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

Cyclic three-level-pulse-area theorem for enantioselective state transfer of chiral molecules

Yu Guo Hunan Key Laboratory of Nanophotonics and Devices, School of Physics and Electronics, Central South University, Changsha 410083, China Hunan Provincial Key Laboratory of Flexible Electronic Materials Genome Engineering, School of Physics and Electronic Science, Changsha University of Science and Technology, Changsha 410114, China The Key Laboratory of Low-Dimensional Quantum Structures and Quantum Control of Ministry of Education, Hunan Normal University, Changsha 410081, China    Xun Gong Hunan Provincial Key Laboratory of Flexible Electronic Materials Genome Engineering, School of Physics and Electronic Science, Changsha University of Science and Technology, Changsha 410114, China    Songshan Ma Hunan Key Laboratory of Nanophotonics and Devices, School of Physics and Electronics, Central South University, Changsha 410083, China    Chuan-Cun Shu [email protected] he/him/his Hunan Key Laboratory of Nanophotonics and Devices, School of Physics and Electronics, Central South University, Changsha 410083, China
Abstract

We derive a pulse-area theorem for a cyclic three-level system, an archetypal model for exploring enantioselective state transfer (ESST) in chiral molecules driven by three linearly polarized microwave pulses. By dividing the closed-loop excitation into two separate stages, we obtain both amplitude and phase conditions of three control fields to generate high fidelity of ESST. As a proof of principle, we apply this pulse-area theorem to the cyclohexylmethanol molecules (C7H14O\text{C}_{7}\text{H}_{14}\text{O}), for which three rotational states are connected by the aa-type, bb-type, and cc-type components of the transition dipole moments in both center-frequency resonant and detuned conditions. Our results show that two enantiomers with opposite handedness can be transferred to different target states by designing three microwave pulses that satisfy the amplitude and phase conditions at the transition frequencies. The corresponding control schemes are robust against the time delays between the two stages. We suggest that the two control fields used in the second stage should be applied simultaneously for practical applications. This work contributes an alternative pulse-area theorem to the field of quantum control, which has the potential to determine the chirality of enantiomers in a mixture.

I Introduction

Since Louis Pasteur first reported molecular chirality in 1848 Pasteur1848 , the theoretical and experimental study of chiral molecules has drawn increasing interest because of its fundamental importance in modern chemical and biochemical industries as well as quantum science Quack2008 ; review1 ; review2 ; review3 . Two enantiomers of chiral molecules with opposite handedness have the same components and configuration for spatial reflection. It implies that distinguishing enantiomers from each other, highly related to the discrimination, separation, and purification of chiral molecules, remains a formidable task by comparing general physical and chemical properties, such as boiling points, melting points, and densities. Based on chemical mechanisms and enantiomer-specific interactions with auxiliary substances, many spectroscopic techniques were established to detect enantiomers of chiral molecules with different handedness Vogt ; Hazen2003 ; Patterson2014 . The traditional techniques, such as crystallization and chiral chromatography, are complicated and expensive and require significantly more protracted than seconds, leading to discrimination of chiral molecules out of reach.
By taking advantage of the sign difference property of optical rotations, it has become promising to select enantiomers by designing coherent quantum optical schemes Heppke1999 ; Brumer2001 ; Thanopulos2003 ; Frishman2004 ; Li2007 ; Li2008 ; Ye2018 ; Ye2019 ; Yachmenev2019 ; Wu2020 ; Torosov2020 ; Torosov2020-2 ; Ye2020 ; Wu2020-OE ; Chen2020 ; Xu2020 ; liyong2021 ; Tutunnikov2021 . The concept of the adiabatic passage techniques, such as stimulated Raman adiabatic passages stirap1 ; stirap2 and shortcuts to adiabaticity sta1 ; sta2 , was proposed to generate efficient and robust detection and separation of chiral molecules Shapiro2000 ; Kral2001 ; Kral2003 ; Gerbasi2001 ; Vitanov2019 ; Wu2019 . In order to meet adiabatic criteria, the adiabatic passage techniques involve strict limitations on the control fields, and therefore, the corresponding control processes are usually slow and complicated. To that end, nonadiabatic schemes using much shorter durations of control pulses than the adiabatic approaches have been proposed to reach fast enantioselective excitation of chiral molecules Frishman2003 ; Hamedi2019 . Experimentally, it has been demonstrated by using resonant microwave three-wave mixing (M3WM) techniques M3WM0 ; M3WM1 ; M3WM2 ; M3WM3 ; M3WM4 ; M3WM5 ; M3WM6 ; M3WM7 . A common feature of both adiabatic and nonadiabatic control schemes usually involves a closed-loop quantum system by cyclic coupling of three molecular (i.e., rotational or rovibrational) states (as shown in Fig. 1), which are resonantly driven through the aa-type, bb-type, and cc-type components of the transition dipole moments by a combination of three orthogonally polarized and phase-controlled microwave fields Leibscher2019 . Since one of three cyclic couplings differs in the sign of the transition dipole moment in two opposite enantiomers, direct one-photon transition path from the ground state to a given target state constructively or destructively interferes with indirect two-photon transition path through an intermediate state, leading to enantioselective state transfer (ESST). Although the pulse areas of the three control fields that can generate enantioselective excitation have been experimentally examined in M3WM experiments, there is still lacking a general pulse-areas theorem that can be used to directly calculate the amplitudes and phases of three time-dependent control pulses so as to gain insights into the underlying coherent quantum control mechanism.
In this paper, we focus on ESST and present a three-level pulse-areas theorem analysis. Previous works had derived the pulse area theorems for the three-level quantum systems with the ladder-, Λ\Lambda- and VV-type configurations Sugny1 ; Guo2008 ; Shchedrin2015 ; shupra2019 , leading to many successful applications in coherent quantum control simulations and experiments Sola2018 ; jpca2019 ; shuprl2019 ; shupra2020 ; shuol2020 ; shupra2021 . Here, we take a strategy by dividing the closed-loop three-level excitation into two separated stages, i.e., combining a two-level excitation and a time-delayed-open-loop three-level transition. We obtain a pulse-areas theorem of the three-level system with a Δ\Delta-type configuration without applying the rotating-wave approximation. The derived pulse-areas theorem can calculate the exact amplitude and phase conditions for generating efficient ESST to the desired quantum state, which is examined in cyclohexylmethanol molecules (C7H14O\text{C}_{7}\text{H}_{14}\text{O}) with different pulse sequences. This work provides an essential reference for coherent control of ESST using closed-loop three-level interaction schemes.
The remainder of this paper is organized as follows. In Sec. II, we describe the theoretical methods for obtaining a three-level pulse area theorem with cyclic coupling. We perform the numerical simulations to examine the derived pulse-areas theorem in Sec. III. Finally, we conclude with a summary in Sec. IV.

II Theoretical method and numerical model

II.1 Closed-loop three-level model

Refer to caption
Figure 1: Schematic illustration of enantioselective state transfer (ESST). (a) The space orientations of the transition dipole moments μa/b/c\mu_{a/b/c} with respect to the polarization directions of three linearly polarized microwave pulses a/b/c(t)\mathcal{E}_{a/b/c}(t) . (b) The corresponding closed-loop transitions within three rotational states |A|A\rangle, |B|B\rangle and |C|C\rangle. The transition frequencies and couplings between states are identical, except for a difference due to the sign difference in μa\mu_{a} The parameters of the cyclohexylmethanol molecules (C7H14O\text{C}_{7}\text{H}_{14}\text{O}) are used for numerical simulations.

To describe our model, three rotational states of asymmetric top molecules, as shown in Fig. 1, are labeled as |A|A\rangle, |B|B\rangle and |C|C\rangle with a subscript L/RL/R for the left/right-handed enantiomer. The energies EAE_{A}, EBE_{B} and ECE_{C} of three rotational states are identical for enantiomers. We mark three microwave control fields as a/b/c(t)\overrightarrow{\mathcal{E}}_{a/b/c}\left(t\right), which drive the three states with the transition dipole moments μa\overrightarrow{\mu}_{a}, μb\overrightarrow{\mu}_{b} and μc\overrightarrow{\mu}_{c}. As demonstrated in Refs. Leibscher2019 ; kevin2018 , the ESST scheme with the use of linearly polarized control fields requires that a combination of control fields a\overrightarrow{\mathcal{E}}_{a}, b\overrightarrow{\mathcal{E}}_{b} and c\overrightarrow{\mathcal{E}}_{c} with three orthogonal polarization directions along the directions of three dipole moment components μa\overrightarrow{\mu}_{a}, μb\overrightarrow{\mu}_{b} and μc\overrightarrow{\mu}_{c}. To that end, we describe the time- and polarization-dependent electric fields of the three control fields by

