Keldysh Field Theory for Driven Open Quantum Systems

L. M. Sieberer,1 M. Buchhold,2 and S. Diehl2, 3, 4 1Department of , Weizmann Institute of Science, Rehovot 7610001, Israel 2Institute of Theoretical Physics, TU Dresden, D-01062 Dresden, Germany 3Kavli Institute for Theoretical Physics, University of California, Santa Barbara, CA 93106, USA 4Institut f¨urTheoretische Physik, Universit¨atzu K¨oln,D-50937 Cologne, Germany Recent experimental developments in diverse areas — ranging from cold atomic to light-driven semi- conductors to microcavity arrays — move systems into the focus which are located on the interface of quantum optics, many-body physics and . They share in common that coherent and driven-dissipative quantum dynamics occur on an equal footing, creating genuine non-equilibrium scenarios without immediate counterpart in equilibrium condensed matter physics. This concerns both their non-thermal stationary states, as well as their many-body . It is a challenge to theory to identify novel instances of universal emergent macroscopic phenomena, which are tied unambiguously and in an observable way to the microscopic drive conditions. In this review, we discuss some recent results in this direction. Moreover, we provide a sys- tematic introduction to the open system Keldysh functional integral approach, which is the proper technical tool to accomplish a merger of quantum optics and many-body physics, and leverages the power of modern quantum field theory to driven open quantum systems.

CONTENTS 2 Applications 33

I. Introduction2 III. Non-equilibrium stationary states: spin models 33 A. New phenomena3 A. Ising spins in a single-mode cavity 33 B. Theoretical concepts and techniques4 1. Large-spin Holstein-Primakoff representation 34 C. Experimental platforms5 2. Effective Ising spin action for the single-mode cavity 36 1. Cold atoms in an optical cavity, and microcavity arrays: driven-dissipative spin-boson models5 B. Ising spins in a random multi-mode cavity 37 1. Keldysh action and saddle point equations 38 2. - systems: driven open interacting bosons7 2. Non-equilibrium glass transition 39 3. Thermalization of photons and atoms in the glass 3. Cold atoms in optical lattices: heating 40 dynamics9 4. Quenched and Markovian bath coupling in the D. Outline and scope of this review 10 Keldysh formalism 41

IV. Non-equilibrium stationary states: bosonic models 42 1 Theoretical background 11 A. Density-phase representation of the Keldysh action 44 II. Keldysh functional integral for driven open systems 11 B. Critical dynamics in 3D 46 A. From the quantum master equation to the Keldysh 1. Effective action for driven-dissipative functional integral 12 condensation 47 B. Examples 16 2. Non-equilibrium FRG flow equations 48 1. Single-mode cavity, and some exact properties 16 3. Scaling solutions and critical behavior 49 2. Driven-dissipative condensate 20 C. Absence of algebraic order in 2D 51 C. Semiclassical limit of the Keldysh action 22 D. KPZ scaling in 1D 53 D. Symmetries of the Keldysh action 24 1. Thermodynamic equilibrium as a symmetry of V. Universal heating dynamics in 1D 54 arXiv:1512.00637v2 [cond-mat.quant-] 4 Aug 2016 the Keldysh action 24 A. Heating an interacting Luttinger 54 2. Semiclassical limit of the thermal symmetry 27 B. Kinetic equation 56 3. The Noether theorem in the Keldysh formalism 28 C. Self-consistent Born approximation 58 4. Closed systems: energy, momentum and particle 1. Kinetics for small momenta 59 number conservation 29 2. Scaling of the self-energy 60 5. Extended continuity equation in open systems 30 D. Heating and universality 60 6. Spontaneous symmetry breaking and the 1. densities 60 Goldstone theorem 31 2. Non-equilibrium scaling 61 E. Open system functional renormalization group 32 3. Observability 61 2

VI. Conclusions and outlook 61 Systems located at this new interface are not guaranteed to ther- malize, due to the absence of energy conservation and the re- VII. Acknowledgements 62 sulting breaking of detailed balance by the external drive at the microscale. They rather converge to non-equilibrium stationary A. Functional differentiation 63 states of matter, creating scenarios without counterpart in con- densed, equilibrium matter. This rules out conventional theoreti- B. Gaussian functional integration 63 cal equilibrium concepts and techniques to be used, and calls for the development of new theoretical tools. The physical frame- References 64 work sparks broader theoretical questions on the existence of new phases of bosonic [18, 19] and fermionic [20–22] matter, the nature of phase transitions in such driven systems [10, 23– I. INTRODUCTION 26], and the observable consequences of at the largest scales [27, 28]. Beyond stationary states [29], a fun- damental challenge is set by the time evolution of interacting Understanding the quantum many-particle problem is one of quantum systems, which is currently explored theoretically [30– the grand challenges of modern physics. In thermodynamic 36] and experimentally in cold atomic [37–41] and photonic sys- equilibrium, the combined effort of experimental and theoretical tems [42]. A key goal is to identify universal dynamical regimes research has made tremendous progress over the last decades, that hold beyond specific realizations or precise initial condi- revealing the key concepts of emergent phenomena and uni- tions. Combining idealized closed system evolution with the in- versality. This refers to the observation, that the relevant de- trinsic open system character of any real world experiment takes grees of freedom governing the macrophysics may be vastly dif- this setting to the next stage, and exhibits emergent dynamics ferent from those of the microscopic physics, but on the other markedly different from closed systems both for short [43, 44] hand are constrained by basic symmetries on the short distance and long evolution times [45–51]. scale, restoring predictive power. Regarding the role and power of these concepts in out-of-equilibrium situations, there is a The interplay of coherent and driven-dissipative dynamics can large body of work in the context of classical near equilibrium be a natural consequence of the driving necessary to maintain a and non-equilibrium many-body physics and statistical mechan- certain many-body state. Going one step further, it is possible ics [1–5]. However, analogous scenarios and theoretical tools to exploit and further develop the toolbox of quantum optics for for non-equilibrium quantum systems are much less developed. the driven-dissipative manipulation of many-body systems. Re- This review addresses recent theoretical progress in an im- cently, it has been recognized that the concept of dissipative state portant and uprising class of dynamical non-equilibrium phases preparation in quantum optics [52, 53] can be developed into a of quantum matter, which emerge in driven open quantum sys- many-body context, both theoretically [54–57] and experimen- tems, where a Hamiltonian is not the only resource of dynamics. tally [58–60]. Suitably tailored dissipation then does not neces- This concerns both non-equilibrium stationary states, but also sarily act as an adversary to subtle quantum mechanical correla- the dynamics of such ensembles. Strong motivation for the ex- tions such as phase coherence or entanglement [61–66]. In con- ploration of non-thermal stationary states comes from a recent trast, it can even create these correlations, and dissipation then surge of experiments in diverse areas: in cold atomic gases [6– represents the dominant resource of many-body dynamics. In 8], hybrid light-matter systems of Bose-Einstein condensates particular, even topologically ordered states in spin systems [67] placed in optical cavities are created [9, 10], or driven Ryd- or of fermionic matter [68? ] can be induced dissipatively ([56] berg ensembles are prepared [11]; light-driven and [70–72], respectively). These developments open up a new heterostructures realize Bose-Einstein condensation of exciton- arena for many-body physics, where the quantum mechanical [12, 13]; coupled microcavity arrays are pushed to microscopic origin is of key importance despite a dominantly a regime demonstrating both strong coupling of light and mat- dissipative dynamics. ter [14] and scalability [15]; large ensembles of trapped ions implement varieties of driven spin models [16, 17]. All those Summarizing, this is a fledging topical area, where first results systems are genuinely made of quantum ingredients, and share underpin the promise of these systems to exhibit genuinely new in common that coherent and driven-dissipative dynamics oc- physics. In the remainder of the introduction, we will discuss cur on equal footing. This creates close ties to typical setups in in some more detail the three major challenges which emerge in quantum optics. But on the other hand, they exhibit a continuum these systems. The first one concerns the identification of novel of degrees of freedom, characteristic for many-body physics 1. macroscopic many-body phenomena, which witness the micro- scopic driven open nature of such quantum systems. Second, we anticipate the theoretical machinery, which allows us to perform the transition from micro- to macrophysics in a non-equilibrium 1 This includes the cases of extended spatial continuum and lattice systems. In context in practice. Third, we describe some representative ex- both cases, a continuum of momentum modes obtains, underlying the charac- perimental platforms, which motivate the theoretical efforts, and teristics of many-body problems at long wavelength. in which the predictions can be further explored. 3

A. New phenomena breaks detailed balance [26]. In particular, in three dimen- sions and close to the critical point of driven-dissipative Bose- As pointed out above, one of the key goals of the research re- condensation, a degeneracy of critical exponents indicates a uni- viewed here is the identification of new macroscopic many-body versal asymptotic thermalization, in the sense of an emergent phenomena, which can be uniquely traced back to the micro- thermal FDR. A similar phenomenology is observed in a disor- scopic driven open nature of such systems, and do not have an dered multimode extension of the Dicke model, see Sec. III B. immediate counterpart in equilibrium systems. This underlines the strongly attractive nature of the thermal equi- The driven nature common to the systems considered here can librium fixed point at low frequencies. Still in these systems, always be associated to the fact that the underlying microscopic non-equilibrium conditions leave their traces in the dynamical Hamiltonian is time dependent, with a time dependence relating response of the system, in terms of information that does not to external driving fields such as lasers. When such an ensemble enter the FDR at leading order. For example, the critical behav- (i) in addition has a natural partition into a “system” and a “bath” ior is characterized by a fine structure in a new and indepen- — a continuum of modes well approximated by harmonic oscil- dent critical exponent, which measures decoherence, and whose lators with short memory —, (ii) the system-bath coupling is value distinguishes equilibrium and non-equilibrium dynamics, weak compared to a typical energy scale of the “system” Hamil- see Sec.IVB. tonian and (iii) linear in the bath creation and annihilation oper- ators (so that they can be integrated out straightforwardly), then Instead of emergent thermal behavior indicating the fadeout of an effective (still microscopic) description in terms of combined non-equilibrium conditions upon coarse graining, also the oppo- Hamiltonian and driven-dissipative Markovian quantum dynam- site behavior is possible. For example, low dimensional (d ≤ 2) ics of the “system” ensues. The “system” dynamics obtains by bosonic systems at low noise level, such as exciton-polaritons tracing out the bath variables. well above threshold, are not attracted to the equilibrium fixed The resulting effective microdynamics is “non-equilibrium,” point as their three dimensional counterparts, but rather flow to in a sense sharpened in Sec. II D 1. This not only concerns the the non-equilibrium fixed point of the Kardar-Parisi-Zhang [73] time evolution, but also holds for the non-equilibrium stationary universality class, see Sec.IVC. This can be interpreted as a states. More precisely, the above situation implies an explicit universal and indefinite increase of the non-equilibrium strength, breaking of detailed balance, since the “system” energy is not which is triggered even if the violation of detailed balance at the conserved due to the explicitly time-dependent drive. microscopic level is very small. What do we actually mean by “detailed balance” and “ther- mal equilibrium?” In an operational sense, the principle of de- Universal non-equilibrium phenomena can also occur in the tailed balance states that there is a partition invariance for the time evolution of driven open systems. For example, intriguing temperature (or, more generally, the noise level) present in the scaling laws describing algebraic decoherence [46], anomalous system: an arbitrary bipartition of the system can be chosen, one diffusion [74], or glass-like behavior [45, 75–77] have been iden- part can be traced out, and the resulting subsystem will be at the tified in the long time asymptotics of driven spin systems close same temperature (noise level) as the total system. This parti- to the stationary states. Conversely, the short time behavior of tion invariance is the condition for a globally well-defined tem- driven open lattice bosons shows universal scaling laws directly perature characteristic for systems in thermal equilibrium. More witnessing the non-equilibrium drive, see Sec.V. This scaling formally, thermal equilibrium can be detected by means of so- can be related to a strongly pronounced non-equilibrium shape called fluctuation-dissipation relations (FDRs). These connect of the time-dependent distribution function in the early stages of the two fundamental observables in physical systems — correla- evolution, and be traced back to conservation laws of the driven- tion and response functions (see Sec.IIB). In the case of thermal dissipative generator of dynamics. equilibrium, the connection is dictated by the particle statistics alone. It is then given by the Bose- and Fermi-distributions, re- The above discussion mainly focuses on the difference be- spectively. Deviations from this universal form, which has only tween equilibrium and non-equilibrium systems on the macro- two free parameters (temperature and chemical potential, relat- scopic level of observation. Another direction, still much less ing to the typical conserved quantities energy and particle num- developed, concerns the distinction between classical and quan- ber), provide a necessary requirement for non-equilibrium con- tum effects. Again, although the quantum mechanical descrip- ditions. tion is necessary at a microscopic level, the persistence of quan- In non-equilibrium stationary states, no such general form ex- tum effects at the macroscale is not guaranteed. This is mainly ists. We will encounter a concrete and simple example in the due to the Markovian noise level inherent to such quantum sys- context of the driven Dicke model (a cavity photon coupled to tems. Nevertheless, systems with suitably engineered driven- a collective spin) in Sec. III A, where the form of the FDR de- dissipative dynamics show typical quantum mechanical phe- pends on the observable we are choosing (e.g., the position or nomena such as phase coherence [54, 55, 61, 63, 64, 78], en- momentum correlations and responses). tanglement [58–60, 62], or topological order [56, 66, 70–72]. On the other hand, thermal FDRs can emerge at long Especially fermionic systems, which do not possess a classical wavelength, even though the microscopic dynamics manifestly limit, are promising in this direction. 4

B. Theoretical concepts and techniques techniques, such as diagrammatic perturbation theory including sophisticated resummation schemes. But it also encompasses The development of theoretical tools needed to perform the non-perturbative approaches, which often capitalize on the flex- transition from micro- to macro-physics in driven open quantum ibility of the functional integral when it comes to picking the systems is still in progress, as a topic of current research. A rea- relevant degrees of freedom for a given problem. This is the son for the preliminary status of the theory lies in the fact that challenge of emergent phenomena, whose solution typically is two previously rather independent disciplines — quantum op- strongly scale dependent [86]. Familiar examples include an tics and many-body physics — need to be unified on a technical efficient description of emergent Cooper pair or molecular de- level. grees of freedom in interacting fermion systems, or vortices which conveniently parameterize the long-wavelength physics Quantum optical systems are well described microscopically of interacting bosons in two dimensions. The description of the in terms of Markovian quantum master equations, which treat change of physics with scale was given its mathematical foun- coherent Hamiltonian and driven-dissipative dynamics on equal dation in terms of the renormalization group already mentioned footing. To solve such equations both for their dynamics above, yet another tool developed and most clearly formulated in and their stationary states, powerful techniques have been de- a functional language. Finally, the functional integral based on vised. This comprises efficient exact numerical techniques for a single scalar quantity — the system’s action, which encodes small enough systems, such as the quantum trajectories ap- all the dynamics on the microscopic scale — is a convenient proach [48, 79, 80]. But it also includes analytical approaches framework when it comes to the classification of symmetries and such as perturbation theory for quantum master equations, e.g., associated conservation laws, and their use in devising approxi- in the frame of the Nakajima-Zwanzig projection operator tech- mation schemes respecting them. nique [81, 82], or mappings to P, W, or Q representations [83], casting the problem from a second quantized formulation into To put it short, while the driven open many-body systems partial differential equations. are well described by microscopic master equations, the tradi- A characteristic feature of traditional quantum optical sys- tional techniques of quantum optics cannot be used efficiently tems is the finite spacing of the few energy levels which play – at least not in the case where the generic complications of a role. When considering systems with a spatial continuum of many-body systems start to play a role. Conversely, their driven degrees of freedom instead, the energy levels become continu- open character makes it impossible to approach these problems ous. This does not mean that the microscopic modelling in terms in the framework of equilibrium many-body physics. This sit- of a quantum master equation is inappropriate: for the driven- uation calls for a merger of the disciplines of quantum optics dissipative terms in such an equation, the assumption of spa- and many-body physics on a technical level. On the numerical tially independent dissipative processes (such as atomic loss or side, progress has been made in one spatial dimension recently spontaneous emission) is still valid as long as the emitted wave- by combining the method of quantum trajectories with power- length of radiation is well below the spatial resolution at the ful density matrix renormalization group algorithms [87–90], scale where the microscopic model is defined. Indeed, in this see [48] for an excellent review on the topic. For more analyti- situation, destructive interference of radiation justifies the de- cal approaches, the Keldysh functional or path integral [91–96] scription of driven dissipation in terms of incoherent processes. is ideally suited (but see [97] for a recent systematic perturbative However, under these circumstances the smallness of a micro- approach to lattice Lindblad equations with extensions to sophis- scopic expansion parameter no longer guarantees the smallness ticated resummation schemes [98], and [99–101] for advanced of the associated perturbative correction. Here the reason is that variational techniques). Conceptually, the latter captures the in perturbation theory, one is summing over intermediate states most general situation in many-body physics — the dynamics with propagation amplitudes down to the longest wavelengths. of a density matrix under an arbitrary temporally local generator This can lead to infrared divergences in naive perturbation the- of dynamics. We refer to [102, 103] for an introduction to the ory — a circumstance that found its physical interpretation and Keldysh functional integral. In our context, it can be derived by technical remedy in equilibrium in terms of the renormalization a direct functional integral quantization of the Markovian quan- group [84, 85]. We emphasize that it is precisely this situation of tum master equation. This procedure results in a simple transla- long wavelength dominance which underlies much of the univer- tion table from the master equation to the key player in the as- sality, i.e., insensitivity to microscopic details, which is encoun- sociated Keldysh functional integral, the Markovian action. At tered when moving from the microscale to macroscopic observ- first sight, the complexity of the non-equilibrium Keldysh func- ables in many-particle systems. tional integral is increased by the characteristic “doubling” of The modern framework to understand many-particle prob- degrees of freedom compared to thermal equilibrium. However, lems in thermodynamic equilibrium is in terms of the functional it should be noted that it is precisely this feature which relates integral formulation of quantum field theory. The spectrum of the Keldysh functional integral much closer to real time (von its application covers a remarkable range of energy scales, from Neumann) evolution familiar from quantum mechanics. We will ultracold atomic gases to condensed matter systems with strong demonstrate this in Sec.IIA, making the Markovian action a correlations to quantum chromodynamics and the quantum the- rather intuitive object to work with. Furthermore, when prop- ory of gravity. It provides us with a well-developed toolbox of erly harnessing symmetries, the complexity of calculations can 5 often be made comparable to thermodynamic equilibrium. Most ory. The reason is the finite Markovian noise level that such importantly however, the powerful toolbox of quantum field the- systems exhibit generically, as explained in Sec.IIC. The prefix ory is opened in this way. It is thus possible to leverage the full “semi” refers to the fact that phase coherence may still persist power of sophisticated techniques from equilibrium field theory in such circumstances — the situation is comparable to a Bose- to driven open many-body systems. Einstein condensate at finite temperature. Recently however, sit- Relation to classical dynamical field theories — The quan- uations have been identified where the drastic simplifications of tum mechanical Keldysh formulation reduces to the so-called the semi-classical limit do not apply. In particular, this occurs Martin-Siggia-Rose (MSR) functional integral in the (semi-) in systems with dark states — pure quantum states which are classical limit, in turn equivalent to a stochastic Langevin equa- dynamical fixed points of driven-dissipative evolution [54, 110]. tion formulation [102]. This statement will be made precise and In these cases, classical dynamical field theories are inappropri- discussed in Sec.IIC. A large amount of work has been dedi- ate, which calls for the development of quantum dynamical field cated to this limit in the past. On the one hand, this concerns theories [28]. These developments are just in their beginnings. dynamical aspects of equilibrium statistical mechanics, and we In this review, we concentrate on systems composed of refer to the classic work by Hohenberg and Halperin [1] for an bosonic degrees of freedom. However, it is also possible to ad- overview. This work shows that, while the static universal crit- dress spin systems in terms of functional integrals [111], and ical behavior is determined by the symmetries and the dimen- simple models systems have been analyzed in this way, see [112] sionality of the problem, the dynamical critical behavior is sen- and Sec. III A. More sophisticated approaches to spin systems sitive to additional dynamical conservation laws. This leads to were elaborated in the context of multimode optical cavities a fine structure, defining dynamical universality classes which in [113], and systematically for various symmetries for lattice are denoted by models A–J [1]. These models also provide a systems in [114]. Fermi statistics is also conveniently imple- convenient framework to describe the statistics of work, sum- mented in the functional integral formulation. This is relevant, marized in Jarzynski’s work theorem [104, 105] and Crooks’ re- e.g., for driven open Fermi gases in optical cavities [21, 22, 115] lation [106]. On the other hand, non-equilibrium situations are or lattices [116], or dissipatively stabilized topological fermion captured as well. Here, we highlight in particular genuine non- matter [70–72]. equilibrium universality classes, which are not smoothly con- nected to the equilibrium models. Among them is the problem of reaction-diffusion models [107] including directed percola- C. Experimental platforms tion [4], which is relevant to certain chemical processes (for an implementation of this universality class with driven Rydberg The progress in controlling, manipulating, detecting and scal- gases, see [50]). Another key example is surface growth, de- ing up driven open quantum systems to many-body scenarios has scribed by the Kardar-Parisi-Zhang equation [73], giving rise been impressive over the last decade. Here we sketch the basic to a non-equilibrium universality class which is at the heart of physics of three representative platforms, and indicate the rele- driven phenomena such as the growth of bacterial colonies or vant microscopic theoretical models in the frame of the Marko- the spreading of fire. vian quantum master equation. In later sections, we will trans- In the same class of approaches ranges the so-called Doi-Peliti late this physics into the language of the Keldysh functional inte- functional integral [108, 109], which is a functional representa- gral. For each of these platforms, excellent reviews exist, which tion of classical master equations, and may be viewed as an MSR we refer to at the end of Sec.ID together with further litera- theory with a specific, highly non-linear appearance of the field ture on open systems. The purpose of this section is to give variables. It reduces to the conventional MSR form in a leading an overview only, and to put the respective platforms into their order Taylor expansion of the field non-linearities. A compre- overarching context as driven open quantum systems with many hensive overview of models, methods, and physical phenomena degrees of freedom. in the (semi-)classical limit is provided in [5]. We also note that the usual mean field theory, where corre- lation functions are factorized into products of field amplitudes 1. Cold atoms in an optical cavity, and microcavity arrays: and which is often used as an approximation to the quantum driven-dissipative spin-boson models master equation in the literature, corresponds to a further for- mal simplification of the semi-classical limit. Here the effects of Cavity quantum electrodynamics (cavity QED), with its fo- noise are neglected completely. Conversely, the semi-classical cus on strong light-matter interactions, is a growing field of re- limit represents a systematic extension of mean field theory, search, which has experienced several groundbreaking advances which includes the Markovian noise fluctuations. This level of in the past few years. Historically, these systems were devel- approximation is referred to as optimal path approximation in oped as few or single atom experiments, detecting the radiation the literature on MSR functional integrals [5, 102]. properties of atoms, which are strongly coupled to a quantized In many cases, even though the microscopic description is light field. The focus has recently been shifted towards loading in terms of a quantum master equation, at long distances the more and more atoms inside a cavity. Thereby, not only single Keldysh field theory reduces to a semi-classical MSR field the- particle dynamics in strong radiation fields can be probed, but 6 also collective, macroscopic phenomena, which are driven by for N → ∞, it features a , which spontaneously light-matter interactions. For an excellent review on this topic, breaks the Ising symmetry. Crossing the transition, the system covering important experimental and theoretical developments, enters a superradiant phase, characterized by finite expectation x see [10]. In the experiments, cold atoms are loaded inside an values hai , 0, hσi i , 0. This describes condensation of the optical or microwave cavity, for which the coherent interaction cavity photons, i.e., the formation of a macroscopically occu- between the atomic internal states and a single cavity mode dom- pied, coherent intra-cavity field, and a “ferromagnetic” ordering inates over dissipative processes [117]. The atoms absorb and of the atoms in the x-direction. emit cavity photons, thereby changing their internal states. Due Although the Dicke model is a standard model√ for cavity QED to this process, the spatial modulation of the intra-cavity light experiments in the ultra-strong coupling limit Ng > ωzω0, it field induces a coupling of the cavity photons to the atomic in- has been realized only very recently in cold atom experiments, ternal state as well as their motional degree of freedom [118]. where an entire BEC was placed inside an optical cavity [9]. It In this way, cavity photons mediate an effective atom-atom in- has been shown that this setup maps to a Dicke model, with a teraction, which leads to a back-action of a single atom on the “collective” spin degree of freedom [120]. Here, the detuning motion of other atoms inside the cavity. This cavity mediated of the pump laser was chosen such that the atoms effectively re- interaction is long-ranged in space and represents the source of main in the internal ground state, but acquire a characteristic re- collective effects in cold atomic clouds within cavity QED ex- coil momentum when scattering with a cavity photon. This scat- periments. One hallmark of collective dynamics in cavity QED tering creates a collective, motionally excited state, which re- has been the observation of self-organization of a Bose-Einstein places the role of an individual, internally excited atom. The ex- condensate in an optical cavity. This is accompanied by a Dicke perimental realization of a superradiance transition in the Dicke phase transition via breaking of a discrete Z2-symmetry of the model is usually inhibited, since the required coupling strength underlying model [9, 119–121]. by far exceeds the available value of the atomic dipole coupling. Although the dominant dynamics in these systems is coherent, However, for the BEC in the cavity, the energy scales of the ex- there are dissipative effects which cannot be discarded. These cited modes are much lower than the optical scale of the atomic are the loss of cavity photons due to imperfections in the cav- modes. In this way, the superradiance transition indeed became ity mirrors, or spontaneous emission processes of atoms, which experimentally accessible. This was inspired by a theoretical emit photons into transverse modes. These effects modify the proposal using two balanced Raman channels between different dynamics of the system on the longest time scales. Therefore, internal atomic states inside an optical cavity, which reduced the they become relevant for macroscopic phenomena, such as phase effective level splitting of the internal states to much lower en- transitions and collective dynamics. For instance, dissipative ef- ergy scales [129]. fects have been shown theoretically [122, 123] and experimen- In addition to the unitary dynamics represented by the Dicke tally [124] to modify the critical exponent of the Dicke transi- model, the cavity is subject to permanent photon loss due to im- tion compared to its zero temperature value. This illustrates that perfections in the cavity mirrors. For high finesse cavities, the for the analysis of collective phenomena in cavity QED experi- coupling of the intra cavity photons to the surrounding vacuum ments, the dissipative nature of the system has to be taken into radiation field is very weak, and the latter can be eliminated in a account properly [112, 125]. Born-Markov approximation [81]. This results in a Markovian Typically, for a cavity field which has a very narrow spectrum, quantum master equation for the system’s density matrix the atomic internal degrees of freedom can be reduced to two internal states, whose transitions are nearly resonant to the pho- ∂tρ = −i[HD, ρ] + Ldρ, (2) ton frequency. The operators acting on these two internal states can be represented by Pauli matrices, making them equivalent where ρ is the density matrix for the intra cavity system and Ld to a spin-1/2 degree of freedom. A very important model in the adds dissipative dynamics to the coherent evolution of the Dicke framework of cavity QED is the Dicke model [126–128], which model. For a vacuum radiation field, it is given by describes N atoms (i.e., two-level systems) coupled to a single   L † − 1 { † } quantized photon mode. This is expressed by the Hamiltonian dρ = κ LρL 2 L L, ρ . (3) (here and in the following we set ~ = 1) The Lindblad operator L acting on the density matrix describes N N pure photon loss (L = a) with an effective loss rate κ; the lat- † ωz X z g X x  †  HD = ω0a a + σi + √ σi a + a . (1) ter depends on system specific parameters, but is typically the 2 N i=1 i=1 smallest scale in the master equation (2)[10]. z Here, ω0 is the photon frequency, σ = |1ih1| − |0ih0| repre- In generic cavity experiments, there are also atomic sponta- sents the splitting of the two atomic levels with energy differ- neous emission processes. The atoms scatter a laser or cavity x ence ωz. σ = |0ih1| + |1ih0| describes the coherent excitation photon out of the cavity, and this represents a source of deco- and de-excitation of the atomic state proportional to the atom- herence. This process has been considered in Ref. [112], and photon coupling strength g. The Dicke model features a dis- leads to an effective decay rate of the atomic excited state and crete Z2 Ising symmetry: it is invariant under the transforma- therefore to a dephasing of the atoms, described by additional † x † x z tion (a , a, σi ) 7→ (−a , −a, −σi ). In the thermodynamic limit, Lindblad operators Li = σi [130]. However, these losses are 7 at least three orders of magnitude smaller than the cavity de- using multiple qubits [140]. An advantage of such schemes is cay rate and typically not considered [9, 125]. The basic model that the symmetries of the underlying dynamics are preserved for cavity QED with cold atoms is therefore represented by the or less severely corrupted in this way (cf. the discussion in Dicke model with dissipation, formulated in terms of the master Sec.IID). In other platforms, such as exciton-polariton systems equation (2). Its dynamics is discussed in Sec. III, including the (cf. the subsequent section), incoherent pumping is more natural Dicke superradiance transition. from the outset. The Dicke model Hamiltonian takes the form HD = Hc + Hs + All the systems discussed here represent genuine instances of Hcs, where Hc,s,cs represent the bosonic cavity, spin, and spin- driven open many-body systems. Besides the coherent drive, boson sectors, respectively. There are many directions to go be- they undergo dissipative processes, which have to be taken into yond the Dicke model with single collective spin, still keeping account for a proper understanding of their time evolution and the basic feature of coupling spin to boson (cavity photon) de- stationary states. A generic feature of these processes is their grees of freedom. One direction – relevant to future cold atom locality. For the effective spin degrees of freedom, the typical − experiments – is to consider multimode cavities instead of a sin- processes are qubit decay (local Lindblad operators Li = σi ) and z gle one. In particular, in conjunction with quenched disorder, dephasing (Li = σi ). For the bosonic component, local single- intriguing analogies to the physics of quantum glasses can be photon loss is dominant, Li = ai. established in this way [113, 131]. Here, the global coupling of Under various circumstances, such as a low population of all spins to a single mode gi ≡ g is replaced by random couplings the excited spin states, the latter degrees of freedom can be gi,`, where the index ` now refers to a collection of cavity modes. integrated out. In this limit, their physical effect is to gen- The many-body physics of such an open system is discussed in erate Kerr-type bosonic non-linearities, giving rise to driven Sec. III. open variants of the celebrated Bose-Hubbard model. These The basic building block of the Dicke model is the spin-boson models can even be brought into the correlation dominated † x term of the form Hsc ∼ (a + a )σ , i.e. a Rabi type non- regime [24, 132, 133, 141]. Oftentimes, these approximations on linearity that preserves the Z2 symmetry of Hc,s. In the con- the spin sector actually apply, and it is both useful and interesting text of circuit quantum electrodynamics, a natural many-body to study these effective low frequency bosonic theories instead generalization of Hamiltonians with a spin-boson interaction is of the full many-body spin-boson problems, see Refs. [142–146] to consider entire arrays of microcavities (instead of consider- for recent work in this direction. ing many modes within a single cavity). These cavities can be coupled to each other by single photon tunnelling processes between adjacent cavities, giving rise to Hubbard-type hopping † 2. Exciton-polariton systems: driven open interacting bosons terms ∼ Jai a j + h. c., where i, j now label the spatial index of the cavities. In cirquit QED, strong non-linearities can be gen- erated, e.g., by coupling to adjacent qubits made of Cooper pair Exciton-polaritons are an extremely versatile experimental boxes [132, 133]. This gives rise to many-body variants of the platform, which is documented by the richness of physical phe- Rabi model [134], whose phase diagrams have been studied re- nomena that have been studied in these systems both in the- cently [135–137]. Furthermore, for the implementation of lat- ory and experiment. For a comprehensive account of the sub- tice Dicke models with large collective spins, the use of hybrid ject, we refer to a number of excellent review articles [13, 147, quantum systems consisting of superconducting cavity arrays 148]. A Keldysh functional integral approach is discussed in coupled to -state spin ensembles have been proposed [138]. Refs. [149, 150], which provides both a microscopic deriva- Spontaneous collective coherence in driven-dissipative cavity ar- tion of an exciton-polariton model and a mean field analysis rays has been studied in [139]. including Gaussian fluctuations. At this point, we content our- In many physical situations (away from the ultra-strong cou- selves with a short introduction, with the aim of showing that pling limit), a form of the spin-cavity interaction alternative in a suitable parameter regime, exciton-polaritons very natu- to the Rabi term is more appropriate, with non-linear building rally provide a test-bed to study Bose condensation phenom- + † − block Hcs ∼ aσ + a σ . In fact, this form results naturally ena out of thermal equilibrium. Similar physics can also arise from the weak coupling rotating wave approximation of a driven in a variety of other systems, including condensates of pho- spin-cavity problem. For a single cavity mode and spin, the re- tons [151], [152], and potentially [153]. Re- sulting model is the Jaynes-Cummings model, which in contrast markably, even cold atoms could be brought to condense in a to the Dicke model possesses a continuous U(1) phase rotation non-equilibrium regime, where continuous loading of atoms bal- symmetry under a → eiθa, σ− → eiθσ−. ances three-body losses [154], or in atom laser setups [155–157]. Clearly, when such systems are driven coherently via a Hamil- A basic experimental setup for exciton-polaritons consists of † tonian Hd = Ω(a + a ) or suitable multimode generalizations a planar semiconductor microcavity embedding a quantum well thereof, both the U(1) and even the Z2 symmetries of the above (see Fig.1 (a)). This setting allows for a strong coupling of cav- models are broken explicitly. Coherent drive is usually the sim- ity light and matter in the quantum well, as originally proposed plest way to compensate for unavoidable losses due to cavity in [158]. The free dynamics of the elementary excitations of this leakage, although incoherent pumping schemes are conceivable system — i.e., of cavity photons and Wannier-Mott excitons — 8 is described by the quadratic Hamiltonian [13] (a) photon H0 = HC + HX + HC−X, (4) exciton e where the parts of the Hamiltonian involving only photons and excitons, respectively, take the same form, which is given by h (here the index α labels cavity photons, α = C, and excitons, α = X, respectively)2 Z dq X † Braggmirror QW Braggmirror Hα = ωα(q)a (q)aα,σ(q). (5) (2π)2 α,σ σ (b) upper ! † polariton Field operators aα,σ(q) and aα,σ(q) create or destroy a photon or exciton (note that both are bosonic excitations) with in-plane photon momentum q and polarization σ (there are two polarization excitation states of the exciton which are coupled to the cavity mode [13]). relaxation For simplicity, we neglect polarization effects leading to an ef- exciton fective spin-orbit coupling [13]. Due to the confinement in the transverse (z) direction, i.e., along the cavity axis, the motion of lower photons in this direction is quantized as qz,n = πn/lz, where n is polariton a positive integer, and lz is the length of the cavity. In writing emission the Hamiltonian (5), we are assuming that only the lowest trans- q verse mode is populated, which leads to a quadratic dispersion q 2 2 as a function of the in-plane momentum q = |q| = qx + qy: Figure 1. (a) Schematic of two Bragg mirrors forming a microcavity, in which a quantum well (QW) is embedded. In the regime of strong q q2 light-matter interaction, the cavity photon and the exciton hybridize and ω (q) = c q2 + q2 = ω0 + + O(q4). (6) C z,1 C 2m form new eigenmodes, which are called exciton-polaritons. (b) Energy C dispersion of the upper and lower polariton branches as a function of in- 0 plane momentum q. In the experimental scheme illustrated in this figure Here, c is the speed of sound, ωC = cqz,1, and the effective mass of the photon is given by m = q /c. Typically, the value of the (cf. Ref. [12]), the incident laser is tuned to highly excited states of the C z,1 quantum well. These undergo relaxation via emission of and photon mass is orders of magnitude smaller than the mass of the scattering from polaritons, and accumulate at the bottom of the lower exciton, so that the dispersion of the latter appears to be flat on polariton branch. In the course of the relaxation process, coherence is the scale of Fig.1 (b). quickly lost, and the effective pumping of lower polaritons is incoher- Upon absorption of a photon by the semiconductor, an exciton ent. is generated. This process (and the reverse process of the emis- sion of a photon upon radiative decay of an exciton) is described by non-radiative decay rate of excitons, it is appropriate to think of exciton-polaritons as the elementary excitations of the system. Z dq X   H = Ω a† (q)a (q) + H.c. , (7) In experiments, it is often sufficient to consider only lower C−X R 2 X,σ C,σ polaritons in a specific spin state, and to approximate the disper- (2π) σ sion as parabolic [13]. Interactions between exciton-polaritons where ΩR is the rate of the coherent interconversion of photons originate from various physical mechanisms, with a dominant into excitons and vice versa. The quadratic Hamiltonian (4) can contribution stemming from the screened Coulomb interactions be diagonalized by introducing new modes — the lower and between electrons and holes forming the excitons. Again, in the upper exciton-polaritons, ψLP,σ(q) and ψUP,σ(q) respectively, low-energy scattering regime, this leads to an effective contact which are linear combinations of photon and exciton modes. The interaction between lower polaritons. As a result, the low-energy dispersion of lower and upper polaritons is depicted in Fig.1 description of lower polaritons takes the form (in the following (b). In the regime of strong light-matter coupling, which is we drop the subscript indices in ψLP,σ)[13] reached when ΩR is larger than both the rate at which pho- Z " 2 ! # tons are lost from the cavity due to mirror imperfections and the † 0 ∇ † 2 2 HLP = dx ψ (x) ωLP − ψ(x) + ucψ (x) ψ(x) . (8) 2mLP While this Hamiltonian is quite generic and arises also, e.g., in 2 In Ref. [149, 150], a different model for excitons is used: they are assumed cold bosonic atoms in the absence of an external potential, the to be localized by disorder, and interactions are included by imposing a hard- peculiarity of exciton-polaritons is that they are excitations with core constraint. relatively short lifetime. In turn, this necessitates continuous re- 9 plenishment of energy in the form of laser driving in order to 3. Cold atoms in optical lattices: heating dynamics maintain a steady population. In Fig.1 (b), we consider the case in which the excitation laser is tuned to energies well above In recent years, experiments with cold atoms in optical lat- the lower polariton band. The thus created high-energy excita- tices have shown remarkable progress in the simulation of many- tions are deprived of their excess energy via phonon-polariton body model systems both in and out of equilibrium. A par- and stimulated polariton-polariton scattering. Eventually, they ticular strength of cold atom experiments is the unprecedented accumulate at the bottom of the lower polariton band. As a con- tuneability of model parameters, such as the local interaction sequence of multiple scattering processes, the coherence of the strength and the lattice hopping amplitude. This becomes possi- incident laser field is quickly lost, and the effective pumping of ble by, e.g., manipulation of the lattice laser and external mag- lower polaritons is incoherent. netic fields. It comes along with a very weak coupling of the A phenomenological model for the dynamics of the lower po- system to the environment, such that the dynamics can often be lariton field, which accounts for both the coherent dynamics gen- seen as isolated on relevant time scales for typical measurements erated by the Hamiltonian (8) and the driven-dissipative one de- of static, equilibrium correlations. However, more and more ex- scribed above, was introduced in Ref. [159]. It involves a dissi- periments start to investigate the realm of non-equilibrium phe- pative Gross-Pitaevskii equation for the lower polariton field that nomena with cold atoms, e.g., by letting systems prepared in a is coupled to a rate equation for the reservoir of high-energy ex- non-equilibrium initial condition relax in time towards a steady citations. However, for the study of universal long-wavelength state [37, 39, 163–165]. With these experiments, time scales are behavior in Sec.IV, this degree of microscopic modeling is ac- reached, for which the dissipative coupling to the environment tually not required: indeed, any (possibly simpler) model that becomes visible in experimental observables. Such dissipation possesses the relevant symmetries (see the discussion at the be- may even hinder the system from relaxation towards a well de- ginning of Sec.IV) will yield the same universal physics. Such fined steady state. a model can be obtained by describing incoherent pumping and A relevant example of a dissipative coupling is decoherence losses of lower polaritons by means of a Markovian master equa- of an atomic cloud induced by spontaneous emission of atoms tion:3 in the lattice [166, 167]. In this way, the many-body system is heated up, and therefore driven away from the low entropy ∂tρ = −i[HLP, ρ] + Ldρ, (9) state in which it was prepared initially. A detailed discussion of the microscopic physics and its long time dynamics can be where Ldρ encodes incoherent single-particle pumping and found in [45–47], see also the review article [48]. For bosonic losses, as well as two-body losses: atoms in optical lattices, the coherent dynamics is described by the Bose-Hubbard Hamiltonian [168, 169] Z  † 2  X † U X Ldρ = dx γpD[ψ(x) ]ρ + γlD[ψ(x)]ρ + 2udD[ψ(x) ]ρ , H = −J b b + n (n − 1), (12) BH l m 2 l l (10) hl,mi l where which models bosonic atoms in terms of the creation and annihi- † lation operators [b , bm] = δlm in the lowest band of a lattice with 1 l D[L]ρ = LρL† − {L†L, ρ} (11) site indices l, m. The atoms hop between neighboring lattice sites 2 with an amplitude J, and experience an on-site repulsion U. The lattice potential V(x), which leads to the second quantized form reflects the Lindblad form, and γp, γl, 2ud are the rates of single- of the Bose-Hubbard model, is created by the superposition of particle pumping, single-particle loss, and two-body loss, re- counter-propagating laser beams in each spatial dimension. The spectively. The inclusion of the non-linear loss term ensures coherent laser field couples two internal atomic states via stimu- saturation of the pumping. An analogous mechanism is im- lated absorption and emission, which leads to the single particle plemented in the above-mentioned Gross-Pitaevskii description. Hamiltonian More precisely, in the spirit of universality, the above quantum 2 pˆ ∆ z x Ω(xˆ) master equation (9) and the above mentioned phenomenologi- Hatom = − σ − σ , (13) cal model reduce to precisely the same low frequency model for 2m 2 2 bosonic degrees of freedom upon taking the semiclassical limit where pˆ is the atomic momentum operator, ∆ is the detuning of in the Keldysh path integral associated to Eq. (9) (see Sec.IIC the laser from the atomic transition frequency, and Ω(xˆ) is the for its implementation), and integrating out the upper polariton laser field at the atomic position. For large detuning, the excited reservoir in the phenomenological model. state of the atom can be traced out, which leads to the lattice Hamiltonian pˆ 2 |Ω(xˆ)|2 Heff = + . (14) atom 2m 4∆ 3 While this approach captures the universal behavior, we note that non- This describes a lattice potential for the ground state atoms Markovian effects can be of key importance for other properties [160–162]. generated by the spatially modulated AC Stark shift. Adding 10

cial example for universality in out-of-equilibrium dynamics, which can be probed by cold atom experiments.

