Stochastic quantum for finite and extended systems

Heiko Appela,b,,∗, Massimiliano Di Ventrab

aFritz-Haber-Institut der Max-Planck-Gesellschaft, Faradayweg 4-6, D-14195 Berlin, Germany bUniversity of California San Diego, La Jolla, California 92093, USA cEuropean Theoretical Spectroscopy Facility

Abstract We present a detailed account of the technical aspects of stochastic quantum molecular dynamics, an ap- proach introduced recently by the authors [H. Appel and M. Di Ventra, Phys. Rev. B 80 212303 (2009)] to describe coupled electron-ion dynamics in open quantum systems. As example applications of the method we consider both finite systems with and without ionic motion, as well as describe its applicability to extended systems in the limit of classical ions. The latter formulation allows the study of important phenomena such as decoherence and energy relaxation in bulk systems and surfaces in the presence of time-dependent fields.

1. Introduction the measurement apparatus itself which necessarily projects non-unitarily the state of the system onto Time-dependent density-functional theory states of the observables. This is generally true for (TDDFT) calculations in the linear-response both electrons and ions, so that a first-principles limit are currently enjoying a large popularity description of their coupled dynamics in the presence due to their efficiency and success in describing of one or more environments is of fundamental low-lying excitation energies in molecular systems importance in order to describe phenomena and [1]. Beyond linear-response, many applications have compare with experiments. At this point, it is worth been investigated with TDDFT. Examples include noting that present quantum molecular dynamics electronic transport [2, 3, 4, 5], nonlinear optical (QMD) approaches, (e.g., the Born-Oppenheimer, response [6], or atoms and molecules in strong laser Ehrenfest or Car-Parrinello methods) either do not fields [7, 8]. In the latter cases, the time-dependent allow excited states dynamics (Born-Oppenheimer Kohn-Sham (TDKS) equations are usually evolved and Car-Parrinello methods) or, if they do (e.g., in real-time. However, the majority of these stud- Ehrenfest QMD), they do not permit electronic ies pertains to the description of closed quantum coupling to external environments. Indeed, in all systems, since the corresponding TDKS equations these approaches, energy dissipation and thermal describe a set of N particles evolving coherently coupling to the environment are usually described in time. 1 On the other hand, most experimental with additional thermostats coupled directly to the situations involve some level of energy dissipation classical nuclear degrees of freedom, which fall short and/or decoherence induced by either some envi- of describing the numerous physical phenomena ronments to which the given system is coupled, or associated with quantum decoherence and energy dissipation. In order to overcome these shortcomings, we have ∗Corresponding author: Email address: [email protected] (Heiko Appel) recently introduced a novel time-dependent den- 1Notable exceptions are the references [4, 9, 10, 11, 12]. sity functional approach based on stochastic time-

Preprint submitted to Elsevier March 24, 2011 dependent Kohn-Sham equations [13], where we allow bulk systems and surfaces. We are in the process of the coupling of both electrons and (in principle quan- implementing SQMD for extended systems and we tum) ions with external baths. This approach - we will report these results in a forthcoming publication have named stochastic quantum molecular dynamics [14]. (SQMD) - extends the previously introduced stochas- The paper is organized as follows. In Section 2 we tic time-dependent-current density-functional theory give an introduction to the theory of stochastic quan- (STDCDFT) [9, 10] to the coupled dynamics of elec- tum molecular dynamics. For completeness, this in- trons and ions. The latter was formulated to account cludes general aspects of open quantum systems as for electrons interacting with external environments, well as the basic theorem of SQMD. In Section 3 we without however including atomic motion. There- discuss the aspects of a practical implementation of fore, SQMD combines and improves on the strengths SQMD. Finally, in Section 4 we illustrate with some of STDCDFT and present QMD methods by greatly examples the application of SQMD to finite systems expanding the physical range of applications of these with and without ionic motion, and outline its exten- methods. sion to periodic systems. Conclusions are reported in Clearly, from a practical point of view the present Section 5. approach suffers - like all density-functional theory (DFT) based methods - from our limited knowledge of the properties of the exact exchange-correlation 2. Theory functional. Furthermore, in the present case, the ex- act functional depends not only on the electronic de- 2.1. Stochastic Schr¨odingerequation grees of freedom, but also on the ionic and bath(s) de- grees of freedom [13]. Nevertheless, due to the weak In the following, we consider an electron-ion many- system-bath(s) coupling assumption of the present body system coupled to a bosonic bath. For simplic- theory, as well as the limited number of systems ity, we will consider only a single bath, but the for- where quantum nuclear effects are of disproportion- mulation is trivially extended to the case of several ate importance, we may start by considering the limit environments. The total Hamiltonian of the entire of SQMD to classical nuclei and adopt the available system is then functionals of standard closed-system TDDFT. Like in any other practical application of DFT, it is the ˆ ˆ ˆ ˆ ˆ ˆ predictions that we obtain and comparison with ex- H = HS ⊗ IB + IS ⊗ HB + λHSB. (1) periments that will be the ultimate judge of the range of validity of the approximate functionals used. Our system of interest is described by the many-body ˆ In Ref. [13] we have outlined the details of the proof Hamiltonian HS and the environment degrees of free- ˆ of the theorem at the core of SQMD, and provided a dom are given in terms of HB. The interaction of the simple example of the relaxation dynamics of a finite system with the environment is given by the Hamil- ˆ system (a molecule) prepared in some excited state tonian HSB and is assumed to be weak in the sense and embedded in a thermal bath. However, there are that a perturbation expansion in terms of this cou- some technical details behind an actual implementa- pling can be performed. With λ we denote the cor- tion of this approach we have not reported yet, and responding coupling parameter for the system-bath which are nonetheless important if one is interested interaction. ˆ in using this method for practical computations. In The total system described by the Hamiltonian H this work we then present all the technical aspects follows a unitary time-evolution, which can be for- for a practical implementation and use of SQMD. In mulated for pure states either in terms of the time- addition, we present the theory behind its applicabil- dependent Schr¨odingerequation (TDSE), withh ¯ = 1 ity to extended systems which is of great importance in the study of decoherence and energy relaxation in i∂tΨ(t) = Hˆ (t)Ψ(t), (2)