a/b/c(t)=ea/b/ca/b/cfa/b/c(t)cos[ωa/b/c(tta/b/c)+ϕa/b/c]\overrightarrow{\mathcal{E}}_{a/b/c}\left(t\right)=\overrightarrow{e}_{a/b/c}\mathcal{E}_{a/b/c}f_{a/b/c}\left(t\right)\cos\left[\omega_{a/b/c}(t-t_{a/b/c})+\phi_{a/b/c}\right] (1)

where ea/b/c\overrightarrow{e}_{a/b/c}, a/b/c\mathcal{E}_{a/b/c}, fa/b/c(t)f_{a/b/c}\left(t\right), ωa/b/c\omega_{a/b/c}, ta/b/ct_{a/b/c}, and ϕa/b/c\phi_{a/b/c} denote the polarization direction, the strength, envelope function, center frequency, center time, and phase of a/b/c(t)\overrightarrow{\mathcal{E}}_{a/b/c}\left(t\right), respectively. The triple product μa(μb×μc)\overrightarrow{\mu}_{a}\cdot(\overrightarrow{\mu}_{b}\times\overrightarrow{\mu}_{c}) is independent on the choice of the inertia principle axes aa, bb, and cc, but is of opposite sign for the left- and right-handed enantiomers. For convenience, we specify that μb\overrightarrow{\mu}_{b} and μc\overrightarrow{\mu}_{c} are identical for two enantiomers, whereas μa\overrightarrow{\mu}_{a} changes sign with the handedness. Thus, the Hamiltonian of two different handed-enantiomers in the presence of the control fields can be written as

HL,R(t)\displaystyle H^{L,R}\left(t\right) =\displaystyle= (EA±Ωa(t)Ωb(t)±Ωa(t)EBΩc(t)Ωb(t)Ωc(t)EC),\displaystyle\left(\begin{array}[]{ccc}E_{A}&\pm\Omega_{a}\left(t\right)&\Omega_{b}\left(t\right)\\ \pm\Omega_{a}\left(t\right)&E_{B}&\Omega_{c}\left(t\right)\\ \Omega_{b}\left(t\right)&\Omega_{c}\left(t\right)&E_{C}\end{array}\right), (5)

where the three cyclic couplings read Ωa(t)=μaa(t)\text{$\Omega_{a}\left(t\right)=-\mu_{a}$}\mathcal{E}_{a}\left(t\right), Ωb(t)=μbb(t)\text{$\Omega_{b}\left(t\right)=-\mu_{b}$}\mathcal{E}_{b}\left(t\right) and Ωc(t)=μcc(t)\text{$\Omega_{c}\left(t\right)=-\mu_{c}$}\mathcal{E}_{c}\left(t\right).
We now analyze how to achieve enantioselective excitation of a given target state from a given initial state. We assume that two enantiomers are initially in the state |A|A\rangle, and the control target can be either the excited state |B|B\rangle or |C|C\rangle. For the choice of |C|C\rangle as the target, there are a direct one-photon transition |A|C|A\rangle\leftrightarrow|C\rangle and an indirect two-photon transition |A|B|C|A\rangle\leftrightarrow|B\rangle\leftrightarrow|C\rangle, which form the closed-loop interaction scheme. If we take |B|B\rangle as the target, two transition paths correspond to a direct one |A|B|A\rangle\leftrightarrow|B\rangle and an indirect one |A|C|B|A\rangle\leftrightarrow|C\rangle\leftrightarrow|B\rangle. Since it remains difficult to derive an analytical solution by directly using the Hamiltonian in Eq. (5), we use a strategy by dividing the excitation processes into two stages.

II.2 Control conditions for ESST to |C|C\rangle

II.2.1 Analytical solution for a two-level system

For the ESST to |C|C\rangle, we assume that the coupling Ωa\Omega_{a} is turned on at the initial time t0t_{0} and off it at a time t1t_{1} before the couplings Ωb\Omega_{b} and Ωc\Omega_{c}. Thus, the system is reduced into a two-level system, and the corresponding Hamiltonian reads

H1L,R(t)=(EA±Ωa(t)±Ωa(t)EB).\displaystyle H_{1}^{L,R}\left(t\right)=\left(\begin{array}[]{cc}E_{A}&\pm\Omega_{a}\left(t\right)\\ \pm\Omega_{a}\left(t\right)&E_{B}\end{array}\right). (8)

Without using the rotating wave approximation, the evolution of the system in the interaction picture can be described by using the unitary operator

U1L,R(t,t0)=U1L,R(t0,t0)it0t𝑑tH1IL,R(t)U1L,R(t,t0)U_{1}^{L,R}\left(t,t_{0}\right)=U_{1}^{L,R}\left(t_{0},t_{0}\right)-i\int_{t_{0}}^{t}dt^{\prime}H_{1I}^{L,R}\left(t^{\prime}\right)U_{1}^{L,R}\left(t^{\prime},t_{0}\right) (9)

where H1IL,R(t)=exp(iH10t)[±Ωa(t)(|AB|+|BA|)]exp(iH10t)H_{1I}^{L,R}\left(t\right)=\exp\left(iH_{10}t\right)\left[\pm\Omega_{a}\left(t\right)\left(|A\rangle\langle B|+|B\rangle\langle A|\right)\right]\exp\left(-iH_{10}t\right) with the field-free Hamiltonian of the two-level system H10=EA|AA|+EB|BB|H_{10}=E_{A}\left|A\right\rangle\left\langle A\right|+E_{B}\left|B\right\rangle\left\langle B\right|. By involving the first-order the Magnus expansion pr:470:151 , the time-dependent unitary operator U1L,R(t,t0)U_{1}^{L,R}\left(t,t_{0}\right) can be given by Shchedrin2015 ; shupra2019 ; shuprl2019 ; shupra2020

U1L,R(t,t0)\displaystyle U_{1}^{L,R}\left(t,t_{0}\right) =\displaystyle= cos|θa(t)|(|AA|+|BB|)\displaystyle\cos\left|\theta_{a}(t)\right|\left(\left|A\right\rangle\left\langle A\right|+\left|B\right\rangle\left\langle B\right|\right) (10)
isin|θa(t)|[|θa(t)|θa(t)|BA|+|θa(t)|θa(t)|AB|]\displaystyle\mp i\sin\left|\theta_{a}(t)\right|\left[\frac{\left|\theta_{a}\text{$\left(t\right)$}\right|}{\theta_{a}^{*}\text{$\left(t\right)$}}\left|B\right\rangle\left\langle A\right|+\frac{\left|\theta_{a}\text{$\left(t\right)$}\right|}{\theta_{a}\text{$\left(t\right)$}}\left|A\right\rangle\left\langle B\right|\right]

in terms of the complex pulse area θa(t)=t0tΩa(t)exp(iωABt)𝑑t\theta_{a}\text{$\left(t\right)$}=\int_{t_{0}}^{t}\Omega_{a}\left(t^{\prime}\right)\exp(i\omega_{AB}t^{\prime})dt^{\prime} with ωAB=EBEA\omega_{AB}=E_{B}-E_{A}. By considering the left and right-handed enantiomers initially in the ground state |A\left|A\right\rangle, an analytic solution for the wave-function of the two-level system can be obtained by

|ψ1L,R(t)\displaystyle\left|\psi_{1}^{L,R}\text{$\left(t\right)$}\right\rangle =\displaystyle= U1L,R(t,t0)|A\displaystyle U_{1}^{L,R}\left(t,t_{0}\right)\left|A\right\rangle (11)
=\displaystyle= cos|θa(t)||Ai|θa(t)|θa(t)sin|θa(t)||B.\displaystyle\cos\left|\theta_{a}\left(t\right)\right||A\rangle\mp i\frac{\left|\theta_{a}\left(t\right)\right|}{\theta_{a}^{*}\left(t\right)}\sin\left|\theta_{a}\left(t\right)\right||B\rangle.

