This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Diffusive limit of non-Markovian quantum jumps

Kimmo Luoma [email protected] Institut für Theoretische Physik, Technische Universität Dresden, D-01062, Dresden, Germany    Walter T. Strunz Institut für Theoretische Physik, Technische Universität Dresden, D-01062, Dresden, Germany    Jyrki Piilo Turku Center for Quantum Physics, Department of Physics and Astronomy, University of Turku, FI-20014, Turun Yliopisto, Finland
Abstract

We solve two long standing problems for stochastic descriptions of open quantum system dynamics. First, we find the classical stochastic processes corresponding to non-Markovian quantum state diffusion and non-Markovian quantum jumps in projective Hilbert space. Second, we show that the diffusive limit of non-Markovian quantum jumps can be taken on the projective Hilbert space in such a way that it coincides with non-Markovian quantum state diffusion. However, the very same limit taken on the Hilbert space leads to a completely new diffusive unraveling, which we call non-Markovian quantum diffusion. Further, we expand the applicability of non-Markovian quantum jumps and non-Markovian quantum diffusion by using a kernel smoothing technique allowing a significant simplification in their use. Lastly, we demonstrate the applicability of our results by studying a driven dissipative two level atom in a non-Markovian regime using all of the three methods.

Introduction.—

Deriving and solving the equations of motion for driven dissipative quantum systems is a notoriously hard task, especially in the presence of quantum memory effects. In this Letter, we open new avenues to tackle these problems of broad on-going interest. Currently, state-of-the-art experiments explore driven dissipative open quantum systems Bloch et al. (2008), non-equilibrium phase transitions in a Rydberg gas has been observed Gutiérrez et al. (2017), simulation of general open system dynamics with trapped ions has been reported Barreiro et al. (2011); Müller et al. (2011) – and even the statistical likelihood of a physical process (a statistical arrow of time) has been experimentally characterized using superconducting qubit systems Harrington et al. (2019). Similar type of open quantum systems appear also in the context of photosynthesis Engel et al. (2007); Lee et al. (2007) and in general in molecular aggregates Brixner et al. .

One of the main difficulties in analyzing driven open quantum systems has its origin in the lack of a typical time scale, such as an energy gap of the system Hamiltonian. One possible solution is to try to model the open system and environment dynamics exactly, as in non-Markovian quantum state diffusion Diósi et al. (1998); Strunz et al. (1999), where a stochastic Schrödinger equation describes the dynamics of the open system and the effects of the environment are contained in the statistical properties of the driving noise. Typically approximation methods are required to solve the resulting equations of motion Yu et al. (1999); Suess et al. (2014); Hartmann and Strunz (2017). This type of approach has been successfully used to describe energy Roden et al. (2009); Ritschel et al. (2011); Abramavicius and Abramavicius (2014) and charge transport Gao and Eisfeld (2019); Zhong and Zhao (2013) in molecular aggregates.

Alternatively, starting from a microscopic model an effective time local master equation can be derived Breuer et al. (2002) and unravelled, for example, with non-Markovian quantum jumps Piilo et al. (2008, 2009); Härkönen (2010). Quantum jump methods have been used earlier, e.g., to study excitonic energy transport with Ai et al. (2014); Tao et al. (2016) and without driving Rebentrost et al. (2009); Oh et al. (2019) and even to understand singlet fission in molecular crystals, which may help to design more efficient solar panels Renaud and Grozema (2015).

On the theoretical side, our motivation is to look for the missing connection between the quantum jump Piilo et al. (2008) and quantum state diffusion Strunz et al. (1999) approaches in the non-Markovian regime – and with the help of these results expand significantly their applicability of the former for complex practical problems. First, we formulate both approaches in the projective Hilbert space, thus extending the well known results from the Markov Breuer and Petruccione (1995) to the non-Markovian regime. Then a diffusive limit of the quantum jumps is taken in such a way that it coincides with quantum state diffusion in the projective Hilbert space and in the non-Markovian regime. Interestingly, the same limit in Hilbert space results in a completely new unraveling, which we call non-Markovian quantum diffusion (see Fig. 1). We enhance the quantum jumps and quantum diffusion approaches with kernel smoothing techniques Ghosh (2017), which allows us to handle driven dissipative systems with quantum memory effects easily. Lastly, we apply all of the methods to the driven dissipative two level atom.

Open systems and projective Hilbert space.—

A typical model for open systems in the interaction picture with respect to environment Hamiltonian HB=λωλaλaλH_{B}=\sum_{\lambda}\omega_{\lambda}a_{\lambda}^{\dagger}a_{\lambda} is

H(t)=\displaystyle H(t)= HS(t)+λgλLaλeiωλt+gλLaλeiωλt,\displaystyle H_{S}(t)+\sum_{\lambda}g_{\lambda}La_{\lambda}^{\dagger}e^{i\omega_{\lambda}t}+g_{\lambda}^{*}L^{\dagger}a_{\lambda}e^{-i\omega_{\lambda}t}, (1)

where the creation- and annihilation operators aλa_{\lambda}^{\dagger} and aλa_{\lambda} of a bath mode labeled by λ\lambda satisfy the bosonic commutation relations [aλ,aμ]=δλμ[a_{\lambda},a_{\mu}^{\dagger}]=\delta_{\lambda\mu}. We assume that the coupling operator LL is traceless, i.e. tr{L}=0\text{tr}\left\{L\right\}{}=0.