2 or, alternatively for mixed states, in terms of the with xB the bath’s coordinates (including possibly Liouville-von Neumann equation spin), and expand the total wavefunction of Eq. (2) d h i in the complete set of orthonormal states formed by ρˆ(t) = −i Hˆ (t), ρˆ(t) , (3) {χ x } 2 dt n( B) , namely whereρ ˆ is the statistical operator. X Ψ(x , x ; t) = φ (x ; t)χ (x ), (5) Since Hˆ is a many-body Hamiltonian and we have S B n S n B n to deal, in principle, with infinitely many degrees of with φ (x ; t) some functions (not necessarily nor- freedom (due to the bath) it is not possible to solve n S malized) in the Hilbert space of the system S. Eq. (2) or (3) in practice, except for a few simple In order to see that in the presence of a bath the model cases. In addition, in most cases of interest the functions φ (x ; t) form a statistical ensemble de- microscopic knowledge about the bath is limited and n S scribing the properties of the subsystem S, let us only its macroscopic thermodynamic properties are proceed as follows. For a general observable Oˆ of known, e.g., one typically assumes that the bath is in S the system S we find after simple algebra (and using thermal equilibrium. However, we are only interested the orthonormality of the bath states χ (x )) in the dynamics of the system degrees of freedom. It n B is therefore desirable to find an effective description hΨ(xS, xB; t)|OˆS|Ψ(xS, xB; t)i = for the system only. X hφ x t |Oˆ |φ x t i. (6) To accomplish this we may trace out the bath de- n( S; ) S n( S; ) grees of freedom at the level of the statistical opera- n tor, namely we perform the operationρ ˆS = TrB{ρˆ}, Let us now normalize the functions φn(xS; t) by writ- ρ ing where ˆS is called the reduced statistical operator of p system S. It is worth pointing out here that this ψn(xS, t) ≡ φn(xS; t)/ pn(t) (7) procedure does not generally lead to a closed equa- with tion of motion for the reduced statistical operator and pn(t) = hφn(xS; t)|φn(xS; t)i, (8) one needs further approximations. Depending on the which, according to Eq. (5), is nothing other than the approximations involved, one may arrive at an effec- probability for the bath to be in the state χ (x ). tive quantum master equation for the reduced den- n B We can now define the following statistical opera- sity operatorρ ˆ [15, 16, 17]. As we will discuss later, S tor this approach has some drawbacks when used within X a density-functional formulation, both fundamental ρˆS ≡ pn(t)|ψn(xS; t)ihψn(xS; t)| - in view of the theorems of DFT - and practical, (9) ≡ |ψ(t)ihψ(t)|, since solving for the density matrix is computation- ally more demanding than solving directly for state and immediately recognize that the average (6) can vectors. be re-written as We therefore take here a different route. Instead of hΨ(x , x ; t)|Oˆ |Ψ(x , x ; t)i = Tr{ρˆ O }, (10) working with a derived/composite quantity like the S B S S B S S statistical operator, we summarize briefly how the bath degrees of freedom can be traced out directly 2 At first sight this expansion might seem formally similar to at the level of the wavefunction. The derivation that the factorization used for the Born-Oppenheimer (BO) approx- follows has been reported elsewhere in the literature imation. However, in the BO case the expansion coefficients (see, e.g., Ref. [18]). We repeat some steps here for depend on the dynamical variables of both subsystems, elec- trons and nuclei. This dependence originates from the fact that completeness and to clarify our starting point. the electronic Hamiltonian in the Born-Oppenheimer approxi- To this end let us consider the set of eigenfunctions mation has a parametric dependence on the nuclear degrees of {χn(xB)} of the bath Hamiltonian freedom. In the present case we assume that the partitioning of the total Hamiltonian in Eq. (1) is such that expansion (5) HˆBχn(xB) = εnχn(xB), (4) becomes exact.

3 namely, due to the interaction with the bath, the right hand side is a memory term that is recording system S is necessarily in a mixture of states de- the history of the time evolution.3 fined by the macrostate {pn(t), ψn(xS; t)}. We thus Note that, up to this point, we have made no expect that the equation of motion for the repre- approximations, i.e., the time evolution given by sentative wave-function ψ(t) of the subsystem S to Eq. (15) is still fully coherent. However, the solution be “naturally” stochastic, namely we expect to find of Eq. (15) is very involved and, apart from model an equation of motion that provides the macrostate systems, not feasible in practice. Furthermore, a so- {pn(t), ψn(xS; t)}. lution would require the initial conditions for all the In order to show this we follow the Feshbach microscopic degrees of freedom of the bath. These projection-operator method [19, 20] and define the cannot all be determined simultaneously by a mea- following projection operators surement. In practice, one rather has only knowl- edge about macroscopic thermodynamic properties Pˆn := IˆS ⊗ | χn ih χn | , (11) of the bath, like temperature and pressure. It is X therefore common to perform the following additional Qˆ := Iˆ ⊗ | χ ih χ | , (12) n S k k approximations which are motivated by the form of k6=n the system-bath interaction and the thermodynamic properties of the bath: (i) due to the assumed weak where Iˆ is the identity in the Hilbert space of the S coupling between system and bath the source and system. The rationale behind the choice of the above memory terms are expanded up to second order in operators is to obtain the equation of motion of a the system-bath coupling parameter λ, (ii) the bath representative coefficient φ (x ; t). n S and subsystem S are assumed to be uncorrelated at By acting with these projection operators on the the initial time, (iii) a random phase approximation many-body TDSE for the combined system and bath is performed for the phases in the source and memory in Eq. (2) we arrive at terms4, and (iv) it is assumed that the bath degrees of freedom form a dense energy spectrum and are in i∂ Pˆ Ψ(t) = Pˆ Hˆ Pˆ Ψ(t) + Pˆ Hˆ Qˆ Ψ(t), (13) t n n n n n local thermal equilibrium characterized by

1 ˆ ρ e−βHB , i∂ Qˆ t Qˆ Hˆ Qˆ t Qˆ Hˆ Pˆ t . ˆB = ˆ (16) t nΨ( ) = n nΨ( ) + n nΨ( ) (14) Tr(e−βHB )

Equation (14) can be formally solved. Inserting the where β = 1/kBT . result back into Eq. (13) we obtain

−iQˆHˆ Qtˆ 3These equations have a formal similarity to the quantum i∂tPˆΨ(t) = PˆHˆ PˆPˆΨ(t) + PˆHˆ Qeˆ QˆΨ(0) transport formulation introduced by Kurth et. al. [21]. How- Z t iQˆHˆ Qˆ(τ−t) ever, in this case the projection operators project on the real − i PˆHˆ Qeˆ QˆHˆ PˆPˆΨ(τ) dτ, space regions of leads and central molecular device. Also the 0 bath is fermionic in the quantum transport case (leads) and (15) electrons can be exchanged between system and “bath”. This is in contrast to the present case where we consider bosonic where we have omitted the index n for brevity. The baths and only energy and momentum can be exchanged be- first term on the right hand side of Eq. (15) con- tween system and bath. 4The random phase approximation invoked in the deriva- tains only projections on the system manifold, and tion of the Markovian stochastic Schr¨odingerequation might describes the coherent evolution of the system degrees seem at first sight surprising. The derivation of the Lindblad of freedom. The second term is a source term that equation from the Markovian stochastic Schr¨odingerequation carries a dependence on initial conditions (QˆΨ(0) are on the other hand shows, that both describe the exact same dynamics if the Hamiltonian does not depend on internal de- the initial conditions of all system’s states except the grees of freedom or any time-dependent or stochastic field (see one we are considering), and the third term on the section. 2.2).