II.2.2 Analytical solution for a three-level system

After the coupling Ωa\Omega_{a} off at t1t_{1}, we turn on the couplings Ωb\Omega_{b} and Ωc\Omega_{c}. The Hamiltonian in Eq. (5) is reduced into

H2L,R(t)\displaystyle H^{L,R}_{2}\left(t\right) =\displaystyle= (EA0Ωb(t)0EBΩc(t)Ωb(t)Ωc(t)EC).\displaystyle\left(\begin{array}[]{ccc}E_{A}&0&\Omega_{b}\left(t\right)\\ 0&E_{B}&\Omega_{c}\left(t\right)\\ \Omega_{b}\left(t\right)&\Omega_{c}\left(t\right)&E_{C}\end{array}\right). (15)

The corresponding time-dependent unitary operator can be given by

U2L,R(t,t1)=U2L,R(t1,t1)it1t𝑑tH2IL,R(t)U2L,R(t,t1)U^{L,R}_{2}(t,t_{1})=U^{L,R}_{2}(t_{1},t_{1})-i\int_{t_{1}}^{t}dt^{\prime}H_{2I}^{L,R}\left(t^{\prime}\right)U^{L,R}_{2}(t^{\prime},t_{1}) (16)

where H2IL,R(t)=exp(iH20t)[Ωb(t)(|AC|+|CC|)+Ωc(t)(|CB|+|BC|)]exp(iH20t)H_{2I}^{L,R}\left(t\right)=\exp\left(iH_{20}t\right)\left[\Omega_{b}(t)(|A\rangle\langle C|+|C\rangle\langle C|)+\Omega_{c}(t)(|C\rangle\langle B|+|B\rangle\langle C|)\right]\exp\left(-iH_{20}t\right) with the field-free Hamiltonian of the three-level system H20=EA|AA|+EB|BB|+EC|CC|H_{20}=E_{A}|A\rangle\langle A|+E_{B}|B\rangle\langle B|+E_{C}|C\rangle\langle C|. By making the first-order Magnus expansion of the unitary operator U2L,R(t,t1)U^{L,R}_{2}(t,t_{1}), the time-dependent wave function of two enantiomers can be given by

|ψ2L,R(t)\displaystyle|\psi_{2}^{L,R}\left(t\right)\rangle =\displaystyle= U2L,R(t,t1)|ψ1L,R(t1)\displaystyle U_{2}^{L,R}\left(t,t_{1}\right)\left|\psi_{1}^{L,R}\text{$\left(t_{1}\right)$}\right\rangle (17)
=\displaystyle= [cos|θa(t1)||θc(t)|2+|θb(t)|2cosθ(t)θ2(t)isin|θa(t1)ζ(t)||θa(t1)|θa(t1)]|A\displaystyle\left[\cos\left|\theta_{a}\left(t_{1}\right)\right|\frac{\left|\theta_{c}\left(t\right)\right|^{2}+\left|\theta_{b}\left(t\right)\right|^{2}\cos\theta\left(t\right)}{\theta^{2}\left(t\right)}\mp i\sin\left|\theta_{a}\left(t_{1}\right)\zeta\left(t\right)\right|\frac{\left|\theta_{a}\left(t_{1}\right)\right|}{\theta_{a}^{*}\left(t_{1}\right)}\right]\left|A\right\rangle
+[cos|θa(t1)|ζ(t)isin|θa(t1)||θa(t1)|θa(t1)|θb(t)|2+|θc(t)|2cosθ(t)θ2(t)]|B\displaystyle+\left[\cos\left|\theta_{a}\left(t_{1}\right)\right|\zeta^{*}\left(t\right)\mp i\sin\left|\theta_{a}\left(t_{1}\right)\right|\frac{\left|\theta_{a}\left(t_{1}\right)\right|}{\theta_{a}^{*}\left(t_{1}\right)}\frac{\left|\theta_{b}\left(t\right)\right|^{2}+\left|\theta_{c}\left(t\right)\right|^{2}\cos\theta\left(t\right)}{\theta^{2}\left(t\right)}\right]\left|B\right\rangle
sinθ(t)θ(t)[icos|θa(t1)|θb(t)±sin|θa(t1)|θc(t)|θa(t1)|θa(t1)]|C,\displaystyle-\frac{\sin\theta\left(t\right)}{\theta\left(t\right)}\left[i\cos\left|\theta_{a}\left(t_{1}\right)\right|\theta_{b}\left(t\right)\pm\sin\left|\theta_{a}\left(t_{1}\right)\right|\theta_{c}\left(t\right)\frac{\left|\theta_{a}\left(t_{1}\right)\right|}{\theta_{a}^{*}\left(t_{1}\right)}\right]\left|C\right\rangle,

where ζ(t)=θc(t)θb(t)[cosθ(t)1]/θ2(t)\zeta\left(t\right)=\theta_{c}\left(t\right)\theta_{b}^{*}\left(t\right)\left[\cos\theta\left(t\right)-1\right]/\theta^{2}\left(t\right) and θ(t)=|θb(t)|2+|θc(t)|2\theta\left(t\right)=\sqrt{\left|\theta_{b}\left(t\right)\right|^{2}+\left|\theta_{c}\left(t\right)\right|^{2}} in terms of the complex pulse areas θb(t)=t1tΩb(t)exp(iωACt)𝑑t\theta_{b}\left(t\right)=\int_{t_{1}}^{t}\Omega_{b}\left(t^{\prime}\right)\exp(i\omega_{AC}t^{\prime})dt^{\prime} and θc(t)=t1tΩc(t)exp(iωBCt)𝑑t\theta_{c}\left(t\right)=\int_{t_{1}}^{t}\Omega_{c}\left(t^{\prime}\right)\exp(i\omega_{BC}t^{\prime})dt^{\prime} with the transition frequencies ωBC=ECEB\omega_{BC}=E_{C}-E_{B} and ωAC=ECEA\omega_{AC}=E_{C}-E_{A}.

To entirely transfer the left-handed enantiomer to the state |C|C\rangle while keeping the right-handed one unpopulated at the final time tft_{f}, the complex pulse areas should satisfy the following two relations

|sinθ(tf)θ(tf)[iθb(tf)cos|θa(t1)|+|θa(t1)|θc(tf)θa(t1)sin|θa(t1)|]|\displaystyle\left|\frac{\sin\theta\left(t_{f}\right)}{\theta\left(t_{f}\right)}\left[i\theta_{b}\left(t_{f}\right)\cos\left|\theta_{a}\left(t_{1}\right)\right|+\frac{\left|\theta_{a}\left(t_{1}\right)\right|\theta_{c}\left(t_{f}\right)}{\theta_{a}^{*}\left(t_{1}\right)}\sin\left|\theta_{a}\left(t_{1}\right)\right|\right]\right| =\displaystyle= 1,\displaystyle 1, (18)
|sinθ(tf)θ(tf)[iθb(tf)cos|θa(t1)||θa(t1)|θc(tf)θa(t1)sin|θa(t1)|]|\displaystyle\left|\frac{\sin\theta\left(t_{f}\right)}{\theta\left(t_{f}\right)}\left[i\theta_{b}\left(t_{f}\right)\cos\left|\theta_{a}\left(t_{1}\right)\right|-\frac{\left|\theta_{a}\left(t_{1}\right)\right|\theta_{c}\left(t_{f}\right)}{\theta_{a}^{*}\left(t_{1}\right)}\sin\left|\theta_{a}\left(t_{1}\right)\right|\right]\right| =\displaystyle= 0,\displaystyle 0, (19)

From the Eq. (19), we can derive

θc(tf)θa(t1)=iθb(tf)cos|θa(t1)||θa(t1)|sin|θa(t1)|.\frac{\theta_{c}\left(t_{f}\right)}{\theta_{a}^{*}\left(t_{1}\right)}=\frac{i\theta_{b}\left(t_{f}\right)\cos\left|\theta_{a}\left(t_{1}\right)\right|}{\left|\theta_{a}\left(t_{1}\right)\right|\sin\left|\theta_{a}\left(t_{1}\right)\right|}. (20)

By inserting Eq. (20) into Eq. (18), we can obtain a relation

4|θb(tf)|2sin2θ(tf)cos2|θa(t1)|=θ2(tf).4\left|\theta_{b}\left(t_{f}\right)\right|^{2}\sin^{2}\theta\left(t_{f}\right)\cos^{2}\left|\theta_{a}\left(t_{1}\right)\right|=\theta^{2}\left(t_{f}\right). (21)

This relation can be fulfilled when the three control fields satisfy the amplitude conditions

|θb(tf)|\displaystyle\left|\theta_{b}\left(t_{f}\right)\right| =|θc(tf)|=12(k+12)π,(kN)\displaystyle=\left|\theta_{c}\left(t_{f}\right)\right|=\frac{1}{\sqrt{2}}\left(k+\frac{1}{2}\right)\pi,(k\in N)
|θa(t1)|\displaystyle\Big{|}\theta_{a}\left(t_{1}\right)\Big{|} =(k+14)π,(kN).\displaystyle=\left(k^{\prime}+\frac{1}{4}\right)\pi,(k^{\prime}\in N). (22)

Furthermore, we insert Eq. (22) into Eq. (20) with θb=|θb|exp(iϕb)\theta_{b}=-\left|\theta_{b}\right|\exp\left(-i\phi_{b}\right), θc=|θc|exp(iϕc)\theta_{c}=-\left|\theta_{c}\right|\exp\left(-i\phi_{c}\right) and θa=|θa|exp(iϕa)\theta_{a}=-\left|\theta_{a}\right|\exp\left(-i\phi_{a}\right). We find that the three control fields satisfy the following phase condition,

ϕa+ϕcϕb=(2l+12)π,(lZ).\phi_{a}+\phi_{c}-\phi_{b}=\left(2l+\frac{1}{2}\right)\pi,(l\in Z). (23)

Similarly, we can use the same amplitude conditions as that in Eq. (22) to achieve complete ESST to |C|C\rangle of the right-handed enantiomer by using the phase condition of ϕa+ϕbϕc=(2l1/2)π,(lZ)\phi_{a}+\phi_{b}-\phi_{c}=(2l-1/2)\pi,(l\in Z). It implies that a π\pi flip of the phase on one of three control fields can result in opposite ESST. To that end, the handedness of enantiomers can be determined by measuring the population in the state |C|C\rangle.