D. Outline and scope of this review

The remainder of this review is split into two parts. Figure 2. Illustration of the coherent (a) and the incoherent (b) con- Part1 develops the theoretical framework for the e fficient de- tribution to the dynamics stemming from the coupling of the atoms to scription of driven open many-body quantum systems. In Sec.II, a laser with amplitude |Ω| and large detuning ∆. (a) Via the AC Stark we begin with a direct derivation of the open system Keldysh effect, stimulated absorption and emission lead to a coherent periodic functional integral from the many-body quantum master equa- |Ω|2 potential with amplitude 4∆ . (b) Stimulated absorption and subsequent tion (Sec.IIA). We then discuss in Sec. II B 1 in detail a simple spontaneous emission lead to effective decoherence of the atomic state example: the damped and driven optical cavity. This allows the |Ω|2 with rate Γ 4∆ , where Γ is the microscopic spontaneous emission rate. reader to familiarize with the functional formalism. In particu- lar, the key players in terms of observables — correlation and re- sponse functions — are described. We also point out a number of an atomic interaction potential and expanding both the single exact structural properties of Keldysh field theories, which hold particle Hamiltonian (14) and the interaction in terms of Wan- beyond the specific example. Another example is introduced in nier states, the leading order Hamiltonian is the Bose-Hubbard Sec. II B 2: there we discuss the mean field theory of conden- model (12)[169]. sation in a bosonic many-body system with particle losses and In the semi-classical treatment of the atom-laser interaction, pumping. The semi-classical limit of this model and its valid- spontaneous emission events are neglected, as their probabil- ity are the content of Sec.IIC. This is followed by a discussion ity is typically very small (see below for a more precise state- of symmetries and conservation laws in the Keldysh formalism ment). They can be taken into account on the basis of optical in Sec.IID. In particular, we point out a symmetry that allows Bloch equations [166], which leads, after elimination of the ex- one to distinguish equilibrium from non-equilibrium conditions. cited atomic state, to an additional, driven-dissipative term in Finally, an advanced field theoretical tool — the open system the atomic dynamics. It describes the decoherence of the atomic functional renormalization group — is introduced in Sec.IIE. state due to spontaneous emission, i.e., position dependent ran- Part2 harnesses this formalism to generate an understand- dom light scattering. The leading order contribution to the dy- ing of the physics in different experimental platforms. We be- namics in the basis of lowest band Wannier states is captured by gin in Sec. III with simple but paradigmatic spin models with the master equation discrete Ising Z2 symmetry in driven non-equilibrium station- X ary states. In particular, we discuss the physics of the driven ∂tρ = −i[HBH, ρ] − γ [nl, [nl, ρ]], (15) open Dicke model in Sec. III A. This is followed by an extended l variant of the latter in the presence of disorder and a multi-

† mode cavity, which hosts an interesting spin and photon glass where nl = bl bl is the local atomic density, and ρ is the many- phase in Sec. III B. Sec.IV is devoted to the non-equilibrium 2 |Ω| U body density matrix. For a red detuned laser the rate γ = Γ 4∆2 is stationary states of bosons with a characteristic (1) phase ro- proportional to the microscopic spontaneous emission rate Γ and tation symmetry: the driven-dissipative condensates introduced the laser amplitude |Ω|. Note the suppression of the scale γ by in Sec. II B 2. After some additional technical developments re- a factor Γ/∆  1 for large detuning, compared to the strength lating to U(1) symmetry in Sec.IVA, we discuss critical be- eff of Hatom. The coherent and incoherent contribution of the atom- havior at the Bose condensation transition in three dimensions. laser coupling to the dynamics is illustrated in Fig.2. In particular, we show the decrease of a parameter quantifying The dissipative term in the master equation (15) leads to an non-equilibrium strength in this case (Sec.IVB). The opposite energy increase hHBHi(t) ∼ t linear in time, and thus to heat- behavior is observed in two (Sec.IVC) and one (Sec.IVD) ing. Furthermore, it introduces decoherence in the number state dimensions. Finally, leaving the realm of stationary states, in basis, i.e., it projects the local density matrix on its diagonal in Sec.V we discuss an application of the Keldysh formalism to Fock space and leads to a decrease of the coherences in time [45– the time evolution of open bosonic systems in one dimension, 48]. Starting from a low entropy state at t = 0, the heating leads which undergo number conserving heating processes. We set up to a crossover from coherence dominated dynamics at short and the model in Sec.VA, derive the kinetic equation for the dis- intermediate times [43] to a decoherence dominated dynamics at tribution function in Sec.VB, and discuss the relevant approxi- long times [45–47]. In one dimension, both regimes have been mations and physical results in Secs.VC andVD, respectively. analyzed extensively both numerically (with a focus on deco- Conclusions are drawn in Sec.VI. Finally, brief introductions herence dominated dynamics) as well as analytically, and dis- to functional differentiation and Gaussian functional integration play several aspects of non-equilibrium universality, see Sec.V. are given in AppendicesA andB. Therefore, heating in interacting lattice systems represents a cru- Reflecting the bipartition of this review, the scope of it is 11 twofold. On the one hand, it develops the Keldysh functional in- ouvillian (sometimes this term is reserved for the second con- tegral approach to driven open quantum systems “from scratch”, tribution on the RHS of Eq. (16) alone). There are two contri- in a systematic and coherent way. It starts from the Marko- butions to the Liouvillian: first, the commutator term, which is vian quantum master equation representation of driven dissipa- familiar from the von Neumann equation, describes the coher- tive quantum dynamics [81, 82, 170], and introduces an equiv- ent dynamics generated by a system Hamiltonian H; the second alent Keldysh functional integral representation. Direct contact part, which we will refer to as the dissipator D 4, describes the is made to the language and typical observables of quantum op- dissipative dynamics resulting from the interaction of the sys- tics. It does not require prior knowledge of quantum field the- tem with an environment, or “bath.” It is defined in terms of a ory, and we hope that it will find the interest of — and be use- set of so-called Lindblad operators (or quantum jump operators) ful for — researchers working on quantum optical systems with Lα, which model the coupling to that bath. The dissipator has a many degrees of freedom. On the other hand, this review doc- characteristic Lindblad form [174, 175]: it contains an anticom- uments some recent theoretical progresses made in this concep- mutator term which describes dissipation; in order to conserve tual framework in a more pedagogical way than the original lit- the norm tr(ρ) of the system density matrix, this term must be ac- erature. We believe that this not only exposes some interesting companied by fluctuations. The corresponding term, where the physics, but also demonstrates the power and flexibility of the Lindblad operators act from both sides onto the density matrix, Keldysh approach to open quantum systems. is referred to as recycling or quantum jump term. Dissipation This work is complemetary to excellent reviews putting more occurs at rates γα which are non-negative, so that the density emphasis on the specific experimental platforms partially men- matrix evolution is completely positive, i.e., the eigenvalues of tioned above: The physics of driven Bose-Einstein condensates ρ remain positive under the combined dynamics generated by H in optical cavities is reviewed in [10]. A general overview of and D [176]. If the index α is the site index in an optical lattice driven ultracold atomic systems, specifically in optical lattices, or in a microcavity array, or even a continuous position label (in is provided in [48], and systems with engineered dissipation are which case the sum is replaced by an integral), in a translation described in [57]. Detailed accounts for exciton-polariton sys- invariant situation there is just a single scale γα = γ for all α tems are given in [13, 147, 148]; specifically, we refer to the associated to the dissipator. review [149] working in the Keldysh formalism. The physics of The quantum master equation (16) provides an accurate de- microcavity arrays is discussed in [14, 24, 134, 141], and trapped scription of a system-bath setting with a strong separation of ions are treated in [17, 171]. We also refer to recent reviews scales. This is generically the case in quantum optical systems, on additional upcoming platforms of driven open quantum sys- which are strongly driven by external classical fields. More pre- tems, such as Rydberg atoms [172] and opto-nanomechanical cisely, there must be a large energy scale in the bath (as com- settings [173]. For a recent exposition of the physics of quan- pared to the system-bath coupling), which justifies to integrate tum master equations and to efficient numerical techniques for out the bath in second-order time-dependent perturbation the- their solution, see [48]. ory. If in addition the bath has a broad bandwidth, the combined Born-Markov and rotating-wave approximations are appropri- ate, resulting in Eq. (16). A concrete example for such a set- Part 1 ting is provided by a laser-driven atom undergoing spontaneous emission. Generic condensed matter systems do not display such Theoretical background a scale separation, and a description in terms of a master equa- tion of the type (16) is not justified.5 However, the systems dis- cussed in the introduction, belong to the class of systems which II. KELDYSH FUNCTIONAL INTEGRAL FOR DRIVEN OPEN SYSTEMS permit a description by Eq. (16). We refer to them as driven open many-body quantum systems. Due to the external drive these systems are out of thermody- In this part, we will be mainly concerned with a Keldysh field namic equilibrium. This statement will be made more precise theoretical reformulation of the stationary state of Markovian in Sec. II D 1 in terms of the absence of a dynamical symme- many-body quantum master equations. As we also demonstrate, try which characterizes any system evolving in thermodynamic this opens up the powerful toolbox of modern quantum field the- equilibrium, and which is manifestly violated in dynamics de- ory for the understanding of such systems. scribed by Eq. (16). Its absence reflects the lack of energy con- The quantum master equation, examples of which we have servation and the epxlicit breaking of detailed balance. As stated already encountered in Sec.IC, describes the time evolution of a reduced system density matrix ρ and reads [81, 82]

! 4 We use the term “dissipation” here for all kinds of environmental influences X † 1 † ∂ ρ = Lρ = −i[H, ρ] + γ L ρL − {L L , ρ} , (16) on the system which can be captured in Lindblad form, including effects of t α α α 2 α α α decay and of dephasing/decoherence. 5 Davies’ prescription [177] allows one to also describe equilibrium systems where the operator L acts on the density matrix ρ “from both in terms of operatorial master equations, however with collective Lindblad sides” and is often referred to as Liouville superoperator or Li- operators ensuring detailed balance conditions (cf. also [82]). 12 in the introduction, the main goal of this review is to point out terms of a state vector evolution, even in the case of purely co- the macroscopic, observable consequences of this microscopic herent Hamiltonian dynamics.6 Instead, it is necessary to study violation of equilibrium conditions. the evolution of a state matrix, which transforms according to the integral form of the von Neumann equation in the second line of Eq. (17). Therefore, we have to apply the Trotter formula A. From the quantum master equation to the Keldysh functional and coherent state insertions on both sides of the density matrix. integral This leads to the doubling of degrees of freedom, characteris- tic of the Keldysh functional integral. Moreover, time evolution In this section, starting from a many-body quantum master can now be interpreted as occurring along two branches, which equation Eq. (16) in the operator language of second quantiza- we denote as the forward and a backward branches, respectively tion, we derive an equivalent Keldysh functional integral. We (cf. Fig.3 b)). Indeed, this is an intuitive and natural feature of focus on stationary states, and we discuss how to extract dynam- evolving matrices instead of vectors. ics from this framework in Sec.V. Our derivation of the Keldysh So far, we have concentrated on closed systems which evolve functional integral applies to a theory of bosonic degrees of free- according to purely Hamiltonian dynamics. However, we can dom. If spin systems are to be considered, it is useful to first per- allow for a more general generator of dynamics and still proceed form the typical approximations mapping them to bosonic fields, along the two-branch strategy. The most general (time local) and then proceed along the construction below (see Sec. III and evolution of a density matrix is given by the quantum master Refs. [112, 114]). Clearly, this amounts to an approximate treat- equation (16). Its formal solution reads ment of the spin degrees of freedom; for an exact (equilibrium) (t−t0)L N functional integral representation for spin dynamics, taking into ρ(t) = e ρ(t0) ≡ lim (1 + δtL) ρ(t0). (19) N→∞ account the full non-linear structure of their commutation rela- tions, we refer to Ref. [178]. For fermionic problems, the con- The last equality gives a meaning to the formal solution in terms struction is analogous to the bosonic case presented here. How- of the Trotter decomposition: at each infinitesimal time step, the ever, a few signs have to be adjusted to account for the fermionic exponential can be expanded to first order, such that the action anticommutation relations [102, 178]. of the Liouvillian superoperator is just given by the RHS of the The basic idea of the Keldysh functional integral can be de- quantum master equation (16); At finite times, the evolved state veloped in simple terms by considering the Schrodinger¨ vs. the is given by the concatenation of the infinitesimal Trotter steps. von Neumann equation, If we restrict ourselves to the stationary state of the system,7 but want to evaluate correlations at arbitrary time differences, i∂ |ψ(t)i = H|ψ(t)i ⇒ |ψ(t)i = U(t, t )|ψ(t )i, t 0 0 we should extend the time branch from an initial time in the dis- † ∂tρ(t) = −i[H, ρ(t)] ⇒ ρ(t) = U(t, t0)ρ(t0)U (t, t0), tant past t0 → −∞ to the distant future, t f → +∞. In analogy to (17) thermodynamics, we are then interested in the so-called Keldysh partition function Z = tr ρ(t) = 1. The trace operation contracts −iH(t−t0) where U(t, t0) = e is the unitary time evolution opera- the indices of the time evolution operator as depicted in Fig.3 tor. In the first case, a real time path integral can be constructed c), giving rise to the closed time path or Keldysh contour. Con- along the lines of Feynman’s original path integral formulation servation of probability in the quantum mechanical system is of quantum mechanics [179]. To this end, a Trotter decomposi- reflected in the time-independent normalization of the partition tion of the evolution operator function. In order to extract physical information, again in anal-

−iH(t−t0) N ogy to statistical mechanics, below we introduce sources in the e = lim (1 − iδtH) , (18) N→∞ partition function. This allows us to compute the correlation and response functions of the system by taking suitable variational with δ = t−t0 , is performed. Subsequently, in between the t N derivatives with respect to the sources. factors of the Trotter decomposition, completeness relations in After this qualitative discussion, let us now proceed with the terms of coherent states are inserted in order to make the (nor- explicit construction of the Keldysh functional integral for open mal ordered) Hamilton operator a functional of classical field systems, starting from the master equation (16). As mentioned variables. This is illustrated in Fig.3 a), and we will perform above, the Keldysh functional integral is an unraveling of Li- these steps explicitly and in more detail below in the context ouvillian dynamics in the basis of coherent states, and we first of open many-body systems. Crucially, we only need one set of field variables representing coherent Hamiltonian dynamics, which corresponds to the forward evolution of the Schrodinger¨ state vector. It is also clear that — noting the formal analogy of 6 Of course, the evolution of an M × M matrix, where M is the dimension of the the operators e−iH(t−t0) and e−βH — this construction can be lever- Hilbert space, can be formally recast into the evolution of a vector of length aged over to the case of thermal equilibrium, where the “Trotter- M × M. While such a strategy is often pursued in numerical approaches [48], ization” is done in imaginary instead of real time. it does in general not allow for a physical interpretation. In contrast to these special cases, the von Neumann equation 7 We assume that it exists. We thus exclude scenarios with dynamical limit for general mixed state density matrices cannot be rewritten in cycles, for simplicity. 13

a) |Ωi represents the vacuum in Fock space. (Note that accord- ing to this definition, which is usually adopted in the discus- (t0) t t t0 | i sion of field integrals [180], the state |ψi is not normalized.) U A key property of coherent states is that they are eigenstates of the annihilation operator, i.e., b |ψi = ψ |ψi, with the com- b) plex eigenvalue ψ. Clearly, this implies the conjugate relation ⇢(t0) hψ| b† = hψ| ψ∗.8 The overlap of two non-normalized coherent t t ∗ U U † t states is given by hψ|φi = eψ φ, and the completeness relation R ∗ ∗ reads 1 = dψdψ e−ψ ψ |ψi hψ|. c) + contour π The starting point of the derivation is Eq. (19), and we focus ⇢(t ) ⇢(t ) f 0 first on a single time step, as in the usual derivation of the co- tf =+ - contour t = 1 0 1 herent state functional integral [180]. That is, we decompose the time evolution from t0 to t f into a sequence of small steps of du- Figure 3. Idea of the Keldysh functional integral. a) According to the ration δt = (t f − t0)/N, and denote the density matrix after the Schrodinger¨ equation, the time evolution of a pure state vector is de- n-th step, i.e., at the time tn = t0 + δtn, by ρn = ρ(tn). We then −iH(t−t0) scribed by the unitary operator U(t, t0) = e . In the Feynman have functional integral construction, the time evolution is chopped into in- finitesimal steps of length δt, and completeness relations in terms of ρ = eδtLρ = (1 + δ L) ρ + O(δ2). (20) coherent states are inserted in between consecutive time steps. This n+1 n t n t As anticipated above, we proceed to represent the density matrix insertion is signalled by the red arrows. b) In contrast, if the state is mixed, a density matrix must be evolved, and thus two time branches in the basis of coherent states. For instance, ρn at the time tn can are needed. As explained in the text, the dynamics need not necessarily be written as be restricted to unitary evolution. The most general time-local (Marko- Z ∗ ∗ vian) dynamics is generated by a Liouville operator in Lindblad form. dψ+,ndψ+,n dψ−,ndψ−,n − ∗ − ∗ ρ = e ψ+,nψ+,n ψ−,nψ−,n c) For the analysis of the stationary state, we are interested in the real n π π time analog of a partition function Z = tr ρ(t f ), starting from t0 = −∞ × hψ |ρ |ψ− i |ψ i hψ− | . (21) and running until t f = +∞. The trace operation connects the two time +,n n ,n +,n ,n branches, giving rise to the closed Keldysh contour. As a next step, we would like to express the matrix element hψ+,n+1|ρn+1|ψ−,n+1i, which appears in the coherent state repre- collect a few important properties of those. Coherent states are sentation of ρn+1, in terms of the corresponding matrix element defined as (for simplicity, in the present discussion we restrict at the previous time step tn. Inserting Eq. (21) in Eq. (20), we ourselves to a single bosonic mode b) |ψi = exp(ψb†) |Ωi, where find that this requires us to evaluate the “supermatrixelement”

 hψ+,n+1|L(|ψ+,ni hψ−,n|)|ψ−,n+1i = −i hψ+,n+1|H|ψ+,ni hψ−,n|ψ−,n+1i − hψ+,n+1|ψ+,ni hψ−,n|H|ψ−,n+1i " # X 1   + γ hψ |L |ψ i hψ |L† |ψ i − hψ |L† L |ψ i hψ |ψ i + hψ |ψ i hψ |L† L |ψ i . (22) α +,n+1 α +,n −,n α −,n+1 2 +,n+1 α α +,n −,n −,n+1 +,n+1 +,n −,n α α −,n+1 α Without loss of generality, we assume that the Hamiltonian is normal ordered. Then, a matrix element hψ|H|φi of the Hamiltonian between coherent states can be obtained simply by replacing the creation operators by ψ∗ and the annihilation operators by φ. The same is true for matrix elements of L, L†, and L†L, after performing the commutations which are necessary to bring these operators to the form of sums of normal ordered expressions (see Ref. [181] for a detailed discussion of subtleties related to normal ordering). Then we obtain by re-exponentiation

Z ∗ ∗   dψ+,ndψ+,n dψ−,ndψ−,n iδ −ψ i∂ ψ∗ −ψ∗ i∂ ψ −iL(ψ∗ ,ψ ,ψ∗ ,ψ ) 2 hψ |ρ |ψ i = e t +,n t +,n −,n t −,n +,n+1 +,n −,n+1 −,n hψ |ρ |ψ i + O(δ ), (23) +,n+1 n+1 −,n+1 π π +,n n −,n t

8 Note that the creation operator cannot have eigenstates due to the fact that there is a minimal occupation number of a bosonic state. In particular, the 14

Keldysh partition function for stationary states — When we where we are using the shorthand suggestive notation ∂tψ±,n = are interested in a stationary state, but would like to obtain infor- (ψ±,n+1 − ψ±,n)/δt. The time derivative terms emerge from the mation on temporal correlation functions at arbitrarily long time overlap of neighboring coherent states at time steps n and n + 1, differences, it is useful to perform the limit t0 → −∞, t f → +∞ combined with the weight factor in the completeness relation for in Eq. (25). In an open system coupled to several external baths, ∗ ∗ step n; the quantity L(ψ+,n+1, ψ+,n, ψ−,n+1, ψ−,n) is the superma- it is typically a useful assumption that the initial state in the in- trixelement in Eq. (22), divided by the above-mentioned over- finite past does not affect the stationary state — in other words, laps: there is a complete loss of memory of the initial state. Under this physical assumption, we can ignore the boundary term, i.e., the hψ+,n+1|L(|ψ+,ni hψ−,n|)|ψ−,n+1i ∗ ∗ matrix element hψ+(t0)|ρ(t0)|ψ−(t0)i of the initial density matrix L(ψ+,n+1, ψ+,n, ψ−,n+1, ψ−,n) = . hψ+,n+1|ψ+,ni hψ−,n|ψ−,n+1i in Eq. (25), and obtain for the final expression of the Keldysh (24) partition function By iteration of Eq. (23), the density matrix can be evolved from Z ρ(t0) at t0 to ρ(t f ) at t f = tN . This leads in the limit N → ∞ (and ∗ ∗ iS Z = D[ψ+, ψ+, ψ−, ψ−] e = 1, (29) hence δt → 0) to Z ∞ ∗ ∗ ∗ ∗  (t f −t0)L − − L Zt ,t = tr ρ(t f ) = tr e ρ(t0) S = dt ψ+i∂tψ+ ψ−i∂tψ− i (ψ+, ψ+, ψ−, ψ−) . (30) f 0 −∞ Z (25) ∗ ∗ iS = D[ψ+, ψ+, ψ−, ψ−] e hψ+(t0)|ρ(t0)|ψ−(t0)i , This setup allows us to study stationary states far away from thermodynamic equilibrium as realized in the systems intro- where the integration measure is given by duced in Sec.I, using the advanced toolbox of quantum field theory. For the discussion of the time evolution of the system’s N ∗ ∗ initial state, the typical strategy in practice is not to start directly ∗ ∗ Y dψ+,ndψ+,n dψ−,ndψ−,n D[ψ+, ψ+, ψ−, ψ−] = lim , (26) from Eq. (25) — strictly speaking, this would necessitate knowl- N→∞ π π n=0 edge of the entire density matrix of the system, which in a gen- uine many-body context is not available. Rather, the Keldysh and the Keldysh action reads functional integral is used to derive equations of motion for a

Z t f given set of correlation functions. The initial values of the cor- ∗ ∗ ∗ ∗  S = dt ψ+i∂tψ+ − ψ−i∂tψ− − iL(ψ+, ψ+, ψ−, ψ−) . (27) relation functions have to be taken from the physical situation t0 under consideration. For interacting theories, the set of corre- lations functions typically corresponds to an infinite hierarchy. The coherent state representation of L in the exponent in The possibility of truncating this hierarchy to a closed subset Eq. (23) comes with a prefactor δ , so that to leading order for t usually involves approximations, which have to be justified from δ → 0 it is consistent to ignore the difference stemming from t case to case. the bra vector at n + 1 and the ket vector at n in Eq. (24). As- The Keldysh partition function is normalized to 1 by con- suming all operators are normally ordered in the sense discussed struction. As anticipated above, correlation functions can be ob- above, we obtain ∗  tained by introducing source terms J± = j±, j± that couple to ∗  ∗ ∗ the fields Ψ± = ψ±, ψ± (here and in the following we denote L(ψ+, ψ+, ψ−, ψ−) = −i (H+ − H−) spinors of a field and its complex conjugate by capital letters), " # X 1   ∗ ∗ ∗ Z R   + γα Lα,+Lα,− − Lα,+Lα,+ + Lα,−Lα,− , (28) iS +i J† Ψ −J† Ψ 2 t,x + + − − α Z[J+, J−] = D[Ψ+, Ψ−] e  R   (31) ∗ i J† Ψ −J† Ψ where H± = H(ψ±, ψ±) contains fields on the ± contour only, = e t,x + + − − , and the same is true for Lα,±. We clearly recognize the Lindblad superoperator structure of Eq. (16): operators acting on the den- where weR abbreviateR R (switching now to a spatial continuum of sity matrix from the left (right) reside on the forward, + (back- fields) = dt ddx, the average is taken with respect to the ward, -) contour. This gives a simple and direct translation ta- t,x action S , and we have the normalization Z[J = 0, J = 0] = 1 ble from the bosonic quantum master equation to the Markovian + − in the absence of sources. Physically, sources can be realized, Keldysh action (28), with the crucial caveat of normal ordering e.g., by coherent external fields such as lasers; this will be made to be taken into account before performing the translation. more concrete in the following section. The source terms can be thought of as shifts of the original Hamiltonian operator, justify- ing the von Neumann structure indicated above. Keldysh rotation — With these preparations, arbitrary correla- coherent states are not eigenstates. We rather have the relations b† |ψi = tion functions can be computed by taking variational derivatives ∗ (∂/∂ψ) |ψi , hψ| b = (∂/∂ψ ) hψ|. with respect to the sources. However, while the representation in 15 terms of fields residing on the forward and backward branches discussed in the subsequent subsection by means of a concrete allows for a direct contact to the second quantized operator for- example. malism, it is not ideally suited for practical calculations. In fact, Keldysh effective action — At this point we need one last the above description contains a redundancy which is related to technical ingredient: the effective action, which is an alternative the conservation of probability (this statement and the origin of way of encoding the correlation and response information of a the redundancy is detailed below in Sec. II B 1). This can be non-equilibrium field theory [182, 183]. We first introduce the avoided by performing the so-called Keldysh rotation, a unitary new generating functional transformation in the contour index or Keldysh space according to W[Jc, Jq] = −i ln Z[Jc, Jq]. (35) 1 1 Differentiation of the functional W generates the hierarchy φc = √ (ψ+ + ψ−) , φq = √ (ψ+ − ψ−) , (32) 2 2 of so-called connected field averages, in which expectation values of lower order averages are subtracted; for example, and analogously for the source terms. The index c (q) stands for ∗ 0 0 ∗ 0 0 ∗ 0 0 hφc(t, x)φc(t , x )i = hφc(t, x)φc(t , x )id − hφc(t, x)idhφc(t , x )id “classical” (“quantum”) fields, respectively. This terminology describes the density of particles which are not condensed. In a signals that the symmetric combination of fields can acquire a compact notation, where W(2) denotes the second variation as in (classical) field expectation value, while the antisymmetric one Eq. (34), cannot. In terms of classical and quantum fields, the Keldysh partition function takes the form W(2)(t − t0, x − x0) Jc=Jq=0 Z R  † †  ! iS +i J Φ +J Φ K 0 0 R 0 0 Z[J , J ] = [Φ , Φ ]e t,x q c c q G (t − t , x − x ) G (t − t , x − x ) c q D c q = − A 0 0 . (36) (33) G (t − t , x − x ) 0  R  † †  i Jq Φc+Jc Φq = e t,x ; The effective action Γ is obtained from the generating functional W by a change of active variables. More precisely, the active note in particular the coupling of the classical field Φ = φ , φ∗ c c c variables of the functional W, which are the external sources Jc,q,  ∗ to the quantum source Jq = jq, jq , and vice versa. Apart from i.e., W = W[Jc, Jq], are replaced by the field expectation values ¯ ¯ ∗ ¯ ¯ removing the redundancy mentioned above, a further key ad- Φν = φν, φν where ν = c, q, i.e., Γ = Γ[Φc, Φq]. (We remind vantage of this choice of basis is that taking variational deriva- the reader that capital letters denote spinors of fields and their tives with respect to the sources produces the two basic types complex conjugates.) The field expectation values are defined of observables in many-body systems: correlation and response as (the notation indicates that these expectation values are taken functions. Of particular importance is the single particle Green’s in the presence of the sources taking non-zero values) function, which has the following matrix structure in Keldysh ∗ ∗  ¯ 0 space (for a brief introduction to functional differentiation see Φν = hΦνi|Jc,Jq = δW/δJν0 = δW/δ jν0 , δW/δ jν , (37) AppendixA; note that here we are talking functional deriva- 0 tives with respect to the components of the spinors of sources where ν = q for ν = c and vice versa. Switching to these new J = j , j∗ where ν = c, q), variables is accomplished by means of a Legendre transform fa- ν ν ν miliar from classical mechanics, ! hφ t, x φ∗ t0, x0 i hφ t, x φ∗ t0, x0 i Z c( ) c( ) d c( ) q( ) d  † †  ∗ 0 0 ∗ 0 0 Γ[Φ¯ , Φ¯ ] = W[J , J ] − J Φ¯ + J Φ¯ . (38) hφq(t, x)φ (t , x )id hφq(t, x)φ (t , x )id c q c q c q q c c q t,x  δ2Z δ2Z   ∗ 0 0 ∗ 0 0   δ jq(t,x)δ jq(t ,x ) δ jq(t,x)δ jc(t ,x )  We now proceed to show that the difference between the action = −  2 2   δ Z δ Z  δ j∗(t,x)δ j (t0,x0) δ j∗(t,x)δ j (t0,x0) S and the effective action Γ lies in the inclusion of both statis- c q c c Jc=Jq=0 (34) ! tical and quantum fluctuations in the latter. To this end, instead GK(t − t0, x − x0) GR(t − t0, x − x0) i d d . of working with Eq. (38), we represent Γ as a functional inte- = A − 0 − 0 Gd (t t , x x ) 0 gral [182], with the result

In the last equality, in addition to stationarity (time translation in- Z ¯ ¯ δΓ δΓ iS [Φc+δΦc,Φq+δΦq]−i T δΦq−i T δΦc R/A/K iΓ[Φ¯ c,Φ¯ q] δΦ¯ δΦ¯ variance) we have assumed spatial translation invariance. G e = D[δΦc, δΦq] e c q . are called retarded, advanced, and Keldysh Green’s function, (39) and — in the terminology of statistical mechanics — the index On the right hand side, we have used the property of the Legen- d stands for disconnected averages obtained from differentiating dre transformation the partition function Z; the zero component is an exact prop- erty and reflects the elimination of redundant information (see ¯ ∗ ¯ ∗ δΓ/δΦc = −Jq, δΓ/δΦq = −Jc . (40) Sec. II B 1 below). We anticipate that the retarded and advanced components describe responses, and the Keldysh component the Equation (39) is obtained by exponentiating (i times) Eq. (38), correlations. The physical meaning of the Green’s function is and using the explicit functional integral representation of the 16

Keldysh partition function in W = −i ln Z, Eq. (33). We have in- ation of the effective action is precisely the full inverse Green’s ¯ troduced the notation Φν = Φν+δΦν. This implies hδΦνi|Jc,Jq = 0 function. by construction (the average is taken in the presence of non- Typically, an exact evaluation of the effective action is not vanishing sources), and moreover allows us to use D[Φc, Φq] = possible — it would constitute the full solution of the interacting D[δΦc, δΦq] in the functional measure. Note the appearance of non-equilibrium many-body problem. However, powerful ana- the field fluctuation δΦν in the last term in the functional inte- lytical tools have been developed for the analysis of such prob- gral (39). This is due to the explicit subtraction of the source lems, ranging from systematic diagrammatic perturbation the- term in Eq. (38). ory over the efficient introduction of emergent degrees of free- Equation (39) simplifies when we consider vanishing exter- dom to genuine non-perturbative approaches such as the func- nal sources Jc = Jq = 0. This case is particularly intuitive, as tional renormalization group. The latter technique is discussed it shows that the effective action Γ[Φ¯ c, Φ¯ q] corresponds to sup- in Sec.IIE. Examples in which such strategies were put into plementing the action S [Φ¯ c, Φ¯ q] by all possible fluctuation con- practice are discussed in Part2 of this review. tributions, expressed by the functional integration over δΦ¯ c and δΦ¯ q. The variational condition on Γ, Eq. (40) (technically reflect- B. Examples ing the change of active variables via Eq. (38)), precisely takes the form of the equation of motion derived from the principle of In this section, we bring the rather formal considerations of least action (in the presence of an external source or force term) the previous one to life by considering explicit examples for familiar form classical mechanics. Here, however, it governs the driven-dissipative systems. To be specific, in Sec. II B 1 we dynamics of the full effective action. In light of the above in- consider the decay of a single-mode cavity. This is arguably terpretation of the effective action, Eq. (40) thus promotes the one of the simplest examples that combines coherent and dissi- conventional classical action principle to full quantum and sta- pative dynamics: the system itself consists of a single bosonic tistical status. mode (the cavity photon), has a quadratic Hamiltonian and lin- Conversely, neglecting the quantum and statistical fluctua- ear Lindblad operators. Hence, the Keldysh action is quadratic tions in Eq. (39), we directly arrive at Γ[Φ¯ c, Φ¯ q] = S [Φ¯ c, Φ¯ q]. 9 and the functional integral can be solved exactly. Hence, we This approximation is appropriate at intermediate distances in are able to obtain an explicit expression for the generating func- the case where there is a macroscopically√ occupied conden- tional defined in Eq. (33), which allows us to conveniently study ¯ sate: we may then count Φc = O( N), with N the exten- several properties of the Keldysh formalism: what is known as sive number of particles in the condensate, while fluctuations the causality structure, the analyticity properties of the Green’s ¯ δΦc, δΦq = O(1), as well as Φq = O(1) for the noise field, which functions, and the intuitive and transparent way in which spectral cannot acquire a non-vanishing expectation value. This crude and statistical properties are encoded in the formalism. More- approximation of the full effective action reproduces the stan- over, we compare these findings with the case of a bosonic mode dard Gross-Pitaevski mean-field theory, if we consider a generic in equilibrium. The presence of the properties discussed in this Hamiltonian with kinetic energy and local two-body collisions, section is not restricted to the non-interacting case. Much rather, and drop the dissipative contributions to the action. they prevail also in the presence of interactions, and are thus ex- The functions generated by the effective action functional (via act properties of non-equilibrium field theories. We point this ¯ ¯ taking variational derivatives with respect to its variables Φc, Φq) out alongside the examples. are called the one-particle irreducible (1PI) or amputated vertex In Sec. II B 2, we consider an example for an interacting functions [182]. A crucial and useful relation between the 1PI bosonic many-body system, which contains non-linearities not vertex functions and connected correlation functions (generated only due to particle-particle interactions, but also in the dissi- by W) is pative contribution to the dynamics. These features are realized experimentally in exciton-polariton systems. In the present sec- Z  δ2Γ   0 δφ¯∗ t,x δφ¯ t00,x00  tion, we restrict ourselves to a mean-field analysis of this system −  c( ) q( )   δ2Γ δ2Γ  t00,x00  ∗ 00 00 ∗ 00 00  and defer a discussion of the role of fluctuations to Sec.IV. δφ¯q(t,x)δφ¯c(t ,x ) δφ¯q(t,x)δφ¯q(t ,x )  δ2W δ2W   δ j∗ t00,x00 δ j t0,x0 δ j∗ t00,x00 δ j t0,x0  ×  q( ) q( ) q( ) c( )  − 0 − 0 1  δ2W  = δ(t t )δ(x x ) . (41)  ∗ 00 00 0 0 0  1. Single-mode cavity, and some exact properties δ jc(t ,x )δ jq(t ,x )

This relation follows directly from Eqs. (37) and (40), using the The master equation describing the decay of photons in a chain rule. Combined with Eq. (36), it states that the second vari- single-mode cavity takes the general form of Eq. (16), with † † H = ω0a a, where a and a are creation and annihilation op- erators of photons, and ω0 is the frequency of the cavity mode. 9 At very long wavelengths, gapless fluctuations of the Goldstone mode lead to Assuming the external electromagnetic field to be in the vacuum infrared divergences in perturbation theory and invalidate this power counting state, there is only a single term in the sum over α with L = a, argument [184, 185]. describing the decay of the cavity field at a rate 2κ (the factor of 17

2 is chosen for convenience). The corresponding Keldysh action trace in the operator based master equation Eq. (16): in this way, is given by [112] using the cyclic property of the trace allows us to shift all opera- tors to one side of the density matrix, leading to the cancellation Z  ∗ ∗ of terms such that ∂t tr ρ(t) = 0. S = a+ (i∂t − ω0) a+ − a− (i∂t − ω0) a− t We still need to argue that the above property of the classical  ∗ ∗ ∗  action also holds for the full theory, i.e., the effective action, −iκ 2a+a− − a+a+ + a−a− , (42) ∗ ∗ ∗ Γ[¯ac, a¯ , a¯q = 0, a¯ = 0] = 0, (48) where a±, a± represent the complex photon field. Performing c q the basis rotation to classical and quantum fields as in Eq. (32) ¯ ¯ in Sec.IIA, and going to Fourier space, the action becomes or, more schematically in the notation of Sec.IIA, Γ[Φc, Φq = 0] = 0. To this end, we note that W[Jc, Jq = 0] = 0 Z ! !   0 PA(ω) a (ω) (Z[Jc, Jq = 0] = 1) holds actually for arbitratry classical sources S = a∗(ω), a∗(ω) c , (43) c q R K Jc (whereas in Sec.IIA we worked additionally with Jc = 0 ω P (ω) P aq(ω) for conceptual clarity): any term ∼ Jc can be absorbed into the R R ∞ ≡ dω underlying Hamiltonian, describing nothing but a Hamiltonian where we used the shorthand ω −∞ 2π . Furthermore, we have contribution in the presence of a classical external potential, and thus it cannot affect the normalization property of the theory. R A ∗ K P (ω) = P (ω) = ω − ω0 + iκ, P = 2iκ. (44) The above properties are equivalent: ¯ ¯ PR/A are the inverse retarded and advanced Green’s functions, Γ[Φc, Φq = 0] = 0 ⇔ W[Jc, Jq = 0] = 0. (49) and PK is the Keldysh component of the inverse Green’s func- This is seen using the definition of the Legendre transform tion. To see this, we evaluate the generating functional (33) Eq. (38) and the definition of the quantum field in terms of by Gaussian integration (cf. AppendixB). Then, the generating Eq. (37) for one direction of the mutual implication. The other functional (35) for connected correlation functions is given by direction results from the involutory property of the Legendre 10 Z K R ! ! transform .  ∗ ∗  G (ω) G (ω) jq(ω) W[Jc, Jq] = − jq(ω), jc(ω) A . (45) Sometimes, the property of conservation of probability – ex- ω G (ω) 0 jc(ω) pressed in the effective action formalism as Γ[Φ¯ c, Φ¯ q = 0] = 0 – According to Eq. (36), the second variation (i.e., the matrix in is referred to as “causality structure” in the literature. the above equation, see AppendixA) represents the Green’s Hermeticity properties of the Green’s functions — As can be R A function. It is obtained by inversion of the matrix in the action read off from Eqs. (46) and (47), G (ω) and G (ω) are Hermi- K in Eq. (43), tian conjugates, and G (ω) is anti-Hermitian. These properties are exact, as can be read off from the definition of the Green’s R A ∗ 1 1 function in terms of functional derivatives in Eq. (34). Equiva- G (ω) = G (ω) = R = , (46) P (ω) ω − ω0 + iκ lently, these properties hold for the corresponding components i2κ of the inverse Green’s function (cf. Eq. (41)). GK(ω) = −GR(ω)PKGA(ω) = − , (47) R 2 2 Analytic structure — The poles of G (ω) are located at ω = (ω − ω0) + κ ω0 − iκ, i.e., in the lower half of the complex plane (accordingly, i.e., the matrix in the action indeed represents the inverse Green’s the poles of GA(ω) are in the upper half). In the real time domain, function. We now summarize a few key structural properties that this implies that GR describes the retarded response (and ac- can be gleaned from this explicit discussion. Indeed, as we argue cordingly, GA the advanced): indeed, taking the inverse Fourier below, these properties are valid in general. transform, we obtain Conservation of probability — There is a zero matrix entry R −(iω +κ)t in Eq. (43), or, equivalently, in Eq. (45). Technically, as antic- G (t) = −iθ(t)e 0 , (50) ipated above, this property reflects a redundancy in the ± basis and eliminates it. This simplifies practical calculations in the where θ(t) is the Heaviside step function. Hence, the response of Keldysh basis. Physically, this property ensures the normaliza- the system, if it is perturbed at t = 0, is retarded (non-zero only tion of the partition function (Z = tr ρ(t) = 1), and can thus be for t > 0) and decays. interpreted as manifestation of the conservation of probability The analytic structure of retarded and advanced Green’s func- tions is a general property of Keldysh actions, too; one may then (∂t tr ρ(t) = 0), which is an exact property of physical problems. This can be seen as follows: consider the more general property, think of these Green’s functions as renormalized, full single par- which implies the vanishing matrix element in the quadratic sec- ticle Green’s functions connected to the bare, microscopic ones ∗ ∗ tor, S [ac, ac, aq = 0, aq = 0] = 0. Any Keldysh action associated to the Liouville operator Eq. (16), has this property, as can be seen by setting ψ+ = ψ− (i.e., φq = 0) in Eq. (28). Indeed, this operation on the Keldysh action may be interpreted as taking the 10 We thank F. Tonielli for pointing out this compact argument. 18 via renormalization, and thus preserving the analyticity proper- 0] = 1, and as a consequence, expectation values of n-point func- ties. We note, however, that the pole of the retarded bosonic tions of the fields a∗, a can be expressed via n-th order functional Green’s function can approach the real axis from below via tun- derivatives of the partition function with respect to the source ing of microscopic parameters. The touching point typically sig- fields (see AppendixA). For example nals a physical instability: beyond that point, the description of the system must be modified qualitatively. Such a scenario is δZ( jc, jq) hac(t)i = i , discussed in the next subsection. δ j∗(t) q jc= jq=0 Connection to the operator formalism — It is sometimes use- δ2Z( j , j ) ful to restore the precise relation between the operator formalism K 0 − h ∗ 0 i c q G (t, t ) = i ac(t)ac(t ) = i ∗ 0 , (57) and the functional integral description at the level of the Green’s δ j (t)δ jq(t ) q jc= jq=0 functions. At the single particle level, they read δ2Z( j , j ) R 0 − h ∗ 0 i c q R 0 0 † 0 G (t, t ) = i ac(t)aq(t ) = i ∗ 0 . G (t, t ) = −iθ(t − t )h[a(t), a (t )]i, (51) δ j (t)δ jc(t ) q jc= jq=0 GK(t, t0) = −ih{a(t), a†(t0)}i, (52) Due to causality, the field jq has to be zero in any physical setup and we note again that in stationary state GR/A/K(t, t0) = and the introduction of this field only serves as a technical tool to GR/A/K(t − t0). These relations are exact and can be obtained compute expectation values via derivatives. On the other hand, from going back to the ± basis: the retarded Green’s function is the field jc can, in principle, be different from zero and we will see in the following what the physical meaning of this classical R 0 ∗ 0 G (t, t ) = −ihac(t)aq(t )i source field is, and in which respect it generates the response i function. = − h(a (t) + a (t)) (a∗ (t0) − a∗ (t0))i 2 + − + − Responses: The retarded Green’s function, often called syn- i   onymously response function, describes the linear response of a = − hTa(t)a†(t0)i + h[a(t), a†(t0)]i − hTa˜ (t)a†(t0)i system which is perturbed by a weak external source field. As an 2 0 † 0 illustrative example, consider a driven cavity system consisting = −iθ(t − t )h[a(t), a (t )]i, of atoms and photons, which is considered to be in a stationary (53) state. Due to imperfections in the cavity mirrors, there is a finite rate with which a photon is escaping the cavity, or an external where T, T˜ are the time-ordering, anti-time-ordering operators, 0 photon is entering the cavity. This process is expressed by the which lead to a cancellation of the commutator for t > t (see Hamiltonian Ref. [102] for a more detailed discussion of time ordering on the √ Keldysh contour). The Keldysh Green’s function in the operator  † †  Hp = 2κ b a + a b , (58) formalism is obtained the same way, † 0 ∗ 0 where the operators b, b represent photons outside the cavity. GK(t, t ) = −iha (t)a (t )i c c Shining a laser through the cavity mirror, the operators b, b† can i ∗ 0 ∗ 0 be replaced by the coherent laser field j(t), j∗(t), which oscillates = − h(a+(t) + a−(t)) (a+(t ) + a−(t ))i 2 with the laser frequency ω j. The corresponding Hamiltonian, i   = − hTa(t)a†(t0)i + h{a(t), a†(t0)}i + hTa˜ (t)a†(t0)i describing the complete system is (for the present purposes we 2 can leave the Hamiltonian H of photons and atoms in the cavity = −ih{a(t), a†(t0)}i. unspecified) (54) √  ∗ † H j = H + 2κ j (t)a + j(t)a . (59) Response vs. correlation functions — In order to generate re- sponse and correlation functions directly from the partition func- Since the fields j, j∗ are classical external fields, they are equal tion, it is convenient to introduce source fields j+, j− and express on the plus and minus contour and the Keldysh action in the the partition function as the average (cf. Eq. (31)) presence of the laser is Z Z − ∗ ∗ −   Z[ j , j ] = he iS j i = D[a , a , a , a ]eiS iS j , (55) ∗ ∗ + − + − + − S j = S − j (t)aq(t) + j(t)aq(t) . (60) t where the source action is defined as This action is very similar to the action in Eq. (56), with a lin- Z Z ear source term, which is however j = j = j. In the present ∗ ∗   ∗ ∗  + − S j = j+a+ − j−a− + c.c. = jcaq + jqac + c.c. . (56) example, the meaning of the source term is physically very trans- t t parent: it is nothing but the coherent laser field that is coupled to In the second step, we have performed a Keldysh rotation. Due the cavity photons. For a weak source field, one is interested in to the normalization of the Keldysh path integral Z[ jc = 0, jq = the first order correction of observables induced by the coupling 19

A(ω) = 2πδ(ω − ω0). As these considerations illustrate, the re- tarded Green’s function contains essential information on the system’s response towards an external perturbation, and on the spectral properties. Correlations: The Keldysh Green’s function contains elemen- tary information on the system’s correlations and the occupation of the individual quantum mechanical modes. A prominent ex- ample of a correlation function in quantum optics is the photonic g(2) correlation function. It is defined as the four-point correlator † † Figure 4. Illustration of a homodyne detection measurement which (2) ha (t)a (t + τ)a(t + τ)a(t)i 0 g (t, τ) = , (64) determines the response function GR(t, t ) of the cavity photons, see |ha†(t)a(t)i|2 Ref. [113]. The system of atoms and photons inside the cavity is per- turbed by the laser field η(t), entering the cavity through the left mirror. and it is proportional to the intensity fluctuations of the intracav- The response of the system is encoded in the light field which is leaking ity radiation field g(2)(t, τ) ∝ hI(t)I(t + τ)i. In the limit of τ → 0, out of the cavity on the right mirror with a rate κ. It can be measured the g(2) correlation function reveals the statistics of the cavity by a standard homodyne detection measurement in which the reference photons, i.e., it demonstrates super-Poissonian (g(2)(0) > 1) or laser β(t) and a beam splitter are used in order to obtain information on sub-Poissonian statistics (g(2)(0) < 1) as an effect of the light- the system’s coherence hac(t)i. Figure copied from Ref. [113]. (Copy- matter interactions. In the absence of interactions (as in our ex- right (2013) by The American Physical Society.) ample), it is straightforward to show that |GK(t, τ + t) + GR(t, τ + t) − GA(t, τ + t)|2 g(2)(t, τ) = 1 + . (65) to the source. In the present case, this is the coherent light field |iGK(t, t) − 1|2 inside the cavity hac(t)i. Up to first order in j(t) it is given by For equal times, the retarded and advanced Green’s functions √ Z satisfy GR(t, t) − GA(t, t) = −i, indicating the expected photon- h i h i − h ∗ 0 i 0 ac(t) j = ac(t) j=0 i κ ac(t)aq(t ) j=0 j(t ) bunching at τ → 0. On the other hand, in the presence of in- t0 √ Z teractions, Eq. (64) is modified and contains additional higher R 0 0 = hac(t)i j=0 + κ G (t − t ) j(t ). order terms, as well as off-diagonal contributions. However, it t0 remains a measure of the photonic statistics in the cavity due to The retarded Green’s function is therefore a measure of the sys- the physical relation to the intensity fluctuations. tem’s response to an external perturbation. An experimental A particularly relevant and instructive limit of the two-time 0 setup, with which one can measure the coherent light field and Keldysh Green’s function Eq. (52) concerns equal times t = t , therefore the response function of the cavity via so-called homo- which in general describes static correlation functions, or covari- dyne detection, is illustrated in Fig.4. A closely related function ances in the quantum optics language. In the cavity example, it of interest is the spectral function yields the mode occupation number, iGK(t, t) = 2ha†ai + 1. (66) A(ω) = −2 Im GR(ω). (61) The appearance of the combination 2ha†ai + 1 is rather intuitive It is the distribution of excitation levels of the system, i.e., when when taking into account the relation to the operatorial formal- adding a single photon with frequency ω to the system, A(ω) is ism. Indeed, in operator language, 2a†a + 1 = {a, a†} = {a†, a} the probability to hit the system at resonance. Indeed, it can be is invariant under permutation of the operators a, a†, and there- shown [178] that the spectral function is positive, A(ω) > 0, and fore lends itself to a direct functional integral representation, fulfills the sum rule which in fact carries no information on operator ordering.11 At Z the same time, the anticommutator carries all the physical infor- A(ω) = h[a, a†]i = 1. (62) mation on state occupation, while the commutator carries none, ω † † 1 † † [a, a ] = 1 (note a a = 2 ({a, a } − [a, a ])). For our example of a single-mode cavity, the spectral function is Then, the explicit form of the Keldysh Green’s function stated given by in Eq. (47) leads to the result Z ! 2κ † 1 K A(ω) = , (63) ha ai = i G (ω) − 1 = 0, (67) 2 2 2 ω (ω − ω0) + κ i.e., it is a Lorentzian, which is centered at the cavity frequency ω0 and has a half-width at half-maximum given by κ. Note that 11 The same structural argument based on operator ordering insensitivity of the for κ → 0, the photon number states become exact eigenstates functional representation demonstrates why an ordinary Euclidean functional integral yields the time ordered Green’s functions, cf., e.g., Ref. [180]. and the spectral density reduces to a δ-function peaked at ω0, 20 showing that the cavity is empty in the steady state—which is equilibrium Green’s functions not surprising in the absence of external pumping. 1 We note that in the master equation formalism, information on GR(ω) = GA(ω)∗ = , (69) − the static or spatial correlations is most easily accessible — only ω ω0 + iδ K the state (density matrix) has to be known, but not the dynamics G (ω) = −i2πδ(ω − ω0) (2n(ω) + 1) . (70) acting on it. Temporal correlation functions can be extracted using the quantum regression theorem [186]. In the Keldysh Note that the Green’s functions obey a thermal fluctuation- formalism, spatial and temporal correlation functions are treated dissipation relation (FDR), which for the present example of a on an equal footing. single bosonic mode reads   We can concisely summarize the above discussion of the re- GK(ω) = (2n(ω) + 1) GR(ω) − GA(ω) . (71) sponse and Keldysh Green’s functions of the system, at the risk of oversimplifying: This is discussed in more detail in Sec. II D 1. For the time being, let us mention that this construction to describe thermodynamic equilibrium by adding infinitesimal dissipative terms does not Responses GR/A ↔ spectral information on excitations: only work in the present case of a quadratic action but can also which excitations are there? be applied in the interacting case. Then, the construction ensures Correlations GK ↔ statistical information on excitations: that the free Green’s functions (i.e., the ones obtained by ignor- how are the excitations occupied? ing the interactions) obey an FDR. If the non-linear terms in the action obey the equilibrium symmetry discussed in Sec. II D 1 (which is the case for generic interaction terms), this property is Relation to thermodynamic equilibrium — Before we move shared by the full Green’s functions of the non-linear system. on, let us briefly discuss how we have to modify the formal- General Fluctuation-Dissipation Relations — While Eq. (71) ism in order to describe a cavity in thermodynamic equilibrium. is valid only in thermodynamic equilibrium, it is always possible In general, a system is in thermodynamic equilibrium at a tem- to parameterize the (anti-Hermitian) Keldysh Green’s function perature T = 1/β, if it is in a Gibbs state ρ = e−βH/Z where in terms of the retarded and advanced Green’s functions (which, −βH as discussed above, are Hermitian conjugates of each other) and Z = tr e , and its dynamics is coherent and generated by the † Hamiltonian H. (In particular, dissipative dynamics described a Hermitian matrix F = F in the form [102, 178] by a term in the Lindblad form in Eq. (16) is not compatible GK = GR ◦ F − F ◦ GA, (72) with equilibrium conditions: see Refs. [187–189], and the dis- cussion in Sec. II D 1.) Specifying the thermal density matrix where ◦ denotes convolution. In this parametrization, F is the at t0 in the Keldysh functional integral in Eq. (25) explicitly in distribution function, which describes the distribution of (quasi- terms of its matrix elements is rather inconvenient, especially if ) particles over the modes of the system. For a non-equilibrium one is interested in steady state properties and wants to take the steady state, F is time-translational invariant and its Fourier limit t0 → −∞. An alternative is suggested by the observation transform in frequency space F(ω) represents the energy re- that thermodynamic equilibrium can be established in the system solved occupation of (quasi-) particle modes. On the other hand, if it is weakly coupled to a thermal bath. Then, the finite decay as we discuss in detail in Sec.VB, for the case of a time- rate κ in the retarded Green’s function in Eq. (46) should be re- evolving system, for which time translation invariance is absent, placed by an infinitesimally weak system-bath coupling δ → 0. the Wigner transform of F corresponds to the instantaneous local Additionally, the Keldysh Green’s function has to be modified in distribution function. For the important case of thermodynamic such a way that the integral in Eq. (67) yields the Bose distribu- equilibrium (of a bosonic system), it is F(ω) = coth(βω/2) = tion function n(ω0) implying thermal occupations of the cavity 2n(ω) + 1, and Eq. (72) reduces to (71). For the case where the mode, bosonic Green’s function has a matrix structure in Nambu space, a subtlety concering the preservation of the symplectic structure † 1 of that space arises, cf. Sec.VB. There, also a time-dependent ha ai = n(ω0) = . (68) eβω0 − 1 variant of the non-equilibrium fluctuation-dissipation relations is discussed. This can be achieved by replacing the Keldysh component of the inverse Green’s function in Eq. (43) by PK(ω) = i2δ (2n(ω) + 1). We emphasize the key structural difference of the thermal 2. Driven-dissipative condensate Keldysh component, which is strongly frequency dependent, to the Markovian case discussed previously, where this entry is fre- In the previous section, we discussed the simple case of quency independent — this gives a strong hint that Markovian a quadratic Keldysh action, which allowed us to perform the systems can behave quite differently from systems in thermal Keldysh functional integral explicitly. An additional simplifi- R A ∗ equilibrium. With P (ω) = P (ω) = ω − ω0 − iδ we obtain the cation resulted from the fact that we were considering only a 21 single bosonic mode. Let us now consider a genuine many-body The solution φc = φ0 is determined by the imaginary part of problem, which is non-linear and in which the system consists Eq. (75): for rd ≥ 0, in the so-called symmetric phase, the clas- 2 of a continuum of modes. To be specific, in this section we dis- sical field expectation value is zero, ρ0 = |φ0| = 0, whereas for cuss the model introduced in Sec.IC2 for a bosonic many-body rd < 0 we have a finite condensate density ρ0 = −rd/ud. Tak- system with interactions and non-linear loss processes (i.e., with ing the real part of Eq. (75), we obtain for the parameter rc the a loss rate that is proportional to the density) in addition to the relation rc = ucrd/ud. This condition can always be satisfied by linear dissipative terms which were already present in the exam- proper choice of a rotating frame, i.e., by performing a gauge −iωt ple of the single-mode cavity. Then, for a specific value of the transformation φc 7→ φce such that rc 7→ rc − ω. In the orig- mean-field density ρ0, non-linear loss and linear pump exactly inal (laboratory) frame this simply means that the condensate balance each other and the system reaches a stationary state. If amplitude oscillates at a finite frequency. ρ0 is different from zero, this signals the presence of a conden- In a first step beyond mean field theory, quadratic fluctu- sate, which is accompanied by the breaking of a specific phase ations around the mean-field order parameter can be investi- rotation symmetry as will be discussed in detail in Sec.IID, and gated within a Bogoliubov or tree-level expansion: we set φc = the establishment of long-range order. In Sec.IV, we give a φ0 +δφc, φq = δφq in the action Eq. (73) and expand the resulting detailed account of the influence of fluctuations on this driven- expression to second order in the fluctuations δφc,q. The inverse dissipative condensation transition. Here, we content ourselves retarded, advanced and Keldysh Green’s functions now become with a mean-field analysis, which serves to illustrate some of the ∗ 2 × 2 matrices in the space of Nambu spinors δΦν = δφν, δφν . field theoretical concepts introduced in Sec.IIA — in particular, In particular, we have in the frequency and momentum domain the effective action and field equations — in a simple setting. In the basis of classical and quantum fields, the Keldysh action 2 ! R ω − Kcq − (uc − iud) ρ0 − (uc − iud) ρ0 associated with the quantum master equation (9) reads P (ω, q) = 2 , − (uc + iud) ρ0 −ω − Kcq − (uc + iud) Z A R † n ∗  2  P (ω, q)) = P (ω, q) , S = φq i∂t + Kc∇ − rc + ird φc + c.c. t,x PK = iγ1. h  ∗ ∗ 2 ∗ ∗ 2 i − (uc − iud) φqφcφc + φqφcφq + c.c. (76) ∗  ∗ o +i2 γ + 2udφ φc φ φq , (73) c q The excitation spectrum is obtained from the condition 0 det PR(ω, q) = 0. Indeed, this is the condition for the field equa- where Kc = 1/(2mLP) and rc = ω ; as additional parameters, LP tion of the fluctuations PR(ω, q)δΦ (ω, q) = 0 to have nontrivial we introduced the noise level γ = (γl + γp)/2 and the spec- c solutions. This yields [159] tral mass or gap rd = (γl − γp)/2. Hence, the rates of losses and pumping add up to the total noise level; in contrast, the dif- q ference of these rates enters in the spectral gap r , which be- d ωR = −iu ρ ± K q2 K q2 + 2u ρ  − (u ρ )2. (77) comes negative when the rate of incoherent pumping exceeds the 1,2 d 0 c c c 0 d 0 single-particle loss rate, signaling the physical instability against condensation. In a mean-field analysis of the condensation tran- We note that due to the tree-level shifts ∝ ρ0 the above described sition, we perform a saddle-point approximation of the func- instability for rd < 0 is lifted: both poles are consistently lo- tional integral in Eq. (39). To leading order, fluctuations around cated in the lower complex half-plane, indicating a physically the field expectation values are completely neglected. The ex- stable situation with decaying single-particle excitations. For pectation values are then obtained as spatially homogenous and ud = 0, Eq. (77) reduces to the standard Bogoliubov result [190], R stationary solutions to the classical field equations (here, “classi- where for q → 0 the dispersion is phononic, ω = ±cq p 1,2 cal” refers to the fact that these field equations are derived from (c = 2Kcucρ0 the speed of sound), whereas particle-like be- S R 2 the classical (or: bare, microscopic) action discarding fluctua- havior ω1,2 ∼ Kcq is obtained at high momenta. Here, due to tions, in contrast to the field equations in Eq. (40), which involve the presence of two-body loss ud , 0, the dispersion is qualita- the effective action) tively modified: while at high momenta the dominant behav- ior is still ωR ∼ K q2, at low momenta we find purely in- δS δS 1,2 c R Kcuc 2 ∗ = 0, ∗ = 0. (74) coherent diffusive, non-propagating modes ω ∼ −i q and δφ δφ 1 ud c q R R ω2 ∼ −i2udρ0. In particular, for q = 0 we have ω1 = 0: this As already mentioned above in Sec. II B 1, there are no terms is a dissipative Goldstone mode [159, 191, 192], associated with ∗ in the action Eq. (38) with zero power of both φq and φq, and the spontaneous breaking of the global U(1) symmetry in the or- ∗ the same is clearly true for δS/δφc. Therefore, the first equation dered phase. The existence of such a mode is not bound to the in (74) is solved by φq = 0. Inserting this condition into the mean-field approximation, but rather is an exact property guar- second equation, we have anteed by the U(1) invariance of the effective action, even in the present case of a driven-dissipative condensate. We will come h 2i −rc + ird − (uc − iud) |φ0| φ0 = 0. (75) back to this point in Sec. II D 6. 22