4 Let us then write the interaction Hamiltonian as Equation (20) is a general non-Markovian stochastic Schr¨odingerequation. Indeed, it still contains a time- ˆ X ˆ ˆ HSB = Sα ⊗ Bα, (17) integral over the past dynamics which is originating α from the memory term of Eq. (15). Even though where Sˆα and Bˆα are - in the most general case - the theorem of SQMD could be formulated with non- many-body operators that act on the Hilbert spaces Markovian baths we will focus in the following only of the system and bath, respectively. In the following on the Markovian limit we will also assume that the average of the opera- 0 0 tors Bˆα vanishes on the n-th eigenstate of the bath, Cαβ(t − t ) = δαβδ(t − t ), (23) namely X Sˆαh χn | Bˆα | χn i = 0. (18) namely, we consider baths that are δ-correlated. α Physically, this means that the bath does not retain If this is not the case we simply redefine the system memory of the interaction with the system which is Hamiltonian via valid when the typical thermalization time-scales in- side the bath are much faster than the thermalization ˆ 0 ˆ X ˆ ˆ HS = HS + λ Sαh χn | Bα | χn i, (19) time-scales of the system. This approximation is well α justified for a large number of bath degrees of free- ˆ 0 ˆ dom. If this assumption does not hold, one has to and the interaction Hamiltonian as HSB = HSB − P ˆ ˆ ˆ resort to the solution of the more involved Eq. (20). λ α Sαh χn | Bα | χn i. The term h χn | Bα | χn i thus contributes to the unitary evolution of the system by By inserting the Markov approximation, Eq. (23), renormalizing its eigenvalues (a typical example of into Eq. (20) we then arrive at the stochastic this is the Lamb shift [16, 17]). Schr¨odingerequation in the Born-Markov limit With these approximations in place, the source i X † term can be regarded as a stochastic driving term i∂tψ(t) =HˆS(t)ψ(t) − Sˆ Sˆαψ(t) 2 α [18]. This is because, the system’s state we have α (24) singled out in Eq. (11) now interacts with a (prac- X + lα(t)Sˆαψ(t), tically infinite) large set of bath states densely dis- α tributed in energy. The previously coherent Eq. (15) then has to be regarded as a non-Markovian stochas- where the parameter λ has been absorbed in the tic Schr¨odingerequation for the general state vector operators Sˆα. The first term on the right hand ψ(t) ≡ φn(xS; t)/hφn(xS; t)|φn(xS; t)i [18] side of Eq. (24) is the usual unitary evolution of X the system under the action of the system Hamil- i∂tψ(t) = HˆSψ(t) + λ lα(t)Sˆαψ(t) tonian HˆS, the second term describes the dissipa- α t Z ˆ tion effects introduced by the bath and would in- 2 X ˆ† −iHS (t−τ) ˆ − iλ Cαβ(t − τ)Sαe Sβψ(τ)dτ deed make the probability density generated by ψ(t) αβ 0 decay in time. The last term, however, introduces + O(λ3), fluctuations so that the norm of the state vector (20) ψ(t) averaged over the ensemble is conserved, namely hψ(t)|ψ(t)i = 1 + O(λ4). where lα(t) are stochastic processes with zero ensem- Due to the stochastic nature of this equation, the ble average, lα(t) = 0, and correlation functions stochastic process described by Eq. (24) has to be 0 simulated in terms of an ensemble of state vectors lα(t)lβ(t ) = 0, (21) ψ(t). Each member ψ(t) of the ensemble evolves dif- ferently in time due to the random variables lα(t) in ∗ 0 0 lα(t)lβ(t ) = Cαβ(t − t ). (22) the third term on the rhs. of Eq. (24). If we consider

5 ˆ ˆ j an initial mixed state degrees of freedom - Hamiltonians HS =6 HS[{|ψki}] - or do not depend explicitly on some stochastic field, X 0 ρˆS(0) = pk | ψk(0) ih ψk(0) | , (25) like e.g., a stochastic thermostat [22], it is possible to k derive the Lindblad equation [23, 24, 16, 17] from the where p0 are the probabilities (with P p0 = 1) of stochastic Schr¨odingerequation (24). k k k For notational clarity, let us denote in the following finding the state ψk(0) in the ensemble, the statistical discussion with |ψi a single member of the stochastic average over all members of the ensemble allows us to j construct the reduced density operator for the system ensemble {|ψki}. If we consider for simplicity the case degrees of freedom of a single bath operator in Eq. (24), and observe that in the Markovian limit X 0 ρˆS(t) = pk | ψk(t) ih ψk(t) | . (26) Z t 0 0 k W (t) = l(t )dt (28) 0 Here, we use the symbol ··· to indicate the statistical average over all members of the ensemble of state is a Wiener process [15] with properties dW = 0 i and dW †dW = dt, we can formulate the stochastic vectors ψk(t), namely the ensemble {ψk(t)} of state vectors with initial conditions ψk(0). Schr¨odingerequation (24) for a single bath in differ- The expectation value of a general physical observ- ential form according to able of the system S, OˆS, can then be computed as   1 † in Eq. (10), i.e., d|ψi = −iHˆS|ψi − Sˆ Sˆ|ψi dt − iSˆ|ψidW. (29) 2 ˆ ˆ h OS i = Tr(ˆρS(t)OS) Next, we employ Itˆostochastic calculus in order to X 0 ˆ compute the following differential = Tr( pk | ψk(t) ih ψk(t) | OS) (27) k d|ψihψ| = (d|ψi)hψ|+|ψi(dhψ|)+(d|ψi)(dhψ|). (30) X 0 ˆ = pkh ψk(t) | OS | ψk(t) i, k Unlike in normal calculus, we also have to keep the third term in the product rule above. This becomes where the last step shows that the construction of necessary, since a statistical average over the Wiener ρˆS(t) is not actually required: we can compute expec- increment dW †dW is proportional to dt, which will tation values of observables directly from the wave- cause terms quadratic in dW to contribute to first functions in the usual way, followed by a statistical order in dt. Inserting Eq. (29) and its Hermitian average over all members of the ensemble of state vec- conjugate into Eq. (30) we arrive after elementary tors. It is also important to note that this approach algebra at provides directly the full distribution of the given ob- servable at any given time, provided we can compute d|ψihψ| = − iSˆ|ψihψ|dW + h.c. a large enough set of realizations of the stochastic h i 1n † o − i HˆS, |ψihψ| dt − Sˆ S,ˆ |ψihψ| dt processes lα(t) (for an example of this see, e.g., Ref. 2 [13]). From this distribution we can then compute all + Sˆ|ψihψ|Sˆ†dW †dW higher moments and/or cumulants (e.g., the variance, ˆ ˆ skewness, etc.) some of which are directly accessible + S|ψihψ|HS dW dt + h.c. (31) experimentally. i + Sˆ|ψihψ|Sˆ†Sˆ dW dt + h.c. 2 2.2. Derivation of the Lindblad equation and stochas- 2 1 † † 2 + HˆS|ψihψ|HˆSdt + Sˆ Sˆ|ψihψ|Sˆ Sdtˆ tic Hamiltonians 4 For many-body Hamiltonians which are not i n † o 2 + HˆS, |ψihψ|Sˆ Sˆ dt . stochastic, namely they do not depend on internal 2

6 In order to construct the statistical operator from We have thus shown that the stochastic j the state vectors of the statistical ensemble {|ψki}, Schr¨odinger equation of Eq. (24) and the mas- we perform in the next step the statistical average ter equation (34) lead to the same statistical over all members in the ensemble, i.e. operator, if and only if the Hamiltonian is not stochastic. However, in order to prove any DFT dρˆ = d|ψihψ|. (32) theorem one is led to consider the dynamics of the actual many-body system and that of any auxiliary dW dW dt Taking the properties = 0, = 0 and one (including the Kohn-Sham system) with different dW †dW dt l t = of the stochastic process ( ) into ac- interaction potentials, but reproducing the exact count, we see that only the second and third line in many-body density or current density. It is then dt Eq. (31) contribute to first order in and we arrive at this stage that a choice has to be made - in at the case of a many-body system open to one or h i 1n † o more environments - regarding the basic equation of dρˆ = − i HˆS, |ψihψ| dt − Sˆ S,ˆ |ψihψ| dt 2 motion to work with. If we choose to work with a + Sˆ|ψihψ|Sˆ†dt + O(dt2). (33) quantum master equation of the type (34), then we are assuming from the outset that the Kohn-Sham At this point, note that this equation of motion is Hamiltonian is not stochastic. But this is an hypoth- not necessarily closed forρ ˆ = |ψihψ| because the esis that constitutes part of the final thesis, namely first term on the right hand side of Eq. (33) is not we have to prove that this statement is correct, not h i assume it a priori [27]. This issue does not arise with equal to the commutator −i HˆS, ρˆS unless HˆS =6 the stochastic Schr¨odingerequation (24), because in ˆ j ˆ HS[{|ψki}], or HS does not depend on any stochas- that case we can consider all possible Hamiltonians, tic field, or the system is in a pure state at all times including those that are stochastic. - which would amount to the case Sˆ = 0.5 How- ever, if the Hamiltonian is stochastic, one has to deal with an ensemble of Hamiltonians, and the statisti- cal average of the first term on the right hand side of In addition to the above important point, we also Eq. (33) involves also a statistical average over these recall that for arbitrary time-dependent operators Hamiltonians (see, e.g., Refs. [25, 26]). Sˆ(t) and Hˆ (t), Eq. (34) may not yield a positive- ˆ S For the moment being, let us assume that HS =6 definite statistical operator at all times (see, e.g., ˆ j HS[{|ψki}] and furthermore that the Hamiltonian Ref. [28]). This is a major limitation in practical HˆS does not depend on some external stochastic field. calculations, since loss of positivity (which precludes In this case we find a statistical interpretation of physical observables) should then be checked at every instant of time. Note h i 1 n † o † ∂tρˆS = −i HˆS, ρˆS − Sˆ S,ˆ ρˆS + SˆρˆSSˆ (34) that such a limitation does not pertain to the stochas- 2 tic Schr¨odingerequation which can be equally applied which is the well-known quantum master equation in to arbitrary time-dependent operators without possi- Born-Markov limit (or Lindblad equation if the bath ble loss of positivity of the ensuing statistical opera- operators and the Hamiltonian, do not depend on tor. The construction of the statistical operator from time) [23, 24, 16, 17]. stochastic trajectories, cf. Eq. (9), effectively selects only the physical solutions of the associated quantum master equation while the latter also permits non- 5A further complication would arise if the operators Sˆ de- ˆ ˆ j physical solutions. Therefore, the above issues make pended on internal degrees of freedom, i.e., S = S[{|ψki}]. In that case, the average over the statistical ensemble in the sec- the equation of motion of the statistical operator a ond and third terms on the right hand side of Eq. (33) has to less solid starting point for a DFT theory of open be performed over the operators Sˆ as well. quantum systems.