NMQJNMQDNMQSDLMEKME2𝒫()\mathcal{P}(\mathcal{H})𝒫()\mathcal{P}(\mathcal{H})ε0\varepsilon\to 0ε0\varepsilon\to 0
Figure 1: Relation between non-Markovian quantum jumps (NMQJ), non-Markovian quantum diffusion (NMQD) and non-Markovian quantum state diffusion (NMQSD). In 𝒫()\mathcal{P}(\mathcal{H}), NMQJ corresponds to a Liouville master equation (LME), whereas NMQSD is associated with a 2nd order Kramers-Moyal expansion (KME2) of the LME. In other words, the diffusive limit can be taken in such a way that the LME associated with NMQJ transforms to a KME2 associated with NMQSD. However, when the very same limit is taken in \mathcal{H}, it results to a completely new unraveling, which we call non-Markovian Quantum Diffusion (NMQD).

In a projective Hilbert space 𝒫()\mathcal{P}(\mathcal{H}), each point is associated with a projector |ψψ||\psi\rangle\!\langle\psi| Breuer et al. (2002); Chruscinski and Jamiolkowski (2004). Given a separable Hilbert space, coordinates ψi\psi_{i}\in\mathbb{C} on a 𝒫()\mathcal{P}(\mathcal{H}) can be easily constructed with respect to a fixed orthonormal basis as |ψ=iψi|i|\psi\rangle=\sum_{i}\psi_{i}|i\rangle. For more information on 𝒫()\mathcal{P}(\mathcal{H}), see the Supplementary Material Note (1).

Non-Markovian quantum state diffusion.—

Reduced system dynamics can be represented exactly for a large class of models, even beyond Eq. (1Link and Strunz (2017), with the following linear non-Markovian quantum state diffusion (NMQSD) equation

t|ψ(t,z)=\displaystyle\partial_{t}|\psi(t,z^{*})\rangle= [iHS(t)+ztL]|ψ(t,z)\displaystyle\left[-iH_{S}(t)+z_{t}^{*}L\right]|\psi(t,z^{*})\rangle
L0tdsα(ts)δ|ψ(t,z)δzs.\displaystyle-L^{\dagger}\int\limits_{0}^{t}\textrm{d}s\,\alpha(t-s)\frac{\delta|\psi(t,z^{*})\rangle}{\delta z_{s}^{*}}. (2)

Here, LL is the coupling operator between the system and the bath and HS(t)H_{S}(t) is an arbitrary Hamiltonian acting on the open system Diósi et al. (1998); Strunz et al. (1999). NMQSD is driven by a complex valued colored Gaussian noise ztz_{t}^{*}, completely characterized by the correlations

𝖬[ztzs]=α(ts),𝖬[zt]=𝖬[ztzs]=0,\displaystyle\mathsf{M}\left[z_{t}z_{s}^{*}\right]=\alpha(t-s),\qquad\mathsf{M}\left[z_{t}^{*}\right]=\mathsf{M}\left[z_{t}z_{s}\right]=0, (3)

where 𝖬[]\mathsf{M}\left[\cdot\right] is the average over the noise process ztz_{t}^{*}. Solutions |ψ(t,z)|\psi(t,z^{*})\rangle are analytic functionals of the whole noise process ztz_{t}^{*} up to time tt.

In the remainder of this Letter, we will make the following restriction. We assume that the functional derivative satisfies, at least approximately Yu et al. (1999)

δδzs|ψ(z,t)=f(t,s)L|ψ(z,t).\displaystyle\frac{\delta}{\delta z_{s}^{*}}|\psi(z^{*},t)\rangle=f(t,s)L|\psi(z^{*},t)\rangle. (4)

Eq. (4) guarantees that the mean state will evolve according to a closed form master equation. However, the NMQSD method itself works perfectly well even if no such master equation exist for the reduced state.

The above stochastic Schrödinger equation (Non-Markovian quantum state diffusion.—) satisfies the ordinary rules of calculus since the noise process has a finite correlation time. The dynamics of the average state ρ(t)=𝖬[|ψ(z,t)ψ(z,t)|]\rho(t)=\mathsf{M}\left[|\psi(z^{*},t)\rangle\!\langle\psi(z^{*},t)|\right] described by Eq. (Non-Markovian quantum state diffusion.—) with assumption (4) reads

ρ˙(t)\displaystyle\dot{\rho}(t) =i[HS(t)+S(t)LL,ρ(t)]+2γ(t)Lρ(t)L\displaystyle=-i[H_{S}(t)+S(t)L^{\dagger}L,\rho(t)]+2\gamma(t)L\rho(t)L^{\dagger}
γ(t){LL,ρ(t)},\displaystyle-\gamma(t)\left\{L^{\dagger}L,\rho(t)\right\}, (5)

where F(t)=γ(t)+iS(t)F(t)=\gamma(t)+iS(t) and F(t)=0tdsα(ts)f(t,s)F(t)=\int_{0}^{t}\textrm{d}s\,\alpha(t-s)f(t,s).

To look for a connection between NMQSD and non-Markovian quantum jumps, we first have to derive a representation of the former in the projective Hilbert space. The probability density functional can be expressed as

PQ[ψ,t]\displaystyle P_{Q}[\psi,t] =𝖬[δ(ψψ(z,t))].\displaystyle=\mathsf{M}\left[\delta(\psi-\psi(z^{*},t))\right]. (6)

