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

Generalized Adiabatic Theorems: Quantum Systems Driven by Modulated Time-Varying Fields

Amro Dodin Department of Chemistry, Massachusetts Institute of Technology, Cambridge, Massachusetts, 02139, USA adodin@mit.edu    Paul Brumer Chemical Physics Theory Group, Department of Chemistry and Center for Quantum Information and Quantum Control, University of Toronto, Toronto, Ontario, M5S 3H6, Canada Paul.Brumer@utoronto.ca
(August 7, 2025)
Abstract

We present generalized adiabatic theorems for closed and open quantum systems that can be applied to slow modulations of rapidly varying fields, such as oscillatory fields that occur in optical experiments and light induced processes. The generalized adiabatic theorems show that a sufficiently slow modulation conserves the dynamical modes of time dependent reference Hamiltonians. In the limiting case of modulations of static fields, the standard adiabatic theorems are recovered. Applying these results to periodic fields shows that they remain in Floquet states rather than in energy eigenstates. More generally, these adiabatic theorems can be applied to transformations of arbitrary time-dependent fields, by accounting for the rapidly varying part of the field through the dynamical normal modes, and treating the slow modulation adiabatically. As examples, we apply the generalized theorem to (a) predict the dynamics of a two level system driven by a frequency modulated resonant oscillation, a pathological situation beyond the applicability of earlier results, and (b) to show that open quantum systems driven by slowly turned-on incoherent light, such as biomolecules under natural illumination conditions, can only display coherences that survive in the steady state.

I Introduction

Quantum systems driven by slowly varying external fields play an important role in atomic, molecular and optical physics. The ubiquity of these processes has led to sustained interest in powerful adiabatic theorems (AT’s) that characterize their dynamics Born and Fock (1928); Kato (1950); Sarandy and Lidar (2005); Messiah (2017). The best known AT states that a system initialized in an energy eigenstate will remain in that state when driven by a slowly varying field. Adiabatic processes have found far reaching utility in quantum dynamics Landau (1932); Zener (1932) and in the development of Adiabatic Quantum Computing (AQC) where they are used to reliably realize quantum state transformations Albash and Lidar (2018); Barends et al. (2016). Moreover, adiabaticity conditions bound the allowable speed of state transformations, setting the time cost of AQC algorithms and spurring efforts to find shortcuts to adiabaticity Torrontegui et al. (2017); del Campo (2013); Torrontegui et al. (2013); Campo and Boshier (2012). However, ATs fail for oscillating fields Amin (2009); Du et al. (2008); Tong et al. (2005); Marzlin and Sanders (2004), since these fields can induce transitions between energy eigenstates, precisely the opposite of what is required in traditional adiabatic theorems. Hence traditional ATs can not, in general, be applied to light-induced processes in either isolated or open quantum systems.

Adiabatic dynamics have been successfully deployed in experimental settings through Adiabatic Rapid Passage (ARP)Malinovsky and Krause (2001) and Stimulated Raman Adiabatic Passage (STIRAP) Vitanov et al. (2017); Bergmann et al. (2015) techniques. In these experiments, the adiabatic theorem provides a robust mechanism for high fidelity state preparation that is not sensitive to slight perturbations in the system properties and has been used, for example, to study molecular reaction dynamics Kaufmann et al. (2001) and in the preparation of ultra-cold molecules Külz et al. (1996); Aikawa et al. (2010). However, the domain of these techniques has been limited by the applicability of the AT, limiting the forms of driving fields and transformation times that can be realized. By relaxing key restrictions imposed by the traditional AT, the AMT introduced below significantly broadens the range of experimental techniques, unlocking the potential for faster, more flexible state preparation and system dynamics.

Numerous scenarios do not admit traditional adiabatic approaches. Natural light-induced excitation of biological molecules (e.g., in photosynthesis or vision) provides a particularly important example of a challenge to standard ATs insofar as it involves incoherent excitation of open quantum systems with turn-on times that are exceedingly long compared to the time scale of molecular dynamics. For example, human blinking, the turn-on time of light in vision, occurs over 0.2 sec as compared to the timescale of molecular isomerization induced by incident light in vision which can, under pulsed laser excitation, occur in 50 fs or less. Thus, challenging questions related to the creation of molecular coherences in such natural processes requires an AT adapted to oscillating incoherent fields in open quantum systems Dodin and Brumer (2019); Dodin et al. (2016); Cheng and Fleming (2009); Chenu and Scholes (2015).

In this paper we derive generalized AT’s, termed Adiabatic Modulation Theorems (AMTs), for open and closed quantum systems that apply to oscillatory and other rapidly varying fields, extending adiabaticity conditions to optically driven processes and providing faster pathways to adiabaticity, with the potential to accelerate AQC. Moreover, we show that the adiabatic transformation time is limited not by the energy gaps in the systems but rather by the frequency differences of the instantaneous normal modes, that are often easier to manipulate. These results exploit the fact that many complicated fields of interest are time-dependent modulations (e.g., frequency and amplitude modulated oscillations) of simpler fields that have well understood dynamics.

The AMT’s derived below show that a system subjected to a sufficiently slow modulation of a time-dependent field conserves the dynamical modes of time-dependent reference Hamiltonians. These theorems contain traditional ATs in the limit of modulations of static fields, and generalize easily to modulations of periodic fields, which are shown to preserve Floquet states rather than energy eigenstates. Moreover, we show that the adiabatic transformation time is limited not by the energy gaps in the system but rather by the frequency differences of the instantaneous normal modes, that are often easier to manipulate. Remarkably, these constructions allow for the design of experimental techniques that go beyond the preparation of time-independent states and allow for the preparation of specific dynamics using the well developed intuition of adiabatic theory.

The AMT theorems are widely applicable and two significant examples are provided below. In the isolated system example, we show how the theorem allows for control of dynamics on the entire Bloch sphere in a two level qubit system. In applications to open systems, the theorem resolves a longstanding controversy regarding the role of light-induced coherent oscillations in biophysical processes. The generality of the theorems ensure applications to a wide variety of alternative systems.

This paper is organized as follows: Sect. II proves the adiabatic modulation theorems for both isolated and open systems. Sample applications, designed to demonstrate the theorems and their applications are provided in Sect. III.

II The Adiabatic Modulation Theorems

II.1 Isolated Systems

Consider a family of time-dependent reference Hamiltonians H^0(t;λ)\hat{H}_{0}(t;\lambda) indexed by the parameter λ\lambda (e.g. λ\lambda may be an amplitude H^0(t;λ)=λh^0(t)\hat{H}_{0}(t;\lambda)=\lambda\hat{h}_{0}(t) or frequency H^0(t;λ)=exp(iλt)h^0+h.c.\hat{H}_{0}(t;\lambda)=\exp(-i\lambda t)\hat{h}_{0}+h.c.). For each value of λ\lambda, setting =1\hbar=1, we define the instantaneous normal modes as

|ψn(t;λ)eiθn(t;λ)|n(t;λ),\ket{\psi_{n}(t;\lambda)}\equiv e^{-i\theta_{n}(t;\lambda)}\ket{n(t;\lambda)}, (1)

where θn(t;λ)0t𝑑sϵn(s;λ)\theta_{n}(t;\lambda)\equiv\int^{t}_{0}ds\epsilon_{n}(s;\lambda) is the dynamical phase, ϵn(t;λ)\epsilon_{n}(t;\lambda) is an instantaneous quasienergy and the |n(t;λ)\ket{n(t;\lambda)} is the instantaneous normal mode of H^0(t;λ)\hat{H}_{0}(t;\lambda).

