Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Robust and high-efficiency dynamical method of enantio-specific state transfer

Open Access Open Access

Abstract

We propose a simple dynamical method to realize fast enantio-specific state transfer (ESST) of chiral molecules. Driven by three external electromagenetic fields, the chiral molecules are modeled as cyclic three-level systems, where the overall phase differs by π for the left- and right-handed chiral molecules. We unveil that the ESST is allowed when the amplitudes of three Rabi frequencies in the cyclic three-level systems are equal. Our method is robust and highly efficient in the sense that the external fields can have arbitrary waveforms. This thus provides the opportunity of simplifying the experimental implementations of ESST through pulse design.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

It is known that a chiral molecule cannot be transformed into its mirror counterpart by only spatial translations and/or rotations [1]. The left- and right-handed chiral molecules, also known as enantiomers, are ubiquitous and play important roles in chemistry [2], biotechnologies [3], and pharmaceutics [4] due to their peculiar properties. Even though the two enantiomers may exhibit identical physical properties, they can however have completely distinct chemical properties. This holds critical significance in the field of pharmaceutics. Therefore, enantiodiscrimination [5] (as well as enantioseparation [6,7] and enantioconversion [8]) of chiral molecules is an essential and urgent task. Conventional methods of enantiodiscrimination are based on the interference between magnetic-dipole (or electrical-quadrupole) and electronic-dipole interactions [912] between light and molecules, which suffers from the fact that magnetic-dipole or electrical quadrupole interactions are typically very weak.

About twenty years ago, cyclic three-level systems of chiral molecules were proposed and investigated [13,14], where three electromagnetic fields are applied to couple to the three electronic-dipole transitions, respectively. Such cyclic three-level systems are similar for the left- and right-handed chiral molecules except for a difference $\pi$ between the overall phases of their closed-loop transition paths [13,14]. This chirality dependence of the overall phase makes the dynamics of enantiomers different, allowing for enantio-specific state transfer (ESST) of chiral molecules (i.e., the left- and right-handed chiral molecules can be transferred to different-energy target states from the same-energy initial ones, e.g. the ground ones) [1329]. In experiments, ESSTs of three-level chiral molecules have been demonstrated in gaseous samples by applying three well-designed microwave fields [3035]. In addition, a variety of theoretical methods of direct enantiodiscrimination [3648] and enantioconversion [14,4953] (together with spatial enantioseparation [5458]) have also been proposed for chiral molecules with cyclic three-level (sub-) structures.

Among the above seminal works the method of coherently controlled adiabatic passage [13,14] appears to be a pathbreaking one, yet it typically leads to slow and complicated ESST processes due to the requirement of intrinsic adiabaticity. Subsequently, fast ESST methods based on resonant ultrashort pulses and the resulting nonadiabatic dynamics were proposed [1518]. Moreover, Vitanov et al. proposed recently a dynamical method based on shortcuts to adiabaticity to realize the fast ESST [19]. Inspired by this, a series of theoretical extensions have been proposed and developed [2029].

Most recently, a fast ESST method was proposed based on designing a desired population transfer process for only one enantiomer [26], in contrast to the previous methods [13,1623] which focused on designing population transfer processes for both enantiomers simultaneously. The key element of the method in Ref. [26] is that the three-level systems of the chiral molecules possess specific SU(2) algebraic structures (where two of the Rabi frequencies are equal and the overall phase is fixed as $\pm \pi /2$). Such a structure leads to the high-efficiency ESST where the populations transferred to the two excited states are exchanged between the left- and right-handed chiral molecules if they are initially prepared in the ground states of the three-level systems.

Here we further investigate the ESST of cyclic three-level chiral molecules where all the Rabi frequencies have identical magnitude. This requirement makes it possible to realize the desired state transfer process from the ground state to the target excited one for one enantiomer (e.g., the left-handed chiral molecules) by using external fields with arbitrary pulse waveforms and appropriate values of pulse area of the Rabi frequency. Simultaneously, the other enantiomer (e.g., the right-handed chiral molecules) can be automatically transferred to the other excited state from the ground one. Therefore, by tuning the pulse area of the external fields, the left- and right-handed chiral molecules can evolve to different excited states from the initial ground states and thus exhibit fast ESST. Our current method shows the flexibility in controlling the waveform and overall phase of the pulses. In contrast, the method proposed in Ref. [26] requires precise control over these parameters. This flexibility makes our fast ESST method easier to implement and control, benefiting from its robustness and high efficiency.

In the previous ESST methods [1518,31,35], the ESST processes of the cyclic three-level chiral molecules are typically accomplished by resorting to several successive pulses. In contrast, here we use three pulses which are applied simultaneously to the enantiomers to realize fast ESST. In addition, compared with the shortcuts-to-adiabaticity methods [19,27] where all the three electromagnetic fields are also applied simultaneously, our method shows the advantage in terms of the arbitrary pulse waveforms.

2. Cyclic three-level systems of enantiomers

The two enantiomers (i.e., the left- and right-handed chiral molecules) can be modeled as cyclic three-level systems based on the electric-dipole transitions when they are subject to three external electromagnetic fields [13,59]. Here, we only consider the case where all the three external fields are coupled resonantly to the corresponding electric-dipole transitions, respectively. Under the rotating-wave approximation, the Hamiltonian of the cyclic three-level systems can be written in the basis $\{|m\rangle _{Q}\}~(m=1,2,3)$ as ($\hbar =1$)

$$\hat{H}^{Q}(t)=\Omega_{12}(t)|1\rangle_{QQ}\langle2|+\Omega_{23}(t)|2\rangle_{QQ}\langle3| +\Omega_{13}(t)e^{{-}i\phi_{Q}}|1\rangle_{QQ}\langle3|+\textrm{H.c.},$$
where the scripts $Q=L,R$ have been introduced to denote the left- and right-handed chiral molecules, respectively. Without loss of generality, we assume real Rabi frequencies $\Omega _{mn}(t)~(3\geq n>m\geq 1)$ for all the transitions $|m\rangle _{QQ}\langle n|$. Here $\phi _{Q}$ are the overall phases of the cyclic three-level systems, which differ by $\pi$ for the left- and right-handed chiral molecules due to the mirror-reflection symmetry of the electric-dipole moments [35]. The chirality of the enantiomers is specified by choosing the overall phases
$$\phi_{L}=\phi,~\phi_{R}=\phi+\pi.$$

Here we only focus on the case

$$\Omega_{12}(t)=\Omega_{13}(t)=\Omega_{23}(t)\equiv\Omega(t),$$
which simply means that all the Rabi frequencies have identical magnitude. This requirement does not depend on specific molecular parameters, hence the cyclic three-level systems in Fig. 1 can be readily constructed in experiments.

 figure: Fig. 1.

Fig. 1. Cyclic three-level systems of (a) left- and (b) right-handed chiral molecules. Three external electromagnetic fields are coupled resonantly to the electric-dipole transitions, respectively, with overall phase $\phi _{Q}$ and equal Rabi frequency $\Omega$.

