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

Incoherent control of optical signals; quantum heat engine approach

Md Qutubuddin and Konstantin E. Dorfman [email protected] State Key Laboratory of Precision Spectroscopy, East China Normal University, Shanghai 200062, China
Abstract

Optical pump-probe signals can be viewed as work done by the matter while transferring the energy between two coherent baths (from pump to probe). In thermodynamics a heat engine, such as laser, is a device which performs similar work but operating between two thermal baths. We propose an “incoherent” control procedure for the optical signals using the physics of quantum heat engine. By combining a coherent laser excitation of electronic excited state of molecule with thermal relaxation we introduce an effective thermal bath treating stimulated emission of probe photons as work performed by the heat engine. We optimize power and efficiency for the pump-probe signal using control parameters of the pump laser utilizing four level molecular model in strong and weak coupling regime illustrating its equivalence with the thermodynamic cycle of the heat engine.

pacs:
Valid PACS appear here

The advancement in the field of quantum heat engines (QHE) attracted a lot of attention in the last decade due to its connection with the real physical systems, such as lasers, solar cells sd11 , and biological systems dv13 . The maximum quantum efficiency for a three-level maser QHE, first introduced by Scovil and Schulz-DuBois ssd59 ; ssdap ; ssd67 is governed by the Carnot bound obtained by invoking the detailed balance condition. Since then many theoretical proposals for thermodynamical heat engine in quantum regime have been discussed al79 ; prl98 ; pra74 ; ul15 ; agl7 ; lk17 ; sprl ; cgg17 . In addition effects of profound quantum nature, such as quantum coherence and correlation and there influence on the performance of QHEs have been further investigated sz03 ; sd11 ; dv13 . Recent experiments demonstrated that the QHE physics rd16 ; dd18 ; pb19 can be studied using pump-probe optical measurements kb17 where the working fluid is a radiation produced by resonantly driven electronic transitions in the material. The spectroscopic setup can be therefore viewed as the QHE which transfers energy from one heat bath (pump pulse) to another (probe pulse), while the work performed by the system is measured in the form of detected probe photons. Furthermore we employ the reservoir engineering pprl96 ; kqs99 ; cprl01 which has been previously studied in the context of thermodynamics zpra03 ; vnp09 ; ppra10 ; anj11 ; mepl11 as a control method for optical measurements utilizing the analogy between spectroscopic setups and quantum heat engines. While this analogy is not complete, since in QHE the system is in contact with thermal environment, while spectroscopic measurements are performed with coherent laser sources, there is a possibility to connect the two approaches in a sensible way. In particular, we introduce an effective thermal bath that mimics all the dynamical properties of the thermal bath by replacing a coherent laser excitation followed by a dissipative phonon-assisted relaxation due to internal degrees of freedom with an incoherent pumping. In the effective two-level system this can be achieved by matching the populations of the electronic states with two types of baths (coherent plus relaxation vs thermal). This consequently defines the range of parameters of the laser source which can be used to mimic thermal operation of the effective QHE. Once the laser parameters are fixed one can calculate the power and efficiency of the QHE assuming either strong or weak coupling between the molecular system and the probe field. The former represents a conventional QHE regime kpra18 , while the later is a typical case for spectroscopic measurements muknl , where weak field-matter interactions allow the perturbative treatment of the signals.

Various methods have been developed to optimize spectroscopic signals by carefully engineering phase relations in the system-baths interactions. For instance, pulse shaping techniques allow to modify light absorption pathways using constructive and destructive interference by manipulating phase of the optical pulses gos3 . Quantum control theory ys09 is yet another method, which gradually evolved from the studies of linear, closed, and small-scale systems to nonlinear, open, and large-scale networks. This optimization process relies on an elaborate numerical feedback loop algorithms. The utilization of the quantum control sb11 governs the system evolution through the Hamiltonian (unitary) dynamics and has therefore limited applicability for an open quantum systems. This limitation can be overcome by the addition of active manipulation of non-unitary (i.e., incoherent) evolution described by Liouville operator responsible for the effects of system environment. The control methods applicable to the dissipative (incoherent) dynamics due to the environment must be therefore developed. These new methods must be based on a different set of assumptions and are distinct from the coherent control methods applicable to the unitary system evolution pr06 as well as the thermal reservoir engineering lsr20 with feedback control mepj57 ; mpra04 ; spra06 .

Refer to caption
Figure 1: (color online)(a) Schematic of the pump-probe measurement in molecular system consisting of two electronic states. Pump field resonant with electronic transition gg-2 excites a vibrational wavepacket in the higher energy vibrational state 2 which relaxes to the lower energy vibrational state 1. Probe field then stimulates the emission from the state 11 to the excited vibrational level 0 of the ground electronic state. Finally, vibrational relaxation brings the system back to its ground state gg. (b) Equivalent three-level QHE with energy levels transitions g1g-1 and g0g-0 driven by hot (at temperature ThT_{h}) and cold (at TcT_{c}) heat baths. The single-mode stimulated emission occurs at 101-0 transition with the coupling strength λ\lambda.

We consider a two-level molecular system with the ground state gg and the excited electronic state ee shown in Fig. 1a. We further consider two vibrational states of the excited electronic state 11 and 22. A coherent pump field excites resonantly the transition g2g-2 with Rabi frequency Ωp\Omega_{p}. The upper vibrational state 2 then relaxes to lower state 1 via the phonon emission. The stimulated emission 1-0 via interaction with the probe field with Rabi frequency λ\lambda followed by the thermal relaxation via interaction with the cold bath 0g0-g then brings the system to its initial ground state. The corresponding equation of motion for the density matrix is given by

ρ˙gg\displaystyle\dot{\rho}_{gg} =\displaystyle= 2Γc[(nc+1)ρ00ncρgg]+iΩp(ρg2ρ2g),\displaystyle 2\Gamma_{c}[(n_{c}+1)\rho_{00}-n_{c}\rho_{gg}]+i\Omega_{p}(\rho_{g2}-\rho_{2g}),
ρ˙00\displaystyle\dot{\rho}_{00} =\displaystyle= 2Γc[(nc+1)ρ00ncρgg]+iλ(ρ01ρ10),\displaystyle-2\Gamma_{c}[(n_{c}+1)\rho_{00}-n_{c}\rho_{gg}]+i\lambda(\rho_{01}-\rho_{10}),
ρ˙11\displaystyle\dot{\rho}_{11} =\displaystyle= Γ2[(n2+1)ρ22n2ρ11]iλ(ρ01ρ10),\displaystyle\Gamma_{2}[(n_{2}+1)\rho_{22}-n_{2}\rho_{11}]-i\lambda(\rho_{01}-\rho_{10}),
ρ˙22\displaystyle\dot{\rho}_{22} =\displaystyle= Γ2[(n2+1)ρ22n2ρ11]iΩp(ρg2ρ2g),\displaystyle-\Gamma_{2}[(n_{2}+1)\rho_{22}-n_{2}\rho_{11}]-i\Omega_{p}(\rho_{g2}-\rho_{2g}),
ρ˙g2\displaystyle\dot{\rho}_{g2} =\displaystyle= [Γ2(n2+1)/2+Γcnci(ω2gωp)]ρg2\displaystyle-[\Gamma_{2}(n_{2}+1)/2+\Gamma_{c}n_{c}-i(\omega_{2g}-\omega_{p})]\rho_{g2}
\displaystyle- iΩp(ρggρ22),\displaystyle i\Omega_{p}(\rho_{gg}-\rho_{22}),
ρ˙2g\displaystyle\dot{\rho}_{2g} =\displaystyle= [Γ2(n2+1)/2+Γcnc+i(ω2gωp)]ρ2g\displaystyle-[\Gamma_{2}(n_{2}+1)/2+\Gamma_{c}n_{c}+i(\omega_{2g}-\omega_{p})]\rho_{2g}
+\displaystyle+ iΩp(ρggρ22),\displaystyle i\Omega_{p}(\rho_{gg}-\rho_{22}),
ρ˙01\displaystyle\dot{\rho}_{01} =\displaystyle= [Γ2n2/2+Γc(nc+1)i(ω10ωk)]ρ01\displaystyle-[\Gamma_{2}n_{2}/2+\Gamma_{c}(n_{c}+1)-i(\omega_{10}-\omega_{k})]\rho_{01}
\displaystyle- iλ(ρ11ρ00),\displaystyle i\lambda(\rho_{11}-\rho_{00}),
ρ˙10\displaystyle\dot{\rho}_{10} =\displaystyle= [Γ2n2/2+Γc(nc+1)+i(ω10ωk)]ρ10\displaystyle-[\Gamma_{2}n_{2}/2+\Gamma_{c}(n_{c}+1)+i(\omega_{10}-\omega_{k})]\rho_{10} (1)
+\displaystyle+ iλ(ρ11ρ00),\displaystyle i\lambda(\rho_{11}-\rho_{00}),

where Γ2/2\Gamma_{2}/2 is the dephasing rate and n2=[exp(ω21/kBTc)1]1n_{2}=[\exp(\hbar\omega_{21}/k_{B}T_{c})-1]^{-1} is the average phonon occupation number corresponding to 121\leftrightarrow 2 transition at ambient temperature TcT_{c}. Eqs. (1) has been derived using Born-Markov approximation assuming weak near-resonant pump field ΩpΓ2n2\Omega_{p}\ll\Gamma_{2}n_{2}. In this work we follow the pattern established in the series of earlier works which defined the framework for the laser QHE sz03 ; sd11 ; kpra18 , where we assume system in contact with the thermal phonon bath with relaxation described by a constant rate Γ2\Gamma_{2}. This represents a limit of the fast nuclear dynamics. Generally the connection between lineshape function g(t)g(t) and the spectral density is given by muknl g(t)=12π𝑑ωC(ω)ω2[exp(iωt)+iωt1]g(t)=\frac{1}{2\pi}\int_{-\infty}^{\infty}d\omega\frac{C(\omega)}{\omega^{2}}[\exp(-i\omega t)+i\omega t-1], where C(ω)C(\omega) is a spectral density function. Representing a bath by a collections of oscillators we obtain C(ω)=2κωΛω2+Λ2C(\omega)=2\kappa\frac{\omega\Lambda}{\omega^{2}+\Lambda^{2}}, where κ\kappa is a reorganization energy, and Λ\Lambda is a nuclear dynamics timescale. Assuming nuclear dynamics to be fast compared to the coupling strength that governs magnitude of fluctuations Λ22κkT\hbar\Lambda^{2}\gg 2\kappa kT we obtain the homogeneous dephasing limit Re[g(t)]=Γ2tRe[g(t)]=\Gamma_{2}t such that Γ2=κkT/(Λ)\Gamma_{2}=\kappa kT/(\hbar\Lambda) is the homogeneous dephasing limit. In more general case of spectral density, the solution of Eq. (1) can be found numerically as shown in kb17 .