These normal modes and corresponding quasienergies are found by solving the eigenproblem

H^F(t;λ)|n(t;λ)=ϵn(t;λ)|n(t;λ),\hat{H}_{F}(t;\lambda)\ket{n(t;\lambda)}=\epsilon_{n}(t;\lambda)\ket{n(t;\lambda)}, (2)

where H^F(t;λ)H^0(t;λ)i/t\hat{H}_{F}(t;\lambda)\equiv\hat{H}_{0}(t;\lambda)-i\partial/\partial t. That is, the normal modes are particular solutions to the Time-Dependent Schrödinger Equation (TDSE),

it|Ψ(t)=H^(t)|Ψ(t),i\frac{\partial}{\partial t}\ket{\Psi(t)}=\hat{H}(t)\ket{\Psi(t)}, (3)

and are well understood for many driving fields. For example, for time-independent Hamiltonians, the normal modes of static Hamiltonians are energy eigenstates, while for periodic Hamiltonians they are Floquet states.

A modulation of a time-dependent Hamiltonian, H^0(t;λ)\hat{H}_{0}(t;\lambda), is a transformation that varies λλt\lambda\to\lambda_{t} over a time interval [0,τ][0,\tau] through a modulation protocol Λ{λt|t[0,τ]}\Lambda\equiv\{\lambda_{t}|t\in[0,\tau]\}. The resultant modulated field H~^Λ(t)H^0(t;λt)\hat{\widetilde{H}}_{\Lambda}(t)\equiv\hat{H}_{0}(t;\lambda_{t}) sweeps through the Hamiltonian family defined above (e.g. a field with a time-dependent amplitude for H^0=λh0(t)\hat{H}_{0}=\lambda h_{0}(t)). Note, as an aside, that this is an operator generalization of modulations in signal processing and encoding theory Byrne (1993). Generally, the modulated Hamiltonian can be significantly more complicated than H0H_{0} since the modulation function may be highly nonlinear and the modulation parameter can vary non-trivially with time.

The dynamics induced by H~^Λ(t)\hat{\widetilde{H}}_{\Lambda}(t) can now be expressed in terms of instantaneous normal modes of H^0\hat{H}_{0}, Eq. (1), to give

|Ψ(t)ncn(t)eiθn(t;Λ)|n(t;λt)\ket{\Psi(t)}\equiv\sum_{n}c_{n}(t)e^{-i\theta_{n}(t;\Lambda)}\ket{n(t;\lambda_{t})} (4)

where θn(t;Λ)0t𝑑sϵn(s;λs)/\theta_{n}(t;\Lambda)\equiv\int^{t}_{0}ds\epsilon_{n}(s;\lambda_{s})/\hbar. The instantaneous state of the system can always be expressed in this form since H^F(t;λt)\hat{H}_{F}(t;\lambda_{t}) is always Hermitian and therefore has a basis of eigenstates {|n(t;Λt)}\{\ket{n(t;\Lambda_{t})}\} that can be used to expand |Ψ(t)\ket{\Psi(t)} with coefficient an(t)n(t;λt)|Ψ(t)=cn(t)exp(iθn(t))a_{n}(t)\equiv\braket{n(t;\lambda_{t})|\Psi(t)}=c_{n}(t)\exp(-i\theta_{n}(t)) 111The modified Hamiltonian is defined using a partial time derivative and therefore neglects the implicit time-dependence of λt\lambda_{t}..

This approach mirrors that used to derive traditional ATs, but replacing the eigenstates with normal modes and energies with quasienergies. We therefore proceed through a similar path to derive adiabaticity conditions under which a system initialized in a normal mode of H^0(0;λ0)\hat{H}_{0}(0;\lambda_{0}) will remain in that normal mode at all times. Mathematically, this occurs when cn(t)c_{n}(t) are decoupled in the TDSE, Eq. (3). Substituting Eq. (4) into Eq. (3) we find

n(c˙n(t)|n(t;λt)+cn(t)λ˙tλt|n(t;λt))eiθn(t;Λ)=0,\sum_{n}\left(\dot{c}_{n}(t)\ket{n(t;\lambda_{t})}+c_{n}(t)\dot{\lambda}_{t}\frac{\partial}{\partial\lambda_{t}}\ket{n(t;\lambda_{t})}\right)e^{-i\theta_{n}(t;\Lambda)}=0, (5)

where we have used the identity dθn(t;Λ)/dt=ϵn(t;λt)d\theta_{n}(t;\Lambda)/dt=\epsilon_{n}(t;\lambda_{t}) and Eq. (2) to cancel terms proportional to ϵn(t;λt)\epsilon_{n}(t;\lambda_{t}). Projecting onto an |m(t;λt)\ket{m(t;\lambda_{t})}, yields equations of motion for the coefficients,

c˙m(t)=cm(t)λ˙tm|λt|n+nmλ˙teiθnmm|H~^λ|nϵmϵn,\dot{c}_{m}(t)=-c_{m}(t)\dot{\lambda}_{t}\bra{m}\frac{\partial}{\partial\lambda_{t}}\ket{n}+\sum_{n\neq m}\frac{\dot{\lambda}_{t}e^{-i\theta_{nm}}\braket{m|\frac{\partial\hat{\widetilde{H}}}{\partial\lambda}|n}}{\epsilon_{m}-\epsilon_{n}}, (6)

where we have suppressed the tt and λt\lambda_{t} dependence for brevity, and θnmθnθm\theta_{nm}\equiv\theta_{n}-\theta_{m}. To obtain Eq. (6), Eq. (2) was differentiated with respect to λt\lambda_{t} to obtain m|/λt|n=m|H~^/λt|n/(ϵnϵm)\braket{m|\partial/\partial\lambda_{t}|n}=\braket{m|\partial\hat{\widetilde{H}}/\partial\lambda_{t}|n}/(\epsilon_{n}-\epsilon_{m}).

Equation (6) takes the same form as the original adiabatic theorem with cross-coupling between normal modes scaled by a rapidly oscillating term Sarandy and Lidar (2005). As a result normal modes evolve independently of one another when the following condition is satisfied:

max0tτ|λ˙tm|H^(t;λt)λt|nϵnm(t;λt)|min0tτ|ϵnm(t;λt)|\max_{0\leq t\leq\tau}\left|\frac{\dot{\lambda}_{t}\braket{m|\frac{\partial\hat{H}(t;\lambda_{t})}{\partial\lambda_{t}}|n}}{\epsilon_{nm}(t;\lambda_{t})}\right|\ll\min_{0\leq t\leq\tau}\left|\epsilon_{nm}(t;\lambda_{t})\right| (7)

where ϵnm(t;λt)ϵn(t;λt)ϵm(t;λt)\epsilon_{nm}(t;\lambda_{t})\equiv\epsilon_{n}(t;\lambda_{t})-\epsilon_{m}(t;\lambda_{t}). Consequently, if λt\lambda_{t} changes much more slowly than the difference in dynamical mode frequencies, the slow modulation will not cross couple the normal modes of the modulated Hamiltonian.

