k; t2 − t ), 0 0

R 0 K 0 where Σ (k; t − t ) and Σ (k; t, t ) are retarded self-energy the Brillouin zone extends from kx = −π/a to kx = π/a and and Keldsyh self-energy respectively for each mode. The self ky = −π/a to ky = π/a. energies can arise if the system is coupled to an external bath We study the dynamics of these phonon modes starting or from the interaction between the modes of the scalar fields. from an athermal initial condition. Since the coupling to the We will now use this formalism to study the non-equilibrium bath will anyway generate non-trivial dynamics, we consider dynamics of scalar fields coupled to external baths. While the initial state of the system to be a Fock state correspond- this formalism is valid for any scalar fields, we will look into ing to the bare phonon dispersion Ω(k) i.e. we consider the q q the concrete example of phonons coupled to various kinds of Ωk 1 ˙ annihilation operator ak = φ(k) + φ(k) and the external baths to illustrate the use of this formalism. 2 2Ωk † occupation number states corresponding to nˆk = akak. In one dimension we consider a lattice of even number of sites III. DYNAMICS OF PHONONS COUPLED TO EXTERNAL 2L with momenta k = ±πi/L with integer i. The initial BATH Fock state then correspond to nk = 1 for even values of i and 0 for odd values of i. Similarly in two dimensions We now consider the dynamics of a system of phonons we consider a square lattice of (2L)2 sites with momenta (quantized lattice vibrations) initialized to a non-thermal state (kx, ky) = (±πi/L, ±πj/L) with integer i and j. The ini- and coupled to an external bath. In one dimension, the bare tial Fock state then corresponds to nk = 1 for even values of i dispersion of the phonons is given by and nk = 0 for odd values of i, so we have strips of occupation numbers 0 and 1 along y-axis. We note that these states are q quite different from a thermal state of the bosons. The thermal Ω = ω2 sin2(ka/2) + m2, (18) k 0 state has a large occupancy near k = (0, 0) which decreases monotonically as k is increased. We will now consider differ- where a is the lattice spacing and m is the mass, which can ent kinds of baths and study their effects on the dynamics of be set to 0 to describe longitudinal phonons. For the mass- the phonons. less phonons, the long wavelength modes have a linear dis- ω0a persion Ωk ∼ csk, with the sound velocity cs = 2 . Since we are working on a lattice, the Brillouin zone extends from k = −π/a to k = π/a. The bandwidth of the massless A. Phonons Coupled to an Ohmic Bath phonons is then given by ω0, which controls both the ul- traviolet and the infrared dispersions in this simple model. We first consider phonons in one dimension coupled to a We also note that this kind of dispersion is not specific to bosonic bath. The effect of the bath on the phonons is con- phonons, there exist other systems (for example magnons or trolled by the bath spectral function J(ω). Here we consider quantized spin waves), which can be described by similar the case where all the phonon modes see the same Ohmic 2 − ω scalar field dispersion. Similarly the massive fields, which can bath with J(ω) = ωe σ2 . This bath spectral function is be used to represent optical phonons, have a low energy dis- linear in energy for low ω(<< σ) which is characteristic of p 2 2 2 k2 52,78 persion Ωk ∼ csk + m ∼ m + 2m∗ , where the curvature the Ohmic bath . The parameter σ ensures that the bath is ∗ 2 m ∼ m/cs. well behaved in the ultra-violet limit with the spectral function In two dimensions, we consider phonons on a square lattice smoothly decaying to zero at large energy. So σ plays the role of lattice spacing a, with the bare dispersion given by of an effective bandwidth for the bath without producing non- analyticities which lead to non-Markovian dynamics79. The q 2 2 2 2 bath spectral function J(ω) is plotted as a function of energy Ωk = ω0[sin (kxa/2) + sin (kya/2)] + m . (19) for five different values of σ ranging from σ = ω0 to σ = 5ω0 Once again, for the massless scalar fields, the long wavelength in Fig.1(a). It is clear that the bath spectral function deviates from its linear behaviour at progressively lower energies as σ modes have a linear dispersion Ωk ∼ cs|k|, with speed cs = ω0a , while for the massive fields, the low energy dispersion is decreased. We note that we have absorbed constants into 2 q the system bath coupling here, so that both κ2 and J(k, ω) 2 2 2 k2 ∗ Ωk ∼ csk + m ∼ m + 2m∗ , where m has the same have dimensions of energy in this problem. form as in one dimensional case. We note for a square lattice, In this case the momentum independent retarded self en- 7

(a) (b) (c) | ⟩ ) t , 2 ⟩

) σ ω σ ω = 2 k = 5 0 t 0 , − ( k σ ω l = 4 0 c

− π a

0.15 / ϕ ( ) l

σ ω t = 3 0 c 0 , ϕ k ) ω ( t l / c , ) k ϕ ω ( ⟨ l

( π a δ

c 0.25 / J ) ϕ a ⟨ σ = 2ω /

0 ) 0 a ω

/ π a

0.5 / ( 0 σ = ω0 | e ω

( k π a π a = 0.25 /

/ log 0 0 2 ω t ω t ω/ω0 0 0 0.1 (d) (e) (f) | ⟩

) 0.03 t , k 0.1 − ( l c κ2 = 0.2ω ϕ 0 ) t , 0 0 k ( ω ω l / / c 2 k k κ ω ϕ γ σ = 5ω γ = 0.098 0 ⟨ 0 δ ) σ = 4ω0 a / σ ω 0 0.01 = 3 0 ω ( σ = 2ω 2 | 0 κ = 0.05ω0 e k π a 2 = 0.5 / σ = ω0 κ = 0.05ω σ ω 0 0.0 = 5 0 log