We now investigate the four-level QHE based on Scovil and Schultz-Dubois maser ssd59 shown in Fig. 1b. Quantum heat engine regime is characterized by a strong coupling to the output radiation which makes it distinct from the spectroscopic regime, where probe field is typically weak. A hot reservoir at temperature ThT_{h} of the system resonant with the g1g-1 transition and a cold reservoir at temperature TcT_{c} coupled with the g0g-0 transition. The time evolution of the system in rotating frame is given by kpra18 ; eln02

ρ˙=i[H0H¯+VR,ρR]+c[ρR]+h[ρR],\displaystyle\dot{\rho}=-\frac{i}{\hbar}[H_{0}-\bar{H}+V_{R},\rho_{R}]+\mathcal{L}_{c}[\rho_{R}]+\mathcal{L}_{h}[\rho_{R}], (2)

where H0=i=g,0,1ωi|ii|H_{0}=\hbar\sum_{i=g,0,1}\omega_{i}|i\rangle\langle i|, H¯=ωg|gg|+ω2(|11||00|)\bar{H}=\hbar\omega_{g}|g\rangle\langle g|+\frac{\hbar\omega}{2}(|1\rangle\langle 1|-|0\rangle\langle 0|) and interaction with the probe field is described by VR=λ(|10|+|01|)V_{R}=\hbar\lambda(|1\rangle\langle 0|+|0\rangle\langle 1|), where λ=μ10pr\lambda=\mu_{10}\mathcal{E}_{pr} is the probe-matter coupling which is strong compared to other relaxation processes, where μ10\mu_{10} is a transition dipole moment and pr\mathcal{E}_{pr} is the classical amplitude of probe field. Master equation (2) contains two parts. The first, unitary part contains interaction with the coherent probe field and is governed by a commutator term. The last two terms governed by Liouville operator are describing interaction with the thermal reservoirs (see Appendix A). While the latter assumes weak field-matter interaction which is typical for the thermal radiation, the former has no assumptions about the strength of the probe field. Following the pioneering work on laser QHE ssd59 and more recent work of Scully and others sz03 ; sprl ; sd11 ; dv13 ; kpra18 we adopt the strong coupling to the probe field which represents the so-called QHE limit which corresponds to e.g. cavity radiation of the laser. The Liouville operator for the system-bath interaction is given by w[ρ]=Γw(nw+1)[2|gg|ρlwlw|lwlw|ρρ|lwlw|]+Γwnw[2|lwlw|ρgg|gg|ρρ|gg|],\mathcal{L}_{w}[\rho]=\Gamma_{w}(n_{w}+1)[2|g\rangle\langle g|\rho_{l_{w}l_{w}}-|l_{w}\rangle\langle l_{w}|\rho-\rho|l_{w}\rangle\langle l_{w}|]+\Gamma_{w}n_{w}[2|l_{w}\rangle\langle l_{w}|\rho_{gg}-|g\rangle\langle g|\rho-\rho|g\rangle\langle g|], where w=cw=c(cold) and hh(hot) baths and lwl_{w} is 0 and 11 for the hot and cold baths, respectively. Note, that the strong field is generally defined with respect to the diagonal terms of the Hamiltonian (bare eigen energies), rather than Liouville operator (relaxation process). Furthermore, the QHE power, heat flux and other characteristics are defined via a trace operator, which is invariant with respect to the choice of the basis set. In refs. sd11 ; sz03 ; sprl ; dv13 ; kpra18 ; mprl10 it has been demonstrated how to find the system density matrix using bare system eigenstates. In other works djpc15 ; vnp09 ; ppra10 ; anj11 authors chose to use dressed states, while others xc16 ; agl7 consider strong coupling to the phonon bath and redefined master equation in polaron frame.

Before proceeding to the QHE model which is based on the solution of the complete set of equations given by Eq. (Incoherent control of optical signals; quantum heat engine approach) we first introduce an effective heat bath. To that end we assume that the pump is relatively weak ΩpΓ2n2\Omega_{p}\ll\Gamma_{2}n_{2} and the coupling to the probe field is much stronger than the coupling to the phonon bath that governs 212-1 transition which itself is stronger than that of the bath driving 0g0-g transition: λΓ2n2Γcnc\lambda\gg\Gamma_{2}n_{2}\gg\Gamma_{c}n_{c}. The latter condition can be obtained in variety of molecular systems hb89 . Under these conditions one can eliminate the state 0 from the total system of equations (Incoherent control of optical signals; quantum heat engine approach) and consider only three states such that the coherent excitation g2g-2 followed by a relaxation 212-1. The solution of this reduced system for the ground gg and lowest excited state 11 populations read

ρ11c(t)=𝒩c(t)(n2+1)(1eΓ~t)ρggc=1ρ11c,\displaystyle\rho^{c}_{11}(t)=\mathcal{N}_{c}(t)(n_{2}+1)(1-e^{-\tilde{\Gamma}t})\quad\rho_{gg}^{c}=1-\rho_{11}^{c}, (3)

where superscript cc indicates the coherent bath and Γ~=Γ2(n2+1)/4Γ22(n2+1)2/162Ωp2\tilde{\Gamma}=\Gamma_{2}\left(n_{2}+1\right)/4-\sqrt{\Gamma_{2}^{2}\left(n_{2}+1\right)^{2}/16-2\Omega_{p}^{2}}. Normalization function 𝒩c(t)=[1+2n2+n2eΓ~t]1\mathcal{N}_{c}(t)=[1+2n_{2}+n_{2}e^{-\tilde{\Gamma}t}]^{-1} ensures that the population of an effective two-level system consisting of states gg and 11 is conserved (see Appendix B). In the high temperature limit n21n_{2}\gg 1, and, assuming ΩpΓ2n2\Omega_{p}\ll\Gamma_{2}n_{2} we obtain Γ~4Ωp2Γ2(n2+1)\tilde{\Gamma}\simeq\frac{4\Omega_{p}^{2}}{\Gamma_{2}(n_{2}+1)}. Generally, the condition n21n_{2}\gg 1 represents the high temperature limit. This is the optimum regime for operation of the QHE klr14 . For a room temperature “cold bath” Tc=0.0259T_{c}=0.0259 eV. High temperature limit is valid for various e.g. acoustic-like phonon modes, with the energy scale up to meV range.

The combined effect of the coherent excitation g2g\to 2 followed by the phonon relaxation 212\to 1 can be replaced by an effective thermal bath at temperature ThT_{h} with the average photon number nh=[exp(ω1g/kBTh)1]1n_{h}=[\exp(\hbar\omega_{1g}/k_{B}T_{h})-1]^{-1} and dephasing Γh\Gamma_{h}. In this case the state 22 can be eliminated and the corresponding equation of motion for the populations of gg and 11 read

ρ˙11=Γh[(nh+1)ρ11nhρgg],ρ˙gg=ρ˙11,\displaystyle\dot{\rho}_{11}=-\Gamma_{h}[(n_{h}+1)\rho_{11}-n_{h}\rho_{gg}],\quad\dot{\rho}_{gg}=-\dot{\rho}_{11}, (4)

which yields the time-dependent solution:

ρ11th(t)=𝒩thnh(1eΓh(1+2nh)t),ρggth=1ρ11th,\displaystyle\rho^{th}_{11}(t)=\mathcal{N}_{th}n_{h}(1-e^{-\Gamma_{h}(1+2n_{h})t}),~{}\rho^{th}_{gg}=1-\rho^{th}_{11},\ \ (5)

where superscript thth indicates the thermal bath and the normalization 𝒩th=[1+2nh]1\mathcal{N}_{th}=[1+2n_{h}]^{-1}. In order to match the solutions of the effective thermal bath in Eq. (5) with that of a coherent bath in Eq. (3) the corresponding nhn_{h} and Γh\Gamma_{h} must satisfy

nh(t)=n2+1n2eΓ~t1,Γh(t)=4Ωp2Γ2(n2+1)(2nh(t)+1).\displaystyle n_{h}(t)=\frac{n_{2}+1}{n_{2}e^{-\tilde{\Gamma}t}-1},\quad\Gamma_{h}(t)=\frac{4\Omega_{p}^{2}}{\Gamma_{2}(n_{2}+1)(2n_{h}(t)+1)}. (6)

Eq. (6) describes the time-dependent parameters of the effective thermal bath, which yields a complete match between the two baths (coherent and thermal) at any given time (see Fig. 2b). More detailed description of the environment will result in more complicated relation which can be solved numerically xu18 . One can further provide an approximate relation between the bath parameters that is time independent and is easy to analyze. Assuming n21n_{2}\gg 1 and arbitrarily fixing the time: t=Γ~1log(n2/2)t^{*}=\tilde{\Gamma}^{-1}\log(n_{2}/2) we obtain

nhnh(t)=n2,ΓhΓh(t)=2Ωp2Γ2n22.n^{*}_{h}\equiv n_{h}(t^{*})=n_{2},\quad\Gamma_{h}^{*}\equiv\Gamma_{h}(t^{*})=\frac{2\Omega_{p}^{2}}{\Gamma_{2}n^{2}_{2}}. (7)