Equation (7) has several important features. First, if H^0(t;λ)\hat{H}_{0}(t;\lambda) are tt-independent, then their normal modes are energy eigenstates and all Hamiltonian time-dependence is contained in the modulation. In this case, the AMT reduces to the traditional AT, simply relabeling the time variable as λ\lambda. The AMT then extends previous AT’s by recognizing them as statements about how time-dependent transformations of Hamiltonian fields impact dynamical modes of the TDSE. These coincide with eigenstates for static Hamiltonians but the same insight can be applied to any field with well-understood dynamics. For example, if H^0(t;λ)\hat{H}_{0}(t;\lambda) are periodic, Eq. (7) states that slow modulations do not couple Floquet states. One key distinction of the AMT is that its speed limits are set by the differences in normal mode frequencies which are often easier to manipulate than energy gaps between eigenstates, allowing for faster adiabatic transformations. A numerical example is discussed in Sect. III.1.1.

II.2 Open Systems

Consider now an adiabatic theorem for open quantum systems. To prove this theorem we take an approach inspired by Ref. Sarandy and Lidar (2005). The dynamics of an open quantum system are governed by the Liouville von-Neumann (LvN) equation

ddt|ρ=^^(t)|ρ\frac{d}{dt}|\rho\rangle\rangle=\hat{\hat{\mathcal{L}}}(t)|\rho\rangle\rangle (8)

in the time-convolutionless form Breuer and Petruccione (2007); Alicki and Lendi (2007); Blum (2012). The double-ket notation ||\cdot\rangle\rangle is indicates an operator Liouville space with a trace inner product A|BTr{A^B^}\langle\langle A|B\rangle\rangle\equiv\operatorname{Tr}\{\hat{A}^{\dagger}\hat{B}\}. By analogy with the isolated system we define modulations by starting with a family of time-dependent Liouvillians ^^0(t;λ)\hat{\hat{\mathcal{L}}}_{0}(t;\lambda) and allowing the modulation parameter to vary over time to give ^^Λ(t)^^0(t;λt)\hat{\hat{\mathcal{L}}}_{\Lambda}(t)\equiv\hat{\hat{\mathcal{L}}}_{0}(t;\lambda_{t}). A modified Liouvillian can then be defined by ^^F(t;λ)^^(t;λ)/t\hat{\hat{\mathcal{L}}}_{F}(t;\lambda)\equiv\hat{\hat{\mathcal{L}}}(t;\lambda)-\partial/\partial t.

The LvN equation appears similar to the TDSE [Eq. (3)], suggesting that we may be able to apply the same analysis, but replacing Hamiltonians with Liouvillians. However, the Liouvillian and it’s corresponding modified operator are completely positive but not necessarily Hermitian Alicki and Lendi (2007) and therefore require more care in defining their normal modes. A given M×MM\times M modified Liouvillian, ^^F\hat{\hat{\mathcal{L}}}_{F}, has an incomplete set of NMN\leq M left and right eigenvectors 𝒬α(t)|\langle\langle\mathcal{Q}_{\alpha}(t)| and |𝒫α(t)|\mathcal{P}_{\alpha}(t)\rangle\rangle. Associated with each eigenvector is a Jordan Block comprised of nαn_{\alpha} generalized eigenvectors that combine to give an orthonormal basis of Liouville space defined by the generalized eigenproblem

𝒬α(j)|^^F=𝒬α(j)|χα+𝒬α(j+1)|\langle\langle\mathcal{Q}_{\alpha}^{(j)}|\hat{\hat{\mathcal{L}}}_{F}=\langle\langle\mathcal{Q}_{\alpha}^{(j)}|\chi_{\alpha}+\langle\langle\mathcal{Q}_{\alpha}^{(j+1)}| (9a)
^^F|𝒫α(j)=χα|𝒫α(j)+|𝒫α(j1)\hat{\hat{\mathcal{L}}}_{F}|\mathcal{P}_{\alpha}^{(j)}\rangle\rangle=\chi_{\alpha}|\mathcal{P}_{\alpha}^{(j)}\rangle\rangle+|\mathcal{P}_{\alpha}^{(j-1)}\rangle\rangle (9b)

where 𝒬α(nα1)|𝒬α|\langle\langle\mathcal{Q}_{\alpha}^{(n_{\alpha}-1)}|\equiv\langle\langle\mathcal{Q}_{\alpha}|,𝒬α(nα)|0|\langle\langle\mathcal{Q}_{\alpha}^{(n_{\alpha})}|\equiv\langle\langle 0|, |𝒫α(0)|𝒫α|\mathcal{P}_{\alpha}^{(0)}\rangle\rangle\equiv|\mathcal{P}_{\alpha}\rangle\rangle, and |𝒫α(1)|0|\mathcal{P}_{\alpha}^{(-1)}\rangle\rangle\equiv|0\rangle\rangle, where |0|0\rangle\rangle is the zero operator. Dynamically, the LvN Equation does not cross couple the Jordan blocks to one another, but can lead to cross coupling within states of the same Jordan block, thereby defining decoupled dynamical subspace. (For a concise summary of Jordan canonical form in the context of quantum adiabatic theorems, see ref. Sarandy and Lidar (2005).)

Motivated by the isolated systems derivation, we consider the modified Liouvillian and aim to expand the dynamics of the open system in terms of its instantaneous generalized eigenstates. These are given by the (right) generalized equation

^^F(t;λt)|𝒫α(j)=χα|𝒫α(j)+|𝒫α(j1).\hat{\hat{\mathcal{L}}}_{F}(t;\lambda_{t})|\mathcal{P}_{\alpha}^{(j)}\rangle\rangle=\chi_{\alpha}|\mathcal{P}_{\alpha}^{(j)}\rangle\rangle+|\mathcal{P}_{\alpha}^{(j-1)}\rangle\rangle. (10)

The left generalized eigenstates can be similarly defined as the right eigenstates of ^^F\hat{\hat{\mathcal{L}}}_{F}^{\dagger}.

Allowing the modulation parameter to change with time, the instantaneous state of the system can be expanded as a superposition over the generalized eigenstates in the form

|ρ(t)β=1Nj=0nβ1rβ(j)(t)|𝒫β(j)(t;λt)|\rho(t)\rangle\rangle\equiv\sum_{\beta=1}^{N}\sum_{j=0}^{n_{\beta}-1}r_{\beta}^{(j)}(t)|\mathcal{P}_{\beta}^{(j)}(t;\lambda_{t})\rangle\rangle (11)

where rβ(j)(t)𝒬β(j)(t;λt)|ρ(t)r_{\beta}^{(j)}(t)\equiv\langle\langle\mathcal{Q}_{\beta}^{(j)}(t;\lambda_{t})|\rho(t)\rangle\rangle is a complex valued expansion coefficient. Similarly, to Eq. (4), any density operator |ρ|\rho\rangle\rangle can be expressed in this form since the generalized eigenbasis that generates Jordan canonical form is complete and orthonormal.