7 2.3. Theorem of Stochastic QMD for the ions. The total particle and current density We are now in a position to state the basic theorem operators of the system can then be written as of SQMD. Before doing this let us define the basic X quantities we work with. The many-body system we Nˆ(x, t) =n ˆ(r, t) + Nˆα(R, t) (39) are interested in consists of Ne electrons with coor- α r ≡ {r } N P N dinates j and n = s s,n nuclei, where for the particle number, and each nuclear species s comprises Ns,n particles with Z M j ...N X charges s,j, masses s,j, = 1 s,n, and coor- Jˆ(x, t) = ˆj(r, t) + Jˆα(R, t) (40) dinates R ≡ {Rs,j}, respectively. Their dynamics - α subject to an arbitrary classical electromagnetic field, whose vector potential is A(t) - is described by the for the current density. To simplify the notation we Hamiltonian have also denoted with x ≡ {R, r} the combined set of electronic and nuclear coordinates, and we use the ˆ ˆ ˆ ˆ ˆ HS(t) = Te(r, t) + Wee(r) + Uext,e(r, t) + Wen(r, R) combined index α = {s, j} for the nuclear species. + Tˆn(R, t) + Wˆ nn(R) + Uˆext,n(R, t), (35) We now formulate the theorem for a single bath operator. It trivially extends to many operators. ˆ ˆ where Te(t) and Tn(t) are the kinetic energies of Theorem.— For a given bath operator Sˆ, many-body electrons and ions, with velocitiesv ˆk(t) = [ˆpk + initial state Ψ(x, t = 0) (not necessarily pure) and ex- ˆ ˆ ˆ eA(ˆrk, t)]/m and Vα(t) = [Pα − ZαA(Rα, t)]/Mα, ternal vector potential A(x, t), the dynamics of the ˆ ˆ respectively and Uext,e(r, t), Uext,n(R, t) the external stochastic Schr¨odingerequation in Eq. (24) generates potentials acting on electrons and ions. The particle- ensemble-averaged total particle and current densi- particle interactions are given by ties N(x, t) and J(x, t). Under reasonable physical 0 Ne Ne assumptions, any other vector potential A (x, t) (but 1 X e2 X Wˆ ee(r) = ≡ wee(ˆrj − ˆrk), same initial state and bath operator) that leads to 4π0 |ˆrj − ˆrk| j

Nn rent density, has to coincide, up to a gauge transfor- 1 X Zα Zβ Wˆ nn(R) = mation, with A(x, t). 4π0 ˆ ˆ α<β |Rα − Rβ| A sketch of the proof of this theorem can be found N in the original paper [13]. We thus refer the reader Xn ≡ wnn(Rˆ α − Rˆ β), to this publication for more details. Here, we just α<β mention an important point. As in Ref. [9, 10] we

Ne Nn are implicitly assuming that given an initial condi- 1 X X e Zα Wˆ en(r, R) = − tion, bath operator, and ensemble-averaged current 4π0 ˆ k=1 α=1 |ˆrk − Rα| density, a unique ensemble-averaged density can be N N obtained from its equation of motion: Xe Xn ≡ wen(ˆrk − Rˆ α). ∂N(x, t) k=1 α=1 = − ∇ · J(x, t) (41) (36) ∂t  1 1  We then define the charge current operator + Sˆ†nˆSˆ − Sˆ†Sˆnˆ − nˆSˆ†Sˆ . e X 2 2 ˆj(r, t) = {vˆk(t), δ(r − ˆrk)} (37) 2m k If we write this equation in the compact form for the electrons, and ∂tN(x, t) = −∇ · J(x, t) + FB(x, t) (42) ˆ Zα X ˆ ˆ Jα(R, t) = {Vβ(t), δ(R − Rβ)}, (38) F x, t 2Mα the above amounts to saying that B( ) is a func- β Zα=Zβ ,Mα=Mβ tional of N(x, t) and J(x, t), or better of J(x, t) alone,

8 and that Eq. (41) admits a unique physical solution. where |ΨKSi is a Slater determinant of single-particle Therefore, unlike what has been recently argued [12], wave-functions and the density is not independent of the current den- N 2 sity, and our theorem establishes a one-to-one corre- X [ˆpk + eAeff (ˆrk, R, t)] HˆKS = (44) spondence between current density and vector poten- 2m k=1 tial. If this were not the case, namely that the parti- cle and current densities were independent functions, is the Hamiltonian of non-interacting particles with then FB(x, t) would not be completely determined by the sole knowledge of N(x, t) and J(x, t) [27]. Aeff (ˆrk, R, t) = Aext(ˆrk, R, t) + Ahxc(ˆrk, R, t), (45)

where Aext is the vector potential applied to the true 2.4. The limit of classical nuclei and Kohn-Sham many-body system, and Ahxc is the vector potential scheme whose only scope is to mimic the correct dynamics of the ensemble-averaged current density, and we have At this stage we may formulate a Kohn-Sham lumped in it also the Hartree interaction potential in (KS) scheme of SQMD where an exchange-correlation addition to the xc one. All these potentials depend (xc) vector potential Axc - functional of the initial on the instantaneous classical nuclear coordinates R. states, bath operator(s), and ensemble-averaged cur- We immediately note that for a general many-body rent density - acting on non-interacting species, al- bath operator acting on many-body states one can- lows to reproduce the exact ensemble-averaged den- not reduce Eq. (43) to a set of independent single- sity and current densities of the original interacting particle equations, as done in the usual DFT schemes many-body system. The ensuing charge and current for closed systems. In other words, this would gen- densities would contain all possible correlations in the erally require the solution of an equation of motion system - if the exact functional were known. of Slater determinants, which is still computation- However, in the present case, we could construct ally quite demanding. To see this point, suppose we several schemes - based on corresponding theorems - have N particles and retain M single-particle states. by defining different densities and current densities. M We then need to solve for CN − 1 elements of the For instance, we could collect all nuclear densities M state vector (with CN = M!/N!(M − N)! and the into one quantity as done in Ref. [29]. This, by no −1 comes from the normalization condition). In ad- means is a limitation of this approach. Rather, it al- dition, one has to average over an amount, call it m, lows us to “specialize” the given schemes to specific of different realizations of the stochastic process. The physical problems. Instead, a much more serious lim- problem thus scales exponentially with the number of itation relates to the construction of xc functionals particles. If this seems prohibitive let us also recall for the chosen scheme. Therefore, as anticipated in that a density-matrix formalism would be even more the introduction, we will restrict ourselves here to the computationally demanding, requiring the solution of limit of classical nuclei. M M (CN +2)×(CN −1)/2 coupled differential equations, Let us then assume that we know the vector po- even after taking into account the constraints of her- tential Aeff that generates the exact current density miticity and unit trace of the density matrix. in the non-interacting system. By construction, the It was recently suggested in Ref. [30] that for op- ˆ P ˆ system follows the dynamics induced by the stochas- erators of the type O = j Oj, namely operators tic Schr¨odinger equation (for a single bath operator) that can be written as sum over single-particle opera- tors (like the density or current density), the expecta- ˆ   tion value of O over a many-particle non-interacting 1 † d|ΨKSi = −iHˆKS − Sˆ Sˆ |ΨKSidt state with dissipation can be approximated as a sum 2 (43) of single-particle expectation values of Oˆj over an − iSˆ|ΨKSidW ensemble of N single-particle systems with specific