We show in Sec. LABEL:sec:deriv-fokk-planck of 111Supplementary Material, that the probability density functional satisfies the following second order partial differential equation

tPQ[ψ,t]=\displaystyle\partial_{t}P_{Q}[\psi,t]= k=1dψkck(ψ)PQ[ψ,t]+ψkck(ψ)PQ[ψ,t]\displaystyle\sum_{k=1}^{d}\partial_{\psi_{k}}c_{k}(\psi)P_{Q}[\psi,t]+\partial_{\psi_{k}^{*}}c_{k}^{*}(\psi)P_{Q}[\psi,t]
+k,l=1dψkψl2dkl(ψ)PQ[ψ,t],\displaystyle+\sum_{k,l=1}^{d}\partial^{2}_{\psi_{k}\psi_{l}^{*}}d_{kl}(\psi)P_{Q}[\psi,t], (7)

where the drift and diffusion coefficients are ck(ψ)=k|(iHF(t)LL)|ψc_{k}(\psi)=\langle k|\left(-iH-F(t)L^{\dagger}L\right)|\psi\rangle and dkl(ψ)=(F(t)+F(t))k|L|ψψ|L|ld_{kl}(\psi)=\left(F(t)+F^{*}(t)\right)\langle k|L|\psi\rangle\langle\psi|L^{\dagger}|l\rangle, respectively. NMQSD thus corresponds to a 22nd order Kramers - Moyal expansion in 𝒫()\mathcal{P}(\mathcal{H}) Risken (2012). If the diffusion coefficient F(t)+F(t)=2γ(t)F(t)+F^{*}(t)=2\gamma(t) is not negative for any time tt the KME2 equation is, in fact, a proper Fokker-Planck equation Gardiner (2009).

Non-Markovian Quantum Jumps.—

Master equations of the form

ρ˙(t)=\displaystyle\dot{\rho}(t)= i[HS(t)+ksk(t)LkLk,ρ]+k2γk(t)LkρLk\displaystyle-i[H_{S}(t)+\sum_{k}s_{k}(t)L_{k}^{\dagger}L_{k},\rho]+\sum_{k}2\gamma_{k}(t)L_{k}\rho L_{k}^{\dagger}
γk(t){LkLk,ρ(t)},\displaystyle-\gamma_{k}(t)\left\{L_{k}^{\dagger}L_{k},\rho(t)\right\}, (8)

can be unravelled with non-Markovian quantum jumps (NMQJ) Piilo et al. (2008, 2009); Härkönen (2010), 222We demand that at time t=0t=0 all of the decay rates γk(t)\gamma_{k}(t) have to be non-negative. It is a piecewise deterministic process in the Hilbert space of the open system. Here we present a linear version of the process (LNMQJ) given by the following Ito stochastic differential equation

|dψ=\displaystyle|\textrm{d}\psi\rangle= iG(t)|ψdt+k(Lk𝟙)|ψdN+k(t)\displaystyle-iG(t)|\psi\rangle\textrm{d}t+\sum_{k}(L_{k}-\mathbbm{1})|\psi\rangle\textrm{d}N_{+}^{k}(t)
+dψ(|ψ|ψ)dN,ψk(t),\displaystyle+\int\textrm{d}\psi^{\prime}\,\left(|\psi^{\prime}\rangle-|\psi\rangle\right)\textrm{d}N_{-,\psi^{\prime}}^{k}(t), (9)

where G(t)=HS(t)+ksk(t)LkLkiγk(t)[LkLk𝟙]G(t)=H_{S}(t)+\sum_{k}s_{k}(t)L_{k}^{\dagger}L_{k}-i\gamma_{k}(t)[L_{k}^{\dagger}L_{k}-\mathbbm{1}]. Increments of the Poisson processes, dN+k(t)\textrm{d}N_{+}^{k}(t) and dN,ψl(t),\textrm{d}N_{-,\psi^{\prime}}^{l}(t), are mutually independent dN+k(t)dN+l(t)=δkldN+k(t)\textrm{d}N_{+}^{k}(t)\textrm{d}N_{+}^{l}(t)=\delta_{kl}\textrm{d}N_{+}^{k}(t), dN,ψk(t)dN,ψ′′l(t)=δklδ(ψψ′′)dN,ψ(t)\textrm{d}N_{-,\psi^{\prime}}^{k}(t)\textrm{d}N_{-,\psi^{\prime\prime}}^{l}(t)=\delta_{kl}\delta(\psi^{\prime}-\psi^{\prime\prime})\textrm{d}N_{-,\psi^{\prime}}(t) and dN+k(t)dN,ψl(t)=0\textrm{d}N_{+}^{k}(t)\textrm{d}N_{-,\psi^{\prime}}^{l}(t)=0. The mean values of the increments are 𝖤[dN+k(t)]=2γ+k(t)dt{\mathsf{E}}\left[\textrm{d}N_{+}^{k}(t)\right]=2\gamma_{+}^{k}(t)\textrm{d}t and 𝖤[dN,ψl(t)]=2γl(t)P[ψ,t]P[ψ,t]δ(ψLlψ)dt{\mathsf{E}}\left[\textrm{d}N_{-,\psi^{\prime}}^{l}(t)\right]=2\gamma_{-}^{l}(t)\frac{P[\psi^{\prime},t]}{P[\psi,t]}\delta(\psi-L_{l}\psi^{\prime})\textrm{d}t, where γk(t)=γk+(t)γk(t)\gamma_{k}(t)=\gamma_{k}^{+}(t)-\gamma_{k}^{-}(t). It is easy to see that the average evolution reproduces Eq. (Non-Markovian Quantum Jumps.—).

In NMQJ, the memory effects reside in the jump probability from a source state ψ\psi to a target state ψ\psi^{\prime} via channel kk when γk(t)<0\gamma_{k}(t)<0. In particular, a “reverse jump” can occur from ψ\psi to ψ\psi^{\prime} iff Lk|ψ=|ψL_{k}|\psi^{\prime}\rangle=|\psi\rangle. The probability of such jumps depends on the ratio P[ψ,t]/P[ψ,t]P[\psi^{\prime},t]/P[\psi,t]. In order to compute the jump probability, the knowledge of the whole ensemble is required Piilo et al. (2008). This poses a serious challenge since a state |ψψ||\psi\rangle\!\langle\psi| has measure zero in 𝒫()\mathcal{P}(\mathcal{H}). We describe later a method to overcome this.