Parameters in Eq. (7) results in the population dynamics shown in Fig. 2d. It yields a qualitatively good agreement at initial time (ρgg(0)=1\rho_{gg}(0)=1, ρ11(0)=0\rho_{11}(0)=0) as well as near the steady state (ρgg()=ρ11()=1/2\rho_{gg}(\infty)=\rho_{11}(\infty)=1/2) as seen in Fig. 2a. The error of 8%\sim 8\% associated with the choice of tt^{*} becomes apparent at intermediate times as shown in Fig. 2c, where a comparison with the exact numerical solution of Eq. (Incoherent control of optical signals; quantum heat engine approach) for the full four-level system is presented. While the general expression in Eq. (5) yields negative steady state value for nhn_{h} when tt\to\infty, the relevant population dynamics is determined by Γ~\tilde{\Gamma} which is a small number. Therefore the choice of tt^{*} that is displayed in Fig. 2d corresponds to the value of time at which populations approach the steady state values, while nhn_{h} is a growing function. Note, that the choice of tt^{*} near the steady state, can be used in the following thermodynamic analysis. Note, that the result of Eq. (6) - (7) is applicable for a wide variety of parameters and is valid in the high temperature limit only when n21n_{2}\gg 1. In the low temperature limit, coherent bath creates a population inversion (ρ11c()=1\rho_{11}^{c}(\infty)=1, ρggc()=0\rho_{gg}^{c}(\infty)=0), while the thermal bath yields weak excitation: (ρ11th()=0\rho_{11}^{th}(\infty)=0, ρggth()=1\rho_{gg}^{th}(\infty)=1), which corresponds to the low efficiency regime and thus it won’t be considered any further.

Refer to caption
Figure 2: (Color online ) (a) The population of ground gg and lowest excited state 11 obtained using coherent bath in Eq. (3) (solid lines) and thermal bath ρth\rho^{th} using parameters in Eq. (6) (--dashed) and ρth\rho^{{}^{\prime}th} using Eq. (7)( dot-dashed). (b) The difference between populations of coherent and thermal baths ρcρth\rho^{c}-\rho^{th} for parameters in Eq. (6). (c) same as (b) but for exact solution of Eq. (Incoherent control of optical signals; quantum heat engine approach) (ρ(1)\rho^{(1)}) vs approximate solutions of Eqs. (3) and (7). (d) The time evolution of the microscopic occupation number nh(t)n_{h}(t) in Eq. (6) vs its approximation nh=n2n^{*}_{h}=n_{2} in Eq. (7). The parameters for the simulations are Ωp=0.0001\Omega_{p}=0.0001 eV, Γ2=0.025\Gamma_{2}=0.025ps-1 and n2=1000n_{2}=1000.

The output power, and efficiency of a QHE described by Eq. (Incoherent control of optical signals; quantum heat engine approach) are given by prl98 ; pra74

PQ=iTr{[H0,VR]ρR},η=PQ˙h,\displaystyle P^{Q}=-\frac{i}{\hbar}\text{Tr}\{[H_{0},V_{R}]\rho_{R}\},\quad\eta=-\frac{P}{\dot{Q}_{h}}, (8)

where superscript QQ indicates the QHE power (not to be confused with spectroscopic power in Eq. (12), the heat flux is Qh˙=Tr{h[ρR]H0}\dot{Q_{h}}=\text{Tr}\{\mathcal{L}_{h}[\rho_{R}]H_{0}\}. Following the general approach outlined in kpra18 summarized in Appendix C we assume the high temperature limit and expand the occupation numbers nh=n2=Tc/ω21n_{h}=n_{2}=T_{c}/\omega_{21}, nc=Tc/ωcn_{c}=T_{c}/\omega_{c} where ωc=ω0g\omega_{c}=\omega_{0g}. We then introduce an effective temperature of the hot bath Th=Ωpωc(Γ2Γc/2)1/2T_{h}=\Omega_{p}\omega_{c}(\Gamma_{2}\Gamma_{c}/2)^{-1/2}, dimensionless temperature scale: τ=Tc/Th\tau=T_{c}/T_{h}, pump energy scale: cp=ωp/ωcc_{p}=\omega_{p}/\omega_{c}, and coupling scale: λ=λωc/(ΓcTh)\lambda^{\prime}=\lambda\omega_{c}/(\Gamma_{c}T_{h}). We then optimize the power in Eq. (C7) with respect to dimensionless variable ω21/ωc\omega_{21}/\omega_{c} and obtain

PmaxQ=P0Q2λ2{2cpλ2τ2+cp(λ2+τ2)2λτ𝒞}3(λ2τ2)2,\displaystyle P_{max}^{Q}=P_{0}^{Q}\frac{2\lambda^{{}^{\prime}2}\{2c_{p}\lambda^{{}^{\prime}2}\tau^{2}+c_{p^{\prime}}(\lambda^{{}^{\prime}2}+\tau^{2})-2\lambda^{\prime}\tau\mathcal{C}\}}{3(\lambda^{{}^{\prime}2}-\tau^{2})^{2}}, (9)

where P0Q=ΓcωcP_{0}^{Q}=\Gamma_{c}\omega_{c}, cp=cp1c_{p^{\prime}}=c_{p}-1, 𝒞=(λ2cp+cp)(cpτ2+cp)\mathcal{C}=\sqrt{(\lambda^{{}^{\prime}2}c_{p}+c_{p^{\prime}})(c_{p}\tau^{2}+c_{p^{\prime}})} and λτ\lambda^{\prime}\neq\tau. The efficiency corresponding to the maximum output power defined in Eq. (C9) is thus given by

η=1[cp+λτ(λτcp𝒞)cp(λ2+τ2+1)1]1.\eta^{*}=1-\left[c_{p}+\frac{\lambda^{{}^{\prime}}\tau(\lambda^{{}^{\prime}}\tau c_{p}-\mathcal{C})}{c_{p}\left(\lambda^{{}^{\prime}2}+\tau^{2}+1\right)-1}\right]^{-1}. (10)

Efficiency (10) is evaluated at fixed ωc\omega_{c} since this is a parameter of a given molecular system whereas ωh\omega_{h} can be manipulated via scanning of the pump frequency ωp\omega_{p}.

Refer to caption
Figure 3: (Color online) 2D mapping of the efficiency at maximum power η\eta^{*} in Eq. (11) vs Carnot efficiency ηC=1τ\eta_{C}=1-\tau -(a) 3D mapping of η\eta^{*} vs ηC\eta_{C} vs dimensionless pump frequency cpc_{p} - (b). 2D mapping of the cpc_{p} vs ηC\eta_{C} corresponding to (a).
Bound η\eta^{*} cpc_{p}
I 0 22
I/II ηC/2\eta_{C}/2 4/(2ηC)4/(2-\eta_{C})
II/III ηCA\eta_{CA} 2/1ηC2/\sqrt{1-\eta_{C}}
III/IV ηC/(2ηC)\eta_{C}/(2-\eta_{C}) (2ηC)/(1ηC)(2-\eta_{C})/(1-\eta_{C})
IV ηC\eta_{C} 2/(1ηC)2/(1-\eta_{C})
Table 1: Parameters of the coherent bath corresponding to the QHE efficiency bounds shown in Fig. 3.

We now compare the efficiency at maximum power (10) with the corresponding high temperature bounds of the QHE obtained in kpra18 . We first assume that the coupling with probe field λ\lambda is the largest coupling in the system: λΓhnh,Γcnc\lambda\gg\Gamma_{h}n_{h},\Gamma_{c}n_{c} which yields

ηSC=1[cp+τ(τ1+τ2)]1,\eta^{*}_{SC}=1-\left[c_{p}+\tau(\tau-\sqrt{1+\tau^{2}})\right]^{-1}, (11)

where the subscript SCSC indicates the strong coupling. Entire parameter space corresponding to the efficiency given by Eq. (11) can be separated in four regions summarized in Table I represented by the colorful 2D shapes in Fig. 3a. The characteristic values describing the boundaries between the four regions correspond to 0, ηC/2\eta_{C}/2 (between I and II regions), ηC/(2ηC)\eta_{C}/(2-\eta_{C}) (between III and IV), Carnot efficiency ηC=1τ\eta_{C}=1-\tau upper bound of IV) and Curzon-Ahlborn (CA) limit klr14 ηCA=1τ1/2\eta_{CA}=1-\tau^{1/2} (between II and III). Using dimensionless pump frequency cpc_{p} as a control parameter which depends on the effective temperature ratio τ\tau one can obtain the corresponding three dimensional parameter space for each of the regions that include {τ,cp,η}\{\tau,c_{p},\eta^{*}\} as shown in Fig. 3b. Note, that the two parameters of the pump field: frequency ωp\omega_{p} and the Rabi frequency Ωp\Omega_{p} which defines an effective hot bath temperature ThT_{h} can be controlled experimentally. Thus 2D parameter space {τ,cp}\{\tau,c_{p}\} shows a constrained relation between the two as seen in Fig. 3c. For instance, CA limit is obtained when cp2/τc_{p}\simeq 2/\sqrt{\tau}. This corresponds to the Rabi frequency Ωp(CA)ωp2ΓcΓ2/2Tc/(4ωc)\Omega_{p}^{(CA)}\simeq\omega_{p}^{2}\sqrt{\Gamma_{c}\Gamma_{2}/2}T_{c}/(4\omega_{c}). Similarly one can analyze other bounds corresponding to the standard QHE model kpra18 . There exists an additional constraint corresponding to parameter γ=Γh/Γc\gamma=\Gamma_{h}/\Gamma_{c}. For instance, CA limit is obtained when γ0\gamma\to 0 which corresponds to ΩpΓcΓ2n2\Omega_{p}\ll\sqrt{\Gamma_{c}\Gamma_{2}}n_{2}. In another limit η=ηC/(2ηC)\eta^{*}=\eta_{C}/(2-\eta_{C}) corresponds to γ\gamma\to\infty which yields ΩpΓcΓ2n2\Omega_{p}\gg\sqrt{\Gamma_{c}\Gamma_{2}}n_{2}. Therefore, by setting the specific relation between the pump frequency and intensity one can control energy conversion between the pump and the probe pulse near the maximum of the corresponding thermodynamic cycle. Note, that one can exceed not only standard efficiency at maximum power limits such as CA, but also the Carnot limit. This is not a surprising result, since in the strong coupling limit QHE is approaching lasing threshold, which corresponds to the gain regime and the input coherent drive is not at thermodynamic equilibrium. Thus the system under coherent drive cannot be generally treated as a closed system. Therefore, Eq. (2) describes the system driven by the thermal pump and is therefore bounded by the Carnot limit applicable as long as cp<2(1ηC)1c_{p}<2(1-\eta_{C})^{-1} according to regime IV in Table I.