C. Semiclassical limit of the Keldysh action as canonical scaling analysis, or power counting. A useful ap- proximation consists in neglecting all irrelevant couplings. On the other hand, relevant couplings which are compatible with the The theoretical description of a many-body system depends symmetries of the model – even those, which are not present in crucially on the scale (this could be a length, time, or energy the microscopic description — should be included in the long- scale — in practice these scales can be expressed in terms of wavelength effective action. each other) on which it is observed and, in particular, on the re- lation between this observation scale and the intrinsic scales of Let us perform such an analysis for the Keldysh action in the system. For example, at finite temperature T above a quan- Eq. (73). We can anticipate the canonical scaling dimensions tum critical point, non-trivial quantum critical behavior can be by noting that they are just the physical dimensions, measured observed at moderate energy scales which are larger than T but in powers of the momentum. We first focus on the vicinity of ∼ − → smaller than the Ginzburg scale where fluctuations start to dom- the threshold for condensation, i.e., when rd γl γp 0. inate over the mean field effects [111]. On the other hand, clas- Then the retarded and advanced inverse Green’s functions scale R/A ∼ 2 ∼ 2 sical thermal critical behavior, which is perfectly described by as P q (note that ω q for low-momentum excitations). taking the semiclassical limit of the underlying quantum theory, The canonical dimension is thus positive, dPR/A = 2. On the sets in at energy scales below T. Out of thermodynamic equilib- other hand, as anticipated above, Markovian noise analogous to rium, the analogue of a finite temperature is Markovian noise. a finite temperature corresponds to a momentum-independent ∝ ∗ In the Keldysh action, this corresponds to a constant term in noise vertex, i.e., the term φqφq in the Keldysh action. In the Keldysh sector of the inverse Green’s function, i.e., a con- other words, the Keldysh component of the inverse Green’s func- stant noise vertex. Such a term is present in the action given in tion has vanishing canonical scaling dimension and classifies as K ∼ 0 Eq. (73) and, therefore, we expect that in the long-wavelength marginal, P = i2γ q . We furthermore use the natural scal- d ∼ −d ∼ ∼ −2 limit this action can actually be simplified by taking the semi- ing d x q and dt 1/ω q , and the condition of scale ∼ 0 classical limit [102, 178]. We refer to it as the semiclassical invariance of the action, S q — this is a requirement stem- limit, as, e.g., effects of quantum mechanical phase coherence ming from the fact that the action appears in the exponent of the are not necessarily suppressed in this limit, as we will see. A functional integral (25), and thus must be dimensionless. We useful equilibrium analogy is the physics of Bose-Einstein con- then find the scaling dimensions of the fields from the Gaussian, densates at finite but low temperatures, where phase coherence quadratic part of the action to be still persists. (d−2)/2 (d+2)/2 φc ∼ q , φq ∼ q . (78) Formally, the suitability of the semiclassical limit can be un- derstood in terms of ideas that lie at the basis of the renormal- This in turn allows us to derive the scaling dimensions of the ization group (RG). The latter provides a recipe for finding an quartic terms, and to classify their degree of relevance as pointed effective description that is valid on large length scales, starting out above. In particular, we find that in three spatial dimensions, from a microscopic theory. Various RG schemes exist, which al- any quartic term that includes more than a single quantum field low to systematically eliminate fluctuations on short scales and is irrelevant; the only non-irrelevant term of higher order in the infer their influence on the effective description of the system on noise field is the quadratic noise vertex discussed above. There- large scales; cf. Sec.IIE. The most basic but still informative fore, omitting irrelevant terms amounts to keeping only the clas- RG approach however consists in simply ignoring the effect of sical vertex in the action Eq. (73), i.e., to taking the semiclassical short-scale fluctuations: one subdivides momentum space into a limit [102, 178]. Then, the Keldysh action takes the form slow and a fast region, with qs ∈ [0, Λ/b] and q f ∈ [Λ/b, Λ] where Λ is the UV cutoff (given by the inverse of some mi- Z n ∗ h 2 i croscopic length scale below which the theoretical description S = φq i∂t + (Kc − iKd) ∇ − rc + ird φc + c.c. is not valid any more, e.g., a lattice spacing) and b > 1, and t,x h ∗ ∗ 2 i ∗ o omits all contributions to the action which involve fluctuations − (uc − iud) φqφcφc + c.c. + i2γφqφq , (79) with fast momenta. In a second step, one rescales all momenta with b to restore the original range of momenta q ∈ [0, Λ], and where in addition to omitting irrelevant contributions we have 2 the thus obtained effective long-wavelength description can be added an effective diffusion term ∝ iKd∇ . Such a term can compared to the original one. As a result, due to this simple RG and will be generated upon integrating out short-scale fluctua- transformation couplings in the action are rescaled as g 7→ gbdg , tions [26, 181]. Moreover, a complex prefactor of the term in- where dg is known as the canonical scaling dimension of g. Ev- volving the derivative with respect to time, which also emerges idently, for dg > 0, g grows under renormalization and hence upon renormalization, can be absorbed into a redefinition of the g is a relevant coupling, while it shrinks under renormalization fields. and, hence, is irrelevant at long wavelength for dg < 0. For the Strictly speaking, the above analysis is valid in the vicinity of case of a marginal coupling with dg = 0, this level of approx- the critical point only, where the inverse retarded and advanced imation is not conclusive as to whether the coupling becomes Green’s function show scaling. However, as long as one is close larger or smaller under renormalization. Classifying the cou- enough to threshold |γl − γp|/(γl + γp)  1, the canonical power pling parameters of an action according to this scheme is known counting is expected to give a useful orientation in the problem. 23

quantum quantum Keldysh functional microscopic Langevin equation master equation integral

semiclassical limit

Langevin MSR functional quantum mesoscopic Fokker-Planck equation equation integral problem

renormalization group

long-wavelength macroscopic effective action scale

Figure 5. Equivalent descriptions on varying length scales. The quantum and classical Langevin equations are stochastic equations of motions for the field operators, or for classical field variables. In contrast, the descriptions in the middle column are deterministic equations of motion, where either a density operator or a probability distribution (diagonals of a density matrix) are evolved. In the functional integral formulation, the basic object in both quantum and classical cases is an action, which is averaged over all possible realizations of field configurations. The semiclassical limit is valid at mesoscopic scales as dicsussed in Sec.IIC. An e ffective description at macroscopic scales can be obtained by means of renormalization group methods (see Sec.IIE). Generically in Markovian systems, in order to reduce the complexity of the problem it is useful to first perform the semiclassical limit before doing a renormalization group computation. However, in driven open quantum systems there are also circumstances where this is inappropriate, and the full quantum problem has to be analyzed, cf. [28].

Apart from providing a significant simplification, taking the can be performed, resulting in a δ-functional, semiclassical limit also allows us to establish the connec- Z R ∗ ∗ − 1 ξ∗ξ tion [193, 194] between the Keldysh functional integral for- 2γ t,x Z = D[φc, φc, ξ, ξ ] e malism and the more traditional formulation of dynamics close h 2 2i  to a continuous phase transition in terms of Langevin equa- × δ i∂t + (Kc − iKd) ∇ − rc + ird − (uc − iud) |φc| φc − ξ tions [1,5]. In fact, the action in Eq. (79) is fully equivalent h 2 2i ∗ ∗ to the following Langevin equation: ×δ −i∂t + (Kc + iKd) ∇ − rc − ird − (uc + iud) |φc| φc − ξ . (83) h 2 2i This expression can be interpreted as follows: for a given i∂tφc = − (Kc − iKd) ∇ + rc − ird + (uc − iud) |φc| φc + ξ, (80) realization of the noise field ξ, the δ-functional restricts the where ξ is a Markovian Gaussian noise source with zero mean, functional integral over φc to the manifold of solutions to the hξ(t, x)i = 0, and second moment hξ(t, x)ξ(t0, x0)i = 2γδ(t − Langevin equation (80). The statistics of the noise field is de- t0)δ(x − x0). This equivalence can be established by means termined by the Gaussian weight factor in Eq. (83). Correlation of a Hubbard-Stratonovich transformation of the noise ver- functions of classical fields can then be calculated by picking a ∗ random realization of ξ and calculating the corresponding φc that tex [102, 178], i.e., the term i2γφqφq. To wit, in the Keldysh functional integral solves the Langevin equation; evaluating the correlation function for this solution and finally averaging the result according to the Z Gaussian distribution of the noise field. Equation (80) is, there- ∗ ∗ ∗ ∗ iS [φc,φc,φq,φq] fore, equivalent to the functional integral (83) in that it allows Z = D[φc, φc, φq, φq] e , (81) for the evaluation of arbitrary correlation functions. Originally, Langevin equations like Eq. (80) have been intro- where S is the semiclassical action in Eq. (79), we write the duced as a phenomenological description of the coarse-grained noise vertex as a Gaussian integral over an auxiliary variable ξ dynamics of the order parameter, and only later a functional in- (cf. AppendixB), tegral approach — known as the MSR approach — has been developed [195–198]. The action derived in these references is Z formally equivalent to the one in Eq. (79), with φc taking the role R ∗ 1 R ∗ R ∗ ∗ −2γ φqφq ∗ − ξ ξ−i (φqξ−ξ φq) e t,x = D[ξ, ξ ] e 2γ t,x t,x . (82) of the order parameter field, while the field that corresponds to φq is known as the response field. In addition to the descriptions of semiclassical dynamical As a result, the exponent in the functional integral (81) becomes models in terms of a semiclassical Keldysh (or MSR) func- linear in the quantum field, and the corresponding integration tional integral or a Langevin equation, there exists yet another 24 equivalent approach, in which the stochastic Langevin equation Finally, another consequence of a global continuous symme- for the classical field or order parameter field is replaced by try is the existence of gapless modes in case that this symmetry a deterministic evolution equation for the probability distribu- is broken spontaneously, as in a Bose condensation transition. tion of the latter, known as the Fokker-Planck equation. The In a driven-dissipative condensate in which the number of parti- derivation of the Fokker-Planck equation can be found, e.g., in cles is not conserved (cf. Sec. II B 2), we show that the symme- Refs. [5, 102, 178, 199]. Figure5 illustrates the equivalence of try which is broken spontaneously at the condensation transition these approaches, which are valid on a (coarse-grained) meso- is the classical phase rotation symmetry, and we work out the scopic scale, and the relations to microscopic and macroscopic Goldstone theorem, which guarantees the existence of a mass- descriptions. less mode, for this case in Sec. II D 6.

D. Symmetries of the Keldysh action 1. Thermodynamic equilibrium as a symmetry of the Keldysh action

Symmetries take center stage in field theories. Translating the A system is in thermodynamic equilibrium at a temperature physics of the quantum master equation to the Keldysh func- T = 1/β, if (i) the density matrix is given by ρ = e−βH/Z where tional integral allows us to leverage the power of symmetry con- Z = tr e−βH, and (ii) the very same Hamiltonian operator H ap- siderations over to the context of open systems. In this section, pearing in ρ generates the unitary time evolution of the system, we discuss three different aspects of symmetries. U = e−iHt. Condition (ii) implies that static correlations are in The presence of a first, discrete symmetry, considered in general not sufficient to prove that a system is in thermal equi- Sec. II D 1, allows us to conclude whether or not a quantum or librium; however, dynamical correlations are: this can be in- classical system is situated in the realm of thermodynamic equi- ferred from the fact that the static correlations of a physical sys- librium. This symmetry is equivalent to the validity of thermal tem can always be encoded in a density matrix ρ, and the latter fluctuation-dissipation relations (cf. Sec. II B 1) for correlation can always be parameterized formally as an equilibrium density 0 functions of any order and can be connected to energy conser- matrix, ρ = e−βH /Z0, with a Hermitian operator H0. (In other vation. Thermal equilibrium can thus be diagnosed by means words, any state can be thought of as a thermal state with respect a of simple symmetry test on the Keldysh action. In particular, to some Hamiltonian.) On the other hand, static (i.e., purely we argue that Markovian quantum master equations explicitly momentum- or space-dependent) properties do not allow us to violate this symmetry, indicating non-equilibrium conditions. In discriminate whether the generator of dynamics coincides with the semiclassical limit, we obtain a simple and intuitive geomet- H0. In sharp contrast, any dynamical (i.e., frequency- or time- ric interpretation of the symmetry and its absence under non- dependent) observable is manifestly governed by this generator, equilibrium drive in terms of the location of the coupling con- as is easily seen in the Heisenberg picture (or a suitable general- stants of the effective action in the complex plane. ization to open systems). Response functions at finite frequency A second fundamental consequence of the presence of sym- are such dynamical observables, and the equilibrium conditions metries — now, continous global symmetries — is the Noether formulated above are reflected in the fact that in thermodynamic theorem, stating that such symmetries imply conserved charges. equilibrium the response of the system to an external perturba- In Sec. II D 3, we discuss the Noether theorem in the context tion at a frequency ω is related to thermal fluctuations within the of the Keldysh formalism. Working on the Keldysh contour, system at the same frequency by what is known as a fluctuation- we have to distinguish between two symmetry transformations, dissipation relation [200] (FDR; for a discussion in the context for each the forward and backward branches of the closed time of non-equilibrium Bose-Einstein condensation see Ref. [201]). path. In the basis of classical and quantum fields, we can iden- Such a relation, which is equivalent to the combination of the tify “classical” symmetry transformations, which act in the same Kubo-Martin-Schwinger (KMS) condition [202, 203] with time way on fields on the forward and backward branches, and “quan- reversal [204], is valid for any pair of operators. In particular, tum” transformations, for which the transformation of the fields for the case that these are the basic field operators of a single- on the backward branch is the inverse of the transformation on component Bose system at the temperature T = 1/β, the FDR the forward branch. Non-trivial conservation laws follow only reads (note that this is a generalization of Eq. (71) to the case of from the symmetry of the Keldysh action under quantum trans- a spatial continuum of degrees of freedom) formations. We illustrate this point with the example of the sym- metry of the action of a closed system with respect to space-time  ω  GK(ω, q) = coth (GR(ω, q) − GA(ω, q)) (84) translations and phase rotations in Sec. II D 4, which entails con- 2T servation of energy and momentum and the number of particles, respectively. On the other hand, in open systems, in which the (in the presence of a chemical potential µ, in the argument of the number of particles is not conserved, only classical phase rota- trigonometric function we have to shift ω → ω − µ). In quan- tions are a symmetry of the action, and the continuity equation tum field theory, relations among Green’s functions often follow that is implied by particle number conservation has to be ex- as consequences of a symmetry of the action. Then, they are tended as detailed in Sec. II D 5. known as Ward-Takahashi identities associated with the sym- 25 metry [86, 205].12 This raises the question, whether FDRs and By generalizing this argument to arbitrary field averages, one hence the presence of thermodynamic equilibrium conditions can establish the full equivalence between the infinite hierar- are also connected to a symmetry of the Keldysh action. chy of FDRs and the symmetry property of the Keldysh ac- To make the point clear, let us rephrase the question: above we tion (85)[189]. Thus, the symmetry of the Keldysh action under made the observation, that the defining property of (canonical) this transformation is a direct proof of the presence of thermo- thermal equilibrium is the validity of FDRs. Importantly, these dynamic equilibrium conditions. The existence of a symmetry FDRs hold for correlation functions of arbitrary order, i.e., not that is related to thermal equilibrium has first been realized in the just the two-point Keldysh Green’s function in Eq. (84) can be context of classical stochastic models [197, 198, 206–208], and expressed through response functions, but also any higher order these considerations have been extended to the realm of quantum correlation function is determined by corresponding higher or- systems in Refs. [189, 209]. der response functions. Then, the question — which we answer What is the explicit form of the transformation Tβ? We can in the affirmative below — is, whether this infinite hierarchy of guess its essential ingredients by reminding ourselves that, as FDRs expressing thermal equilibrium is related to a structural stated above Eq. (84), FDRs can be obtained from the KMS property of the theory. Indeed, this structural property is a sym- condition [202, 203]. The latter reads, for operators A(t) = metry of the Keldysh action, i.e., a transformation of the fields eiHtA(0)e−iHt and B(t0) in the Heisenberg representation, Ψ 7→ TβΨ such that hA(t)B(t0)i = hB(t0)A(t + iβ)i. (87) S [Ψ] = S˜ [TβΨ]. (85) ∗ ∗  Comparing this with Eq. (86) indicates that Tβ involves a trans- Here, we denote Ψ = ψ+, ψ+, ψ−, ψ− , and we specify the pre- lation of t into the complex plane by an amount proportional to T cise form of β below in Eq. (88). The tilde on the RHS of the inverse temperature β. Moreover, the order of operators on the equation indicates that all external fields appearing, S have the RHS of the KMS condition is reversed as compared to the to be replaced by their corresponding time-reversed values. To LHS. The original time order can be restored by means of a time name an example, the signs of magnetic fields have to be in- reversal transformation [210] — this step is necessary in order verted [189]. Evidently, discussing a single symmetry instead of to obtain a time-ordered expression which can be expressed as a an infinite hierarchy of equations is much more elegant and prac- Keldysh field integral (by construction, field integrals yield time- tical: checking whether a given Keldysh action obeys Eq. (85) is ordered averages [180]). A careful analysis [189] leads to the straightforward and can be accomplished in finite time, in con- precise form of the thermal symmetry transformation (σ = +(−) trast to verifying the full set of FDRs. for fields on the forward (backward) branch): It is easily seen, how these FDRs can be deduced as a con- sequence of the symmetry of the Keldysh action (85). To this ∗ Tβψσ(t, x) = ψσ(−t + iσβ/2, x), end, we note by a Keldysh rotation (32) and after Fourier trans- (88) T ψ∗ (t, x) = ψ (−t + iσβ/2, x) formation, the Green’s functions appearing in Eq. (84) can be β σ σ h ∗ 0 0 i expressed as a sum of averages of the form ψσ(t, x)ψσ0 (t , x ) . (in the presence of a chemical potential µ, we have to multiply Writing these explicitly as field integrals, and anticipating that a the RHS of the first (second) line by eσβµ/2(e−σβµ/2)). The trans- change of integration variables Ψ → TβΨ leaves the functional formation Tβ is a composition of complex conjugation of the measure [Ψ] invariant [189], we find D fields ψσ and inversion of the sign of the time t — both origi- Z nating from the time reversal transformation —, and translation ∗ 0 0 ∗ 0 0 iS [Ψ] hψσ(t, x)ψσ0 (t , x )i = D[Ψ] ψσ(t, x)ψσ0 (t , x )e of the value of t by an amount iσβ/2. The latter is induced by Z the KMS condition Eq. (87). We note that the translation of t in ∗ 0 0 iS [TβΨ] = D[TβΨ] Tβψσ(t, x)Tβψσ0 (t , x )e Eq. (88) takes opposite signs depending on whether a field on the forward or on the backward branch is being transformed. As we ∗ 0 0 = hTβψσ(t, x)Tβψσ0 (t , x )i. show in Sec. II D 4 below, a similar form of time translations is (86) connected to conservation of energy: indeed, if the Keldysh ac- tion is invariant under time translations ψσ(t, x) 7→ ψσ(t + σs, x) In the last equality, we used the symmetry of the Keldysh ac- with s ∈ R, the total energy in the system is conserved. A cru- tion (85) (assuming for simplicity that there are no external cial difference from the time translation that is part of Tβ is that fields). Inserting here the explicit form of the symmetry transfor- energy conservation requires invariance under shifts by an ar- mation specified in Eq. (88) below where β = 1/T is the inverse bitrary real value s, whereas Tβ involves a shift by the purely temperature, it can be seen that Eq. (86) is in fact equivalent to imaginary value iβ/2, where β = 1/T is fixed and determined by the FDR (84)[189]. the temperature. Under which conditions does the Keldysh action (30) with L defined in Eq. (28) have the thermal symmetry, i.e., under which 12 Usually, the term “Ward-Takahashi identity” is reserved for relations that fol- conditions does it describe a system in thermal equilibrium? Let low from a continuous symmetry. Here, we use it also in the context of dis- us first consider the parts of the action corresponding to uni- crete symmetry transformations. tary time evolution, i.e., the first two terms in Eq. (30) and the 26

first line in Eq. (28). It is straightforward to check [189] that the Markov and rotating-wave approximations, the description these terms are symmetric if the Hamiltonian densities H± do of the system dynamics in terms of a quantum master equation not explicitly depend on time. On the other hand, adding an will lead to the wrong prediction that fluctuations in the system external classical driving field such as a laser, and thus break- do not obey an FDR. ing time translational invariance by making H± time dependent, A simple example that illustrates the above discussion is given 13 also the thermal symmetry is broken . The violation of the by a single bosonic mode a with energy ω0, driven coherently at thermal symmetry by classical driving fields on the level of a frequency ωl, and coupled to a bath of harmonic oscillators a microscopic Hamiltonian description indicates that quantum bµ in thermal equilibrium. The associated Keldysh action can master equations correspond to genuine non-equilibrium condi- be decomposed as S = S s + S sb + S b, where the action for the tions [187–189]. Physically, this is due to the fact that a system system consists of two parts, S s = S 0 + S l which read for which such a description is appropriate is necessarily driven. Z X ∗ In an effective description of a driven-dissipative system in S 0 = σ aσ(ω) (ω − ω0) aσ(ω), (89) terms of a Markovian master equation in a rotating frame, there σ ω is often no explicit time dependence — indeed, our deriva- and tion of the dissipative Keldysh action in Sec.IIA started from Z Eq. (16) with time-independent Hamiltonian. Even then, the X   S = Ω σ dt a (t)eiωlt + a∗ (t)e−iωlt thermal symmetry can be used to diagnose non-equilibrium con- l σ σ σ ditions [189]. Again, it is sufficient to study only the time- (90) X ∗  translation part of Tβ: for time-independent H± and Lα,± in = Ω σ aσ(ωl) + aσ(ωl) . Eq. (28), the Keldysh action is still invariant under time trans- σ lations of the form ψσ(t, x) 7→ ψσ(t + s, x). Importantly, this Ω is the amplitude of the driving field, and due to the harmonic T differs from the time translation that occurs in β by the absence time dependence of the drive it affects only the component a(ωl) of a factor of σ, meaning that t is shifted by the same amount in frequency space. The action for the bath is expressed most on the forward and backward branches. Then, by means of a conveniently in the basis of classical and quantum fields (cf. simple shift of the integration variable t in Eq. (30) the origi- Eq. (32)). In addition to the coherent part stemming from the nal form of the action can be restored. This strategy fails for oscillator frequencies, it involves infinitesimal dissipative regu- the dissipative contributions in Eq. (28) that couple forward and larization terms specifying the thermal equilibrium state of the backward branches, when the transformation involves the con- bath at inverse temperature β = 1/T as discussed in Sec. II B 1, tour index σ.14 Again we reach the conclusion that quantum Z master equations describe genuine non-equilibrium conditions, X  ∗ ∗  even though by a slightly different argument than in the driven S b = bµ,c(ω), bµ,q(ω) µ ω but purely Hamiltonian setting. ! ! 0 ω − ω − iδ b (ω) In some cases, the Markov and rotating-wave approximations × µ µ,c . (91) leading to a description in terms of a quantum master equation or ω − ωµ + iδ i2δ coth(βω/2) bµ,q(ω) equivalent Keldysh action are applied in the absence of external driving fields. Then, these approximations might still be justified Finally, the system-bath interaction with coupling strength λ cor- to study the behavior of specific observables, even though they responds to the following contribution to the action: Z explicitly break the symmetry [189]. To give an example, if one X X  ∗ ∗  attempts to study thermalization of a system due to the coupling S sb = λ σ aσ(ω)bµ,σ(ω) + aσ(ω)bµ,σ(ω) . (92) with a heat bath, and one integrates out the bath using the above- σ µ ω mentioned approximations, the resulting dynamics of the system To check whether the transformation (88) is a symmetry of the will still lead to a thermal stationary state. Correspondingly, all action even in the presence of the driving term in Eq. (90), it static properties of the system will appear thermal. However, as is most convenient to rewrite the transformation in frequency discussed at the beginning of the present section, to unambigu- space. Moreover, for future reference, we give the form of the ously prove the presence of thermal equilibrium conditions, one transformation including a chemical potential: has to consider dynamical signatures such as FDRs. Then, as a −σβ(ω−µ)/2 ∗ consequence of the explicit violation of the thermal symmetry by Tβ,µaσ(ω) = e a (ω), σ (93) ∗ σβ(ω−µ)/2 Tβ,µaσ(ω) = e aσ(ω).

13 The more precise statement is that there exists no rotating frame in which the It is straightforward to check that both S b and S sb are invariant explicit time dependence fully disappears from the problem. under this transformation with µ = 0 [189]; the same holds true 14 We note that also dissipative contributions corresponding to a system in ther- for the contribution S 0 to the action of the system. However, the mal equilibrium, as described below Eq. (68) in Sec. II B 1, couple the forward driving part (90) becomes after the transformation and backward branches. However, the special form of these terms conspire X   with the fact that the temporal shift in Tβ is determined by the inverse temper- σβωl/2 ∗ −σβωl/2 S l = Ω σ aσ(ωl)e + aσ(ωl)e , (94) ature β = 1/T to make them invariant under Tβ. σ 27 where the appearance of the exponentials shows that the sym- FDRs for cross-correlations between system and bath observ- metry is violated. One might wonder whether it is possible to ables would reveal that there is no true equilibrium in the sense restore the symmetry in a rotating frame in which the explicit of a single global temperature and chemical potential. Such time dependence of the action is eliminated, i.e., by introduc- cross-correlations, however, are suppressed in the limit λ → 0. ing new variablesa ˜σ(t) and b˜σ(t) rotating at the frequency of the In this limit, both subsystems decouple, and each of them ex- iω t driving field,a ˜σ(t) = aσ(t)e l and analogously for b˜σ(t). In hibits thermal behavior. terms of these variables, the driving part of the action becomes Z X ∗  2. Semiclassical limit of the thermal symmetry S l = Ω σ dt a˜σ(t) + a˜σ(t) σ (95) X ∗  For many applications, as discussed in Sec.IIC, the Keldysh = Ω σ a˜σ(0) + a˜σ(0) , action in the semiclassical limit (79) is appropriate. Correspond- σ ingly we should consider the semiclassical limit of the transfor- mation (88). In the limit of high temperatures β = 1/T → 0, we i.e., the drive couples to the zero-frequency component ofa ˜σ(ω). Clearly, applying again the same transformation Eq. (93) with can perform an expansion of the transformed fields, with their arguments shifted by iσβ/2, in terms of derivatives. To leading µ = 0 to the new variables, the driving term, as well as S 0 in order,16 this yields Eq. (89) and the system-bath coupling S sb in Eq. (92) are in- variant. However, in the action for the bath (91), as a conse- T Φ (t, x) = σ Φ (−t, x), quence of the transformation to the rotating frame the distribu- β c x c   tion function acquires an effective chemical potential and be- i (97) TβΦq(t, x) = σx Φq(−t, x) + ∂tΦc(−t, x) , comes coth(β(ω − ωl)/2). Hence, to leave this part of the ac- 2T tion invariant, the transformation Eq. (93) has to be applied with where σ is the usual Pauli matrix. “High temperatures” thus µ = ω , and again the full action is not invariant under the equi- x l means that the typical frequency scale of the field is much librium transformation. No frame exists in which the reference smaller than temperature; indeed this recovers the intuition of to the driving scale ω were eliminated. l a semiclassical limit. If we replace the quantum field Φ by While this simple example demonstrates explicitly that gener- q the response field Φ˜ = −iσ Φ , Eq. (97) takes the form of the ically external driving takes a system out of thermal equilib- z q classical symmetry introduced in Ref. [207]. The FDR in the rium, and how this is manifest in the symmetry properties of semiclassical limit, which may be derived as a Ward-Takahashi the Keldysh action, it is interesting to note that there are surpris- of the symmetry Eq. (97), simplifies to the Raleigh-Jeans form ing exceptions to this rule, emerging as limiting cases. One of them has been identified in Ref. [211]. Along the lines of this 2T   GK(ω, q) = GR(ω, q) − GA(ω, q) . (98) reference, we discard the driving term S l and instead consider a ω parametric system-bath coupling of the form15 The thermal FDR is just one consequence of the presence of Z X X   the symmetry Eq. (97). A second one concerns the possible val- S = λ σ dt a∗ (t)b (t)e−iωpt + a (t)b∗ (t)eiωpt sb σ µ,σ σ µ,σ ues of the coupling constants defining the action in the semiclas- σ µ sical limit. This allows us to state precisely in which sense the X X Z  ∗ ∗  driven-dissipative systems represent a genuine non-equilibrium = λ σ aσ(ω + ωp)bµ,σ(ω) + aσ(ω + ωp)bµ,σ(ω) , µ σ ω situation. To this end, it is most convenient to discuss the equiv- (96) alent Langevin equation Eq. (80). We rewrite it by splitting the deterministic parts on the RHS into reversible (coherent) and ir- in the limit λ → 0. Then, the full action S = S 0 + S sb + S b, reversible (dissipative) contributions according to (here we re- where S 0 and S b are as above (Eqs. (89) and (91)), is invariant if place φc = ψ) the system fields are transformed with Tβ,µ=ωp and the bath oscil- δHc δHd lators with Tβ,µ=0. In other words, the parametric coupling acts − i∂tψ = ∗ i ∗ + ξ (99) to thermalize the system at the temperature of the bath while at δψ δψ ω the same time shifting the chemical potential by p with respect with effective coherent and dissipative Hamiltonians (α = c, d) to the bath chemical potential. Concomitantly, the FDR for the system variables takes the form of Eq. (84) with ω → ω − ω , Z   p 2 2 uα 4 while for the bath it is Eq. (84) without modification, and only Hα = Kα |∇ψ| + rα |ψ| + |ψ| . (100) t,x 2

15 16 In realistic systems the system-bath coupling usually also contains terms of the Note that a contribution ∼ ∂tΦq in the first line is suppressed according to form aσ(t)bσ(t) that are neglected in the rotating-wave approximation [211]. canonical power counting, cf. Sec.IIC, Eq. (78). 28

It can be shown [181, 212] that the presence of the symme- metry is present on the microscopic scales on which the quantum try (97), or, in physical terms, relaxation of a system to ther- master equation (16) or the corresponding Keldysh action (27) modynamic equilibrium (a state with global detailed balance, provide a suitable description of the system, in several cases it where arbitrary subparts are in equilibrium with each other) re- has been found that driven-dissipative systems appear as approx- quires the condition imately thermal at low frequencies [25, 33, 54, 110, 112, 122, 191, 213, 214]. In driven-dissipative condensates in three spa- Kc uc Hc = rHd ⇔ r = = . (101) tial dimensions, this thermalization at low frequencies is partic- Kd ud ularly sharply reflected via the emergence of the thermal sym- metry in the RG flow in this regime [26, 181], cf. Fig.6 (c); this (Note that there is no condition on the ratio rc/rd since the ef- is discussed in more detail in Sec.IV. On the other hand, there fective chemical potential rc can always be adjusted by a gauge −iωt are also cases where the opposite behavior occurs, and the non- transformation ψ 7→ ψe such that rc 7→ rc − ω without chang- ing the physics.) In equilibrium dynamics, the ratio of real (re- equilibrium character becomes more pronounced as one coarse versible) and imaginary (dissipative) parts is thus locked to one grains to the macroscale. This occurs in driven two dimensional common value for all couplings. This is illustrated in Fig.6 (a). systems, as explained in Sec.IV. In the complex plane spanned by real and imaginary parts of the couplings K = Kc + iKd and u = uc + iud, they lie on one single ray. The intuition behind this seeming fine-tuning is the follow- 3. The Noether theorem in the Keldysh formalism ing: a microscopically reversible dynamics starts from a Hamil- tonian functional Hc alone, i.e., all couplings are located on the A global continuous symmetry of the Keldysh action is a ∗ ∗ T real axis. Coarse graining the system from the microscopic to transformation Tα of the fields Ψ = ψ+, ψ+, ψ−, ψ− that leaves the macroscopic scales introduces irreversible dynamics in the the value of the action invariant, i.e., form of finite imaginary parts, however preserving their loca- S [TαΨ] = S [Ψ], (102) tion on a single ray: the ray just rotates under coarse graining, but does not spread out. The geometric constraint is thus not where α is a real time- and space independent parameter, and for 1 due to fine tuning, but results from the microscopic “initial con- α = 0 the transformation is the identity, T0 = . The Noether dition” for the RG flow, in combination with the presence of a theorem states that any such global continuous symmetry entails µ symmetry. In stark contrast, in a driven non-equilibrium system, the existence of a current j with components j , which obeys a µ the microscopic origins of reversible and irreversible dynamics continuity equation on average, i.e., h∂µ j i = 0 (here and in the are independent, as illustrated in Fig.6 (b). For example, in the following summation over repeated indicies is implied), where microscopic description of Eq. (73) the rates can be tuned fully ∂0 = ∂t, and ∂1,2,...d are derivatives with respect to spatial coor- R 0 independently from the Hamiltonian parameters — they have dinates. Then, the integral over space Q = x j — the Noether completely different physical origins. charge — is an integral of motion, and we have hdQ/dti = 0. In the following, we prove this relation, which states that Q is con- (a) (b) (c) served on average, in the framework of the Keldysh formalism. We note however, that a global continuous symmetry implies the Im Im Im u u even stronger statement dQ/dt = 0 of conservation of Q on the operator level [86]. u In order to prove the Noether theorem, it is sufficient to con- K ⇤ K K sider infinitesimal transformations. Then, for α  1 we expand ⇤ the transformation as Re Re Re 2 equilibrium non-equilibrium fixed point TαΨ = Ψ + αGΨ + O(α ), (103) where G is called the generator of the transformation. In general, Figure 6. Location of the couplings in the Langevin equation (99) in the G × complex plane. Real (imaginary) parts describe reversible (irreversible) is a 4 4 matrix with derivative operators as entries. We contributions to the dynamics. (a) An equilibrium system is charac- consider specific examples below in Sec. II D 4. In the following terized by the location of couplings on a single ray, reflecting detailed we assume that as in Eq. (30) the action SR can be written in balance. The irreversible dynamics is not independent of the underlying terms of a Lagrangian density L as S = t,x L , and that the reversible Hamiltonian dynamics, but rather generated by it. (b) In con- Lagrangian density is a local function of the fields and their first trast, in a driven non-equilibrium system, generically there is a spread derivatives with respect to time and space. This assumption is in the location of couplings, because the reversible and dissipative dy- appropriate for most practical purposes. namics have different physical origins. (c) Near a critical point in three In an expansion of the LHS of Eq. (102) in powers of α, each dimensions, the couplings flow strongly with scale and approach the coefficient has to vanish individually. The first-order contribu- imaginary axis (decoherence). The RG fixed point is purely dissipative. tion yields the relation (Figure adapted from [26].) Z ! ∂L ∂L G G T Ψ + T ∂µ Ψ = 0, (104) Finally, we note that, while an explicit violation of the sym- t,x ∂Ψ ∂∂µΨ 29

∗ ∗ T where ∂L /∂Ψ ≡ ∂L /∂ψ+, ∂L /∂ψ+, ∂L /∂ψ−, ∂L /∂ψ− . translationally invariant Hamiltonian alone. Indeed, the con- In the cases we consider, Eq. (104) holds true because the in- struction of the Keldysh action for a Markovian master equation tegrand can be written as the divergence of a vector field f µ, i.e., in Sec.IIA shows that terms which couple the forward and back- we have ward branches are only contained in the dissipative parts of the action (30). In other words, they arise upon coupling the system ∂L ∂L ∂ f µ = GΨ + ∂ GΨ. (105) to a bath and integrating out the latter. On the other hand, the µ ∂ΨT ∂∂ ΨT µ µ Keldysh action for a closed system with unitary dynamics does To proceed we consider local transformations, i.e., we con- not contain such terms and can be written as sider Tα with α = α(t, x). We perform a change of integration X Z variables Ψ → TαΨ in the partition function. Then, assuming S = σ Lσ, (111) that the functional measure is invariant with respect to the local σ t,x transformation, we have Z Z where L+ and L− are the Lagrangian densities evaluated with Z = D[Ψ]eiS [Ψ] = D[Ψ]eiS [TαΨ]. (106) fields on the forward and backward branches, respectively. Given this structure of the action, let us consider a transforma- As before, we expand the RHS of this equality in a power se- tion Tα which does not mix fields on the forward and backward G ries in α. Since by assumption the latter is a function of (t, x), branches. Then, the generator has the block-diagonal struc- Eq. (102) does not hold true any more. Instead, we find ture Z ! ! ∂L g+ 0 µ − G 2 G = . (112) S [TαΨ] = S [Ψ] + α∂µ f T Ψ + O(α ), (107) 0 g− t,x ∂∂µΨ where we used Eq. (105) and integration by parts to write the The cases of physical interest are the classical and quantum RHS in a form that does not contain derivatives of α explicitly. transformations mentioned above, corresponding to the choices Inserting Eq. (107) in the exponential on the RHS of Eq. (106), g+ = g− and g+ = −g−, respectively. In both cases, the Noether and expanding the latter to first order in α, we obtain the condi- current can be represented as superposition of currents on the tion forward and backward branches, which we define in terms of Z * + g ≡ g+ as ∂L µ − G α∂µ f T Ψ = 0. (108) ∂∂µΨ ∂Lσ t,x µ g − µ jσ = T Ψσ fσ , (113) Since there are no restrictions on the choice of α, this equation ∂∂µΨσ implies that the divergence of the expectation value vanishes. where The latter is just the quantum Noether current, ∂Lσ ∂Lσ ∂L ∂ f µ = gΨ + ∂ gΨ . (114) jµ = GΨ − f µ. (109) µ σ T σ T µ σ T ∂Ψσ ∂∂µΨσ ∂∂µΨ

Note that jµ is not unique: for constant a, bµ, the combination With this definition, j− can be obtained from j+ simply by re- a jµ + bµ is also a conserved current. The associated Noether placing all instances of Ψ+ appearing in j+ by Ψ−. Then, the charge Q corresponds to the integral over space of the zeroth symmetry of the action under a quantum transformation yields component of the Noether current, a classical Noether current and vice versa. These currents are Z Z ! given by (as noted above we are free to choose a convenient ∂L 0 G − 0 multiplicative normalization of the currents) Q = j = T Ψ f . (110) x x ∂∂tΨ 1 j = ( j + j ) , j = j − j . (115) While this derivation is quite general, a more concrete expres- c 2 + − q + − sion for the Noether current can be obtained by restricting the form of the generator G in Eq. (103). In particular, in the next As pointed out in Sec. II B 2, due to causality only the classi- section we consider the specific cases of classical and quantum cal component of the field can acquire a finite expectation value. transformations. The same is true for the currents defined in Eq. (115): in the presence of an external field that induces a current, we have h j+i = h j−i = h jci , 0, whereas h jqi = 0. Hence, as anticipated µ 4. Closed systems: energy, momentum and particle number above, the continuity equation h∂µ jqi = 0, which follows from conservation the symmetry under a classical transformation, does not entail a non-trivial Noether charge. In the next section, we show that Energy and momentum conservation – This property is ex- such a symmetry nevertheless does have physical consequences, pected for systems, whose classical action is determined by a as it can be broken spontaneously in a condensation transition. 30