II.3 Control conditions for ESST to |B|B\rangle

For ESST to |B|B\rangle, we apply the coupling Ωb\Omega_{b} before the couplings Ωa\Omega_{a} and Ωc\Omega_{c}, which results in a coherent superposition state of |A|A\rangle and |C|C\rangle. As demonstrated above by involving the first-order Magnus expansion and further mathematical derivations, an analytic wave-function of the three-level Λ\Lambda-type system can be given by

|ψ2L,R(t)=[cos|θb(t1)||θc(t)|2+|θa(t)|2cosθ(t)θ2isin|θb(t1)|ξ(t)|θb(t1)|θb(t1)]\displaystyle|\psi_{2}^{L,R}\left(t\right)\rangle=\left[\cos\left|\theta_{b}\left(t_{1}\right)\right|\frac{\left|\theta_{c}\left(t\right)\right|^{2}+\left|\theta_{a}\left(t\right)\right|^{2}\cos\theta\left(t\right)}{\theta^{2}}\mp i\sin\left|\theta_{b}\left(t_{1}\right)\right|\xi^{*}\left(t\right)\frac{\left|\theta_{b}\left(t_{1}\right)\right|}{\theta_{b}^{*}\left(t_{1}\right)}\right] |A\displaystyle\left|A\right\rangle
sinθ(t)θ(t)[sin|θb(t1)|θc(t)|θb(t1)|θb(t1)±icos|θb(t1)|θa(t)]\displaystyle-\frac{\sin\theta\left(t\right)}{\theta\left(t\right)}\left[\sin\left|\theta_{b}\left(t_{1}\right)\right|\theta_{c}^{*}\left(t\right)\frac{\left|\theta_{b}\left(t_{1}\right)\right|}{\theta_{b}^{*}\left(t_{1}\right)}\pm i\cos\left|\theta_{b}\left(t_{1}\right)\right|\theta_{a}\left(t\right)\right] |B\displaystyle\left|B\right\rangle
[isin|θb(t1)||θb(t1)|θb(t1)|θa(t)|2+|θc(t)|2cosθ(t)θ2(t)cos(|θb(t1)|)ξ(t)]\displaystyle-\left[i\sin\left|\theta_{b}\left(t_{1}\right)\right|\frac{\left|\theta_{b}\left(t_{1}\right)\right|}{\theta_{b}^{*}\left(t_{1}\right)}\frac{\left|\theta_{a}\left(t\right)\right|^{2}+\left|\theta_{c}\left(t\right)\right|^{2}\cos\theta\left(t\right)}{\theta^{2}\left(t\right)}\mp\cos\left(\left|\theta_{b}\left(t_{1}\right)\right|\right)\xi\left(t\right)\right] |C,\displaystyle\left|C\right\rangle, (24)

where ξ(t)=θc(t)θa(t)[cosθ(t)1]/θ2(t)\xi\left(t\right)=\theta_{c}\left(t\right)\theta_{a}\left(t\right)\left[\cos\theta\left(t\right)-1\right]/\theta^{2}\left(t\right) and θ(t)=|θa(t)|2+|θc(t)|2\theta\left(t\right)=\sqrt{\left|\theta_{a}\left(t\right)\right|^{2}+\left|\theta_{c}\left(t\right)\right|^{2}}.
To entirely transfer the left-handed enantiomer to the state |B|B\rangle at the final time tft_{f}, but the right-handed one is not populating, we have

|sinθ(tf)θ(tf)[sin|θb(t1)|θc(tf)|θb(t1)|θb(t1)+icos|θb(t1)|θa(tf)]|\displaystyle\left|\frac{\sin\theta\left(t_{f}\right)}{\theta\left(t_{f}\right)}\left[\sin\left|\theta_{b}\left(t_{1}\right)\right|\theta_{c}^{*}\left(t_{f}\right)\frac{\left|\theta_{b}\left(t_{1}\right)\right|}{\theta_{b}^{*}\left(t_{1}\right)}+i\cos\left|\theta_{b}\left(t_{1}\right)\right|\theta_{a}\left(t_{f}\right)\right]\right| =\displaystyle= 1,\displaystyle 1, (25)
|sinθ(tf)θ(tf)[sin|θb(t1)|θc(tf)|θb(t1)|θb(t1)icos|θb(t1)|θa(tf)]|\displaystyle\left|\frac{\sin\theta\left(t_{f}\right)}{\theta\left(t_{f}\right)}\left[\sin\left|\theta_{b}\left(t_{1}\right)\right|\theta_{c}^{*}\left(t_{f}\right)\frac{\left|\theta_{b}\left(t_{1}\right)\right|}{\theta_{b}^{*}\left(t_{1}\right)}-i\cos\left|\theta_{b}\left(t_{1}\right)\right|\theta_{a}\left(t_{f}\right)\right]\right| =\displaystyle= 0.\displaystyle 0. (26)

Furthermore, we can obtain the amplitude condition for the three control fields

|θa(tf)|\displaystyle\left|\theta_{a}\left(t_{f}\right)\right| =|θc(tf)|=12(k+12)π,(kN)\displaystyle=\left|\theta_{c}\left(t_{f}\right)\right|=\frac{1}{\sqrt{2}}\left(k+\frac{1}{2}\right)\pi,(k\in N)
|θb(t1)|\displaystyle\Big{|}\theta_{b}\left(t_{1}\right)\Big{|} =(k+14)π,(kN).\displaystyle=\left(k^{\prime}+\frac{1}{4}\right)\pi,(k^{\prime}\in N). (27)

The ESST to |B|B\rangle of the left-handed enantiomer can be reached by using the phase condition

ϕa+ϕcϕb=(2l12)π,(lZ).\phi_{a}+\phi_{c}-\phi_{b}=\left(2l-\frac{1}{2}\right)\pi,(l\in Z). (28)

The amplitude conditions by Eq. (27) in forms are the same as Eq. (21) with different orders. That is, the orders of the three pulses are interchangeable, dependent on the choice of the target state. From the phase condition in Eq. (28), we can find that a π\pi flip of the phase on one of three control fields can also lead to opposite ESST. Since our schemes that divide the closed-loop excitation schemes into two stages are different from previous works Wu2020 ; Torosov2020 ; Torosov2020-2 ; Wu2019 ; M3WM5 , which turned on the direct one-photon transition path before the two-photon one, these amplitude and phase conditions provide an alternative way to achieve ESST within a cyclic three-level systems in chiral molecules.
Note that the amplitude conditions of |θa/b(tf)|=π/4|\theta_{a/b}(t_{f})|=\pi/4 and |θ(tf)|=|θb/a(tf)|2+|θc(tf)|2=π/2|\theta(t_{f})|=\sqrt{|\theta_{b/a}(t_{f})|^{2}+|\theta_{c}(t_{f})|^{2}}=\pi/2 are equivalent to that with the use of π/2\pi/2 and π\pi pulses, for which a scale of 1/21/2 factor comes from the definition of the complex pulse-areas without using the rotating wave approximation and the resonant excitation conditions. To show the advantage of using the complex pulse-areas, we can have a frequency-domain analysis for the control fields

a/b/c=1πt0tf𝑑ω𝒜a/b/c(ω)eiϕa/b/c(ω)eiωt\mathcal{E}_{a/b/c}=\frac{1}{\pi}\int_{t_{0}}^{t_{f}}d\omega\mathcal{A}_{a/b/c}(\omega)e^{i\phi_{a/b/c}(\omega)}e^{i\omega t} (29)

where 𝒜a/b/c(ω)\mathcal{A}_{a/b/c}(\omega) and ϕa/b/c(ω)\phi_{a/b/c}(\omega) are the spectral amplitude and phase, respectively. We can find that the values of θa/b/c(tf)\theta_{a/b/c}(t_{f}) depend only on 𝒜a/b/c(ω)\mathcal{A}_{a/b/c}(\omega) and ϕa/b/c(ω)\phi_{a/b/c}(\omega) of three control fields at transition frequencies ωAB\omega_{AB}, ωAC\omega_{AC} and ωBC\omega_{BC}. Thus, our definitions of the complex pulse areas can also be applied to the pulsed control fields whose center frequencies are detuned away from the transition frequencies. In Sec. III, we present simulations to examine the amplitude and phase conditions for both center-frequency resonant and detuned microwave excitation schemes.