Now, Eq. (Non-Markovian quantum state diffusion.—) is equivalent to Eq. (Non-Markovian Quantum Jumps.—) with 2m(1k2m)2m\,(1\leq k\leq 2m) time dependent rates and time independent jump operators defined as

sk(t)=\displaystyle s_{k}(t)= s(t)2m|ξk|2ε2,γk(t)=γ(t)2m|ξk|2ε2,\displaystyle\frac{s(t)}{2m|\xi_{k}|^{2}\varepsilon^{2}},\qquad\gamma_{k}(t)=\frac{\gamma(t)}{2m|\xi_{k}|^{2}\varepsilon^{2}},
Lk=\displaystyle L_{k}= 𝟙+εξkL,s.t.ξk+ξk+m=0,\displaystyle\mathbbm{1}+\varepsilon\xi_{k}L,\,\,{\rm s.t.}\,\xi_{k}+\xi_{k+m}=0, (10)

where ξk\xi_{k}\in\mathbb{C}, |ξk|=|ξ||\xi_{k}|=|\xi| and ε>0\varepsilon>0. The deterministic part G(t)G(t) of the quantum jump process in Eq. (Non-Markovian Quantum Jumps.—) transforms under (Non-Markovian Quantum Jumps.—) to G(t)=HS(t)+s(t)LLiγ(t)LL+Θ(t)𝟙G^{\prime}(t)=H_{S}(t)+s(t)L^{\dagger}L-i\gamma(t)L^{\dagger}L+\Theta(t)\mathbbm{1}. The last term Θ(t)=k=12ms(t)2m|ξk|2ε2\Theta(t)=\sum_{k=1}^{2m}\frac{s(t)}{2m|\xi_{k}|^{2}\varepsilon^{2}} is a global phase factor, which can be neglected. If εL<1||\varepsilon L||<1, then operators LlL_{l} are invertible 333Ll1=j=0(1ξlε)jLj=𝟙εξlL+𝒪(ε2){L_{l}}^{-1}=\sum_{j=0}^{\infty}(-1\xi_{l}\varepsilon)^{j}L^{j}=\mathbbm{1}-\varepsilon\xi_{l}L+\mathcal{O}(\varepsilon^{2}), when εL<1{||\varepsilon L||}<1.. In this case, the transformed statistics of the Poisson increments eventually become

𝖤[dN+k(t)]\displaystyle{\mathsf{E}}\left[\textrm{d}N_{+}^{k}(t)\right] =γ+(t)mε2|ξk|2dt,\displaystyle=\frac{\gamma_{+}(t)}{m\varepsilon^{2}|\xi_{k}|^{2}}\textrm{d}t,
𝖤[dN,ψl(t)]\displaystyle{\mathsf{E}}\left[\textrm{d}N_{-,\psi^{\prime}}^{l}(t)\right] =γ(t)mε2|ξk|2P[ψ,t]P[ψ,t]δ(Ll1ψψ)|detLl||detLl|dt.\displaystyle=\frac{\gamma_{-}(t)}{m\varepsilon^{2}|\xi_{k}|^{2}}\frac{P[\psi^{\prime},t]}{P[\psi,t]}\frac{\delta(L_{l}^{-1}\psi-\psi^{\prime})}{|\det L_{l}||\det L_{l}^{\dagger}|}\textrm{d}t. (11)

Remarkably, after the transformation the increment dN,ψl(t)\textrm{d}N_{-,\psi^{\prime}}^{l}(t) does not depend on the target state of the jump,|ψ|\psi^{\prime}\rangle, anymore. This arises because a reverse jump corresponds to a mapping |ψLl1|ψ=|ψ|\psi\rangle\mapsto L_{l}^{-1}|\psi\rangle=|\psi^{\prime}\rangle, i.e. the target state of the jump is given by the action of the inverse operator on the source state |ψ|\psi\rangle.

Therefore, we can write the transformed process as

|dψt=\displaystyle|\textrm{d}\psi_{t}\rangle= iG(t)|ψdt+k[(Lk𝟙)|ψtdM+k(t)\displaystyle-iG^{\prime}(t)|\psi\rangle\textrm{d}t+\sum_{k}\Bigg{[}(L_{k}-\mathbbm{1})|\psi_{t}\rangle\textrm{d}M_{+}^{k}(t)
+(Lk1𝟙)|ψtdMk(t)],\displaystyle+\left({L_{k}}^{-1}-\mathbbm{1}\right)|\psi_{t}\rangle\textrm{d}M_{-}^{k}(t)\Bigg{]}, (12)

with mutually independent Poisson increments dM±k\textrm{d}M_{\pm}^{k} with statistics 𝖤[dM+k(t)]=γ+(t)mε2|ξk|2dt{\mathsf{E}}\left[\textrm{d}M_{+}^{k}(t)\right]=\frac{\gamma_{+}(t)}{m\varepsilon^{2}|\xi_{k}|^{2}}\textrm{d}t and 𝖤[dMk(t)]=P[Lk1ψ,t]P[ψ,t]|detLk||detLk|γ(t)mε2|ξk|2dt{\mathsf{E}}\left[\textrm{d}M_{-}^{k}(t)\right]=\frac{P[{L_{k}}^{-1}\psi,t]}{P[\psi,t]|\det L_{k}||\det L_{k}^{\dagger}|}\frac{\gamma_{-}(t)}{m\varepsilon^{2}|\xi_{k}|^{2}}\textrm{d}t, which are just relabeled increments of Eq. (Non-Markovian Quantum Jumps.—). To assert that this equation is still valid, we compute the average evolution of |ψψ||\psi\rangle\!\langle\psi| which coincides with the master equation (Non-Markovian quantum state diffusion.—) (see Sec. LABEL:sec:average-evolution of the Note (1)).