Download Full Size | PDF

To illustrate our method, we use the 1,2-propanediol molecule as an example to construct the cyclic three-level system by selecting appropriate rotational states $|J,\tau,M\rangle$ [59] (with electronic and vibrational ground state). Here, $J$ represents the total angular momentum, $\tau$ ranges from $-J$ to $J$ in ascending energy order, and $M$ corresponds to degenerate magnetic sub-levels. The corresponding cyclic three-level system, as described in Refs. [35,59], consists of $|1\rangle =|0,0,0\rangle$, $|2\rangle =|1,-1,0\rangle$, and $|3\rangle =(|1,1,1\rangle +|1,1,-1\rangle )/\sqrt {2}$. Three microwave pulses are utilized to couple resonantly with the three transitions in the three-level system respectively and simultaneously. Specifically, the transition $|1\rangle \leftrightarrow |2\rangle$ (with frequencies $2\pi \times 6431.06\, \textrm {MHz}$) is couple resonantly to a Z-polarized microwave pulse. Meanwhile, the transitions $|1\rangle \leftrightarrow |3\rangle$ and $|2\rangle \leftrightarrow |3\rangle$ (with frequencies $2\pi \times 12212.15\, \textrm {MHz}$ and $2\pi \times 5781.09\, \textrm {MHz}$, respectively) are couple resonantly to Y- and X-polarized microwave pulses. Achieving equal Rabi frequencies in such a cyclic three-level system becomes very simple due to the comparable magnitudes of dipole moments associated with these three rotational transitions [59]. We would like to remark that the technique involving three microwaves has already been implemented in many experiments of cyclic three-level systems for chiral molecules [3035].

3. Enantio-specific state transfer via dynamical processes

We assume that the two enantiomers stay in their ground states $|1\rangle _{Q}$ initially. In this section (as well as Sec. 4.), we consider the case of $\phi =\pi /2$ as an example. The dynamical evolutions of the two enantiomers can be captured by the Schrödinger equation $i\partial _{t}|\psi (t)\rangle _{Q}=\hat {H}^{Q}(t)|\psi (t)\rangle _{Q}$. For instance, one can assume the evolved state $|\psi (t)\rangle _{L}=\sum _{m=1,2,3}C^{L}_{m}(t)|m\rangle _{L}$ for the left-handed molecules, where $C_{m}^{L}$ represents the excitation amplitude of the state $|m\rangle _{L}$. Under the initial condition $C^{L}_{1}(0)=1, C^{L}_{2}(0)=0$, and $C^{L}_{3}(0)=0$, the evolved state can be obtained by solving the dynamical equations as

$$\begin{aligned} C_{1}^{L}(t)&=\frac{1}{3}\left(1+2\cos[\sqrt{3}A(t)]\right),\\ C_{2}^{L}(t)&={-}\frac{i}{3}\left(1-\cos[\sqrt{3}A(t)]+\sqrt{3}\sin[\sqrt{3}A(t)]\right),\\ C_{3}^{L}(t)&={-}\frac{1}{3}\left(1-\cos[\sqrt{3}A(t)]-\sqrt{3}\sin[\sqrt{3}A(t)]\right), \end{aligned}$$
where the pulse area
$$A(t)=\int_{0}^{t}\Omega(t')dt'.$$

Likewise, the evolved state $|\psi (t)\rangle _{R}=\sum _{m=1,2,3}C^{R}_{m}(t)|m\rangle _{R}$ of the right-handed chiral molecules can be obtained by solving

$$\begin{aligned} C_{1}^{R}(t)&=\frac{1}{3}\left(1+2\cos[\sqrt{3}A(t)]\right),\\ C_{2}^{R}(t)&=\frac{i}{3}\left(1-\cos[\sqrt{3}A(t)]-\sqrt{3}\sin[\sqrt{3}A(t)]\right),\\ C_{3}^{R}(t)&={-}\frac{1}{3}\left(1-\cos[\sqrt{3}A(t)]+\sqrt{3}\sin[\sqrt{3}A(t)]\right) \end{aligned}$$
under the initial condition $C^{R}_{1}(0)=1, C^{R}_{2}(0)=0$, and $C^{R}_{3}(0)=0$. Comparing Eqs. (4) and (6), we have
$$C_{1}^{L}(t)=C_{1}^{R}(t),\,C_{2}^{L}(t)=iC_{3}^{R}(t),\,C_{3}^{L}(t)=iC_{2}^{R}(t),$$
which lead to
$$P_{1}^{L}(t)=P_{1}^{R}(t),~~P_{2}^{L}(t)=P_{3}^{R}(t),~~P_{3}^{L}(t)=P_{2}^{R}(t)$$
with $P_{m}^{Q}(t)=|C_{m}^{Q}(t)|^{2}~(m=1,2,3)$ being the populations in the state $|m\rangle _{Q}$. It is clear that the ground state populations of the two enantiomers are always equal, while their populations in the two excited states are exactly exchanged. This means that one can design a concise scheme of population transfer for one enantiomer (e.g., the left-handed chiral molecules) such that it can evolve from the initial ground state to a target excited state. Simultaneously, the other enantiomer (e.g., the right-handed ones) would evolve to the other excited state satisfying Eq. (8), thus leading to the realization of ESST. When the left-handed chiral molecules exhibit perfect population transfer from $|1\rangle _{L}$ to $|2\rangle _{L}$, the right-handed ones would be completely transferred from $|1\rangle _{R}$ to $|3\rangle _{R}$. This occurs as long as the pulse area satisfies
$$A=\frac{2\sqrt{3}}{3}k\pi+\frac{2\sqrt{3}\pi}{9}~\,(k= 0,\pm1,\ldots)$$
in the case of $\phi =\pi /2$. This further implies that the ESST can be realized by tuning the pulse area $A$ of the external fields.

4. Flexibility of the pulse waveforms

In Sec. 3, we have identified in Eq. (9) the condition for realizing ESST of the three-level chiral molecules. Such a condition is clearly irrelevant to the shape of the pulse waveform. Our scheme is thus flexible in that one can in principle use arbitrary types of pulse waveforms to realize fast ESST.

Here we assume that the characteristic length of the molecular sample is much smaller than the beam radii of the electromagnetic fields, so that the Rabi frequencies are almost uniform for the molecules. For the sake of simplicity, one can choose pulses of square waveforms with typical experimentally available Rabi frequency [31,32]

$$\Omega(t)=\Omega_{0}=2\pi\times10\,\text{MHz}.$$

In order to achieve the ESST, the duration time of the pulse is set to be $T_{0}=19.2\,\text {ns}$ so that the related pulse area $A=\frac {2\sqrt {3}\pi }{9}$. Note that using square pulses typically requires switching the pulses on and off abruptly, which presents a challenge in precisely controlling the pulse area to achieve a specific value. In practical applications, it is preferable to smoothly turn on and off the applied pulses. In view of this, the waveform can be chosen as a single-period cosine-like one, i.e.,

$$\Omega(t)=\frac{\Omega_{0}}{2}\left(1-\cos\frac{2\pi t}{T}\right)$$
with $T=38.5\,\text {ns}$ being the period. For a pulse in Eq. (11), the corresponding duration time can be set to be $t=T=38.5\,\text {ns}$ to ensure $A=\frac {2\sqrt {3}\pi }{9}$. Moreover, one can also choose a Gaussian waveform
$$\Omega(t)=\Omega_{0}\,\mathrm{exp}\left[-(t-3T')^{2}/T'^{2}\right]$$
with $T'=10.8\,\text {ns}$ being the width. When the corresponding duration time is set to be $t=6T'=64.8\,\text {ns}$, the integral in Eq. (12) leads to $A\simeq \frac {2\sqrt {3}\pi }{9}$.