ω0t ka ka

2 − ω FIG. 1. Non-equilibrium dynamics of phonons coupled to an Ohmic bath: (a) The bath spectral function J(ω) = ωe σ2 as a function of ω for different values of σ. Here ω0 is the bandwidth of bare phonons. J(ω) deviates from linear behaviour at progressively smaller energies as σ is decreased. (b) Thermalizing dynamics of correlation function iDK (k; t, t) for four different momentum modes at a low temperature 2 T = 0.02ω0 with system-bath coupling strength κ = 0.05ω0. Here σ = 2ω0. The time dependent correlation functions (solid lines) relax to their thermal values (dashed lines) at long times. (c) and (d) Difference of the time dependent correlation function and its thermal value as a function of time in a log plot for two momentum modes. (c) k = 0.25π/a, whose initial occupation is 0 and (d) k = 0.5π/a, whose initial occupation is 1. The decay rate is extracted from the slope of the envelope in the log plot. (e) Decay rate (γk) vs momentum (k) for different 2 values of the effective bath bandwidth scales (σ). γk approaches the constant value κ /2 independent of k as σ/ω0 is increased and the bath approaches the limit of pure white noise. (f) Decay rate (γk) vs momentum (k) for different values of system bath coupling. σ is fixed at the 2 highest value 5ω0 we considered. We see that γk goes very close to the constant value κ /2. Note that with increasing system bath coupling γk starts to have a weak k dependence. ergy is given by which follows from fluctuation-dissipation theorem49. Here T is the temperature of the bath. Z ∞   R 2 dz 1 1 If the effective bandwidth of the bath σ is much larger than Σ (ω) = −κ J(z) + + + . −∞ 2π z − ω − i0 z + ω + i0 the phonon bandwidth ω0, then all the phonon modes effec- (20) tively see a Gaussian white noise and one would expect the Note that Im ΣR(ω) = −κ2J(ω) and Re ΣR(ω) = modes to thermalize at the same rate γ = κ2/2. On the other 2 2 x 2 √κ  ω  −x R y hand, if σ is less than or comparable to ω then the relaxation − π σ − 2ωFD σ , where FD(x) = e 0 dy e is 0 Dawson’s function80. of different phonon modes depends on their energies with the slowest relaxation rate for the highest phonon energy. To see The Keldysh self energy is this, we track the dynamics of the phonons coupled to this  ω  Ohmic bath (T = 0.2ω0 and σ = 2ω0) with a system-bath ΣK (ω) = −iκ22 coth J(ω), (21) 2 2T coupling strength κ = 0.05ω0. In Fig.1(b), we plot the time 8 evolution of the correlation function hφcl(k, t)φcl(−k, t)i for where κ is the system bath coupling. Here φ(q) has the dimen- four different values of momentum k. The correlation func- p sion√ of [ volume/energy], while ψ(k) has the dimension of tion decays to its thermal value with damped oscillations. volume, hence κ2 has the dimension of [energy]3 × volume. While the steady state is independent of the initial condition, We would like to note that the system bath coupling in this the modes which are populated initially approach the steady problem has different dimensions than in the problem with value from above while the correlation in the unpopulated the Ohmic bath. modes oscillate around the long time value. We plot the abso- The non-unitary dynamics of the phonons is governed by lute value of the deviation of the correlation function from its the retarded self energy ΣR(k; t − t0) and the Keldysh self en- steady value on a log scale for k = 0.25π/a (nk = 0 at t = 0) ergy ΣK (k; t, t0). The real part of ΣR leads to the dressing of in Fig-1(c) and for k = 0.5π/a (nk = 1 at t = 0) in Fig.1(d). phonon dispersion while its imaginary part controls the dis- An exponential fit to the envelop of such curves is used to de- sipation in the system. the Keldysh self energy ΣK controls termine the momentum dependent decay rate γk. In Fig.1(e), the stochastic noise from the external bath. It is clear from we plot γk as a function of k for the five different bath band- Eq. (23) that the phonons actually couple to the particle hole widths ranging from σ = ω0 to σ = 5 ω0. For σ = ω0 we excitations of the fermionic system. The retarded self energy find that the decay rate strongly depends on k, decreasing by is then given by, a factor of 2.5 as we go from the zone centre to the edge of the Brillouin zone. As σ is increased, the dependence of γk on X F (k+q) − F (q) ΣR(k, ω) = κ2 , (24) k is weakened with an almost k-independent γ for σ = 5ω0. ω + iη + q − k+q The large bandwidth bath thus behaves like a source of Gaus- q sian white noise53. In the large bandwidth limit we consider F (x) = tanh x−µ  ΣR the effect of the system bath coupling on the decay rate γ in where 2T . The imaginary part of , which Fig.1(f). As expected, the system decays faster as system controls the dissipation in the system, is related to the spectral bath coupling is increased. We note that with increasing sys- function of the effective bath J(k, ω) by tem bath coupling γ acquires a weak k-dependence even for Im[ΣR(k, ω)] = −κ2J(k, ω) σ = 5ω0. 2 X = −πκ δ(ω − k+q + q)[F (k+q) − F (q)]. q B. Phonons Coupled to a Fermionic Bath (25) Using fluctuation-dissipation theorem, the Keldysh self en- We now consider the dynamics a system of scalar fields ergy is given by initialized to a Fock state and coupled to a bath of non-  ω  interacting Fermions. This is prototype of a system which ΣK (k, ω) = 2i coth Im[ΣR(k, ω)] can be found in varied contexts in nature like phonons cou- 2T 81–83 84,85 X pled to electrons , light coupled to metallic systems , = i2πκ2 δ(ω −  +  )[F ( )F ( ) − 1]. 86 k+q q k+q q ultracold atoms in cavity , descriptions of early and q 87 multi-component dark matter systems . For concreteness we (26) will consider phonons coupled to non-interacting fermions. We consider a bath of non-interacting spinless Fermions For the non-equilibrium evolution of the phonon correla- with a Hamiltonian given by tors, we need to the self-energies and use Eq. (18) to evolve the Dyson equations. X † HF = k ψ (k)ψ(k), (22) The particle-hole excitations to which the scalar fields cou- k ples, have qualitatively different density of states (DOS) in one and two dimensions. This difference is reflected in the where ψ(k) are Fermion annihilation operators in mode k dynamics of the scalar fields through the self-energies. So, and k is the corresponding dispersion. We consider fermions we will consider the case of one and two dimensional systems hopping on a linear chain in one dimension and on a square separately lattice in two dimensions with only nearest neighbour hop- ping which results in the dispersions k = −B cos ka and k = −B[cos kxa + cos kya] respectively, where a is the lat- tice constant. Here, the bandwidth of the fermions is given by 1. One dimensional systems 2B (1-D) and 4B (2-D) respectively. The Fermionic bath is characterized by a temperature T and a chemical potential µ, There are two reasons why one dimensional systems show which fixes the particle density in the bath. atypical behaviour: (a) energy-momentum conservation rela- The phonons couple to the fermionic bath through the tions impose tight constraints on possible processes in one di- Hamiltonian mension and (b) the density of states of excitations have strong singular behaviour in one dimension. As we will see in this X † Hint = κ ψ (k + q)ψ(q)φ(k), (23) section, both these factors play an important role in the ab- k,q sence of thermalization for long wavelength phonons in one 9

(a) (b) (c)

2 | t s ⟩ ) ⟩ T = 0.017ϵB t ) , k

T k

= 0 − ( − l ( c l k π a c 1 = 0.05 / ϕ ) ϕ B ) k ϵ t ( / l , c k ω ( ϕ l ⟨ c δ

k ϕ |

2 ⟨ ) ) a a / / k π a B B 2 = 0.5 / ϵ ϵ ( k1 ( 0 0 2kFa π ka ϵBt ka

FIG. 2. Non-equilibrium dynamics of phonons coupled to a low temperature fermionic bath. (a) The region of finite density of particle-hole excitations in the fermionic bath at T = 0 is shown in the k − ω plane as a shaded region. The solid brown lines define the upper and lower threshold energies. Note that the region collapses to a line (with slope v ) at low momenta. The phonon dispersion for c = √1 v is also F s 3 F shown as a dashed line. The two momenta k1 and k2, shown by the dots are chosen such that for k1, the phonon energy lies below the particle- K hole bath while for k2, the energy lies within the bath. (b) Time evolution of the correlation function iD (k; t, t) = hφcl(k, t)φcl(−k, t)i for k1 = 0.05π/a, where the phonons do not see a bath (red solid line) and for k2 = 0.5π/a, where the phonons see an effective bath (solid blue line). k1 and k2 are are shown in (a). The correlator at k1 oscillates about its initial value and never approaches its thermal value shown by the dashed red line. On the other hand, the mode at k2 is damped and the correlator approaches its thermal value (dashed blue line) at long 2 3 times. We have used a bath with T = 0.017B and µ = 0.5B and set the system-bath coupling strength κ = 0.005B a. (c) The deviation of the long time value of the correlator from its thermal value plotted as a function of momenta. The long wavelength modes do not thermalize while the higher momentum correlators approach their thermal value. The value of the correlator for modes that do not thermalize depend on the initial conditions. The oscillations corresponds to the initial (1,0,1,0,1,0...) pattern of occupation of the momentum modes.