Substituting the trial form of Eq. (11) into the time convolutionless Liouville-von Neumann equation [Eq. (8)] and projecting onto the left generalized eigenstate 𝒬α(i)|\langle\langle\mathcal{Q}_{\alpha}^{(i)}|, gives the following equation of motion for the expansion coefficients:

r˙α(i)=χαrα(i)+rα(i+1)β=1Nj=0nβ1λ˙trβ(j)𝒬α(i)|λt|𝒫β(j)\dot{r}_{\alpha}^{(i)}=\chi_{\alpha}r_{\alpha}^{(i)}+r_{\alpha}^{(i+1)}-\sum_{\beta=1}^{N}\sum_{j=0}^{n_{\beta}-1}\dot{\lambda}_{t}r_{\beta}^{(j)}\langle\langle\mathcal{Q}_{\alpha}^{(i)}|\frac{\partial}{\partial\lambda_{t}}|\mathcal{P}_{\beta}^{(j)}\rangle\rangle (12)

where we have used the Floquet Equation (10) and the orthonormality condition of generalized eigenstates. By convention, we take rα(nα)=0r_{\alpha}^{(n_{\alpha})}=0.

Differentiating Eq. (10) for some right eigenstate |𝒫β(j)|\mathcal{P}_{\beta}^{(j)}\rangle\rangle with respect to λt\lambda_{t} at fixed tt and projecting onto left eigenstate 𝒬α(i)|\langle\langle\mathcal{Q}_{\alpha}^{(i)}| with αβ\alpha\neq\beta gives

𝒬α(i)|^^λt|𝒫β(j)+𝒬α(i)|^^λt|𝒫β(j)=χβλt𝒬α(i)|𝒫β(j)+χβ𝒬α(i)|𝒫˙β(j)+𝒬α(i)|𝒫˙β(j1).\langle\langle\mathcal{Q}_{\alpha}^{(i)}|\frac{\partial\hat{\hat{\mathcal{L}}}}{\partial\lambda_{t}}|\mathcal{P}_{\beta}^{(j)}\rangle\rangle+\langle\langle\mathcal{Q}_{\alpha}^{(i)}|\hat{\hat{\mathcal{L}}}\frac{\partial}{\partial\lambda_{t}}|\mathcal{P}_{\beta}^{(j)}\rangle\rangle=\frac{\partial\chi_{\beta}}{\partial\lambda_{t}}\langle\langle\mathcal{Q}_{\alpha}^{(i)}|\mathcal{P}_{\beta}^{(j)}\rangle\rangle+\chi_{\beta}\langle\langle\mathcal{Q}_{\alpha}^{(i)}|\dot{\mathcal{P}}_{\beta}^{(j)}\rangle\rangle+\langle\langle\mathcal{Q}_{\alpha}^{(i)}|\dot{\mathcal{P}}_{\beta}^{(j-1)}\rangle\rangle. (13)

This expression can be simplified by first noting that for αβ\alpha\neq\beta orthonormality of the basis eliminates the 𝒬α(i)|𝒫β(j)\langle\langle\mathcal{Q}_{\alpha}^{(i)}|\mathcal{P}_{\beta}^{(j)}\rangle\rangle term on the right hand side of Eq. (13). Equation (9a) can then be used to expand the second term, 𝒬α(i)|^^(/λt)|𝒫β(j)\langle\langle\mathcal{Q}_{\alpha}^{(i)}|\hat{\hat{\mathcal{L}}}({\partial}/{\partial\lambda_{t}})|\mathcal{P}_{\beta}^{(j)}\rangle\rangle yielding a recursive expression for the projected change of the normal modes.

𝒬α(i)|λt|𝒫β(j)=(1χβα(𝒬α(i)|^^λt|𝒫β(j)+𝒬α(i+1)|λt|𝒫β(j)𝒬α(i)|λt|𝒫β(j1)),\begin{split}\langle\langle\mathcal{Q}_{\alpha}^{(i)}|\frac{\partial}{\partial\lambda_{t}}|\mathcal{P}_{\beta}^{(j)}\rangle\rangle&=\left(\frac{1}{\chi_{\beta\alpha}}(\langle\langle\mathcal{Q}_{\alpha}^{(i)}|\frac{\partial\hat{\hat{\mathcal{L}}}}{\partial\lambda_{t}}|\mathcal{P}_{\beta}^{(j)}\rangle\rangle+\langle\langle\mathcal{Q}_{\alpha}^{(i+1)}|\frac{\partial}{\partial\lambda_{t}}|\mathcal{P}_{\beta}^{(j)}\rangle\rangle\right.\\ &\left.-\langle\langle\mathcal{Q}_{\alpha}^{(i)}|\frac{\partial}{\partial\lambda_{t}}|\mathcal{P}_{\beta}^{(j-1)}\rangle\rangle\right),\end{split} (14)

where χβαχβχα\chi_{\beta\alpha}\equiv\chi_{\beta}-\chi_{\alpha}. Iterating recursively through Eq. (14), the change in the normal modes can be related to the change in the modulated Liouvillian by

𝒬αi|λ|𝒫β(j)=p=1nαik1=0jS0kp=0jSp𝒬α(i+p1)|^^λ|𝒫β(jSp)(1)Spχβαp+Sp\langle\langle\mathcal{Q}_{\alpha}^{i}|\frac{\partial}{\partial\lambda}|\mathcal{P}_{\beta}^{(j)}\rangle\rangle=\sum_{p=1}^{n_{\alpha}-i}\sum_{k_{1}=0}^{j-S_{0}}\cdot\cdot\cdot\sum_{k_{p}=0}^{j-S_{p}}\frac{\langle\langle\mathcal{Q}_{\alpha}^{(i+p-1)}|\frac{\partial\hat{\hat{\mathcal{L}}}}{\partial\lambda}|\mathcal{P}_{\beta}^{(j-S_{p})}\rangle\rangle}{(-1)^{S_{p}}\chi_{\beta\alpha}^{p+S_{p}}} (15)

where Sqs=1qksS_{q}\equiv\sum_{s=1}^{q}k_{s}, and ksk_{s} are the summation indexes over states in the Jordan Block.