Note that in this section we have assumed $\phi =\pi /2$ and identical magnitude for the three Rabi frequencies of the cyclic three-level systems. In Fig. 2, we plot the evolved populations in $|2\rangle _{Q}$ and $|3\rangle _{Q}$ for the two enantiomers in this case. It shows that the chiral molecules in the initial state $|1\rangle _{L}\,(|1\rangle _{R})$ are transferred to the state $|2\rangle _{L}\,(|3\rangle _{R})$ eventually, regardless of the explicit waveform of the pulses. All the three waveforms in Eqs. (10)–(12) enable the desired ESST, where the left-handed molecules are excited into $|2\rangle _{L}$ and the right-handed ones are excited into $|3\rangle _{R}$ from the initial ground state $|1\rangle _{Q}$ after the evolution with $A=\frac {2\sqrt {3}\pi }{9}$. Such robustness allows us to employ a large variety of pulses, thus offering the flexibility in implementing ESST.

 figure: Fig. 2.

Fig. 2. (a), (c), and (e) Pulse waveforms given in Eqs. (10), (11), and (12), respectively. (b), (d), and (f) Evolved populations for the left- and right-handed chiral molecules: $P_2^L(t)=P_3^R(t)$ (red solid line) and $P_3^L(t)=P_2^R(t)$ (blue dashed line), corresponding to the waveforms in panels (a), (c), and (e), respectively. Here the overall phase is $\phi =\pi /2$ and the initial states are assumed as $|1\rangle _{L,R}$.

Download Full Size | PDF

5. Flexibility of the overall phase and the pulse area

In Sec. 3 and Sec. 4, we focused on the case where the overall phase $\phi$ is fixed as $\pi /2$. In fact, the overall phase can be tuned by adjusting the initial phases of the three electromagnetic fields. Note that when the three Rabi frequencies have equal magnitude, the Hamiltonians $H^{Q}(t)$ in Eq. (1) at different time instants always commute with each other, i.e., $[H^{Q}(t),H^{Q}(t')]=0$, regardless of the overall phase. This allows to describe the evolution of the system using a time-evolution operator (instead of a general time-ordering operator)

$$\begin{aligned} \hat{U}^{Q}(t)&=e^{{-}i\int_{0}^{t} \hat{H}^{Q}(t')dt'}\\ &=\textrm{exp}[{-}iA(t)(|1\rangle_{QQ}\langle2|+|2\rangle_{QQ}\langle3| +e^{{-}i\phi_{Q}}|1\rangle_{QQ}\langle3|+\textrm{H.c.})], \end{aligned}$$
which would make the calculation easier.

To better illustrate the discussions, we now consider a general cyclic three-level system where the Rabi frequencies have equal magnitude. The corresponding Hamiltionian can be expressed, in the basis of $\{|1\rangle, |2\rangle, |3\rangle \}$, as

$$\hat{H}(t)=\Omega \left( \begin{array}{ccc} 0 & 1 & e^{{-}i\phi} \\ 1 & 0 & 1 \\ e^{i\phi} & 1 & 0\\ \end{array} \right).$$

By solving the Schrödinger equation $i\partial _{t}|\psi (t)\rangle =\hat {H}(t)|\psi (t)\rangle$ with the initial condition $|\psi (0)\rangle =|1\rangle$, the evolved state can be formally obtained as

$$\begin{aligned} |\psi(t)\rangle&=e^{{-}i\int_{0}^{t} \hat{H}(t')dt'}|\psi(0)\rangle\\ &=C_{1}(t)|1\rangle+C_{2}(t)|2\rangle+C_{3}(t)|3\rangle. \end{aligned}$$

If we perform the transformation $|1'\rangle =ie^{-i\phi }|1\rangle, |2'\rangle =-i|3\rangle$, and $|3'\rangle =i|2\rangle$, the Hamiltonian in Eq. (14) can be re-written as

$$\hat{H'}(t)=\Omega \left( \begin{array}{ccc} 0 & -1 & e^{i\phi} \\ -1 & 0 & -1 \\ e^{{-}i\phi} & -1 & 0\\ \end{array} \right)',$$
where the prime is used to label the new basis $\{|1'\rangle, |2'\rangle, |3'\rangle \}$ after the transformation. Similarly, the corresponding evolved state $|\psi (t)\rangle '$ can be obtained by solving the Schrödinger equation $i\partial _{t}|\psi (t)\rangle '=\hat {H}'(t)|\psi (t)\rangle '$ with the initial condition $|\psi (0)\rangle '=-ie^{i\phi }|1\rangle '$ as
$$\begin{aligned} |\psi(t)\rangle'&=e^{{-}i\int_{0}^{t} \hat{H'}(t')dt'}|\psi(0)\rangle'\\ &=C'_{1}(t)|1\rangle'+C'_{2}(t)|2\rangle'+C'_{3}(t)|3\rangle'. \end{aligned}$$

Since Eqs. (15) and (17) should give the same evolved states, one has

$$C_{1}(t)=ie^{{-}i\phi}C'_{1}(t),~C_{2}(t)=iC'_{3}(t), C_{3}(t)={-}iC'_{2}(t).$$

Here we note that the equation $-i\partial _{t}|\psi ^{*}(t)\rangle '=\hat {H}'^{*}(t)|\psi ^{*}(t)\rangle '$ should be satisfied via taking complex conjugate for the Schrödinger equation $i\partial _{t}|\psi (t)\rangle '=\hat {H}'(t)|\psi (t)\rangle '$. That means the complex conjugate $|\psi ^{*}(t)\rangle '$ of the state in Eq. (17), which follows

$$\begin{aligned} |\psi^{*}(t)\rangle'&=e^{{-}i\int_{0}^{t}[ -\hat{H}'^{*}(t')]dt'}|\psi^{*}(0)\rangle'\\ &=C'^{*}_{1}(t)|1\rangle'+C'^{*}_{2}(t)|2\rangle'+C'^{*}_{3}(t)|3\rangle', \end{aligned}$$
is clearly subject to the equivalent Hamiltonian
$$-\hat{H}'^{*}(t)=\Omega \left( \begin{array}{ccc} 0 & 1 & e^{{-}i(\phi+\pi)} \\ 1 & 0 & 1 \\ e^{i(\phi+\pi)} & 1 & 0\\ \end{array} \right)'.$$

In this way, one can identify the relation between the populations $P_{m}(t)=|C_{m}(t)|^{2}$ for Hamiltonian (14) and $P'_{m}(t)=|C'_{m}(t)|^{2}$ for Hamiltonian (16,20), i.e.,

$$P_{1}(t)=P'_{1}(t),~P_{2}(t)=P'_{3}(t),~P_{3}(t)=P'_{2}(t).$$

We would like to point out that when the Hamiltonian of the left-handed chiral molecules in the basis of $\{|m\rangle _{L}\}$ has the same form as Eq. (14), the Hamiltonian of the corresponding right-handed ones in the basis of $\{|m\rangle _{R}\}$ should have the same form as Eq. (20). According to Eq. (21), for left- and right-handed chiral molecules evolved from the initial ground states (i.e. $|1\rangle _{L,R}$, up to a constant phase factor), we have

$$P_{1}^{L}(t)=P_{1}^{R}(t),~P_{2}^{L}(t)=P_{3}^{R}(t),~P_{3}^{L}(t)=P_{2}^{R}(t).$$

This implies that the populations occupying the two excited states are always exchanged between the two enantiomers during the evolution, as long as the magnitude of the three Rabi frequencies are equal, regardless of the overall phase $\phi$.