dimension. To see this, we consider the case where√ the density of bath 3 electrons gives µ = 0.5B, so that vF = Ba. We choose To understand the novel behaviour of the system, we focus 2 a phonon dispersion with ω0 = B so that cs = 0.5Ba, on the polarization function of a one dimensional free Fermi 1 i.e. cs = √ vF . Here the phonon dispersion lies below T = 0 3 gas. We first consider . In Fig.2(a), the region in the the electronic dispersion at low momenta, as shown by the k − ω J(k, ω) T = 0 plane where is finite at is shown as a dashed line in Fig.2(a). We consider a low bath temperature shaded region bounded by solid lines. The upper limit is given of T = 0.017 to illustrate the low temperature behaviour ω = (v /a) sin ka + µ(1 − cos ka) k ≤ B by the curves F for of the system. We have set the system-bath coupling strength |2k − π| ω = 2 sin (ka/2) k > |2k − π| F and B for F , while κ2 = 0.0053 a. In Fig.2(b), we plot the time dependence of ω = (v /a) sin ka−µ(1−cos ka) B the lower limit is given by F the correlation function iDK (k, t, t) = hφ (k, t)φ (−k, t)i k ≤ 2k ω = −(v /a) sin ka + µ(1 − cos ka) cl cl for F and F for for the setup given above (solid lines) two different momenta: k > 2k k F , where F is the Fermi wave-vector of the Fermions a low momentum k = 0.05π/a, where the phonon dispersion v 1 and F is the Fermi velocity. At low momenta, both the upper is outside the energy range of the bath and a high momentum ∼ v k and lower limit F , and thus the region actually merges k = 0.5π/a, where the phonon dispersion lies within the v 2 into a single line with a slope of F . The width of the region bath energy range. The thermal value of the correlator is also ∼ k2 at low momenta is . This is a consequence of the strong indicated in this plot with dashed lines. It is clear that the constraints of energy and momentum conservation in one di- mode at k thermalizes, whereas the long wavelength mode mension. In this case, the long-wavelength phonons will not 2 k1 does not thermalize in this case. The correlator at k1 un- see any bath and undergo unitary quantum dynamics unless dergoes unitary quantum dynamics with a dressed phonon en- c v the phonon velocity s is exactly equal to F . Note that, in an ergy, oscillating about its initial value and never approach- interacting system, where one would expect a Luttinger liquid ing the thermal value shown with the dashed red line. On like behaviour88, one would once again get a linearly dispers- the other hand, the damped correlator at k2 approaches the ing mode with the velocity tuned by the Luttinger parameter. thermal value (dashed blue line) at long times. To see which Further, at low temperatures, the spectral weight outside this modes are thermalizing, we plot the difference between the region is exponentially small. Since this fine tuning is impos- long time value of the non-equilibrium correlators and its ther- sible in real systems, the long wavelength phonons will not K mal value. iδD (k; t, t) = δhφcl(k, t)φcl(−k, t)i, as a func- thermalize at low temperature (or more precisely have an ex- tion of k in Fig.2(c). We use the value of the correlator at ponentially long thermalization time) in one dimension. 10

(a) (b) (c)

2 | t

s T ϵ

= 0.17 B ) ⟩ k π a t = 0.05 / ) i

Dressed k ωmax(k) n u undamped − . ( l c modes b k = 0.04π/a r ϕ B ) a ϵ ( k / ( l ) c ω ω ϕ , ⟨ k δ ( | ) � a k π a / = 0.03 / B

T > 0 ϵ ( 0 0.0 0.1 0 π ka ka ω/ϵB (d) (e) (f)

k π a 2

= 0.04 / B a a / / ϵ / 2 B π π ] ϵ ) / ] ω ) 0.04 0.07 , ω k = = ( ) , 1 k k k ω ( − , ) R ω ϵ k R / B Σ ( [ J D e ( a [ R B e ϵ R k = 0.04π/a

ω/ϵB ω/ϵB ω/ϵB

FIG. 3. Dynamics of phonons coupled to high temperature fermionic bath and emergence of dressed undamped modes. (a) Particle-hole density of states (DOS) of the fermionic bath is finite only in the shaded region in the k − ω plane at finite temperature. It is bounded above by the solid black line given by ωmax(k) = 2B sin(ka/2) . The phonon dispersion with cs = 0.5B a is also plotted as a dashed blue line. The dots above ωmax(k) correspond to the undamped “polarinon” modes formed due to strong fermion-phonon coupling. (b) The deviation of the correlation function at long times from its thermal value at a moderate temperature T = 0.17B and system-bath coupling strength 2 3 κ = 0.005B a. A few long wavelength phonon modes do not thermalize. (c) Phonon spectral function for a few low momentum modes. The broader peak represents the original dresses damped modes of the phonons. The sharp peaks indicates undamped modes. (d) Bath spectral functions for k = 0.04π/a and k = 0.07π/a are plotted in solid red and blue lines respectively. They have sharp cut-offs at ω = ωmax(k) indicated by dashed vertical lines. The spectral function has a √ 1 singularity near this edge. (e) ReΣR(k, ω) as function of ω for ωmax−ω k = 0.04π. It diverges at ω = ωmax(k) (vertical green dashed line) from above and goes to a constant value inside. (f) The inverse for the phonon modes dressed by the fermionic bath. the inverse propagator has two zero crossings, one near the bare phonon frequency, corresponding to damped dressed phonons, and one above the bath edge [ωmax(k), shown as dashed green line] corresponding to undamped polarinons [Inset shows the region near ωmax(k) clearly showing the second zero crossing].

Bt = 12000 as the “long time” value, and average over a all k. The region in the k − ω plane where J(k, ω) is finite timescale of Bt ∼ 500 to smooth out effects of the oscil- at high temperatures is shown as shaded region in Fig.3(a). lations in modes that do not thermalize. It is clear that low Note that the finite bandwidth of the Fermions still results in momentum modes that lie outside the effective bath do not an ultraviolet cut-off in the bath; i.e. for each k, there is a max- thermalize. For these modes the long time value of the corre- imum energy ωmax(k) = 2B sin(ka/2) beyond which the lator depends on initial condition, and the oscillations with k bath has no spectral weight at any temperature. This is plotted represent the (1,0,1,0,1,0...) pattern of initial occupation. as a solid line in Fig.3(a). One would thus expect the low The energy conservation constraints are relaxed as we in- momentum modes to thermalize once the bath has sufficiently crease the bath temperature. In this case, the bath spectral high temperature, provided the phonon dispersion lies below function gains weight at low energies starting from ω = 0 for this maximum energy, i.e. the phonon velocity cs < Ba, 11

(a) (b) ues of k at T = 0.17B and µ = 0.5B. Using Kramers- Kronig relations91, one can easily show that this would also p R k = 0.04π/a | t T = 0.17ϵB lead to a ∼ 1/ ω − ωmax(k) divergence in ReΣ (k, ω) s ⟩ ) )

t ω k

max( ) k i T = 0.4ϵB for frequencies just above the band edge. This is shown in n − R ( u l . c Fig.3(e), where it is evident that Re Σ (k, ω) → +∞ as T = 0.4ϵB ϕ b ) r

k we approach the band edge from above. In contrast, the real a ( ( l c )

ϕ part of the self energy approaches a constant as the frequency ⟨ ω , δ k |

( comes close to the band edge from below. The real part of the ) a � T = 0.17ϵB /