III Results and discussion

We perform the numerical simulations in the cyclohexylmethanol molecules. Three rotational states of |101,|202\left|1_{01}\right\rangle,\left|2_{02}\right\rangle, and |212\left|2_{12}\right\rangle are used as |A,|B\left|A\right\rangle,\left|B\right\rangle, and |C\left|C\right\rangle. The transition frequencies between states are ωAB=4720MHz,ωBC=2339MHz\omega_{AB}=4720\text{MHz},\omega_{BC}=2339\text{MHz}, and ωAC=7059MHz\omega_{AC}=7059\text{MHz}, and the transition dipole moments take the values of μa=0.4Debye,μb=1.2Debye\mu_{a}=0.4\text{Debye},\mu_{b}=1.2\text{Debye} and μc=0.8Debye\mu_{c}=0.8\text{Debye} Wu2020 ; M3WM7 . In our simulations, we take three control fields with the Gaussian profile as follows

a/b/c(t)=2πAa/b/cτ0exp[(tta/b/c)22τ02]cos[ωq/p/s(tta/b/c)+ϕa/b/c].\mathcal{E}_{a/b/c}\left(t\right)=\sqrt{\frac{2}{\pi}}\frac{A_{a/b/c}}{\tau_{0}}\exp{\Bigg{[}-\frac{(t-t_{a/b/c})^{2}}{2\tau_{0}^{2}}}\Bigg{]}\cos\left[\omega_{q/p/s}(t-t_{a/b/c})+\phi_{a/b/c}\right]. (30)

This description of the control fields is convenient for determining the field strengths a/b/c\mathcal{E}_{a/b/c} for any accessible duration τ0\tau_{0}. By choosing constant values of Aa/b/cA_{a/b/c}, we can see that the complex pulse-areas θa/b/c(tf)\theta_{a/b/c}(t_{f}) with such descriptions do not depend on the duration τ0\tau_{0}. Thus, this scheme avoids the strict limitations by the adiabatic criterion, providing a way to design fast control schemes using much shorter pulse duration than the adiabatic scenario. For practical applications, however, we need to balance the choice of pulse duration τ0\tau_{0}, so that unwanted transitions to neighboring energy levels could be avoided by using narrowband pulses.

Refer to caption
Figure 2: Chiral dependence of ESST on the phase for the target state |C|C\rangle. (a) The final population of the left-handed enantiomer PCL(tf)P_{C}^{L}(t_{f}) versus the duration τ0\tau_{0} and phase ϕa\phi_{a} in a range of [0,2π][0,2\pi], and the corresponding cut lines at (b) τ0=35\tau_{0}=35ns and (c) ϕa=π/2\phi_{a}=\pi/2. (d)-(f) The same plots as (a)-(c) for the right-handed enantiomer except for ϕa=3π/2\phi_{a}=3\pi/2 in (f).

For the cyclohexylmethanol molecules, there exists a rotational state |111\left|1_{11}\right\rangle with the energy slightly below the state |B|B\rangle, referred as the state |B|B^{\prime}\rangle, which can be connected to the excited state |C|C\rangle via the aa-type transition in ωBC=4484MHz\omega_{B^{\prime}C}=4484\text{MHz} and to the ground state |A|A\rangle via the cc-type transition in ωAB=2575MHz\omega_{AB^{\prime}}=2575\text{MHz}. To this end, we include this state with a four-level model to perform the numerical simulations. The corresponding field-molecule interaction Hamiltonian reads.

HcL,R(t)\displaystyle H^{L,R}_{c}\left(t\right) =\displaystyle= (0Ωc(t)±Ωa(t)Ωb(t)Ωc(t)00±Ωa(t)±Ωa(t)00Ωc(t)Ωb(t)±Ωa(t)Ωc(t)0),\displaystyle\left(\begin{array}[]{cccc}0&\Omega^{{}^{\prime}}_{c}(t)&\pm\Omega_{a}(t)&\Omega_{b}(t)\\ \Omega^{{}^{\prime}}_{c}(t)&0&0&\pm\Omega_{a}^{{}^{\prime}}(t)\\ \pm\Omega_{a}(t)&0&0&\Omega_{c}(t)\\ \Omega_{b}(t)&\pm\Omega_{a}^{{}^{\prime}}(t)&\Omega_{c}(t)&0\end{array}\right), (35)

where we take the additional couplings Ωa(t)=μaa(t)\Omega_{a}^{{}^{\prime}}(t)=-\mu^{\prime}_{a}\mathcal{E}_{a}(t) and Ωc(t)=μcc(t)\Omega_{c}^{{}^{\prime}}(t)=-\mu^{\prime}_{c}\mathcal{E}_{c}(t) with μa=μa\mu^{\prime}_{a}=\mu_{a} and μc=μc\mu^{\prime}_{c}=\mu_{c} in our simulations. The time-dependent unitary operator in the interaction picture can be numerically computed by

UL,R(t,t0)=UL,R(t0,t0)it0t𝑑tHIL,R(t)UL,R(t,t0),U^{L,R}(t,t_{0})=U^{L,R}(t_{0},t_{0})-i\int_{t_{0}}^{t}dt^{\prime}H_{I}^{L,R}\left(t^{\prime}\right)U^{L,R}(t^{\prime},t_{0}), (36)

where UL,R(t0,t0)=𝕀U^{L,R}(t_{0},t_{0})=\mathbb{I} and HIL,R(t)=exp(iH0t)[HcL,R(t)]exp(iH0t)H_{I}^{L,R}\left(t\right)=\exp\left(iH_{0}t\right)[H^{L,R}_{c}\left(t\right)]\exp\left(-iH_{0}t\right) with the field-free Hamiltonian H0=X=ACEX|XX|H_{0}=\sum_{X=A}^{C}E_{X}|X\rangle\langle X|. By projecting the unitary operator UL,R(t,t0)U^{L,R}(t,t_{0}) onto the initial state |A|A\rangle, we can obtain the time-dependent wave function of the system |ψL,R(t)=UL,R(t,t0)|A|\psi^{L,R}(t)\rangle=U^{L,R}(t,t_{0})|A\rangle without using the first-order Magnus expansion. Thus, the time-dependent population in the state |X|X\rangle can be calculated by PXL/R(t)=|X|ψL,R(t)|2P^{L/R}_{X}(t)=\left|\langle X|\psi^{L,R}(t)\rangle\right|^{2} with X=A,B,B,CX=A,B^{\prime},B,C.

III.1 ESST to |C|C\rangle

Refer to caption
Figure 3: The time-dependent populations of four rotational states for ESST to the state |C|C\rangle. The simulations for (a) the left-handed and (b) right-handed enantiomers with the duration τ0=35\tau_{0}=35ns and phase ϕa=π/2\phi_{a}=\pi/2.

For the target state |C|C\rangle, we set the parameters Aa=π/(4μa)A_{a}=\pi/(4\mu_{a}), Ab=π/(22μb)A_{b}=\pi/(2\sqrt{2}\mu_{b}), and Ac=π/(22μc)A_{c}=\pi/(2\sqrt{2}\mu_{c}). It is easily to verify that the three control fields defined by Eq. (30) with different values of τ0\tau_{0} exactly satisfy the amplitude conditions in Eq. (22) at transition frequencies by fixing the center frequencies ωa=ωAB\omega_{a}=\omega_{AB}, ωb=ωAC\omega_{b}=\omega_{AC} and ωc=ωBC\omega_{c}=\omega_{BC}. Figure 2 shows the results of PCL,R(tf)P_{C}^{L,R}(t_{f}) versus τ0\tau_{0} and ϕa\phi_{a} with ϕb=ϕc=0\phi_{b}=\phi_{c}=0. As expected, the ESST to the state |C|C\rangle appears and depends highly on the phase values of ϕa\phi_{a}. The fidelity of PCL,R(tf)>0.999P_{C}^{L,R}(t_{f})>0.999 can be reached for τ0>35\tau_{0}>35 ns, indicating that the unwanted transition to the neighboring state |B|B^{\prime}\rangle can be ignored. Figures 2 (b) and (e) plot the dependence of PCL,R(tf)P_{C}^{L,R}(t_{f}) on the phase ϕa\phi_{a} for the case of τ0=35\tau_{0}=35 ns. There is no ESST at ϕa=0\phi_{a}=0 and π\pi. The entire ESST to the left-handed molecule occurs at ϕa=π/2\phi_{a}=\pi/2 and a phase change to ϕa=3π/2\phi_{a}=3\pi/2 results in an opposite transfer to the right-handed one. The similar features can be observed by changing the values of ϕb\phi_{b} ( or ϕc\phi_{c}) while choosing ϕa=ϕc=0\phi_{a}=\phi_{c}=0 ( or ϕa=ϕb=0\phi_{a}=\phi_{b}=0). These results are in good agreement with the theoretical predication by the phase conditions as well as previous M3WM experiments.