It is worth stressing that when γk(t)<0\gamma_{k}(t)<0, the quantum jumps are given by the inverse jump operator Lk1L_{k}^{-1}. Contrary to the original approach in Piilo et al. (2008), the quantum jumps and reverse quantum jumps are exactly inverses of each other. The quantum memory effects are contained in the probability for these jumps which still depends on the ratio P[Lk1ψ,t]/P[ψ,t]P[L_{k}^{-1}\psi,t]/P[\psi,t].

LNMQJ in projective Hilbert space.—

In the projective Hilbert space LNMQJ corresponds to the following Liouville master equation Härkönen (2010)

tP[ψ,t]\displaystyle\partial_{t}P[\psi,t] =ikψk(k|G(t)|ψP[ψ,t])\displaystyle=i\sum_{k}\partial_{\psi_{k}}\left(\langle k|G^{\prime}(t)|\psi\rangle P[\psi,t]\right)
ψk(ψ|G(t)|kP[ψ,t])\displaystyle-\partial_{\psi_{k}^{*}}\left(\langle\psi|G^{\prime\dagger}(t)|k\rangle P[\psi,t]\right)
+dϕ(R[ψ|ϕ]P[ϕ,t]R[ϕ|ψ]P[ψ,t]),\displaystyle+\int\textrm{d}\phi\,\left(R[\psi|\phi]P[\phi,t]-R[\phi|\psi]P[\psi,t]\right), (13)

where the jump rates R[ϕ|ψ]R[\phi|\psi] are

R[ϕ|ψ]=\displaystyle R[\phi|\psi]= k=12mγ+(t)mε2|ξk|2δ(ϕLkψ)\displaystyle\sum_{k=1}^{2m}\frac{\gamma_{+}(t)}{m\varepsilon^{2}|\xi_{k}|^{2}}\delta(\phi-L_{k}\psi)
+γ(t)mε2|εk|2P[ϕ,t]P[ψ,t]δ(ψLkϕ).\displaystyle+\frac{\gamma_{-}(t)}{m\varepsilon^{2}|\varepsilon_{k}|^{2}}\frac{P[\phi,t]}{P[\psi,t]}\delta(\psi-L_{k}\phi). (14)

When comparing the drift terms in Fokker-Planck equation (Non-Markovian quantum state diffusion.—) and in the Liouville master equation (LNMQJ in projective Hilbert space.—), we see that they are equal. The jump part takes the form dϕ(R[ψ|ϕ]P[ϕ,t]R[ϕ|ψ]P[ψ,t])=k=12mγ(t)mε2|ξk|2Fk[ψ]2γ(t)ε2|ξ|2P[ψ,t]\int\textrm{d}\phi\,\left(R[\psi|\phi]P[\phi,t]-R[\phi|\psi]P[\psi,t]\right)=\sum_{k=1}^{2m}\frac{\gamma(t)}{m\varepsilon^{2}|\xi_{k}|^{2}}F_{k}[\psi]-\frac{2\gamma(t)}{\varepsilon^{2}|\xi|^{2}}P[\psi,t], where

Fk[ψ]=P[Lk1ψ,t]|detLk||detLk|.\displaystyle F_{k}[\psi]=\frac{P[L_{k}^{-1}\psi,t]}{|\det L_{k}||\det L_{k}^{\dagger}|}. (15)

After expanding Fk[ψ]F_{k}[\psi] to second order in ε\varepsilon and assuming m>2m>2 we find dϕ(R[ψ|ϕ]P[ϕ,t]R[ϕ|ψ]P[ψ,t])\int\textrm{d}\phi\,\left(R[\psi|\phi]P[\phi,t]-R[\phi|\psi]P[\psi,t]\right){\to}
k,l=0dψkψl2(2γ(t)k|L|ψψ|L|lP[ψ,t])\sum_{k,l=0}^{d}\partial^{2}_{\psi_{k}\psi_{l}^{*}}\left(2\gamma(t)\langle k|L|\psi\rangle\langle\psi|L^{\dagger}|l\rangle P[\psi,t]\right), while ϵ0\epsilon\to 0 444For the computation we use the results of  Note (1) to compute the determinant, to expand the probability density, to expand a rational polynomial and choose ξk=eiπm(k1)\xi_{k}=e^{i\frac{\pi}{m}(k-1)} with m2m\geq 2. We thus have proven the validity of the part LMEε0FPE\mathrm{LME}\xrightarrow{\varepsilon\to 0}\mathrm{FPE} of the diagram in Fig. 1.

Non-Markovian quantum diffusion.—

Next we take the above diffusion limit directly on the piecewise deterministic LNMQJ process in the Hilbert space. Full details can be found in the Supplement Note (1). First, Eq. (Non-Markovian Quantum Jumps.—) is expanded to first order in ε\varepsilon, resulting in

|dψt=\displaystyle|\textrm{d}\psi_{t}\rangle= iG(t)|ψdt+k[ξkL|ψtεdM+k(t)\displaystyle-iG^{\prime}(t)|\psi\rangle\textrm{d}t+\sum_{k}\Bigg{[}\xi_{k}L|\psi_{t}\rangle\varepsilon\textrm{d}M_{+}^{k}(t)
ξkL|ψtεdMk(t)]+𝒪(ε2).\displaystyle-\xi_{k}L|\psi_{t}\rangle\varepsilon\textrm{d}M_{-}^{k}(t)\Bigg{]}+\mathcal{O}(\varepsilon^{2}). (16)