So far we analyzed the case of the strong coupling to the probe field (λ1\lambda^{\prime}\gg 1), which corresponds to the laser QHE regime. We now focus on the regime typical for spectroscopic signals derived using perturbative expansions over light-matter interactions. We derive the pump-probe signal to second order in the pump as well as in the probe described by the double sided Feynman diagrams shown in Fig. D1 of Appendix D. Diagrams a and b illustrate how the two interactions with the pump pulse excite molecular system from its ground state gg to the excited state population 22 which then relaxes to 11 via phonon scattering. The consequent two interactions with the probe bring the system to 0 state population. In diagram c the system remain in the ground state population gg after a Raman process initiated by the pump (g2gg-2-g). Interaction with a cold bath promotes the system to the population of 0 state and the consequent Raman process initiated by the probe (0100-1-0) brings it back to the 0 state population. The corresponding power is defined as a rate of change of the probe photon number multiplied by a photon energy Ps=ω10ddtE^pr(t)E^pr(t)P_{s}=\omega_{10}\frac{d}{dt}\langle\hat{E}_{pr}^{\dagger}(t)\hat{E}_{pr}(t)\rangle, which takes the form

Ps=ω10pu2pr2Im[R(ωpr,ωp)],\displaystyle P^{s}=\omega_{10}\mathcal{E}_{pu}^{2}\mathcal{E}_{pr}^{2}\text{Im}[R(-\omega_{pr},\omega_{p})], (12)

where superscript ss indicates spectroscopic regime, pu\mathcal{E}_{pu} and pr\mathcal{E}_{pr} are the classical amplitudes of the pump and probe fields, respectively. Note, that both definitions of the output power in Eq. (8) and (12) are equivalent and represent the rate of change of the probe photon energy flux according to Ref. sd11 . However, while Eq. (8) contains definition originated from the general solution of the equations of motion for an arbitrary pump and probe intensities, Eq. (12) results from the perturbative expansion with respect to pump- and probe-matter interactions. The molecular response function has three terms R(ωpr,ωp)=j=a,b,cRj(ωpr,ωp)R(-\omega_{pr},\omega_{p})=\sum_{j=a,b,c}R_{j}(-\omega_{pr},\omega_{p}) defined by Eqs. (D1) - (D3) of Appendix D corresponding to the three diagrams in Fig. D1 muknl ; mcr9 . Following a similar optimization of the power defined by Eq. (D12) with respect to ω21\omega_{21} that results in Eq. (9) we obtain for the maximum power

Pmaxs=λ2γcP0sf2τ4(f+2γc)2(cp+γc+fγc/2),\displaystyle P_{max}^{s}=\frac{\lambda^{{}^{\prime}2}\gamma_{c}P_{0}^{s}}{f^{2}\tau^{4}}(f+2\sqrt{\gamma_{c}})^{2}(c_{p^{\prime}}+\gamma_{c}+f\sqrt{\gamma_{c}}/2), (13)

where P0s=6Γc2Tc2/(σpωc)P_{0}^{s}=6\Gamma_{c}^{2}T_{c}^{2}/(\sigma_{p}\omega_{c}), σp\sigma_{p} is a bandwidth of the pump field, f=γc8cp+9γcf=\sqrt{\gamma_{c}}-\sqrt{8c_{p^{\prime}}+9\gamma_{c}}, and γc=Γ2/Γc\gamma_{c}=\Gamma_{2}/\Gamma_{c}. Since the spectroscopic signal represents power rather than the efficiency we will be comparing maximum power PmaxQP_{max}^{Q} in Eq. (9) vs PmaxsP_{max}^{s} in Eq. (13). We first note, that while PsP^{s} has been obtained to second order in the probe field, the QHE power PQP^{Q} in Eq. (9) is derived to all orders in the probe. Therefore, we expand PQP^{Q} to second order in λ\lambda^{\prime}. Second, PmaxsP_{max}^{s} depends on extra parameter γc\gamma_{c} which is absent in the QHE result. Maximized powers become equivalent PmaxQ=PmaxsP_{max}^{Q}=P_{max}^{s} at γc=2cp2/(9τ2)\gamma_{c}=2c_{p}^{2}/(9\tau^{2}) assuming for brevity that P0Q=P0s=1P_{0}^{Q}=P_{0}^{s}=1. On the other hand the maximum of Eq. (13) with respect to γc\gamma_{c} is achieved at γc=cp/2\gamma_{c}^{*}=c_{p^{\prime}}/2. Thus, Eqs. (9) and (13) reduce to

PmaxQ=2λ2cp3τ2,Pmaxs=λ2cp216τ4.\displaystyle P_{max}^{Q*}=\frac{2\lambda^{{}^{\prime}2}c_{p^{\prime}}}{3\tau^{2}},\quad P_{max}^{s*}=\frac{\lambda^{{}^{\prime}2}c_{p^{\prime}}^{2}}{16\tau^{4}}. (14)

To further compare expressions in (14) one can identify the pump pulse parameters (ωp\omega_{p}) and Ωp\Omega_{p} corresponding to limiting cases. For instance PmaxQ=PmaxsP_{max}^{Q*}=P_{max}^{s*} at cp=32τ2/3c_{p^{\prime}}^{*}=32\tau^{2}/3 which corresponds to Ωp=4Tc[Γ2Γc/(3ωcωp)]1/2\Omega_{p}^{*}=4T_{c}[\Gamma_{2}\Gamma_{c}/(3\omega_{c}\omega_{p})]^{1/2}. Furthermore, by setting cp=16τ2/3c_{p^{\prime}}^{**}=16\tau^{2}/3 (i.e. Ωp=Ωp/2\Omega_{p}^{**}=\Omega_{p}^{*}/\sqrt{2}) we obtain PmaxQ=2Pmaxs=32/9P_{max}^{Q**}=2P_{max}^{s**}=32/9 which corresponds to the maximum of the difference between the two expressions. Thus, in the weak coupling regime one can also achieve the degree of control over the pump-probe signal by identifying the constrained parameter space for the pump field.

In summary, we have developed a novel method for control of the optical signals based on the analogy with quantum heat engines where energy transfer occurs from the pump to the probe field. We found that the yield of the spectroscopic measurement can be improved when the corresponding regime is close to the thermodynamic cycle. The proposed model can provide a reasonable qualitative guidance for the experimental implication in molecular systems consisting of two electronic states where the process of coherent excitation and consequent relaxation can be viewed as an effective thermal heat bath environment. This apparent connection between the thermodynamics of the QHE and the spectroscopy may emerge as a novel “incoherent control” tool for optimization of the optical measurements, which can enhance the yield of the fluorescence, the pump-probe measurements, and improve the signal-to-noise ratio in a wide class of the optical signals. The rich physics of the QHE can be extended to the Otto-cycle engines mimicking the time-domain nonlinear optical signals which will be the subject of the future studies.

We gratefully acknowledge the support from the National Science Foundation of China (No. 11934011), the Zijiang Endowed Young Scholar Fund, the East China Normal University and the Overseas Expertise Introduction Project for Discipline Innovation (111 Project, B12024). MQ thanks P. Saurabh for valuable discussion and acknowledges the support from CSC Scholarship (CSC No. 2018 DFH 007778).

Appendix A: Derivation of the dissipative Lindblad superoperator

We consider a four-level system interacting resonantly with the classical pump and probe fields. Four-level system is in contact with the ambient reservoir at room temperature TcT_{c}. This thermal reservoir drives the relaxation between states 2-1 and 0-g. The Hamiltonian for the system-bath interaction for the transition 212-1 reads

V^(t)=gkb^kei(ωνk)t|21|+H.c.,\displaystyle\hat{V}(t)=\hbar\sum\text{g}_{k}\hat{b}_{k}e^{i(\omega-\nu_{k})t}|2\rangle\langle 1|+\text{H.c.,} (A1)

and the Hamiltonian for the cold bath interacting with 0g0-g transition is given by

𝒱^(t)=Gqb^qei(ωνq)t|0g|+H.c.,\displaystyle\hat{\mathcal{V}}(t)=\hbar\sum\text{G}_{q}\hat{b}_{q}e^{i(\omega-\nu_{q})t}|0\rangle\langle g|+\text{H.c.,} (A2)

where the system-bath coupling constant gk=𝒫21.ϵ^kk/\text{g}_{k}=\mathcal{P}_{21}.\hat{\epsilon}_{k}\mathcal{E}_{k}/\hbar and Gq=𝒫og.ϵ^qq/\text{G}_{q}=\mathcal{P}_{og}.\hat{\epsilon}_{q}\mathcal{E}_{q}/\hbar with the dipole moment for transition 212\leftrightarrow 1 and 0g0\leftrightarrow g is given by 𝒫21\mathcal{P}_{21} and 𝒫og\mathcal{P}_{og}, respectively, and the polarization of the field is denoted as ϵk^\hat{\epsilon_{k}} and ϵg^\hat{\epsilon_{g}}, respectively. The electric field per photon is k=(νk/2ϵ0Vph)1/2\mathcal{E}_{k}=(\hbar\nu_{k}/2\epsilon_{0}V_{ph})^{1/2} and q=(νq/2ϵ0Vph)1/2\mathcal{E}_{q}=(\hbar\nu_{q}/2\epsilon_{0}V_{ph})^{1/2} respectively, where VphV_{ph} is the photon quantization volume. We assume that the system interacts with the reservoir represented by the reservoir density operator ρR\rho_{R}. The equation of motion for the system density operator ρ\rho is given by