Now we turn to study ESST for other general values of $\phi$ (i.e., $\phi \neq \pi /2$). We plot in Fig. 3(a) the populations of $|2\rangle _{L}$ or $|3\rangle _{R}$ after the evolution following Eq. (13) as a function of the pulse area $A$ and the overall phase $\phi$. As expected, one can find many regions where the efficiency of the ESST is very high (over 95%). In addition to the case of $\phi = \pi /2$, the nearly perfect ESST (with approximately 100% efficiency) can also be realized for many other values of $\phi$, such as $\phi = 0.18\pi$ (the black solid line) as shown in Fig. 3(b). This is accomplished by choosing an appropriate pulse area $A$ for a given $\phi$ (e.g., $A\simeq 3.2$ for $\phi = 0.18\pi$).

 figure: Fig. 3.

Fig. 3. (a) Populations of $|2\rangle _{L}$ or $|3\rangle _{R}$ in the final state versus the overall phase $\phi$ and the pulse area $A$. (b) Populations of $|2\rangle _{L}$ and $|3\rangle _{R}$ in the final state versus the pulse area $A$ with two chosen values of the overall phase $\phi =0.18\pi$ (the black solid line) and $\phi =0.5\pi$ (the green dashed line), corresponding to the black solid and green dashed lines in panel (a), respectively. The initial states are assumed as $|1\rangle _{L,R}$.

Download Full Size | PDF

Moreover, we conclude from Fig. 3(a) that our ESST scheme is robust against weak perturbations in both the overall phase and the pulse area. For example, a high-efficiency ESST (with the populations in the target states more than 95%) can still be achieved even if the overall phase deviates from the predetermined value of $\phi =0.5\pi$ by $\pm 0.1\pi$, or if the pulse area deviates from the value of $A\simeq 1.2$ by $\pm 0.15$. In the current scheme, the overall phase provides an alternative controllable knob, with which the high-efficiency ESST can be achieved despite some possible implementation imperfections.

We would like to point out that the previous method proposed in Ref. [26] also ensures the interesting exchange symmetry of the excited-state populations for the two enantiomers. It suffers from a limitation, however, that the overall phase is required to be $\pm \pi /2$ to protect such an exchange symmetry. In contrast, our current method enables the exchange symmetry regardless of the value of the overall phase. Moreover, the previous method [26] demands well-designed pulse waveform shapes to accomplish the ESST. As a result, it is extremely sensitive to variations in the pulse waveforms. However, our current method releases this limitation and thus simplifies the experimental implementation of ESST. Consequently, our current method shows advantages in terms of the robustness and experimental complexity.

6. Robustness

So far we have examined our ESST method with ideal parameters. In practice, however, the systems are inevitably influenced by various perturbations. Therefore, it is important to investigate the effects of possible perturbations which may arise from the implementation imperfections of ESST. Now let us consider the effect of the relaxation and dephasing to the ESST process under consideration.

Here we take pulses with single-period cosine-like waveforms as an example to realize ESST. In experiments, such pulses can be readily generated by using standard microwave techniques. However, random amplitude noises are typically unavoidable due to various implementation imperfections. Therefore, it is necessary to take into account such noises, which cause the Rabi frequencies to perturb randomly. Here we model the noises by using the random Rabi frequencies

$$\Omega^{\text{rand}}_{mn}(t)=\frac{\Omega_{0}}{2}\,\left(1-\cos\frac{2\pi t}{T}\right)[1+\text{rand}_{mn}(t,\eta_{mn})]$$
with $\Omega _{0}=2\pi \times 10\,\text {MHz}$ and the period $T=38.5\,\text {ns}$. In Eq. (23), the function “rand” generates a random sequence in the range of $[-\eta _{mn},\eta _{mn}]$. In practice, the amplitudes of the three pulses are supposed to perturb randomly and independently. We introduce three different random perturbations, i.e., $\eta _{12}=0.3$, $\eta _{23}=0.4$, and $\eta _{13}=0.5$, respectively, for the three pulses. It is clear that in the absence of the perturbations the three Rabi frequencies have equal magnitude and the corresponding pulse area is $A=\frac {2\sqrt {3}\pi }{9}$ for the pulse duration time $t=T$.

As shown in Fig. 4, although the waveforms deviate from the ideal cosine-like shape due to the noises, the time evolutions of the populations in $|2\rangle _{Q}$ and $|3\rangle _{Q}$ are nearly unaffected in all three cases. The ESST can still show a nearly $100{\%}$ efficiency ESST as long as $t=T$ and $A=\frac {2\sqrt {3}\pi }{9}$. The exchange symmetry of the excited-state populations of the two enantiomers is also almost unaffected by the random amplitude noises, as shown in Fig. 4(b). This robustness is guaranteed by the zero mean value of the perturbations, with which the pulse areas remain almost unchanged.

 figure: Fig. 4.

Fig. 4. (a) Rabi frequencies of the three single-period cosine-like pulses with random amplitude perturbations [given in Eq. (23)], with perturbation coefficients $\eta _{12}=0.3$, $\eta _{23}=0.4$, and $\eta _{13}=0.5$, respectively. (b) Time evolutions of the corresponding populations in $|2\rangle _{Q}$ and $|3\rangle _{Q}$. Here the overall phase is $\phi =\pi /2$ and the initial states are assumed as $|1\rangle _{Q}$.

Download Full Size | PDF

We would also like to compare the robustness of our present scheme with the previous ESST scheme (e.g. Ref. [26]) where three driving fields with well-designed time-dependent amplitudes are applied simultaneously. We show that the ESST scheme in Ref. [26] is vulnerable to random perturbations in their waveforms, as seen in Appendix A. In contrast, our current method is highly robust against certain types of noises that do not change the pulse areas significantly.

In our scheme, the presumed initial state is the ro-vibrational ground state of an asymmetric-top molecule. In principle, the molecules are initially prepared naturally in the ro-vibrational ground state if their temperature is sufficiently low (e.g., below $0.01$ K for 1,2-propanediol molecule) [24,37,59]. Two possible techniques for cooling the molecular sample are buffer-gas cooling and supersonic expansion [31,3740]. In recent experiments, the molecular samples can be cooled down to $1-2\,\text {K}$ by using cryogenic buffer gas cells [31,37,38] and cooled down to $1-2\,\text {K}$ by resorting to the supersonic expansion technique [39,40]. In current experiments, it is challenging to achieve the sufficiently low temperature to prepare these chiral molecules in their ground states directly. An innovative approach with an additional depletion process has been proposed in Ref. [35] to address this challenge, allowing for the preparation of chiral molecules in their rotational ground state without lowering the molecular temperature significantly. The chiral molecules are cooled to low temperatures (around 1-2 K) through a supersonic expansion technique by using Argon as the carrier gas. Subsequently, a continuous-wave ultraviolet (UV) laser is employed to selectively deplete the population of the target rotational state. By following UV excitation, the molecules predominantly emit toward higher vibrational levels and other rotational states that fall outside the selected cyclic three-level system. With the help of this approach, one can optically pump and deplete the target rotational states $J=1$ by the UV laser, thereby avoiding the influence of thermal populations and effectively preparing the chiral molecules in the rotational ground state $|0,0,0\rangle$.