Finally, by substituting Eq. (15) into Eq. (12) and considering only the terms that couple non-degenerate Jordan blocks αβ\alpha\neq\beta, we obtain an adiabaticity condition for slowly modulated open quantum systems:

max0tT|p=1nαik1=0jS0kp=0jSp𝒬α(i+p1)|^^λ|𝒫β(jSp)(1)Spχβαp+Sp|min0tT|λ˙t1|.\begin{split}\max_{0\leq t\leq T}&\left|\sum_{p=1}^{n_{\alpha}-i}\sum_{k_{1}=0}^{j-S_{0}}\cdot\cdot\cdot\sum_{k_{p}=0}^{j-S_{p}}\frac{\langle\langle\mathcal{Q}_{\alpha}^{(i+p-1)}|\frac{\partial\hat{\hat{\mathcal{L}}}}{\partial\lambda}|\mathcal{P}_{\beta}^{(j-S_{p})}\rangle\rangle}{(-1)^{S_{p}}\chi_{\beta\alpha}^{p+S_{p}}}\right|\\ &\ll\min_{0\leq t\leq T}|\dot{\lambda}_{t}^{-1}|.\end{split} (16)

This expression is admittedly unwieldy but provides a completely general criteria for open system adiabaticity. A number of simpler but less tight bounds can be obtained for the adiabaticity condition by extending the approach discussed in Sarandy and Lidar (2005) that treated constant Liouvillian. The simplest of these results states that if λ˙t0\dot{\lambda}_{t}\to 0, e.g., in Eq. (12), then the open quantum system undergoes adiabatic dynamics. Moreover, similarly to the closed system adiabatic theorem, Eq. (16) can be used to derive the earlier non-modulated adiabatic theorem of Ref. Sarandy and Lidar (2005) by assuming a constant ^^0(t)\hat{\hat{\mathcal{L}}}_{0}(t).

A significant example of the open system AMT is provided in Sect. III.2 where the theorem is applied to the slowly turned-on incoherent (e.g., solar radiation) excitation of molecular systems.

III Computational Examples

III.1 Isolated Systems

III.1.1 Rabi type system

In this section, we consider a two-level system (TLS) driven by a frequency modulated oscillatory field. We begin by defining an extension of the Rabi model, the family of Hamiltonians,

H^0(t;ω)=(Δ0VeiωtVeiωtΔ0),\hat{H}_{0}(t;\omega)=\begin{pmatrix}-{\Delta_{0}}&Ve^{i\omega t}\\ V^{*}e^{-i\omega t}&{\Delta_{0}}\end{pmatrix}, (17)

that characterize driving by a field with frequency ω\omega. Here Δ0\Delta_{0} is the energy difference between the states |0\ket{0} and |1\ket{1}, the coupling coefficient is given by V=𝑬𝝁𝟎𝟏V=\bm{E}\cdot\bm{\mu_{01}}, 𝑬\bm{E} is the electric field vector driving the system, 𝝁𝟎𝟏\bm{\mu_{01}} is the transition dipole moment between the two states and we have set =1\hbar=1. This family of Hamiltonians comprises the standard Rabi model with well understood dynamics for all values of ω\omega.

It is useful to briefly review the dynamics induced by Eq. (17) from the perspective of Floquet’s theorem to highlight the normal modes that play an important role in the generalized adiabatic theorem. Since H^0(t)\hat{H}_{0}(t) is periodic with period T=2π/ωT=2\pi/\omega, it’s dynamical normal modes are two TT-periodic Floquet states, |±(t;ω)\ket{\pm(t;\omega)} with time-independent quasienergies ϵ±\epsilon_{\pm}. These states are eigenstates of the Floquet Hamiltonian, satisfying

H^F(t;ω)|±(t;ω)=ϵ±|±(t;ω)\hat{H}_{F}(t;\omega)\ket{\pm(t;\omega)}=\epsilon_{\pm}\ket{\pm(t;\omega)} (18a)
H^F(t;ω)H^0(t;ω)it\hat{H}_{F}(t;\omega)\equiv\hat{H}_{0}(t;\omega)-i\frac{\partial}{\partial t} (18b)
.

For the Hamiltonian in Eq. (17), these Floquet states and quasi energies are given by

|+(t;ω)=(sinθ2eiω2tcosθ2eiϕeiω2t)\ket{+(t;\omega)}=\begin{pmatrix}\sin\frac{\theta}{2}e^{i\frac{\omega}{2}t}\\ \cos\frac{\theta}{2}e^{-i\phi}e^{-i\frac{\omega}{2}t}\end{pmatrix} (19a)
|(t;ω)=(cosθ2eiω2tsinθ2eiϕeiω2t)\ket{-(t;\omega)}=\begin{pmatrix}\cos\frac{\theta}{2}e^{i\frac{\omega}{2}t}\\ \sin\frac{\theta}{2}e^{-i\phi}e^{-i\frac{\omega}{2}t}\end{pmatrix} (19b)
ϵ±=±Ω±Δ2+|V|2\epsilon_{\pm}=\pm\Omega\equiv\pm\sqrt{\Delta^{2}+|V|^{2}} (19c)

where ΔΔ0ω/2\Delta\equiv\Delta_{0}-\hbar\omega/2 is the detuning, ϕ=argV\phi=\arg{V} is the coupling phase, θarccos(Δ/Ω)=arcsin(|V|/Ω)\theta\equiv\arccos(\Delta/\Omega)=\arcsin(|V|/\Omega) is the mixing angle and Ω\Omega is the generalized Rabi frequency.

Consider then driving this system by a frequency modulated field with time varying ωωt\omega\to\omega_{t}. The resulting Hamiltonian can be expressed as a modulation of the form discussed above with H~^(t)=H^0(t;ωt)\hat{\widetilde{H}}(t)=\hat{H}_{0}(t;\omega_{t}), that sweeps through Hamiltonians in the family described by Eq. (17). In general, this problem does not admit a closed form solution, but is tractable through direct numerical propagation of the TDSE. However, in the limit where the modulation changes sufficiently slowly the dynamics can be solved using the generalized adiabatic theorem for isolated systems [Eq. (7)].

To define the domain in which this theorem applies consider the modulation derivative,

ωH~^(t;ω)=(0itVeiωtitVeiωt0)\frac{\partial}{\partial\omega}\hat{\widetilde{H}}(t;\omega)=\begin{pmatrix}0&itVe^{i\omega t}\\ -itV^{*}e^{-i\omega t}&0\end{pmatrix} (20)

which characterizes the effect of the modulation on the driving field. Given Eq. (19), we have

|(t;ω)ωH~^(t;ω)+(t;ω)|=iVΔΩt,\ket{-(t;\omega)}\frac{\partial}{\partial\omega}\hat{\widetilde{H}}(t;\omega)\bra{+(t;\omega)}=i\frac{V\Delta}{\Omega}t, (21)

required for the generalized adiabatic theorem.

Substituting Eqs. (21) and (20) into Eq. (7) shows that the generalized adiabatic theorem applies when

max0tτ{|ω˙t||V|Δt2Ωt2t}min0tτ{2|Ωt|},\max_{0\leq t\leq\tau}\left\{\frac{|\dot{\omega}_{t}||V|\Delta_{t}}{2\Omega_{t}^{2}}t\right\}\ll\min_{0\leq t\leq\tau}\{2|\Omega_{t}|\}, (22)

where we have indicated the parameters that change over time upon frequency modulation using the subscript tt. Here |Ωt|=Δt2+|V|2|\Omega_{t}|=\sqrt{\Delta_{t}^{2}+|V|^{2}}.