21[ρ]\displaystyle\mathcal{L}_{21}[\rho] =\displaystyle= iTrR[V(t),ρ(t0)ρR]\displaystyle-\frac{i}{\hbar}\text{Tr}_{R}\left[V(t),\rho(t_{0})\otimes\rho_{R}\right] (A3)
\displaystyle- 12TrRt0t[V(t)[V(t),ρ(t)ρR(t0)]]𝑑t.\displaystyle\frac{1}{\hbar^{2}}\text{Tr}_{R}\int_{t_{0}}^{t}\left[V(t)\left[V(t^{\prime}),\rho(t^{\prime})\otimes\rho_{R}(t_{0})\right]\right]dt^{\prime}.\ \ \ \ \ \

We note the bath operator represents the thermal state, i.e. b^k=b^k=0,b^kb^k=b^kb^k=0,b^kb^k=n¯kδkk\langle\hat{b}_{\textbf{k}}\rangle=\langle\hat{b}^{\dagger}_{\textbf{k}}\rangle=0,\langle\hat{b}_{\textbf{k}}\hat{b}_{\textbf{k}^{\prime}}\rangle=\langle\hat{b}^{\dagger}_{\textbf{k}}\hat{b}^{\dagger}_{\textbf{k}^{\prime}}\rangle=0,\langle\hat{b}^{\dagger}_{\textbf{k}}\hat{b}_{\textbf{k}^{\prime}}\rangle=\bar{n}_{\textbf{k}}\delta_{\textbf{kk}^{\prime}} and b^kb^k=(n¯k+1)δkk\langle\hat{b}_{\textbf{k}}\hat{b}^{\dagger}_{\textbf{k}^{\prime}}\rangle=(\bar{n}_{\textbf{k}}+1)\delta_{\textbf{kk}^{\prime}} where n¯𝐤=[exp(βcϵ𝐤)1]1\bar{n}_{\mathbf{k}}=[\exp(\beta_{c}\epsilon_{\mathbf{k}})-1]^{-1} is the phonon occupation number corresponding to the cold temperature Tc=βc1T_{c}=\beta_{c}^{-1}. Inserting V(t)^\hat{V(t)} in Eq. (A3), we obtain

21[ρ]\displaystyle\mathcal{L}_{21}[\rho] =\displaystyle= 12t0tdtkkgkgk{(X+X+ρ(t)2X+\displaystyle-\frac{1}{\hbar^{2}}\int_{t_{0}}^{t}dt^{\prime}\sum_{kk^{\prime}}\text{g}_{k}\text{g}_{k^{\prime}}\big{\{}(\text{X}_{+}\text{X}_{+}\rho(t^{\prime})-2\text{X}_{+} (A4)
ρ(t)X++ρ(t)X+X+)bkbkei(ωνk)t+\displaystyle\rho(t^{\prime})\text{X}_{+}+\rho(t^{\prime})\text{X}_{+}\text{X}_{+})\langle b_{k}b_{k^{\prime}}\rangle e^{i(\omega-\nu_{k})t+}
+i(ωνk)t(X+Xρ(t)Xρ(t)X+)bkbk\displaystyle{}^{i(\omega-\nu_{k^{\prime}})t^{\prime}}+(\text{X}_{+}\text{X}_{-}\rho(t^{\prime})-\text{X}_{-}\rho(t^{\prime})\text{X}_{+})\langle b_{k}b^{\dagger}_{k^{\prime}}\rangle
ei(ωνk)ti(ωνk)t+(XX+ρ(t)X+ρ(t)\displaystyle e^{i(\omega-\nu_{k})t-i(\omega-\nu_{k^{\prime}})t^{\prime}}+(\text{X}_{-}\text{X}_{+}\rho(t^{\prime})-\text{X}_{+}\rho(t^{\prime})
X)bkbkei(ωνk)t+i(ωνk)t}+H.c.,\displaystyle\text{X}_{-})\langle b^{\dagger}_{k}b_{k}\rangle e^{-i(\omega-\nu_{k})t+i(\omega-\nu_{k})t}\big{\}}+\text{H.c.},

where X+=|21|X_{+}=|2\rangle\langle 1| and X=|12|X_{-}=|1\rangle\langle 2|. On substituting the various expectation values from above paragraph in Eq. (A4), we obtain

21[ρ]\displaystyle\mathcal{L}_{21}[\rho] =\displaystyle= 12t0tdtkgk2{nk(|11|ρ(t)+ρ(t)\displaystyle-\frac{1}{\hbar^{2}}\int_{t_{0}}^{t}dt^{\prime}\sum_{k}\text{g}^{2}_{k}\big{\{}n_{k}\big{(}|1\rangle\langle 1|\rho(t^{\prime})+\rho(t^{\prime}) (A5)
|11|2|22|ρ11)ei(ωνk)(tt)+(nk+1)\displaystyle|1\rangle\langle 1|-2|2\rangle\langle 2|\rho_{11}\big{)}e^{-i(\omega-\nu_{k})(t-t^{\prime})}+(n_{k}+1)
(|22|ρ(t)+ρ(t)|22|2|11|ρ22)\displaystyle\big{(}|2\rangle\langle 2|\rho(t^{\prime})+\rho(t^{\prime})|2\rangle\langle 2|-2|1\rangle\langle 1|\rho_{22}\big{)}
ei(ωνk)(tt)}.\displaystyle e^{i(\omega-\nu_{k})(t-t^{\prime})}\big{\}}.

The sum over k is replaced by an integral through prescription

kVphπ20𝑑kk2.\sum_{\textbf{k}}\rightarrow\frac{V_{ph}}{\pi^{2}}\int_{0}^{\infty}dkk^{2}. (A6)

We can assume that the density matrix varying slowly with the time and using the Markov approximation ρ(t)ρ(t)\rho(t^{\prime})\approx\rho(t). We then extend the upper limit of integration to infinity. Then the time integration yields

t0𝑑tei(ω2νk)(tt)=πδ(ωνk)+iP1ωνk,\int_{t_{0}}^{\infty}dt^{\prime}e^{i(\omega_{2}-\nu_{k})(t-t^{\prime})}=\pi\delta(\omega-\nu_{k})+i\text{P}\frac{1}{\omega-\nu_{k}}, (A7)

where P denotes the Cauchy principal value. The above expression split the correlation into real and imaginary part to obtain the decays rate and Lamb shift, respectively. So, the Cauchy part will not affect our result and for simplicity, we consider the transition between the atomic levels only. In the Weisskopf-Wigner approximation (Born-Markov plus RWA approximations), we replace gkgk0\text{g}_{k}\approx\text{g}_{k0} and nknk0=n2n_{k}\approx n_{k0}=n_{2} assuming broadband coupling, and obtain the final form of the Liouville operator driving 2-1 transition

21[ρ]\displaystyle\mathcal{L}_{21}[\rho] =\displaystyle= γ2(n2+1)(|11|ρ22|22|ρ+ρ|22|2)\displaystyle\gamma_{2}(n_{2}+1)\big{(}|1\rangle\langle 1|\rho_{22}-\frac{|2\rangle\langle 2|\rho+\rho|2\rangle\langle 2|}{2}\big{)} (A8)
+\displaystyle+ γ2n2(|22|ρ11|11|ρ+ρ|11|2),\displaystyle\gamma_{2}n_{2}\big{(}|2\rangle\langle 2|\rho_{11}-\frac{|1\rangle\langle 1|\rho+\rho|1\rangle\langle 1|}{2}\big{)},

where γ2=2k0Vphgk02πc\gamma_{2}=\frac{2k_{0}V_{ph}\text{g}^{2}_{k0}}{\pi c}. Similarly we derive the c[ρ]\mathcal{L}_{c}[\rho] describing interaction with cold bath of the 0-g transition, which yields

c[ρ]\displaystyle\mathcal{L}_{c}[\rho] =\displaystyle= γc(nc+1)(2|gg|ρ00|00|ρρ|00|)\displaystyle\gamma_{c}(n_{c}+1)\big{(}2|g\rangle\langle g|\rho_{00}-|0\rangle\langle 0|\rho-\rho|0\rangle\langle 0|\big{)} (A9)
+\displaystyle+ γcnc(2|00|ρgg|gg|ρρ|gg|),\displaystyle\gamma_{c}n_{c}\big{(}2|0\rangle\langle 0|\rho_{gg}-|g\rangle\langle g|\rho-\rho|g\rangle\langle g|\big{)},

where γc=4q0Vphgq02πc\gamma_{c}=\frac{4q_{0}V_{ph}\text{g}^{2}_{q0}}{\pi c}, has an extra factor of 22 compared with γ2\gamma_{2} to be consistent with the earlier works kpra18 . Using equations (A8) and (A9), we obtain Eq. (Incoherent control of optical signals; quantum heat engine approach).