In realistic scenarios, the ESST process of our scheme exists with relaxation and dephasing. In this case, the time evolution of the density matrix operator $\rho _{Q}$ is governed by the master equation

$$\frac{d\rho_{Q}}{dt}={-}i[\hat{H}^{Q}(t),\rho_{Q}]+\mathcal{L}_{rl}\rho_{Q}+\mathcal{L}_{dp}\rho_{Q}.$$

In Eq. (24), the population relaxation is described by the Lindblad superoperator $\mathcal {L}_{rl}$, which satisfies [52,60]

$$\mathcal{L}_{\textrm{rl}}\rho_{Q}= \sum_{3 \geq m>n\geq1}\gamma_{mn} (2\hat{\sigma}_{nm}\rho_{Q}\hat{\sigma}_{mn} -\hat{\sigma}_{mn}\hat{\sigma}_{nm}\rho_{Q}-\rho_{Q}\hat{\sigma}_{mn}\hat{\sigma}_{nm})$$
with $\gamma _{mn}$ the relaxation rates and $\hat {\sigma }_{mn}\equiv |m\rangle \langle n|~(m,n=1,2,3)$. The related pure dephasing is described by [52,60]
$$\mathcal{L}_{\textrm{dp}}\rho_{Q}={-}\tilde{\gamma}\sum_{m\neq n}\hat{\sigma}_{mm}\rho_{Q}\hat{\sigma}_{nn}.$$

For the sake of simplicity, here we assume state-independent pure dephasing rate $\tilde {\gamma }$.

To quantify the efficiency of the ESST in the presence of the relaxation and dephasing effects, we define the enantiomeric excess of the ESST as [24]

$$\varepsilon(t)\equiv\Big|\frac{P^{L}_{2}(t)-P^{R}_{2}(t)}{P^{L}_{2}(t)+P^{R}_{2}(t)}\Big|.$$

According to state-of-the-art experiments, the typical pure dephasing rate is chosen as $\tilde {\gamma }=2\pi \times 0.1\,\textrm {MHz}$ [31,32] and the relaxation rates for the typical rotational transitions are chosen as $\gamma _{12}=\gamma _{23}=2\pi \times 0.1\,\textrm {MHz}$ and $\gamma _{13}=2\pi \times 0.15\,\textrm {MHz}$ [31,32]. We solve numerically the master Eq. (24) and plot the time evolutions of $P_2^L(t)$ and $P_2^R(t)$ in the presence of relaxation and dephasing effects in Fig. 5. Here we use pulses with a single-period cosine waveform given as $\Omega (t)$ in Eq. (11), the initial states are $|1\rangle _{Q}$. With appropriate parameters, one can see from Fig. 5 that the enantiomeric excess grows up to $\varepsilon = 99.15{\% }$ at $t=T=38.5\,\textrm {ns}$. Moreover, it shows that most left-handed chiral molecules remain in the state $|2\rangle _{L}$, while only a tiny fraction of the right-handed chiral molecules remain in the corresponding same-energy excited state $|2\rangle _{R}$. We thus conclude that the impact of the relaxation and dephasing is negligible in our method for typical experimental parameters.

 figure: Fig. 5.

Fig. 5. Time evolutions of the populations $P_2^L(t)$ (red solid line) and $P_2^R(t)$ (blue dashed line) and the corresponding enantiomeric excess $\varepsilon (t)$ (black dotted line). Here the systems are driven by external electromagnetic fields with pulse waveforms in Eq. (11) and also subject to the relaxation and dephasing effects. The initial states are $|1\rangle _{Q}$.

Download Full Size | PDF

7. Conclusion

In summary, we have proposed a robust and highly-efficient dynamical method to realize fast ESST for chiral molecules with cyclic three-level structures. This can be accomplished by using pulses of arbitrary waveforms, as long as the amplitudes of the three Rabi frequencies of the cyclic three-level systems are equal. The equal Rabi frequencies for all three transitions serve the purpose of ensuring the populations occupying the two excited states are always exchanged between the enantiomers during the evolution, regardless of the overall phase and the waveform. Based on this symmetry, we focus on designing the suitable pulse area of microwave fields with the equal Rabi frequencies to facilitate the population transfer evolution to the target excited state for one enantiomer (e.g., left-handed chiral molecules). Correspondingly, employing the same microwave fields allows the right-handed chiral molecules to automatically undergo population transfer to the other excited state simultaneously. Due to the fact that the exchange symmetry is independent of the waveform and overall phase of the pulses, it ensures the flexibility in controlling the waveform and overall phase of the pulses, facilitating the achievement of ESST over a broader parameter range.

We have demonstrated the robustness of our ESST method in the presence of specific types of noise. Additionally, we have shown that the potential impact of relaxation and dephasing on the ESST within our system is negligible under typical experimental parameters. Consequently, despite its simplicity, our current method provides a promising tool for implementing the fast ESST of chiral molecules.

Appendix

A. Discussion about the robustness of the ESST protocol in Ref. [26]

Here we consider the robustness of the ESST protocol in Ref. [26] by introducing the random perturbations with a similar manner as that in this work. In Ref. [26], three driving fields with well-designed time-dependent amplitudes are applied simultaneously to couple to the three transitions $|1\rangle \leftrightarrow |2\rangle$, $|2\rangle \leftrightarrow |3\rangle$, $|1\rangle \leftrightarrow |3\rangle$, respectively, with the corresponding coupling strengths (Rabi frequencies) “$\Omega _{x}$”, “$\pm \Omega _{y}$”, and “$\Omega _{z}e^{i\phi }$”. For the protocol 2 in Ref. [26], at the presence of the random perturbations, the coupling strengths are modified as

$$\begin{aligned} \Omega_{x}(t)&=\dot{\theta}(t)\rightarrow\Omega^{\text{rand}}_{x}(t)=\dot{\theta}(t)[1+\text{rand}_{x}(t,\eta_{x})],\\ \pm\Omega_{y}(t)&={\mp}\dot{\theta}(t)\frac{\sin\theta(t)}{\cos[\theta(t)-\epsilon]}\rightarrow\pm\Omega^{\text{rand}}_{y}(t)={\mp}\dot{\theta}(t)\frac{\sin\theta(t)}{\cos[\theta(t)-\epsilon]}\ [1+\text{rand}_{y}(t,\eta_{y})],\\ \Omega_{z}(t)e^{i\phi}&=\dot{\theta}(t)e^{i\phi}\rightarrow\Omega^{\text{rand}}_{z}(t)e^{i\phi}=\dot{\theta}(t)[1+\text{rand}_{z}(t,\eta_{z})]e^{i\phi},\\ \end{aligned}$$
where $\theta (t)=\frac {\pi }{2}-\frac {3\pi }{2T^{2}}t^{2}+\frac {\pi }{T^{3}}t^{3}$, $\epsilon =0.04$ is a chosen small value, and the overall phase is taken as $\phi =\pi /2$. In Eq. (28), the random perturbations (i.e., certain noises) with $\eta _{x}=0.3$, $\eta _{y}=0.4$, and $\eta _{z}=0.5$ are introduced, similar to that in Eq. (23). And the “+” (“–”) sign before the Rabi frequency $\Omega _{y}(t)$ corresponds to left- (right-) handed chiral molecules, as shown in Ref. [26]. Without the presence of random perturbations, the Rabi frequencies in Eq. (28) had been shown in Ref. [26] to achieve fast ESST with $P^{L}_{2}(T)~[P^{R}_{3}(T)]=0.9946$ for $T=100~\text {ns}$ in the final target state.