We define new processes dV±k=εdM±kεE[dM±k]\textrm{d}V_{\pm}^{k}=\varepsilon\textrm{d}M_{\pm}^{k}-\varepsilon E\left[\textrm{d}M_{\pm}^{k}\right] Pellegrini and Petruccione (2009) and by using the Ito rules, we have 𝖤[dV±k]=0{\mathsf{E}}\left[\textrm{d}V_{\pm}^{k}\right]=0, 𝖤[dVkdV+l]=0{\mathsf{E}}\left[\textrm{d}V_{-}^{k}\textrm{d}V_{+}^{l}\right]=0 and 𝖤[dV+kdV+l]=δkl(εdV+k+γ+(t)m|ξk|2dt){\mathsf{E}}\left[\textrm{d}V_{+}^{k}\textrm{d}V_{+}^{l}\right]=\delta_{kl}(\varepsilon\textrm{d}V_{+}^{k}+\frac{\gamma_{+}(t)}{m|\xi_{k}|^{2}}\textrm{d}t). We then define limε0dV±k=dW±k\lim_{\varepsilon\to 0}\textrm{d}V_{\pm}^{k}=\textrm{d}W_{\pm}^{k}, where the increments dW±k\textrm{d}W_{\pm}^{k} satisfy the following Ito rules

𝖤[dW±k]=0,𝖤[dWkdW+l]=0,\displaystyle{\mathsf{E}}\left[\textrm{d}W_{\pm}^{k}\right]=0,\qquad{\mathsf{E}}\left[\textrm{d}W_{-}^{k}\textrm{d}W_{+}^{l}\right]=0,
𝖤[dW+kdW+l]=δklγ+(t)m|ξk|2dt,\displaystyle{\mathsf{E}}\left[\textrm{d}W_{+}^{k}\textrm{d}W_{+}^{l}\right]=\delta_{kl}\frac{\gamma_{+}(t)}{m|\xi_{k}|^{2}}\textrm{d}t, (17)
𝖤[dWkdWl]=δklγ(t)m|ξk|2dt.\displaystyle{\mathsf{E}}\left[\textrm{d}W_{-}^{k}\textrm{d}W_{-}^{l}\right]=\delta_{kl}\frac{\gamma_{-}(t)}{m|\xi_{k}|^{2}}\textrm{d}t.

The goal is now to express the stochastic Schrödinger equation (Non-Markovian quantum diffusion.—) in terms of Wiener increments dW±k\textrm{d}W_{\pm}^{k}. After some simplification steps (detailed in the Supplement Note (1)), we find in the limit ε0\varepsilon\to 0

|dψ=\displaystyle|\textrm{d}\psi\rangle= (iG(t)+2γ(t)n=0dψ|L|nlnP[ψ,t]ψnL)|ψdt\displaystyle\bigg{(}-iG^{\prime}(t)+2\gamma_{-}(t)\sum_{n=0}^{d}\langle\psi|L^{\dagger}|n\rangle\frac{\partial\ln P[\psi,t]}{\partial\psi_{n}^{*}}L\bigg{)}|\psi\rangle\textrm{d}t
+L|ψdZ+L|ψdZ,\displaystyle+L|\psi\rangle\textrm{d}Z_{+}-L|\psi\rangle\textrm{d}Z_{-}, (18)

where dZ±=kξkdW±k\textrm{d}Z_{\pm}=\sum_{k}\xi_{k}\textrm{d}W_{\pm}^{k}. The complex noise dZ±\textrm{d}Z_{\pm} satisfies

𝖤[dZ±]=0,𝖤[dZ±dZ±]=0,\displaystyle{\mathsf{E}}\left[\textrm{d}Z_{\pm}\right]=0,\quad{\mathsf{E}}\left[\textrm{d}Z_{\pm}\textrm{d}Z_{\pm}\right]=0,
𝖤[dZ±dZ±]=2γ±(t)dt.\displaystyle{\mathsf{E}}\left[\textrm{d}Z_{\pm}\textrm{d}Z_{\pm}^{*}\right]=2\gamma_{\pm}(t)\textrm{d}t. (19)

The average evolution of NMQD equation (Non-Markovian quantum diffusion.—) corresponds to Eq. (Non-Markovian Quantum Jumps.—) as we show in the Supplement Note (1).

Interestingly, both noises dZ±\textrm{d}Z_{\pm} couple to the system via LL but with a different phase. Nevertheless, both noise terms produce “sandwich” terms 2γ±(t)LρLdt2\gamma_{\pm}(t)L\rho L^{\dagger}\textrm{d}t on the average evolution. The drift term with logarithmic derivative compensates the term 2γ(t)LρLdt2\gamma_{-}(t)L\rho L^{\dagger}\textrm{d}t on average such that the correct sandwich term 2(γ+(t)γ(t))LρLdt2(\gamma_{+}(t)-\gamma_{-}(t))L\rho L^{\dagger}\textrm{d}t emerges. The term proportional to the logarithmic derivative can be seen as the change in the stochastic entropy of the system which contributes to the deterministic evolution Seifert (2005).