B retarded self energy dresses the spectrum, and its divergence ϵ

( R−1 2 2 R ensures that D (k, ω) = 2(ω − Ωk) − Σ (k, ω) has ze- 0.0 0.1 roes just above the upper edge of the bath; i.e. close to ω ∼ ω ϵ ka R−1 / B ωmax(k). This is shown in Fig.3(f), where Re D (k, ω) is plotted as a function of ω, revealing the two zero cross- FIG. 4. Effect of temperature on non-thermalizing behaviour of ings. The lower energy zero crossing is related to the phonon phonon modes. (a) Phonon spectral function at T = 0.17B and mode (now dressed by the bath), while the higher frequency T = 0.4B for the mode k = 0.04π/a which does not thermalize. zero crossing (shown more clearly in the inset) is related to As temperature is increased spectral weight in the polarinon mode divergence of the bath spectral function. Thus the strong di- becomes larger. (b) Absolute difference of the long time correlation vergence of the bath density of states leads to an additional function from its thermal value at long time at a high temperatures pole in the dressed phonon Green’s function above the upper T = 0.17B (solid red line) and T = 0.4B (dashed purple line). edge of the bath. This undamped “polarinon” mode, which The deviation is higher at the higher temperature. We have set the system-bath coupling strength κ2 = 0.0053 a. shows up as a sharp feature in Fig.3(c), does not thermalize B even at high temperatures. This explains the lack of thermal- ization seen in Fig.3(b) even for high temperatures where the bare phonon sees a substantial bath spectral density. which is expected to be valid in generic one dimensional ma- The dispersion of the polarinon mode ωpl is plotted as a 89,90 k terial systems like nanowires . In Fig.3(b), we plot the dotted line in Fig.3(a). As momentum of the mode increases, difference between the long time value and the thermal value the gap between the bare phonon dispersion Ωk and the bath of the correlator, iδDK (k; t, t) = δhφ (k, t)φ (−k, t)i for cl cl band edge ωmax(k) increases. Hence the dispersion of the vF the phonon modes with cs = √ , for which we had earlier R pl pl 2 2 3 polarinon, given by Re Σ (k, ωk ) = 2[(ωk ) − Ωk], comes looked at thermalization at low bath temperatures. The disper- pl closer to the bath band edge (ωk needs to be close to ωmax(k) sion of this mode is shown as a dashed blue line in Fig.3(a). R pl We see that the low momentum modes do not thermalize even for Re Σ (k, ωk ) to be large). We note that the quasiparticle residue for the polarinon at a reasonably high temperature of T = 0.17B, although all the modes now see a particle-hole bath with sufficient spec- −1 " R # 3 tral weight. We note that this lack of thermalization remains pl pl ∂ReΣ (k, ω) h pl i 2 Z (k) = 4ωk − ∼ ωk − ωmax(k) if the temperature of the bath is further increased (see later for ∂ω pl ωk temperature dependence). (27) To get an insight into the lack of thermalization of low mo- vanishes rapidly as its dispersion approaches the band edge. mentum modes even at high temperatures, in Fig.3(c), we Thus at large momenta, the effect of the polarinon modes on plot the spectral function of the dressed phonons in the steady the non equilibrium dynamics of the phonons is minimal. This state, A(k, ω) = −(1/π)ImDR(k, ω), as a function of ω for is why the absence of thermalization at high temperatures is several values of k. In Fig.3(c), the spectral functions for dif- restricted to low momenta. While most quasiparticle modes ferent modes are shifted by arbitrary amounts to make them are smeared out in presence of finite temperature, a unique visible. In addition to the original phonon mode, which is feature of the polarinon modes is the fact that their spectral dressed and has a width, we find another sharp mode with weight increases with temperature. This is shown in Fig.4(a), no damping at an energy above the boundary of the particle- where we plot the spectral function of a phonon mode with hole bath. This mode appears due to the strong coupling be- k = 0.04π/a at two different temperatures, T1 = 0.17B tween the phonons and the electron-hole pairs, similar to po- and T2 = 0.4B. Thus, with increasing temperature, the de- lariton modes which occur due to strong coupling between viations of the phonon correlators from their thermal values photons and excitons68,69. We will call these modes the “po- increases, as seen in Fig.4(b). larinon” modes. A key difference with polaritons is that un- To summarize, in one dimension, phonons coupled to like excitons, the particle-hole excitations are not coherent; Fermionic baths do not thermalize, especially the modes at i.e. a simple mode coupling theory would not work here. low momenta. At low temperature, this is primarily due to Rather the origin of this mode can be traced back to the strong strong energy momentum constraints, which leads to a very p ∼ 1/ ωmax(k) − ω divergence of the spectral density of the narrow energy band of particle-hole excitations at low mo- particle-hole excitations at the band edge in 1-d, which gives a menta. However, the lack of thermalization at high tempera- similar divergence in ImΣR(k, ω). This is shown in Fig.3(d), tures is dominated by the formation of the undamped polar- where we plot the bath spectral function for two different val- inon modes with energies above the bath band edge due to 12 strong divergence of the bath density of states near the band of self energy and hence there are no additional undamped edge. The effect of these modes increases both with temper- dressed states (poles in the Green’s function) of the system. ature and with electron-phonon coupling in the system. The Thus the system continues to thermalize as temperature is in- effect also increases if the phonons are stiffer, so that the gap creased, unlike the one dimensional phonons. between the phonon dispersion and the band edge is reduced. The thermalization of the scalar fields can be prohibited if the fields are massive. In the context of phonons, this would correspond to optical phonons in the system98. The low mo- 2. Two dimensional systems mentum dispersion of the massive fields, with m = 0.2B is plotted in Fig.5(a) as a dashed purple line. As k → 0, In this section we will consider the non-equilibrium dy- the dispersion of the phonons lies above the bandwidth of the namics of phonons coupled to fermions in two dimensions particle-hole excitations and hence one would expect that the and see how the dynamics differ from that in one dimen- long wavelength modes would not thermalize. As the momen- sion. As we have seen before, the dynamics is governed tum is increased, the dispersion enters the region of finite bath by the density of states of particle-hole pairs of fermions. spectral density and these higher momentum modes thermal- Contrary to one dimension, the energy momentum conserva- ize. This is clearly seen in Fig.5(d) where we plot the time tion relations can be satisfied much more easily in two di- evolution of the correlator hφcl(k, t)φ(−k, t) (solid lines) for mensions. Hence even at T = 0, the density of states of two different values of k along with the thermal values of the particle-hole excitations with a given momentum transfer k correlators (dashed lines). The mode with the low momentum remains finite at arbitrary low energies. The finite bandwidth (0.03π/a, 0.03π/a) (shown by red line) does not thermalize of the lattice fermions leads to an upper threshold energy as its energy lies above the bath band. It oscillates about its √ initial value sufficiently far from the thermal value (dashed ω (k) = 2vF sin ka+µ(1−cos ka), which disperses lin- max a red line). On the other hand, the correlator of the mode at early at low |k|, i.e. ω (k) ∼ v k, where v is the Fermi max F F (0.08π/a, 0.08π/a), which lies within the bath energy ranges, velocity of the fermions in the bath. Beyond this threshold approaches its thermal value in the long time limit. Fig.5(e) energy the density of states of particle-hole excitations vanish. plots the deviation of the long time correlation functions from In Fig.5(a) the particle-hole DOS is sketched schematically in their thermal value as a function of |k| along the (1, 1) direc- the k − ω plane along the direction k = k in k-space up-to x y tion. It is clear that there is a sharp cut-off in momentum, be- k = k = 0.1π/a. In the figure, the upper bound ω (k) is x y max low which the modes do not thermalize for the massive scalar represented by a solid brown line. In the shaded region below field. This corresponds to the lowest momentum for which the this line we have finite particle-hole DOS. dressed dispersion lies below ω (k). Linearly dispersing long wavelength phonon modes with max cs < vF would see a bath with substantial spectral den- sity even at very low temperatures and would thermalize. In Fig.5(a), we also plot the bare dispersion of a longitudi- IV. CONCLUSION nal phonon mode with cs = 0.73vF as a green dashed line to illustrate this point. Note that the criterion cs < vF is In this paper, we have proposed a new method to study satisfied for most material systems other than compensated non-equilibrium dynamics of scalar fields starting from non- semimetals92–94 or bilayer graphene95 like systems near their thermal initial conditions. This extends earlier work on charge neutrality point. In these systems, by tuning the car- Schrodinger¨ Bosons to the case of scalar fields. The method rier density, one can possibly see a lack of thermalization works by adding a source to the bilinears of quantum fields at of the phonons, although at these very low densities, dis- the initial time in a Schwinger Keldysh field theoretic formal- order and interaction would play a very important role96,97 ism. The correlation functions are calculated in presence of and the simple picture of a non-interacting bath has to be this source. One then takes a set of derivatives of this corre- suitably modified. In Fig.5(b) we have plotted the dynam- lation function with respect to this source (with the set deter- ics of the correlation function hφcl(k, t)φ(−k, t) as a func- mined by the initial conditions) to obtain the physical correla- tion of time for massless phonon modes (with cs = 0.73vF ) tors. The key difference between the Schrodinger¨ bosons and for two different values of momenta, k1 = (0.03π, 0.03π) the scalar fields is the coupling of the sources to both the fields (red line) and k2 = (0.08π, 0.08π) (blue line) at a low tem- and their time derivatives, reflecting the nature of the classical perature T = 0.017B and system-bath coupling strength equations of motion. 2 3 2 κ = 0.005Ba . The thermal values of the correlators are We use this method to study non-equilibrium dynamics indicated by the dashed lines. It is clear in this case that both of massless and massive scalar fields, initialized to athermal these modes thermalize at long times. states, and coupled to external baths. For concreteness, we We would like to note that the bath density of states in consider a system of phonons with the massive fields corre- two dimensions do not have the strong divergence at the band sponding to optical phonons and the massless fields corre- edges that was present in the one dimensional case. This is sponding to longitudinal phonons. We first consider coupling shown in Fig.5(c), where we plot the spectral density of the the system to an Ohmic bath with a smooth ultraviolet cut- particle-hole excitations for two different momentum trans- off. In this case, the system thermalizes with the one par- fers. The softening of the band edge non-analyticity means ticle distributions relaxing to their thermal values. The relax- that there is no corresponding divergence in the real part ation rate is momentum dependent when the ultraviolet cut-off 13