Appendix B: Effective thermal bath

The two electronic energy levels of the molecule shown in Fig. 1. Two vibrational states of the electronic ground state gg and 0 while 11 and 22 are two vibrational state of the electronically excited state. Coherent pump pulse excites the transition g2g-2 with Rabi frequency Ωp\Omega_{p}. The corresponding equation of motion for the density matrix is given in Eq. (Incoherent control of optical signals; quantum heat engine approach)(coherent bath) and steady state solution for the populations of gg, 11 and 22 states are given by

ρggc(ss)=ρ22c(ss)=n23n2+1,ρ11c(ss)=n2+13n2+1.\displaystyle\rho_{gg}^{{}^{\prime}c(ss)}=\rho_{22}^{{}^{\prime}c(ss)}=\frac{n_{2}}{3n_{2}+1},\ \ \rho_{11}^{{}^{\prime}c(ss)}=\frac{n_{2}+1}{3n_{2}+1}.\ \ \ (B1)

In order to eliminate the state 22 we normalize the above solution for the population of states gg and 11 such that ρggc(ss)+ρ11c(ss)=1\rho_{gg}^{c(ss)}+\rho_{11}^{c(ss)}=1, we then obtain

ρggc(ss)=n22n2+1,ρ11c(ss)=n2+12n2+1.\displaystyle\rho_{gg}^{c(ss)}=\frac{n_{2}}{2n_{2}+1},\quad\rho_{11}^{c(ss)}=\frac{n_{2}+1}{2n_{2}+1}. (B2)

Similarly we obtain steady state solution for the thermal bath using Eq. (5):

ρggth(ss)=n0+12n0+1,ρ11th(ss)=n02n0+1.\displaystyle\rho_{gg}^{th(ss)}=\frac{n_{0}+1}{2n_{0}+1},\quad\rho_{11}^{th(ss)}=\frac{n_{0}}{2n_{0}+1}. (B3)

We then solve Eq.(Incoherent control of optical signals; quantum heat engine approach) non perturbatively over the Ωp\Omega_{p} by factorizing Eqs. (1) into two uncoupled system of equations: one part containing the ground state population ρgg\rho_{gg} and coherences ρg2\rho_{g2} and ρ2g\rho_{2g} and another containing the excited state populations ρ11\rho_{11} and ρ22\rho_{22}. Since electronic coherence ρg2\rho_{g2} evolves fast, we can assume stationary population for the excited state 22 while keeping the ground state evolution exactly. To have correct solution in the steady state we match the value of the stationary population for the excited state 22 with its steady state solution: ρ22ρ22c(ss)\rho_{22}\to\rho_{22}^{{}^{\prime}c(ss)} while initial conditions are ρgg(0)=1\rho_{gg}(0)=1, ρg2(0)=ρ2g(0)=0\rho_{g2}(0)=\rho_{2g}(0)=0. In this case solution for the coherences yields:

ρg2c(t)=4i(2n2+1)Ωpe14Γ2(n2+1)tsinh(14t𝒵)(3n2+1)𝒵,\displaystyle\rho_{g2}^{c}(t)=\frac{4i\left(2n_{2}+1\right)\Omega_{p}e^{-\frac{1}{4}\Gamma_{2}\left(n_{2}+1\right)t}\sinh\left(\frac{1}{4}t\mathcal{Z}\right)}{\left(3n_{2}+1\right)\mathcal{Z}}, (B4)

where 𝒵=Γ22(n2+1)232Ωp2\mathcal{Z}=\sqrt{\Gamma_{2}^{2}\left(n_{2}+1\right){}^{2}-32\Omega_{p}^{2}}. while the ground state population reads (assuming 42ΩpΓ2(n2+1)4\sqrt{2}\Omega_{p}\ll\Gamma_{2}(n_{2}+1)

ρggc(t)=n23n2+1+(2n2+1)eΓ~t3n2+1,\displaystyle\rho_{gg}^{c}(t)=\frac{n_{2}}{3n_{2}+1}+\frac{\left(2n_{2}+1\right)e^{-\tilde{\Gamma}t}}{3n_{2}+1}, (B5)

We now calculate the population of states 11 and 22 using the solutions for ρg2\rho_{g2} and ρ2g\rho_{2g} given by Eq. (B4), and initial conditions are ρ11(0)=ρ22(0)=0\rho_{11}(0)=\rho_{22}(0)=0 which yields

ρ11(t)=(n2+1)(1eΓ~t)3n2+1,ρ22(t)=n2(1eΓ~t)3n2+1.\displaystyle\rho_{11}(t)=\frac{\left(n_{2}+1\right)\left(1-e^{-\tilde{\Gamma}t}\right)}{3n_{2}+1},\rho_{22}(t)=\frac{n_{2}\left(1-e^{-\tilde{\Gamma}t}\right)}{3n_{2}+1}. (B6)

Appendix C: Three level QHE

A detailed derivation of the matrix equations for a three level atom and induced coherence given in the Supplementary of sd11 . Here we consider a three level system g,0g,0 and 11 and the Hamiltonian for the three level system is given by

H0=i=g,0,1ωi|ii|,H_{0}=\sum_{i=g,0,1}\omega_{i}|i\rangle\langle i|, (C1)

while the probe-system interaction is described by kpra18

V(t)=λ(eiωt|10|+eiωt|01|),V(t)=\lambda(e^{i\omega t}|1\rangle\langle 0|+e^{-i\omega t}|0\rangle\langle 1|), (C2)

where λ\lambda is a probe-matter coupling. Transition 0g0-g is in contact with the cold reservoir at temperature TcT_{c}. The corresponding Liouville operator for the cold bath-system interaction is given by Eq. (A9).

Effective hot bath at temperature ThT_{h} is driving g1g-1 transition described by the Liouville operator

h[ρ]\displaystyle\mathcal{L}_{h}\left[\rho\right] =\displaystyle= Γh(nh+1)[2|gg|ρ11|11|ρρ|11|]\displaystyle\Gamma_{h}(n_{h}+1)\left[2|g\rangle\langle g|\rho_{11}-|1\rangle\langle 1|\rho-\rho|1\rangle\langle 1|\right] (C3)
+\displaystyle+ Γhnh[2|11|ρgg|gg|ρρ|gg|].\displaystyle\Gamma_{h}n_{h}\left[2|1\rangle\langle 1|\rho_{gg}-|g\rangle\langle g|\rho-\rho|g\rangle\langle g|\right].

Here the average occupation numbers are given by

nc=(eωcTc1)1,nh=(eωhTh1)1.n_{c}=(e^{\frac{\omega_{c}}{T_{c}}}-1)^{-1},n_{h}=(e^{\frac{\omega_{h}}{T_{h}}}-1)^{-1}. (C4)

The evolution of the density matrix reads

ϱ00˙\displaystyle\dot{\varrho_{00}} =\displaystyle= iλ(ρ10ρ01)2Γc(1+nc)ρ00+2Γcncρgg,\displaystyle-i\lambda(\rho_{10}-\rho_{01})-2\Gamma_{c}(1+n_{c})\rho_{00}+2\Gamma_{c}n_{c}\rho_{gg},
ϱ11˙\displaystyle\dot{\varrho_{11}} =\displaystyle= iλ(ρ10ρ01)2Γh(1+nh)ρ11+2Γhnhρgg,\displaystyle i\lambda(\rho_{10}-\rho_{01})-2\Gamma_{h}(1+n_{h})\rho_{11}+2\Gamma_{h}n_{h}\rho_{gg},
ϱ01˙\displaystyle\dot{\varrho_{01}} =\displaystyle= iΔρ01iλ(ρ11ρ00)Γc(1+nc)ρ01\displaystyle-i\Delta\rho_{01}-i\lambda(\rho_{11}-\rho_{00})-\Gamma_{c}(1+n_{c})\rho_{01}
\displaystyle- Γh(1+nh)ρ01,\displaystyle\Gamma_{h}(1+n_{h})\rho_{01},
ϱ10˙\displaystyle\dot{\varrho_{10}} =\displaystyle= iΔρ01+iλ(ρ11ρ00)Γc(1+nc)ρ10\displaystyle i\Delta\rho_{01}+i\lambda(\rho_{11}-\rho_{00})-\Gamma_{c}(1+n_{c})\rho_{10} (C5)
\displaystyle- Γh(1+nh)ρ10,\displaystyle\Gamma_{h}(1+n_{h})\rho_{10},

where Δ=ω+ω0ω1\Delta=\omega+\omega_{0}-\omega_{1} is the detuning between the output laser radiation and molecular level transition; nhn_{h} and ncn_{c}, Γh\Gamma_{h} and Γc\Gamma_{c} are the average occupation numbers and dephasing rates for the hot and cold baths, respectively.
Solving Eqs.(C5) in the steady state assuming Δ=0\Delta=0 we obtain for the power and efficiency defined by Eq. (8) as well as the heat fluxkpra18

P=iλ(ωcωh)(ϱ01ϱ10),\displaystyle P=i\hbar\lambda(\omega_{c}-\omega_{h})(\varrho_{01}-\varrho_{10}),
Q˙h=iωhλ(ϱ01ϱ10),\displaystyle\dot{Q}_{h}=i\hbar\omega_{h}\lambda(\varrho_{01}-\varrho_{10}),
η=1ωcωh.\displaystyle\eta=1-\frac{\omega_{c}}{\omega_{h}}. (C6)

where ωh=ω1ωg\omega_{h}=\omega_{1}-\omega_{g} and ωc=ω0ωg\omega_{c}=\omega_{0}-\omega_{g}. After solving Eq. (C5) the output power reads

P=23λ2ΓhΓc(ncnh)(ωcωh)(Γhnh+Γcnc)(λ2+ΓhΓcncnh).P=\frac{2}{3}\frac{\lambda^{2}\Gamma_{h}\Gamma_{c}(n_{c}-n_{h})(\omega_{c}-\omega_{h})}{(\Gamma_{h}n_{h}+\Gamma_{c}n_{c})(\lambda^{2}+\Gamma_{h}\Gamma_{c}n_{c}n_{h})}. (C7)

In the high temperature limit nc=n1Tc/ωcn_{c}=n_{1}\simeq T_{c}/\omega_{c}, n2Tc/ω21n_{2}\simeq T_{c}/\omega_{21} (both n2n_{2} and ncn_{c} have the same ”cold” temperature since they are related to the same ambient (phonon) environment). Introducing dimensionless parameters: ωh=cωc\omega_{h}=c\omega_{c} and c21=(cpc)ωcc_{21}=(c_{p}-c)\omega_{c} where cp=ωp/ωcc_{p}=\omega_{p}/\omega_{c} and c21=ω21/ωcc_{21}=\omega_{21}/\omega_{c}, the effective hot bath temperature Th=ωc2Ω2/γ2ΓcT_{h}=\omega_{c}\sqrt{2\Omega^{2}/\gamma_{2}\Gamma_{c}} Eq. (C7) reads

𝒫=2P0λ2c21(cp1)(1c21)3(λ2+c21)(τ2+c21),\mathcal{P}=\frac{2P_{0}\lambda^{{}^{\prime}2}c_{21}(c_{p}-1)(1-c_{21})}{3\left(\lambda^{{}^{\prime}2}+c_{21}\right)\left(\tau^{2}+c_{21}\right)}, (C8)

where P0=ΓcωcP_{0}=\Gamma_{c}\omega_{c}, λ=λωc/(ΓcTh)\lambda^{\prime}=\lambda\omega_{c}/(\Gamma_{c}T_{h}), τ=Tc/Th\tau=T_{c}/T_{h} and we assume that cpc21c_{p}\gg c_{21} valid for e.g. visible range for the pump field and IR phonon range. Using the above dimensionless parameters we recast the efficiency in Eq. (Incoherent control of optical signals; quantum heat engine approach) as

η=11cpc21.\displaystyle\eta=1-\frac{1}{c_{p}-c_{21}}. (C9)

Further maximization of Eq. (C8) with respect to c21c_{21} yields Eq. (9). The corresponding efficiency at maximum power is given by Eq. (10).