9 single-particle dissipation operators. In particular, which ground-state orbitals (occupied/unoccupied) it was found that the approximate single-particle have to be taken into account for the bath operator scheme provides an excellent approximation for the at a given temperature. current density compared to the exact many-body For the present paper, we further assume that the calculation. [30] We refer the reader to Ref. [30] for rates are independent of space and orbital indices, i.e, the numerical demonstration of this scheme and its γjk(r) ∝ 1/τ, where τ is a relaxation time. The bath analytical justification. operators of this model are sufficient for the illustra- From now on, for numerical convenience, we will tion purposes of the present work but they clearly then adopt the same ansatz which in the present case provide only a simplified picture of the full system- reads, bath interaction. We emphasize here, that a rigorous form of the bath operators and the associated relax- N Xe ation rates can always be derived from the micro- h |Oˆ| i ' hφj |Oˆ |φj i, ΨKS ΨKS KS j KS (46) scopic form of the complete Hamiltonian of system j=1 and bath and their mutual interaction, Eq. (1). For example, in the case of a phonon bath, the system- with |φj i single-particle KS states solutions of KS bath interaction Hamiltonian could be taken into ac- 2 count in terms of e.g., a Fr¨ohlich interaction. In that j (ˆp + eAeff (ˆr, R, t)) j d|φKSi = − i |φKSidt (47) case the relaxation rates can be extracted from the 2m electron-phonon coupling matrix elements. The sit- 1 ˆj† ˆj j ˆj j − SspSsp|φKSidt − iSsp|φKSidW (t), uation is much simpler is the case of a photon bath, 2 where the Einstein rates of stimulated and sponta- ˆj neous emmission can be used. with Ssp an operator acting on single particle states. 2.6. Forces on ions 2.5. Model for the system-bath interaction Once we have the single-particle KS states and the corresponding Slater determinant Ψ (x, t) at hand The single-particle operators Sj in Eq. (47) that KS sp we can compute the forces on the nuclei as [13] we employ in the present work are given by the fol- lowing time-independent projectors 6 Fα(t) = −hΨKS(x, t)|∇αHˆKS(x, t)|ΨKS(x, t)i ,

j q (49) ˆ 0 0 7 Skk0 (r) = δkj(1 − δkk ) γjk (r)fD(k) for each realization of the stochastic process. Note, GS GS that this force is stochastic in nature since the wave- × | ψj (r) ih ψk0 (r) | , (48) functions in the above expectation value are solu- h  i−1 tions of a stochastic Schr¨odingerequation. Since ap- εk−µ where fD(εk) = 1 + exp denotes the kB T proximations to the xc functional of the Kohn-Sham usual Fermi- distribution and δkj(1 − δkk0 ) de- Hamiltonian may make the latter stochastic then the note Pauli blocking factors. The projectors in Eq. force one would obtain using a density matrix ap- (48) cause a relaxation of the system back to the proach - e.g., by solving the quantum master equa- GS ground-state orbitals | ψj (r) i with a rate given by tion (34) - would not be necessarily equal to the av- the rate constants γjk(r) (generally space and orbital dependent), while the temperature in the Fermi fac- tor is modeling the temperature of the bath. We em- 7Note that Eq. (49) is not the expression for the force phasize, that the Fermi factor in Eq. (48) determines one would obtain from the Hellmann-Feynman theorem. This is because we are considering a system out of equilibrium. Rather, Eq. (49) is the total time derivative of the average of the ion momentum operator over the state of the system 6Cf. also to the examples in Refs. [30, 31, 10, 13]. (see, e.g., Ref. [32]).

10 erage force obtained from Eq. (49) by averaging over Higher-order approximations can be easily derived the ensemble of realizations. from the Magnus series and appropriate quadrature points and weights, but experience shows that the 3. Simulation Algorithms second order gives a good balance between speed and accuracy for many applications. In the present work We have now outlined the general theory behind we use this approximation for the piecewise deter- SQMD and we are ready to move on to the descrip- ministic evolution that we are going to introduce in tion of its actual implementation. the next section.

3.1. Real-time propagation The standard real-time propagation of the Kohn- 3.2. Quantum Jump Algorithm Sham orbitals for a closed quantum system is based We are now left with the actual solution of the on numerical approximations for the time-ordered stochastic Schr¨odingerequation (24). In past work, evolution operator this has been done by directly integrating this equa- Z t+∆t ! tion with standard approaches - e.g., with appropri- ˆ ˆ ˆ U(t + ∆t, t) = T exp −i HKS(τ)dτ . (50) ately modified Runge-Kutta methods (see e.g. [10, 35] t and references therein). These approaches are rea- There are several approaches employed in standard sonable when we deal with a small number of ac- TDDFT computer packages, like, e.g., octopus [33] cessible states or short propagation times. However, to evaluate Eq. (50) numerically. Here, we have cho- they become increasingly unstable with an increasing sen the Magnus propagator as basic building block number of states or for very long timescales, which for our stochastic simulations. [34] The Magnus se- is the case for realistic systems, like molecular struc- ries [34] provides an exact expression for the time- tures, surfaces or solids. As an alternative, we have evolution operator (50) as a time-unordered exponen- thus adopted the quantum jump algorithm pioneered tial of so called Magnus operators Ωj in the form in the work of Di´osi[36], Dalibard [37], Zoller and   Gardiner and collaborators [38, 39, 40] as well as Uˆ(t + ∆t, t) = exp Ωˆ 1 + Ωˆ 2 + Ωˆ 3 + ··· , (51) Carmichael [41]. At the price of introducing the prop- agation of auxiliary states, the quantum jump algo- ˆ where the Ωj are given in terms of time-integrals over rithm provides improved stability for systems with nested commutators of the Hamiltonian at different a large number of states/particles and, due to the points in time piecewise deterministic evolution, also a stable prop- Z t+∆t agation scheme for long timescales. Ωˆ 1 = − i HˆKS(τ)dτ This algorithm works as follows. Consider the t deterministic time-evolution given by the follwing Z t+∆t Z τ1 (52) norm-preserving non-linear Schr¨odingerequation Ωˆ 2 = [HˆKS(τ1), HˆKS(τ2)]dτ2dτ1 t t   . d i 2 . ψj(t) = −i HˆS + ||Sψˆ || ψj(t), (54) . dt 2 The time-integrals can be evaluated numerically with e.g., a Gauss-Legendre quadrature. In the simplest where the non-Hermitian Hamiltonian HS is given by case, which is accurate up to second order in the time- i † step, one arrives at the exponential midpoint rule HˆS = HˆS − S S. (55) 2   ˆ ˆ 3 U(t + ∆t, t) = exp Ω1 + O(∆t ) ˆ ˆ (53) As before, the operators HS and S denote the Her- 3 Ωˆ 1 = −iHˆKS(t + ∆t/2) + O(∆t ). mitian system Hamiltonian and the bath operator