(a) (b) (c) 0.68 ⟩ T ϵ m a /

) T ϵ = 0.017 B = 0 a = 0.017 B t / π , π k − 0.08 ( 0.03 l ) c kx ky π a = = = 0.03 / = ω ϕ y B ) , y k t ϵ k k / , ( =

m ϵ k J = ω = 0.2 B ( x 2 l x k c a k ϕ B

⟨ k k π a x = y = 0.08 / ϵ ) 2 a /

m T B = 0 = 0 ϵ 0 ( 0 0.1π ka ϵBt ω/ϵb

(d) (e) | ⟩ t ) T ϵ m ϵ s t B B T ϵ = 0.017 = 0.2 ⟩ = 0.017 B , ) k k m = 0.2ϵB − − ( ( l l c k k π a c ϕ x = y = 0.08 / ϕ ) ) t k , ( l k c ( l ϕ c ⟨ ϕ δ ⟨ | ) kx = ky = 0.03π/a ) 2 2 a / a / B B ϵ ( ϵ (

ϵBt ka

FIG. 5. Non-equilibrium dynamics of phonons coupled to fermionic bath in two dimensions. (a) Schematic depiction of particle-hole density of states (DOS) of the fermionic bath on k − ω plane along k = k direction at T = 0. The shaded region, from ω = 0 to ω (k) = √ x y max 2vF a sin ka + µ(1 − cos ka) (shown in brown solid line) has finite DOS. Dispersion of acoustic (massless) phonons (m = 0) with cs = B a = 0.73vF is given by the dashed green line. The dashed purple line is the dispersion of optical (massive) phonons with m = 0.2B K and cs = B a = 0.73vF . (b) Time evolution of correlation function iD (k; t, t) = hφcl(k, t)φcl(−k, t)i at k = (0.03π/a, 0.03π/a) (red solid line) and k = (0.08π/a, 0.08π/a) (blue solid line) for massless phonons are plotted. Both the correlators relax to their thermal values (shown with dashed lines) at long times. (c) Bath spectral functions for k = (0.03π/a, 0.03π/a) and k = (0.08π/a, 0.08π/a) are plotted in solid red and blue lines respectively. It’s finite value till ω = 0 indicates the absence of a lower bound in two dimensions. Unlike one dimension, there is no divergence in the bath spectral function in two dimensions. (d) Dynamics of correlation function iDK (k; t, t) at k = (0.03π, 0.03π) (red line) and k = (0.08π, 0.08π) (blue line) for massive phonons with m = 0.2B and cs = B a = 0.73vF . the dispersion of the mode at k = (0.03π/a, 0.03π/a) lies above the band threshold and hence the correlator does not relax to its thermal value (shown by dashed red line). On the other hand the energy of the modes at k = (0.08π/a, 0.08π/a) lies within the region of finite bath spectral function. So the corresponding correlator relaxes to its thermal value at long times. (e) The deviation of the long time value of the correlation function iDK (k; t, t) from its thermal value as a function of momenta. The low momentum modes, with dispersion above the band threshold do not thermalize. All data in this figure is obtained for a bath temperature T = 0.017B and chemical potential µ = 0.5B and system-bath 2 3 2 coupling strength κ = 0.005B a . is small, and approaches a momentum independent Gaussian this is due to the effectively zero bandwidth of the particle- white noise limit as the cut-off is increased. hole excitations with small momentum transfer, as a result of which the long wavelength phonon modes do not see any bath We then consider the dynamics of these phonon modes and undergoes unitary quantum motion. At high temperatures, coupled to a system of non-interacting fermions, where the the lack of thermalization is dominated by the formation of phonons couple to the particle-hole excitations of the system. undamped dressed modes just above the upper threshold of the In one dimension we find that the long wavelength phonons bath. These modes are neither phonon, nor fermion modes; fail to thermalize at all temperatures. At low temperatures, 14 rather these “polarinon” modes arise due to strong fermion We note that the method we have constructed is much more phonon coupling in the system together with divergences in widely applicable than the models we have considered in this the density of states of one dimensional particle-hole excita- paper, including the study of non equilibrium dynamics of in- tions. These modes gain spectral weight with increasing tem- teracting scalar field theories. We believe this method will find perature; as a result the long wavelength modes fail to ther- much wider applications in the future. malize at any temperature.