Before, let us derive the Noether charge associated with spatial components of the current are given by the symmetry of a Hamiltonian Keldysh action under contour- 1 dependent — i.e., quantum – translations in space-time. To j = ψ∗ ∇ψ − ψ ∇ψ∗  , (119) σ i2m σ σ σ σ be specific, we consider the transformation Ψσ(X) 7→ Ψσ(X + σαeν), where X = (t, x), and eν is the basis vector in direc- which is just the ordinary quantum mechanical current for par- tion of the νth coordinate of d + 1-dimensional space-time. This ticles with quadratic dispersion relation, evaluated on the closed is a symmetry of the action (111): the shift of the coordinates time path. What is the physical meaning of the classical and can be absorbed by performing a change of integration variables quantum components, which are defined in Eq. (115), of the cur- X → X − σαeν in the integral over the Lagrangian density Lσ. rent given in Eq. (119)? If we introduce in the Hamiltonian an Clearly, the symmetry is violated in the presence of dissipative externally imposed gauge field, in the Keldysh action it would terms that couple the forward and backward branches. couple to the quantum current, in analogy to the source terms in For space-time translations, the generators g± in Eq. (112) ac- the generating functional Eq. (33). However, the observable ef- quire an additional index ν and are given by g±,ν = ±∂ν. It fect of such a gauge field is that it can induce a classical current µ follows that the vector field fσ,ν in Eq. (114) takes the form h jci 0. µ µ , fσ,ν = δν Lσ, and the classical Noether current, which we de- We close this section with a comment on the status of the note as Tc,ν, is given by Uq(1) symmetry. It will be present on the microscopic level for an action corresponding to a Hamiltonian which commutes with X µ 1 µ µ ∂Lσ µ the number operator. Since the symmetry considerations are T = T , T = ∂νΨσ − δ Lσ, (116) c,ν 2 σ,ν σ,ν ∂∂ ΨT ν σ µ σ properly applied to the microscopic action (and the functional measure), the conservation law ensues. However, this does not µ where T±,ν are the components of the energy-momentum ten- imply that the effective action must manifestly preserve a Uq(1) sor [86, 205] on the forward and backward branches. In particu- symmetry. In fact, classical models of number conserving dy- 0 0 lar, the component T±,0 is the energy density, whereas T±,i is the namics [1] do not exhibit this symmetry. This leads to the pic- density of momentum in the spatial direction i = 1, 2,..., d. The ture that this symmetry is broken spontaneously under RG. The spatial integrals over these densities yield the associated Noether precise workings of such a mechanism are, to the best of our charges, e.g., if we express the Lagrangian density Lσ in terms knowledge, not settled so far. of a Hamiltonian density (as in Eq. (30), where L is given by Eq. (28) with γα = 0) as 5. Extended continuity equation in open systems 1 † Lσ = Ψσiσz∂tΨσ − Hσ, (117) 2 The conservation of particle number in a closed system fol- lows from the continuity equation for the classical Noether cur- we obtain the conserved energy density E , which is as expected rent that is associated with the symmetry under quantum phase given by the sum of the Hamiltonian densities on the forward rotations. In an open system, this symmetry is absent, and the and backward branches, continuity equation has to be extended to account for the ex- Z Z change of particles with the environment, as we discuss in the 1 X 0 1 E = T = (H + H−) . (118) 2 σ,0 2 + following. σ x x To be specific, we consider the Keldysh action (73) for a sys- Along the same lines, conservation of angular momentum fol- tem with single-particle pump as well as single-particle and two- lows from the quantum symmetry of the Keldysh action with body loss. In order to derive the extended continuity equation, respect to rotations of the spatial coordinates. as in Eq. (106) we perform a change of integration variables → Number conservation – As another example, we consider Ψ UqΨ, where Uq is a local quantum phase rotation, in the phase rotations and their relation to particle number conser- Keldysh partition function. Since the partition function is invari- ant under this transformation, the coefficients in an expansion of vation. In this case the classical transformation reads ψ± 7→ iα the partition function in powers of the phase shift must vanish. Ucψ± = e ψ± and the quantum transformation is ψ± 7→ Uqψ± = ±iα In linear order, we find the condition e ψ±. Classical phase rotations are a symmetry of the Keldysh ∗ action, if the fields appear only in the combinations ψ ψσ0 . The µ ∗ ∗ ∗ 2 σ h∂µ jc i − γphψ ψ−i + γlhψ−ψ+i + 4udh ψ−ψ+ i = 0. (120) quantum transformation is more restrictive: in this case only + 0 products with σ = σ are allowed. Hence, the symmetry with This should be compared to Eq. (108): if Uq were a symme- regard to quantum phase rotations implies the classical symme- try of the action, we would have found an ordinary continuity try. The generators of Uq are given by g± = ±iσz. Then, with equation. This would have been the case in the absence of the Eq. (10) and inserting the representation of the Lagrangian den- pumping and loss terms proportional to γp, γl, and ud. The inter- sity in Eq. (117) in the expression for the Noether current (113), pretation of these terms is as follows: in a system that is perfectly T we find that the latter can be written as jσ = (ρσ, jσ) , where isolated from its environment, the temporal change of the den- 2 ρσ = |ψσ| is the density on the contour with index σ and the sity at a given point is due to the motion of particles towards to 31 or away from this point, i.e., the rate of change of density is a can be detected by checking whether the determinant of the mass source or sink for the mass flow as measured by the divergence matrix M vanishes. The latter is determined by the inverse prop- of the current h∇ · jci. In an open system, on the other hand, agator there is in addition a dissipative current [215, 216] due to the ! exchange of particles between the system and its environment. 0 PA(ω, q) P(ω, q) = = G−1(ω, q), (122) In particular, in Eq. (120) this dissipative current has contribu- PR(ω, q) PK(ω, q) tions from the pumping and loss terms. It is interesting to note that Eq. (120) can also be ob- given exemplarily in Eq. (76), at ω = q = 0, i.e., M = −P(0, 0). tained from the equation of motion of the local density n(x) = Therefore, it is sufficient to consider frequency- and momentum- ψ†(x)ψ(x) in the operator formalism. From the master equation independent field configurations or, in other words, homoge- ∂tρ = Lρ with Liouvillian L in Lindblad form given in Eq. (9), neous field configurations, so that instead of the full effective it follows that the time evolution of n(t, x) in the Heisenberg rep- action Eq. (39) we only have to deal with the effective potential ∗ resentation is given by ∂tn(t, x) = L n(t, x) with the adjoint Li- U = − Γ|hom. /Ω, where Ω is the quantization volume in space- ouvillian defined by tr(ALB) = tr(BL∗A). We find time. Since the effective potential U inherits the symmetries of

† the effective action, it is a function of precisely these combina- ∂tn(t, x) = −∇ · j(t, x) + γpψ(t, x)ψ (t, x) tions of fields which are invariant under classical or quantum † 2 2 − γln(t, x) − 4udψ (t, x) ψ(t, x) , (121) phase rotations. In the basis of classical and quantum fields, ∗ these Uc-symmetric combinations are ρνν0 = φνφν0 , where the where first term encodes coherent dynamics and corresponds to indices ν and ν0 can take the values c and q. In the following, we the Heisenberg commutator i[HLP, n(t, x)], whereas the remain- show that Uc invariance is sufficient to guarantee the existence ing contributions incorporate the dissipative parts. Taking the of a massless mode. average of this relation and expressing the expectation values as For convenience we switch to a basis of real fields χ = Keldysh functional integrals yields Eq. (120). (χc,1, χc,2, χq,1, χq,2) which correspond to the real and imagi- Finally, we can obtain some intuition on the dissipative cor- nary parts of the complex classical and quantum fields, i.e., rection to the current expectation value in mean field theory, 1  φν = √ χν, + iχν, for ν = c, q. In this basis, the mass ma- where the correlators in Eq. (120) factorize into products of sin- 2 1 2 trix reads gle field expectation values hψ+i = hψ−i = ψ0. We then recog- ∗ nize in the non-derivative terms (ψ times) the LHS of Eq. (75) 2 " 2 # " 2 # √ 0 ∂ U X ∂ ρa ∂U X ∂ρa ∂ρb ∂ U (note that due to the factor of 1/ 2 in the Keldysh rotation (32) Mi j = = + , √ ∂χi∂χ j ∂χi∂χ j ∂ρa ∂χi ∂χi ∂ρa∂ρb ss the expectation values are related as hφ i = φ = 2 hψ i, and ss a ss a,b c 0 ± (123) that r = (γ − γ )/2), which equals zero in a homogeneous sit- d l p where the subscript ss indicates that the fields should be set to uation within mean field theory. In this way, we see that there is their average values in the stationary state. The indices i and no particle number current on average in a homogeneous driven- j label the components of the four-vector χ defined above (i.e., dissipative system in stationary state, as expected. χ1 = χc,1, χ2 = χc,2 etc.), and a and b are double indices, taking the values cc, cq, qc, qq. Let us consider the first term on the 6. Spontaneous symmetry breaking and the Goldstone theorem RHS of Eq. (123): in the ordered phase, the classical field has a finite expectation value in the stationary state. Without loss of generality we assume that this value is real. Then, the field Even though the pumping and loss terms in the Keldysh ac- equations tion in Eq. (73) break the symmetry with respect to quantum phase rotations Uq, the action is still symmetric under Uc trans- " # ∂U X ∂ρa ∂U formations. As a result, even in the absence of particle num- = = 0, (124) ∂χ ∂χ ∂ρ ber conservation there is a possibility of spontaneous symmetry i ss a i a ss breaking. In particular, a finite average value hφci , 0 breaks the which actually determine the average value of the fields χ in classical phase rotation symmetry in a non-equilibrium conden- p i sation transition. Here we show that this spontaneous symmetry stationary state, have the solution χi|ss = 2ρ0δi,1. Performing breaking is accompanied by the appearance of a massless mode, the derivatives ∂ρa/∂χi in Eq. (124) explicitly and inserting χi|ss, i.e., a mode with vanishing frequency and decay rate for zero we obtain the following conditions: momentum, which is known as the Goldstone boson [217–219]. ∂U ∂U ∂U We obtain this result by carrying over the corresponding deriva- = = = 0. (125) tion from the equilibrium formalism (see, e.g., Ref. [182]) to the ∂ρcc ss ∂ρcq ss ∂ρqc ss Keldysh framework. The single-particle excitation spectrum is encoded in the poles Therefore, only the derivative uqq = [∂U/∂ρqq]ss contributes of retarded and advanced Green’s functions, and such a pole at to the first term on the RHS of Eq. (123). Denoting mixed ω = q = 0 is dubbed a massless mode. Poles of the Green’s second derivatives of the effective potential as uab = uba = 2 functions correspond to roots of the determinant of inverse prop- [∂ U/∂ρa∂ρb]ss, we find that the mass matrix can be written as agator (see Sec.IIA). Hence, the presence of a massless mode 32

    0 0 0 0   0 0 ucc,cq + ucc,qc i ucc,cq − ucc,qc         0 0 0 0  0 0 0 0        M =   + ρ0  1 i  . (126) 0 0 uqq 0   ucc,cq + ucc,qc 0 ucq,cq + 2ucq,qc + uqc,qc ucq,cq − uqc,qc       2    2  0 0 0 uqq  i 1  i ucc,cq − ucc,qc 0 2 ucq,cq − uqc,qc − 2 ucq,cq − 2ucq,qc + uqc,qc

An entry ucc,cc is not ruled out by symmetry, however, it must mulation. This approach offers the particularly attractive feature vanish due to conservation of probability, cf. Sec. II B 1. Cru- of not being restricted to the critical point. cially, the retarded and advanced sectors feature one zero eigen- The functional renormalization group equation (for reviews value. This proves the existence of a massless mode as a conse- see [229–233]) constitutes an exact reformulation of the func- quence of spontaneous symmetry breaking for a theory with Uc tional integral representation of a quantum many-body prob- invariance. lem in terms of a functional differential equation. In this, it is While this analysis guarantees the existence of a zero mode strongly distinct from, e.g., perturbative field theoretical renor- (i.e., the complex dispersion relation has the property ω(q = malization group approaches, which concentrate exclusively on 0) = 0), it does not provide a statement on the functional de- the critical surface of a given problem. Instead, it may be viewed pendence ω(q). The latter could be inferred along the lines of as an alternative and potentially more tractable tool for the anal- Refs. [220–224]. As we found in Sec. II B 2 in Bogoliubov ap- ysis of the complete many-body problem, also on length scales proximation, in an open system without (i.e., broken) Uq sym- well below the correlation length near criticality. Indeed, it has metry and spontaneously broken Uc symmetry, the leading be- proven a very versatile tool in many different physical context, havior at low momenta is diffusive, ω(q) ∼ −iDq2, with a real ranging from quantum dots [234–236], ultracold atoms [233], diffusion coefficient D. This has to be contrasted to closed sys- strongly correlated electrons [237], classical stochastic mod- tems, in which microscopically both Uq and Uc symmetry are els [238, 239], quantum chromodynamics [232, 240], to quan- present. There, the leading behavior in a phase with sponta- tum gravity [241]. Here we give a brief overview of the gen- neously broken Uc symmetry is that of coherent sound waves, eral concept adapted to non-equilibrium systems [26, 181, 234– ω(q) ∼ cq, with real speed of sound c. A phenomenological 236, 242–254]. It is used for the discussion of critical behavior justification of this behavior is given in Ref. [1], where this can in driven open quantum systems in Sec.IV, with applications be understood as a consequence of a coupling to additional slow to a broader non-equilibrium many-body context left for future modes relating to particle number conservation. In these phe- work. nomenological models, Uq symmetry is absent. This suggests a The transition from the action S to the effective action Γ con- scenario of an additional spontaneous breakdown of Uq symme- sists in the inclusion of both statistical and quantum fluctuations try upon coarse graining, but this issue seems not to be settled to into the latter (cf. Eq. (39)). In the functional renormalization date. group approach based on the Wetterich equation [228], the func- tional integral over fluctuations is carried out stepwise. To this end, an infrared regulator is introduced, which suppresses the E. Open system functional renormalization group fluctuations with momenta less than an infrared cutoff scale Λ. This is achieved by adding to the action in (33) a term In Sec.IVB, we investigate dynamical criticality of the Z ! ! Bose condensation transition in driven open systems, moti-  ∗ ∗ 0 RΛ φc ∆S Λ = φc, φq ∗ , (127) vated by many-body ensembles such as exciton-polariton con- t,x RΛ 0 φq densates [12, 13, 159, 192, 225]. There again, coherent dy- namics naturally competes with dissipation in the form of in- with a cutoff or regulator function RΛ. Some key structural prop- coherent particle losses and pumping. The situation parallels a erties are indicated below, but apart from these properties the laser threshold [226, 227], however with a continuum of spa- choice of the cutoff is flexible and problem-specific. The re- tial degrees of freedom. This ingredient, however, causes the sulting cutoff-dependent generating functional and its logarithm characteristic long-wavelength divergences of many-body prob- (cf. Eq. (35)) are denoted by, respectively, ZΛ and WΛ. Then, the lems in their symmetry broken phase, or at a critical point. It scaled-dependent effective action ΓΛ is defined by modifying the implies that perturbation theory necessarily breaks down, even Legendre transform in Eq. (38) according to when the interaction constants are small, due to a continuum of ¯ ¯ modes without a gap, which form the intermediate states and ΓΛ[Φc, Φq] = WΛ[Jc, Jq] Z are summed over in many-body perturbation theory. This calls  † ¯ † ¯  ¯ ¯ − Jc Φq + Jq Φc − ∆S Λ[Φc, Φq]. (128) for the development of efficient functional integral techniques t,x able to cope with these problems. Our method of choice is the functional renormalization group based on the Wetterich equa- The subtraction of the ∆S Λ on the RHS guarantees that the only tion [228], which we briefly introduce here in its Keldysh for- difference between the functional integral representations for Γ 33 and ΓΛ is the inclusion of the cutoff term in the latter, in terms canonical scaling dimensions, keeping only those cou- plings which are — in the sense of the renormalization group — Z ¯ ¯ ¯ ¯ relevant or marginal at the phase transition, cf. the discussion in eiΓΛ[Φc,Φq] = [δΦ , δΦ ] eiS [Φc+δΦc,Φq+δΦq] D c q Sec.IIC, and see Sec.IVB for applications. δΓ δΓ −i ¯ T δΦq−i T δΦc+i∆S Λ[δΦc,δΦq] × e δΦc δΦq . (129) Part 2 Physically, ΓΛ can thus be viewed as the effective action for av- erages of fields over a coarse-graining volume with a size ∼ Λ−d, Applications where d is the spatial dimension. Note that we chose the form of the cutoff action ∆S such Λ III. NON-EQUILIBRIUM STATIONARY STATES: SPIN that it modifies the inverse retarded and advanced propagators MODELS in Eq. (127) only. This is sufficient to regularize possible in- frared divergences, which result from poles of the retarded and advanced propagators being located at the origin of the complex In this section, we discuss the steady state properties of many- frequency plane. A typical choice in practical calculations is body systems consisting of atoms, which are coupled to the radi- ation field of a cavity, in turn subject to dissipation in the form of 2 2 permanent photon loss. The corresponding low frequency field RΛ(q ) ∼ Λ , q/Λ → 0, (130) theory, a 0 + 1-dimensional path integral for real valued, Ising giving the inverse propagators a mass ∼ Λ2. In this way, fluctu- type fields corresponds to the simplest, non-trivial field theoretic ations with wavelength & Λ−1 are effectively cut off. Therefore, models and is therefore particular useful to get used to applica- for any finite Λ, the technical problem of infrared divergences is tions of the Keldysh formalism for relevant physical setups. The under control. basic model describing the dynamics of atoms in a cavity is the The main usefulness of the so-modified effective action, how- Dicke model (1) with dissipation, as described in Sec.IC1, as ever, lies in the fact that it smoothly interpolates between the well as its extension to multiple cavity photon modes. Despite action S for Λ → Λ0, where Λ0 is the ultraviolet cutoff of the the simplicity of the underlying field theory, it is a non-trivial problem (e.g., the inverse lattice spacing), and the full effective task to solve it for its rich many-body dynamics. This includes action Γ for Λ → 0. This is ensured by the following require- the critical behavior of a single mode cavity at the superradi- ments on the cutoff [244]: ance transition as well as universal dynamics in the formation of spin glasses in multi-mode cavities. We discuss these and fur- 2 2 RΛ(q ) ∼ Λ0, Λ → Λ0, ther features of the Dicke model here by putting a focus on the (131) theoretical framework of solving the corresponding Ising model R (q2) → 0, Λ → 0. Λ on the Keldysh contour.

Under the condition that Λ0 exceeds all energy scales in the ac- tion, for Λ → Λ we may evaluate the functional integral (129) 0 A. Ising spins in a single-mode cavity in the stationary phase approximation. Then, we find to lead- ing order ΓΛ0 ∼ S — in the absence of fluctuations (suppressed by the cutoff mass gap ∼ Λ2), the effective action approaches As for the equilibrium path integral, the Keldysh field theory 0 for spin models which obey the standard Ising Z symmetry, is the classical, or microscopic one. The evolution of ΓΛ from this 2 starting point in the ultraviolet to the full effective action in the formulated in terms of real fields, fluctuating in time and space. infrared for Λ → 0 is described by an exact flow equation — the These models have become rather important in the field of quan- Wetterich equation [228] — which was adapted to the Keldysh tum optics, where the typical situation consists of a set of atoms, framework in [234, 235, 244, 246]. It reads modeled as two-level systems, coupled to the radiation field of a high finesse cavity. One important model in this context is the i  −1  Dicke model, which has been introduced in Eq. (1) as the generic ∂ Γ = Tr Γ(2) + R ∂ R , (132) Λ Λ 2 Λ Λ Λ Λ model for cavity QED experiments with strong light-matter cou- pling. Its Hamiltonian is (2) where Γ and RΛ denote, respectively the second variations of Λ N N the effective action and the cutoff action ∆S Λ. As anticipated † ωz X z g X x  †  H = ωpa a + σ + √ σi a + a . (133) above, the flow equation provides us with an alternative but fully 2 i N equivalent formulation of the functional integral (129) as a func- i=1 i=1 tional differential equation. Like the functional integral, the flow The Dicke Hamiltonian represents an effective model which de- equation can not be solved exactly for most interesting prob- scribes the dynamics of a set of two-level atoms, labeled by the lems. It is, however, amenable to numerous systematic approxi- index i, inside a cavity, which is pumped by a transverse laser. mation strategies. For example, in the vicinity of a critical point In this scenario, the coupling constant g describes the scattering it is possible to perform an expansion of the effective action ΓΛ of laser photons into the cavity, as well as the reverse process, 34 and therefore rotates in time with the laser frequency ωl. In a ro- regimes for which both approaches can be applied, they are com- tating frame, this time dependence is gauged away in the Hamil- pletely equivalent, as we demonstrate. tonian, resulting in an effective shift ωp = ωc − ωl of the bare cavity frequency ωc. In addition to this external drive, the cavity is subject to permanent photon loss, due to imperfections in the 1. Large-spin Holstein-Primakoff representation cavity mirrors. The photon loss is described by a weak coupling of the intra-cavity photons to the environment, i.e., the vacuum The single-mode Dicke Hamiltonian (133) has the property radiation field outside the cavity. For typical experimental pa- that all atomic variables couple to the cavity photon mode via rameters, this coupling, which represents the rate with which the same coupling constant g. Consequently, the coupling of the intra-cavity field and the environment exchange photons, is the photons to the sum of the individual Pauli matrices can be much smaller than the typical relaxation time of the environment replaced by the coupling of the cavity photon mode to a large and the latter can be traced out under the Born and Markov ap- spin, proximation. This results in a Markovian master equation for the system’s density matrix, which reads (cf. Eqs. (2,16) repeated † z 2g x  †  H = ωpa a + ωzS + √ S a + a , (136) for convenience) N where S z, S x are spin operators in a spin-N/2 Hilbert space. ∂tρ = −i[H, ρ] + Lρ. (134) In order to find a path integral representation for the Hamilto- In this equation, ρ is the density matrix, H is the Dicke Hamilto- nian (136), these spin operators can be transformed into bosonic nian (133) and L is the dissipative Liouvillian, acting as on the operators via the common Schwinger-boson or Holstein- density matrix as Primakoff transformations. For weak coupling g, the system re- ! mains strongly polarized |hS zi|  1 and the Holstein-Primakoff 1 Lρ = κ aρa† − {a†a, ρ} (135) transformation around the non-interacting ground state, which 2 has the eigenvalue S z = −N/2, is the most natural choice. It reads with loss rate κ for the cavity photons. N  p   p  Given the Markovian master equation (134), there are two S z = b†b − , S x = N − b†b b + b† N − b†b (137), common ways to derive a corresponding path integral descrip- 2 tion for the dissipative Dicke model, which are based on different where b, b† are bosonic operators. In the thermodynamic limit representations of the atomic degree of freedom in terms of field N → ∞, the square root in the S x operator is expanded in pow- variables: the representation of the atomic degrees of freedom ers of 1/N, and the Hamiltonian is subsequently formulated on as a collective spin and subsequent bosonization in terms of a the Keldysh contour. To zeroth order in 1/N, the corresponding Holstein-Primakoff transformation [122, 123, 128], and the rep- Keldysh action is resentation of the atomic variables in terms of individual Ising fields. We discuss both approaches in the following, putting a   −1  Z  0 GA  focus on the more general Ising representation, which can also †  4×4  S = Φ (ω)   −1  Φ(ω), (138) ω  R K  be applied in the case of multiple cavity modes and individual G4×4 Σ4×4 atomic loss processes, where a representation of the atoms in terms of a single, collective spin is no longer possible. In the with the combined Nambu-Keldysh spinor

 ∗ ∗ ∗ ∗ T Φ(ω) = ac(ω), ac(−ω), bc(ω), bc(−ω), aq(ω), aq(−ω), bq(ω), bq(−ω) , (139) the inverse retarded Green’s function in Nambu space

 ω − ω + iκ 0 −g −g   p  −1    R   0 −ω − ωp − iκ −g −g  G4×4 =   (140)  −g −g ω − ωz 0    −g −g 0 −ω − ωz

and the Keldysh self-energy Due to the photon decay terms ∼ κ and the atom-photon cou- pling g, the above theory is regularized even without infinites- imal imaginary contributions in the atomic sector. When inte- ΣK = 2iκ diag(1, 1, 0, 0). (141) 35 grating out the photons, the latter infinitesimal contributions are overwritten in any case by the finite imaginary part of the photon Green’s function, and it is therefore reasonable to leave them out from the start. The excitation spectrum of the atom-photon system is en- coded in the retarded Green’s function, and the poles, marking the excitation energies, fulfill the requirement

!  R −1 2 2 2 2 2 0 = det G4×4 = (ω − ωz )[(ω + iκ) − ωp] − 4g ωpωz. (142) Figure 7. Illustration of the pole structure of the system’s eigenmodes. In the absence of atom-photon coupling, i.e., for g = 0, these The poles represent the location of the mode frequencies ω in the com- are the non-interacting poles ω1,2 = ±ωz, corresponding to the plex plane, the real part Re(ω) represents the energy of the mode, while atomic transition frequencies and ω3,4 = ±ωp+iκ, corresponding the imaginary part Im(ω) is the mode’s decay rate. Without interactions, ± to the energy of a photon ωp and its decay rate κ. For g > 0, the g = 0, the system is described by the bare atomic, ω = ωz, and pho- modes begin to hybridize and the excitation energies are slightly tonic modes ω = ±ωp−iκ. For finite interactions, the atoms and photons modified compared to their non-interacting values, see Fig.7. hybridize and form polaritonic modes, two of which move closer to the For small coupling strength, the elementary excitations can still imaginary axis. At the critical point g = gc, one single mode becomes critical at ω = 0 and the system undergoes a phase transition from a be seen as weakly dressed atoms and photons, with excitation disordered phase to the Z2 symmetry breaking superradiant phase. The energies close to the non-interacting values, while they strongly classical nature of the transition in the presence of dissipation is ex- hybridized, inseparable degrees of freedom for strong coupling pressed by the fact that the two zero energy modes Re(ω) = 0 become strengths. purely dissipative for couplings g < gc already (indicated with the red Above a critical coupling strength gc, the ground state of the arrows). system breaks the Z2 symmetry and the atoms form a macro- scopic spin aligned in the x-direction, hS xi = O(N), which is ac- companied with a coherent macroscopic population of the intra- cavity mode hai , 0. This symmetry broken phase is called the which is diagonal in the photon modes. The approach to this superradiant phase of the cavity system. In the case of a macro- limit is, however, algebraic ∼ ω−3, in contrast to an exponen- scopic expectation of S x, the orthogonal S z component can no tial approach according to large frequency behavior of the Bose- longer be macroscopically large, which renders any expansion distribution function in equilibrium. On the other hand, for small 2 of the Holstein-Primakoff operators (137) in 1/N invalid. This frequencies ω  g , the occupation becomes essentially ther- is expressed by an instability of the quadratic theory at the su- ωz mal and purely off-diagonal F = 2Teff σx, with low energy effec- perradiance transition, revealed by the presence of a zero energy ω g2 mode, i.e., a pole at ω = 0 in the excitation spectrum. According tive temperature Teff = , cf. Eq. (84). In this low frequency ωz to Eq. (142), this happens at limit, the elementary excitations are strongly hybridized polari- s ton modes, which are diagonal in the photon quadratures x and 2 2 (κ + ωp)ωz p. The absence of a global temperature scale in the present gc = . (143) 4ωp system reflects the fact that due to the Markovian dissipation, this interacting system with a discrete set of degrees of freedom The mode structure in the vicinity of the transition has been ana- is not able to achieve detailed balance between the particular lyzed in Ref. [112], where it has been found that in the presence modes, i.e. the x and p quadratures, for all energy scales. of photon decay, the critical mode becomes purely imaginary before the transition happens, and the corresponding critical dy- One should note that Teff is not proportional to the decay rate namics corresponds to a classical finite temperature transition, κ and consequently the zero temperature equilibrium limit is not see also Fig.7. This is mirrored by the fact that the photons simply obtained by taking the limit κ → 0. The reason is that in effectively thermalize in the low energy regime and their corre- the present setting, the atom-photon coupling g leads to a com- lations can be described by a low energy effective temperature petition of unitary and dissipative dynamics, which is reflected

Teff. The effective temperature can be obtained via a fluctua- in the fact that for g = 0, the eigenstate of the Hamiltonian is a tion dissipation relation, as discussed in Eq. (72) but promoted steady state of the dynamics while this is no longer the case for to Nambu space (see Sec.VB for a discussion of the FDR in finite g. As a consequence, the low energy effective temperature Nambu space). Solving this equation yields the photon distribu- depends on g and vanishes in the limit g → 0. tion function in Nambu representation A further consequence of the dissipative nature of the tran- 2g2 ω2 sition is that the critical exponents correspond to the classical F = σz + σx z . (144) ph 2 2 finite temperature equilibrium transition. One such critical ex- ωzω ω − ωz 2 ponent of the superradiant transition is the so-called photon flux For large frequencies ω  g , this corresponds to the zero ωz exponent νx, which describes the divergence of the cavity pho- z −ν temperature distribution of the non-interacting system F = σ , ton number at the transition, i.e., n ∼ |g − gc| x . It is obtained 36 by frequency integration over the photonic correlation function 2. Effective Ising spin action for the single-mode cavity Z ! K −1 2n = iGph (ω) − 1 ∼ |g − gc| . (145) The Dicke model defined by the master equation (134) obeys ω the common Ising Z2 symmetry, which is spontaneously bro- ken by the ground state of the superradiant phase. Another ap- The exponent νx = 1 found for the present non-equilibrium transition coincides with the classical, finite temperature expo- proach to express the Dicke model in a Keldysh path integral nent for the equilibrium Dicke model in line with the discussion approach is therefore to represent the spin operators in terms above. On the other hand, at zero temperature, the exponent of real Ising field variables, σx(t) → φ(t). In contrast to the of the corresponding quantum phase transition is found to be Holstein-Primakoff transformation applied in the previous sec- tion, the Ising representation does not require a single, large spin νx = 1/2. The scaling behavior of the critical mode at the transition is to have a well defined field theory representation, and is there- expressed in terms of the dynamical critical exponent ν . In the fore also applicable in situations in which the atom-photon cou- t → presence of dissipation, the excitations in the system decay ex- pling is spatially dependent g g(x) and a large spin represen- ponentially in time, which results in an asymptotic exponential tation becomes impossible. Furthermore, we use this approach decay of real-time correlation functions to describe the symmetry broken phase of the problem. The atomic sector of the Hamiltonian (133) describes Ising † K −t/ξt spins in a transverse field and the corresponding universal low h{a(t), a (0)}i = Gph (t) ∼ e . (146) energy description for frequencies below the level spacing ωz is The characteristic decay time ξt is determined by the slowest obtained by transforming the quantum spin operators to classical decaying mode in the system, i.e., by the mode with the smallest real fields. According to the quantum to classical mapping [255] imaginary part in the frequency spectrum. In the vicinity of the this is possible in the scaling regime of the corresponding, i.e., phase transition, this is the critical mode , which shows scaling for the present case in the vicinity of the 2nd order superradiance and glass transitions. The present model represents a general- 1 ξt ∼ , (147) ization of the zero-dimensional Ising model in a transverse field, |g − gc| for which the quantum to classical mapping is known [255] and consists of the replacements i.e., a dynamical critical exponent νt = 1. This behavior is in stark contrast to a zero temperature quantum phase transition, x σi (t) → φi(t) (148) where correlation functions do not decay over time but oscillate with characteristic frequencies ωc, the smallest of which indi- and cates the distance to the phase transition and encodes the critical z 2 2 σ (t) → 1 − 2 (∂tφi(t)) . (149) scaling behavior. The fact that in the present setting the critical i ωz mode becomes purely imaginary is a generic feature of dissipa- The dynamic constraint (σx(t))2 = 1 is implemented via the non- tive phase transitions, which is discussed further in Sec.IV. i linear constraint All of the above discussed results are encoded in the quadratic Z 2 Keldysh action, described by Eqs. (138-141). The resulting crit- 2 iλl(t)(φ (t)−1) δ(φ (t) − 1) = Dλl(t) e i , (150) ical behavior is then that of a non-interacting system, and is i associated accordingly to a Gaussian fixed point of the renor- which introduces dynamical Lagrange parameters λi(t) for each malization group flow. This is in contrast to critical behavior spin variable. In this framework, the purely atomic part of the in interacting systems, which are described by an interacting action on the (±)-contour is Wilson-Fisher fixed point, not smoothly connected to the pre- Z vious one (cf. the discussion in Sec.IVB). This action describes 1 X h  2  2i S at = α λαi φ − 1 − (∂tφαi) . (151) ωz αi the problem up to 1/N corrections resulting from the expansion ω α=± of the Holstein-Primakoff bosons in 1/N. It contains correctly i=1,..,N the physics in the thermodynamic limit N → ∞ and captures The action for the atom-photon coupling and the bare photonic the essential dynamics of the superradiance transition. For finite part is straightforwardly derived according to the previous sec- systems, the higher order terms in 1/N can not be neglected and tions, and the corresponding Keldysh action is have to be taken into account properly. Analytical approaches, relying on an expansion of the Holstein-Primakoff operators up Z X 2! ! 1 λqi(t) λci(t) + ∂t φci to first order in the ratio 1/N, have shown that the first order cor- S = (φci, φqi) 2 ω λci(t) + ∂ λqi(t) φqi rection term leads to a quartic contribution in the action, whose z t i t Z self-consistently determined one-loop corrections are already in g X   + √ φ (a∗ + a ) + φ (a∗ + a ) very good agreement with numerical results [112]. The latter ci q q qi c c N t i have been obtained from a Monte-Carlo wave function (MCWF) Z ! ! simulation of the master equation and agreed with the analytical ∗ ∗ 0 i∂t − ωp − iκ ac + (ac, aq) − . (152) results even on the level of N = 10 atoms. t i∂t ωp + iκ 2iκ aq 37

∗ It is quadratic in the fields φ, a, a as it was the case for the In order to stabilize the system, for coupling strengths g > gc Holstein-Primakoff action in the large-N limit. However, the the steady state breaks the Z2 symmetry of the action, which is non-linear constraint, i.e., the fact that λq,c is a dynamic vari- expressed in terms of a finite, symmetry breaking order parame- able, introduces higher order terms in the Ising fields. The ap- ter φc(ω) → ψδ(ω) + φc(ω), and equivalently in the photon basis proach corresponding to the large-N expansion of the Holstein- ac(ω) → aδ(ω) + ac(ω). These order parameters correspond to x Primakoff fields in the present formalism (in the sense that the finite expectation values hσi (t)i = ψ and ha(t)i = a in the sys- action becomes quadratic) is to treat the Lagrange parameters as tem’s ground state. They modify the soft-spin condition (153) static mean field variables, i.e., λci(t) → λc and λqi(t) → λq. Due according to to causality, a static mean quantum field must be zero, λq = 0. Z ! On the other hand, the value of λc has to be chosen such that the x 2 2 K h(σi (t)) i = ψ + iGat(ω) = 1, (158) non-linear constraint is preserved on average ω Z K x 2 2 K ! where Gat is the Keldysh Green’s function of the fluctuating vari- h(σi (t)) i = hφc(t)i = iGat (ω) = 1. (153) ω able φ. As a consequence, in order to fulfill Eq. (158), the La- grange multiplier λ becomes a continuous function of the order This is equivalent to a vanishing variation of the action with re- c parameter, and therefore an implicit function of the coupling g spect to the Lagrange parameters λ and is known as the “soft- c,q in the superradiant phase. The dependence of λ can be deter- spin” approach to spin models (i.e. the constraint is treated on c mined from Eq. (157). At the superradiant transition at g = g , the mean field level). It becomes exact in the thermodynamic c this equation has one critical solution, i.e., the propagator has a limit N → ∞ [256, 257]. pole at ω = 0, see Fig.7. For larger coupling strengths, this pole Integrating out the N individual atoms leads to the effective crosses the origin and obtains a positive imaginary part, thereby photon action rendering the system unstable. In order to compensate for this   −1  Z  A  mechanism, λc is modified such that the pole does not cross the 1 †  0 G2×2  S = A (ω)   −1  A(ω), (154) real axis, i.e., remains at its value ω = 0 in the superradiant 2 ω  R  2 G2×2 2iκ diag(1, 1) 4g ωpωz phase. Consequently λc = 2 2 for g > gc. Via Eq. (158), this κ +ωp with the Nambu spinor p reveals the scaling of ψ ∼ |g − gc| when the transition is ap-   proached from inside the superradiant phase, which is the same  ac(ω)   ∗  critical behavior of the order parameter as for the equilibrium  ac(−ω)  A(ω) =   (155) transition.  aq(ω)   ∗  The results on the critical properties of the Ising field theory at aq(−ω) the superradiance transition conclude the discussion of the single and the inverse retarded Green’s function mode Dicke model in the framework of the Keldysh path inte-  2g2ω 2g2ω  gral. We have shown that both the Holstein-Primakoff approach  −1  ω − ω + iκ − z − z  R  p ω2−λ ω2−λ  G2×2 =  2 2  . in terms of complex bosonic variables as well as the Ising ap-  − 2g ωz − − − 2g ωz  ω2−λ ω ωp iκ ω2−λ proach in terms of real Ising variables are equivalent on the level (156) of the quadratic, large-N limit of the theory. The critical point The poles of the Green’s function are again determined by the of the transition, as well as the mode structure and the critical zeros of the determinant of the inverse Green’s function, i.e., by scaling behavior have been directly derived from the quadratic the roots of the equation Keldysh action, which illustrates the strength of the present field   theory approach. In the following section, we discuss the exten- ! − 2 2 − 2 2 0 = (λc ω ) (ω + iκ) ωp + 4ωpωzg . (157) sion of the Dicke model to multiple cavity modes, which con- For the non-interacting theory, i.e., for g = 0, this determines tains the possibility of frustrated atom-atom interactions and the √ formation of a in the cavity system. the poles of the atomic sector to be ω = ± λc and in turn fixes 2 λc = ωz to be consistent with the microscopic theory. As long as the system is not superradiant, the atom-photon interaction does not modify the integral (153), and for the entire param- B. Ising spins in a random multi-mode cavity 2 eter range of the normal phase, we find λc = ωz as for the non-interacting case. This again yields the critical value of the In the previous section, we saw that dissipation renders the su- coupling strength (143) we obtained above from the Holstein- perradiant transition essentially a classical phase transition with Primakoff approach. Consequently, the results from the previous critical exponent corresponding to a Z2 symmetry breaking, fi- section are recovered in the Ising representation of the Keldysh nite temperature transition. However, no signature of the non- path integral. equilibrium nature of the steady state could be found in the long- We now turn to the description of the ordered phase. Di- wavelength dynamics governing the transition since the phonon rectly at the superradiant transition, the system becomes unsta- distribution function shows a classical Rayleigh-Jeans diver- ble, which is indicated by a critical mode with frequency ω = 0. gence F ∼ 1/ω at small frequencies. This changes drastically 38 when the setup allows for multiple dissipative cavity modes, frequency space which couple atoms and photons via a random interaction po- Z 2! ! tential. The random interaction potential is realized by freezing 1 X 0 λc − ω φcs S= (φcs, φqs) 2 ω λc − ω 0 φqs the atoms on random positions inside the cavity and by exclud- z ω s ing atomic self-organization with a large set of modes. Z A ! ! X 0 Λ (ω) φci The effective, cavity mediated interactions in the case of mul- + Jsi (φcs, φqs) R K . (162) Λ (ω) Λ (ω) φqi tiple cavity modes are able to induce frustration in the effective si ω atom-atom interactions. At a critical frustration level, these will The frequency dependent terms are the symmetrized photon drive the system into an Ising spin glass phase. Due to the pres- Green’s functions ence of drive an dissipation in the quantum optical realization of this spin glass phase, it features no analogue in condensed matter −ωp ΛR(ω) = , physics and the corresponding spin glass transition does not cor- 2 2 (ω + iκ) − ωp respond to the classical finite temperature transition [113]. The 2iκ(ω2 + κ2 + ω2 ) (163) corresponding model is motivated by recent works on ultracold K p Λ (ω) = atoms in cavities [113, 131, 258–260] and is described by the 2 2 2 (ω + iκ) − ωp master equation (134), where now the Hamiltonian

with an average photon frequency of the cavity ωp. The effective ω XN XM X coupling H = z σz + ω a†a + g (a† + a )σx (159) 2 s l l l sl l l s s=1 l=1 s,l XM g g J = sl il (164) si 4 describes the energy of M photon modes with frequency ωl and l=1

N Ising spins with level spacing ωz, interacting via coupling con- 2 Mg0 Mg0 stants gsl = g0 cos(kl xs), which depend on the atomic position xs fluctuates between the values − 4 ≤ Jsi ≤ 4 and can be ei- and the photon mode function kl, such that −g0 ≤ gsl ≤ g0. The ther ferromagnetic Jsi < 0 or antiferromagnetic Jsi > 0 depend- dissipator describes the decay of each of the M photon modes ing on the s, i. Assuming gsl to be independently and equally with a uniform decay rate κ, distributed for each atom, the couplings Jsi are for sufficiently large M distributed according to a Gaussian distribution func- M ! tion. We set their average value J¯ ≡ hJ i = 0 as it lifts the frus- X 1 si Lρ = κ a ρa† − {a†a , ρ} . (160) tration in the system caused by fluctuating couplings but does not l l 2 l l l=1 modify the universal behavior of the system at the glass transi- tion [113, 131]. The averaged partition function of the system, including the probability distribution for the couplings J is 1. Keldysh action and saddle point equations si Z iS +iS P Z = D[φs]D[Jsi]e , (165) The corresponding Keldysh action is similar to Eq. (152) and reads where ! ! Z X 2 2 1 0 λc + ∂t φcs N X J S= (φcs, φqs) 2 si ω λc + ∂ 0 φqs iS P = − (166) z t s t 2 K Z s,i X  ∗ ∗  + gsl φcs(a + aql) + φqs(a + acl) (161) ql cl is the action of the random couplings Jsi, with correlations t sl Z X ! ! K ∗ ∗ 0 i∂t − ωl − iκ acl h i + (acl, aql) . Jsi Jlm = (δs,lδi,m + δs,mδi,l) . (167) i∂t − ωl + iκ 2iκ aql N t l In this sense, the coupling of the atomic degrees of freedom to Here, the soft spin approximation has been performed and the the random variables Jsi has the same structure as the coupling of scaling of the coupling constants gsl in the thermodynamic limit the atoms to an external bath. However, the significant difference − 1 is implicit, i.e. gsl ∼ N 2 . The thermodynamic limit is reached to a Markovian bath, which would be δ-correlated in space and by taking an extensive number of atoms N → ∞ but leaving the time is that the quenched disorder, while correlated locally in number of photon modes M < ∞ finite since the consideration of space, has infinite correlation time. The latter is expressed by an extensive number of photon modes is physically unrealistic. the fact that the variables Jsi are time-independent. In order to obtain a large-N effective action, the photon degrees Averaging over all realizations of the couplings is done by a of freedom are integrated out, leading to the atomic action in Gaussian integration (cf. App.B), and transforms the disorder 39 part, i.e., the second line of Eq. (162), to a time non-local quartic In the glass phase, the spins attain temporally frozen configu- contribution rations, expressed by an infinite correlation time of the atomic iK Z X correlator, which is expressed by a non-zero Edwards-Anderson S = ΦΛ (ω)ΦΛ (ν), (168) parameter at-at N l,m m,l ω,ν l,m 1 X x x qEA ≡ lim hσ (t)σ (0)i. (178) with t→∞ N l l l ! ! 0 ΛA(ω) φ ΦΛ (ω) = (φ , φ ) cm . (169) Consequently, the correlation function QK(ω) consists of a regu- l,m cl ql ΛR(ω) ΛK(ω) φ qm lar part, describing the short time dynamics and a non-vanishing The double sum in (168) is decoupled via a Hubbard- contribution at ω = 0. It can be expressed via the modified fluc- Stratonovich transformation, which introduces the macroscopic tuation dissipation relation [261] 0 fields Qαα0 , with α, α = c, q being Keldysh indices [113, 131]. K K This transformation is in general not unique but can be made Q (ω) = 4iπqEAδ(ω) + Qr (ω) (179) unique by requiring the equivalence in terms of the order parameter and a regular contribution QK ω δ 1 X r ( ). The soft-spin condition (177) together with Eq. (176) ⇔ 0 h 0 i Z = 0 Qαα (ω, ν) = N φαl(ω)φα l(ν) . (170) fixes the value of the Lagrange parameter λ and therefore fully δQαα0 l determines the spectrum of the system. Similar to the superradi- This identifies the Hubbard-Stratonovich field with the average ant transition, the Edwards-Anderson parameter becomes non- atomic correlation function. Since one is interested in the sta- zero when the poles of the system become critical (approach the tionary, time translational invariant state, real axis) and its emergence is a mechanism to stabilize the crit- ical modes in the ordered phase. Qαα0 (ω, ν) = 2πδ(ω + ν)Qαα0 (ω). (171) After the Hubbard-Stratonovich decoupling, the resulting action 2. Non-equilibrium glass transition is quadratic in the atomic fields and they are integrated out, lead- ing to the macroscopic action The variance of the effective, long-range atom-atom interac- " # 1 tion K is a measure of the frustration in the system and for suf- S = iNTr KΛQΛQ − ln G˜ (ω). (172) 2 ficiently strong frustration, the system enters a glass phase, de- scribed by an infinite autocorrelation time of the spins and the Q G˜ Here Λ and are matrices in Keldysh space and is defined as emergence of a non-zero Edwards-Anderson parameter qEA > 0.  −1 This goes hand in hand with a critical continuum of modes ˜ −1 G(ω) = G0 (ω) − 2KΛ(ω)Q(ω)Λ(ω) . (173) reaching zero, which is the characteristic feature of a critical phase of matter. The phase diagram for the fully coherent model In the thermodynamic limit, the partition function is determined has been analyzed in [131] and in [113] it was shown that the by the saddle point value of the action, i.e. by the condition glass phase persists in the presence of dissipation. However,

δS ! the dissipative model shows new universal features of the glass = 0 (174) phase, which correspond neither to a zero nor to a finite temper- δQ 0 αα ature equilibrium transition. This is attributed to the fact that the 0 for all α, α . This yields the values of the fields Qαα0 (ω) at the critical modes are described by poles in the complex plane which saddle-point, which can be identified with the atomic response are neither purely real as in the zero temperature (or quantum) and correlation functions. The saddle-point equations for the case nor purely imaginary as for the finite temperature transition. atomic response and correlation functions are In the presence of dissipation, a photon that is emitted from !−1 an atom can either be absorbed by another atom, leading to an 2(λ − ω2)  2 QR(ω) = − 4K ΛR(ω) QR(ω) (175) effective atom-atom interaction, which is infinite range in space, ωz or decay from the cavity. The latter happens on a characteristic time scale τp = 1/κ, which is as well the time scale for which and atoms can interact with each other by exchanging a photon be- K 4K|QR|2ΛK (QAΛA+QRΛR) fore the photon decays. As a consequence, the photon decay Q (ω) = R R 2 . (176) 1−4K|Q Λ | reduces the effective range of the atom-atom interaction in time Additionally, due to Eq. (170), the soft-spin constraint maps to and reduces the strength of frustration in the system. Close to the Q-fields according to the glass transition and in the glass phase, this leads to the emer- −1 gence of a crossover scale ωc = τ , above which the spectral Z 2 X c i QK(ω) = hφ (−ω)φ (ω)i = 2. (177) properties of the system correspond exactly to a zero tempera- N cl cl ω l ture spin glass. Below this frequency, the spectrum reveals the 40