When Eq. (22) is well satisfied, then a system initialized in one of the generalized adiabats |±\ket{\pm} will remain in that state at all times. To demonstrate this, we consider a simple linear modulation ωt=ω0+ω˙t\omega_{t}=\omega_{0}+\dot{\omega}t, where ω˙\dot{\omega} is the constant frequency sweep velocity, and where the field, initially in resonance Δt=0=0\Delta_{t=0}=0, is swept linearly to Δt=τ=4Δ0\Delta_{t=\tau}=4\Delta_{0}. In Fig. 1, numerically exact dynamics are obtained for a range of sweep velocities through Runge-Kutta integration of the TDSE. These dynamics are then compared to the adiabatic theorem predictions, showing excellent agreement for slow modulations. (We note that the time taken to perform this modulation varies significantly for different sweep velocities, and we have plotted the dynamics in Fig. 1 on a normalized time axis t/τt/\tau.)

Refer to caption
Figure 1: Dynamics of a two-level system driven by a frequency modulated field. The frequency of the driving field is swept linearly from a resonant frequency ω0=Δ0\hbar\omega_{0}=\Delta_{0} to ωτ=5Δ0\hbar\omega_{\tau}=5\Delta_{0} with sweep velocity ω˙\dot{\omega} (i.e. ωt=Δ0+ω˙t\omega_{t}=\Delta_{0}+\dot{\omega}t). This transformation takes place over a time interval τ=4Δ0/ω˙\tau=4\Delta_{0}/\dot{\omega}. The yellow to blue traces show the dynamics under a range of sweep velocities while the red dashed line shows the adiabatic prediction expected for infinitely slow sweep velocities.

This example is helpful in highlighting the construction and flexibility of the AMT. The adiabatic speed limit is not set by the energy gap but rather by the Rabi frequency Ω\Omega on the right hand side of Eq. (22). This feature, typical of the AMT, is remarkable since it allows for manipulation of the adiabatic time scales simply by increasing the intensity of the driving field VV, and hence |Ωt||\Omega_{t}|, and allows for adiabatic transformations of near degenerate systems on manageable timescales. Moreover, the normal modes, given by the Floquet states |±\ket{\pm}, follow the dynamics of the system under fixed frequency driving. By following these oscillatory dynamics the oscillations of the driving field that typically lead to the failure of traditional AT’s are removed from consideration Amin (2009); Du et al. (2008); Tong et al. (2005); Marzlin and Sanders (2004). In this particularly simple case where only one frequency of light drives the system, this is equivalent to moving to a rotating reference frame using the interaction picture. The instantaneous normal modes can therefore be thought of as a more general method for following the dynamics of the system when an interacting reference frame cannot be defined.

III.1.2 Generalizing Adiabatic Experiments

As an indication of the additional role of the AMT in isolated systems, note that it allows a wide range of experimental applications beyond that of the traditional AT. Currently, experimental Adiabatic Rapid Passage (ARP) and Stimulated Raman Adiabatic Passage (STIRAP) techniques have combined the rotating wave approximation with a rotating reference frame in order to apply the traditional AT to optically driven systems. While this approach can be effective for treating systems driven by near-resonant monochromatic lasers or transform limited pulses, it imposes several restrictions on the types of experiments that can be realized. The AMT, however, removes many of these limitations, allowing for driving beyond the Rotating Wave Approximation, driving of a transition by multichromatic fields, and driving of more simultaneous transitions than is allowed under standard conditions. The removal of these limitations can have significant effects on possible experimental techniques. (For example, a recently derived adiabaticity condition Wang and Plenio (2016) allowed for the use of specifically designed pulse sequences to realize adiabatic transformations in finite time. Xu et al. (2019)) Moreover, the AMT allows for the adiabatic treatment of new phenomena that lie far outside of the rotating wave regime, such as Sisyphus cooling, and can address experimental challenges that are intractable using simpler driving protocols, such implementing Stimulated Hyper-Raman Adiabatic Passage in the presence of Autler-Townes shifts from spectator states Guérin et al. (1998); Yatsenko et al. (1998).

To appreciate the advantages afforded by the AMT, consider issues in the application of the traditional AT to oscillatory Hamiltonians. This has been carried out in the Interaction Picture by moving to a rotating reference frame where the oscillatory Hamiltonian time-dependence can be removed. If this can not be done, then the Marzlin-Sanders inconsistency Marzlin and Sanders (2004) prevents application of the traditional adiabatic theorem. Hence, applications of the traditional AT are limited to systems where an appropriate choice of basis and reference Hamiltonian can be found to remove all oscillatory time-dependence.

Consider then the limits of the transformations possible by an appropriate choice of reference Hamiltonian in the Interaction picture. Let H(t)H(t) be an arbitrary time-dependent Hamiltonian

H(t)=iEi(t)|ii|+ijVij(t)|ij|,H(t)=\sum_{i}E_{i}(t)\ket{i}\bra{i}+\sum_{ij}V_{ij}(t)\ket{i}\bra{j}, (23)

where Ei(t)E_{i}(t) are the time-dependent diagonal energies and Vij(t)V_{ij}(t) the off-diagonal couplings.

An interaction picture, such as the commonly used rotating reference frame, is defined by a choice of a reference Hamiltonian H0H_{0}. For simplicity, we can work in the eigenbasis of the reference Hamiltonian to give the representation H0=iEi(0)|ii|H_{0}=\sum_{i}E^{(0)}_{i}\ket{i}\bra{i}. Equation (23) can then be rewritten in the interaction picture to give

HI(t)=eiH0(t)t(H(t)H0)eiH0(t)t=i(Ei(t)Ei(0))|ii|+ijVij(t)eiEij(0)t|ij|H_{I}(t)=e^{i\frac{H_{0}(t)}{\hbar}t}(H(t)-H_{0})e^{-i\frac{H_{0}(t)}{\hbar}t}=\sum_{i}(E_{i}(t)-E^{(0)}_{i})\ket{i}\bra{i}+\sum_{ij}V_{ij}(t)e^{-i\frac{E^{(0)}_{ij}}{\hbar}t}\ket{i}\bra{j} (24)

where Eij(0)=Ei(0)Ej(0)E^{(0)}_{ij}=E^{(0)}_{i}-E^{(0)}_{j}.

Considering the conditions under which oscillatory time-dependence can be removed from Eq. (24) provides a set of conditions for when AT based approaches can be applied. This requires a basis in which the diagonal elements of H(t)H(t) have no oscillatory time-dependence, defining the eigenbasis of the reference Hamiltonian H0H_{0}. The off diagonal elements show that the only way to eliminate oscillatory time-dependence in Vij(t)V_{ij}(t) is if VijeiωijtV_{ij}\propto e^{-i\omega_{ij}t} for a frequency ωij=Eij(0)/\omega_{ij}=E^{(0)}_{ij}/\hbar. This condition that prescribes a specific functional form for the time-dependence of VijV_{ij} is very restrictive and suggests that the strategy of removing oscillatory Hamiltonian time-dependence will only be effective in specific situations. However, this approach has been successful since the oscillatory functional form VijeiωijtV_{ij}\propto e^{-i\omega_{ij}t} is precisely what is seen when modelling optically driven transitions in the rotating wave approximation.