We finally consider phonons coupled to fermionic baths in ACKNOWLEDGMENTS two dimensions. Here simultaneous energy momentum con- servation does not lead to stringent criteria and the bath band- The authors are grateful to Ahana Chakraborty for useful width is finite at low momenta at all temperatures. Further discussions and suggestions. The authors acknowledge the the strong divergence of density of states is also absent. As use of computational facilities at the Department of Theoreti- a result, we recover the typical thermalizing behaviour of the cal Physics, Tata Institute of Fundamental Research, Mumbai phonon modes in two dimensions. for this paper.

Appendix A: Green’s functions for single Harmonic Oscillator

In developing the extended Keldysh formalism for the harmonic oscillators in the position basis, a crucial step was the inversion of the matrix that appears in the u dependent action, Eq. (6). This leads to the u dependent one-particle correlator D(u), which finally leads to the physical correlation functions in the system. In this appendix, we present the details of this non-trivial inversion. The matrix in Eq. (6) can be separated two parts, one of which is independent of the source u. This decomposition can be written as

D−1(t, t0; u) = D−1(t, t0; 0) − ∆(t, t0; u), (A1) where

 2 2  −1 0 0 (−∂t − Ω ) + δt0(−∂t + iω) 0 D (t, t ; 0) = δ(t − t ) 2 2 , (A2) 0 (∂t + Ω ) + δt0(∂t + iω) and

0 2u x ∆(t, t ; u) = iω [−u + σ ]δ δ 0 . (A3) 1 − u2 I2 t0 t 0

−1 0 0 Here I2 is 2 × 2 identity matrix. Inverting D (t, t ; 0) we get the components of D(t, t ; 0) 1 i  ω2  D+−(t, t0, 0) = sin Ω(t − t0) − cos Ωt cos Ωt0 + sin Ωt sin Ωt0 2Ω 2ω Ω2 1 i  ω2  D−+(t, t0, 0) = − sin Ω(t − t0) − cos Ωt cos Ωt0 + sin Ωt sin Ωt0 2Ω 2ω Ω2 D++(t, t0, 0) = Θ(t − t0)D−+(t, t0, 0) + Θ(t0 − t)D+−(t, t0, 0) and D−−(t, t0, 0) = Θ(t − t0)D+−(t, t0, 0) + Θ(t0 − t)D−+(t, t0, 0). (A4)

−1 00 00 0 0 0 One can easily check that D (t, t ; 0)D(t , t ; 0) = δ(t − t )I2. Now D(t, t ; u) can be found using the D(t, t0; u) = D(t, t0; 0) + D(t, 0; 0)[∆(0, 0; u) + ∆(0, 0; u)D(0, 0; 0)∆(0, 0; u) + ...]D(0, t0; 0) (A5)

i x Using the fact that D(0, 0; 0) = 2ω [I + σ ], the series inside the brackets of Eq. (A5) can be evaluated using standard Pauli  2u  x matrix identities as iω 1−u σ . Using this and Eq. (A4) we get the source dependent Green’s functions in Eq. (8). One then takes the required u derivatives to get the physical correlation function. 15

(a) (b)

K G (q, Ω) ψ†(q + k, Ω + ω) κ 1

ϕcl(k, ω) (−k, − ω) (k, ω) ψ1(q, Ω)

ψ†(q + k, Ω + ω) κ 2

ϕcl(k, ω) GR k q ω ψ2(q, Ω) ( + , + Ω)

ψ†(q + k, Ω + ω) GA q κ 2 ( , Ω)

ϕq(k, ω) k ω k ω ψ1(q, Ω) (− , − ) ( , )

ψ†(q + k, Ω + ω) κ 1

ϕq(k, ω) ψ q K 2( , Ω) G (k + q, ω + Ω)

FIG. 6. Coupling of phonons to a fermionic bath. (a) Allowed interaction vertices in terms of the Keldysh rotated fields. The solid line represents the fermionic field ψ1 and the dashed line represents ψ2. Solid wavy line represent the scalar field φcl and the dotted wavy line stand for the scalar field φq. κ is the coupling strength between the scalar field and the fermions. (b) Diagrams for retarded self energy ΣR(k, ω) of phonons coupled to a fermionic bath. Note that the retarded self energy of phonons is related to the polarization function of the fermions.

Appendix B: Phonon self energy from the fermionic bath

In this appendix we present the details of the derivation of the phonon self-energy when the phonons are coupled to a bath of free fermions through the Hamiltonian

X † Hint = κ ψ (k + q, t)ψ(q, t)φ(k, t), (B1) k,q

On the Keldysh contour49 (+/- basis) its contribution to the action is given by

Z +∞ X h † † i Sint = −κ dt ψ+(k + q, t)ψ+(q, t)φ+(k, t) − ψ−(k + q, t)ψ−(q, t)φ−(k, t) (B2) k,q −∞ 16

It is useful to work in the Keldysh rotated basis49. Keldysh rotated basis for scalar fields: 1 φcl(k, t) = [φ+(k, t) + φ−(k, t)] 2 (B3) 1 φ (k, t) = [φ (k, t) − φ (k, t)]. q 2 + −

Keldysh rotated basis for fermions: 1 1 ψ1(k, t) = √ [ψ+(k, t) + ψ−(k, t)] ψ2(k, t) = √ [ψ+(k, t) − ψ−(k, t)] 2 2 (B4) 1 1 ψ†(k, t) = √ [ψ† (k, t) − ψ† (k, t)] ψ†(k, t) = √ [ψ† (k, t) + ψ† (k, t)]. 1 2 + − 1 2 + − In the Keldysh rotated basis the part of action is given by

Z +∞ X † † Sint = −κ dt[ψ1(k + q, t)ψ1(q, t)φcl(k, t) + ψ2(k + q, t)ψ2(q, t)φcl(k, t) k,q −∞ (B5) † † + ψ1(k + q, t)ψ2(q, t)φq(k, t) + ψ2(k + q, t)ψ1(q, t)φq(k, t)].

In frequency space it turns into

X Z +∞ dω Z +∞ dΩ S = −κ [ψ†(k + q, ω + Ω)ψ (q, Ω)φ (k, ω) + ψ†(k + q, ω + Ω)ψ (q, Ω)φ (k, ω) int 2π 2π 1 1 cl 2 2 cl k,q −∞ −∞ (B6) † † + ψ1(k + q, ω + Ω)ψ2(q, Ω)φq(k, ω) + ψ2(k + q, ω + Ω)ψ1(q, Ω)φq(k, ω)]. The terms in Eq. (B6) gives allowed vertices for system-bath coupling. They are sketched in Fig.6(a). Diagrams for retarded self energy of the phonons are sketched in Fig.6(b). They are given in terms of the vertices and free fermionic Green’s functions. The retarded self energy is

X Z ∞ dΩ ΣR(k, ω) = −iκ2 [GR(k + q, ω + Ω)GK (q, Ω) + GK (k + q, ω + Ω)GA(q, Ω)]. 2π (B7) q −∞

Here Green’s functions of the free fermions at thermal equilibrium are

R A ∗ 1 G (k, ω) = [G (k, ω)] = + and (B8) 0 ω−k+i0 K G0 (k, ω) = −2πiF (ω)δ(ω − k), (B9) ω−µ  where F (ω) = tanh 2T . T is the temperature of the bath and µ is the chemical potential.