Refer to caption
Figure 4: The dependence of ESST on the time delays of control fields for the target state |C|C\rangle. (a) The final population of the left-handed enantiomer PCL(tf)P_{C}^{L}(t_{f}) versus the time delays tba=tbtat_{ba}=t_{b}-t_{a} and tca=tctat_{ca}=t_{c}-t_{a}. (b) The cut line plot of PCL(tf)P_{C}^{L}(t_{f}) along tba=tcat_{ba}=t_{ca}.

To visualize the underlying quantum state transfer mechanism, Fig.3 shows the time-dependent populations of the four states induced by the control fields for the cases of τ0=35\tau_{0}=35 ns and ϕa=π/2\phi_{a}=\pi/2. For the two enantiomers, there are no visible populations in the state |B|B^{\prime}\rangle during the whole process. The four-level system is equivalent to the present closed-loop three-level model with the used pulse parameters. The control field a(t)\mathcal{E}_{a}(t) with the pulse-areas θa(t1)=π/4\theta_{a}(t_{1})=\pi/4 drives the system to the maximal coherent superposition of |A|A\rangle and |B|B\rangle with PAL,R(t1)=PBL,R(t1)=0.5P_{A}^{L,R}(t_{1})=P_{B}^{L,R}(t_{1})=0.5 for both enantiomers. Due to the sign difference of the transition from |A|A\rangle to |B|B\rangle, a(t)\mathcal{E}_{a}(t) with a phase ϕa=π/2\phi_{a}=\pi/2 will result in the phase of the state |B|B\rangle in 0 and π\pi for the left- and right-handedness, respectively, as described by Eq. (17). For the left-handedness, the transition path from |A|A\rangle to |C|C\rangle induced by b(t)\mathcal{E}_{b}(t) will constructively with the path from |B|B\rangle to |C|C\rangle by c(t)\mathcal{E}_{c}(t), leading to entire ESST to |C|C\rangle. For the right-handedness, however, the two paths are destructive, which keeps the molecules in the states |A|A\rangle and |C|C\rangle, as shown in Fig. 3 (a) and (b).
As can be seen from our theoretical derivations, we divide the closed-loop excitation scheme into two time-separated stages. To see whether the amplitude and phase conditions can be applied to the overlapped cases, as an example, Fig. 4 plots the landscape of PCL(tf)P_{C}^{L}(t_{f}) with respect to the time delays tba=tbtat_{ba}=t_{b}-t_{a} and tca=tctat_{ca}=t_{c}-t_{a} while fixing the center time tat_{a} unchanged. PCL(tf)P_{C}^{L}(t_{f}) strongly depends on the overlap between two control fields of the second stage. Interestingly, PCL(tf)P_{C}^{L}(t_{f}) remains the value of PCL(tf)>0.999P_{C}^{L}(t_{f})>0.999 for tba=tca>2τ0t_{ba}=t_{ca}>2\tau_{0} when b(t)\mathcal{E}_{b}(t) and c(t)\mathcal{E}_{c}(t) are turned on simultaneously with tba=tcat_{ba}=t_{ca}. Even the three control fields are applied without any delays, high fidelity of PCL(tf)>0.90P_{C}^{L}(t_{f})>0.90 holds, as shown in Fig. 4 (b). The phenomena can also be observed for the right-handedness (not shown here).

III.2 ESST to |B|B\rangle

Refer to caption
Figure 5: Chiral dependence of ESST on the phase for the target state |B|B\rangle. (a) The final population of the left-handed enantiomer PCL(tf)P_{C}^{L}(t_{f}) versus the duration τ0\tau_{0} and phase ϕb\phi_{b} in a range of [0,2π][0,2\pi], and the corresponding cut lines at (b) τ0=35\tau_{0}=35 ns and (c) ϕb=π/2\phi_{b}=\pi/2. (d)-(f) The same plots as (a)-(c) for the right-handed enantiomer except for ϕa=3π/2\phi_{a}=3\pi/2 in (f).

Figure 5 examines the same simulations as Fig. 2 but for the target |B|B\rangle with ϕa=ϕc=0\phi_{a}=\phi_{c}=0. In our simulations, we choose the parameters Ab=π/(4μa)A_{b}=\pi/(4\mu_{a}), Aa=π/(22μb)A_{a}=\pi/(2\sqrt{2}\mu_{b}), and Ac=π/(22μc)A_{c}=\pi/(2\sqrt{2}\mu_{c}) so as to the all fields satisfy the amplitude conditions. The influence of the state |B|B^{\prime}\rangle looks more visible than that in Fig. 2 in the short duration regime, which becomes rather weak by increasing duration τ0\tau_{0}. The final population PBL,R(tf)P_{B}^{L,R}(t_{f}) can also reach high fidelity for τ0>35\tau_{0}>35ns. As demonstrated in Fig. 2, the landscape of PBL,R(tf)P_{B}^{L,R}(t_{f}) exhibits a chiral symmetry with respect to the phase ϕb\phi_{b}, for which the control field b(t)\mathcal{E}_{b}(t) with ϕb=π/2\phi_{b}=\pi/2 leads to entire ESST to |B|B\rangle of the left-handedness. For the right-handedness, however, it requires to ϕb=3π/2\phi_{b}=3\pi/2. This dependence of PBL,R(tf)P_{B}^{L,R}(t_{f}) on the phase is consistent with the theoretical predication.

Refer to caption
Figure 6: The time-dependent populations of four rotational states for ESST to the state |B|B\rangle. The simulations for (a) the left-handed and (b) right-handed enantiomers with the duration τ0=35\tau_{0}=35ns and phase ϕb=π/2\phi_{b}=\pi/2.

Figure 6 plots the time-dependent populations of the system with τ0=35\tau_{0}=35 and ϕb=π/2\phi_{b}=\pi/2. Since the transition moments μb\mu_{b} are identical for the two enantiomers without a difference of sign, b(t)\mathcal{E}_{b}(t) plays the same role in the first stage, generating the same maximal coherent superposition of |A|A\rangle and |C|C\rangle. The constructive or destructive interference that depends on μa\mu_{a} occurs between the transition paths from |A|A\rangle and |C|C\rangle to |B|B\rangle. As a result, the left-handed enantiomer is fully transferred to the state |B|B\rangle, whereas the right-handed one is still in the coherent states |A|A\rangle and |B|B\rangle at end of three pulses.

Refer to caption
Figure 7: The dependence of ESST on the time delays of control fields for the target state |B|B\rangle. (a) The final population of the left-handed enantiomer PBL(tf)P_{B}^{L}(t_{f}) versus the time delays tba=tbtat_{ba}=t_{b}-t_{a} and tca=tctat_{ca}=t_{c}-t_{a}. (b) The cut line plot of PBL(tf)P_{B}^{L}(t_{f}) along tba=tcat_{ba}=t_{ca}.

To see the robustness of the scheme on the time delays, Fig. 7 examines the dependence of PBL(tf)P_{B}^{L}(t_{f}) on the time delays tab=tatbt_{ab}=t_{a}-t_{b} and tcb=tctbt_{cb}=t_{c}-t_{b}. Similar behaviors can be observed, indicating that both excitation schemes do not require strict separations between the control fields. The identical delays of the second stage fields are beneficial for the control. As a result, the amplitude and phase conditions can also be used for the overlapped control fields, leading to the high selectivity of handedness.

III.3 ESST with center frequency-detuned pulses

Finally, we examine the amplitude and phase conditions of control fields whose center frequencies ωa/b/c\omega_{a/b/c} are not exactly resonant with the transition center frequencies ωAB\omega_{AB}, ωBC\omega_{BC} and ωAC\omega_{AC}. As can be seen from the definitions of the complex pulse areas θa/b/c(tf)\theta_{a/b/c}(t_{f}), when the center frequencies are detuned away from resonances, the values of |θa/b/c(tf)||\theta_{a/b/c}(t_{f})| will be decreased while keeping the parameters Aa/b/cA_{a/b/c} unchanged as used in resonant cases. We can increase the values of Aa/b/cA_{a/b/c} to revive the values of θa/b/c(tf)\theta_{a/b/c}(t_{f}) at the transition frequencies so as to satisfy the amplitude conditions. That is, ESST in principle could be achieved by using the center-frequency-detuned pulses, as long as they satisfy the amplitude and phase conditions a/b/c\mathcal{E}_{a/b/c} and ϕa/b/c\phi_{a/b/c} at transition frequencies.

Refer to caption
Figure 8: The dependence of ESST on the detunings for the target state |B|B\rangle. and (b) The final population PBL(tf)P_{B}^{L}(t_{f}) versus the scale factor α\alpha of the coupling Ωb(t)\Omega_{b}(t) for different values of detuning Δb=ωbωAC\Delta_{b}=\omega_{b}-\omega_{AC}. (c) and (d) The final population PBL(tf)P_{B}^{L}(t_{f}) versus the scale factor β\beta of the couplings Ωa(t)\Omega_{a}(t) and Ωc(t)\Omega_{c}(t) by taking Δa=ωaωAB=Δc=ωcωBC=Δ\Delta_{a}=\omega_{a}-\omega_{AB}=\Delta_{c}=\omega_{c}-\omega_{BC}=\Delta. The analytical simulations (left panels) are compared with the exact results (right panels).