The impact of the random perturbations in Eq. (28) has been plotted in Fig. 6. It shows such an ESST scheme in Ref. [26] is vulnerable. For example, the populations in $|2\rangle _{L}$ is $P^{L}_{2}(T)=0.9464$ and the populations in $|3\rangle _{R}$ is $P^{R}_{3}(T)=0.8502$. In contrast, the final populations in the target states ($|2\rangle _{L}$ and $|3\rangle _{R}$) are larger than 0.990 with the similar random amplitude perturbations of the three pulses in our current scheme. Thus our current method exhibits high robustness against random perturbations in the pulse waveforms.

 figure: Fig. 6.

Fig. 6. (a) The designed Rabi frequencies of the three pulses with random amplitude noises [given in Eq. (28)], with perturbation coefficients $\eta _{x}=0.3$, $\eta _{y}=0.4$, and $\eta _{z}=0.5$, respectively. (b) Time evolutions of the corresponding populations in $|2\rangle _{Q}$ and $|3\rangle _{Q}$. Here the overall phase is $\phi =\pi /2$, $\epsilon =0.04$, $T=100~\text {ns}$ and the initial states are assumed to be $|1\rangle _{Q}$.

Download Full Size | PDF

Funding

National Natural Science Foundation of China (12074030, 12088101, 12274107, U223040003); Hainan University (KYQD(ZR)23010).

Disclosures

The authors declare no conflicts of interest.

Data Availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. R. G. Woolley, “Quantum theory and molecular structure,” Adv. Phys. 25(1), 27–52 (1976). [CrossRef]  

2. K. T. Barrett, A. J. Metrano, P. R. Rablen, and S. J. Miller, “Spontaneous transfer of chirality in an atropisomericallyenriched two-axis system,” Nature 509(7498), 71–75 (2014). [CrossRef]  

3. T. J. Leitereg, D. G. Guadagni, J. Harris, T. R. Mon, and R. Teranishi, “Sensory evaluation spectrum method as a descriptive sensory analysis,” J. Agric. Food Chem. 19(4), 785–787 (1971). [CrossRef]  

4. A. R. Ribeiro, P. M. L. Castro, and M. E. Tiritan, “Environmental fate of chiral pharmaceuticals: Determination, degradation and toxicity,” Environ. Chem. Lett. 10(3), 239–253 (2012). [CrossRef]  

5. R. McKendry, M. E. Theoclitou, T. Rayment, and C. Abell, “Chiral discrimination by chemical force microscopy,” Nature 391(6667), 566–568 (1998). [CrossRef]  

6. G. L. J. A. Rikken and E. Raupach, “Chirality, magnetism and light,” Nature 405(6789), 932–935 (2000). [CrossRef]  

7. S. Ahuja, Chiral Separation Methods for Pharmaceutical and Biotechnological Products (New York, 2011).

8. H. Zepik, E. Shavit, M. Tang, T. R. Jensen, K. Kjaer, G. Bolbach, L. Leiserowitz, I. Weissbuch, and M. Lahav, “Chiral amplification of oligopeptides in two-dimensional crystalline self-assemblies on water,” Science 295(5558), 1266–1269 (2002). [CrossRef]  

9. N. Berova and K. Nakanishi, Circular Dichroism: Principles and Applications (New York, 2000).

10. L. A. Nafie, T. A. Keiderling, and P. J. Stephens, “Vibrational circular dichroism,” J. Am. Chem. Soc. 98(10), 2715–2723 (1976). [CrossRef]  

11. R. Kondru, P. Wipf, and D. Beratan, “Atomic contributions to the optical rotation angle as a quantitative probe of molecular chirality,” Science 282(5397), 2247–2250 (1998). [CrossRef]  

12. L. D. Barron, F. Zhu, L. Hecht, G. E. Tranter, and N. W. Isaacs, “Vibrational circular dichroism, raman optical activity and raman spectra of amphetamine species-quantum chemical model calculations and experiments,” J. Mol. Struct. 7, 834 (2007).

13. P. Král and M. Shapiro, “Cyclic population transfer in quantum systems with broken symmetry,” Phys. Rev. Lett. 87(18), 183002 (2001). [CrossRef]  

14. P. Král, I. Thanopulos, M. Shapiro, and D. Cohen, “Two-step enantio-selective optical switch,” Phys. Rev. Lett. 90(3), 033001 (2003). [CrossRef]  

15. Y. Li and C. Bruder, “Dynamic method to distinguish between left- and right-handed chiral molecules,” Phys. Rev. A 77(1), 015403 (2008). [CrossRef]  

16. W. Z. Jia and L. F. Wei, “Distinguishing left- and right-handed molecules using two-step coherent pulses,” J. Phys. B 43(18), 185402 (2010). [CrossRef]  

17. B. T. Torosov, M. Drewsen, and N. V. Vitanov, “Efficient and robust chiral resolution by composite pulses,” Phys. Rev. A 101(6), 063401 (2020). [CrossRef]  

18. B. T. Torosov, M. Drewsen, and N. V. Vitanov, “Chiral resolution by composite raman pulses,” Phys. Rev. Res. 2(4), 043235 (2020). [CrossRef]  

19. N. V. Vitanov and M. Drewsen, “Highly efficient detection and separation of chiral molecules through shortcuts to adiabaticity,” Phys. Rev. Lett. 122(17), 173202 (2019). [CrossRef]  

20. C. Ye, Q. Zhang, Y.-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(4), 043403 (2019). [CrossRef]  

21. 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(1), 014302 (2019). [CrossRef]  

22. 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. Appl. 13(4), 044021 (2020). [CrossRef]  

23. M. Leibscher, E. Pozzoli, C. Pérez, M. Schnell, M. Sigalotti, U. Boscain, and C. P. Koch, “Full quantum control of enantiomer-selective state transfer in chiral molecules despite degeneracy,” Commun. Phys. 5(1), 110 (2022). [CrossRef]  

24. Q. Zhang, Y.-Y. Chen, C. Ye, and Y. Li, “Evading thermal population influence on enantiomeric-specific state transfer based on a cyclic three-level system via ro-vibrational transitions,” J. Phys. B: At., Mol. Opt. Phys. 53(23), 235103 (2020). [CrossRef]  

25. M. Leibscher, J. Kalveram, and C. P. Koch, “Rational pulse design for enantiomer-selective microwave three-wave mixing,” Symmetry 14(5), 871 (2022). [CrossRef]  

26. J.-J. Cheng, Y.-Y. Chen, Y. Li, and L. Zhang, “Enantiomer-specific state transfer of chiral molecules in cyclic three-level systems with SU(2) structures,” Phys. Rev. A 107(1), 013718 (2023). [CrossRef]  

27. J.-J. Cheng, C. Ye, and Y. Li, “Enantio-specific state transfer of chiral molecules through enantio-selective shortcut-to-adiabaticity paths,” J. Chem. Phys. 158(16), 164303 (2023). [CrossRef]  

28. Y. Guo, X. Gong, S. Ma, and C.-C. Shu, “Cyclic three-level-pulse-area theorem for enantioselective state transfer of chiral molecules,” Phys. Rev. A 105(1), 013102 (2022). [CrossRef]  

29. C. Ye, Y. F. Sun, Y. Li, and X. D. Zhang, “Single-shot nondestructive quantum sensing for gaseous samples with hundreds of chiral molecules,” J. Phys. Chem. Lett. 14(30), 6772–6777 (2023). [CrossRef]  