Kernel smoothing.—

A Gaussian kernel KK is defined

K[ψ]=1πd+1eψ2,|ψd+1.\displaystyle K[\psi]=\frac{1}{\pi^{d+1}}e^{-||\psi||^{2}},|\psi\rangle\in\mathbb{C}^{d+1}. (20)

Given an ensemble of stochastic states {|ψν}ν=1M\{|\psi^{\nu}\rangle\}_{\nu=1}^{M}, we estimate the probability density P[ψ]P[\psi] in the projective Hilbert space with

Pσ[ψ]=1M(σ2πd+1)ν=1MK[(ψψν)/σ],\displaystyle P_{\sigma}[\psi]=\frac{1}{M(\sigma^{2}\pi^{d+1})}\sum_{\nu=1}^{M}K[(\psi-\psi^{\nu})/\sigma], (21)

where σ>0\sigma>0 is a free parameter. A rule of thumb for choosing the variance is that σ=M1d+5\sigma=M^{\frac{-1}{d+5}} Ghosh (2017), where dd is the real dimension of the underlying Hilbert space. Using the estimated density, we can compute the logarithmic derivative of the density appearing in Eq. (Non-Markovian quantum diffusion.—) as

lnPσ[ψ]ψn=ν=1Meψnψnν2ψnψnνσ2ν=1Meψnψnν2=ψnψnσ2,\displaystyle\frac{\partial\ln P_{\sigma}[\psi]}{\partial\psi_{n}^{*}}=-\frac{\sum_{\nu=1}^{M}e^{-||\psi_{n}-\psi_{n}^{\nu}||^{2}}\frac{\psi_{n}-\psi_{n}^{\nu}}{\sigma^{2}}}{\sum_{\nu=1}^{M}e^{-||\psi_{n}-\psi_{n}^{\nu}||^{2}}}=-\frac{\psi_{n}-\langle\langle\psi_{n}\rangle\rangle}{\sigma^{2}}, (22)

where average \langle\langle\cdot\rangle\rangle is taken with respect to distribution pσ=1𝒵eψψν2σ2p_{\sigma}=\frac{1}{\mathcal{Z}}e^{-\frac{||\psi-\psi^{\nu}||^{2}}{\sigma^{2}}}, with 𝒵=ν=1Meψnψnν2\mathcal{Z}=\sum_{\nu=1}^{M}e^{-||\psi_{n}-\psi_{n}^{\nu}||^{2}}. Kernel estimation can be also used to evaluate the ratios Pσ[ψ]Pσ[ψ]=ν=1MK[(ψψν)/σ]ν=1MK[(ψψν)/σ]\frac{P_{\sigma}[\psi^{\prime}]}{P_{\sigma}[\psi]}=\frac{\sum_{\nu^{\prime}=1}^{M}K[(\psi^{\prime}-\psi^{\nu^{\prime}})/\sigma]}{\sum_{\nu=1}^{M}K[(\psi-\psi^{\nu})/\sigma]}. Therefore, after performing the transformation (Non-Markovian Quantum Jumps.—) on the NMQJ and using the smoothed estimate for P[ψ,t]P[\psi,t] we can compute the reverse jump probabilities easily. The reason for this simplification is that the target state of the jump is directly given by the inverse jump operator and the ratio of probabilities for the target and the source state to occur can be efficiently evaluated from the estimate.

Example: Driven dissipative two level atom.—

An open system with HS=ω2σz+Ω2σxH_{S}=\frac{\omega}{2}\sigma_{z}+\frac{\Omega}{2}\sigma_{x} and L=σL=\sigma_{-} corresponds to an amplitude damped two level atom with driving and is not solvable in closed form. We assume that the bath correlation function takes the following exponential form

α(t,s)=gΓ2eiωc(ts)Γ|ts|,\displaystyle\alpha(t,s)=g\frac{\Gamma}{2}e^{-i\omega_{c}(t-s)-\Gamma|t-s|}, (23)

where Γ\Gamma is the inverse of the bath correlation time τc=Γ1\tau_{c}=\Gamma^{-1}, ωc\omega_{c} is the bath resonance frequency and g>0g>0 is a dimensionless parameter describing the overall system bath coupling strength. The limit Γ\Gamma\to\infty leads to a singular bath correlation function α(t,s)gδ(ts)\alpha(t,s)\to g\delta(t-s) and to a Gorini-Kossakowski-Sudarshan-Lindblad master equation with time independent decay rate gg Gorini et al. (1976); Lindblad (1976). The chosen correlation function can emerge from a microscopic model where the driven two level system is placed inside a leaky cavity near absolute zero temperature such that thermal excitations can be neglected. When the bath correlation time is short, Eq. (4) is approximately true Yu et al. (1999). Within this approximation the NMQSD equation takes the following form

t|ψt(z)=\displaystyle\partial_{t}|\psi_{t}(z^{*})\rangle= iHS|ψt(z)+ztσ|ψt(z)\displaystyle-iH_{S}|\psi_{t}(z^{*})\rangle+z_{t}^{*}\sigma_{-}|\psi_{t}(z^{*})\rangle
F(t)σ+σ|ψt(z),\displaystyle-F(t)\sigma_{+}\sigma_{-}|\psi_{t}(z^{*})\rangle, (24)

with α(t,s)=ztzs\alpha(t,s)=\langle z_{t}z_{s}^{*}\rangle being the only non-zero correlation of the complex noise. Then the average state obeys the following master equation

tρ=\displaystyle\partial_{t}\rho= i[ω2σz+Ω2σx+s(t)σ+σ,ρ]+2γ(t)σ+ρσ\displaystyle-i[\frac{\omega}{2}\sigma_{z}+\frac{\Omega}{2}\sigma_{x}+s(t)\sigma_{+}\sigma_{,}\rho]+2\gamma(t)\sigma_{+}\rho\sigma_{-}
γ(t){σ+σ,ρ},\displaystyle-\gamma(t)\{\sigma_{+}\sigma_{-},\rho\}, (25)

where γ(t)+is(t)=F(t)\gamma(t)+is(t)=F(t).