Appendix D: Perturbative pump-probe signal

Refer to caption
Figure D1: (color online) The double sided Feynman-diagram representing the pump-probe signal. There are total six pathways that contribute to the pump-probe signal (diagrams a, b, c and their complex conjugate).

The pump-probe signal is given by Eq. (12). The response function can be read off the diagrams in Fig. D1, which yields

a(ωpr,\displaystyle\mathcal{R}_{a}(-\omega_{pr^{\prime}}, ωpr,ωp,ωp)=\displaystyle\omega_{pr},-\omega_{p^{\prime}},\omega_{p})=
=𝒱00,10𝒢10,10(ωpr+ωpωp)𝒱10,11𝒢11,22(ωpωp)𝒱22,2g𝒢2g,2g(ωp)𝒱2g,gg,\displaystyle=\langle\mathcal{V}_{00,10}\mathcal{G}_{10,10}({\omega_{pr}}+\omega_{p}-\omega_{p^{\prime}})\mathcal{V}_{10,11}\mathcal{G}_{11,22}(\omega_{p}-\omega_{p^{\prime}})\mathcal{V}_{22,2g}\mathcal{G}_{2g,2g}({\omega_{p}})\mathcal{V}_{2g,gg}\rangle, (D1)
b(ωpr,\displaystyle\mathcal{R}_{b}(-\omega_{pr^{\prime}}, ωpr,ωp,ωp)=\displaystyle\omega_{pr},\omega_{p^{\prime}},-\omega_{p})=
=𝒱00,10𝒢10,10(ωprωp+ωp)𝒱10,11𝒢11,22(ωp+ωp)𝒱22,g2𝒢g2,g2(ωp)𝒱g2,gg,\displaystyle=\langle\mathcal{V}_{00,10}\mathcal{G}_{10,10}({\omega_{pr}}-\omega_{p}+\omega_{p^{\prime}})\mathcal{V}_{10,11}\mathcal{G}_{11,22}(-\omega_{p}+\omega_{p^{\prime}})\mathcal{V}_{22,g2}\mathcal{G}_{g2,g2}(-{\omega_{p}})\mathcal{V}_{g2,gg}\rangle, (D2)
c(ωpr,\displaystyle\mathcal{R}_{c}(-\omega_{pr^{\prime}}, ωpr,ωp,ωp)=\displaystyle\omega_{pr},\omega_{p^{\prime}},-\omega_{p})=
=𝒱00,10𝒢10,10(ωprωp+ωp)𝒱10,00𝒢00,gg(ωp+ωp)𝒱gg,g2𝒢g2,g2(ωp)𝒱g2,gg.\displaystyle=\langle\mathcal{V}_{00,10}\mathcal{G}_{10,10}({\omega_{pr}}-\omega_{p}+\omega_{p^{\prime}})\mathcal{V}_{10,00}\mathcal{G}_{00,gg}(-\omega_{p}+\omega_{p^{\prime}})\mathcal{V}_{gg,g2}\mathcal{G}_{g2,g2}(-{\omega_{p}})\mathcal{V}_{g2,gg}\rangle. (D3)

Perturbative result in Eqs.(D1) - (D3) contains both population and coherence Green’s functions. The coherence Green’s functions for 1-0 transition is given by 𝒢10,10(ω)=[i(ωω10)Γ10]1\mathcal{G}_{10,10}(\omega)=-[i(\omega-\omega_{10})-\Gamma_{10}]^{-1}, for gg-2 transition 𝒢2g,2g(ω)=[i(ωω2g)Γ2g]1\mathcal{G}_{2g,2g}(\omega)=-[i\left(\omega-\omega_{2g}\right)-\Gamma_{2g}]^{-1}, where Γ10=[Γc(nc+1)+Γ2n2]/2\Gamma_{10}=[\Gamma_{c}(n_{c}+1)+\Gamma_{2}n_{2}]/2 and Γ2g=[Γcnc+Γ2(n2+1)]/2\Gamma_{2g}=[\Gamma_{c}n_{c}+\Gamma_{2}(n_{2}+1)]/2. The population Green’s function is a solution of the coupled transport (relaxation) equations:

ρ22˙\displaystyle\dot{\rho_{22}} =\displaystyle= Γ2(n2+1)ρ22+Γ2n2ρ11,\displaystyle-\Gamma_{2}(n_{2}+1)\rho_{22}+\Gamma_{2}n_{2}\rho_{11},
ρ11˙\displaystyle\dot{\rho_{11}} =\displaystyle= Γ2(n2+1)ρ22Γ2n2ρ11,\displaystyle\Gamma_{2}(n_{2}+1)\rho_{22}-\Gamma_{2}n_{2}\rho_{11}, (D4)
ρ00˙\displaystyle\dot{\rho_{00}} =\displaystyle= Γc(nc+1)ρ00+Γcncρgg,\displaystyle-\Gamma_{c}(n_{c}+1)\rho_{00}+\Gamma_{c}n_{c}\rho_{gg},
ρgg˙\displaystyle\dot{\rho_{gg}} =\displaystyle= Γc(nc+1)ρ00Γcncρgg.\displaystyle\Gamma_{c}(n_{c}+1)\rho_{00}-\Gamma_{c}n_{c}\rho_{gg}. (D5)

Eqs. (D4) - (D5) can be recast as a Pauli master equation:

ρ˙ii(t)=ii,jjκii,jjρjj(t),\displaystyle\dot{\rho}_{ii}(t)=-\sum_{ii,jj}\kappa_{ii,jj}\rho_{jj}(t), (D6)

where, κii,jj\kappa_{ii,jj} is the population transport matrix. In Eq. D6, the diagonal elements, i=ji=j, κii,ii\kappa_{ii,ii} are positive, whereas the off-diagonal elements, iji\neq j, κii,jj\kappa_{ii,jj} are negative. The population transport matrix satisfies the population conservation: iκii,jj=0\sum_{i}\kappa_{ii,jj}=0. The evolution of the diagonal elements is defined by the population Green function, ρjj(t)=i𝒢jj,ii(t)ρii(0)\rho_{jj}(t)=\sum_{i}\mathcal{G}_{jj,ii}(t)\rho_{ii}(0). where 𝒢jj,ii(t)\mathcal{G}_{jj,ii}(t) is given mcr9

𝒢jj,ii(t)=nξjn(R)Dnn1exp(λnt)ξni(L),\displaystyle\mathcal{G}_{jj,ii}(t)=\sum_{n}\xi_{jn}^{(R)}D_{nn}^{-1}\exp(-\lambda_{n}t)\xi_{ni}^{(L)}, (D7)

where λn\lambda_{n} is the nthnth eigenvalue of left and right eigenvector (ξn(L),ξn(R))(\xi_{n}^{(L)},\xi_{n}^{(R)}) and D=ξLξRD=\xi^{L}\xi_{R} is a diagonal matrix. Using Eq. (D7) we obtain for the population Green’s functions:

𝒢00,gg(t)\displaystyle\mathcal{G}_{00,gg}(t) =\displaystyle= nc(1et(1+2nc)Γc)(1+2nc),\displaystyle\frac{n_{c}(1-e^{-t(1+2n_{c})\Gamma_{c}})}{(1+2n_{c})}, (D8)
𝒢11,22(t)\displaystyle\mathcal{G}_{11,22}(t) =\displaystyle= (1+n2)(1et(1+2n2)Γ2)(1+2n2).\displaystyle\frac{(1+n_{2})(1-e^{-t(1+2n_{2})\Gamma_{2}})}{(1+2n_{2})}. (D9)

Assuming narrowband resonant pump and probe fields ωp=ωp=ω2g\omega_{p^{\prime}}=\omega_{p}=\omega_{2g} and ωpr=ωpr=ω10\omega_{pr^{\prime}}=\omega_{pr}=\omega_{10} Eqs. (D1) - (D3) read

a=b=4|μ10|2|μ2g|2(1+n2)(ncΓc+n2Γ2)2(1+2n2)σp,\displaystyle\mathcal{R}_{a}=\mathcal{R}_{b}=\frac{4|\mu_{10}|^{2}|\mu_{2g}|^{2}(1+n_{2})}{(n_{c}\Gamma_{c}+n_{2}\Gamma_{2})^{2}(1+2n_{2})\sigma_{p}}, (D10)
c=4nc|μ10|2|μ2g|2(ncΓc+n2Γ2)2(1+2nc)σp,\displaystyle\mathcal{R}_{c}=\frac{4n_{c}|\mu_{10}|^{2}|\mu_{2g}|^{2}}{(n_{c}\Gamma_{c}+n_{2}\Gamma_{2})^{2}(1+2n_{c})\sigma_{p}}, (D11)

where σp\sigma_{p} is an infinitesimal parameter of the order of the pump bandwidth required for the convergence of the Fourier transform. Taking high temperature limit nc1n_{c}\gg 1, n21n_{2}\gg 1 the total power (12) reads

Ps=12(ωhωc)λ2Ωp2(ncΓc+n2Γ2)2σp,\displaystyle P^{s}=\frac{12(\omega_{h}-\omega_{c})\lambda^{2}\Omega_{p}^{2}}{(n_{c}\Gamma_{c}+n_{2}\Gamma_{2})^{2}\sigma_{p}}, (D12)

where we used ω10=ωhωc\omega_{10}=\omega_{h}-\omega_{c}, λ=μ10pr\lambda=\mu_{10}\mathcal{E}_{pr}, Ωp=μ2gp\Omega_{p}=\mu_{2g}\mathcal{E}_{p}. Expanding occupation numbers in the high temperature limit nc=Tc/ωcn_{c}=T_{c}/\omega_{c}, n2=Tc/ω21n_{2}=T_{c}/\omega_{21} and maximizing the power with respect to c21=ω21/ωcc_{21}=\omega_{21}/\omega_{c} we obtain Eq. (13).