30. J. U. Grabow, “Fourier transform microwave spectroscopy: Handedness caught by rotational coherence,” Angew. Chem., Int. Ed. 52(45), 11698–11700 (2013). [CrossRef]  

31. S. Eibenberger, J. Doyle, and D. Patterson, “Enantiomer-specific state transfer of chiral molecules,” Phys. Rev. Lett. 118(12), 123002 (2017). [CrossRef]  

32. C. Pérez, A. L. Steber, S. R. Domingos, A. Krin, D. Schmitz, and M. Schnell, “Coherent enantiomer-selective population enrichment using tailored microwave fields,” Angew. Chem., Int. Ed. 56(41), 12512–12517 (2017). [CrossRef]  

33. K. K. Lehmann, “Influence of spatial degeneracy on rotationalspectroscopy: Three-wave mixing and enantiomeric state separation of chiral molecules,” J. Chem. Phys. 149(9), 094201 (2018). [CrossRef]  

34. C. Pérez, A. L. Steber, A. Krin, and M. Schnell, “State-specific enrichment of chiral conformers with microwave spectroscopy,” J. Phys. Chem. Lett. 9(16), 4539–4543 (2018). [CrossRef]  

35. J. Lee, J. Bischoff, A. O. Hernandez-Castillo, B. Sartakov, G. Meijer, and S. Eibenberger-Arias, “Quantitative study of enantiomer-specific state transfer,” Phys. Rev. Lett. 128(17), 173001 (2022). [CrossRef]  

36. W. Z. Jia and L. F. Wei, “Probing molecular chirality by coherent optical absorption spectra,” Phys. Rev. A 84(5), 053849 (2011). [CrossRef]  

37. D. Patterson, M. Schnell, and J. M. Doyle, “Enantiomer-specific detection of chiral molecules via microwave spectroscopy,” Nature 497(7450), 475–477 (2013). [CrossRef]  

38. D. Patterson and J. M. Doyle, “Sensitive chiral analysis via microwave three-wave mixing,” Phys. Rev. Lett. 111(2), 023008 (2013). [CrossRef]  

39. 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(4), 1152–1155 (2014). [CrossRef]  

40. 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(1), 196–200 (2015). [CrossRef]  

41. V. A. Shubert, D. Schmitz, C. Pérez, C. Medcraft, A. Krin, S. R. Domingos, D. Patterson, and M. Schnell, “Chiral analysis using broadband rotational spectroscopy,” J. Phys. Chem. Lett. 7(2), 341–350 (2016). [CrossRef]  

42. C. Ye, Q. Zhang, Y.-Y. Chen, and Y. Li, “Determination of enantiomeric excess with chirality-dependent ac stark effects in cyclic three-level models,” Phys. Rev. A 100(3), 033411 (2019). [CrossRef]  

43. C. Ye, Y. S. Sun, and X. D. Zhang, “Entanglement-assisted quantum chiral spectroscopy,” J. Phys. Chem. Lett. 12(35), 8591–8597 (2021). [CrossRef]  

44. Y.-Y. Chen, J.-J. Cheng, C. Ye, and Y. Li, “Enantiodetection of cyclic three-level chiral molecules in a driven cavity,” Phys. Rev. Res. 4(1), 013100 (2022). [CrossRef]  

45. M.-R. Cai, C. Ye, H. Dong, and Y. Li, “Enantiodetection of chiral molecules via two-dimensional spectroscopy,” Phys. Rev. Lett. 129(10), 103201 (2022). [CrossRef]  

46. Y.-Y. Chen, C. Ye, and Y. Li, “Enantio-detection via cavity-assisted three-photon processes,” Opt. Express 29(22), 36132 (2021). [CrossRef]  

47. F. Zou, Y. Y. Chen, B. Liu, and Y. Li, “Enantiodiscrimination of chiral molecules via quantum correlation function,” Opt. Express 30(17), 31073 (2022). [CrossRef]  

48. C. Ye, Y. F. Sun, L. Fu, and X. D. Zhang, “Phase-matched locally chiral light for global control of chiral light–matter interaction,” Opt. Lett. 48(21), 5511 (2023). [CrossRef]  

49. M. Shapiro, E. Frishman, and P. Brumer, “Coherently controlled asymmetric synthesis with achiral light,” Phys. Rev. Lett. 84(8), 1669–1672 (2000). [CrossRef]  

50. P. Brumer, E. Frishman, and M. Shapiro, “Principles of electric-dipole-allowed optical control of molecular chirality,” Phys. Rev. A 65(1), 015401 (2001). [CrossRef]  

51. C. Ye, Q. Zhang, Y.-Y. Chen, and Y. Li, “Fast enantioconversion of chiral mixtures based on a four-level double-Δ model,” Phys. Rev. Res. 2(3), 033064 (2020). [CrossRef]  

52. C. Ye, B. Liu, Y.-Y. Chen, and Y. Li, “Enantio-conversion of chiral mixtures via optical pumping,” Phys. Rev. A 103(2), 022830 (2021). [CrossRef]  

53. B. A. Stickler, M. Diekmann, R. Berger, and D. Wang, “Enantiomer superpositions from matter-wave interference of chiral molecules,” Phys. Rev. X 11, 031056 (2021). [CrossRef]  

54. Y. Li, C. Bruder, and C. P. Sun, “Generalized stern-gerlach effect for chiral molecules,” Phys. Rev. Lett. 99(13), 130403 (2007). [CrossRef]  

55. X. Li and M. Shapiro, “Theory of the optical spatial separation of racemic mixtures of chiral molecules,” J. Chem. Phys. 132(19), 194315 (2010). [CrossRef]  

56. A. Jacob and K. Hornberger, “Effect of molecular rotation on enantioseparation,” J. Chem. Phys. 137(4), 044313 (2012). [CrossRef]  

57. N. Kravets, A. Aleksanyan, and E. Brasselet, “Chiral optical stern-gerlach newtonian experiment,” Phys. Rev. Lett. 122(2), 024301 (2019). [CrossRef]  

58. B. Liu, C. Ye, C. P. Sun, and Y. Li, “Spatial enantioseparation of gaseous chiral molecules,” Phys. Rev. A 104(1), 013113 (2021). [CrossRef]  

59. C. Ye, Q. Zhang, and Y. Li, “Real single-loop cyclic three-level configuration of chiral molecules,” Phys. Rev. A 98(6), 063401 (2018). [CrossRef]  

60. D. Patterson and J. M. Doyle, “Cooling molecules in a cell for ftmw spectroscopy,” Mol. Phys. 110(15-16), 1757–1766 (2012). [CrossRef]  

Data Availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (6)