The latter leads to a characteristic algebraic decay of the photon correlation function, as discussed below. The universality class of the present dissipative quantum glass transition is determined by the critical exponents at the transition, which describe the scaling of the parameters α, γ, Z as a function of δ = |K − Kc|. It is summarized in the equations √ 2 2 3 2(ωp+κ ) δ 2 αδ = √ 8 K3κ log(δ) 2 2 2 ωp+κ δ , (183) γδ = 16K2κ log(δ) 2 2 3 ωp(ωp+κ ) δ Zδ = √ 8 K5κ2 log(δ) which show the typical logarithmic correction to scaling at a   quantum glass transition and identify the critical exponents. In- Figure 8. Spectral density A(ω) = i QR(ω) − QA(ω) of the atoms in dicated by the scaling behavior, dissipative and coherent dynam- the glass phase for parameters K = 0.01, ωz = 2 and varying photon√ pa- ics rival each other when approaching the transition, separating rameters ωp, κ. For frequencies below the crossover ω < ωc, A ∼ ω this glass transition from equilibrium transitions and the previ- is overdamped, corresponding to Eq. (182), while it recovers the typi- ously discussed Dicke transition, which are either fully coherent A ∼ cal thermal glass behavior ω at intermediate frequencies ω > ωc. (quantum) with γ = 0 or fully dissipative (classical) with α = 0 Figure copied from Ref. [113]. (Copyright (2013) by The American Physical Society) sufficiently close to the transition. The present glass transition has therefore no counterpart in static equilibrium physics and is termed dissipative quantum glass transition. Similar behavior is breaking of time reversal symmetry by the dissipation and for- present in system bath settings in equilibrium, where the bath mally corresponds to the dynamical universality class of dissipa- however not only imprints a finite temperature to the system but tive quantum glasses, other examples of which are spin glasses as well modifies its spectral properties [259, 261, 262]. What coupled to an ohmic bath [261–263] or a bath of metallic elec- these situations share in common is that the effective theory, after trons [257, 259]. The crossover frequency is elimination of the bath variables, obeys no time reversal symme- try, which ensures the same asymptotic universal behavior as in   2 −1 ω2 + 1 κ2 ω2 + κ2  the case of the Markovian photon loss in the present system.  p 2 p  ωc = κ  + √  . (180)  ω2 + κ2 2   p 2 Kωz  3. Thermalization of photons and atoms in the glass phase and the modifications of the spectrum below this scale are due to the damping introduced by the Markovian bath. On the normal or disordered side of the phase diagram, the corresponding low As typical for many open quantum systems (for exceptions, frequency propagator is see the discussion in Sec.IV), the statistical properties of the excitations are described by a low energy effective temperature  −1 R 2 2 (LET) and a corresponding thermal distribution of the excita- Qn (ω) = Z (ω + iγ) − α , (181) tions. However, as has been found for the single mode Dicke describing the Ising spins as damped harmonic oscillators with model in the normal as well as in the superradiant phase, the decay rate γ > 0 and frequency α, with the physical meaning LET for the photonic and the atomic subsystem did not coincide, of an inverse lifetime and energy gap of the atomic excitations. i.e., the subsystems have not thermalized but are held at differ- When the glass transition is approached, the gap, the inverse ent temperatures corresponding to their individual coupling to a lifetime and the residue Z scale to zero simultaneously and the bath [112]. This dramatically changes in the present system with atomic response in the glass phase multiple photon modes. As frustration is increased by driving  − 1 the variance K towards the critical value Kc, atoms and photons R ¯ 2 2 Qg = Z ω + γ¯|ω| (182) begin to thermalize towards the same, shared LET. The distri- bution function F of photonic or atomic modes is obtained by is non-analytic and can no longer be interpreted as a harmonic solving the fluctuation dissipation relation oscillator. The broken time reversal invariance manifests itself in both parameters γ, γ¯ 0 being non-zero, which modifies the K R A , Q (ω) = Q (ω)Fat(ω) − Fat(ω)Q (ω) (184) low frequency dynamics in the entire glass phase towards a dis- sipative quantum glass. This is for instance expressed by a spec- for the atomic Green’s functions. It leads to an atomic LET tral density, which features an anomalous square root behavior √ 2 2 A(ω) = −2Im(QR(ω)) ∼ |ω| for small frequencies and there- ωp + κ Teff = (185) fore has a non-analytic response at zero frequency, see Fig.8. 4ωp 1 1

S(ν) Sinc(ν) Sc(ν) ν

S(ν) Sinc(ν) Sc(ν) ν

2κ 1 (ω) ω ωξδ ω ξδ A ω0√2Kδκ ∝ − 1 (ω) ω ωξδ ω ξδ A ω0√Kδ ∝ − S(ν) Sinc(ν) Sc(ν) ν 1 S(ν) Sinc(ν) Sc(ν) ν 3 κ =0.03,ω0 =0.9 κ =0.1,ω0 =1 κ = 10− ,ω0 =0.8 3 κ =0.03,ω0 =0.9 κ =0.1,ω0 =1 κ = 10− ,ω0 =0.8

2κ S(ν) Sinc2κ(ν) Sc(ν) ν (ω) ω (ω)ωξω δ ω ωξξδ δ ω ξδ A ω0√AKδ ω∝0√Kδ−ω ∝ − ∝ ω ∝

3 κ =0.03,ω0 =0.9 2κκ=0.1,ω0 =1 κ = 10− ,ω0 =0.8 3 (ω) ωκ =0.03,ω0ωξ=0.9 κ =0.1ω,ω0 =1ξ κ = 10− ,ω0 =0.8 J √Kκδ =0.2 κ =0δ.05 κ =0 A ω0√Kδ ∝ − J √Kκ=0 .2 κ =0.05 κ =0 ω ∝ ω ∝ 3 3 c κ =0.03,ωκ0=0=0.1.,ω9 0 κ=0=0.7 .1κ,ω=00 .=103,ω0 κ=0=.9 10−κ =,ω 100 −=0,ω.08=0.6 √ωωQG 3 ∝ c κ =0.1,ω =0.7 κ =0.03,ω =0.9 κ = 10− ,ω =0.6 √ωω 0 0 0 ∝ QG J √Kκ=0.2 κ =0.05 κ =0 ω2 + ω2 + κ2 ωF(ω) 0 J =0.1 J =0.11 J =0.14 J =0.17 ω2 +2ωωJ2 +√κKκ2 =0.2 κ =0.05 κ =0 ωF(ω) 00 J =0.1 J =0.11 J =0.14 J =0.17 2ωω0 ∝ 3 c κ =0.1,ω =0.7 κ =0.03,ω =0.9 κ = 10− ,ω =0τ .6 √ωω 0 0 1+0 0 ∝ QG ττ0 1+  3 c κ2=0.21,ω02=0.7 κ =0.03,ω0 =0.9τκ = 10− ,ω0 =0.6 √ωωQG ω + ω0 + κ ∝ ωF(ω) J =0.1 J =0.11 J τω=0.14 J =0.17 2√ω0 0 J Kκ=0.2 κ =0.05 κ =0τω 41 ω2 + ω2 + κ2 2π 0 =1, =-6, =0.4, J=0.4,0 Wurzel(K)=0.4 ωF(ω0 ) τ0 J =0.21π J =0.11 J =0.14 J =0.17 1.2 1+2ω0 finiteq ˜EA implies long temporal memory in photon autocorrela- τ (2)   g (τ) tion functions and an extensive number of photons permanently g(2)(τ) occupying a continuum of high energy modes. The latter is re- τω0 3 τ0 c κ =0.1,ω =0.7 κ =0.03,ω =0.9 κ = 10−1+,ω =0.6 √vealedωω by an algebraic decay of the system’s correlation func- 0 0 2π 0 2κτ QG (2) 1+2e− τ ∝ tions, as for instance for the four point (or g ) correlation func-  2κτ 1+2e− tion, which is shown in Fig.9. (2) 2 2 2 g (τ) τω0 The photon response and correlation functions are easily ac- ω + ω0 + κ 2π cessible via cavity photon output measurements, and therefore ωF(ω) 1.1 J =0.1 J =0.11τc =J2=0π .14 J =0.17 2ω0 ω2cπ represent the natural observables for the detection of the glass 2κτ τ = 1+2e− c dynamics in cavity QED experiments. In fact, in Ref. [113], it ωc g(2)(τ) has been demonstrated that a complete characterization of the τ20π glass phase can be performed by different but standard photon 1+τc = output measurements, such as homodyne detection and inten- τωc 2κτ sity correlation measurements. Formally, the photon Green’s   1+2e− 1 functions can be obtained from the Keldysh action formalism by 0 10 τω0 20 30 introducing source fields µαα0 in the microscopic action, which couple to the photon variables. After integrating out the pho- 2π 2π τc = ton variables in the action, the photon Green’s functions are then ωc Figure 9. Algebraic decay of the photon four point correlation function obtained via functional derivatives with respect to these source (2) † † 2 fields. This standard field theory technique relates the photon g (τ) = ha (0)a (τ)a(τ)a(0)i/ng(2)at( longτ) times, for parameters ωp = 1, κ = 0.4, ωz = 6, K = 0.16, see also Ref. [113]. The algebraic decay Green’s functions to the solution for the atomic response and sets in at the inverse crossover frequency ωc, given by Eq. (180). The correlation functions [113]. corresponding exponential decay of g(2) in the normal phase is plotted 2κτ for comparison. Figure copied1+2 frome Ref.− [113]. (Copyright (2013) by The American Physical Society) 4. Quenched and Markovian bath coupling in the Keldysh formalism

2π The Keldysh path integral formalism represents a theoretical in the glass phase, which coincidesτc = with the photonic LET [113]. approach which incorporates in a very straightforward way the ωc The thermalization of the two subsystems is a consequence of coupling of the system to a non-thermal bath. Here we com- the disorder induced effective long range interactions, which re- pare quenched disorder and Markovian baths in more detail. As distribute energy between the different modes and enable equili- shown above in the present section, the coupling of the system bration even in the presence of the Markovian dissipation in the variables to quenched disorder, realized by the variables Jlm, is photon sector. In the paramagnetic phase, the distribution func- equivalent to the coupling to a bath with infinite correlation time, tions of atoms and photons are identical for frequencies ω > α larger than the gap, but deviate from each other for lower fre- 0 K hJ (t)J 0 0 (t )i = (δ 0 δ 0 + δ 0 δ 0 ). (187) quencies. The elementary excitations above this frequency are lm l m N ll mm lm ml strongly correlated and can not be seen as weakly dressed pho- tons or atoms. As a consequence, the observables in the critical This represents the opposite limit of a Markovian bath, which glass phase are dominated by the universal low energy behavior has a typical correlation time that is much shorter than the sys- of the glass propagator (182) and thermal statistics of the excita- tem correlation time. As a consequence, at each single system- tions. bath interaction event, the Markovian bath immediately equi- The strong light-matter interactions not only lead to thermal- librates and looses all its memory on the interaction. On the ization of the atomic and photonic subsystems, they also fea- other hand, the quenched bath equilibrates infinitely slowly and ture a complete imprint of the glass dynamics of the atoms onto keeps memory on each single system-bath interaction. As a con- the cavity photons. This results in universal spectral proper- sequence, it introduces effective temporally long range interac- ties of the photonic response, as has been shown in Ref. [113], tions between the system variables. These tend to slow down the and concerns a complete locking of both subsystem’s low en- system, pronouncing its ω = 0 correlations as expressed by the ergy response properties. The photons form a photon glass state, Edwards-Anderson parameter. which features a large, incoherent population of photon modes, In an equilibrium formalism, disordered systems have to be signaled by a finite Edwards-Anderson parameter in the photon dealt with a computationally demanding replica method [256, correlation function 257], which is furthermore physically far less transparent than the Keldysh formalism, which thus represents a convenient ap- K K proach to disordered systems. In open quantum systems, where Gph (ω) = 4πδ(ω)˜qEA + Greg(ω), (186) disorder, i.e., the coupling to a quenched bath, comes together which shows the same scaling behavior as the atomic Edwards- with the dissipation introduced by a Markovian bath, both types

Anderson parameter close to the glass transitionq ˜EA ∼ qEA.A of bath couplings compete with each other, leading in the present 42 glassy system to strongly modified spectral properties, which driven-dissipative Bose-Einstein condensation with systems in mirror the effect of the Markovian bath by a crossover from non- thermal equilibrium can be given most conveniently on the basis equilibrium to equilibrium scaling behavior. of the semiclassical limit of Sec.IIC: indeed, the equation of motion (80) for the classical field takes the form of the Langevin equations that are used to model universal dynamics in ther- IV. NON-EQUILIBRIUM STATIONARY STATES: BOSONIC mal equilibrium phenomenologically [1]. Structurally, equa- MODELS tion (80) is similar to model A for a non-conserved order pa- rameter, with the additional inclusion of reversible mode cou- plings [268]. However, the equilibrium symmetry discussed in Recent years have seen tremendous progress in experiments Sec. II D 1, which is present in all models of Ref. [1], is violated on exciton-polaritons in semiconductor microcavities (for re- in driven-dissipative systems. This violation is due to pumping views see, e.g., Refs. [13, 148]), making these systems the and loss terms, which moreover lead to the absence of parti- prime candidates for studying condensation phenomena under cle number conservation. The latter typically is associated with non-equilibrium conditions. As we have already mentioned in the symmetry under quantum phase rotations (see Sec. II D 4). Sec.IC2, the fundamental di fference from conventional con- This is another crucial difference to Bose-Einstein condensation densates is due to the finite life-time of exciton-polaritons, which in equilibrium: there, particle number conservation entails the makes continuous driving necessary to maintain a steady popu- existence of a slow dynamical mode that modifies the universal lation. properties, and is taken into account in model F of Ref. [1]. To What makes these systems genuinely driven-dissipative, is summarize, driven-dissipative condensation differs from Bose- that they are coupled to several baths, or to time-dependent driv- Einstein condensation in equilibrium by the absence of the equi- ing fields, with which they exchange particles and energy. In librium symmetry and the symmetry under quantum phase ro- the case of exciton-polaritons, which are hybrid tations. This difference lies at the heart of the novel universal composed of a Wannier-Mott exciton and a photon, the photonic behavior out of equilibrium discussed in the following. component can leak out through the mirrors forming the cav- In Sec. II B 2, we analyzed driven-dissipative condensates ity. Thus, the electromagnetic vacuum outside the cavity serves within mean-field theory, disregarding fluctuations around the as a reservoir into which particles are lost. These losses have homogeneous condensate mode. However, in order to access to be compensated by laser driving. In many experiments, the universal aspects such as the behavior of correlations of the con- driving laser is far blue-detuned from the bottom of the lower densate field at large distances or dynamical critical phenomena polariton band, and thus coherently creates high-energy excita- at the condensation transition, one has to carefully include the tions. During the relaxation of the latter, which is caused by effect of fluctuations. This can be done gradually — first inte- phonon-polariton and stimulated polariton-polariton scattering, grating out short-scale fluctuations and moving on to account for coherence is quickly lost. In other words, lower polaritons can fluctuations on larger and larger scales — by means of RG meth- be regarded as being pumped not directly by the laser but rather ods, such as the FRG discussed in Sec.IIE. Then, an intrigu- by a reservoir of high energy excitons. Several approaches have ing question is, whether the non-equilibrium nature of driven- been used to model this effectively incoherent pumping mecha- dissipative condensates at the microscopic scale becomes more nism [159, 192, 264]. In this context, the description of a driven- or less pronounced under renormalization. In the former case, dissipative condensate introduced in Sec. II B 2 might be con- effective equilibrium is established at large scales, while in the sidered as a toy model, which hides all the microscopic details latter case, the universal physics is expected to be profoundly associated with coherent excitation and subsequent relaxation of different from its equilibrium counterpart. Which of these sce- high energy excitons in the coupling of the system to several narios is realized depends crucially on the spatial dimensionality independent baths. of the system: in 3D the long-wavelength regime is effectively In the following, we review recent investigations of the uni- thermal, whereas in one- and two-dimensional systems the devi- versal long-wavelength scaling properties of driven-dissipative ation from equilibrium is relevant in the RG sense, and these sys- condensates [19, 26, 181, 265–267]. Then, the precise details tems are governed by strongly non-equilibrium RG fixed points. of the chosen model cease to matter: according to the power- To see how this comes about, we require a quantitative mea- counting arguments given in Sec.IIC, for any choice of a mi- sure for the deviation from equilibrium conditions in a driven- croscopic model the effective long-wavelength description is dissipative condensate. In Sec. II D 1 we discussed that thermal given by the semiclassical Langevin equation for the conden- equilibrium requires the ratios of coherent to corresponding dis- sate field (80). In fact, the latter can be derived in the spirit of sipative couplings to take a common value. This is illustrated in Ginzburg-Landau theory for continuous phase transitions [190] Fig.6 and formalized in Eq. (101). In turn, it implies that any by writing down the most general equation that is compatible deviation Kc/Kd , uc/ud in the values of these ratios directly with the symmetries of the problem. The reasoning behind such indicates a deviation from equilibrium conditions. Hence, the an approach is that the universal properties are fully determined quantity λ defined by [19] by the spatial dimensionality and symmetries of a physical sys- ! tem [84, 86, 182]. Kduc λ = −2Kc 1 − (188) A comparison in terms of symmetries of systems exhibiting Kcud 43 can serve as a quantitative measure for the departure from ther- g mal equilibrium. Note that λ = 0 in equilibrium, and that λ is well-defined also for Kd = 0, which is the microscopic value of the diffusion constant Kd in the model for driven-dissipative condensates introduced in Sec. II B 2. It turns out to be most convenient to combine λ with the quantities [19]

!  2  Kd uc γ  uc  D = Kc + , ∆ = 1 +  (189) K u ρ  2  c d 2 0 ud 1 (see Sec. II B 2 for the definition of the noise strength γ and the mean-field condensate density ρ0) to define the dimensionless non-equilibrium strength as [5]: 1 2 3 d

λ2∆ Figure 10. Equilibrium vs. non-equilibrium phase diagram for driven- g = Λd−2 , (190) 0 D3 dissipative condensates (cf. Ref. [252]). The line g = 1, where g is defined in Eq. (190) and measures the deviation from equilibrium con- where Λ0 is the UV momentum cutoff. Thus, the answer to ditions, separates the close-to-equilibrium regime for g < 1 from the the question of whether equilibrium vs. non-equilibrium uni- strong-coupling, far-from-equilibrium regime at g > 1. Red dots indi- versal behavior is realized in driven-dissipative condensates, is cate the fixed-point values of g that are reached if the RG flow is initial- encoded in the RG flow of g. ized in the close-to-equilibrium regime. In dimensions one and two, the In the condensed phase, the RG flow of g is driven domi- equilibrium fixed point at g = 0 is unstable, and the RG flow along the nantly by fluctuations of the gapless Goldstone mode discussed dashed lines is directed towards the blue line of strong-coupling fixed in Sec. II D 6, i.e., by fluctuations of the phase of the conden- points. Thus, a system that is microscopically close to equilibrium will exhibit strongly non-equilibrium behavior at large scales. On the other sate field. The latter were shown [269–272] to be governed by hand, in three spatial dimensions, an initially small value of g is dimin- the Kardar-Parisi-Zhang (KPZ) equation [73], in which λ de- ished under renormalization, and the universal large-scale behavior is fined in Eq. (188) appears as the coefficient of the characteristic governed by the effective equilibrium fixed point at g = 0. The green non-linear term, see Eq. (199). Below in Sec.IVA, we present line indicates the existence of a critical value gc in d > 2, corresponding an alternative mapping of the long-wavelength condensate dy- to a transition between the effective equilibrium phase and a true non- namics to the KPZ equation, starting from the Keldysh action in equilibrium phase that is realized for large microscopic values g > gc. Eq. (73) and integrating out the gapped density mode within the Keldysh functional integral. As a consequence of the mapping to the KPZ equation, the RG flow of g in d spatial dimensions is RG flow of g that is found within this approach is illustrated at the one-loop level given by [5] qualitatively in Fig. 10, which shows that also in 2D the flow is out of the shaded close-to-equilibrium regime with g < 1, and (2d − 3) Cd 2 ∂`g = − (d − 2) g + g , (191) towards a strong-coupling, non-equilibrium fixed point. The sit- 2d uation is quite different in 3D: in this case, if the microscopic where ` = ln(Λ/Λ0), Λ is the running momentum cutoff, and value of g is small, at large scales an effective equilibrium with a 1−d −d/2 Cd = 2 π Γ(2 − d/2) is a geometric factor. The key role renormalized value g → 0 is reached. However, for d > 2, there that is played by spatial dimensionality becomes manifest in the exists a critical line of unstable fixed points gc, separating the canonical scaling of g, which is encoded in the first term on the basins of attraction of the equilibrium and non-equilibrium fixed RHS of the flow equation: to wit, g is relevant in 1D where points, for g < gc and g > gc respectively. Thus, in addition d−2 < 0, marginal in 2D, and irrelevant in 3D since then d−2 > to the effective equilibrium phase, a true non-equilibrium phase 0. In 2D, the loop correction — the second term on the RHS of may be reached in systems that are far from equilibrium even at Eq. (191) — is positive, making g marginally-relevant. This has the microscopic level also in 3D [273]. The properties of this far-reaching consequences for a driven-dissipative condensate in phase have not been explored so far. which the microscopic value of g is small, i.e., which is close to The rest of this section is organized as follows: in Sec.IVB, equilibrium: upon increasing the scale at which the system is ob- we review the dynamical critical behavior at the driven- served, the non-equilibrium nature is more pronounced in one- dissipative condensation transition in 3D [26, 181, 274], which, and two-dimensional systems, whereas effective equilibrium is according to the above discussion, is governed by an effective established on large scales in three-dimensional systems. In 1D equilibrium fixed point. Signatures of the non-equilibrium na- the canonical scaling towards strong coupling is balanced at an ture of the microscopic model are present in the asymptotic fade- attractive strong-coupling fixed point (SCFP) g∗ by the loop cor- out of the deviation from equilibrium at large scales. In contrast, rection. This term vanishes at d = 3/2, and for d > 3/2 the one- the universal scaling behavior of driven-dissipative condensates loop equation does not have a stable SCFP, which, however, is in both 2D [19] and 1D [265–267] is quite distinct from the equi- recovered in a non-perturbative FRG approach [252–254]. The librium case and governed by the SCFP of the KPZ equation. 44

We review the resultant physical picture in Secs.IVC andIVD, The low-energy effective action for the Goldstone boson θ in a respectively. driven-dissipative system can be derived by integrating out fluc- tuations of the density ρ in Eq. (192) in the Keldysh partition function with action S given by Eq. (73), A. Density-phase representation of the Keldysh action Z Z ∗ ∗ iS ∗ iS Z = D[φc, φc, φq, φq]e = D[ρ, θ, ζ, ζ ]e . (193) Especially in reduced dimensions, at large scales the proper- The equality of the integrals over complex classical and quan- ties of condensates are vitally influenced by fluctuations of the tum fields and the variables introduced in the transforma- Goldstone mode. To name an example, in 2D, both in [275] tion Eq. (192) follows from the fact that this transforma- and out of equilibrium [19] these fluctuations lead to a sup- tion leaves the integration measure invariant, i.e., we have pression of long range correlations. In Ref. [19], an effective ∗ ∗ ∗ D[φc, φ , φq, φ ] = D[ρ, θ, ζ, ζ ]. Note that this would not be the long-wavelength description of a driven-dissipative condensate c q case if instead of the density we introduced the amplitude of φ with the condensate phase as the single dominant gapless de- c as a degree of freedom. Our goal is then to treat the integrals over gree of freedom (cf. Sec. II D 6) has been formulated starting ρ and ζ in Eq. (193) in a saddle-point approximation, which, as from the Langevin equation (80) for the complex order parame- we show below, is justified since fluctuations of the density are ter. Here we present an alternative and direct derivation within gapped in the ordered phase with r < 0 and hence expected to the Keldysh functional integral formalism [102, 178]. It is based d be small. The first step is thus to find the saddle point, i.e., the on the Keldysh action in terms of which the microscopic theory solutions to the classical field equations (cf. Eqs. (74)) of a bosonic many-body system with particle loss and gain is formulated in Sec. II B 2. This differs from the derivation pre- δS δS = 0, = 0. (194) sented in [19], which was based on the detour over the Langevin δρ δζ equation (80) that effectively captures the physics on a meso- scopic scale (cf. Fig.5 and the discussion in Sec.IIC). On this For rd < 0, we recover the mean-field solution of Sec. II B 2, − − level, amplitude fluctuations can be eliminated leading to the given by ρ = ρ0 = rd/ud = rc/uc (note that the last equality KPZ equation for the phase. In this sense, we establish here a can always be satisfied by performing a gauge transformation closer link between microscopic and mesoscopic theories, which to adjust the value of rc as described in Sec. II B 2) and ζ = is both appealing from a theoretical point of view and brings 0. We proceed to expand the action Eq. (73) to second order about a number of technical advantages. For example, physical in fluctuations of ρ and ζ around the saddle point. Note that observables are usually represented by quantum mechanical op- the quantum vertex in the action Eq. (73) that is cubic in the erators or equivalently in terms of fields in a Keldysh functional quantum fields does not contribute at this order. Denoting the − integral description of the microscopic theory; here our approach density fluctuations as π = ρ ρ0 we find comes in handy as it yields the effective long-wavelength form Z  √ n h 2i 2 of generating functionals for expectation values and correlation S = 2 ρ0 −ζ1 ∂tθ + Kc (∇θ) + Kcζ2∇ θ functions of these observables which can then be evaluated fur- t,x o 2 ther utilizing established approximation strategies. − (ucζ1 − udζ2) π + i (γ + 2udρ0) |ζ| , (195) Apart from specific applications, the derivation of the action ζ ζ for the Goldstone mode presented here deepens our understand- where 1 and 2 are, respectively, real and imaginary parts of ζ ζ π ing of general properties of the Keldysh formalism with regard . Here, of all terms involving the products of fluctuations 1 ζ π to phase rotation symmetries. The crucial point is, that classical and 2 we only keep the dominant ones in the long-wavelength phase rotations introduced in Sec. II D 4 are a symmetry of the limit, i.e., we neglect contributions containing temporal deriva- ζ ∂ π, ζ π∂ θ ζ ∇2π, ζ ∇π · ∇θ, Keldysh action both in a closed system and in the presence of tives 2 t 1 t , and spatial derivatives 1 2 and ζ π (∇θ)2 terms that describe incoherent pumping and losses, and as we 1 which are both small as compared to the mass-like u ζ π u ζ π have seen in Sec. II D 6, the spontaneous breaking of this sym- contributions c 1 and d 2 in Eq. (195) for the Goldstone θ ω → q → metry is sufficient to ensure the appearance of a Goldstone mode mode , in the regime 0 for 0. Note that terms π ζ that corresponds to fluctuations of the phase of the order param- of higher order in and are contained in the original action eter [159, 191, 192, 276]. In the basis of classical and quantum Eq. (73) both in contributions involving derivatives and in the co- iα herent and dissipative vertices. The validity of the saddle-point fields φc,q such phase rotations become φc,q 7→ φc,qe , showing that the Goldstone mode corresponds to joint phase fluctuations approximation, therefore, is restricted to the low-frequency and of both the classical and the quantum fields. Therefore, in order low-momentum sector in a weakly interacting system. π to derive the action for the Goldstone boson we represent the The action (195) resulting from this expansion is linear in fields in the form and hence integration over this variable is trivial and yields a δ-functional, which in turn facilitates integration over the imag- √ inary part ζ of ζ, φ = ρeiθ, φ = ζeiθ, (192) 2 c q Z Z 0 − iS ˜ iS KPZ where the density ρ is real whereas ζ is a complex variable. Z = D[θ, ζ1, ζ2] δ[ucζ1 udζ2]e = D[θ, θ] e , (196) 45 where at each step a normalization factor is implicitly included As pointed out above, an alternative derivation of the KPZ in the integration measure, ensuring Z = 1 [102, 178]. In the√ last equation (199) as the effective long-wavelength description of equality we replaced ζ1 by the KPZ response field θ˜ = i2 ρ0ζ1, driven-dissipative condensates starts from the Langevin equa- and the KPZ action S KPZ is given by [5, 102] tion (80)[19]. Then, the coefficients in the KPZ equation are slightly different, and they are given by Eqs. (188) and (189) Z   2 λ 2 instead of Eq. (198). To be specific, the differences are (i) the S KPZ = θ˜ ∂tθ − D∇ θ − (∇θ) − ∆θ˜ , (197) t,x 2 absence in Eq. (189) of the tree-level shift ∝ udρ0 of the noise strength ∆ in Eq. (198), and (ii) the absence in Eq. (198) of the where the diffusion constant, non-linear coupling, and noise terms proportional to Kd in Eqs. (188) and (189). (i) is due to strength, respectively, are expressed in terms of the microscopic the fact that in the Langevin equation (80), which is valid in parameters in the original action Eq. (73) as the semiclassical limit (see Sec.IIC), the quantum vertex that is φ∗φ φ∗φ  2  proportional to c c q q in the action in Eq. (73) is neglected. uc γ + 2udρ0  uc  (ii) results from the absence of the diffusion term K in the mi- D = Kc , λ = −2Kc, ∆ = 1 +  . (198) d u ρ  2  d 2 0 ud croscopic model (73); on the other hand, in the Langevin equa- tion (80) this term is included, as it is generated by integrating Along the lines of the derivation of the Langevin equation (80) out fluctuations with wavelengths below the mesoscopic scale from the action in the semiclassical limit in Eq. (79), the KPZ on which the Langevin equation is valid (cf. Fig.5). action can be seen to be equivalent to the KPZ equation, which Starting from the effective long-wavelength description of the reads condensate dynamics derived in this section, the universal scal- λ ing properties can be obtained from an RG analysis. For the KPZ ∂ θ = D∇2θ + (∇θ)2 + η, (199) equation, this procedure, which leads at lowest order in a pertur- t 2 bative expansion in λ to the one-loop flow equation (191),18 is where the stochastic noise η has zero mean, hη(t, x)i = 0, and described, e.g., in Refs. [5, 102]. is Gaussian with second moment hη(t, x)η(t0, x0)i = 2∆δ(t − For completeness, we note that in the absence of drive and 0 0 t )δ(x − x ). Originally [2, 73], Eq. (199) was suggested by Kar- dissipation, i.e., for Kd = rd = ud = 0, an analogous derivation dar, Parisi, and Zhang as a model to describe the growth of a for a weakly interacting in thermal equilibrium leads surface, e.g., due to the random deposition of atoms. In this con- to an effective action for the phase alone that is given by text, h = θ is the height of the surface, and the origin of the Z   non-linear term is purely geometric [73]: the growth is assumed ˜ 2 2 2 S = θ ∂t − c ∇ θ, (201) to occur in a direction that is locally normal to the surface and t,x at a rate ds/dt = λ; If an increment ds = λdt is added along the p normal, the corresponding change of the surface height is where c = 2Kcucρ0 is the speed of sound, and describes the dissipationless propagation of sound waves with linear disper- q  λ  sion ω = cq. To actually describe thermal equilibrium, this dh = (λdt)2 + (λdt∇h)2 ≈ λ + (∇h)2 dt. (200) 2 action has to be supplemented by infinitesimal regularization terms, discussed at the end of Sec. II B 1. Removing from dh/dt the average deposition rate λ by a trans- Finally, let us comment on the relation of the approach pre- formation to a co-moving frame, h(t, x) 7→ h(t, x) + λt, and sented here to a Bogoliubov expansion in fluctuations of the adding a term D∇2h that describes surface tension, we obtain complex fields around the mean field average values, δφc = the (deterministic part of the) growth equation (199). Intuitively, √ φc − ρ0 and δφq = φq. The latter can be recovered formally it seems clear that the growth of a surface represents a genuine by expanding the transformation Eq. (192) around the arbitrarily non-equilibrium process. More formally, this can be seen by not- chosen value θ = 0, which yields ing that the non-linear term in the KPZ equation does in general not satisfy a potential condition [5].17 √ π √ φc = ρ0 + √ + i ρ0θ, φq = ζ, (202) 2 ρ0

and interpreting π and θ as the real and imaginary parts of 17 The most general Langevin equation describing (near-) equilibrium dynamics δφc. In the KPZ action in Eq. (197), such an expansion in θ contains both (i) relaxational and (ii) reversible contributions to the determin- would amount to neglecting the non-linearity. Without the non- istic dynamics [5]. The linear diffusion term in Eq. (199) is of type (i): it can R linearity, however, the KPZ action describes a free field cou- − H H D ∇ 2 be written as δ /δθ, where = 2 x ( θ) , and this term alone would cor- respond to relaxation to an equilibrium stationary distribution ∝ e−H/∆. The non-linear term, on the other hand, can not be represented as the derivative of a Hamiltonian functional and is hence of type (ii). However, for reversible contributions to be compatible with a thermal stationary state, they have to 18 Due to symmetries of the KPZ equation (for a comprehensive discussion see satisfy a potential condition. This is not the case for the non-linear term in the Ref. [277]), the RG flow is described by the single parameter g defined in KPZ equation in dimensions d > 1. Eq. (190). 46 pled to a thermal bath. In other words, by performing a Bogoli- between these fixed points is reflected by the fact that the values ubov approximation the non-equilibrium character that is intrin- of the critical exponents differ; in particular, at a Gaussian fixed sic to the microscopic model with action (73) is lost in the long- point, all exponents are rational numbers, reflecting the valid- wavelength limit. Formally, this failure is due to the fact that ity of canonical power counting (cf. Sec.IIC), while one ob- the saddle-point approximation is not valid for low-momentum tains non-trivial rational or non-rational numbers mirroring the modes in the Goldstone direction since they are not gapped. importance of strong long-wavelength fluctuation corrections at a Wilson-Fisher fixed point. The latter is more stable than the Gaussian one (where, in fact, all couplings are relevant in the B. Critical dynamics in 3D sense of the RG), and thus the physically relevant one even if interactions are small on the microscopic scale. This is intuitive, taking into account that critical behavior deals with the longest In three spatial dimensions, the deviation from thermody- distances and timescales in a physical system. namic equilibrium conditions, which is quantified by the value of g defined in Eq. (190), is irrelevant in the RG sense (cf. The impact of fluctuations on the values of the critical expo- Eq. (191)). Hence, at large scales, effective equilibrium is es- nents can be described quantitatively in the framework of the tablished. This immediately implies that the overall picture of renormalization group, which provides a systematic way to deal Bose-Einstein condensation in thermal equilibrium in 3D re- with the intricate long-wavelength, low-momentum divergences mains valid also in the driven-dissipative context: in particu- caused by the divergent correlation length. The critical behavior lar, the mean-field analysis of Sec. II B 2, which predicts a con- of driven-dissipative systems in 3D was investigated in [26, 181] tinuous phase transition beyond which the classical phase ro- based on the functional renormalization group approach dis- tation symmetry (cf. Sec. II D 4) is spontaneously broken and cussed in Sec.IIE, and in [274] using the field theoretic per- off-diagonal long-range order is established, is modified quanti- turbative renormalization group [5]. These studies gave rise to tatively but not qualitatively once fluctuations around the mean- the following picture of driven criticality: field condensate are taken into account. (Note, however, that in The RG fixed point is purely dissipative (see Fig.6 (c)), i.e., equilibrium the condensation transition is induced by lowering all complex couplings rotate to the imaginary axis. However, the temperature to a critical value, whereas driven-dissipative it is the approach to this fixed point that contains universal in- condensation is established by tuning the single-particle pump formation, and this makes it possible to distinguish equilibrium rate.) Yet, the non-equilibrium nature of the driven-dissipative from non-equilibrium critical behavior. The following key prop- system leaves its mark on the approach to the long-wavelength erties are identified: thermalized regime in a fully universal way. (i) Asymptotic thermalization of correlation functions. — An equilibrium system, fine-tuned to a critical point, beyond Both the static and dynamical critical exponents, governing, e.g., which spontaneous symmetry breaking occurs via a second or- the asymptotic decay of the spatial and temporal first order co- der phase transition, exhibits universal behavior. This is wit- herence functions, are found to take the same values as in the nessed in non-analyticities in the free energy, and in the long- corresponding equilibrium problem. This can be made plausible range decay properties of correlation and response functions. by the fact that the fixed point couplings indeed lie on a sin- The decay properties are then governed by power laws, freed gle ray in the complex plane — the imaginary axis. More for- from the generic exponential cutoff ∼ e−r/ξ at large distances mally, it is understood in terms of an emergent symmetry imply- r due to the divergent correlation length ξ → ∞, defining the ing asymptotic emergence of detailed balance, or thermalization, critical point. While the concept of a free energy is not mean- at large spatial or temporal distances. ingful in a non-equilibrium context, correlations and responses (ii) Universal decoherence. — The approach to the purely can be considered also in the latter case. Then, universality en- dissipative fixed point still hosts information on the underlying tails that the algebraic long-wavelength, long-time decay of any quantum dynamics. The fadeout of this coherent dynamics is correlation or response function can be characterized in terms described by a new critical exponent, which can be shown to of a set of critical exponents, which do not depend on the mi- be independent of the static and dynamical exponents, i.e., it croscopic details of the problem but rather on symmetries and does not relate to the latter by scaling relations. Moreover, it dimensionality. The critical point is fully characterized by this can be seen that no more independent exponents could possi- set of critical exponents. bly be found [26, 274]. This is because the maximum number We emphasize a crucial difference between the critical be- of independent exponents is determined by the maxium num- havior in strictly non-interacting theories described by quadratic ber of independent microscopic mass scales [84], which are, in Hamiltonians or quantum master equations, and the more the quadratic part of the action (79), given by the real parameters generic — but also much more complex — case of interact- rc, rd, and γ. In addition, a coupling f to an external source field, R  ∗ ∗  ing problems (for an example where a Gaussian fixed point is corresponding to a term t,x f jcφq + jqφc + c.c. in the action, physically important, cf. Sec. III A). The non-interacting criti- has to be taken into account, leading in total to four independent cal theories are described by a so-called Gaussian fixed point parameters. No more independent scales can be added to the of the RG flow, while the interacting ones are associated to a quadratic part of the action without violating the conditions of so-called Wilson-Fisher fixed point. The structural difference (Anti-)Hermiticity and conservation of probability as explained 47 in Sec. II B 1. Correspondingly, the set of four independent crit- tion S at Λ = Λ0 to the full effective action Γ at Λ → 0, as ical exponents (the correlation length exponent ν, the anoma- fluctuations with momenta larger than the cutoff Λ are integrated lous dimension η, the dynamical critical exponent z, and new out, and the latter is gradually lowered. It is an exact non-linear exponent) is maximal. Finally, the value of this exponent distin- differential equation for the functional ΓΛ. In the absence of an guishes equilibrium from non-equilibrium conditions. This can exact solution, suitable schemes to approximate the functional be understood from Fig.6: the exponent describes the fadeout differential equation have to be found. For the analysis of crit- of coherent couplings, i.e., the approach to the imaginary axis, ical behavior, progress can be made by choosing an ansatz for −ηr governed by a power law ∼ Λ , where ηr is the new exponent. ΓΛ that has a similar structure as the microscopic action S in In the equilibrium case, all couplings rotate uniformly, giving Eq. (73). In this way, the effective action ΓΛ is parameterized rise to ηr ≈ −0.143 within the FRG approach of Refs. [26, 181]. in terms of the coupling constants appearing in the ansatz, and In contrast, in the non-equilibrium case, the slowest approach to consequently, the functional differential equation for ΓΛ can be the real axis is given by ηr ≈ −0.101. The value of the non- cast in the form of a set of ordinary differential equations for the equilibrium exponent can also be determined analytically from couplings, as we describe in detail below. the field theoretic perturbative renormalization group in a dimen- Any such ansatz will contain only a finite number of cou- 19 sional expansion, yielding the result (to two-loop order) [274] plings, and will hence truncate the most general structure of ΓΛ. In other words, by choosing an ansatz of a particular form, one 2 (4 − d)2 4 ηr = − ln . (203) makes an approximation, and the question arises, which cou- 25 3 plings should be included in the ansatz or truncation in order to Specifying to d = 3 dimensions, we obtain ηr ≈ −0.023. The obtain meaningful results. For the study of critical phenomena discrepancy between this value and the one obtained from the at a second order phase transition, the power counting scheme FRG stated above (ηr ≈ −0.101) indicates that the two-loop of Sec.IIC provides a guideline: following the arguments given computation underestimates fluctuation corrections in three di- there, we choose a truncation which includes all couplings that mensions.20 Moreover, (possibly significant) quantitative cor- are not irrelevant, and which therefore takes the form of the rections to the values of critical exponents should also be ex- semiclassical action in Eq. (79): pected from FRG calculations that go beyond the truncation used ( " # ) in Refs. [26, 181]. Z   ∂U¯ ∗ Γ = φ¯∗ iZ∗∂ + K¯ ∗∇2 φ¯ − + c.c. + iγ¯φ¯∗φ¯ . (iii) Observability. — The drive exponent manifests itself, for Λ q t c ¯∗ q q t,x ∂φc example, in the frequency and momentum resolved dynamical (204) single particle response as probed in homodyne detection, see (Recall that the variables of the effective action are the field Ref. [181] for details. It thus corresponds to a direct experi- expectation values Φ¯ ν, ν = c, q defined in Eq. (37).) Here, in mental observable, though its small value poses a challenge for addition to the complex prefactor K¯ = A¯ + iD¯ of the Lapla- experimental observation. cian, we are including a complex wave-function renormalization In summary, it is found that the correlation functions (both Z = ZR + iZI. In the semiclassical limit, the homogeneous (i.e., static and dynamic) thermalize, and show identical univer- not containing derivatives with respect to time or space) part of sal behavior to an equilibrium critical system with the same the action can be written in terms of an effective potential U¯ , symmetries. However, the dynamical response functions con- ¯∗ ¯ which is a function ofρ ¯c = φcφc. Due to the invariance of the mi- tain universal information distinguishing equilibrium from non- croscopic action under classical phase rotations (cf. Sec. II D 4), equilibrium systems. The microscopic drive conditions are thus which is inherited by the effective action, only this combination witnessed even at the largest macroscopic distances in a fully of fields is allowed in the potential. The latter is given by universal way. In the following, we review how the results summarized 1 1 U¯ (¯ρ ) = u¯ (ρ¯ − ρ¯ )2 + u¯ (ρ¯ − ρ¯ )3 . (205) above can be obtained from an open-system functional RG ap- c 2 2 c 0 6 3 c 0 proach [26, 181]. The basic ingredients of this method are dis- ¯ cussed in Sec.IIE. Here, both the two-body and three-body couplings,u ¯2 = λ + iκ¯ andu ¯3 = λ¯ 3+iκ¯3, respectively, are complex. The three-body term is marginal according to power counting, and therefore included 1. Effective action for driven-dissipative condensation in the truncation. In the FRG, it is advantageous to approach the transition from the ordered phase. Then, the form of the effective The Wetterich equation (132) describes how the scale- potential in Eq. (205) corresponds to an expansion around the dependent effective action Γ evolves from the microscopic ac- stationary condensate densityρ ¯0. Indeed, this choice implies, Λ ¯∗ ¯∗ that the field equations δΓΛ/δφc = 0, δΓΛ/δφq = 0 (Eqs. (40) in the absence of sources and evaluated with the scale-dependent 19 effective action Γ ) are solved byρ ¯ = ρ¯ and φ¯ = 0 on all The exponent ηc calculated in Ref. [274] is related ηr via ηr = ηc − η. Λ c 0 q 20 We note that in Ref. [181] erroneously the value of ηc from the field theo- scales Λ. retic RG was directly compared to ηr obtained from the FRG. Similarly, in In the truncation Eq. (204), all couplings — including the con- Ref. [274] the value of η was by mistake compared to η of Ref. [181]. c A densate densityρ ¯0 — are scale dependent. As indicated above, 48 by means of such an ansatz for the effective action ΓΛ, the func- in Eq. (205), the couplingsu ¯2,3 can be obtained by taking further tional differential equation (132) can be rewritten as a set of or- derivatives with respect toρ ¯c. However, instead of projecting the dinary differential equations for these running couplings, with flow equation (208) in this way onto equations foru ¯2,3, it is more initial conditions given by the microscopic action Eq. (79). This convenient to introduce rescaled quantities as outlined above. is achieved by applying projection prescriptions. In the follow- Inserting the representation Eq. (206) of the bare quantum field ing, we summarize this method for the problem at hand. For in the effective action (207) leads to appearance of a factor 1/Z∗ details we refer the reader to Ref. [181]. in front of the term involving the effective potential. This factor can be absorbed by introducing a rescaled potential via U¯ = ZU. The flow equations of the bare and rescaled effective potential 2. Non-equilibrium FRG flow equations are related via ¯ 0 0 0 The main idea of a projection prescription on a specific cou- ∂`U = Z −ηZU + ∂`U , (209) pling is to extract this coupling from the effective action Γ by Λ where η denotes the anomalous dimension associated with taking appropriate derivatives of the latter with respect to the Z the wave-function renormalization (∂ Z is specified below in fields and coordinates, and subsequently setting the fields to ` √ Eq. (222)), their stationary values φ¯c = φ¯0 = ρ¯0 and φ¯q = 0 (as noted in Sec. II D 6, choosing φ¯c to be real does not lead to a loss of ηZ = −∂`Z/Z. (210) generality). Then, applying the very same projection descrip- tion to the Wetterich equation (132) yields the flow equation Then, with ∂ρ¯cq = Z∂ρcq and inserting Eq. (208) on the RHS for the corresponding coupling. For the actual evaluation of the of Eq. (211), the flow equation for the renormalized potential resulting flow equations, it is convenient to introduce rescaled becomes (here, primes denote derivatives with respect to ρc = ∗ fields, which are related to the bare ones by absorbing the wave- φcφc) function renormalization Z in the quantum field: 1 h i 0 0 0 0 − φc = φ¯c, φq = Zφ¯q. (206) ∂tU = ηZU + ζ , ζ = ∂ρcq ∂tΓk,cq ∗ . (211) Ω φq=φq=0 As a consequence of this transformation, and by rescaling all In analogy to Eq. (205), the renormalized effective potential can couplings appropriately, it is possible to obtain a reduced set of be written as flow equations, from which Z is eliminated. In a second step, the number of flow equations can be diminished further by in- 1 2 1 3 U(ρc) = u2 (ρc − ρ0) + u3 (ρc − ρ0) , (212) troducing dimensionless renormalized variables, as detailed in 2 6 Sec. IV B 3 below. with renormalized couplings defined as u = u¯ /Z = λ + iκ and We start by deriving flow equations for the non-linear cou- 2 2 u = u¯/Z = λ + iκ . To obtain flow equations for u , n = 2, 3, plings in the effective potential defined in Eq. (205). To see 3 3 3 n one simply has to take derivatives of the relation u = U(n)(ρ ) how one can project the Wetterich equation onto flow equations n 0 with respect to the cutoff scale `, taking into account that also ρ for these couplings, consider the effective action, evaluated for 0 is a running coupling: homogeneous, i.e., space- and time-independent “background fields:”  (n) (n+1) ∂`un = ∂`U (ρ0) + U (ρ0)∂`ρ0. (213)  0 0∗  ΓΛ,cq = −Ω U¯ ρ¯cq + U¯ ρ¯qc − iγ¯ρ¯q . (207) Inserting Eq. (211) on the RHS of this relation leads us to Here, the subscript cq in ΓΛ,cq indicates that both the classical ∂ u β η u u ∂ ρ ∂ ζ0 , and the quantum fields are set to constant but non-zero values; Ω ` 2 = u2 = Z 2 + 3 ` 0 + ρc ss (214) denotes the quantization volume, and we introduced the follow- 2 0 ∂`u3 = βu3 = ηZu3 + ∂ρ ζ , (215) ing products of fields, which are invariant under classical phase c ss ¯∗ ¯ ∗ ¯∗ ¯ 0 rotations (see Sec. II D 4):ρ ¯cq = φcφq = ρ¯qc andρ ¯q = φqφq. where we evaluate ζ with ρc set to its stationary value ρc|ss = ρ0. From the representation Eq. (207) it becomes immediately clear Finally, flow equations for the real and imaginary parts of u2 and how to project the flow equation (132) for ΓΛ onto a flow equa- u3 can be obtained by taking the real and imaginary parts of tion for the derivative of the potential U¯ in Eq. (205) with respect Eq. (214) and (215), respectively. toρ ¯c: one has to (i) evaluate Eq. (132) for homogeneous fields, The flow equation for the stationary density ρ0 cannot be (ii) take the derivative with respect toρ ¯cq, and finally (iii) set the specified without formal ambiguity: as we have already seen quantum background fields to their stationary state value, in Sec. II B 2, both real and imaginary parts of the field equa- tion (75) yield conditions on ρ0. However, we have seen as well 0 1 h i ∂`U¯ = − ∂ρ ∂`Γ ,cq . (208) that the condition stemming from the real part can always be ¯cq Λ φ¯ φ¯∗ Ω q= q=0 satisfied by choosing the proper rotating frame. Therefore, the Here and in the following, we specify flow equations in terms of physically correct choice is to assume that ρ0 is actually deter- the logarithmic cutoff scale ` = ln(Λ/Λ0). From the potential U¯ mined by the imaginary part of the field equation, i.e., by the 49

0 condition Im U (ρ0) = 0. Taking the derivative of this condition by with respect to the cutoff `, we find  ¯ 2 2 2 h 2  2 2 2 2 i AΛ for q < Λ , 0 00 0 A¯ q + Λ − q θ(Λ − q ) =  (221) ∂ ρ − ∂ U ρ / U ρ − ζ /κ,  2 2 2 ` 0 = Im ` ( 0) Im ( 0) = Im ss (216) Aq¯ for q ≥ Λ , 0 where ζ is the same as in Eq. (211). (and there is an analogous replacement for the terms Dq¯ 2). Having illustrated the main idea, the flow equation for the Hence, fluctuations with momenta below the cutoff scale Λ ac- 2 rescaled noise strength γ = γ/¯ |Z| can easily be obtained along quire a mass ∼ Λ2, and the infrared divergences are lifted. the lines of the derivation of Eqs. (214) and (215), with the result It remains to specify the flow equations for Z and K¯ . They can (for details of the derivation see Ref. [181]; ηZR is the real part be obtained by choosing specific values of the indices i and j in of ηZ) the flow equation (218) for the inverse propagator, and by tak- ing derivatives with respect to the frequency ω and the squared i h i 2 ∂`γ = βγ = 2ηZRγ − ∂ρ ∂`ΓΛ,cq . (217) momentum q , respectively: Ω q ss