References

  • (1) Scully MO, Chapin KR, Dorfman KE, Kim MB, Svidzinsky A (2011) Quantum heat engine power can be increased by noise-induced coherence. Proc Natl. Acad Sci USA 108:15097–15100.
  • (2) K. E. Dorfman, D. V. Voronine, S. Mukamel, and M. O. Scully (2013) Photosynthetic reaction center as a quantum heat engine. Proc. Natl. Acad. Sci. U.S.A. 110, 2746 (2013).
  • (3) Scovil HED, Schulz-DuBois EO (1959) Three-level masers as heat engines. Phys Rev Lett 2:262–263.
  • (4) Geusic JE, Schulz-DuBois EO, De Grasse RW, Scovil HED (1959) Three level spin refrigeration and maser action at 1500 mc/sec. J Appl Phys 30:1113–1114.
  • (5) Geusic JE, Schulz-DuBois EO, Scovil HED (1967) Quantum equivalent of the carnot cycle. Phys Rev 156:343–351.
  • (6) Alicki R (1979) The quantum open system as a model of the heat engine. J Phys A 12:L103–L107.
  • (7) E. Boukobza and D. J. Tannor, Phys. Rev. Lett.98, 240601 (2007).
  • (8) E.Boukobza and D.J.Tannor, Phys.Rev.A74, 063823 (2006).
  • (9) R. Uzdin, A. Levy, and R. Kosloff, Equivalence of Quantum Heat Machines, and Quantum-Thermodynamic Signatures, Phys. Rev. X 5, 031044 (2015).
  • (10) B. K. Agarwalla, J.-H. Jiang, and D. Segal, Quantum efficiency bound for continuous heat engines coupled to noncanonical reservoirs, Phys. Rev. B 96, 104304 (2017).
  • (11) Sheng-Wen Li, Moochan B. Kim, Girish S. Agarwal, and Marlan O. Scully, Quantum statistics of a single-atom Scovil–Schulz-DuBois heat engine Physical review A 96, 063806 (2017).
  • (12) M. O. Scully, Quantum Photocell: Using Quantum Coherence to Reduce Radiative Recombination and Increase Efficiency, Phys. Rev. Lett. 104, 207701 (2010).
  • (13) F. Chen, Y. Gao, and M. Galperin, Molecular Heat Engines: Quantum Coherence Effects Entropy 19, 472 (2017).
  • (14) M.O. Scully, M.S. Zubairy, G.S. Agarwal, and H. Walther, Extracting work from a single heat bath via vanishing quantum coherence, Science 299, 862 (2003).
  • (15) Johannes Roßnagel, Samuel T. Dawkins, Karl N. Tolazzi, Obinna Abah, Eric Lutz, Ferdinand Schmidt-Kaler, and Kilian Singer, “A singleatom heat engine,” Science 352, 325–329 (2016).
  • (16) S. Deng, A. Chenu, P. Diao, F. Li, S. Yu, I. Coulamy, A. del Campo, and H. Wu, “Superadiabatic quantum friction suppression in finite-time thermodynamics,”Science Advances 4 (2018).
  • (17) J. P. S. Peterson, T. B. Batalhão, M. Herrera, A. M. Souza, R. S. Sarthour, I. S. Oliveira, and R. M. Serra, “Experimental characterization of a spin quantum heat engine,” Phys. Rev. Lett. 123, 240601 (2019).
  • (18) J. Klatzow, J. N. Becker, P. M. Ledingham, C. Weinzetl, K. T. Kaczmarek, D. J. Saunders, J. Nunn, I. A. Walmsley, R. Uzdin, and E. Poem, “Experimental Demonstration of Quantum Effects in the Operation of Microscopic Heat Engines,” Phys. Rev. Lett. 122, 110601 (2017).
  • (19) Poyatos, J. F., Cirac, J. I. and Zoller, P. Quantum reservoir engineering with laser cooled trapped ions. Phys. Rev. Lett. 77, 4728–4731 (1996).
  • (20) Kist T. B. L., Orszag M., Brun T. A. and Davidovich L., Stochastic Schro¨\ddot{o}dinger equations in cavity QED: physical interpretation and localization, J. Opt. B: Quantum Semiclass. Opt., 1 (1999) 251.
  • (21) Carvalho A. R. R., Milman P., de Matos Filho R. L. and Davidovich L., Decoherence, Pointer Engineering, and Quantum State Protection, Phys. Rev. Lett., 86 (2001) 4988.
  • (22) MF Santos, ARR Carvalho, Observing different quantum trajectories in cavity QED, EPL, 94, 64003 (2011).
  • (23) Zippilli S., Vitali D., Tombesi P. and Raimond J.-M., Scheme for decoherence control in microwave cavities, Phys. Rev. A, 67 052101 (2003).
  • (24) Verstraete F., Wolf M. M. and Ignacio Cirac J., Quantum computation and quantum-state engineering driven by dissipation, Nat. Phys., 5, 633(2009).
  • (25) Pielawa S., Davidovich L., Vitali D. and Morigi G., Engineering atomic quantum reservoirs for photons, Phys. Rev. A, 81, 043802 (2010).
  • (26) Andre R R Carvalho and Marcelo França Santos, Distant entanglement protected through artificially increased local temperature, New J. Phys. 13, 013010 (2011).
  • (27) K. E. Dorfman, D. Xu, and J. Cao, Efficiency at maximum power of a laser quantum heat engine enhanced by noise-induced coherence, Phys. Rev. E 97, 042120 (2018).
  • (28) Mukamel S (1995) Principles of Nonlinear Optical Spectroscopy (Oxford Univ Press, New York).
  • (29) D. Goswami, “Optical pulse shaping approaches to coherent control,” Physics Reports, vol. 374, no. 6, pp. 385 – 481, 2003.
  • (30) Y. Silberberg, Quantum Coherent Control for Nonlinear Spectroscopy and Microscopy, Annual Review of Physical Chemistry 60, 277 (2009).
  • (31) Moshe Shapiro and Paul Brumer (2011), Quantum Control of Molecular Processes, (WILEY-VCH Verlag GmbH and Co. KGaA).
  • (32) Alexander Pechen and Herschel Rabitz, Teaching the environment to control quantum systems, PhysRevA.73.062102(2006).
  • (33) Louis Garbe, Peregrine Wade, Fabrizio Minganti, Nathan Shammah, Simone Felicetti, Franco Nori, Scientific Reports volume 10, Article number: 13408 (2020).
  • (34) Stockton J K, van Handel R and Mabuchi H, Deterministic Dicke-state preparation with continuous measurement and control Phys. Rev. A 70 022106 (2004).
  • (35) Mancini S and Wang J 2005 Towards feedback control of entanglement Eur. Phys. J. D 32 257.
  • (36) Mancini S, Markovian feedback to control continuous-variable entanglement, Phys. Rev. A 73 010304 (2006).
  • (37) Englert B.-G. and Morigi G., Five lectures on dissipative master equations in Coherent Evolution in Noisy Environments, edited by Buchleitner A. and Hornberger K., Lect. Notes Phys., Vol. 611 (Springer, Berlin, Heidelberg) 2002, pp. 55–106.
  • (38) Massimiliano Esposito, Ryoichi Kawai, Katja Lindenberg, and Christian Van den Broeck, ”Efficiency at Maximum Power of Low-Dissipation Carnot Engines”, Phys. Rev. Lett. 105, 150603 (2010).
  • (39) David Gelbwaser-Klimovsky and Alán Aspuru-Guzik, Strongly Coupled Quantum Heat Machines, J. Phys. Chem. Lett. 2015, 6, 17, 3477–3482 (2015).
  • (40) D. Xu and J. Cao, Non-canonical distribution and non-equilibrium transport beyond weak system-bath coupling regime: A polaron transformation approach, Front. Phys. 11, 110308/1-17 (2016).
  • (41) D. C. Harris and M.D. Bertolucci ”Symmetry and Spectroscopy, An Introduction to Vibrational and Electronic Spectroscopy” Dover Publications, Inc., New York. 1989.
  • (42) R. Kosloff and A. Levy, Quantum heat engines and refrigerators: Continuous devices, Annual Review of Physical Chemistry 65 (2014), pp. 365–393.
  • (43) Y. Y. Xu, B. Chen, and J. Liu, Achieving the classical Carnot efficiency in a strongly coupled quantum heat engine, Phys. Rev. E 97, 022130 (2018).
  • (44) Abramavicius D, Palmieri B, Voronine DV, Sanda F, Mukamel S (2009) Coherent multidimensional optical spectroscopy of excitons in molecular aggregates; quasiparticle versus supermolecule perspectives. Chem Rev 109(6):2350–2408.