Fig. 1.
Fig. 1. Cyclic three-level systems of (a) left- and (b) right-handed chiral molecules. Three external electromagnetic fields are coupled resonantly to the electric-dipole transitions, respectively, with overall phase $\phi _{Q}$ and equal Rabi frequency $\Omega$.
Fig. 2.
Fig. 2. (a), (c), and (e) Pulse waveforms given in Eqs. (10), (11), and (12), respectively. (b), (d), and (f) Evolved populations for the left- and right-handed chiral molecules: $P_2^L(t)=P_3^R(t)$ (red solid line) and $P_3^L(t)=P_2^R(t)$ (blue dashed line), corresponding to the waveforms in panels (a), (c), and (e), respectively. Here the overall phase is $\phi =\pi /2$ and the initial states are assumed as $|1\rangle _{L,R}$.
Fig. 3.
Fig. 3. (a) Populations of $|2\rangle _{L}$ or $|3\rangle _{R}$ in the final state versus the overall phase $\phi$ and the pulse area $A$. (b) Populations of $|2\rangle _{L}$ and $|3\rangle _{R}$ in the final state versus the pulse area $A$ with two chosen values of the overall phase $\phi =0.18\pi$ (the black solid line) and $\phi =0.5\pi$ (the green dashed line), corresponding to the black solid and green dashed lines in panel (a), respectively. The initial states are assumed as $|1\rangle _{L,R}$.
Fig. 4.
Fig. 4. (a) Rabi frequencies of the three single-period cosine-like pulses with random amplitude perturbations [given in Eq. (23)], with perturbation coefficients $\eta _{12}=0.3$, $\eta _{23}=0.4$, and $\eta _{13}=0.5$, respectively. (b) Time evolutions of the corresponding populations in $|2\rangle _{Q}$ and $|3\rangle _{Q}$. Here the overall phase is $\phi =\pi /2$ and the initial states are assumed as $|1\rangle _{Q}$.
Fig. 5.
Fig. 5. Time evolutions of the populations $P_2^L(t)$ (red solid line) and $P_2^R(t)$ (blue dashed line) and the corresponding enantiomeric excess $\varepsilon (t)$ (black dotted line). Here the systems are driven by external electromagnetic fields with pulse waveforms in Eq. (11) and also subject to the relaxation and dephasing effects. The initial states are $|1\rangle _{Q}$.
Fig. 6.
Fig. 6. (a) The designed Rabi frequencies of the three pulses with random amplitude noises [given in Eq. (28)], with perturbation coefficients $\eta _{x}=0.3$, $\eta _{y}=0.4$, and $\eta _{z}=0.5$, respectively. (b) Time evolutions of the corresponding populations in $|2\rangle _{Q}$ and $|3\rangle _{Q}$. Here the overall phase is $\phi =\pi /2$, $\epsilon =0.04$, $T=100~\text {ns}$ and the initial states are assumed to be $|1\rangle _{Q}$.

Equations (28)

Equations on this page are rendered with MathJax. Learn more.

H ^ Q ( t ) = Ω 12 ( t ) | 1 Q Q 2 | + Ω 23 ( t ) | 2 Q Q 3 | + Ω 13 ( t ) e i ϕ Q | 1 Q Q 3 | + H.c. ,
ϕ L = ϕ ,   ϕ R = ϕ + π .
Ω 12 ( t ) = Ω 13 ( t ) = Ω 23 ( t ) Ω ( t ) ,
C 1 L ( t ) = 1 3 ( 1 + 2 cos [ 3 A ( t ) ] ) , C 2 L ( t ) = i 3 ( 1 cos [ 3 A ( t ) ] + 3 sin [ 3 A ( t ) ] ) , C 3 L ( t ) = 1 3 ( 1 cos [ 3 A ( t ) ] 3 sin [ 3 A ( t ) ] ) ,
A ( t ) = 0 t Ω ( t ) d t .
C 1 R ( t ) = 1 3 ( 1 + 2 cos [ 3 A ( t ) ] ) , C 2 R ( t ) = i 3 ( 1 cos [ 3 A ( t ) ] 3 sin [ 3 A ( t ) ] ) , C 3 R ( t ) = 1 3 ( 1 cos [ 3 A ( t ) ] + 3 sin [ 3 A ( t ) ] )
C 1 L ( t ) = C 1 R ( t ) , C 2 L ( t ) = i C 3 R ( t ) , C 3 L ( t ) = i C 2 R ( t ) ,
P 1 L ( t ) = P 1 R ( t ) ,     P 2 L ( t ) = P 3 R ( t ) ,     P 3 L ( t ) = P 2 R ( t )
A = 2 3 3 k π + 2 3 π 9   ( k = 0 , ± 1 , )
Ω ( t ) = Ω 0 = 2 π × 10 MHz .
Ω ( t ) = Ω 0 2 ( 1 cos 2 π t T )
Ω ( t ) = Ω 0 e x p [ ( t 3 T ) 2 / T 2 ]
U ^ Q ( t ) = e i 0 t H ^ Q ( t ) d t = exp [ i A ( t ) ( | 1 Q Q 2 | + | 2 Q Q 3 | + e i ϕ Q | 1 Q Q 3 | + H.c. ) ] ,
H ^ ( t ) = Ω ( 0 1 e i ϕ 1 0 1 e i ϕ 1 0 ) .
| ψ ( t ) = e i 0 t H ^ ( t ) d t | ψ ( 0 ) = C 1 ( t ) | 1 + C 2 ( t ) | 2 + C 3 ( t ) | 3 .
H ^ ( t ) = Ω ( 0 1 e i ϕ 1 0 1 e i ϕ 1 0 ) ,
| ψ ( t ) = e i 0 t H ^ ( t ) d t | ψ ( 0 ) = C 1 ( t ) | 1 + C 2 ( t ) | 2 + C 3 ( t ) | 3 .
C 1 ( t ) = i e i ϕ C 1 ( t ) ,   C 2 ( t ) = i C 3 ( t ) , C 3 ( t ) = i C 2 ( t ) .
| ψ ( t ) = e i 0 t [ H ^ ( t ) ] d t | ψ ( 0 ) = C 1 ( t ) | 1 + C 2 ( t ) | 2 + C 3 ( t ) | 3 ,
H ^ ( t ) = Ω ( 0 1 e i ( ϕ + π ) 1 0 1 e i ( ϕ + π ) 1 0 ) .
P 1 ( t ) = P 1 ( t ) ,   P 2 ( t ) = P 3 ( t ) ,   P 3 ( t ) = P 2 ( t ) .
P 1 L ( t ) = P 1 R ( t ) ,   P 2 L ( t ) = P 3 R ( t ) ,   P 3 L ( t ) = P 2 R ( t ) .
Ω m n rand ( t ) = Ω 0 2 ( 1 cos 2 π t T ) [ 1 + rand m n ( t , η m n ) ]
d ρ Q d t = i [ H ^ Q ( t ) , ρ Q ] + L r l ρ Q + L d p ρ Q .
L rl ρ Q = 3 m > n 1 γ m n ( 2 σ ^ n m ρ Q σ ^ m n σ ^ m n σ ^ n m ρ Q ρ Q σ ^ m n σ ^ n m )
L dp ρ Q = γ ~ m n σ ^ m m ρ Q σ ^ n n .
ε ( t ) | P 2 L ( t ) P 2 R ( t ) P 2 L ( t ) + P 2 R ( t ) | .
Ω x ( t ) = θ ˙ ( t ) Ω x rand ( t ) = θ ˙ ( t ) [ 1 + rand x ( t , η x ) ] , ± Ω y ( t ) = θ ˙ ( t ) sin θ ( t ) cos [ θ ( t ) ϵ ] ± Ω y rand ( t ) = θ ˙ ( t ) sin θ ( t ) cos [ θ ( t ) ϵ ]   [ 1 + rand y ( t , η y ) ] , Ω z ( t ) e i ϕ = θ ˙ ( t ) e i ϕ Ω z rand ( t ) e i ϕ = θ ˙ ( t ) [ 1 + rand z ( t , η z ) ] e i ϕ ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.