1 h  R i Thus far we have specified how to project the Wetterich equa- ∂`Z = − ∂ω tr 1 + σy ∂`P¯ (ω, q) , (222) 2 ω=0,q=0 tion onto flow equations for the couplings that parameterize the  R R  homogeneous part of the effective action given in Eq. (207), ∂`K¯ = −∂q2 ∂`P¯ (ω, q) + i∂`P¯ (ω, q) . (223) 22 12 ω=0,q=0 where the classical and quantum fields are set to constant val- ues. In the following we review the derivation of flow equations Comparison with Eq. (219) shows, that these are indeed correct for the frequency- and momentum-dependent couplings, i.e., the projection prescriptions. Note, however, that there is some ambi- wave-function renormalization Z and the coefficient K¯ of the guity in choosing these projection prescriptions: for example, A¯ ¯ R ¯ R ¯ Laplacian in Eq. (204). This requires us to consider non-constant appears both in P11 and P22. Our choice extracts K correspond- values of the fields. Moreover, we work in a basis of real fields ing to the Goldstone direction, and mixes Goldstone and gapped 1  which we introduced already in Sec. II D 6, φ¯ν = √ χ¯ ν, + iχ¯ ν, directions symmetrically in the projection on Z (see Ref. [181] 2 1 2 for ν = c, q. Hence, the inverse propagator in this basis is given for details). Finally, the flow equation for the renormalized co- by the second variational derivative of the effective action with efficient K = K¯ /Z is given by respect to the fieldsχ ¯ i (cf. Eq. (41); to ease the notation, we   ∂`K = βK = ηZ K + ∂`K¯ /Z. (224) collect these fields in a vectorχ ¯ = χ¯ c,1, χ¯ c,2, χ¯ q,1, χ¯ q,2 ; the com- ponents of this vector are labeled by i = 1,..., 4), and the flow While Eqs. (214), (215), (216), (217), and (224) define a equation of the inverse propagator reads accordingly: closed system of flow equations for the couplings u2, u3, γ, ρ0, and K (as indicated above, the wave-function renormalization Z " # δ2∂ Γ ¯ − 0 − 0 ` Λ drops out of these equations), the explicit evaluation of the var- ∂`Pi j(ω, q)δ(ω ω )δ(q q ) = 0 0 . δχ¯ i(−ω, −q)δχ¯ j(ω , q ) ss ious projections is rather tedious. For details of this calculation (218) we refer the reader to Ref. [181]. In particular, the inverse retarded propagator is given by

2 2 ! R −iZIω − Aq¯ − 2λ¯ρ¯0 iZRω − Dq¯ 3. Scaling solutions and critical behavior P¯ (ω, q) = 2 2 . (219) −iZRω + Dq¯ + 2¯κρ¯0 −iZIω − Aq¯ At the critical point of a continuous phase transition, the cor- Note that the Goldstone theorem (i.e., the existence of a zero relation length diverges, ξ → ∞. Then, instead of exponen- ¯ R eigenvalue of P (ω = 0, q = 0), see Sec. II D 6) is pre- tial decay according to ∼ e−r/ξ, correlation and response func- served during the flow. At the transition, whereρ ¯0 → 0, both tions depend on distance as power laws. This algebraic scaling branches of the excitation spectrum that is encoded in the ze- behavior is reflected in the RG flow: indeed, the critical point ¯ R ros of det P (ω, q) (cf. Eq. (77)) become gapless. As pointed corresponds to a scaling solution to the RG flow equations. In out in Sec.IIE, in the FRG, the resulting infrared divergences practice, finding a scaling solution is facilitated by absorbing the are regularized by introducing an additional contribution ∆S Λ, scaling factors ∼ Λθ (with some exponent θ for each coupling) given in Eq. (127), in the functional integral. In fact, by choosing in new variables, which thus take constant values at the critical the following optimized form of the cutoff function [278, 279] point. Hence, the latter corresponds to a fixed point of the flow (which obviously satisfies the requirements stated in Eqs. (130) equations for the rescaled couplings. Moreover, by means of a and (131)), suitable choice of rescaled variables it is often possible to further   reduce the number of flow equations, ending up with a minimal 2 ¯ 2 2 2 2 RΛ,K¯ (q ) = −K Λ − q θ(Λ − q ), (220) set of independent equations. For the present case, the flow can be specified in terms of just six real couplings. (2) in the regularized inverse propagator ΓΛ + RΛ appearing in the At the beginning of this section, we introduced the quantity Wetterich equation (132), the terms Aq¯ 2 in Eq. (219) are replaced λ in Eq. (188) as a quantitative measure of the deviations from 50 thermal equilibrium conditions. In a similar spirit, the strength A and driven-dissipative condensates. (Note, however, that in of coherent relative to dissipative dynamics, which is encoded in model F [1], which describes condensation with particle num- the real and imaginary parts of the couplings in the microscopic ber conservation in equilibrium, the dynamical exponent takes a Keldysh action (73), is measured by the ratios different value than in model A, where particle number conser- vation is absent.) This confirms the asymptotic thermalization r = r , r , r  = (A/D, λ/κ, λ /κ ) . (225) K u2 u3 3 3 of correlation functions mentioned in Sec.IVB (i). The values The flow equations for these ratios can be obtained straightfor- of the critical exponents we obtain from the truncation (204) are wardly by taking the RG scale derivatives, e.g., ∂`rK = ∂`A/D − 2 ν ≈ 0.716, η ≈ 0.039, z ≈ 2.121, (229) A∂`D/D , and expressing ∂`A as the real part of Eq. (224) etc. In addition to r, we define another three scaling variables as and agree reasonably well with results from more sophisticated ! calculations of ν to and η in the context of the static equilibrium 2κρ γκ γ2κ s (w, κ, κ ) 0 , , 3 . problem [280]. = ˜ ˜3 = 2 2 3 (226) Λ D 2ΛD 4D All the information on the universal properties of the driven- The flow equations for the six dimensionless running couplings dissipative transition, which are the same as in the equilibrium collected in r and s form a closed set. Besides the Gaussian model A, are encoded in s∗ and the block S of the stability ma- fixed point at which the non-linear couplings vanish, these equa- trix in Eq. (228). The non-equilibrium nature of the microscopic tions have a non-trivial fixed point corresponding to the driven- model, on the other hand, is betrayed by the block N, which de- dissipative condensation transition at scribes the flow of δr. This block has three positive eigenvalues,

r∗ = 0, s∗ ≈ (0.475, 5.308, 51.383) . (227) n1 ≈ 0.101, n2 ≈ 0.143, n3 ≈ 1.728, (230) This fixed point is reached in the RG flow, when the parame- indicating that the ratios r are attracted to the fixed point value ters in the microscopic action are chosen such that the system r∗ = 0. The general solution to the linearized flow equation for is tuned precisely to the transition point. (Note that this point r reads corresponds to the renormalized value of w ∝ ρ0 going to zero, 3 and not the bare one.) What are the physical implications of X r = uici, (231) this fixed point? First, the value r∗ indicates, that the effective i=1 action at the fixed point is purely dissipative. As we have al- ready mentioned at the beginning of Sec.IV, for vanishing co- where ui are the eigenvectors of N associated with the eigenval- herent dynamics (or, in the terminology of equilibrium dynam- ues ni in Eq. (230). The coefficients ci, which are referred to n ` n ical models [1]: in the absence of reversible mode couplings), as scaling fields [178], take the scaling form ci ∼ e i ∼ Λ i . the driven-dissipative model reduces to the equilibrium model Hence, for Λ → 0, the dominant contribution to r is given by n1 −ηr A. Thus, the values s∗ are the same as in model A, and there- r ∼ u1Λ = u1Λ , where we identified the drive exponent fore the fixed point itself does not allow to distinguish whether the microscopic starting point of the RG flow was in or out of ηr = −n1 ≈ −0.101. (232) equilibrium. However, the non-equilibrium nature of the driven- As anticipated in Sec.IVB (ii), this exponent governs the univer- dissipative condensate is witnessed in the RG flow towards this sal fade-out of coherent dynamics ∝ r in the driven-dissipative effective equilibrium fixed point. In the following we consider system. Note that the existence of three distinct eigenval- the universal regime of the RG flow, which is reached in the ues (230) is due to the non-equilibrium character of the micro- deep IR (i.e., for Λ/Λ  1). In this regime, when the cou- 0 scopic model. Indeed, in model A with reversible mode cou- plings are close to their values at the fixed point, the RG flow plings, the equilibrium symmetry discussed in Sec. II D 1 allows can be obtained from a linearization of the flow equations in of only one ratio r = r = r = r (cf. Eq. (101) and Fig.6). δs = s − s , δr = r. The stability matrix governing the linearized K u2 u3 ∗ Then, the block N of the stability matrix in Eq. (228) has only flow takes block diagonal form, one single entry, which is given by the “middle” eigenvalue n2 ! ! ! δr N 0 δr in Eq. (230). This shows that also in the equilibrium setting the ∂ = , (228) ` δs 0 S δs dynamics becomes purely dissipative at the largest scales [268], however, the value of the critical exponent that governs universal with 3 × 3 submatrices N and S . This block-diagonal structure decoherence is different. As pointed out above, this is due to the indicates, that the flow of r and s decouples in the IR. Therefore, absence of the equilibrium symmetry in the driven-dissipative the flow of s close to the fixed point is the same as if we would case. The fact that the different values can be traced back to have set r = 0 from the very beginning. In other words, not a difference in symmetry supports the strength of the result, as only the values of the couplings s at the fixed point, but also the different symmetries are known to give rise to quantitatively dif- critical exponents encoded in the flow of s, which are the correla- ferent critical behavior [86]. tion length exponent ν, the anomalous dimension η, and the dy- Decoherence at large scales has clear physical signatures namical exponent z, see Ref. [181], are the same for both model which facilitate probing the drive exponent in experiments; to 51 wit, it implies that low-momentum excitations are diffusing integration collected in AppendixB. The result is rather than propagating (note that this is also predicted by mean- ∗ 0 i(θ(x)−θ(x0)) field theory, cf. the discussion below Eq. (77)). A careful analy- ψ(x)ψ (x ) ≈ ρ0he i sis in the scaling regime reveals that the effective dispersion re- − 1 h(θ(x)−θ(x0))2i = ρ0e 2 (234) lation of single-particle excitations close to criticality takes the 0 −α form [181] ∼ x − x ,

z−ηr z 2.223 2.121 2 ω ∼ A0q − iD0q ∼ A0q − iD0q , (233) where α = m T/(2πρ0). One the other hand, the derivation in Sec.IVA shows that in the case of a driven-dissipative conden- where the diffusive contribution is supplemented by a subdomi- sate the phase-only action is non-linear and given by the KPZ nant (by the small difference of ηr in the exponent) coherent part. action (197). Hence, while the second equality in Eq. (234) still By definition, Eq. (233) is the location of the pole of the retarded applies to leading order in a cumulant expansion, the expecta- Green’s function, and hence the coherent and diffusive parts en- tion value h(θ(x) − θ(x0))2i cannot be calculated directly.21 In code respectively the position and width of the peak of the spec- the original context of the KPZ equation, where θ takes the role tral function defined in Eq. (61). The latter is probed, e.g., in of the height of a randomly growing surface [2, 73], the behav- angle-resolved spectroscopy in exciton-polariton systems [281] ior of this correlation function is parameterized in terms of the or radio-frequency spectroscopy in ultracold atoms [282]. How- roughness exponent χ as ever, the small difference in the scaling of the position and width 0 2 0 2χ of the peak predicted by Eq. (233) poses a challenge to its ex- h θ(x) − θ(x ) i ∼ x − x . (235) perimental observation. The term roughness exponent is due to the fact that its value distinguishes smooth from rough phases: for χ < 0, fluctuations C. Absence of algebraic order in 2D of the surface height die out on large scales, and the surface is smooth; on the other hand, if χ > 0, the interface is called rough. For any finite value of χ, the scaling behavior of the correlation Semiconductor microcavities hosting exciton-polaritons are function in Eq. (235) leads to stretched exponential decay of the effectively two-dimensional, and therefore this case has the correlations of ψ, greatest significance for current experiments. Even in thermal 0 2χ equilibrium, two-dimensional condensates are markedly differ- ψ(x)ψ∗(x0) ∼ e−c|x−x | , (236) ent from their three-dimensional counterparts: first, according to the Mermin-Wagner theorem [275], in a two-dimensional con- where c is a non-universal constant. In the case that χ = 0 one densate there cannot be true off-diagonal long-range order at any usually expects logarithmic growth of h(θ(x) − θ(x0))2i, which finite temperature. Instead, at low temperatures, spatial correla- would lead to the equilibrium result in Eq. (234). However, as tions decay algebraically with distance. Nevertheless, the sys- discussed at the beginning of Sec.IV, in a 2D driven-dissipative tem remains superfluid. Second, the algebraic or quasi-long- condensate we should expect universal behavior that is quite dif- range order is established in an unusual transition, in which vor- ferent from the equilibrium case. Indeed, the FRG analysis re- tices, which proliferate at high temperatures, form bound pairs ported in Refs. [252–254, 277] and numerical simulations [283– as the temperature is tuned below the critical value. How is this 297] find χ ≈ 0.4 for the value of the roughness exponent, scenario modified under non-equilibrium conditions? As a first which implies that for |x − x0| → ∞ correlations in 2D driven- step to answer this question, the issue of spatial correlations in dissipative condensates obey Eq. (236) and not Eq. (234)22, and two-dimensional driven-dissipative condensates is addressed in decay stretched-exponentially. Ref. [19]. In this work, the results of which we describe in the A notable difference between the KPZ equation for randomly following, the influence of non-topological phase fluctuations growing interfaces, and the present context of the effective long- (spin waves) on the behavior of spatial correlations is analyzed, wavelength description of a driven-dissipative condensate is that which leads to the conclusion, that in driven-dissipative conden- sates algebraic decay is possible only on intermediate scales, and crosses over to stretched-exponential decay on the largest scales. 21 The decay of correlations might be found to be even faster (i.e., The non-linearity λ in the KPZ action (197) vanishes when the equilibrium condition (101) is met, leading to algebraically decaying correlations also in exponential), once topological excitations (vortices) are taken this case. This shows, that merely adding dissipation by coupling the system to into account. a bath in thermal equilibrium does not have an adverse effect on correlations. In a weakly interacting Bose gas in thermal equilibrium, the (In the genuine case in which the condition (101) is met, the system is indeed absence of true long-range order is caused by the vanishing en- coupled to a single bath. Otherwise, realizing this condition when the system ergy cost of long-wavelength phase fluctuations. These are gov- is coupled to several baths would require a pathological fine-tuning of the coupling parameters.) It is indeed the combination of independent drive and erned by the quadratic effective low-energy action (201), from dissipation, which leads to the loss of algebraic coherence out of equilibrium. which the behavior of spatial correlations at long distances can 22 Note that as pointed out at the end of Sec.IVA, the KPZ non-linearity is be obtained straightforwardly, e.g., by introducing sources as de- neglected in Bogoliubov theory. Therefore, in 2D, this approach yields power- scribed in Sec.II and using the formulas for Gaussian functional law decay of spatial correlations [192, 298]. 52 the analogue of the interface height in the latter case is a phase, θ, These considerations lead to the finite-size phase diagram re- and as such it is compact, i.e., defined up to multiples of 2π. This ported in Ref. [19]. means that topological defects — vortices — in this field are The pump strengths at which the KT and KPZ crossovers oc- possible.23 Proliferation of vortices would lead to an even faster, cur can be estimated based on the parameters given in Ref. [303]. simple exponential decay of spatial correlations. The present It is convenient to introduce dimensionless pumping and loss analysis does not take the possible presence of vortices into ac- rates as follows [19]: count. PR Rγ How do these findings compare to experimental results? x = − 1, γ¯ = l . (238) Both in experiments on incoherently pumped polariton conden- γRγl γRuc sates [300, 301] and simulations of parametrically pumped sys- Here, P is the rate at which the excitonic reservoir is replen- tems [302], spatial correlations have been found to decay alge- ished, while R is the amplification rate of the condensate due to braically within the confines of the system. This, however, is stimulated scattering of polaritons from the reservoir; γ and γ not in contradiction to the present analysis based on the KPZ l R are, respectively, the decay rates of lower polaritons and reser- equation: indeed, if the microscopic value g of the rescaled 0 voir excitons, and u is the coherent polariton-polariton interac- non-linearity (190) is small, which is actually the case in current c tion [159]. For the parameters in [303], the KT transition should experiments [19], a renormalized value of g = 1 is reached in be expected at [19] x ≈ 0.02. Denoting by x the pumping the RG flow only at the exponentially large scale KT ∗ strength at which the KPZ scale L∗ in Eq. (237) drops below the

8π/g0 system size, we have [19] L∗ = ξ0e , (237) 2 x∗/xKT = γ¯ ln(L/ξ0) ≈ 0.04, (239) where ξ0 is a microscopic scale where the RG flow is initial- ized, e.g., the healing length of the condensate. Indeed, to ob- where we tookγ ¯ ≈ 0.1, ξ0 ≈ 2 µm, and assumed a pump spot tain this result, we have solved Eq. (191) with initial condition size of L ≈ 100 µm. Thus, approaching the transition from g0 at the scale ξ0. Then, in systems of a size L well below L∗, above by lowering the pump power, the critical value xKT is an effective equilibrium description is applicable, leading to the reached first, and the system loses algebraic order through un- observed algebraic decay of correlations (below the equilibrium binding of vortices [300–302]. On the other hand, the crossover KT transition). In other words, even in an infinite system we to the disordered regime will be controlled by KPZ physics once should expect a smooth crossover from algebraic to exponential x∗ ≥ xKT, which can be achieved by increasing the loss rate (i.e., decay at the scale L∗. reducing the cavity Q) toγ ¯ ≈ 0.5. So far, our analysis has been based on the semiclassical While the above analysis shows that algebraic order in 2D Langevin equation (80) for the condensate dynamics, which driven-dissipative condensates prevails only on intermediate according to the arguments given at the beginning of Sec.IV scales, remarkably it can be restored on all scales in strongly correctly captures the universal scaling properties of driven- anisotropic systems [19]: consider a generalization of Eq. (80), dissipative condensates. However, in order to obtain an esti- P i 2 where the gradient terms are replaced by i=x,y Kα∂i φc for α = mate of L∗ for specific experimental parameters, a more micro- c, d. Correspondingly, Eqs. (188) and (189) are replaced by scopic model of exciton-polariton condensates is required. Start- ing from a widely used model, which has been introduced in  Ki   Ki u  i  d uc  i  d c  Ref. [159] and consists of a coupled system of equations for the Di = K  +  , λi = −2K 1 −  , (240) c  Ki u  c  Ki u  lower polariton field and the excitonic reservoir, the bare value c d c d g0 and hence the scale L∗ can be seen to depend on the rate at which for i = x, y are the diffusion constants and non-linearities which the reservoir is replenished [19]. For high pump rates appearing in the anisotropic KPZ equation. The RG flow of this the KPZ scale L∗ grows rapidly, so that by pumping the system equation has been analyzed in Refs. [299, 304]. It can be de- strongly enough, algebraic correlations can be made to extend scribed in terms of the anisotropy parameter Γ = λyDx/λxDy over the entire system for any finite system size L. When the (the system is anisotropic for Γ , 1), and the non-linearity pump rate is reduced, the system can loose its algebraic order in 2 2 p g = λx∆/(Dx DxDy). The flow equations are given by two ways: either through the effect of the KPZ non-linearity if L∗ drops below the system size, or — if L∗ is still much larger 2 dg g  2  than the system size at the critical value of the pump strength for = − Γ + 4Γ − 1 , d` 32π (241) the equilibrium KT transition to occur — through the prolifera- dΓ Γg   tion of vortices. Note that as pointed out above, even when KPZ = − 1 − Γ2 . d` 32π physics becomes relevant, vortices might still modify Eq. (236). (Note that these equations differ from the RG equations in Ref. [19] by the sign on the RHS, which is due to the fact that here we define the logarithmic scale as ` = ln(Λ/Λ0) instead 23 This difference with the conventional KPZ equation also arises in “Active of ` = ln(L/ξ0).) The line Γ = 0 divides the flow into two re- Smectics” [299]. gions with distinct fixed-point structure: for Γ > 0, which we 53 denote as the regime of weak anisotropy, at large scales isotropy driven interfaces, β is called the growth exponent [2]) and c is is restored, i.e., Γ → 1 for ` → ∞, and all the results discussed a non-universal constant. This behavior was confirmed numer- above apply.24 In the regime of strong anisotropy corresponding ically [265–267]. In equilibrium, i.e., for g = λ = 0, the KPZ to Γ < 0, on the other hand, the flow is attracted to an effective equation (199) reduces to a noisy diffusion equation, which in equilibrium fixed point with g = 0 and Γ = −1. Then, algebraic the surface growth context is known as the Edwards-Wilkinson correlations of the condensate field can survive if the effective model [309]. Then, the behavior of spatial correlations is un- temperature at the fixed point, which is given by the renormal- changed, whereas the exponent β governing the decay of tem- p ized value of the dimensionless noise strength κ = ∆/ DxDy, poral correlations takes the value β = 1/4. The distinction be- is below the critical value κc = π for the equilibrium KT transi- tween one-dimensional condensates in equilibrium and driven- tion. Generalizing the microscopic model for exciton-polariton dissipative condensates thus becomes manifest only in the dy- condensates mentioned above to account for spatial anisotropy, namical properties.25 Moreover, in order to actually observe the dependence of the effective temperature on the strength of KPZ scaling in the autocorrelation function, a large value of g, laser pumping can be obtained [19]. Remarkably, the transition corresponding to a system far from equilibrium, is favorable. to the algebraically ordered phase is found to be reentrant: upon This can be achieved by making drive and dissipation the domi- increasing the pump rate the ordered phase is first entered and nant contributions to the dynamics, as is the case in cavities with then left again at even higher values of the pump rate. a reduced Q factor [265]. To be specific, for the parameters re- ported in Ref. [310], the Q factor would have to be reduced by a factor of ≈ 30 (corresponding to a polariton lifetime ≈ 1 ps D. KPZ scaling in 1D instead of ≈ 30 ps achieved in the experiment) in order to make KPZ scaling observable in a system of size ≈ 100 µm. The marginality of g in two spatial dimensions is reflected in Another observable, which conveniently encodes the scaling the emergence of the exponentially large scale L∗ in Eq. (237) properties of the phase θ of the condensate, is defined as beyond which KPZ scaling can be observed. In 1D, on the con- trary, the KPZ non-linearity g is relevant (cf. the discussion at * Z L Z L !2+ 1 2 1 the beginning of Sec.IV), which makes one-dimensional driven- w(L, t) = dx θ(t, x) − dx θ(t, x) . (243) L 0 L 0 dissipative condensates even more promising candidates to ob- serve KPZ universality in finite-size systems. This possibility In the context of growing interfaces, were θ takes the role of was explored numerically in Refs. [265–267], where the scaling the surface height, the quantity w(L, t) is known as the rough- properties of 1D driven-dissipative condensates were studied by ness function. It is a measure of the fluctuations of the sur- simulations of the Langevin equation (80) for the condensate face height over the linear extent of the system L. While the field. roughness function might not be easily accessible in experi- Experimentally, the most directly accessible signatures of ments with driven-dissipative condensates, it allows a very com- KPZ universality are contained in the correlation function of the pact demonstration of both static and dynamic KPZ scaling ex- condensate field (i.e., the Keldysh Green’s function defined in ponents if it is obtained numerically for a range of different Eq. (34)), system sizes [285]. Indeed, the finite-size scaling collapse of w(L, t) in Fig. 11 shows, that after a period of growth during C(t − t0, x − x0) = hψ(t, x)ψ∗(t0, x0)i. (242) which w(L, t) ∼ t2β, the roughness function saturates at the time z Indeed, in experiments with exciton-polaritons, both spatial cor- Ts ∼ L ; the saturation value ws(L) scales with the system size 2χ relations C(0, x) and the autocorrelation function C(t, 0) can be as ws(L) ∼ L , where χ = 1/2 is the value of the roughness ex- obtained by performing interferometric measurements on the ponent in 1D [2,5, 102]. From the growth and roughness expo- photoluminescence emitted from the semiconductor microcav- nents, the usual dynamical exponent can be obtained as z = χ/β. ity [12, 303, 308]. Based on the mapping of the long-wavelength The numerical analysis reported in Ref. [265] was performed condensate dynamics to the KPZ equation, one expects ex- in the regime of weak noise, which is characterized by the ab- ponential decay of spatial correlations C(0, x) and stretched- sence of phase slips in the spatiotemporal range covered by the exponential decay of the autocorrelation function according to simulations. Indeed, the mapping of the condensate dynamics C(t, 0) ∼ exp(−ct2β), where β = 1/3 (in the original con- to the KPZ equation (199), in which the phase is regarded as a text of the KPZ equation, which is the stochastic growth of non-compact variable, does not take into account the possible occurrence of such defects. However, their presence at higher noise levels is expected to affect the scaling properties of driven- dissipative condensates. 24 Current experiments with exciton-polaritons are in fact slightly anisotropic due to the interplay between polarization pinning to the crystal structure, and the splitting of transverse electric and transverse magnetic cavity modes [13, 305]. On the other hand, in experiments using the optical parametric oscillator regime pumping scheme [306, 307] (see also [83, 302]), strong anisotropy is 25 Concomitantly, also within Bogoliubov theory exponential decay of correla- imprinted by the pump wavevector. tions is found [191], cf. the discussion at the end of Sec.IVA and Footnote 22. 54

evolution of heating, with a focus on the short time dynamics 10!1 L#28 following initialization in a pure zero temperature ground state, L#29 where quantum effects are still present. The presence of parti- !2 L#210 cle number conservation leads us to take a different strategy than Α 10 2 11 in the previous sections. Here, accomodating number conserva-

L L#2

#! 12 tion, which is at the heart of the strongly collective behavior of t L#2 , ! 13 one dimensional systems, we first map the quantum master equa- L 3 L#2 " 10 tion in the operatorial formalism to an effective long-wavelength w description in terms of an open . In this way, !4 we can carefully account for the linear sound mode that is ex- 10 pected on the grounds of exact particle number conservation, cf. 10!3 10!2 10!1 100 Sec. II D 6. Only after this procedure, we perform the mapping t Lz to the Keldysh functional integral, similar to our strategy for the spins in Sec. III.

Figure 11. Finite-size scaling collapse of the roughness function w(L, t) ≡ defined in Eq. (243). The values of the roughness exponent α χ = 1/2 A. Heating an interacting Luttinger liquid and the dynamical exponent z = 3!/2 are in the 1D KPZ universality class. Each curve corresponding to a specific system size L is an average over 1000 noise realizations. The simulations were performed after Consider a one-dimensional lattice of interacting bosons for rescaling the Langevin equation (80) to bring it to dimensionless form. which the dynamics is described by the following master equa- For details of the rescaling and the values of the parameters used in tion: the simulations, see Ref. [265]. (Copyright (2014) by The American   X h n 2 oi Physical Society.) ∂tρ = −i H, ρ + γe 2niρni − ni , ρ . (244) i Here, H is a bosonic lattice Hamiltonian, whose long- V. UNIVERSAL HEATING DYNAMICS IN 1D wavelength physics is described by an interacting Luttinger liq- uid. For concreteness, one can consider a Bose-Hubbard model The notion of universality is not restricted to systems in ther- in one dimension away from integer filling, with Hamiltonian mal equilibrium. As discussed in the previous sections, non- X   X H = −J b†b + b† b + U n (n − 1), (245) thermal steady states of driven-dissipative systems can show i i+1 i+1 i 2 i i a large variety of universal features, such as scale invariance i i and effective long-wavelength thermalization. However, in a which describes nearest neighbor hopping of particles with hop- plethora of setups, aspects of universality can even be found in ping amplitude J and local interactions with interaction energies the time evolution, which approaches a steady state only in the U. The dissipative contributions which drive the system out of limit τ → ∞ [45–47, 311–315]. An example which identifies † equilibrium, are the Hermitian jump operators ni = bi bi, mea- generic, universal features in the far from equilibrium dynamics suring the local particle number in terms of bosonic creation and in a strongly interacting one-dimensional system is discussed in † annihilation operators bi , bi. For cold bosonic atoms in optical the present section. lattices, these jump operators represent the leading order contri- The setting we consider here differs from the one in the pre- bution of dissipation induced by spontaneous emission from the vious section not only in its focus on time evolution, but also in lattice drive laser [166] (see also the discussion in Sec.IC3), terms of underlying symmetries. Above we have studied sys- but this kind of dissipation can as well be realized by coupling tems that are open in the sense that both energy and particle the bosonic particles to a phonon reservoir with a large effective number were not conserved, witnessed by the absence of the temperature, which is equivalent to a locally fluctuating chemi- thermal symmetry (cf. Sec. II D 1) and the quantum phase rota- cal potential tion symmetry (cf. Sec. II D 4), and leading to a breaking of de- tailed balance and a low momentum diffusive Goldstone mode µ(x, t) = µ0 + δµ(x, t), 0 0 0 0 (cf. Sec. II D 6), respectively. Here we consider an open sys- hδµ(x, t)δµ(x , t )i = γeδ(t − t )δ(x − x ). tem, where only energy is not conserved, but particle number is. In fact, the absence of energy conservation here is reflected It causes dephasing and leads to a linear increase of the energy by a permanent inflow of energy into the system. The Lindblad in the system, hHit ∝ Jγet [166, 167]. As a consequence of the operators are Hermitian in the present case, and this leads to linear energy increase, the system will never thermalize and is continuous heating and ultimately to an entirely classical, infi- constantly driven away from equilibrium, approaching the T → nite temperature stationary state, described by a density matrix ∞ state described by ρ ∝ 1 at infinite time t. We note that, for ρ ∼ 1, where the latter unit matrix is understood in the entire hermitean Lindblad operators, ρ ∝ 1 is always a solution to the Fock space of the problem. This motivates us to study the time quantum master equation. 55

In order to analyze the heating dynamics for short and tran- properly expressed in the Luttinger framework [43]. The corre- sient times, the master equation is transformed to the Luttinger sponding time regime is set by the condition t < u2(κγ)−1. representation on the operatorial level, and only later on we per- The quadratic part of the equation of motion is diagonalized form the mapping to the Keldysh functional integral. The Lut- by the canonical Bogoliubov transformation tinger description is valid as long as the occupation of quasi- Z particle modes does not exceed a critical value, which is deter- π 1/2 † −iqx θx = θ0 + i ( 2|q|K ) (aq − a−q)e , mined by the Luttinger cutoff Λ [43]. Starting with a zero or q low temperature initial state to which this applies, there exists a Z (249) φ = φ − i ( πK )1/2sgn(q)(a† + a )e−iqx, cutoff time tΛ, up to which the system can be described in terms x 0 2|q| q −q of Luttinger liquid variables. In this regime, one can take the q continuum limit bi → bx and express the bosonic operators in a which leads to the master equation phase and amplitude representation: Z √ h † i iθx bx = ρxe , ∂tρ = −i u|q|aqaq, ρ (246) q ρx = ρ0 + ∂xφx/π, Z h i Z γ|q| † − 1 { † } − (3) + πK AqρAq 2 AqAq, ρ [Hph , ρ]. (250) in terms of the Luttinger variables ∂xφx and θx, which represent q q smooth density and phase fluctuations and fulfill the commuta-   0 In this equation, the dissipative part is expressed via the opera- tion relation ∂xφx, θx0 = iπδ(x − x ). The long-wavelength de- † † scription of the Bose-Hubbard model is expressed by the Hamil- tors Aq = aq + a−q = A−q in terms of bosonic phonon operators † tonian [aq, ap] = δ(q− p). The cubic Hamiltonian incorporates resonant phonon scattering processes 1 Z h i 2 − u 2 2 Z 0 H = uK (∂xθx) K (∂xφx) + κ (∂xφx)(∂xθx) , (247) q p 2π x (3) π | † Hph = 3κ 2K qp(p + q)(ap+qaqap + h.c.), (251) q,p which describes interacting Luttinger phonons on length scales x ≥ (ρ Um)−1/2, above which a continuum representation of the which conserve momentum and the phonon energy. This is ex- 0 R 0 Hamiltonian is appropriate. For weak interactions, the effective pressed by qp, which signals to integrate only over configura-  1/2  1/2 ρ0U π ρ0 tions {p, q} with |q| + |p| = |p + q|. The non-resonant processes parameters can be estimated to be u = m , K = 2 Um . The non-linearity in the Hamiltonian accounts for the leading create only short-lived quantum states which do not contribute in order quasi-particle scattering term in the low energy regime the long time dynamics due to dephasing, and which are there- (κ = 1/m), which is irrelevant in the sense of the renormaliza- fore not relevant for the forward dynamics of the system. tion group and does not modify static, equilibrium correlations. In the Luttinger representation, the dissipative contribution is However, it is vital for the quantitative description of dynamic quadratic and its effect is a constant population of the individual correlation functions, and is non-negligible in a non-equilibrium phonon modes. This can be seen most easily by computing the setting where the dynamics is affected by the non-linearity even time evolution of the phonon densities in the quadratic frame- on a qualitative level. The dissipative part of the master equation work, which yields becomes quadratic in the Luttinger representation, such that the h † i γ|q| equation of motion reads ∂t aqaq t = 2πK , (252) ∂ ha†a† i = − γ|q| − 2iu|q|ha†a† i . (253) Z h n oi t q −q t 2πK q −q t −   γ − 2 ∂tρ = i H, ρ + π2 2(∂xφx)ρ(∂xφx) (∂xφx) , ρ . (248) x For a system prepared initially in the ground or finite temper- ature state, the initial phonon densities are ha†a i = n (u|q|) This decoherence term is the leading order contribution of a q q 0 B h † † i | | βu|q| − −1 U(1)-symmetric (i.e., particle number conserving) decoherence and ∂t aqa−q 0 = 0, where nB(u q ) = (e 1) is the Bose- mechanism in one-dimensional quantum wires, which features Einstein distribution evaluated on-shell. In this case, the solution a linear increase of the energy in time. Furthermore, the U(1) of Eqs. (252), (253) is symmetry guarantees the existence of a linear sound mode and ha†a i = n (u|q|) + γ|q|t , (254) permits the transformation to the Luttinger framework in the co- q q t B 2πK herence dominated regime. From a microscopic perspective, it † † γe−iu|q|t haqa−qit = sin(u|q|t). (255) is evident that the Luttinger description has to break down for 2πuK sufficiently strong decoherence, i.e., after the system has been The first term describes an increase of the phonon density lin- heated up sufficiently. This breakdown can be estimated by the ear in time and momentum and leads to a linear increase of the usual Luttinger criterion nq < Λ/|q|, and leads to a good esti- system energy, i.e., heats up the system according to mate for the relevant time scales up to which the dynamics of (2) (2) uγΛ2t the system is dominated by coherent sound modes and therefore ∆Et = hH it − hH i0 = 4πK . (256) 56

Equation (256) relates the effective long-wavelength Luttinger tailed derivation and discussion of the applicability and limita- parameters u, K, γ of the heating setup to the macroscopic heat- tions of such an approach can be found in [43, 44]. Other forms ing rate ∂t∆Et via the microscopic cutoff Λ. Since this heat- of kinetic equations for interacting Luttinger , which fo- ing rate is not model specific but depends on the individual cus on a different set of non-equilibrium conditions, include a realization of the heating dynamics, it is not surprising that it perturbative treatment of phonon backscattering terms, resulting depends on macroscopic and microscopic parameters. In this from additional disordered or lattice potentials [317, 318], as sense, Eq. (256) should be viewed as the definition of the effec- well as a treatment of cubic phonon interactions in the presence tive heating parameter γ for a generic model in the presence of of a smooth background potential and a curved phonon disper- heating [43]. sion [319]. The off-diagonal phonon density oscillates in the complex The quantity of interest in this section is the time-dependent † † γ h † i plane, thereby taking absolute values |haqa−qit| ≤ 2πuK , which occupation of phonon modes nq,t = aqaq t in the presence are negligibly small in the weak heating regime γ  uK, i.e., in of heating and phonon scattering. For bosonic modes and in the coherence dominated dynamics. It is therefore sufficient to the steady state, the occupation of the modes is related to the consider only the diagonal elements in the phonon basis when Keldysh Green’s function via extending the analysis to the interacting model. Z K The jump operators in the microscopic master equation (244) i Gq,ω = 2nq + 1. (257) ω are the Hermitian, local density operators ni, which preserve the U(1) invariance of the dynamics even in the presence of dissi- For a system in thermal equilibrium, nq = nB(q, T) is the Bose pation. This leads to a decay of the off-diagonal elements of distribution function, see Sec. II D 1, while for a general non- the density matrix, i.e., to decoherence in the local number state equilibrium steady state, nq is a positive function, which has to representation, and an evolution of the density matrix towards be determined from the specific context. One can now introduce its diagonal ensemble. In the Luttinger representation, this de- the hermitian distribution function Fq,ω as in Eq. (72), but gen- coherence expresses itself in a permanent production of photons eralized to a system with a continuum of momentum modes. In (254), i.e. a permanent heating of the Luttinger liquid, which terms of the hermitian distribution function, the anti-hermitian features no compensation mechanism in the quadratic sector and Keldysh Green’s function can be parameterized according to the consequent lack of a well defined steady state. The energy K  R A G 0 = G ◦˜ Fq − Fq ◦˜ G . (258) increases constantly, which will lead to an obvious breakdown q,t,t q q t,t0 † of the Luttinger description as soon as the energy stored in the R A  R long-wavelength modes exceeds a critical value. At this point, Here, Gq , Gq = Gq and Fq are two-time functions eval- the dynamics is no longer dominated by the coherences of ρ but uated at equal momentum q and◦ ˜ represents the convolution by its diagonal elements, including the breakdown of superflu- with respect to time and matrix multiplication according to the idity and quasi-long range order. Nambu structure of the Green’s functions. For a system, which The way in which energy is distributed amongst the long- is diagonal in Nambu space,◦ ˜ = ◦ reduces to a simple multi- plication. In the presence of off-diagonal occupations, i.e. for wavelength modes by the heating (252) is not typical for an in- † † teracting system at low energies, since it deviates strongly from a ha−qaqi , 0, the operation◦ ˜ has to respect the symplectic struc- z Bose-Einstein distribution and the associated detailed balance of ture of bosonic Nambu space and is promoted to◦ ˜ = σ ◦, as (3) described in Ref. [44]. For a bosonic system in steady state, all energy. The phonon scattering terms in Hph favor detailed bal- ance and strongly modify the actual distribution function com- terms in (258) are time-translational invariant and Fourier trans- pared to (254), which makes them non-negligible in the present formation yields non-equilibrium setting. K R z z A Gq,ω = Gq,ωσ Fq,ω − Fq,ωσ Gq,ω. (259) In contrast, for a system out of equilibrium undergoing a non- B. Kinetic equation trivial time evolution, the mode occupations nq,t remain well de- fined, but Eq. (258) is not time-translational invariant anymore. One then has to find a representation for the Green’s functions In order to determine the time evolution of the excitation den- and F, which reveals the time-dependent occupations. This is sities in interacting systems out of equilibrium, a common and done in the following, leading to the Wigner representation of often successful strategy is the so-called kinetic equation ap- the bosonic distribution function. proach [102] (for an application to periodically driven Floquet A convenient parameterization of non-equilibrium correlation systems, cf. Ref. [316]), which determines the time evolution and response functions is the so-called Wigner representation in of the distribution function of the excitations in terms of the sys- time, which introduces a forward and a relative time coordinate tem’s self-energies. For the present setup, this approach has to be (t, δ) for the phonon Green’s functions, according to modified in order to take into account the driven-dissipative na- R † ture of the system and the resonant character of the interactions. Gq,t,δ = −iθ(δ)h[aq,t+δ/2, aq,t−δ/2]i, (260) The latter lead to a breakdown of perturbation theory and require GK = −ih{a , a† }i. (261) non-perturbative techniques beyond one-loop corrections. A de- q,t,δ q,t+δ/2 q,t−δ/2 57

A two-time function C(t1, t2) can always be transformed to should take a closer look at its expansion up to first order in t1+t2 Wigner coordinates, C(t1, t2) = C(t, δ), by defining t = 2 and derivatives: δ = t1 − t2. Here, the explicit dependence on t expresses the for-  R   R  R i ∂t Fq,t,ω ∂ωΣq,t,ω 2 ward time evolution of non-equilibrium systems, while for equi- Σ ◦ F = Σ Fq,t,ω 1 − R + (ω ↔ t) + O(∂ ). q,t,ω q,t,ω 2 Fq,t,ω Σ librium systems in the presence of time-translational invariance, q,t,ω (268) the forward time dependence of generic two-time functions just f ∂t Fq,t,ω The ratio κq ≡ is the rate with which the distribution drops out, C(t, δ) ≡ C(0, δ), for all t. In Wigner coordinates, the Fq,t,ω parametrization of the Keldysh correlation function is F is changing in forward time, i.e., the forward time evolu- tion rate, which is determined by the interplay between heat- K  R A ing and collective quasi-particle scattering. On the other hand, G = G ◦ Fq − Fq ◦ G . (262) q,t,δ q q t,δ  −1 ∂ ΣR r ω q,t,ω κq ≡ R is identified with the inverse rate of the relative Σq,t,ω Here, we choose to neglect the subleading off-diagonal contribu- time dynamics, which is dominated by fast single phonon prop- tions according to the above discussion, and consider only diag- f agation. As a consequence κr  κ and the correction terms onal modes in Nambu space. Eq. (262) contains the full Green’s q q in (268) can be safely neglected. This is a typical situation for functions of the system, which can be expressed via the Dyson many kinetic equation approaches and is termed the Wigner ap- relation in terms of the self-energies ΣR/A/K, proximation. For the present setup, a more careful analysis has !−1 ! shown that the Wigner approximation is indeed satisfied as long GK GR 0 GA − ΣA = 0 . (263) as the Luttinger representation of the problem is valid [43]. GA 0 GR − ΣR −ΣK 0 In order to project Eq. (266) onto the quasi-particle densities, R A R/A/K it is multiplied by the spectral function Aq,t,ω = iGq,t,ω − iGq,t,ω, The self-energies Σ represent the correction to the bare followed by a subsequent integration over frequencies ω. The R/A − | | −1 Green’s functions G0 = (i∂t u q ) due to phonon scatter- spectral function fulfills the sum rules ing and heating events. Inserting the Dyson representation into Z Eq. (262), it can be inverted and rearranged to read † Aq,t,ω = h[aq,t,0, aq,t,0]i = 1, (269) ω K  R A Z Z ∂tFq,t,δ = iΣ − i Σ ◦ Fq − Fq ◦ G . (264) q,t,δ q q t,δ Wigner approx. K Aq,t,ωFq,t,ω = (A ◦ F)q,t,ω = Gq,t,0 = 2nq,t + 1, (270) ω ω Here, the notion (...)t,δ expresses the fact that the whole ex- † pression in brackets should be transformed to Wigner coordi- with nq,t = haq,taq,ti being the phonon density. Applying this to R R R nates after performing the convolution Σ ◦ F ≡ Σ 0 F 0 Eq. (266) in the Wigner approximation yields q q t0 q,t1,t q,t ,t2 in ordinary time representation. The functionals ΣR/A/K are Z i  K R A  the self-energies in the retarded, advanced, and Keldysh sec- ∂tnq,t = 2 Σq,t,ω − Σq,t,ωFq,t,ω + Σq,t,ωFq,t,ω Aq,t,ω.(271) tors, which incorporate the effect of interactions and the heat- ω ing on the quadratic sector. The temporal derivative on the LHS For well defined quasi-particles, i.e., excitations with a well of (264) is the Wigner representation of the bare, non-interacting defined energy-momentum relation, the spectral function re- Green’s functions without heating. Taking the Fourier transform flects the well-defined structure of the excitations and is sharply of Eq. (264) with respect to the relative time coordinate δ yields peaked at the quasi-particle dispersion ω = u|q|, with a typical R the Wigner representation of the distribution function width σq,t  u|q|, which is the imaginary part of the self-energy R R Z evaluated on the mass shell σq,t = −Im(Σq,t,u|q|). If this is the iωδ Fq,t,ω = e Fq,t,δ, (265) case, the full self-energies in Eq. (271) multiplied with the spec- δ tral function can be approximated by their on-shell values, as they are expected to vary only smoothly in the region where A for which R/A/K R/A/K is non-zero. The approximation Σq,t,ω Aq,t,ω ≈ Σq,t,u|q| Aq,t,ω is K  R A ∂tFq,t,ω = iΣq,t,ω − i Σ ◦ F − F ◦ Σ . (266) called the quasi-particle approximation and in the present case, q,t,ω similar to the Wigner approximation, it is applicable in the en- The corresponding transformation for the convolution inside the tire Luttinger regime [43, 44]. The latter is a consequence of parenthesis is the subleading, RG-irrelevant nature of the interactions, which R lead to self-energies σq,t  u|q|. Performing the quasi-particle ← → ← →  i approximation, Eq. (271) obtains the simple form  R  R 2 ∂t ∂ω−∂ω ∂t Σ ◦ F = Σq,t,ωe Fq,t,ω. (267) q,t,ω K R ∂tnq,t = σ˜ q,t − σq,t(2nq,t + 1). (272)

Its explicit evaluation is nontrivial and in most cases simply im- K possible. However, it is possible to approximate the complex In this equation, the anti-Hermitian Keldysh self-energy Σq,t,ω K K exponential by the leading order expansion for many typical re- has been replaced by its on-shell value Σq,t,u|q| = −2iσ˜ q,t, with K laxation dynamics [102]. In order to understand Eq. (267), one the real functionσ ˜ q,t. 58