Figure 8 shows the dependence of PBL(tf)P_{B}^{L}(t_{f}) on the detunings, for which the analytical simulations (left panels) are compared with the exact ones (right panels). We perform the simulations in Figs. 8 (a) and (b) to calculate PBL(tf)P_{B}^{L}(t_{f}) for different values of Δb=ωbωAC\Delta_{b}=\omega_{b}-\omega_{AC} while scaling the coupling Ωb(t)\Omega_{b}(t) with a factor α\alpha, for which the center frequencies ωb\omega_{b} and ωc\omega_{c} are fixed at the resonant conditions. The simulations in Figs. 8 (c) and (d) are accomplished with different detunings Δa=ωaωAB\Delta_{a}=\omega_{a}-\omega_{AB} and Δc=ωcωBC\Delta_{c}=\omega_{c}-\omega_{BC} by scaling the couplings Ωa(t)\Omega_{a}(t) and Ωc(t)\Omega_{c}(t) with a factor β\beta, for which we set Δa=Δc=Δ\Delta_{a}=\Delta_{c}=\Delta while fixing Δb=0\Delta_{b}=0. For both analytical simulations, we can see that the detunings decrease PBL(tf)P_{B}^{L}(t_{f}). By increasing the strengths of the control fields, the maximal value of PBL(tf)P_{B}^{L}(t_{f}) can be revived to the same level as the resonant excitation, as shown in Figs. 8 (a) and (c). For the exact simulations, however, the maximal values can be increased but below the theoretical maximum. The differences can be attributed to the influence of high-order Magnus expansion terms, which are ignored in the analytical model. We also observe similar results for target state |B|B\rangle (not shown here). Thus, the center-frequency detuned excitations with small detunings are also allowed by applying the corresponding amplitude and phase conditions at transition frequencies, whereas the larger detunings will reduce the ESST efficiency due to the optical processes via high-order high-order Magnus terms.

IV Conclusion

We presented a general pulse-areas theorem analysis to explore ESST within a closed-loop three-level system. We considered three rotational states cyclically connected by the aa-type, bb-type, and cc-type components of the transition dipole molecules. Using a strategy that separates the closed-loop excitation into two stages, we derived the amplitude and phase conditions for designing three linearly polarized microwave pulses to generate ESST to different targets. The two-stage strategy we used differs from previous schemes that turned on the direct one-photon transition from the initial state to the target state before the indirect two-photon one. Our schemes firstly switched on one control field involved in the two-photon path by generating maximal coherent supposition between the initial and intermediate states. We examined this three-level pulse-areas theorem in the cyclohexylmethanol molecules and analyzed its applications with both center-frequency resonant and detuned pulse sequences. For the latter, small detunings on the center frequencies of the control pulses would be expected to reduce the influence of high-order Magnus expansion terms. It opens a fundamental question of whether one can design fast and robust quantum control schemes against center-frequency detunings. To that end, optimal and robust control methods combined with artificial intelligence algorithms could be used to search for shaped control pulses subject to multiple constraints shupra2019 ; QOCT2 ; shu6 ; IEEEshu ; Yang2020 .

Acknowledgements.
This work was supported by the National Natural Science Foundations of China (NSFC) under Grant No. 61973317. Y. G. is partially supported by the Scientific Research Fund of Hunan Provincial Education Department under Grant No. 20A025, Changsha Municipal Natural Science Foundation under Grant No. kq2007001, the Opening Project of Key Laboratory of Low Dimensional Quantum Structures and Quantum Control of the Ministry of Education under Grant No. QSQC1905, and the Open Research Fund of Hunan Provincial Key Laboratory of Flexible Electronic Materials Genome Engineering under Grant No. 202009.