It is then easy to see (see Fig.6(b) for the corresponding Feynman diagrams) that the retarded self energy for phonons is just the polarization function of free fermi gas multiplied by factors of system-bath coupling strength. Upon simplifying we get,

R 2 X F (k+q) − F (q) Σ (k, ω) = κ + ω + i0 + q − k+q q (B10) Z ∞ 0 0 2 dω J(k, ω ) = −κ 0 + , −∞ 2π ω − ω − i0 where bath spectral function J(k, ω) is given by X J(k, ω) = π [F (k+q) − F (q)]δ(ω − k+q + q). (B11) q 17

The Keldysh self energy is given by (using fluctuation-dissipation theorem)  ω  ΣK (k, ω) = 2i coth Im[ΣR(k, ω)] 2T X  ω  = −i2πκ2 coth [F ( ) − F ( )]δ(ω −  +  ) 2T k+q q k+q q q        X k+q − µ q − µ k+q − µ q − µ = −i2πκ2 coth − tanh − tanh δ(ω −  +  ) 2T 2T 2T 2T k+q q (B12) q      X k+q − µ q − µ = −i2πκ2 1 − tanh tanh δ(ω −  +  ) 2T 2T k+q q q 2 X = i2πκ [F (k+q) F (q) − 1] δ(ω − k+q + q). q

[email protected] 25 E. Calzetta and B. L. Hu, Phys. Rev. D 37, 2878 (1988). 1 F. Englert and R. Brout, Phys. Rev. Lett. 13, 321 (1964). 26 J. Berges, Nuclear Physics A 699, 847 (2002). 2 P. Higgs, Physics Letters 12, 132 (1964). 27 A. Anisimov, W. Buchmuller,¨ M. Drewes, and S. Mendizabal, 3 P. W. Higgs, Phys. Rev. Lett. 13, 508 (1964). Annals of Physics 324, 1234 (2009). 4 A. R. Liddle, in ICTP Summer School in High-Energy Physics and 28 K. Mukaida and K. Nakayama, Journal of Cosmology and As- Cosmology (1999) arXiv:astro-ph/9901124. troparticle Physics 2013, 017 (2013). 5 B. A. Bassett, S. Tsujikawa, and D. Wands, Rev. Mod. Phys. 78, 29 C. Thomsen, H. T. Grahn, H. J. Maris, and J. Tauc, Phys. Rev. B 537 (2006). 34, 4129 (1986). 6 J. Magana˜ and T. Matos, Journal of Physics: Conference Series 30 J. Qi, X. Chen, W. Yu, P. Cadden-Zimansky, D. Smirnov, N. H. 378, 012012 (2012). Tolk, I. Miotkowski, H. Cao, Y. P. Chen, Y. Wu, S. Qiao, and 7 L. Hui, J. P. Ostriker, S. Tremaine, and E. Witten, Phys. Rev. D Z. Jiang, Applied Physics Letters 97, 182102 (2010). 95, 043541 (2017). 31 D. Novko, F. Caruso, C. Draxl, and E. Cappelluti, Phys. Rev. Lett. 8 F. Chadha-Day, J. Ellis, and D. J. E. Marsh, “Axion dark matter: 124, 077001 (2020). What is it and why now?” (2021), arXiv:2105.01406 [hep-ph]. 32 P.-A. Mante, L. Belliard, and B. Perrin, Nanophotonics 7, 1759 9 S. Doniach and E. H. Sondheimer, Green’s Functions for Solid (2018). State Physicists (Published by Imperial College press and dis- 33 H. Kawashima, M. M. Wefers, and K. A. Nelson, Annual Review tributed by World Scientific Publishing Co., 1998). of Physical Chemistry 46, 627 (1995), pMID: 24341370. 10 C. E. Patton, Physics Reports 103, 251 (1984). 34 A. Weiner and A. Kan’an, IEEE Journal of Selected Topics in 11 P. W. Andrson, Rev. Mod. Phys. 38, 298 (1966). Quantum Electronics 4, 317 (1998). 12 A. Schmitt, Introduction to Superfluidity (Springer International 35 M. Fetterman, D. Goswami, D. Keusters, W. Yang, J.-K. Rhee, Publishing, 2015). and W. Warren, Opt. Express 3, 366 (1998). 13 G. Rempe, R. J. Thompson, R. J. Brecha, W. D. Lee, and H. J. 36 A. M. Weiner, Review of Scientific Instruments 71, 1929 (2000). Kimble, Phys. Rev. Lett. 67, 1727 (1991). 37 A. M. Weiner, Optics Communications 284, 3669 (2011), special 14 R. J. Thompson, G. Rempe, and H. J. Kimble, Phys. Rev. Lett. Issue on Optical Pulse Shaping, Arbitrary Waveform Generation, 68, 1132 (1992). and Pulse Characterization. 15 R. Miller, T. E. Northup, K. M. Birnbaum, A. Boca, A. D. Boozer, 38 P. Ginzburg, Reviews in Physics 1, 120 (2016). and H. J. Kimble, Journal of Physics B: Atomic, Molecular and 39 K. Weiher, E. Agudelo, and M. Bohmann, Phys. Rev. A 100, Optical Physics 38, S551 (2005). 043812 (2019). 16 H. Walther, B. T. H. Varcoe, B.-G. Englert, and T. Becker, Reports 40 A. Dey and M. Kulkarni, Phys. Rev. A 101, 043801 (2020). on Progress in Physics 69, 1325 (2006). 41 O. c. v. Cernotˇ ´ık, A. Dantan, and C. Genes, Phys. Rev. Lett. 122, 17 L. Tonks and I. Langmuir, Phys. Rev. 33, 195 (1929). 243601 (2019). 18 K. Tolpygo, Ukr. J. Phys. 53, 93 (2008). 42 R. Loudon and P. Knight, Journal of Modern Optics 34, 709 19 K. Huang, Nature 167, 779 (1951). (1987). 20 D. L. Mills and E. Burstein, Reports on Progress in Physics 37, 43 A. I. Lvovsky, “Squeezed light,” (2016), arXiv:1401.4118 [quant- 817 (1974). ph]. 21 D. N. Basov, A. Asenjo-Garcia, P. J. Schuck, X. Zhu, and A. Ru- 44 E. A. Sete, H. Eleuch, and S. Das, Phys. Rev. A 84, 053817 bio, Nanophotonics 10, 549 (2021). (2011). 22 M. E. Peskin and D. V. Schroeder, An Introduction to quantum 45 D. Q. Bao, C. J. Zhu, Y. P. Yang, and G. S. Agarwal, Opt. Express field theory (Addison-Wesley, Reading, USA, 1995). 27, 15540 (2019). 23 M. Srednicki, Quantum Theory (Cambridge University 46 J. Schwinger, Journal of Mathematical Physics 2, 407 (1961). Press, 2007). 47 L. V. Keldysh, Sov. Phys. JETP 20, 1018 (1965). 24 A. Altland and B. D. Simons, Condensed Matter Field Theory, 48 A. Kamenev and A. Levchenko, Advances in Physics 58, 197 2nd ed. (Cambridge University Press, 2010). (2009). 18