11 t0 t1 t1 t2 t2 t3 t3 t4 t4

waiting-time distribution φ(t0)= ψ(t0) φ(t1)= Sψˆ (t1) auxilary state

physical state

ψ(t0) ψ(t1) Sψˆ (t1)

Figure 1: The figure illustrates the time-evolution as generated by the quantum jump algorithm. The lower track represents the piecewise deterministic propagation of the physical state which is intercepted at instances in time where the bath operator Sˆ acts on the state. The points in time where this takes place are determined by sampling a waiting-time distribution. The sampling is performed by propagating an auxiliary state (represented in the upper track) with a non-Hermitian Hamiltonian. Uniformly distributed random numbers are drawn and once the norm of the auxiliary state drops below the current random number the propagation of the physical and the auxiliary state is suspended. At this point in time the action of the bath operator on the physical state results in a new state which is then also used to initialize the auxiliary state for the evolution. The simulation of both states is then resumed again.

of Eq. (24) 8. The main objective of the quantum samples the waiting-time distribution on the fly from jump algorithm is to sample the stochastic process the norm decay. given by Eq. (24) in terms of a piecewise determinis- In terms of the Kohn-Sham system, the steps of tic evolution, i.e. a set of deterministic time intervals the algorithm can then be summarized as follows generated by the evolution of Eq. (54) and action of the bath operator Sˆ between two consecutive time 1) Draw a uniform random number ηk ∈ [0, 1] for intervals. A central ingredient of the algorithm is the Kohn-Sham Slater determinant a waiting-time distribution which determines when aux 2) Propagate N auxiliary orbitals φj under the the jumps (i.e., actions of the bath operator) appear non-Hermitian dynamics throughout the simulation.   aux ˆ i ˆ† ˆ aux In order to sample the unknown waiting-time dis- i∂tφj = HKS − S S φj , j = 1 ...N aux 2 tribution, an auxiliary set of wavefunctions φj is aux introduced. The wavefunctions φj are propagated 3) Propagate the orbitals ψKS, j = 1 ...N of the with the non-Hermitian Hamiltonian Hˆ alongside j S Kohn-Sham system with a norm-conserving dy- the actual states ψ . Since the auxiliary system j namics according to evolves with a non-Hermitian Hamiltonian, the norm of the states φaux is not preserved. It can be shown   j KS i † KS 2 KS i∂tψ = HˆKS − Sˆ Sˆ + i||Sψˆ || ψ [16] that the decay of the norm of the auxilary wave- j 2 j j functions is related to the waiting-time distribution. The algorithm makes use of this fact and directly 4) If the norm of the auxiliary Slater determinant aux (formed by the orbitals φj ) drops below the 8For convenience we consider here the case of a single bath. drawn random number ηj, act with the bath op- A generalization to many baths is straightforward. erator(s) on the Kohn-Sham orbitals and update

12 the auxiliary orbitals the situations considered here which involve frequent jumps, we found this degradation of orthogonality ( ψKS(t ) = Sψˆ KS(t ) to be a minor effect. Different models for the bath ||Det{φaux(t )}|| ≤ η → j k j k j k k aux KS operators might require stricter strategies for main- φj (tk) = ψj (tk) taining orthogonality and this should be checked case by case. 5) In case this is not already accounted for by the From our numerical experience so far, the waiting- choice of the bath operator Sˆ, the orbitals have time distribution seems to follow mainly a single ex- to be orthonormalized after step 4). ponential distribution. It would therefore appear ap- 6) Go to step 1) pealing to parametrize this distribution and to draw the waiting times from the analytical expression of The piecewise deterministic evolution that is gen- the parametrization. In this way the propagation of erated by the steps of this algorithm is illustrated the auxiliary states could be avoided and a speedup schematically in Fig. 1. of the propagation by a factor of two could be gained. Averaging at any given time over an ensemble of However, at the moment it is not clear if the wait- stochastic realizations allows then to obtain mean ing times of the Kohn-Sham system follow always an values of physical observables. It is also important exponential distribution. In particular, the shape of to realize that we have a full statistical ensemble at the distribution is unknown, when, e.g., ionic motion hand. This allows us to compute distributions of is involved, or when the system is subject to strong observables, higher order moments, cumulants, etc. external electric or magnetic fields. Therefore, to be quite easily without the need to compute the statis- on the safe side, in the present work we always sam- tical operator. ple the waiting-time distribution by propagating the We also emphasize here, that the interpretation of auxiliary system. a single stochastic trajectory is not meaningful: the It is also worth noting that the average over stochastic realizations have to be considered always stochastic realizations of the ensemble generally con- as an ensemble. When averages over the stochastic verges faster when the system-bath interaction in- ensemble are performed, the “convergence” of all ob- creases. In the opposite limit the convergence is slow. servables of interest has to be checked carefully by When the system-bath interaction is very weak, only increasing the number of realizations of the stochas- a small damping will be exerted by the Sˆ†Sˆ term in tic process. Eq. (56), and hence it takes longer for the norm of the Note that without further constraints the action auxiliary wave functions to drop below their waiting of the bath operator in step 4) of the algorithm can times. This in turn implies that fewer jumps occur in principle lead to a loss of orthogonality. For ex- and hence more stochastic realizations are required ample, all orbitals of the Slater determinant could to converge to a smooth observable distribution. relax to the same orbital shape. The system could loose in this way its fermionic character. In order to maintain the fermionic nature of the Kohn-Sham 4. Applications state vector, we have to ensure that the orbitals of 4.1. Finite Systems the Kohn-Sham Slater determinant remain orthogo- nal. To achieve this, we perform an orthogonalization In the last section we have introduced technical de- of the orbitals after each action of the bath operators. tails for the quantum jump algorithm that we use to This orthogonalization can be thought of as being simulate the stochastic process associated with the part of the definition of the action of the operator Sˆ. stochastic Kohn-Sham equation, Eq. (47). In this When using this prescription, the orthogonality can section we apply the algorithm to molecular systems in principle degrade between two subsequent jumps with and without clamped ions. As first example and is restored only after each jump. However, for we consider a situation of clamped ionic coordinates.

13 Figure 2: Real part of the HOMO (left panel) and LUMO (right panel) orbitals of (1,4)-phenylene-linked zincbacteriochlorin-bacteriochlorin.