The kinetic equation (272) describes the time evolution of the The kinetic equation (274) represents the foundation of the R K phonon density nq,t in terms of the on-shell self-energies σ , σ˜ , analysis of the non-equilibrium dynamics in the presence of which in turn are determined by both the interactions and the heating and phonon scattering. In order to solve for the time- heating term. In order to identify the contribution from the heat- evolution of the phonon densities, one has to compute the self- R/K ing, one has to identify the impact of the dissipative contribution energies σq,t for each momentum mode and at each time step. in Eq. (250) on the action S . Following the steps in Sec.IIA The self-energies have to be determined by a non-perturbative carefully and setting the off-diagonal density contributions to approach, which we discuss in the following. zero, according to Eq. (254), the dissipative contribution to the microscopic action S is Z C. Self-consistent Born approximation γ|p| ∗ S D = i πK aq,p,taq,p,t. (273) p,t For an interacting model of resonantly scattering phonons, the Here, p is the momentum variable and q labels the quantum phonon self-energies are functionals of the phonon density, such component of the Keldysh field variable. The dissipation thus that the RHS of (274) contains an implicit, non-linear depen- enters the action only in the quantum-quantum sector and does dence on nq,t. In order to make this implicit dependence ex- not modify the spectrum in the quantum-classical sector. This plicit, the self-energies are typically evaluated perturbatively at is again a consequence of the Hermitian nature of the Lindblad one loop order and higher order corrections to the time evolu- operators, which lead to a continuously increasing occupation of tion are neglected [102]. However, for the present scenario, the phonons but do not introduce a compensating dissipative mecha- 26 interactions are resonant, i.e., describe scattering events inside a nism in the retarded and advanced sector of the action . This is continuum of degenerate states, and therefore perturbative com- drastically different from the situation in the models of Secs. III putations diverge at any order. This defines the need for non- (cf. Eqs. (141) and (140)) andIV (cf. Eqs. (204) and (205)), perturbative approaches to compute the phonon self-energies, where the interplay of dissipation in the retarded/advanced sec- the simplest of which is the so-called self-consistent Born ap- tors and fluctuation or noise in the Keldysh sector of the action proximation, which we discuss in the following. lead to non-equilibrium fluctuation-dissipation relations describ- The Keldysh action for interacting Luttinger liquids with heat- ing well-defined stationary states different from the trivial state ing is composed of a dissipative part S , which has been dis- ρ ∝ 1. D cussed in Eq. (273), and a Hamiltonian part S , which results With the form of (273), the Keldysh self-energy isσ ˜ K = H q,t from the Hamiltonian dynamics in Eqs. (250), (251). In the γ|p| K K 2πK + σq,t, where the bare σq,t in this form is determined by Keldysh representation, the action is the interactions alone. The resulting kinetic equation consists of three contributions Z ! ! 0 i∂ − u|p| − i0+ a S = (a∗ , a∗ ) t c,p,t γ|q| K R c,p,t q,p,t − | | + γ|p| a ∂tnq,t = 2πK + σq,t − σq,t(2nq,t + 1) . (274) p,t i∂t u p + i0 i πK q,p,t |{z} |{z} | {z } Z 0 in-term heating in-term scattering out-term scattering h ∗ + v(p, k) 2ac,k+p,tac,k,taq,p,t The first term represents the population of phonon modes due to p,k,t ∗   i the constant heating term, while the second and the third term + aq,k+p,t ac,k,tac,p,t + aq,k,taq,p,t + h.c. (275) are effects of the elastic collisions redistributing energy. The second term, proportional to the Keldysh self-energy, describes p π with the vertex function v(p, k) = 3κ 2K |pk(p + k). One way of scattering of phonons into the mode q due to the interactions, computing the one-loop self-energy is to determine the one-loop while the third term, proportional to the retarded self-energy, ∗ correction to the effective action Γ[aα, aα] defined in Eq. (39) for describes scattering of phonons out of the mode q, and is there- general bosonic fields Φ. The effective action in the absence of fore directly proportional to nq,t. Setting the interactions to zero, an external source is defined as both self-energies σR/K = 0 vanish and only the heating term Z remains, rendering the time evolution of the phonon density in ∗ ∗ ∗ iΓ[aα,aα] ∗ iS [aα+δaα,aα+δaα] the absence of interactions into Eq. (252). e = D[δaα, δaα] e , (276)

∗ ∗ δΓ[aα,aα] δΓ[aα,aα] and fulfills the equation of motion ∗ = = 0. The δaα δaα 26 The heating mechanism is operative for generic situations. For example, one-loop effective action is then obtained by expanding the ac- mean-field Mott initial states, which are the exact ground states of the Bose- tion S up to second order in the fluctuation fields and subse- Hubbard model for fine-tuned J = 0, are pure states which are not touched quently integration over the fluctuations by the dissipator considered in this section. One point of view on this phe- Z nomenology is that the infinite temperature state is an attractive fixed point (1-loop) ∗ ∗ i ∗ (2) ∗ ∗ T iΓ [aα,aα] iS [aα,aα] ∗ 2 (δaα,δaα)S [aα,aα](δaα,δaα) of dynamics, but there are other unstable ones which need additional symme- e =e D[δaα, δaα] e tries to be physically relevant (such as a spatially local gauge symmetry in the iS [a∗ ,a ]+ 1 Tr log S (2)[a∗ ,a ] J = 0 example). =e α α 2 ( α α ). (277) 59

This identifies the one-loop effective action (1-loop) ∗ ∗ i  (2) ∗  Γ [aα, aα] = S [aα, aα] − 2 Tr log S [aα, aα] (278) in terms of the microscopic action S and its second variation with respect to the fields  δ2S δ2S  (2) ∗  δa∗ δa 0 δa∗ δa∗  S [a , a ] =  α α α α0  . (279) α α  δ2S δ2S   ∗  δaαδaα0 δaαδaα In order to determine the correction to the bare action, the loga- ∗ rithm in Eq. (278) is expanded in powers of the fields aα, aα. The quadratic self-energy is the second order expansion of the loga- rithm and its matrix elements are determined by the integrals Figure 12. Illustration of the iterative process to compute the time evo- Z 0   lution of the phonon density n . For a specific forward time t, the on- R K 2 − R 2 − A 2 R q,t ΣQ = 2i GP v (q, p)GQ−P+ v (p, q)GP−Q+ v (p, q)GP+Q , shell self-energy σR is determined via Eq. (283) and subsequently the P q,t (280) result is inserted into the kinetic equation (284). In order to integrate the phonon density, a Runge-Kutta solver for differential equations is Z 0 h   K 2 K K R R A A used, which determines nq,t numerically. ΣQ = 2i v (q − p) GP GQ−P + GPGQ−P + GPGQ−P P  i +2v2(p, q) GK GK + GA GR + GR GR , (281) P+Q P P+Q P P+Q P The kinetic equation (284) and the equation for the on-shell self- where we used the collective indices Q = (q, ω), P = (p, ν) for energy (283) determine the forward time evolution of the sys- R momentum and relative frequency. In Wigner approximation, tem’s phonon density nq,t and self-energy σq,t in a self-consistent the Green’s functions are diagonal in forward time and therefore manner. For a general phonon density, both equations have to be evaluated at equal forward time t. The integrals in (280), (281) solved iteratively according to the scheme depicted in Fig. 12. are performed only over resonant momentum configurations, see Before the numerical results for dynamics in the presence of the discussion around Eq. (251). In perturbation theory, the heating, i.e., the numerical solution of Eqs. (283), (284), are dis- Green’s functions under the integral are the bare, non-interacting cussed, it is useful to study certain limiting cases. This facilitates R + −1 Green’s functions GQ = (ω − u|q| + i0 ) , which diverge on the the understanding of the numerical results in the subsequent sec- mass-shell ω = u|q| and lead to a summation of infinities for tion. the self-energy. On the other hand, in self-consistent Born ap- proximation, the bare Green’s functions in Eqs. (280), (281) are − R  R  1 1. Kinetics for small momenta replaced by the full Green’s functions GQ = ω − u|q| − ΣQ . As a consequence, the on-shell Green’s function is regularized by the self-energy ΣR and takes the value For sufficiently small momenta q  1, the kinetic equation simplifies considerably. In this case, the second line of Eq. (284) R  R −1  R−1 Gq,u|q| = − Σq,u|q| = −i σq . (282) can be discarded completely, since the integral is performed over Inserting (282) in the definition for the retarded self-energy and a very small momentum interval 0 < p < q. On the other hand, evaluating the self-energy on-shell, one obtains for the integral in the third line of Eq. (284), all terms (q+ p) ≈ p   for the dominant part of the integral. Therefore Z    qp(q − p) pq(p + q)  R 2 ∂tnp,t    γ  σq,t = v0 σR + 2np,t + 1  R R + R R  q,t σ + σ σ + σ  ∂tnq,t = |q| + Jt , (285) 0

2. Scaling of the self-energy

The fact that the present system is described by a U(1)- invariant, massless field theory is reflected by the absence of a scale in the self-energy equation (283). One important con- q→0 R sequence is that σq,t → 0 generically, i.e., the generation of a mass gap is forbidden by symmetry. A further consequence of Eq. (283) is that, whenever the term in brackets obeys a scaling Figure 13. Time evolution of the phonon density nq,t in a sequence law, 10 (t1, t2, t3) = (2, 3, 4) · 2 in terms of q/Λ and for a heating rate γ = v0Λ ∂ n t p ∼ | |ηn 0.06v0Λ. In the heating regime, for small momenta, the distribution σR + 2np + 1 γn p , (288) p increases linearly in momentum nq,t ∼ |q|, while it decreases as nq,t ∼ −1 R 1/|q| in the interaction dominated thermal regime. The crossover xth ∼ the solution for the self-energy will be a scaling function σq = t−4/5 approaches zero as time evolves. The dashed line represents a ηR γR|p| as well. Inserting this scaling ansatz in Eq. (283) yields Bose-Einstein distribution corresponding to the phonon density at t = 2 t3. The dash-dotted line indicates nB(T(tc)) at the time tc, for which the v γn ηR 0 4+ηn−ηR Luttinger description breaks down, i.e., T(tc) = uΛ. γR|p| = |p| Iηn,ηR (289) γR and identifies was the case for the bare heating (252). The thermal distance η q 4/5 n increases sub-ballistically xth(t) ∼ t in time, with a charac- ηR = 2 + and γR = v0 γnIηn,ηR . (290) 2 teristic heating exponent ηh = 4/5, while at the same time, the effective temperature describing the distribution of the short dis- The dimensionless integral Iη ,η is defined as n R tance modes increases linearly in time T(t) ∼ t. ! Z x(1 − x) x(1 + x) I ηn ηn,ηR = x η η + η η , (291) 0 0 is continu- this regime nq,t = nB(u|q|, T(t)) ≈ u|q| with very good agreement ously increased due to heating. The central result of the analysis and the only indicator of the permanent heating is the continu- R ηR is a scaling solution for the self-energies σq,t ∼ |q| with a new ously increasing temperature T(t) ∼ t. This is contrasted by the non-equilibrium exponent ηR = 5/3, which is observable in the evolution of the phonon density in the low momentum regime, long-wavelength regime, i.e., on distances x > xth(t). Here, xth(t) which is dominated by strong phonon production. In this regime, marks a thermal distance, below which the dynamics is domi- the scattering of high-momentum phonons into low-momentum nated by thermalized short distance modes and above which the modes enhances the effect of the heating and leads, due to the phonon density increases linearly in momentum nq,t ∼ |q|, as it structure of the vertex, to an increase of the linear production rate 61 according to Eq. (285). The pinning of the phonon occupation 3. Observability of the q = 0 mode is an exact result for the underlying U(1)- symmetric, i.e. particle number conserving dynamics. It can The universal scaling of the phonon self-energies as well as be shown that the phonon number fluctuations in the zero mo- the scaling regimes of the phonon distribution function can be mentum mode are proportional to the fluctuations of the global observed in cold atom experiments via Bragg spectroscopy [43], particle number in the system, see Ref. [43], and consequently which gives direct access to the characteristic universal fea- they are integrals of motion for a U(1)-symmetry preserving dy- tures in the heating dynamics. In Bragg experiments, the de- namics. The pinning effect for the low momentum distribution tected Bragg signal is directly proportional to the Fourier trans- at nq=0,t = nq=0,0 leads to a very slow, sub-ballistic thermaliza- form of the two-point density-density correlation function [321– tion of the low momentum regime, since the formation of the R i(qx−ωδ) 324] or dynamical structure factor S q,t,ω = dδdxe h{n(t+ typical, thermal Rayleigh-Jeans divergence nq,t ∼ 1/|q| is only δ/2, x), n(t−δ/2, 0)}i. In the Luttinger framework in the Keldysh achieved by the scaling of the inverse thermal length (x (t))−1 th formalism this translates into S q,t,ω = −hρc(−q, t, −ω)ρc(q, t, ω)i. to zero, instead of a direct filling of these modes. Explicit evaluation of the structure factor yields     (2nq,t+1)|q|K ˜ ω−q S q,t,ω = R f R +(ω→−ω) . (296) 2. Non-equilibrium scaling πσq,t σq,t

Here, f˜(x) = 1/(1 + x2) is a dimensionless scaling function, Away from the crossover scale q x−1, deep in the heating , th which is centered at x = 0 and has unit width. As a consequence, or thermal regimes, the phonon density exhibits scaling behavior the dynamic structure factor is peaked at the mass shell ω = u|q| and can be written as R and has a typical width δω = σq,t, which reveals the scaling of ( −1 the self-energies. c(t)|q| for |q|  xth nq,t = T(t) −1 , (292) The scaling of the distribution function and the time depen- u|q| for |q|  xth dent crossover xth(t) does not necessitate dynamical (frequency where the functions c(t), T(t) have to be determined numerically. resolved) information, and can be obtained from the static struc- In these regimes, according to Sec.VC2, one finds scaling be- ture factor alone. The latter represents the equal time density- havior of the phonon self-energy as well. The corresponding density correlation function. It is determined as the frequency scaling exponent is determined by integrated dynamic structure factor Z |q|K ∂tnq,t η f ≡ + 2n + 1 ∼ |q| n . (293) S q,t = S q,t,ω = π (2nq,t + 1) (297) q,t R q,t ω σq,t and scales quadratically in the momentum in the heating regime In the heating dominated regime, nq,t = c(t)|q| and therefore while approaching a constant in the thermalized regime, thereby 0 0 revealing the crossover between these two regimes. c (t) ηR>1 c (t) 1−ηR 1−ηR fq,t = |q| + c(t)|q| + 1 → |q| (294) γR γR VI. CONCLUSIONS AND OUTLOOK for small momenta, since ηR > 1 is guaranteed by the subleading nature of the interactions. This directly implies ηn = 1−ηR and a 5 super-diffusive exponent ηR = 3 in this regime, which lacks any We have reviewed here recent progress in the theory of driven equilibrium counterpart. open quantum systems, which are at the interface of quantum op- On the other hand, in the large momentum regime fq,t is ob- tics, many-body physics, and statistical mechanics. In particular, viously dominated by the term proportional to nq,t and ηn = −1, we have developed a quantum field theoretical approach based which leads to the known thermal equilibrium exponent ηR = on the Keldysh functional integral for open systems, which un- 3 2 [320] and a thermal scaling behavior for large momenta not derlies these advances. The formalism developed here paves the only in the distribution function but also in the self-energy. The way for many future applications and discoveries. We may struc- crossover between the two scaling regimes can be estimated by ture these into four groups of topics. equating the two relevant terms, which tend to dominate fq,t in Semi-classical regime — A key challenge here is to sharpen the corresponding regimes. This yields the contrast between equilibrium and genuine non-equilibrium physics. One way of succeeding in this respect is to uncover ∂tnq th σR , new links of driven open quantum systems to paradigmatic sit- = qth (295) 2nqth + 1 uations in non-equilibrium (classical) statistical mechanics. We have discussed one such connection between the phase dynamics −1 for qth = xth and can be used to analytically estimate the scaling of driven Bose condensates and surface growth in Sec.IV. Fur- 4/5 of xth(t) in time [43], resulting in xth(t) ∼ t , which agrees well ther instances have been pointed out in the literature recently: with the numerical findings. for example, driven Rydberg gases can be brought into regimes 62 where they connect to the physics of glasses [49] or the univer- it is a fundamental challenge to explore the fate of the quantum sality class of directed percolation [50], and the late stages of Hall effect – or more generally, physical phenomena related to heating of atoms in optical lattices show slow decoherence dy- topology – under general non-equilibrium conditions. Another namics described by non-linear diffusion, reminiscent of glasses angle is provided by theoretical proposals, where drive and/or as well [45, 46, 325]. Connections of such settings to field theo- dissipation do not occur as a small perturbation, but rather as the retical non-linear reaction-diffusion models [107, 326] still await dominant resource of many-body dynamics, guiding the system their exploration. A new kind of equilibrium to non-equilibrium density matrix into topologically nontrivial states, which some- phase transition may be expected in three dimensional driven times even do not have a direct equilibrium counterpart. This bosonic systems on the basis of the phase diagram for surface concerns topologically non-trivial dark states in driven open roughening, and it is intriguing to investigate whether ultracold atomic fermion systems [70, 330] and periodically driven (Flo- atom setups could provide a physical platform to explore such quet) dynamics [331, 332] alike. physics. The density matrices describing such systems typically do Even more ambitiously, such driven open quantum systems not correspond to pure, but rather to mixed states. While such hold the potential for truly new paradigms in non-equilibrium density matrices can still host non-trivial topological proper- statistical mechanics. One example is presented by the fun- ties [216, 333, 334], the extent to which this translates into damental open question on the driven open analogue of the physically observable consequences in the correlations and re- Kosterlitz-Thouless scenario in two dimensions, directly rele- sponses to (artificial) external gauge fields is at the moment vant for experiments with exciton-polariton systems. Techni- a wide open issue. This calls for the development of non- cally, this requires to analyze a KPZ equation with a compact equilibrium topological field theories in the framework of the variable, allowing for the presence of vortex defects. Keldysh Keldysh functional integral: the established equilibrium coun- field theory offers the flexibility to address such questions. terpart has proven to be able to efficiently describe both bulk Quantum regime — The quantum dynamical field theory correlations and responses to external gauge potentials, as well framework developed here also allows us to address prob- as to provide a proper notion of the bulk-boundary correspon- lems where the limit of classical dynamical field theories (see dence (cf. e.g. [69]), giving access to the edge physics in both Sec.IIC) is not applicable. Here one challenge is to identify non-interacting and interacting systems. traces of non-equilibrium quantum effects at macroscopic dis- Dynamics — Addressing the time evolution of open sys- tances. An instance of such a phenomenon has been established tems adds a new twist to the question of thermalization [32, recently in terms of a driven analogue of quantum critical behav- 37, 335, 336] or, more generally, equilibration of quantum sys- ior in a system with a dark state, a state which is decoupled from tems. This is also a necessary step to achieve a realistic de- noise [28]. It remains to be seen whether a full classification of scription of broad classes of experiments, in particular, with ul- driven Markovian (quantum) criticality can be achieved, com- tracold atoms, as well as certain solid states systems in pump- plementing the seminal analysis of equilibrium classical criti- probe setups [337, 338]. While first instances of universal be- cality by Hohenberg and Halperin [1]. Beyond bosonic systems, havior have been identified in low dimension in the short [43] this also includes fermionic systems, which have been shown to and long [45, 47, 51] dynamics, it is certainly fair to say that exhibit critical scaling [61, 63, 64], but so far were analyzed at a general principles so far remain elusive. Gaussian fixed point only. Another challenge in this direction is to identify effects, which unambiguously reveal the microscopic quantum mechanical ori- VII. ACKNOWLEDGEMENTS gin of the underlying dynamics. Here an example was provided recently in the context of driven Rydberg systems, where a short The authors thank E. Altman, L. Chen, A. Chiocchetta, E. distance constraint in the coherent Hamilton dynamics gives rise G. Dalla Torre, A. Gambassi, L. He, S. D. Huber, S. E. Huber, to an additional relevant direction in parameter space, leading to M. Lukin, J. Marino, S. Sachdev, P. Strack, U. Tauber,¨ and J. a new kind of absorbing state phase transition without immedi- Toner for collaboration on projects discussed in this review. We ate counterpart in models of classical origin [327]. are also grateful for inspiring and useful discussions with many Certainly, progress in this respect will necessitate a more com- people at the KITP workshop “Many-Body Physics with Light.” prehensive understanding of the structure of quantum dynamical In particular, we thank A. Altland, B. Altshuler, I. Carusotto, field theories. One relevant issue is to reveal universal aspects A. Daley, A. Gorshkov, M. Hafezi, M. Hartmann, R. Fazio, M. of the low frequency dynamics tied to the presence of conser- Fleischhauer, A. Imamoglu, J. Koch, I. Lerner, I. Lesanovsky, P. vation laws. A concrete goal is the systematic construction of Rabl, A. Rosch, G. Shlyapnikov, M. Szymanska, J. Taylor and dynamical slow modes on the basis of the symmetries (and their H. Tureci¨ for critical remarks and feedback on the manuscript. breaking patterns) of the Keldysh functional integral. We are indebted to J. Koch for proofreading the manuscript. Topology in open quantum systems – Recently, experiments We acknowledge support by the Austrian Science Fund (FWF) with photonic lattices have started to address quantum Hall through the START grant Y 581-N16 and the German Research physics in driven open quantum systems [328, 329]. Although Foundation (DFG) via ZUK 64 and through the Institutional these systems can be idealized to some extent as closed systems, Strategy of the University of Cologne within the German Ex- 63 cellence Initiative (ZUK 81). S. D. also acknowledges support Appendix B: Gaussian functional integration by the European Research Council (ERC) under the European Unions Horizon 2020 research and innovation programme (grant Here we summarize a number of useful formulas for Gaus- agreement No 647434) and by the Kavli Institute for Theoreti- sian functional integration, which can be found in any textbook cal Physics, via the National Science Foundation under Grant on field theory (see, e.g., Refs. [178, 180, 339]). The basic for- No. NSF PHY11-25915, and L. M. S. support from the Euro- d mula for real fields χ(x) = (χ1(x), . . . , χn(x)), x ∈ R (in the ap- pean Research Council through Synergy Grant UQUAM. plications discussed in the main text, the components of χ(x) are fields on the closed time path or — after performing the Keldysh rotation Eq. (32) — classical and quantum fields, and x collects Appendix A: Functional differentiation temporal and spatial coordinates, x = (t, x)), is given by

Z R R i χ(x)T D(x,x0)χ(x0)+i j(x)T χ(x) In this appendix, we give a brief account of functional or vari- D[χ] e 2 x,x0 x ational differentiation, following the presentation in Ref. [233]. R −1/2 − i j(x)T D−1(x,x0) j(x0) The most transparent way to introduce the basic relations of = (det D) e 2 x,x0 , (B1) functional differentiation is by drawing an analogy to the famil- R R d iar formulas of partial differentiation. Indeed, we can consider a where x = d x, j(x) = ( j1(x),..., jn(x)), and the inverse of d 0 field φ(x) with x ∈ R to be the continuum limit of a function φi the integral kernel D(x, x ) is defined by means of the relation defined on lattice points i ∈ Zd. With this identification, relations Z −1 0 0 1 involving partial derivatives can be translated into corresponding D(x, ξ)D (ξ, x ) = δ(x − x ) . (B2) ξ ones for functional derivatives simply by replacing discrete in- P R R d The above formula is valid for invertible symmetric kernels, dices i by continuous ones x, sums i by integrals = d x, x D(x, x0) = D(x0, x)T , with positive semi-definite imaginary part. and partial derivatives ∂/∂φi by functional derivatives δ/δφ(x). If D(x, x0) is not symmetric, on the LHS of Eq. (B1) it can be Following this prescription, the basic formula ∂φi/∂φ j = δi j leads to replaced by the symmetrized kernel 1 h i δφ(x) D˜ (x, x0) = D(x, x0) + D(x0, x)T , (B3) = δ(x − x0). (A1) 2 δφ(x0) leaving the value of the exponent invariant. Then, Eq. (B1) can When we are working with complex fields, φ and φ∗ are usually again be applied. treated as independent variables. Then, the generalization of In case that the integral kernel is translaitionally invariant, i.e., it satisfies D(x, x0) = D(x − x0), it is advantageous to work in   ∂ X  X Fourier space. Then, Eq. (B1) can be written as  ∗  ∗  φi Ai jφ j = Ak jφ j (A2) Z   i R R ∂φk χ(−q)T D(q)χ(q)+i j(−q)T χ(q) i, j j D[χ] e 2 q q R reads − / − i j(−q)T D−1(q) j(q) = (det D) 1 2 e 2 q , (B4) Z Z δ R R φ∗(y)A(y, y0)φ(y0) = A(x, y0)φ(y0). (A3) ≡ dq ∗ where we are using the shorthand notation π d . The δφ (x) 0 0 q (2 ) y,y y Fourier transformation turns the convolution in Eq. (B3) into a 0 multiplication, showing that D−1(q) can be obtained by inversion The expression A(x, x ) is the continuum limit of a matrix Ai j. In particular, it can be a differential operator. As an example, of the matrix D(q), we calculate the second functional derivative for the case that D(q)D−1(q) = 1. (B5) A(x, x0) = ∇2δ(x− x0). By straightforward differentiation we find As above, in Eq. (B4) we are assuming D(q) = D(−q)T . If this is δ2 Z δ Z not the case, D(q) has to be replaced by the symmetrized version φ∗(y)∇2φ(y) = φ∗(y)∇2δ(y − x0) δφ∗(x)δφ(x0) δφ∗(x) 1 h i y y D˜ (q) = D(q) + D(−q)T , (B6) δ 2 = ∇2φ(x0) δφ∗(x) before the formula can be applied. For the case of complex fields ψ(x) = (ψ (x), . . . , ψ (x)) (and = ∇2δ(x − x0). 1 n with corresponding definitions of φ(x) and χ(x)), Eq. (B1) is re- (A4) placed by Z Finally, we note that the chain rule applies also in the case of R † 0 0 R † † ∗ i 0 ψ (x)D(x,x )ψ(x )+i (φ (x)ψ(x)+ψ (x)χ(x)) functional differentiation. This last ingredient is required to per- D[ψ, ψ ] e x,x x form the second variational derivatives in Eq. (34) in order to R † −1 0 −1 −i 0 φ (x)D (x,x )χ(x) obtain the Green’s functions from the generating functional. = (det D) e x,x , (B7) 64   where we are assuming that −i D − D† is positive semi- definite, but D(x, x0) does not have to be symmetric. The cor- responding formula for the case of a translationally invariant in- tegral kernel reads (note the different signs of q in comparison to Eq. (B4))

Z R R ∗ i ψ†(q)D(q)ψ(q)+i (φ†(q)ψ(q)+ψ†(q)χ(q)) D[ψ, ψ ] e q q R − −i φ†(q)D−1(q)χ(q) = (det D) 1 e q . (B8)

[1] P. Hohenberg and B. Halperin, “Theory of dynamic critical phe- “Low-disorder microwave cavity lattices for quantum simulation nomena,” Rev. Mod. Phys. 49, 435–479 (1977). with photons,” Phys. Rev. A 86, 023837 (2012). [2] Timothy Halpin-Healy and Yi-Cheng Zhang, “Kinetic roughen- [16] Joseph W Britton, Brian C Sawyer, Adam C Keith, C-C Joseph ing phenomena, stochastic growth, directed and all Wang, James K Freericks, Hermann Uys, Michael J Biercuk, that. Aspects of multidisciplinary statistical mechanics,” Phys. and John J Bollinger, “Engineered two-dimensional Ising inter- Rep. 254, 215–414 (1995). actions in a trapped-ion quantum simulator with hundreds of [3] H J Jensen, Self Organized Criticality (Cambridge University spins,” Nature 484, 489–92 (2012). Press, Cambridge, 1998). [17] R. Blatt and C. F. Roos, “Quantum simulations with trapped [4] Haye Hinrichsen, “Non-equilibrium critical phenomena and ions,” Nat. Phys. 8, 277–284 (2012). phase transitions into absorbing states,” Adv. Phys. 49, 815–958 [18] Wolfgang Lechner and Peter Zoller, “From classical to quantum (2000). glasses with ultracold polar molecules,” Phys. Rev. Lett. 111, [5] Uwe C. Tauber,¨ Critical Dynamics: A Field Theory Approach to 185306 (2013). Equilibrium and Non-Equilibrium Scaling Behavior (Cambridge [19] Ehud Altman, Lukas M. Sieberer, Leiming Chen, Sebastian University Press, Cambridge, 2014). Diehl, and John Toner, “Two-Dimensional Superfluidity of Ex- [6] Immanuel Bloch, Jean Dalibard, and Wilhelm Zwerger, “Many- citon Polaritons Requires Strong Anisotropy,” Phys. Rev. X 5, body physics with ultracold gases,” Rev. Mod. Phys. 80, 885–964 011017 (2015). (2008). [20] Francesco Piazza and Philipp Strack, “Umklapp superradiance [7] Maciej Lewenstein, Anna Sanpera, Veronica Ahufinger, Bogdan with a collisionless quantum degenerate ,” Phys. Rev. Damski, Aditi Sen(De), and Ujjwal Sen, “Ultracold atomic gases Lett. 112, 143003 (2014). in optical lattices: mimicking condensed matter physics and be- [21] J. Keeling, M. J. Bhaseen, and B. D. Simons, “Fermionic su- yond,” Advances in Physics 56, 243–379 (2007). perradiance in a transversely pumped optical cavity,” Phys. Rev. [8] J. Ignacio Cirac and Peter Zoller, “Goals and opportunities in Lett. 112, 143002 (2014). quantum simulation,” Nat Phys 8, 264–266 (2012). [22] Corinna Kollath, Ameneh Sheikhan, Stefan Wolff, and Ferdi- [9] Kristian Baumann, Christine Guerlin, Ferdinand Brennecke, and nand Brennecke, “Ultracold Fermions in a Cavity-Induced Arti- Tilman Esslinger, “Dicke quantum phase transition with a super- ficial .” Phys. Rev. Lett. 116, 060401 (2016). fluid gas in an optical cavity,” Nature 464, 1301–6 (2010). [23] J. Raftery, D. Sadri, S. Schmidt, H. E. Tureci,¨ and A. A. Houck, [10] Helmut Ritsch, Peter Domokos, Ferdinand Brennecke, and “Observation of a dissipation-induced classical to quantum tran- Tilman Esslinger, “Cold atoms in cavity-generated dynamical sition,” Phys. Rev. X 4, 031043 (2014). optical potentials,” Rev. Mod. Phys. 85, 553–601 (2013). [24] M. J. Hartmann, F. G. S. L. Brandao,˜ and M. B. Plenio, “Quan- [11] Peter Schausz, Marc Cheneau, Manuel Endres, Takeshi tum many-body phenomena in coupled cavity arrays,” Laser Pho- Fukuhara, Sebastian Hild, Ahmed Omran, Thomas Pohl, Chris- tonics Rev. 2, 527–556 (2008). tian Gross, Stefan Kuhr, and Immanuel Bloch, “Observation of [25] Emanuele G. Dalla Torre, Eugene Demler, Thierry Giamarchi, spatially ordered structures in a two-dimensional Rydberg gas,” and Ehud Altman, “Quantum critical states and phase transitions Nature 491, 87–91 (2012). in the presence of non-equilibrium noise,” Nat. Phys. 6, 806–810 [12] J Kasprzak, M Richard, S Kundermann, A Baas, P Jeambrun, (2010). J M J Keeling, F M Marchetti, M H Szymanska,´ R Andre,´ J L [26] L. M. Sieberer, S. D. Huber, E. Altman, and S. Diehl, “Dynam- Staehli, V Savona, P B Littlewood, B Deveaud, and Le Si Dang, ical Critical Phenomena in Driven-Dissipative Systems,” Phys. “Bose-Einstein condensation of exciton polaritons,” Nature 443, Rev. Lett. 110, 195301 (2013). 409–14 (2006). [27] Merlijn van Horssen and Juan P Garrahan, “Open quantum [13] Iacopo Carusotto and Cristiano Ciuti, “Quantum fluids of light,” reaction-diffusion dynamics: Absorbing states and relaxation.” Rev. Mod. Phys. 85, 299–366 (2013). Phys. Rev. E. Stat. Nonlin. Phys. 91, 032132 (2015). [14] Andrew A. Houck, Hakan E. Tureci,¨ and Jens Koch, “On-chip [28] Jamir Marino and Sebastian Diehl, “Driven Markovian Quantum quantum simulation with superconducting circuits,” Nat. Phys. 8, Criticality,” Phys. Rev. Lett. 116, 070407 (2016). 292–299 (2012). [29] J.-T. Hsiang and B.L. Hu, “Nonequilibrium steady state in open [15] D. L. Underwood, W. E. Shanks, J. Koch, and A. A. Houck, quantum systems: Influence action, stochastic equation and 65

power balance,” Ann. Phys. (N. Y). 362, 139–169 (2015). 111, 150403 (2013). [30] Ana Maria Rey, B. L. Hu, Esteban Calzetta, Albert Roura, and [47] J. Schachenmayer, L. Pollet, M. Troyer, and A. J. Daley, “Spon- Charles W. Clark, “Nonequilibrium dynamics of optical-lattice- taneous emission and thermalization of cold bosons in optical loaded Bose-Einstein-condensate atoms: Beyond the Hartree- lattices,” Phys. Rev. A 89, 011601 (2014). Fock-Bogoliubov approximation,” Phys. Rev. A 69, 033610 [48] Andrew J. Daley, “Quantum trajectories and open many-body (2004). quantum systems,” Adv. Phys. 63, 77–149 (2014). [31] J. Berges, Sz. Borsanyi,´ and C. Wetterich, “Prethermalization,” [49] Beatriz Olmos, Igor Lesanovsky, and Juan P. Garrahan, “Facil- Phys. Rev. Lett. 93, 142002 (2004). itated spin models of dissipative quantum glasses,” Phys. Rev. [32] Pasquale Calabrese and John Cardy, “Time dependence of corre- Lett. 109, 020403 (2012). lation functions following a quantum quench,” Phys. Rev. Lett. [50] Matteo Marcuzzi, Emanuele Levi, Weibin Li, Juan P. Garra- 96, 136801 (2006). han, Beatriz Olmos, and Igor Lesanovsky, “Emergence of di- [33] Aditi Mitra and Thierry Giamarchi, “Mode-Coupling-Induced rected percolation in the dynamics of dissipative Rydberg gases,” Dissipative and Thermal Effects at Long Times after a Quantum arXiv:1411.7984 (2014). Quench,” Phys. Rev. Lett. 107, 150602 (2011). [51] Johannes Lang and Francesco Piazza, “Critical Relaxation with [34] Selma Koghee and Michiel Wouters, “Dynamical Casimir emis- Overdamped Quasi-Particles in Driven-Dissipative Systems,” sion from polariton condensates,” Phys. Rev. Lett. 112, 036406 (2016), arXiv:1602.05102. (2014). [52] J. F. Poyatos, J. I. Cirac, and P. Zoller, “Quantum Reservoir En- [35] I. Carusotto, R. Balbinot, A. Fabbri, and A. Recati, “Density cor- gineering with Laser Cooled Trapped Ions,” Phys. Rev. Lett. 77, relations and analog dynamical Casimir emission of Bogoliubov 4728–4731 (1996). phonons in modulated atomic Bose-Einstein condensates,” Eur. [53] S. Bose, P. L. Knight, M. B. Plenio, and V. Vedral, “Proposal for Phys. J. D 56, 391–404 (2010). Teleportation of an Atomic State via Cavity Decay,” Phys. Rev. [36] Pierre-Elie´ Larre´ and Iacopo Carusotto, “Propagation of a quan- Lett. 83, 5158–5161 (1999). tum fluid of light in a cavityless nonlinear optical medium: Gen- [54] S. Diehl, A. Micheli, A. Kantian, B. Kraus, H. P. Buchler,¨ and eral theory and response to quantum quenches,” Phys. Rev. A 92, P. Zoller, “Quantum states and phases in driven open quantum 043802 (2015). systems with cold atoms,” Nat. Phys. 4, 878–883 (2008). [37] M. Gring, M. Kuhnert, T. Langen, T. Kitagawa, B. Rauer, [55] Frank Verstraete, Michael M. Wolf, and J. Ignacio Cirac, “Quan- M. Schreitl, I. Mazets, D. Adu Smith, E. Demler, and tum computation and quantum-state engineering driven by dissi- J. Schmiedmayer, “Relaxation and prethermalization in an iso- pation,” Nat. Phys. 5, 633–636 (2009). lated quantum system,” Science 337, 1318–22 (2012). [56] Hendrik Weimer, Markus Muller,¨ Igor Lesanovsky, Peter Zoller, [38] S. Trotzky, Y-A. Chen, A. Flesch, I. P. McCulloch, and Hans Peter Buchler,¨ “A Rydberg quantum simulator,” Nat. U. Schollwock,¨ J. Eisert, and I. Bloch, “Probing the relax- Phys. 6, 382–388 (2010). ation towards equilibrium in an isolated strongly correlated one- [57] Markus Muller,¨ Sebastian Diehl, Guido Pupillo, and Peter dimensional bose gas,” Nat Phys 8, 325–330 (2012). Zoller, “Engineered open systems and quantum simulations with [39] F. Meinert, M. J. Mark, E. Kirilov, K. Lauber, P. Weinmann, A. J. atoms and ions,” in Advances in Atomic, Molecular, and Optical Daley, and H.-C. Nagerl,¨ “Quantum quench in an atomic one- Physics, Vol. 61, edited by Ennio Arimondo Paul Berman and dimensional ising chain,” Phys. Rev. Lett. 111, 053003 (2013). Chun Lin (Academic Press, 2012) pp. 1 – 80. [40] Chen-Lung Hung, Victor Gurarie, and Cheng Chin, “From Cos- [58] Hanna Krauter, Christine A. Muschik, Kasper Jensen, Wojciech mology to Cold Atoms: Observation of Sakharov Oscillations Wasilewski, Jonas M. Petersen, J. Ignacio Cirac, and Eugene S. in a Quenched Atomic Superfluid,” Science 341, 1213–1215 Polzik, “Entanglement Generated by Dissipation and Steady (2013). State Entanglement of Two Macroscopic Objects,” Phys. Rev. [41] J.-C. Jaskula, G. B. Partridge, M. Bonneau, R. Lopes, J. Ruaudel, Lett. 107, 080503 (2011). D. Boiron, and C. I. Westbrook, “Acoustic analog to the dynam- [59] Julio T. Barreiro, Markus Muller,¨ Philipp Schindler, Daniel Nigg, ical Casimir effect in a Bose-Einstein condensate.” Phys. Rev. Thomas Monz, Michael Chwalla, Markus Hennrich, Christian F. Lett. 109, 220401 (2012). Roos, Peter Zoller, and Rainer Blatt, “An open-system quantum [42] G. Nardin, K. G. Lagoudakis, M. Wouters, M. Richard, A. Baas, simulator with trapped ions,” Nature 470, 486–91 (2011). R. Andre,´ Le Si Dang, B. Pietka, and B. Deveaud-Pledran,´ “Dy- [60] P. Schindler, M. Muller, D. Nigg, J. T. Barreiro, E. A. Martinez, namics of long-range ordering in an exciton-polariton conden- M. Hennrich, T. Monz, S. Diehl, P. Zoller, and R. Blatt, “Quan- sate,” Phys. Rev. Lett. 103, 256402 (2009). tum simulation of dynamical maps with trapped ions,” Nat. Phys. [43] Michael Buchhold and Sebastian Diehl, “Nonequilibrium univer- 9, 361–367 (2013). sality in the heating dynamics of interacting Luttinger liquids,” [61] J. Eisert and T. Prosen, “Noise-driven quantum criticality,” Phys. Rev. A 92, 013603 (2015). arXiv:1012.5013 (2010). [44] Michael Buchhold and Sebastian Diehl, “Kinetic theory for inter- [62] M. J. Kastoryano, F. Reiter, and A. S. Sørensen, “Dissipative acting Luttinger liquids,” The European Physical Journal D 69, Preparation of Entanglement in Optical Cavities,” Phys. Rev. 224 (2015). Lett. 106, 090502 (2011). [45] Dario Poletti, Peter Barmettler, Antoine Georges, and Corinna [63] Birger Horstmann, J. Ignacio Cirac, and Geza´ Giedke, “Noise- Kollath, “Emergence of glasslike dynamics for dissipative and driven dynamics and phase transitions in fermionic systems,” strongly interacting bosons,” Phys. Rev. Lett. 111, 195301 Phys. Rev. A 87, 012108 (2013). (2013). [64]M.H oning,¨ M. Moos, and M. Fleischhauer, “Critical exponents [46] Zi Cai and Thomas Barthel, “Algebraic versus exponential deco- of steady-state phase transitions in fermionic lattice models,” herence in dissipative many-particle systems,” Phys. Rev. Lett. Phys. Rev. A 86, 013606 (2012). 66

[65] Eliot Kapit, Mohammad Hafezi, and Steven H. Simon, “Induced Cambridge Lecture Notes in Physics (Cambridge University self-stabilization in fractional quantum hall states of light,” Phys. Press, Cambridge, 1996). Rev. X 4, 031039 (2014). [86] Jean Zinn-Justin, and Critical Phenom- [66] Nicolai Lang and Hans Peter Buchler,¨ “Exploring quantum ena, 4th ed., International Series of Monographs on Physics No. phases by driven dissipation,” Phys. Rev. A 92, 012128 (2015). 113 (Oxford University Press, Oxford, 2002). [67] Alexei Kitaev, “Anyons in an exactly solved model and beyond,” [87] A. J. Daley, J. M. Taylor, S. Diehl, M. Baranov, and P. Zoller, Ann. Phys. (N. Y). 321, 2–111 (2006). “Atomic three-body loss as a dynamical three-body interaction,” [68] M. Z. Hasan and C. L. Kane, “Colloquium: Topological insula- Phys. Rev. Lett. 102, 040402 (2009). tors,” Rev. Mod. Phys. 82, 3045–3067 (2010). [88] Jean-Sebastien´ Bernier, Peter Barmettler, Dario Poletti, and [69] Xiao-Liang Qi and Shou-Cheng Zhang, “Topological insulators Corinna Kollath, “Emergence of spatially extended pair coher- and superconductors,” Rev. Mod. Phys. 83, 1057–1110 (2011). ence through incoherent local environmental coupling,” Phys. [70] Sebastian Diehl, Enrique Rico, Mikhail A. Baranov, and Peter Rev. A 87, 063608 (2013). Zoller, “Topology by dissipation in atomic quantum wires,” Nat. [89] Lars Bonnes, Daniel Charrier, and Andreas M. Lauchli,¨ “Dy- Phys. 7, 971–977 (2011). namical and steady-state properties of a Bose-Hubbard chain [71] C.-E. Bardyn, M. A. Baranov, E. Rico, A. Imamoglu, P. Zoller, with bond dissipation: A study based on matrix product oper- and S. Diehl, “Majorana Modes in Driven-Dissipative Atomic ators,” Phys. Rev. A 90, 033612 (2014). Superfluids with a Zero Chern Number,” Phys. Rev. Lett. 109, [90] Lars Bonnes and Andreas M. Lauchli,¨ “Superoperators vs. Tra- 130402 (2012). jectories for Matrix Product State Simulations of Open Quantum [72] Jan Carl Budich, Peter Zoller, and Sebastian Diehl, “Dissipa- System: A Case Study,” arXiv:1411.4831 (2014). tive preparation of Chern insulators,” Phys. Rev. A 91, 042117 [91] , “Brownian Motion of a Quantum Oscillator,” (2015). J. Math. Phys. 2, 407 (1961). [73] Mehran Kardar, Giorgio Parisi, and Yi-Cheng Zhang, “Dynamic [92] Julian Schwinger, “Field theory of unstable particles,” Ann. Scaling of Growing Interfaces,” Phys. Rev. Lett. 56, 889–892 Phys. (N. Y). 9, 169–193 (1960). (1986). [93] L. V. Keldysh, “Diagram Technique for Nonequilibrium Pro- [74] Dario Poletti, Jean-Sebastien´ Bernier, Antoine Georges, and cesses,” Sov. Phys. JETP 20, 1018—-1026 (1965). Corinna Kollath, “Interaction-induced impeding of decoherence [94] Kalyana Mahanthappa, “Multiple Production of Photons in and anomalous diffusion,” Phys. Rev. Lett. 109, 045302 (2012). Quantum Electrodynamics,” Phys. Rev. 126, 329–340 (1962). [75] Igor Lesanovsky and Juan P. Garrahan, “Kinetic constraints, hi- [95] Pradip M. Bakshi and Kalyana T. Mahanthappa, “Expectation erarchical relaxation, and onset of glassiness in strongly interact- Value Formalism in Quantum Field Theory. I,” J. Math. Phys. 4, ing and dissipative Rydberg gases,” Phys. Rev. Lett. 111, 215305 1 (1963). (2013). [96] Pradip M. Bakshi and Kalyana T. Mahanthappa, “Expectation [76] Igor Lesanovsky and Juan P. Garrahan, “Out-of-equilibrium Value Formalism in Quantum Field Theory. II,” J. Math. Phys. 4, structures in strongly interacting Rydberg gases with dissipa- 12 (1963). tion,” Phys. Rev. A 90, 011603 (2014). [97] Andy C Y Li, F Petruccione, and Jens Koch, “Perturbative ap- [77] Beatriz Olmos, Igor Lesanovsky, and Juan P. Garrahan, “Out- proach to Markovian open quantum systems,” Sci. Rep. 4, 4887 of-equilibrium evolution of kinetically constrained many-body (2014). quantum systems under purely dissipative dynamics,” Phys. Rev. [98] Andy C. Y. Li, F. Petruccione, and Jens Koch, “Resummation for E 90, 042147 (2014). Nonequilibrium Perturbation Theory and Application to Open [78] S. Diehl, W. Yi, A. J. Daley, and P. Zoller, “Dissipation-induced Quantum Lattices,” arXiv:1511.01551 (2015). d-wave pairing of fermionic atoms in an optical lattice,” Phys. [99] Peter Degenfeld-Schonburg, Carlos Navarrete-Benlloch, and Rev. Lett. 105, 227001 (2010). Michael J. Hartmann, “Self-consistent projection operator the- [79] Jean Dalibard, Yvan Castin, and Klaus Mølmer, “Wave-function ory in nonlinear quantum optical systems: A case study on de- approach to dissipative processes in quantum optics,” Phys. Rev. generate optical parametric oscillators,” Phys. Rev. A 91, 053850 Lett. 68, 580–583 (1992). (2015). [80] M. B. Plenio and P. L. Knight, “The quantum-jump approach to [100] Hendrik Weimer, “Variational principle for steady states of dis- dissipative dynamics in quantum optics,” Rev. Mod. Phys. 70, sipative quantum many-body systems,” Phys. Rev. Lett. 114, 101–144 (1998). 040402 (2015). [81] Crispin W Gardiner and Peter Zoller, Quantum Noise, 2nd ed., [101] S. Finazzi, A. Le Boite,´ F. Storme, A. Baksic, and C. Ciuti, Springer series in synergetics, Vol. 56 (Springer, Berlin Heidel- “Corner-space renormalization method for driven-dissipative berg, 2000). two-dimensional correlated systems,” Phys. Rev. Lett. 115, [82] Heinz-Peter Breuer and Francesco Petruccione, The Theory 080604 (2015). of Open Quantum Systems (Oxford University Press, Oxford, [102] Alex Kamenev, Field Theory of Non-Equilibrium Systems (Cam- 2002). bridge University Press, Cambridge, 2011). [83] Iacopo Carusotto and Cristiano Ciuti, “Spontaneous microcavity- [103] Esteban A. Calzetta and Bei-Lok B. Hu, Nonequilibrium Quan- polariton coherence across the parametric threshold: Quantum tum Field Theory (Cambridge University Press, Cambridge, Monte Carlo studies,” Phys. Rev. B 72, 125335 (2005). 2008). [84] Nigel Goldenfeld, Lectures on Phase Transitions and the Renor- [104] C. Jarzynski, “Nonequilibrium Equality for Free Energy Differ- malization Group, Frontiers in Physics (Perseus Books, Reading, ences,” Phys. Rev. Lett. 78, 2690–2693 (1997). 1992). [105] C. Jarzynski, “Equilibrium free-energy differences from [85] John Cardy, Scaling and Renormalization in Statistical Physics, nonequilibrium measurements: A master-equation approach,” 67