49 A. Kamenev, Field Theory of Non-Equilibrium Systems (Cam- ergy Physics 2016 (2016), 10.1007/jhep05(2016)164. bridge University Press, 2011). 76 G. Mandal, S. Paranjape, and N. Sorokhaibam, “Thermalization 50 J. Rammer, of Non-equilibrium States in 2d critical quench and uv/ir mixing,” (2020), arXiv:1512.02187 (Cambridge University Press, 2007). [hep-th]. 51 A. Chakraborty, P. Gorantla, and R. Sensarma, Phys. Rev. B 99, 77 G. N. Watson, Journal of the London Mathematical Society s1-8, 054306 (2019). 194 (1933). 52 A. J. Leggett, S. Chakravarty, A. T. Dorsey, M. P. A. Fisher, 78 U. Weiss, Quantum Dissipative Systems, 3rd ed. (WORLD SCI- A. Garg, and W. Zwerger, Rev. Mod. Phys. 59, 1 (1987). ENTIFIC, 2008). 53 C. Gardiner and P. Zoller, Quantum Noise (Springer). 79 A. Chakraborty and R. Sensarma, Phys. Rev. B 97, 104306 (2018). 54 H. E. Elsayed-Ali, T. B. Norris, M. A. Pessot, and G. A. Mourou, 80 H. G. Dawson, Proceedings of the London Mathematical Society Phys. Rev. Lett. 58, 1212 (1987). s1-29, 519 (1897). 55 F. C. Wellstood, C. Urbina, and J. Clarke, Phys. Rev. B 49, 5942 81 A. A. Abrikosov, L. P. Gorkov, I. E. Dzyaloshinski, and R. A. Sil- (1994). verman, Methods of Quantum Field Theory in Statistical Physics 56 N. Del Fatti, C. Voisin, M. Achermann, S. Tzortzakis, (Dover, 1975). D. Christofilos, and F. Vallee,´ Phys. Rev. B 61, 16956 (2000). 82 F. Giustino, Rev. Mod. Phys. 89, 015003 (2017). 57 F. Giazotto, T. T. Heikkila,¨ A. Luukanen, A. M. Savin, and J. P. 83 J.-M. Lai, Y.-R. Xie, and J. Zhang, Nano Research 14, 1711 Pekola, Rev. Mod. Phys. 78, 217 (2006). (2021). 58 A. Habib, F. Florio, and R. Sundararaman, Journal of Optics 20, 84 Z. Huang, W. Zhou, J. Huang, J. Wu, Y. Gao, Y. Qu, and J. Chu, 064001 (2018). Scientific Reports 6, 22938 (2016). 59 F. S. Bergeret, M. Silaev, P. Virtanen, and T. T. Heikkila,¨ Rev. 85 H. Yu, Y. Peng, Y. Yang, and Z.-Y. Li, npj Computational Mate- Mod. Phys. 90, 041001 (2018). rials 5, 45 (2019). 60 S. Dal Forno and J. Lischner, Phys. Rev. Materials 3, 115203 86 H. Ritsch, P. Domokos, F. Brennecke, and T. Esslinger, Rev. Mod. (2019). Phys. 85, 553 (2013). 61 C. Giannetti, M. Capone, D. Fausti, M. Fabrizio, F. Parmigiani, 87 E. Masso,´ F. Rota, and G. Zsembinszki, Phys. Rev. D 66, 023004 and D. Mihailovic, Advances in Physics 65, 58 (2016). (2002). 62 L. V. Besteiro, E. Cortes,´ S. Ishii, P. Narang, and R. F. Oulton, 88 T. Giamarchi, Quantum physics in one dimension (Oxford Unver- Journal of Applied Physics 129, 150401 (2021). sity Press, 2003). 63 T. Peterreins, J. Jochum, F. Probst,¨ F. V. Feilitzsch, H. Kraus, and 89 L. Yang, Y. Tao, J. Liu, C. Liu, Q. Zhang, M. Akter, Y. Zhao, T. T. R. L. Mossbauer,¨ Journal of Applied Physics 69, 1791 (1991). Xu, Y. Xu, Z. Mao, Y. Chen, and D. Li, Nano Letters 19, 415 64 J. B. Hertzberg, O. O. Otelaja, N. J. Yoshida, and R. D. Robinson, (2019). Review of Scientific Instruments 82, 104905 (2011). 90 R. Hoffmann, S. Shaik, J. Scott, M.-H. Whangbo, and M. J. Fos- 65 E. Perinati, M. Barbera, A. Collura, S. Serio, and E. Silver, hee, Journal of Solid State Chemistry 34, 263 (1980). Nuclear Instruments and Methods in Physics Research Section 91 G. B. Arfken, H. J. Weber, and F. E. Harris, in Mathematical A: Accelerators, Spectrometers, Detectors and Associated Equip- Methods for Physicists (Seventh Edition), edited by G. B. Arfken, ment 531, 459 (2004). H. J. Weber, and F. E. Harris (Academic Press, Boston, 2013) 66 M. Zukerstein, F. Trojanek,´ B. Rezek, Z. Sobˇ a´n,ˇ M. Kozak,´ and seventh edition ed., pp. 551–598. P. Maly,´ Applied Physics Letters 115, 161104 (2019). 92 I. Pletikosic,´ M. N. Ali, A. V. Fedorov, R. J. Cava, and T. Valla, 67 M. Lakehal, M. Schiro,´ I. M. Eremin, and I. Paul, Phys. Rev. B Phys. Rev. Lett. 113, 216601 (2014). 102, 174316 (2020). 93 J. Jiang, F. Tang, X. C. Pan, H. M. Liu, X. H. Niu, Y. X. Wang, 68 S. Pekar, Journal of Physics and Chemistry of Solids 5, 11 (1958). D. F. Xu, H. F. Yang, B. P. Xie, F. Q. Song, P. Dudin, T. K. Kim, 69 C. Weisbuch, M. Nishioka, A. Ishikawa, and Y. Arakawa, Phys. M. Hoesch, P. K. Das, I. Vobornik, X. G. Wan, and D. L. Feng, Rev. Lett. 69, 3314 (1992). Phys. Rev. Lett. 115, 166601 (2015). 70 N. OCKMAN, W. WANG, and R. ALFANO, International Jour- 94 L.-K. Zeng, R. Lou, D.-S. Wu, Q. N. Xu, P.-J. Guo, L.-Y. Kong, nal of Modern Physics B 05, 3165 (1991). Y.-G. Zhong, J.-Z. Ma, B.-B. Fu, P. Richard, P. Wang, G. T. Liu, 71 J. Flock, T. Dekorsy, and O. V. Misochko, Applied Physics Let- L. Lu, Y.-B. Huang, C. Fang, S.-S. Sun, Q. Wang, L. Wang, Y.-G. ters 105, 011902 (2014). Shi, H. M. Weng, H.-C. Lei, K. Liu, S.-C. Wang, T. Qian, J.-L. 72 J. Hu, K. Igarashi, T. Sasagawa, K. G. Nakamura, and O. V. Mis- Luo, and H. Ding, Phys. Rev. Lett. 117, 127204 (2016). ochko, Applied Physics Letters 112, 031901 (2018). 95 E. McCann and M. Koshino, Reports on Progress in Physics 76, 73 W. Busza, K. Rajagopal, and W. van der Schee, Annual Review 056503 (2013). of Nuclear and Particle Science 68, 339 (2018). 96 S. Das Sarma, E. H. Hwang, and E. Rossi, Phys. Rev. B 81, 74 A. Mitra, Annual Review of Condensed Matter Physics 9, 245 161407 (2010). (2018). 97 O. Vafek and K. Yang, Phys. Rev. B 81, 041401 (2010). 75 S. R. Das, D. A. Galante, and R. C. Myers, Journal of High En- 98 N. W. Ashcroft and N. D. Mermin, Solid State Physics (Holt- Saunders, 1976).