As testcase we investigate a (1,4)-phenylene-linked zincbacteriochlorin-bacteriochlorin complex. Due to an extra Mg atom in the left porphyrin ring of the complex the molecule is not fully symmetric. As a re- sult, the highest-occupied molecular orbital (HOMO) of this molecule is located on the left porphyrin ring, whereas the lowest-unoccupied molecular or- bital (LUMO) is located on the right porphyrin ring, cf. Fig. 2. This system has been used as a model to study charge-transfer excitations in linear response TDDFT [42]. Here instead we consider open and closed system real-time propagation. We prepare the zincbacteriochlorin-bacteriochlorin complex in an en- tangled initial-state, where the orbital of the HOMO Figure 3: Snapshots of the time evolution of the HOMO or- bital of zincbacteriochlorin-bacteriochlorin with clamped ions. is replaced by The plots display the real part of the orbital. In the left col- umn the closed quantum system evolution is shown at different TDKS 1 h GS π GS i −i 2 points in time and the right column displays the evolution of ψHOMO (t = 0) = √ ψHOMO + e ψLUMO , (56) 2 the system with a coupling to a thermal bath. A rather fast re- laxation rate of τ = 1 a.u. has been used for the bath operator GS GS in the open quantum system case. where ψHOMO and ψLUMO denote the ground-state HOMO and LUMO, respectively. For all other or- bitals the ground-state configuration is used at the initial time. Starting from this excited initial Slater LUMO, respectively. Only small nonlinearities arise determinant the system is then evolved freely in time due to the dependence of the Kohn-Sham Hamilto- without any external fields. For the bath operators nian on the time-dependent density. Since the system we employ the model of Eq. (48) introduced in section is propagated as closed quantum system, the phase 2.5 at zero temperature. The dynamics of the system oscillations would continue indefinitely. is illustrated in Fig. 3 where we plot the real part On the other hand, the open quantum system evo- TDKS of the HOMO orbital for different snapshots in time. lution in the right panel of Fig. 3 shows for ψHOMO (t) The left panel summarizes the closed system evolu- a fast relaxation from the entangled initial state back tion and the right panel an ensemble average over to the HOMO which is localized on the left porphyrin 100 stochastic realizations in the open system case. ring. If we now imagine computing Ehrenfest forces Let us first focus on the closed quantum system case. from these orbital contributions, it is clear that the Due to the entangled initial state in Eq. (56), the forces will differ qualitatively in the closed and open TDKS time-dependence of the orbital ψHOMO (t) has mainly quantum system cases. While in the closed quan- oscillatory phase contributions exp(−iHOMOt) and tum system case the forces will be oscillatory, they exp(−iLUMOt) from the ground-state HOMO and will show relaxation behavior similar to the orbitals

14 Figure 4: Left panel: Here we show the ionic positions of a Neon dimer as function of time for a closed quantum system. As initial condition we have selected a stretched configuration of the dimer which results in an indefinite coherent oscillation of the two nuclei. Right panel: Using the same initial state we have evolved with SQMD a stochastic ensemble of trajectories. Shown is the average of the nuclear positions for an ensemble with 100 stochastic realizations. As relaxation time for the simulation we have employed τ = 300 fs. The ionic velocities follow a Maxwell-Boltzmann distribution with a temperature of 290K and the electronic Fermi factors in our model bath operators (cf. Eq. (48)) are employed at the same temperature. in the open quantum case. This example emphasizes tial conditions the open quantum system evolution that the coupling of electronic degrees of freedom to a within SQMD. For the SQMD simulation we have thermal bath yields qualitatively different forces com- employed a relaxation time of τ = 300 fs and an av- pared to standard QMD approaches. erage over 100 stochastic trajectories has been per- This observation motivates our second example, formed which results in a smooth decay of the nuclear where we consider a stochastic QMD simulation for oscillations. We emphasize at this point that the dy- a neon dimer. In this case the ions are not clamped namics of the dimer has to be regarded as an evolu- at the equilibrium configuration. Instead we use tion out of equilibrium. Only for long times we can stretched initial positions for the ions of the dimer address the thermalization of electrons and ions. De- as initial state for the open and closed system propa- pending on the choice of bath operators it might be gation. If we would treat the ions quantum mechan- possible to find electrons and ions at different “effec- ically, then the bath operators would also act on the tive” temperatures during the time-evolution. How- nuclear wavefunctions. However, since we have re- ever, such a study is beyond the scope of the present stricted ourselves here to the limit of classical ions illustration of our approach and will be investigated we replace this action of the bath operators by mod- in a future work. ifying the velocities of the ions. At every occasion when the bath operators act on the electronic wave- 4.2. Extended Systems functions we draw in addition new velocities for the ions from a Maxwell-Boltzmann distribution. This is So far we have considered only finite molecular sys- a simple approach but can be improved with e.g., re- tems. However, a large class of applications requires cently introduced stochastic thermostats for the ions also the treatment of periodic boundary conditions in [22]. one, two or three dimensions. This includes for ex- In the left panel of figure Fig. 4 we show the ionic ample decoherence and dissipation in nanowires, elec- positions of the dimer as function of time for a stan- tronic relaxation on surfaces, or hot electron thermal- dard Ehrenfest TDDFT closed quantum system evo- ization in bulk systems. For these cases it is desirable lution. In the right panel we display for the same ini- to extend our approach to periodic systems. In this

15 section we briefly discuss the necessary steps in order fixed. Again, this could be accomplished in a super- to apply SQMD to extended systems. cell geometry by coupling some “bulk” layers away There are some extra details and conditions that from the surface with a local operator that maintains have to be satisfied in order to treat periodic systems energy equilibrium in that region (an example of such with SQMD. As a first step we expand the stochastic operator is given in Ref. [43]). The rest of the sys- Kohn-Sham orbitals of the periodic system of interest tem is let to follow its own dynamics. If we excite in the complete set of the corresponding ground-state the molecules and/or surface - e.g., by application of Bloch-orbitals ϕk(r) a short electromagnetic field - electrons and ions can then distribute energy and momentum first in the lay- X ψk(r, t) = dkk0 (t)ϕk0 (r). (57) ers adjacent to the surface and then relax energy into k0 the bath, where they would thermalize to the appro- priate canonical distributions. Analogously, we could This gives rise to stochastic expansion coefficients monitor energy relaxation of electrons and ions in a dkk0 (t) which are then propagated in time using, e.g., bulk exited either thermally or electrically and kept the quantum jump algorithm that we have presented at a given temperature by a thermal stage. Impor- in section 3.2. Similar to the case of molecules, the tant phenomena that are then accessible would be, fermionic nature of the electronic subsystem needs to e.g., phase transitions driven by dissipative effects. be taken into account by orthogonalizing the occu- pied states after each application of the bath operator 5. Conclusions (cf. step (5) of the quantum jump algorithm). In addition, care has to be taken for the choice In summary, we have presented a detailed account of gauge for the vector and scalar potentials in the of stochastic quantum molecular dynamics. The ap- Hamiltonian of the extended system. Here, the proach is based on a stochastic Schr¨odingerequation, same restrictions apply as in standard closed-system which may or may not describe Markovian dynam- TDDFT simulations. In practice, we consider only ics - although we have focused the discussion to the purely time-dependent vector potentials which retain Markovian case. Our approach allows us to describe the periodicity of the considered system at all times. the dynamics of electrons and ions coupled to one In the present context we have to assume in addition or many external environments. For simplicity we that the bath operators retain the periodicity of the have restricted our examples to the situation of clas- extended system as well. This restricts the choice of sical ions, but the approach is, in principle, valid also baths represented by local operators that satisfy the for quantum ions. Although we have not reported condition any actual implementation of SQMD for periodic sys- Sˆ(r) = Sˆ(r + R) (58) tems, we have outlined the theory behind its exten- sion to this important case. Work along these lines where R denotes the usual displacement vector of the is in progress and will be reported elsewhere [14]. unit cell. This may exclude certain relaxation mech- This approach is thus amenable to studying many anisms. However, the importance of these relaxation interesting phenomena related to energy relaxation and dephasing channels can always be checked by in- and dephasing of the electronic subsystem in the pres- creasing the size of the supercell that is used in the ence of ionic dynamics, such as local ionic and elec- simulation. tronic heating in laser fields, relaxation processes in While we do not have fully implemented this photochemistry, etc., a feature that is lacking in any scheme yet, we want to argue about important physi- “standard” molecular dynamics approach. cal processes that can be studied with this approach. We acknowledge support from DOE under grant For instance, one could study adsorption of molecules DE-FG02-05ER46204 and Lockheed Martin. on surfaces whose opposite side is set on a thermal stage that keeps the electron and/or ion temperatures [1] M. A. L. Marques, C. A. Ullrich, F. Nogueira,