References

  • (1) L. Pasteur, Recherches sur les relations qui peuvent exister entre la forme cristalline, la composition chimique et le sens de la polarisation rotatoire, Ann. Phys. Chem. 24, 442 (1848).
  • (2) M. Quack, J. Stohner, and M. Willeke, High-resolution spectroscopic studies and theory of parity violation in chiral molecules, Annu. Rev. Phys. Chem. 59,741 (2008).
  • (3) R. Naaman and D. H. Waldeck, Spintronics and chirality: spin selectivity in electron transport through chiral molecules. Annu. Rev. Phys. Chem. 66, 263 (2015).
  • (4) R. Naaman, Y. Paltiel, and D. H. Waldeck, Chiral molecules and the electron spin. Nat. Rev. Chem. 3, 250 (2019).
  • (5) M. R. Wasielewski, M. D. E. Forbes, N. L. Frank, K. Kowalski, G. D. Scholes, J. Yuen-Zhou, M. A. Baldo, D. E. Freedman, R. H. Goldsmith, T. Goodson III, M. L. Kirk, J. K. McCusker, J. P. Ogilvie, D. A. Shultz, S. Stoll, and K. B. Whaley , Exploiting chemistry and molecular systems for quantum information science. Nat. Rev. Chem. 4, 490 (2020).
  • (6) C. Vogt, Separation of D/L-carnitine enantiomers by capillary electrophoresis, J. Chromatogr. A 745,53 (1996).
  • (7) R. M. Hazen and D. S. Sholl, Chiral selection on inorganic crystalline surfaces, Nat. Mater. 2, 367 (2003).
  • (8) D. Patterson and M. Schnell, New studies on molecular chirality in the gas phase: enantiomer differentiation and determination of enantiomeric excess. Phys. Chem. Chem. Phys. 16, 11114 (2014).
  • (9) G. Heppke, A. Jákli, S. Rauch, and H. Sawade, Electric-field-induced chiral separation in liquid crystals, Phys. Rev. E 60,5575 (1999).
  • (10) P. Brumer,E. Frishman,and M. Shapiro, Principles of electric-dipole-allowed optical control of molecular Chirality, Phys. Rev. A 65,015401 (2001).
  • (11) I. Thanopulos, P. Král, and M. Shapiro, Theory of a two-step enantiomeric purification of racemic mixtures by optical means: The D2 S2 molecule, J. Chem. Phys. 119, 5105 (2003).
  • (12) E. Frishman, M.Shapiro, and P.Brumer, Optical purification of racemic mixtures by laser distillation in the presence of a dissipative bath, J. Phys. B:At. Mol. Opt. Phys. 37,2811 (2004).
  • (13) Y. Li, C. Bruder, and C. P. Sun, Generalized Stern-Gerlach effect for chiral molecules, Phys. Rev. Lett. 99,130403 (2007).
  • (14) Y. Li and C. Bruder, Dynamic method to distinguish between left- and right-handed chiral molecules, Phys. Rev. A 77,015403 (2008).
  • (15) C. Ye, Q. Zhang, and Y. Li, Real single-loop cyclic three-level configuration of chiral molecules, Phys. Rev. A 98,063401 (2018).
  • (16) C. Ye, Q. Zhang, Y. Chen, and Y. Li, Effective two-level models for highly efficient inner-state enantioseparation based on cyclic three-level systems of chiral molecules, Phys. Rev. A 100,043403 (2019).
  • (17) A. Yachmenev, J. Onvlee, E. Zak, A. Owens, and J. Küpper, Field-induced diastereomers for chiral separation, Phys. Rev. Lett. 123,243202 (2019).
  • (18) J.-L. Wu, Y. Wang, J.-X. Han, C. Wang, S.-L. Su, Y. Xia, Y. Jiang, and J. Song, Two-path interference for enantiomer-selective state transfer of chiral molecules, Phys. Rev. Applied 13, 044021 (2020).
  • (19) B. T. Torosov, M. Drewsen, and N. V. Vitanov, Efficient and robust chiral resolution by composite pulses, Phys. Rev. A 101, 063401 (2020).
  • (20) B. T. Torosov, M. Drewsen, and N. V. Vitanov, Chiral resolution by composite Raman pulses, Phys. Rev. Research 2, 043235 (2020).
  • (21) C. Ye, Q. Zhang, Y. Chen, and Y. Li, Fast Enantioconversion of chiral mixtures based on a four-level double-Δ\Delta model,Phys. Rev. Research 2,033064 (2020).
  • (22) J.-L. Wu, Y. Wang, S.-L. Su, Y. Xia, Y. Y. Jiang, and J. Song, Discrimination of enantiomers through quantum interfernce and quantum Zeno effect, Opt. Express 28, 33475 (2020).
  • (23) Y. Chen, C. Ye, Q. Zhang, and Y. Li, Enantio-discrimination via light deflection effect, J. Chem. Phys. 152,204305 (2020).
  • (24) X. Xu, C. Ye, Y. Li, and A. Chen, Enantiomeric-excess determination based on nonreciprocal-transition-induced spectral-line elimination, Phys. Rev. A 102,033727 (2020).
  • (25) C. Ye, B. Liu, Y.-Y. Chen, and Y. Li, Enantio-conversion of chiral mixtures via optical pumping, Phys. Rev. A 103,022830 (2021).
  • (26) I. Tutunnikov, L. Xu, R. W. Field, K. A. Nelson, Y. Prior, and I. S. Averbukh, Enantioselective orientation of chiral molecules induced by terahertz pulses with twisted polarization, Phys. Rev. Research 3,013249 (2021).
  • (27) K. Bergmann, H. Theuer, and B. W. Shore, Coherent population transfer among quantum states of atoms and molecules, Rev. Mod. Phys. 70, 1003 (1998).
  • (28) N. V. Vitanov, A. A. Rangelov, B. W. Shore, and K. Bergmann, Stimulated Raman adiabatic passage in physics, chemistry, and beyond, Rev. Mod. Phys. 89, 015006 (2017).
  • (29) X. Chen, I. Lizuain, A. Ruschhaupt, D. Guéry-Odelin, and J. G. Muga, Shortcut to adiabatic passage in two- and three-level atoms, Phys. Rev. Lett. 105, 123003 (2010).
  • (30) D. Guéry-Odelin, A. Ruschhaupt, A. Kiely, E. Torrontegui, S. Martínez-Garaot, and J. G. Muga, Shortcuts to adiabaticity: concepts, methods, and applications, Rev. Mod. Phys. 91, 045001 (2019).
  • (31) M. Shapiro, E. Frishman, and P. Brumer, Coherently controlled asymmetric synthesis with achiral light, Phys. Rev. Lett. 84,1669 (2000).
  • (32) P. Kra’l and M. Shapiro, Cyclic population transfer in quantum systems with broken symmetry, Phys. Rev. Lett. 87, 183002 (2001).
  • (33) P. Kra’l, I. Thanopulos, M. Shapiro, and D. Cohen, Two-step enantio-selective optical switch, Phys. Rev. Lett. 90, 033001 (2003).
  • (34) D. Gerbasi, M. Shapiro, and P. Brumer, Theory of enantiomeric control in dimethylallene using achiral light, J. Chem. Phys. 115,5349 (2001).
  • (35) N. V. Vitanov and M. Drewsen, Highly efficient detection and separation of chiral molecules through shortcuts to adiabaticity, Phys. Rev. Lett. 122,173202 (2019).
  • (36) J. Wu, Y. Wang, J. Song, Y. Xia, S. Su, and Y. Jiang, Robust and highly efficient discrimination of chiral molecules through three-mode parallel paths, Phys. Rev. A 100,043413 (2019).
  • (37) E. Frishman, M. Shapiro, D. Gerbasi, and P. Brumer, Enantiomeric purification of nonpolarized racemic mixtures using coherent light, J. Chem. Phys. 119,7237 (2003).
  • (38) H. R. Hamedi, E. Paspalakis, G. Žlabys, G. Juzeliūnas, and J. Ruseckas, Complete energy conversion between light beams carrying orbital angular momentum using coherent population trapping for a coherently driven double-Λ\Lambda atom-light-coupling scheme, Phys. Rev. A 023811; 102, 019903(E) (2020).
  • (39) D. Patterson and J. M. Doyle, Sensitive chiral analysis via microwave three-wave mixing, Phys. Rev. Lett. 111,023008 (2013).
  • (40) D. Patterson, M. Schnell, and J. M. Doyle, Enantiomer-specific detection of chiral molecules via microwave spectroscopy, Nature 497, 475 (2013).
  • (41) V. A. Shubert, D. Schmitz, D. Patterson, J. M. Doyle, and M. Schnell, Identifying enantiomers in mixtures of chiral molecules with broadband microwave spectroscopy, Angew. Chem. Int. Ed. 53, 1152 (2014).
  • (42) V. A. Shubert, D. Schmitz, C. Medcraft, A. Krin, D. Patterson, J. M. Doyle, and M. Schnell, Rotational spectroscopy and three-wave mixing of 4-Carvomenthenol: a technical guide to measuring chirality in the microwave regime, J. Chem. Phys. 142, 214201 (2015).
  • (43) S. Lobsiger, C. Pérez, L. Evangelisti, K. K. Lehmann, and B. H. Pate, Molecular structure and chirality detection by fourier transform microwave spectroscopy, J. Phys. Chem. Lett. 6, 196 (2015).
  • (44) S. Eibenberger, J. Doyle, and D. Patterson, Enantiomer-specific state transfer of chiral molecules, Phys. Rev. Lett. 118, 123002 (2017).
  • (45) S. R. Domingos, C. Perez, and M. Schnell, Sensing chirality with rotational spectroscopy, Annu. Rev. Phys. Chem. 69, 499 (2018).
  • (46) C. Pérez, A. L. Steber, A. Krin, and M. Schnell, Statespecific enrichment of chiral conformers with microwave spectroscopy, J. Phys. Chem. Lett. 9, 4539 (2018).
  • (47) M. Leibscher, T. F. Giesen, and C. P. Koch, Principles of enantio-selective excitation in three-wave mixing spectroscopy of chiral molecules, J. Chem. Phys. 151,014302 (2019).
  • (48) D. Sugny and C. Kontz, Optimal control of a three-level quantum system by laser fields plus von Neumann measurements, Phys. Rev. A 77,063420 (2008).
  • (49) Y. Guo, L. Zhou, L. M. Kuang, and C. P. Sun, Magneto-optical Stern-Gerlach effect in an atomic ensemble. Phys. Rev. A 78, 013833 (2008).
  • (50) G. Shchedrin, C. O’Brien, Y. Rostovtsev, and M. O. Scully, Analytic solution and pulse area theorem for three-level atoms, Phys. Rev. A 92,063815 (2015).
  • (51) Y. Guo, X. B. Luo, S. Ma, and C.-C. Shu, All-optical generation of quantum entangled state with strict constrained ultrafast laser pulses, Phys. Rev. A 100, 023409 (2019).
  • (52) B. Y. Chang, I. R. Sola, and V. S. Malinovsky, Anomalous rabi oscillations in multilevel quantum systems, Phys. Rev. Lett. 120,133201 (2018).
  • (53) H. Mineo, G.-S. Kim, S. H. Lin, and Y. Fujimura, Dynamic stark-induced coherent π\pi-electron rotations in a chiral aromatic ring molecule: application to phenylalanine, J. Phys. Chem. A. 123,6399 (2019).
  • (54) Y Guo, C.-C. Shu, D. Y. Dong, and F. Nori, Vanishing and revival of resonance Raman scatering, Phys. Rev. Lett. 123,223202 (2019).
  • (55) C.-C. Shu, Q.-Q. Hong, Y. Guo, and N. E. Henriksen, Orientational quantum revivals induced by a single-cycle terahertz pulse, Phys. Rev. A 102, 063124 (2020).
  • (56) C.-C. Shu, Y. Guo, K.-J. Yuan, D. Y. Dong, and A. D. Bandrauk, Attosecond all-optical control and visualization of quantum interference between degenerate magnetic states by circularly polarized pulses. Opt. Lett. 45, 960 (2020).
  • (57) Q.-Q. Hong, L.-B. Fan, C.-C. Shu and N. E. Henriksen, Generation of maximal three-state field-free molecular orientation with terahertz pulses, Phys. Rev. A 104, 013108 (2021).
  • (58) S. Blanes, F. Casas, J. A. Oteo, and J. Ros, The Magnus expansion and some of its applications, Phys. Rep. 470, 151 (2009).
  • (59) K. K. Lehmann, Influence of spatial degeneracy on rotational spectroscopy: three-wave mixing and enantiomeric state separation of chiral molecules, J. Chem. Phys. 149, 094201 (2018).
  • (60) C.-C. Shu, T.-S. Ho, X. Xing, and H. Rabitz, Frequency domain quantum optimal control under multiple constraints. Phys. Rev. A 93, 033417 (2016).
  • (61) Y. Guo, D. Dong, and C.-C. Shu, Optimal and robust control of quantum state transfer by shaping spectral phase of ultrafast laser pulses. Phys. Chem. Chem. Phys. 20, 9498 (2018).
  • (62) D. Dong, C.-C. Shu, J. C. Chen, X. Xing, H. L. Ma, Y. Guo, and H. Rabitz, Learning control of quantum systems using frequency-domain optimization algorithms. IEEE Trans. Control. Syst Technol. 29, 1791 (2021).
  • (63) X. W. Liu, G, J. Zhang, J. Li, G. L. Shi, M. Y. Zhou, B. Q. Huang, Y. J. Tang, X.H. Song, and W.F. Yang, Deep learning for Feynman’s path integral in strong-field time-dependent dynamics, Phys. Rev. Lett. 124, 113202 (2020).