This condition, however, does limit the application of the traditional adiabatic theorem to specific types of optical excitation. First, the excitation must be well modelled using the rotating wave approximation. If this is not the case, then the off-diagonal matrix elements would be real valued (e.g. of the form Vijsin(ωijt)V_{ij}\propto\sin(\omega_{ij}t)) and the interaction picture transformation would leave a residual oscillatory component (e.g. of the form e2iωijte^{2i\omega_{ij}t}). Moreover, it precludes the driving of one transition by more than one frequency of light, preventing the use of multi-color excitation. Finally, for a discrete NN dimensional system, the N(N1)N(N-1) off diagonal elements impose up to N(N1)N(N-1) conditions ωij=Eij(0)\hbar\omega_{ij}=E^{(0)}_{ij} conditions on the NN diagonal elements of H0H_{0}. Beyond two-dimensional spin systems, it is not guaranteed that it will be possible to simultaneously satisfy all of these conditions. As a result, removal of oscillatory time-dependence can only be guaranteed when a total of NN or fewer transitions are driven by an oscillatory external field unless additional resonance conditions are met. The AMT shares none of these limitations.

III.2 Open System Adiabatic Turn-on of Incoherent Light

A particularly important case in open system quantum mechanics involves the adiabatic turn-on of incoherent radiation that is incident on a molecule and the role of quantum coherences in biological processes (e.g photosynthesis and vision). In particular, oscillatory coherences have been observed experimentally in the excitation of biological molecules with pulsed lasers Engel et al. (2007); Zhang et al. (2015); Ishizaki et al. (2010); Johnson et al. (2015). By contrast, we have argued, supported by numerical studies, that natural processes rely upon slowly-turned-on incoherent light, which leads to steady-state transport Jiang and Brumer (1991); Brumer (2018); Dodin and Brumer (2019) with no oscillating coherences. As shown below, the application of the open system adiabatic theorem proves that if a system is driven by very-slowly turned on light, then the only coherences that will be observed are those that survive to the steady state. In particular, the pulsed laser generated coherences noted above do not survive and are irrelevant under natural biological conditions.

Specifically, consider a molecular system initially prepared in the absence of a driving light field. In this case, the system is initially in a simple equilibrium steady state (i.e., with vanishing dynamical frequency). As such, before the radiation field is turned on, the system is found in a Jordan block with zero eigenvalue. At t=0t=0 an incoherent light field is turned on on a time scale much slower than the dynamics of the molecule, a typical circumstance in light induced biophysical processes. This corresponds to a modulation in the limit of λ˙t0\dot{\lambda}_{t}\to 0, ensuring adiabaticity [see Eq. (16] of the underlying dynamics. As a result, at all times, the system is found in a Jordan block with vanishing eigenvalue, that is, in a steady state. This then indicates that the only coherences observed in the slow turn-on limit are those that survive in the steady-state, such as those seen in previous theoretical studies Koyu et al. (2020); Dodin and Brumer (2019); Reppert et al. (2019); Axelrod and Brumer (2018, 2019).

In many cases, consistency with thermodynamics requires the system to have one steady state given by the Gibbs state ρexp(H^/kBT)\rho\propto\exp(-\hat{H}/k_{B}T) which shows no coherences between non-degenerate energy eigenstates. This result suggests that the non-steady state coherent dynamics observed under finite turn-on times are a consequence of the rapid turn on of the incoherent light field. Figure 2 discussed below provides a numerical example of these predictions for a popular generic V-system model Dodin and Brumer (2019); Dodin et al. (2016).

The ability to show that no coherent dynamics survive the slow turn-on limit of incoherent light without requiring any calculation highlights the intuitive power of the adiabatic theorems derived in this letter. Notably, this situation is far beyond the bounds of applicability of previous adiabaticity condition as the system dynamics are driven by a noisy incoherent light field. This produces a driving Hamiltonian that fluctuates extremely rapidly, on the order of 1 fs, much faster than the energy gap between states and consequently the underlying molecular dynamics. A numerical example follows below.

Consider then the basic minimal model: an open three level system under irradiation by slowly turned-on incoherent light. This model has been previously treated where numerical results showed Dodin et al. (2016) that coherent dynamics, i.e. coherences in the excited state, vanished under slow turn-on of the exciting field. The system has a ground energy eigenstate, |g\ket{g}, and two excited states, |e1\ket{e_{1}} and |e2\ket{e_{2}} that are separated by an energy Δ\Delta and is excited by an incoherent light source with time-dependent intensity. The dynamics of this system, in the weak coupling limit, is characterized by the following Partial Secular Bloch-Redfield Master equations for the matrix elements of the density operator ρ^\hat{\rho}:

ρ˙gg=(r1(t)+r2(t))ρgg+(r1(t)+γ1)ρe1e1+(r2(t)+γ2)ρe2e2+2p(r1(t)r2(t)+γ1γ2)ρe1e2R\dot{\rho}_{gg}=-(r_{1}(t)+r_{2}(t))\rho_{gg}+(r_{1}(t)+\gamma_{1})\rho_{e_{1}e_{1}}+(r_{2}(t)+\gamma_{2})\rho_{e_{2}e_{2}}+2p(\sqrt{r_{1}(t)r_{2}(t)}+\sqrt{\gamma_{1}\gamma_{2}})\rho_{e_{1}e_{2}}^{R} (25a)
ρ˙eiei=ri(t)ρgg(ri(t)+γi)ρeieip(r1(t)r2(t)+γ1γ2)ρe1e2R\dot{\rho}_{e_{i}e_{i}}=r_{i}(t)\rho_{gg}-(r_{i}(t)+\gamma_{i})\rho_{e_{i}e_{i}}-p(\sqrt{r_{1}(t)r_{2}(t)}+\sqrt{\gamma_{1}\gamma_{2}})\rho_{e_{1}e_{2}}^{R} (25b)
ρ˙e1e2=12(r1(t)+r2(t)+γ1+γ2+2iΔ)ρe1e2+p2r1(t)r2(t)(2ρggρe1e1ρe2e2)p2γ1γ2(ρe1e1+ρe2e2)\dot{\rho}_{e_{1}e_{2}}=-\frac{1}{2}(r_{1}(t)+r_{2}(t)+\gamma_{1}+\gamma_{2}+2i\Delta)\rho_{e_{1}e_{2}}+\frac{p}{2}\sqrt{r_{1}(t)r_{2}(t)}(2\rho_{gg}-\rho_{e_{1}e_{1}}-\rho_{e_{2}e_{2}})-\frac{p}{2}\sqrt{\gamma_{1}\gamma_{2}}(\rho_{e_{1}e_{1}}+\rho_{e_{2}e_{2}}) (25c)