16 A. Rubio, K. Burke, E. K. U. Gross, Time- [11] X. Zheng, G. Chen, Y. Mo, S. Koo, H. Tian, Dependent Density Functional Theory (Lecture C. Yam, Y. Yan, Time-dependent density func- Notes in Physics), Springer-Verlag, Berlin, 2006. tional theory for quantum transport, J Chem. Phys. 133 (2010) 114101. [2] M. Di Ventra, T. N. Todorov, Transport in nanoscale systems: the microcanonical versus [12] J. Yuen-Zhou, C. Rodr´ıguez-Rosario, grand-canonical picture, J. Phys. Cond. Mat. 16 A. Aspuru-Guzik, Time-dependent current- (2004) 8025. density functional theory for generalized open quantum systems, Phys. Chem. Chem. Phys. 11 [3] N. Bushong, N. Sai, M. Di Ventra, Approach to (2009) 4509. steady-state transport in nanoscale conductors, Nano Letters 5 (2005) 2569. [13] H. Appel, M. Di Ventra, Stochastic quantum molecular dynamics, Phys. Rev. B 80 (2009) [4] K. Burke, R. Car, R. Gebauer, Density Func- 212303. tional Theory of the Electrical Conductivity of Molecular Devices, Phys. Rev. Lett. 94 (2005) [14] H. Appel, M. Di Ventra, Stochastic relaxation in 146803. periodic systems, In preparation. [5] S. Kurth, G. Stefanucci, E. Khosravi, C. Ver- [15] N. G. V. Kampen, Stochastic Processes in dozzi, E. K. U. Gross, Dynamical coulomb block- Physics and Chemistry, Elsevier Science and ade and the derivative discontinuity of time- Technology, 3rd Ed., 2007. dependent density functional theory, Phys. Rev. [16] H. P. Breuer, F. Petruccione, The Theory Lett. 104 (2010) 236801. of Open Quantum Systems, Oxford University [6] F. D. Vila, D. A. Strubbe, Y. Takimoto, X. An- Press, Oxford, 2002. drade, A. Rubio, S. G. Louie, J. J. Rehr, Basis [17] U. Weiss, Quantum Dissipative Systems (Series set effects on the hyperpolarizability of chcl[sub in Modern Condensed Matter Physics), World 3]: -type orbitals, numerical basis sets Scientific Publishing Company; 3 edition, 2008. and real-space grids, J. Chem. Phys. 133 (2010) 034111. [18] P. Gaspard, M. Nagaoka, Non-markovian stochastic Schr¨odingerequation, J. Chem. Phys. [7] M. Petersilka, E. K. U. Gross, Strong-field dou- 111 (1999) 5676. ble ionization of helium: A density-functional perspective, Laser Phys. 9 (1999) 105. [19] H. Feshbach, Unified theory of nuclear reactions, Annals of Physics 5 (1958) 357. [8] T. Kreibich, N. Gidopoulos, R. van Leeuwen, E. Gross, Towards time-dependent density- [20] S. Nordholm, S. A. Rice, A quantum ergodic the- functional theory for molecules in strong laser ory approach to unimolecular fragmentation, J. pulses, Prog. Theor. Chem. and Phys. 14 (2003) Chem. Phys. 62 (1975) 157. 69. [21] S. Kurth, G. Stefanucci, C.-O. Almbladh, A. Ru- [9] M. Di Ventra, R. D’Agosta, Stochastic time- bio, E. K. U. Gross, Time-dependent quan- dependent current-density-functional theory, tum transport: A practical scheme using den- Phys. Rev. Lett. 98 (2007) 226403. sity functional theory, Phys. Rev. B 72 (2005) 035308. [10] R. D’Agosta, M. Di Ventra, Stochastic time- dependent current-density-functional theory: A [22] G. Bussi, D. Donadio, M. Parrinello, Canonical functional theory of open quantum systems, sampling through velocity rescaling, J. Chem. Phys. Rev. B 78 (2008) 165105. Phys. 126 (2007) 014101.

17 [23] G. Lindblad, On the generators of quantum dy- [34] W. Magnus, On the exponential solution of dif- namical semigroups, Commun. Math. Phys. 48 ferential equations for a linear operator, Comm. (1976) 119. Pure Appl. Math. 7 (1954) 649. [24] C. Gardiner, Stochastic Methods: A Handbook [35] P. E. Kloeden, E. Platen, Numerical Solution for the Natural and Social Sciences, Springer Se- of Stochastic Differential Equations (Stochastic ries in Synergetics, Vol. 13, 2nd ed., 1985. Modelling and Applied Probability), Springer, Berlin, 1992. [25] R. Balian, From microphysics to macrophysics, Springer, Berlin, 1991. [36] L. Diosi, Quantum stochastic processes as mod- els for state vector reduction, Journal of Physics [26] M. Di Ventra, Electrical transport in nanoscale A: Mathematical and General 21 (1988) 2885. systems, Cambridge University Press, Cam- bridge, 2008. [37] J. Dalibard, Y. Castin, K. Mølmer, Wave- [27] M. Di Ventra, R. D’Agosta, Comment on ”time- function approach to dissipative processes in dependent current-density functional theory for quantum optics, Phys. Rev. Lett. 68 (1992) 580. generalized open quantum systems” by J. Yuen- [38] R. Dum, A. S. Parkins, P. Zoller, C. W. Gar- Zhou, C. Rodriiguez-Rosario and A. Aspuru- diner, Monte carlo simulation of master equa- Guzik, arXiv http://arxiv.org/abs/0911.3932 tions in quantum optics for vacuum, thermal, (2009) 3. and squeezed reservoirs, Phys. Rev. A 46 (1992) [28] S. Maniscalco, F. Intravaia, J. Piilo, A. Messina, 4382. Misbeliefs and misunderstandings about the [39] C. W. Gardiner, A. S. Parkins, P. Zoller, Wave- non-markovian dynamics of a damped harmonic function quantum stochastic differential equa- oscillator, Journal of Optics B: Quantum and tions and quantum-jump simulation methods, Semiclassical Optics 6 (2004) S98. Phys. Rev. A 46 (1992) 4363. [29] T. Kreibich, E. K. U. Gross, Multicomponent [40] R. Dum, P. Zoller, H. Ritsch, Monte carlo sim- density-functional theory for electrons and nu- ulation of the atomic master equation for spon- clei, Phys. Rev. Lett. 86 (2001) 2984. taneous emission, Phys. Rev. A 45 (1992) 4879. [30] Y. V. Pershin, Y. Dubi, M. Di Ventra, Effective [41] H. J. Carmichael, S. Singh, R. Vyas, P. R. Rice, single-particle order- n scheme for the dynam- Photoelectron waiting times and atomic state re- ics of open noninteracting many-body systems, duction in resonance fluorescence, Phys. Rev. A Phys. Rev. B 78 (2008) 054302. 39 (1989) 1200. [31] N. Bushong, M. Di Ventra, The decay of excited [42] A. Dreuw, M. Head-Gordon, Failure of He from stochastic density-functional theory: a time-dependent density functional theory quantum measurement theory interpretation, J. for long-range charge-transfer excited states: Phys. Cond. Mat. 20 (2008) 395214. The zincbacteriochlorin-bacteriochlorin and [32] M. Di Ventra, S. T. Pantelides, Hellmann- bacteriochlorophyll-spheroidene complexes, J. feynman theorem and the definition of forces Am. Chem. Soc. 126 (2004) 4007. in quantum time-dependent and transport prob- lems, Phys. Rev. B 61 (2000) 16207. [43] Y. Dubi, M. Di Ventra, Thermoelectric effects in nanoscale junctions, Nano Letters 9 (2009) 97. [33] A. Castro, et al., octopus: a tool for the appli- cation of time-dependent density functional the- ory, phys. stat. sol. B 243 (2006) 2465.

18