Phys. Rev. E 56, 5018–5035 (1997). Landig, Tobias Donner, and Tilman Esslinger, “Real-time obser- [106] Gavin E. Crooks, “Entropy production fluctuation theorem and vation of fluctuations at the driven-dissipative Dicke phase tran- the nonequilibrium work relation for free energy differences,” sition.” Proc. Natl. Acad. Sci. U. S. A. 110, 11763–7 (2013). Phys. Rev. E 60, 2721–2726 (1999). [126] Klaus Hepp and Elliott H Lieb, “On the superradiant phase tran- [107] John Cardy, “Renormalisation group approach to reaction- sition for molecules in a quantized radiation field: the Dicke diffusion problems,” arXiv:9607163 (1996). maser model,” Annals of Physics 76, 360 – 404 (1973). [108] M Doi, “Second quantization representation for classical many- [127] Y. K. Wang and F. T. Hioe, “Phase Transition in the Dicke Model particle system,” J. Phys. A. Math. Gen. 9, 1465–1477 (1976). of Superradiance,” Phys. Rev. A 7, 831–836 (1973). [109] L. Peliti, “Path integral approach to birth-death processes on a [128] Clive Emary and Tobias Brandes, “Chaos and the quantum phase lattice,” J. Phys. 46, 1469–1483 (1985). transition in the Dicke model,” Phys. Rev. E 67, 066203 (2003). [110] Sebastian Diehl, Andrea Tomadin, Andrea Micheli, Rosario [129] F. Dimer, B. Estienne, A. S. Parkins, and H. J. Carmichael, “Pro- Fazio, and Peter Zoller, “Dynamical Phase Transitions and Insta- posed realization of the Dicke-model quantum phase transition in bilities in Open Atomic Many-Body Systems,” Phys. Rev. Lett. an optical cavity QED system,” Phys. Rev. A 75, 013804 (2007). 105, 015702 (2010). [130] Kater W. Murch, Kevin L. Moore, Subhadeep Gupta, and [111] Subir Sachdev, Quantum Phase Transitions, 2nd ed. (Cambridge Dan M. Stamper-Kurn, “Observation of quantum-measurement University Press, Cambridge, 2011). backaction with an ultracold atomic gas,” Nat. Phys. 4, 561–564 [112] Emanuele G. Dalla Torre, Sebastian Diehl, Mikhail Lukin, Subir (2008). Sachdev, and Philipp Strack, “Keldysh approach for nonequi- [131] Philipp Strack and Subir Sachdev, “Dicke quantum spin glass of librium phase transitions in quantum optics: Beyond the Dicke atoms and photons,” Phys. Rev. Lett. 107, 277202 (2011). model in optical cavities,” Phys. Rev. A 87, 023831 (2013). [132] Dimitris G. Angelakis, Marcelo Franca Santos, and Sougato [113] Michael Buchhold, Philipp Strack, Subir Sachdev, and Sebastian Bose, “Photon-blockade-induced Mott transitions and XY spin Diehl, “Dicke-model quantum spin and photon glass in optical models in coupled cavity arrays,” Phys. Rev. A 76, 031805 cavities: Nonequilibrium theory and experimental signatures,” (2007). Phys. Rev. A 87, 063622 (2013). [133] Michael J. Hartmann, Fernando G. S. L. Brandao, and Martin B. [114] Mohammad F. Maghrebi and Alexey V. Gorshkov, “Nonequi- Plenio, “Strongly interacting polaritons in coupled arrays of cav- librium many-body steady states via Keldysh formalism,” Phys. ities,” Nat. Phys. 2, 849–855 (2006). Rev. B 93, 014307 (2016). [134] Sebastian Schmidt and Jens Koch, “Circuit QED lattices: To- [115] Francesco Piazza and Philipp Strack, “Quantum kinetics of ultra- wards quantum simulation with superconducting circuits,” Ann. cold fermions coupled to an optical resonator,” Phys. Rev. A 90, Phys. 525, 395–412 (2013). 043823 (2014). [135] M Schiro´ Tureci,¨ M Bordyuh, B Oztop,¨ and H E, “Quantum [116] Jean-Sebastien´ Bernier, Dario Poletti, and Corinna Kollath, phase transition of light in the RabiHubbard model,” J. Phys. B “Dissipative quantum dynamics of fermions in optical lattices: At. Mol. Opt. Phys. 46, 224021 (2013). A slave-spin approach,” Phys. Rev. B 90, 205125 (2014). [136] M Schiro,´ M Bordyuh, B Oztop, and H E Tureci,¨ “Phase transi- [117] H. Mabuchi and A. C. Doherty, “Cavity quantum electrodynam- tion of light in cavity QED lattices.” Phys. Rev. Lett. 109, 053601 ics: Coherence in context,” Science 298, 1372–1377 (2002). (2012). [118] C. J. Hood, T. W. Lynn, A. C. Doherty, A. S. Parkins, and H. J. [137] M. Schiro,´ C. Joshi, M. Bordyuh, R. Fazio, J. Keeling, and H. E. Kimble, “The atom-cavity microscope: Single atoms bound in Tureci,¨ “Exotic attractors of the non-equilibrium Rabi-Hubbard orbit by single photons,” Science 287, 1447–1453 (2000). model,” (2015), arXiv:1503.04456. [119] D. Nagy, G. Konya,´ G. Szirmai, and P. Domokos, “Dicke-Model [138] L. J. Zou, D. Marcos, S. Diehl, S. Putz, J. Schmiedmayer, J. Ma- Phase Transition in the Quantum Motion of a Bose-Einstein jer, and P. Rabl, “Implementation of the Dicke Lattice Model in Condensate in an Optical Cavity,” Phys. Rev. Lett. 104, 130401 Hybrid Quantum System Arrays,” Phys. Rev. Lett. 113, 023603 (2010). (2014). [120] K. Baumann, R. Mottl, F. Brennecke, and T. Esslinger, “Explor- [139] J. Ruiz-Rivas, E. del Valle, C. Gies, P. Gartner, and M. J. Hart- ing Symmetry Breaking at the Dicke Quantum Phase Transition,” mann, “Spontaneous collective coherence in driven dissipative Phys. Rev. Lett. 107, 140402 (2011). cavity arrays,” Phys. Rev. A 90, 033808 (2014). [121] Christoph Maschler and Helmut Ritsch, “Cold atom dynamics in [140] D. Marcos, A. Tomadin, S. Diehl, and P. Rabl, “Photon conden- a quantum optical lattice potential,” Phys. Rev. Lett. 95, 260401 sation in circuit quantum electrodynamics by engineered dissipa- (2005). tion,” New J. Phys. 14, 055005 (2012). [122]B. Oztop,¨ Mykola Bordyuh, Ozg¨ ur¨ E. Mustecaplolu,¨ and [141] A. Tomadin and Rosario Fazio, “Many-body phenomena in Hakan E. Tureci,¨ “Excitations of optically driven atomic conden- QED-cavity arrays,” J. Opt. Soc. Am. B 27, A130–A136 (2010). sate in a cavity: theory of photodetection measurements,” New J. [142] Kosmas V. Kepesidis and Michael J. Hartmann, “Bose-hubbard Phys. 14, 085011 (2012). model with localized particle losses,” Phys. Rev. A 85, 063620 [123] Manas Kulkarni, Baris Oztop,¨ and Hakan E. Tureci,¨ “Cavity- (2012). mediated near-critical dissipative dynamics of a driven conden- [143] Alexandre Le Boite,´ Giuliano Orso, and Cristiano Ciuti, sate,” Phys. Rev. Lett. 111, 220408 (2013). “Steady-State Phases and Tunneling-Induced Instabilities in the [124] R. Landig, F. Brennecke, R. Mottl, T. Donner, and T. Esslinger, Driven Dissipative Bose-Hubbard Model,” Phys. Rev. Lett. 110, “Measuring the dynamic structure factor of a quantum gas un- 233601 (2013). dergoing a structural phase transition,” Nature Communications [144] Jiasen Jin, Davide Rossini, Martin Leib, Michael J. Hartmann, 6, 7046 (2015). and Rosario Fazio, “Steady-state phase diagram of a driven [125] Ferdinand Brennecke, Rafael Mottl, Kristian Baumann, Renate QED-cavity array with cross-Kerr nonlinearities,” Phys. Rev. A 68

90, 023827 (2014). 023633 (2014). [145] Alexandre Le Boite,´ Giuliano Orso, and Cristiano Ciuti, “Bose- [163] Lucia Hackermuller,¨ Ulrich Schneider, Maria Moreno-Cardoner, Hubbard model: Relation between driven-dissipative steady Takuya Kitagawa, Thorsten Best, Sebastian Will, Eugene states and equilibrium quantum phases,” Phys. Rev. A 90, 063821 Demler, Ehud Altman, Immanuel Bloch, and Belen´ Pare- (2014). des, “Anomalous expansion of attractively interacting fermionic [146] Alberto Biella, Leonardo Mazza, Iacopo Carusotto, Davide atoms in an optical lattice,” Science 327, 1621–1624 (2010). Rossini, and Rosario Fazio, “Photon transport in a dissipative [164] Florian Meinert, Manfred J. Mark, Emil Kirilov, Katharina chain of nonlinear cavities,” Phys. Rev. A 91, 053815 (2015). Lauber, Philipp Weinmann, Michael Grobner,¨ Andrew J. Da- [147] Hui Deng, Hartmut Haug, and Yoshihisa Yamamoto, “Exciton- ley, and Hanns-Christoph Nagerl,¨ “Observation of many-body polariton Bose-Einstein condensation,” Rev. Mod. Phys. 82, dynamics in long-range tunneling after a quantum quench,” Sci- 1489–1537 (2010). ence 344, 1259–1262 (2014). [148] Tim Byrnes, Na Young Kim, and Yoshihisa Yamamoto, [165] Philipp M. Preiss, Ruichao Ma, M. Eric Tai, Alexander Lukin, “Exciton-polariton condensates,” Nat. Phys. 10, 803–813 (2014). Matthew Rispoli, Philip Zupancic, Yoav Lahini, Rajibul Islam, [149] Jonathan Keeling, Marzena H. Szymanska,´ and Peter B. Lit- and Markus Greiner, “Strongly correlated quantum walks in op- tlewood, Optical Generation and Control of Quantum Co- tical lattices,” Science 347, 1229–1233 (2015). herence in Semiconductor Nanostructures, edited by Gabriela [166] H. Pichler, A. J. Daley, and P. Zoller, “Nonequilibrium dynamics Slavcheva and Philippe Roussignol, NanoScience and Technol- of bosonic atoms in optical lattices: Decoherence of many-body ogy (Springer Berlin Heidelberg, Berlin, Heidelberg, 2010). states due to spontaneous emission,” Phys. Rev. A 82, 063605 [150] M. H. Szymanska,´ J. Keeling, and P. B. Littlewood, “Non- (2010). Equilibrium Bose-Einstein Condensation in a Dissipative Envi- [167] Hannes Pichler, Johannes Schachenmayer, Andrew J. Daley, and ronment,” arXiv:1206.1784 (2012). Peter Zoller, “Heating dynamics of bosonic atoms in a noisy op- [151] Jan Klaers, Julian Schmitt, Frank Vewinger, and Martin Weitz, tical lattice,” Phys. Rev. A 87, 033606 (2013). “Bose-Einstein condensation of photons in an optical microcav- [168] Matthew P. A. Fisher, Peter B. Weichman, G. Grinstein, and ity,” Nature 468, 545–8 (2010). Daniel S. Fisher, “Boson localization and the superfluid- [152] S. O. Demokritov, V. E. Demidov, O. Dzyapko, G. A. Melkov, transition,” Phys. Rev. B 40, 546–570 (1989). A. A. Serga, B. Hillebrands, and A. N. Slavin, “Bose-Einstein [169] D. Jaksch, C. Bruder, J. I. Cirac, C. W. Gardiner, and P. Zoller, condensation of quasi-equilibrium magnons at room temperature “Cold bosonic atoms in optical lattices,” Phys. Rev. Lett. 81, under pumping,” Nature 443, 430–3 (2006). 3108–3111 (1998). [153] Mathieu Alloing, Mussie Beian, Maciej Lewenstein, David [170] S. M. Barnett and P. M. Radmore, Methods in Theoretical Quan- Fuster, Yolanda Gonzlez, Luisa Gonzlez, Roland Combescot, tum Optics (Oxford University Press, Oxford, 1997). Monique Combescot, and Franois Dubin, “Evidence for a bose- [171] D. Leibfried, R. Blatt, C. Monroe, and D. Wineland, “Quantum einstein condensate of excitons,” EPL (Europhysics Letters) 107, dynamics of single trapped ions,” Rev. Mod. Phys. 75, 281–324 10012 (2014). (2003). [154] Markus Falkenau, Valentin V. Volchkov, Jahn Ruhrig,¨ Axel [172] Matteo Marcuzzi, Emanuele Levi, Sebastian Diehl, Juan P. Gar- Griesmaier, and Tilman Pfau, “Continuous Loading of a Con- rahan, and Igor Lesanovsky, “Universal Nonequilibrium Proper- servative Potential Trap from an Atomic Beam,” Phys. Rev. Lett. ties of Dissipative Rydberg Gases,” Phys. Rev. Lett. 113, 210401 106, 163002 (2011). (2014). [155] M.-O. Mewes, M. R. Andrews, D. M. Kurn, D. S. Durfee, C. G. [173] Florian Marquardt and Steven M Girvin, “Optomechanics,” Townsend, and W. Ketterle, “Output Coupler for Bose-Einstein Physics (College. Park. Md). 2, 40 (2009). Condensed Atoms,” Phys. Rev. Lett. 78, 582–585 (1997). [174] G. Lindblad, “On the generators of quantum dynamical semi- [156] Nicholas P. Robins, Cristina Figl, Matthew Jeppesen, Graham R. groups,” Communications in Mathematical Physics 48, 119–130 Dennis, and John D. Close, “A pumped atom laser,” Nat. Phys. (1976). 4, 731–736 (2008). [175] A. Kossakowski, “On quantum statistical mechanics of non- [157] N.P. Robins, P.A. Altin, J.E. Debs, and J.D. Close, “Atom lasers: hamiltonian systems,” Reports on Mathematical Physics 3, 247 Production, properties and prospects for precision inertial mea- – 274 (1972). surement,” Phys. Rep. 529, 265–296 (2013). [176] Dave Bacon, Andrew M. Childs, Isaac L. Chuang, Julia Kempe, [158] A. Imamoglu, R. J. Ram, S. Pau, and Y. Yamamoto, “Nonequi- Debbie W. Leung, and Xinlan Zhou, “Universal simulation of librium condensates and lasers without inversion: Exciton- Markovian quantum dynamics,” Phys. Rev. A 64, 062302 (2001). polariton lasers,” Phys. Rev. A 53, 4250–4253 (1996). [177] E. B. Davies, “Markovian Master Equations,” Commun. Math. [159] Michiel Wouters and Iacopo Carusotto, “Excitations in a Phys. , 91 (1974). Nonequilibrium Bose-Einstein Condensate of Exciton Polari- [178] Alexander Altland and Ben Simons, Condensed Matter Field tons,” Phys. Rev. Lett. 99, 140402 (2007). Theory, 2nd ed. (Cambridge University Press, Cambridge, 2010). [160] Michiel Wouters and Iacopo Carusotto, “Superfluidity and Crit- [179] R. Feynman, “Space-Time Approach to Non-Relativistic Quan- ical Velocities in Nonequilibrium Bose-Einstein Condensates,” tum Mechanics,” Rev. Mod. Phys. 20, 367–387 (1948). Phys. Rev. Lett. 105, 020602 (2010). [180] John W. Negele and Henri Orland, Quantum Many-Particle Sys- [161] Cristiano Ciuti and Iacopo Carusotto, “Input-output theory of tems (Westview Press, Boulder, 1998). cavities in the ultrastrong coupling regime: The case of time- [181] L. M. Sieberer, S. D. Huber, E. Altman, and S. Diehl, “Nonequi- independent cavity parameters,” Phys. Rev. A 74, 033811 (2006). librium functional renormalization for driven-dissipative Bose- [162] Alessio Chiocchetta and Iacopo Carusotto, “Quantum Langevin Einstein condensation,” Phys. Rev. B 89, 134310 (2014). model for nonequilibrium condensation,” Phys. Rev. A 90, [182] Daniel J Amit and Victor Martin-Mayor, Field Theory, the 69

Renormalization Group, and Critical Phenomena, 3rd ed. (World “Properties of multi-particle Green’s and vertex functions within Scientific, Singapore, 2005). Keldysh formalism,” J. Phys. A Math. Theor. 43, 103001 (2010). [183] E. Calzetta and B. L. Hu, “Nonequilibrium quantum fields: [205] Michael E. Peskin and Daniel V. Schroeder, An Introduction to Closed-time-path effective action, Wigner function, and Boltz- Quantum Field Theory (Westview Press, Boulder, 1995). mann equation,” Phys. Rev. D 37, 2878–2900 (1988). [206] H. K. Janssen, “Field-theoretic method applied to critical dy- [184] A. Z. Patashinskii and V. L. Pokrovskii, “Longitudinal suscepti- namics,” in Dyn. Crit. Phenom. Relat. Top., Lecture Notes in bility and correlations in degenerate systems,” Sov. Phys. JETP Physics, Vol. 104, edited by Charles P. Enz (Springer-Verlag, 37(4), 733 (1973). Berlin, 1979) pp. 25–47. [185] W. Zwerger, “Anomalous fluctuations in phases with a broken [207] Camille Aron, Giulio Biroli, and Leticia F Cugliandolo, “Sym- continuous symmetry,” Phys. Rev. Lett. 92, 027203 (2004). metries of generating functionals of Langevin processes with [186] D. F. Walls and G. J. Milburn, Quantum Optics, 2nd ed. colored multiplicative noise,” J. Stat. Mech. Theory Exp. 2010, (Springer-Verlag, Berlin Heidelberg, 2008). P11018 (2010). [187] Peter Talkner, “The failure of the quantum regression hypothe- [208] Camille Aron, Daniel G. Barci, Leticia F. Cugliandolo, sis,” Ann. Phys. (N. Y). 167, 390–436 (1986). Zochil Gonzalez Arenas, and Gustavo S. Lozano, “Dynami- [188] G. Ford and R. O’Connell, “There is No Quantum Regression cal symmetries of Markov processes with multiplicative white Theorem,” Phys. Rev. Lett. 77, 798–801 (1996). noise,” arXiv:1412.7564 (2014). [189] L. M. Sieberer, A. Chiocchetta, A. Gambassi, U. C. Tauber,¨ and [209] Alexander Altland, Alessandro De Martino, Reinhold Egger, S. Diehl, “Thermodynamic equilibrium as a symmetry of the and Boris Narozhny, “Transient fluctuation relations for time- Schwinger-Keldysh action,” Phys. Rev. B 92, 134307 (2015). dependent particle transport,” Phys. Rev. B 82, 115323 (2010). [190] E. M. Lifshitz and L. P. Pitaevskii, Statistical Physics, Part 2: [210] Albert Messiah, Quantum Mechanics II, 3rd ed. (North-Holland Theory of the Condensed State, 2nd ed. (Pergamon Press, New Pubsishing Company, Amsterdam, 1965). York, 1980). [211] M. Hafezi, P. Adhikari, and J. M. Taylor, “Engineering three- [191] M. Wouters and I. Carusotto, “Absence of long-range coherence body interaction and Pfaffian states in circuit QED systems,” in the parametric emission of photonic wires,” Phys. Rev. B 74, Phys. Rev. B 90, 060503 (2014). 245316 (2006). [212] R. Graham and T. Tel,´ “Steady-state ensemble for the complex [192] M. Szymanska,´ J. Keeling, and P. Littlewood, “Nonequilibrium Ginzburg-Landau equation with weak noise,” Phys. Rev. A 42, Quantum Condensation in an Incoherently Pumped Dissipative 4661–4677 (1990). System,” Phys. Rev. Lett. 96, 230602 (2006). [213] Aditi Mitra, So Takei, Yong Baek Kim, and A. J. Millis, [193] Guang-zhao Zhou, Zhao-bin Su, Bai-lin Hao, and Lu Yu, “Nonequilibrium quantum criticality in open electronic sys- “Closed time path Green’s functions and critical dynamics,” tems.” Phys. Rev. Lett. 97, 236808 (2006). Phys. Rev. B 22, 3385–3407 (1980). [214] Aditi Mitra and Thierry Giamarchi, “Thermalization and dis- [194] Kuang-chao Chou, Zhao-bin Su, Bai-lin Hao, and Lu Yu, “Equi- sipation in out-of-equilibrium quantum systems: A perturba- librium and nonequilibrium formalisms made unified,” Phys. tive renormalization group approach,” Phys. Rev. B 85, 075117 Rep. 118, 1–131 (1985). (2012). [195] P. Martin, E. Siggia, and H. Rose, “Statistical Dynamics of Clas- [215] Ralph Gebauer and Roberto Car, “Current in Open Quantum Sys- sical Systems,” Phys. Rev. A 8, 423–437 (1973). tems,” Phys. Rev. Lett. 93, 160404 (2004). [196] C. De Dominicis, “Techniques de renormalisation de la theorie [216] J. E. Avron, M. Fraas, and G. M. Graf, “Adiabatic Response for des champs et dynamique des phenomenes critiques,” J. Phys. Lindblad Dynamics,” J. Stat. Phys. 148, 800–823 (2012). 37, C247–C253 (1976). [217] J Goldstone, “Field theories with Superconductor solutions,” [197] Hans-Karl Janssen, “On a Lagrangean for classical field dynam- Nuovo Cim. 19, 154–164 (2008). ics and renormalization group calculations of dynamical critical [218] Jeffrey Goldstone, Abdus Salam, and Steven Weinberg, “Broken properties,” Zeitschrift fur¨ Phys. B Condens. Matter Quanta 23, Symmetries,” Phys. Rev. 127, 965–970 (1962). 377–380 (1976). [219] Y. Nambu and G. Jona-Lasinio, “Dynamical Model of Elemen- [198] R. Bausch, H. K. Janssen, and H. Wagner, “Renormalized field tary Particles Based on an Analogy with . I,” theory of critical dynamics,” Zeitschrift fur¨ Phys. B Condens. Phys. Rev. 122, 345–358 (1961). Matter Quanta 24, 113–127 (1976). [220] Yuki Minami and Yoshimasa Hidaka, “Spontaneous symmetry [199] P. M. Chaikin and T. C. Lubensky, Principles of condensed mat- breaking and Nambu-Goldstone modes in dissipative systems,” ter physics (Cambridge University Press, Cambridge, 1995). arXiv:1509.05042 (2015). [200] R Kubo, “The fluctuation-dissipation theorem,” Reports Prog. [221] Haruki Watanabe and Hitoshi Murayama, “Unified Description Phys. 29, 255–284 (1966). of Nambu-Goldstone Bosons without Lorentz Invariance,” Phys. [201] Alessio Chiocchetta, Andrea Gambassi, and Iacopo Carusotto, Rev. Lett. 108, 251602 (2012). “Laser operation and Bose-Einstein condensation: analogies and [222] Yoshimasa Hidaka, “Counting Rule for Nambu-Goldstone differences,” arXiv:1503.02816 (2015). Modes in Nonrelativistic Systems,” Phys. Rev. Lett. 110, 091601 [202] Ryogo Kubo, “Statistical-Mechanical Theory of Irreversible Pro- (2013). cesses. I. General Theory and Simple Applications to Magnetic [223] Haruki Watanabe and Hitoshi Murayama, “Effective Lagrangian and Conduction Problems,” J. Phys. Soc. Japan 12, 570–586 for Nonrelativistic Systems,” Phys. Rev. X 4, 031057 (2014). (1957). [224] Tomoya Hayata and Yoshimasa Hidaka, “Dispersion relations [203] Paul Martin and Julian Schwinger, “Theory of Many-Particle of Nambu-Goldstone modes at finite temperature and density,” Systems. I,” Phys. Rev. 115, 1342–1373 (1959). Phys. Rev. D 91, 056006 (2015). [204] Severin G. Jakobs, Mikhail Pletyukhov, and Herbert Schoeller, [225] D. W. Snoke, Solid State Physics: Essential Concepts (Addison- 70

Wesley, 2008). [245] J. Berges and D. Mesterhazy,´ “Introduction to the nonequilibrium [226] R. Graham and H. Haken, “Laserlight – first example of a functional renormalization group,” Nucl. Phys. B - Proc. Suppl. second-order phase transition far away from thermal equilib- 228, 37–60 (2012). rium,” Zeitschrift fur¨ Phys. 237, 31–46 (1970). [246] Thomas Gasenzer and Jan M. Pawlowski, “Towards far- [227] V. DeGiorgio and Marlan Scully, “Analogy between the Laser from-equilibrium quantum field dynamics: A functional Threshold Region and a Second-Order Phase Transition,” Phys. renormalisation-group approach,” Phys. Lett. B 670, 135–140 Rev. A 2, 1170–1177 (1970). (2008). [228] Christof Wetterich, “Exact evolution equation for the effective [247]L eonie´ Canet, Bertrand Delamotte, Olivier Deloubriere,` and potential,” Phys. Lett. B 301, 90–94 (1993). Nicolas Wschebor, “Nonperturbative Renormalization-Group [229]J urgen¨ Berges, Nikolaos Tetradis, and Christof Wetterich, “Non- Study of Reaction-Diffusion Processes,” Phys. Rev. Lett. 92, perturbative renormalization flow in quantum field theory and 195703 (2004). statistical physics,” Phys. Rep. 363, 223–386 (2002). [248]L eonie´ Canet, Hugues Chate,´ Bertrand Delamotte, Ivan Dornic, [230] Bertrand Delamotte, “An Introduction to the Nonperturbative and Miguel Munoz,˜ “Nonperturbative Fixed Point in a Nonequi- Renormalization Group,” in Renorm. Gr. Eff. F. Theory Ap- librium Phase Transition,” Phys. Rev. Lett. 95, 100601 (2005). proaches to Many-Body Syst. SE - 2, Lecture Notes in Physics, [249]L eonie´ Canet, “Reaction-diffusion processes and non- Vol. 852, edited by Achim Schwenk and Janos Polonyi (Springer perturbative renormalization group,” J. Phys. A. Math. Gen. 39, Berlin Heidelberg, 2012) pp. 49–132. 7901–7912 (2006). [231] Jean-Paul Blaizot, Ramon´ Mendez-Galain,´ and Nicolas´ Wsche- [250]L eonie´ Canet, Hugues Chate,´ and Bertrand Delamotte, “Quan- bor, “A new method to solve the non-perturbative renormaliza- titative Phase Diagrams of Branching and Annihilating Random tion group equations,” Phys. Lett. B 632, 571–578 (2006). Walks,” Phys. Rev. Lett. 92, 255703 (2004). [232] Jan M. Pawlowski, “Aspects of the functional renormalisation [251]L eonie´ Canet, Hugues Chate,´ and Bertrand Delamotte, “Gen- group,” Ann. Phys. (N. Y). 322, 2831–2915 (2007). eral framework of the non-perturbative renormalization group [233] Igor Boettcher, Jan M. Pawlowski, and Sebastian Diehl, “Ul- for non-equilibrium steady states,” J. Phys. A Math. Theor. 44, tracold atoms and the Functional Renormalization Group,” Nucl. 495001 (2011). Phys. B - Proc. Suppl. 228, 63–135 (2012). [252]L eonie´ Canet, Hugues Chate,´ Bertrand Delamotte, and Nicolas´ [234] Severin G. Jakobs, Volker Meden, and Herbert Schoeller, Wschebor, “Nonperturbative Renormalization Group for the “Nonequilibrium Functional Renormalization Group for Inter- Kardar-Parisi-Zhang Equation,” Phys. Rev. Lett. 104, 150601 acting Quantum Systems,” Phys. Rev. Lett. 99, 150603 (2007). (2010). [235] Severin G. Jakobs, Mikhail Pletyukhov, and Herbert [253]L eonie´ Canet, Hugues Chate,´ Bertrand Delamotte, and Nicolas´ Schoeller, “Nonequilibrium functional renormalization group Wschebor, “Nonperturbative renormalization group for the with frequency-dependent vertex function: A study of the single- Kardar-Parisi-Zhang equation: General framework and first ap- impurity Anderson model,” Phys. Rev. B 81, 195109 (2010). plications,” Phys. Rev. E 84, 061128 (2011). [236] C. Karrasch, S. Andergassen, M. Pletyukhov, D. Schuricht, [254]L eonie´ Canet, Hugues Chate,´ Bertrand Delamotte, and Nicolas´ L. Borda, V. Meden, and H. Schoeller, “Non-equilibrium current Wschebor, “Erratum: Nonperturbative renormalization group for and relaxation dynamics of a charge-fluctuating quantum dot,” the Kardar-Parisi-Zhang equation: General framework and first EPL (Europhysics Lett. 90, 30003 (2010). applications [Phys. Rev. E 84 , 061128 (2011)],” Phys. Rev. E 86, [237] Walter Metzner, Manfred Salmhofer, Carsten Honerkamp, 19904 (2012). Volker Meden, and Kurt Schonhammer,¨ “Functional renormal- [255] John B. Kogut, “An introduction to lattice gauge theory and spin ization group approach to correlated fermion systems,” Rev. systems,” Rev. Mod. Phys. 51, 659–713 (1979). Mod. Phys. 84, 299–352 (2012). [256] J. Ye, S. Sachdev, and N. Read, “Solvable spin glass of quantum [238]L eonie´ Canet and Hugues Chate,´ “A non-perturbative approach rotors,” Phys. Rev. Lett. 70, 4011–4014 (1993). to critical dynamics,” J. Phys. A Math. Theor. 40, 1937–1949 [257] N. Read, Subir Sachdev, and J. Ye, “Landau theory of quantum (2007). spin glasses of rotors and ising spins,” Phys. Rev. B 52, 384–410 [239] D. Mesterhazy,´ J. H. Stockemer, L. F. Palhares, and J. Berges, (1995). “Dynamic universality class of Model C from the functional [258] Sarang Gopalakrishnan, Benjamin L. Lev, and Paul M. Goldbart, renormalization group,” Phys. Rev. B 88, 174301 (2013). “Frustration and glassiness in spin models with cavity-mediated [240] Oliver J. Rosten, “Fundamentals of the exact renormalization interactions,” Phys. Rev. Lett. 107, 277201 (2011). group,” Physics Reports 511, 177 – 272 (2012). [259] Markus Muller,¨ Philipp Strack, and Subir Sachdev, “Quantum [241] M. Reuter, “Nonperturbative evolution equation for quantum charge glasses of itinerant fermions with cavity-mediated long- gravity,” Phys. Rev. D 57, 971–985 (1998). range interactions,” Phys. Rev. A 86, 023604 (2012). [242] R. Gezzi, Th. Pruschke, and V. Meden, “Functional renormaliza- [260] Hessam Habibian, Andre´ Winter, Simone Paganelli, Heiko tion group for nonequilibrium quantum many-body problems,” Rieger, and Giovanna Morigi, “Bose-glass phases of ultracold Phys. Rev. B 75, 045324 (2007). atoms due to cavity backaction,” Phys. Rev. Lett. 110, 075304 [243]J urgen¨ Berges, Alexander Rothkopf, and Jonas Schmidt, “Non- (2013). thermal Fixed Points: Effective Weak Coupling for Strongly Cor- [261] Leticia F. Cugliandolo and Gustavo Lozano, “Real-time nonequi- related Systems Far from Equilibrium,” Phys. Rev. Lett. 101, librium dynamics of quantum glassy systems,” Phys. Rev. B 59, 041603 (2008). 915–942 (1999). [244]J urgen¨ Berges and Gabriele Hoffmeister, “Nonthermal fixed [262] L. F. Cugliandolo, D. R. Grempel, G. Lozano, H. Lozza, and points and the functional renormalization group,” Nucl. Phys. B C. A. da Silva Santos, “Dissipative effects on quantum glassy 813, 383–407 (2009). systems,” Phys. Rev. B 66, 014444 (2002). 71

[263] L. F. Cugliandolo, D. R. Grempel, G. Lozano, and H. Lozza, 454, 744–747 (2008). “Effects of dissipation on disordered quantum spin models,” [283] Jin Kim and J. Kosterlitz, “Growth in a restricted solid-on-solid Phys. Rev. B 70, 024422 (2004). model,” Phys. Rev. Lett. 62, 2289–2292 (1989). [264] M. H. Szymanska,´ J. Keeling, and P. B. Littlewood, “Mean-field [284] Vladimir G. Miranda and Fabio´ D. A. Aarao˜ Reis, “Numerical theory and fluctuation spectrum of a pumped decaying Bose- study of the Kardar-Parisi-Zhang equation,” Phys. Rev. E 77, Fermi system across the quantum condensation transition,” Phys. 031134 (2008). Rev. B 75, 195331 (2007). [285] Enzo Marinari, Andrea Pagnani, and Giorgio Parisi, “Critical ex- [265] Liang He, Lukas M. Sieberer, Ehud Altman, and Sebastian ponents of the KPZ equation via multi-surface coding numerical Diehl, “Scaling properties of one-dimensional driven-dissipative simulations,” J. Phys. A. Math. Gen. 33, 8181–8192 (2000). condensates,” Phys. Rev. B 92, 155307 (2015). [286] S. Ghaisas, “Stochastic model in the Kardar-Parisi-Zhang uni- [266] Vladimir N. Gladilin, Kai Ji, and Michiel Wouters, “Spatial co- versality class with minimal finite size effects,” Phys. Rev. E 73, herence of weakly interacting one-dimensional nonequilibrium 022601 (2006). bosonic quantum fluids,” Phys. Rev. A 90, 023615 (2014). [287] Chen-Shan Chin and Marcel den Nijs, “Stationary-state skew- [267] Kai Ji, Vladimir N. Gladilin, and Michiel Wouters, “Temporal ness in two-dimensional Kardar-Parisi-Zhang type growth,” coherence of one-dimensional nonequilibrium quantum fluids,” Phys. Rev. E 59, 2633–2641 (1999). Phys. Rev. B 91, 045301 (2015). [288] Lei-Han Tang, Bruce Forrest, and Dietrich Wolf, “Kinetic sur- [268] C. De Dominicis, E. Brezin,´ and J. Zinn-Justin, “Field-theoretic face roughening. II. Hypercube-stacking models,” Phys. Rev. A techniques and critical dynamics. I. Ginzburg-Landau stochastic 45, 7162–7179 (1992). models without energy conservation,” Phys. Rev. B 12, 4945– [289] T Ala-Nissila, T Hjelt, J M Kosterlitz, and O Venal¨ ainen,¨ “Scal- 4953 (1975). ing exponents for kinetic roughening in higher dimensions,”J. [269] G. I. Sivashinsky, “Nonlinear analysis of hydrodynamic instabil- Stat. Phys. 72, 207–225 (1993). ity in laminar flamesI. Derivation of basic equations,” Acta As- [290] C. Castellano, M. Marsili, M. A. Munoz,˜ and L. Pietronero, tronaut. 4, 1177–1206 (1977). “Scale invariant dynamics of surface growth,” Phys. Rev. E 59, [270] Yoshiki Kuramoto, Chemical Oscillations, Waves, and Turbu- 6460–6475 (1999). lence (Springer, Berlin, 1984). [291] F. D. A. Aarao˜ Reis, “Universality in two-dimensional Kardar- [271] G. Grinstein, David Mukamel, R. Seidin, and Charles Bennett, Parisi-Zhang growth,” Phys. Rev. E 69, 021610 (2004). “Temporally periodic phases and kinetic roughening,” Phys. Rev. [292] Jeffrey Kelling and Geza´ Odor,´ “Extremely large-scale simula- Lett. 70, 3607–3610 (1993). tion of a Kardar-Parisi-Zhang model using graphics cards,” Phys. [272] G. Grinstein, C. Jayaprakash, and R. Pandit, “Conjectures about Rev. E 84, 061150 (2011). phase turbulence in the complex Ginzburg-Landau equation,” [293] Timothy Halpin-Healy, “(2+1)-Dimensional directed in Phys. D Nonlinear Phenom. 90, 96–106 (1996). a random medium: scaling phenomena and universal distribu- [273] Matthew P. A. Fisher and G. Grinstein, “Nonlinear transport and tions.” Phys. Rev. Lett. 109, 170602 (2012). 1/ f α noise in insulators,” Phys. Rev. Lett. 69, 2322–2325 (1992). [294] Andrea Pagnani and Giorgio Parisi, “Numerical estimate of the [274] Uwe C. Tauber¨ and Sebastian Diehl, “Perturbative Field- Kardar-Parisi-Zhang universality class in (2+1) dimensions.” Theoretical Renormalization Group Approach to Driven- Phys. Rev. E. Stat. Nonlin. Soft Matter Phys. 92, 010101 (2015). Dissipative Bose-Einstein Criticality,” Phys. Rev. X 4, 021010 [295] Timothy Halpin-Healy, “Extremal paths, the stochastic heat (2014). equation, and the three-dimensional Kardar-Parisi-Zhang univer- [275] N. D. Mermin and H. Wagner, “Absence of sality class.” Phys. Rev. E. Stat. Nonlin. Soft Matter Phys. 88, or in One- or Two-Dimensional Isotropic 042118 (2013). Heisenberg Models,” Phys. Rev. Lett. 17, 1133–1136 (1966). [296] Timothy Halpin-Healy, “Erratum: Extremal paths, the stochas- [276] Michiel Wouters and Iacopo Carusotto, “Goldstone mode of opti- tic heat equation, and the three-dimensional Kardar-Parisi-Zhang cal parametric oscillators in planar semiconductor microcavities universality class [Phys. Rev. E 88, 042118 (2013)],” Phys. Rev. in the strong-coupling regime,” Phys. Rev. A 76, 043807 (2007). E 88, 069903 (2013). [277] Thomas Kloss, Leonie´ Canet, and Nicolas´ Wschebor, “Nonper- [297] Timothy Halpin-Healy and George Palasantzas, “Universal cor- turbative renormalization group for the stationary Kardar-Parisi- relators and distributions as experimental signatures of (2 + Zhang equation: Scaling functions and amplitude ratios in 1+1, 1)-dimensional Kardar-Parisi-Zhang growth,” EPL (Europhysics 2+1, and 3+1 dimensions,” Phys. Rev. E 86, 051124 (2012). Lett. 105, 50001 (2014). [278] Daniel F. Litim, “Optimisation of the exact renormalisation [298] A. Chiocchetta and I. Carusotto, “Non-equilibrium quasi- group,” Phys. Lett. B 486, 92–99 (2000). condensates in reduced dimensions,” EPL (Europhysics Lett. [279] Daniel F. Litim, “Optimized renormalization group flows,” Phys. 102, 67007 (2013). Rev. D 64, 105007 (2001). [299] Leiming Chen and John Toner, “Universality for Moving Stripes: [280] R Guida and J Zinn-Justin, “Critical exponents of the N -vector A Hydrodynamic Theory of Polar Active Smectics,” Phys. Rev. model,” J. Phys. A. Math. Gen. 31, 8103–8121 (1998). Lett. 111, 088701 (2013). [281] S. Utsunomiya, L. Tian, G. Roumpos, C. W. Lai, N. Ku- [300] Georgios Roumpos, Michael Lohse, Wolfgang H Nitsche, mada, T. Fujisawa, M. Kuwata-Gonokami, A. Lo¨ffler, S. Hofling,¨ Jonathan Keeling, Marzena Hanna Szymanska,´ Peter B Little- A. Forchel, and Y. Yamamoto, “Observation of Bogoliubov exci- wood, Andreas Lo¨ffler, Sven Hofling,¨ Lukas Worschech, Alfred tations in exciton-polariton condensates,” Nat. Phys. 4, 700–705 Forchel, and Yoshihisa Yamamoto, “Power-law decay of the spa- (2008). tial correlation function in exciton-polariton condensates,” Proc. [282] J. T. Stewart, J. P. Gaebler, and D. S. Jin, “Using photoemission Natl. Acad. Sci. U. S. A. 109, 6467–72 (2012). spectroscopy to probe a strongly interacting Fermi gas,” Nature [301] Wolfgang H. Nitsche, Na Young Kim, Georgios Roumpos, 72

Christian Schneider, Martin Kamp, Sven Hofling,¨ Alfred ics of one-dimensional interacting bosons in a disordered poten- Forchel, and Yoshihisa Yamamoto, “Algebraic order and tial: Elastic dephasing and critical speeding-up of thermaliza- the Berezinskii-Kosterlitz-Thouless transition in an exciton- tion,” Phys. Rev. Lett. 113, 010601 (2014). polariton gas,” Phys. Rev. B 90, 205430 (2014). [318] Marco Tavora and Aditi Mitra, “Quench dynamics of one- [302] G. Dagvadorj, J.M. Fellows, S. Matyjaskiewicz,´ F.M. Marchetti, dimensional bosons in a commensurate periodic potential: A I. Carusotto, and M.H. Szymanska,´ “Nonequilibrium Phase quantum kinetic equation approach,” Phys. Rev. B 88, 115144 Transition in a Two-Dimensional Driven Open Quantum Sys- (2013). tem,” Phys. Rev. X 5, 041028 (2015). [319] I. V. Protopopov, D. B. Gutman, M. Oldenburg, and A. D. [303] K. G. Lagoudakis, M. Wouters, M. Richard, A. Baas, I. Caru- Mirlin, “Dissipationless kinetics of one-dimensional interacting sotto, R. Andre,´ Le Si Dang, and B. Deveaud-Pledran,´ “Quan- fermions,” Phys. Rev. B 89, 161104 (2014). tized vortices in an exciton-polariton condensate,” Nat. Phys. 4, [320] A. F. Andreev, “The hydrodynamics of two- and one-dimensional 706–710 (2008). liquids,” Sov. Physics JETP 51, 1038 (1980). [304] Dietrich Wolf, “Kinetic roughening of vicinal surfaces,” Phys. [321] D. M. Stamper-Kurn, A. P. Chikkatur, A. Gorlitz,¨ S. Inouye, Rev. Lett. 67, 1783–1786 (1991). S. Gupta, D. E. Pritchard, and W. Ketterle, “Excitation of [305] I. A. Shelykh, A. V. Kavokin, Yuri G. Rubo, T. C. H. Liew, Phonons in a Bose-Einstein Condensate by Light Scattering,” and G. Malpuech, “Polariton polarization-sensitive phenomena Phys. Rev. Lett. 83, 2876–2879 (1999). in planar semiconductor microcavities,” Semicond. Sci. Technol. [322] Thilo Stoferle,¨ Henning Moritz, Christian Schori, Michael Kohl,¨ 25, 013001 (2010). and Tilman Esslinger, “Transition from a strongly interacting [306] R. M. Stevenson, V. N. Astratov, M. S. Skolnick, D. M. Whit- 1d superfluid to a ,” Phys. Rev. Lett. 92, 130403 taker, M. Emam-Ismail, A. I. Tartakovskii, P. G. Savvidis, J. J. (2004). Baumberg, and J. S. Roberts, “Continuous wave observation of [323] Ulf Bissbort, Soren¨ Gotze,¨ Yongqiang Li, Jannes Heinze, massive polariton redistribution by stimulated scattering in semi- Jasper S. Krauser, Malte Weinberg, Christoph Becker, Klaus conductor microcavities,” Phys. Rev. Lett. 85, 3680–3 (2000). Sengstock, and Walter Hofstetter, “Detecting the Amplitude [307] J. J. Baumberg, P. G. Savvidis, R. M. Stevenson, A. I. Tar- Mode of Strongly Interacting Lattice Bosons by Bragg Scatter- takovskii, M. S. Skolnick, D. M. Whittaker, and J. S. Roberts, ing,” Phys. Rev. Lett. 106, 205303 (2011). “Parametric oscillation in a vertical microcavity: A polariton [324] F. Meinert, M. Panfil, M. J. Mark, K. Lauber, J.-S. Caux, and condensate or micro-optical parametric oscillation,” Phys. Rev. H.-C. Nagerl,¨ “Probing the excitations of a lieb-liniger gas from B 62, R16247–R16250 (2000). weak to strong coupling,” Phys. Rev. Lett. 115, 085301 (2015). [308] A. P. D. Love, D. N. Krizhanovskii, D. M. Whittaker, [325] Bruno Sciolla, Dario Poletti, and Corinna Kollath, “Two-time R. Bouchekioua, D. Sanvitto, S. Al Rizeiqi, R. Bradley, M. S. correlations probing the dynamics of dissipative many-body Skolnick, P. R. Eastham, R. Andre,´ and Le Si Dang, “Intrinsic quantum systems: Aging and fast relaxation,” Phys. Rev. Lett. Decoherence Mechanisms in the Microcavity Polariton Conden- 114, 170401 (2015). sate,” Phys. Rev. Lett. 101, 067404 (2008). [326] John Cardy and Uwe Tauber,¨ “Theory of Branching and Annihi- [309] S. F. Edwards and D. R. Wilkinson, “The Surface Statistics of a lating Random Walks,” Phys. Rev. Lett. 77, 4780–4783 (1996). Granular Aggregate,” Proc. R. Soc. A Math. Phys. Eng. Sci. 381, [327] Matteo Marcuzzi, Michael Buchhold, Sebastian Diehl, and Igor 17–31 (1982). Lesanovsky, “Absorbing state phase transition with competing [310] E. Wertz, A. Amo, D. D. Solnyshkov, L. Ferrier, T. C. H. Liew, quantum and classical fluctuations,” arXiv:1601.07305 (2016). D. Sanvitto, P. Senellart, I. Sagnes, A. Lemaˆıtre, A. V. Kavokin, [328] S Mittal, J Fan, S Faez, A Migdall, J M Taylor, and M Hafezi, G. Malpuech, and J. Bloch, “Propagation and Amplification “Topologically robust transport of photons in a synthetic gauge Dynamics of 1D Polariton Condensates,” Phys. Rev. Lett. 109, field,” Phys. Rev. Lett. 113, 087403 (2014). 216404 (2012). [329] Sunil Mittal, Sriram Ganeshan, Jingyun Fan, Abolhassan Vaezi, [311] Pia Gagel, Peter P. Orth, and Jorg¨ Schmalian, “Universal and Mohammad Hafezi, “Measurement of topological invariants Postquench Prethermalization at a ,” in a 2d photonic system,” Nature Photonics 10, 180 (2016). Phys. Rev. Lett. 113, 220401 (2014). [330] C.-E. Bardyn, M. A. Baranov, C. V. Kraus, E. Rico, [312] Pia Gagel, Peter P. Orth, and Jorg¨ Schmalian, “Universal A. Imamoglu, P. Zoller, and S. Diehl, “Topology by dissipation,” postquench coarsening and aging at a quantum critical point,” New J. Phys. 15, 085001 (2013). Phys. Rev. B 92, 115121 (2015). [331] Netanel H. Lindner, Gil Refael, and Victor Galitski, Nature [313] Alessio Chiocchetta, Marco Tavora, Andrea Gambassi, and Physics 7, 490 (2011). Aditi Mitra, “Short-time universal scaling in an isolated quantum [332] Torsten Karzig, Charles-Edouard Bardyn, Netanel H. Lindner, system after a quench,” Phys. Rev. B 91, 220302 (2015). and Gil Refael, “Topological polaritons,” Phys. Rev. X 5, 031001 [314] Philipp Strack, “Dynamic criticality far from equilibrium: One- (2015). loop flow of Burgers-Kardar-Parisi-Zhang systems with broken [333] O. Viyuela, A. Rivas, and M. A. Martin-Delgado, “Uhlmann Galilean invariance,” Phys. Rev. E 91, 032131 (2015). phase as a topological measure for one-dimensional fermion sys- [315] M Marcuzzi Lesanovsky, E Levi, W Li, J P Garrahan, B Olmos, tems,” Phys. Rev. Lett. 112, 130401 (2014). and I, “Non-equilibrium universality in the dynamics of dissipa- [334] Jan Carl Budich and Sebastian Diehl, “Topology of density ma- tive cold atomic gases,” New J. Phys. 17, 72003 (2015). trices,” Phys. Rev. B 91, 165140 (2015). [316] Maximilian Genske and Achim Rosch, “Floquet-Boltzmann [335] Marcos Rigol, Vanja Dunjko, and Maxim Olshanii, “Thermal- equation for periodically driven Fermi systems,” Phys. Rev. A ization and its mechanism for generic isolated quantum systems,” 92, 062108 (2015). Nature 452, 854–8 (2008). [317] Marco Tavora, Achim Rosch, and Aditi Mitra, “Quench dynam- [336] Michael Moeckel and Stefan Kehrein, “Interaction Quench in the 73

Hubbard Model,” Phys. Rev. Lett. 100, 175702 (2008). time-resolved terahertz spectroscopy,” Rev. Mod. Phys. 83, 543– [337] D. N. Basov, M. M. Fogler, A. Lanzara, Feng Wang, and Yuanbo 586 (2011). Zhang, “Colloquium: Graphene spectroscopy,” Rev. Mod. Phys. [339] Gerald B. Folland, Quantum Field Theory: A Tourist Guide for 86, 959–994 (2014). Mathematicians (American Mathematical Society, Providence, [338] Ronald Ulbricht, Euan Hendry, Jie Shan, Tony F. Heinz, and 2008). Mischa Bonn, “Carrier dynamics in studied with