where γi\gamma_{i} is the spontaneous emission rate and ri(t)r_{i}(t) is the rate of excitation to (and stimulated emission from) state |ei\ket{e_{i}}. The alignment parameter pp characterizes the probability of simultaneous excitation to states |e1\ket{e_{1}} and |e2\ket{e_{2}} and is defined as

p𝝁𝟏𝝁𝟐μ1μ2p\equiv\frac{\bm{\mu_{1}}\cdot\bm{\mu_{2}}}{\mu_{1}\mu_{2}} (26)

where 𝝁𝒊\bm{\mu_{i}} is the transition dipole moment between the ground state |g\ket{g} and excited state |ei\ket{e_{i}}, assumed real for simplicity. The spontaneous emission and excitation rates are related by detailed balance to give ri(t)=γin¯(t)r_{i}(t)=\gamma_{i}\overline{n}(t) where n¯(t)\overline{n}(t) is the mean occupation number of the resonant thermal field mode. It is the slow turn-on of the incoherent light that leads to a time dependent occupation number n¯(t)\overline{n}(t) of the thermal field reflecting it’s time dependent intensity. That is, the slow modulation of the Liouvillian arises due to slow changes in the statistics of the exciting field, in this case in the mean number of photons in the thermal field modes.

These equations of motion can be analytically solved giving several distinct dynamical regimes. In particular, in the limit where γ¯/ΔP1\overline{\gamma}/\Delta_{P}\gg 1, with γ¯=(γ1+γ2)/2\overline{\gamma}=(\gamma_{1}+\gamma_{2})/2 and Δp=Δ2+(1p2)γ1γ2\Delta_{p}=\sqrt{\Delta^{2}+(1-p^{2})\gamma_{1}\gamma_{2}}, a regime examined below, the system shows long-lived quasi-stationary coherences that eventually decay to give an incoherent thermal state Tscherbul and Brumer (2014). Under excitation by a field with time-dependent intensity of the form n¯(t)=n¯(1exp(αt))\overline{n}(t)=\overline{n}(1-\exp(-\alpha t)), the dynamics are given by

ρeiei(t)=12γ¯{rjΔp22γ¯α[Δp22γ¯(1eαt)α(1eΔp22γ¯t)]+ri2γ¯α[2γ¯(1eαt)α(1e2γ¯t)]}\rho_{e_{i}e_{i}}(t)=\frac{1}{2\overline{\gamma}}\left\{\frac{r_{j}}{\frac{\Delta_{p}^{2}}{2\overline{\gamma}}-\alpha}\left[\frac{\Delta_{p}^{2}}{2\overline{\gamma}}\left(1-e^{-\alpha t}\right)-\alpha\left(1-e^{-\frac{\Delta_{p}^{2}}{2\overline{\gamma}}t}\right)\right]+\frac{r_{i}}{2\overline{\gamma}-\alpha}\left[2\overline{\gamma}\left(1-e^{-\alpha t}\right)-\alpha\left(1-e^{-2\overline{\gamma}t}\right)\right]\right\} (27a)
ρe1e2(t)=pr1r22γ¯{1Δp22γ¯α[Δp22γ¯(1eαt)α(1eΔp22γ¯t)]+12γ¯α[2γ¯(1eαt)α(1e2γ¯t)]}\rho_{e_{1}e_{2}}(t)=\frac{p\sqrt{r_{1}r_{2}}}{2\overline{\gamma}}\left\{\frac{1}{\frac{\Delta_{p}^{2}}{2\overline{\gamma}}-\alpha}\left[\frac{\Delta_{p}^{2}}{2\overline{\gamma}}\left(1-e^{-\alpha t}\right)-\alpha\left(1-e^{-\frac{\Delta_{p}^{2}}{2\overline{\gamma}}t}\right)\right]+\frac{1}{2\overline{\gamma}-\alpha}\left[2\overline{\gamma}\left(1-e^{-\alpha t}\right)-\alpha\left(1-e^{-2\overline{\gamma}t}\right)\right]\right\} (27b)

with γ¯\overline{\gamma} and Δp\Delta_{p} defined above. A detailed derivation of Eq. (27) along with an in depth discussion of the underlying physics are given in Ref. Dodin et al. (2016).

While Eq. (25) is complicated, the dynamics are qualitatively simple. When the turn on time τr=α1\tau_{r}=\alpha^{-1} is shorter than the coherence lifetime 2γ¯/Δp22\overline{\gamma}/\Delta_{p}^{2}, the quasi-stationary coherences lead to a potentially long-lived superposition of excited states which rises on a timescale τr\tau_{r}. The coherences then decay on a timescale 2γ¯/Δp22\overline{\gamma}/\Delta_{p}^{2}. As the turn on time becomes comparable to or longer than the coherence lifetime, the quasi-stationary superposition is not fully excited, leading to a decrease in the generated coherences that disappear entirely in the τr\tau_{r}\to\infty limit. These dynamics are shown for a variety of turn on rates, α\alpha, in the solid traces of Fig. 2 for the case of p=1p=1, γ1=γ2\gamma_{1}=\gamma_{2}.

Refer to caption
Figure 2: Dynamics of a V system driven by slowly turned on incoherent light. The intensity of the exciting field, i.e. the mean photon number, turns on with a rate α\alpha following an exponential activation function n¯(t)(1exp(αt))\overline{n}(t)\propto(1-\exp(-\alpha t)). We consider a system driven by a blackbody source with Temperature T=5800KT=5800K with ground to excited state excitation energy of ωge=1.98\omega_{ge}=1.98 eV. Both excited states have a natural line width of γ=1\gamma=1 GHz with a splitting of Δ=0.001γ\Delta=0.001\hbar\gamma. The yellow to blue traces show the dynamics under a range of turn on rates while the dashed red trace shows the adiabatic turn-on prediction.

These traces can be compared to the dashed red trace which shows the predictions of the open system adiabatic theorem. As discussed above, since the system is initialized in a steady state of the zero field case, a sufficiently slow turn on leads to dynamics that remain in the instantaneous steady state at n¯(t)\overline{n}(t). In this example involving only a single bath, the steady state is the thermal Gibbs state with a temperature determined by n¯(t)\overline{n}(t), and therefore shows no coherences. Consequently, as illustrated in this example, the open systems adiabatic theorem guarantees that a sufficiently slow turn-on of an incoherent field will only show coherences that survive in the steady-state. In the case of two baths, steady state coherences associated with transport will persist Dodin and Brumer (2019); Koyu et al. (2020).

IV Summary

We have presented a new set of generalized adiabatic theorems that apply to rapidly fluctuating fields with slowly varying properties. These results express the dynamics of the modulated field in terms of the dynamical normal modes of time dependent reference Hamiltonians. The resulting adiabaticity conditions take a form that is conceptually similar to preexisting theorems but bounds adiabaticity based on the rate of a slowly changing transformation rather than the rapid underlying field. These results significantly extend the applicability of the adiabatic theorem and its underlying intuition to, e.g., the domain of optical excitations that play a central role in atomic, molecular and optical physics.

Acknowledgments This work was supported by the US Air Force Office of Scientific Research Under Contract Number FA9550-17-1-0310 and FA9550-20-1-0354.

References