Refer to caption
Figure 2: Top: Ensemble average over 3000 stochastic trajectories of σx\langle\sigma_{x}\rangle computed with LNMQJ (dotted), NMQD (dashed) with ε=12\varepsilon=\frac{1}{2} and HOPS (dash dotted) with comparison to the master equation solution (ME). HOPS is a numerically exact method and the reasonable agreement shows that the approximations leading to the master equation being unraveled are fairly consistent with the chosen parameters. Bottom: Normalized expectation value for σz\sigma_{z} along a single stochastic trajectory for different values of ε\varepsilon using LNMQJ. The initial state is |+=12(|0+|1)|+\rangle=\sqrt{\frac{1}{2}}(|0\rangle+|1\rangle).

The LNMQJ unraveling (Example: Driven dissipative two level atom.—), in turn, is

d|ψ=\displaystyle\textrm{d}|\psi\rangle= iG(t)|ψdt+k=14εξkσ|ψdM+k(t)\displaystyle-iG(t)|\psi\rangle\textrm{d}t+\sum_{k=1}^{4}\varepsilon\xi_{k}\sigma_{-}|\psi\rangle\textrm{d}M_{+}^{k}(t)
k=14εξkσ|ψdMk(t),\displaystyle-\sum_{k=1}^{4}\varepsilon\xi_{k}\sigma_{-}|\psi\rangle\textrm{d}M_{-}^{k}(t), (26)

where ξ1=1\xi_{1}=1, ξ2=1\xi_{2}=-1, ξ3=i\xi_{3}=i, ξ4=i\xi_{4}=-i and G(t)=(HSiF(t))σ+σG(t)=(H_{S}-iF(t))\sigma_{+}\sigma_{-}. The statistics of the Poisson increments are 𝖤[dM+k]=γ+(t)2ε2dt{\mathsf{E}}\left[\textrm{d}M_{+}^{k}\right]=\frac{\gamma_{+}(t)}{2\varepsilon^{2}}\textrm{d}t and 𝖤[dMk]=P[(𝟙εξkσ)ψ,t]P[ψ,t]γ(t)2ε2dt{\mathsf{E}}\left[\textrm{d}M_{-}^{k}\right]=\frac{P[(\mathbbm{1}-\varepsilon\xi_{k}\sigma_{-})\psi,t]}{P[\psi,t]}\frac{\gamma_{-}(t)}{2\varepsilon^{2}}\textrm{d}t. Subsequently, the diffusive limit of LNMQJ process corresponding to the NMQD process for this system can be written as

d|ψ=\displaystyle\textrm{d}|\psi\rangle= (iG(t)+2γ(t)ψ|σ+|0lnP[ψ,t]ψ0σ)|ψdt\displaystyle\bigg{(}-iG(t)+2\gamma_{-}(t)\langle\psi|\sigma_{+}|0\rangle\frac{\partial\ln P[\psi,t]}{\partial\psi_{0}^{*}}\sigma_{-}\bigg{)}|\psi\rangle\textrm{d}t
+σ|ψ(dZ+dZ),\displaystyle+\sigma_{-}|\psi\rangle(\textrm{d}Z_{+}-\textrm{d}Z_{-}), (27)

where zero mean complex noises satisfy the Ito rules dZ±dZ±=γ±(t)dt\textrm{d}Z_{\pm}\textrm{d}Z_{\pm}^{*}=\gamma_{\pm}(t)\textrm{d}t and dZ±dZ±=dZdZ±=0\textrm{d}Z_{\pm}\textrm{d}Z_{\pm}=\textrm{d}Z_{\mp}\textrm{d}Z_{\pm}^{*}=0. We consider the following parameters in all of the numerical examples ω/Γ=2\omega/\Gamma=2, ωc/Γ=5.5\omega_{c}/\Gamma=5.5, Ω/Γ=0.5\Omega/\Gamma=0.5, g=0.8g=0.8 and we plot all dynamical quantities as a function of the dimensionless time Γt\Gamma t. The decay rate γ(t)\gamma(t) is temporarily negative when 12<Γt<3/2\frac{1}{2}<\Gamma t<3/2 for these parameter values. Figure 2 shows a good agreement between the master equation solution and its unravelings. However, we also solved the dynamics exactly using the HOPS approach to NMQSD Suess et al. (2014); Hartmann and Strunz (2017). The small disagreement shows that the approximations leading to the master equation (Example: Driven dissipative two level atom.—) are not fully consistent with the chosen parameters. Therefore, a word of caution is in place here; within the master equation approach, the quality of the obtained equation is extremely hard to assess555HOPS with hierarchy order 1 corresponds closely to the level of approximation we make when using- and unraveling the master quation with respect to exact dynamics. In Fig. 2 we have truncated the hierarchy after 8 levels.. In the bottom panel, we also show examples of single trajectories with LNMQJ for different values of ϵ\epsilon. The purpose is to demonstrate the agreement of the ensemble averages with the master equation solution – and that with our new methodological results, even driven systems can be very easily handled with the jump method whenever a reliable master equation is available.

Conclusions.—

We have provided a connection between quantum state diffusion and quantum jumps in the non-Markovian regime. As a by product of these investigations we introduced a linear version of the non-Markovian quantum jumps method and a new type of unraveling which we call non-Markovian quantum diffusion. We combined the non-Markovian quantum jumps and non-Markovian quantum diffusion with kernel smoothing techniques thus extending the applicability of these methods dramatically. Moreover, we also demonstrated the power of our approach with the paradigmatic amplitude damped driven two-level atom model. As an outlook, in addition of applying the methods for various state-of-the art complex driven open quantum systems, it would be interesting to investigate, e.g., what role the stochastic entropy term in non-Markovian quantum diffusion plays in quantum stochastic thermodynamics.

References