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

\UseRawInputEncoding
\ensubject

subject

\ArticleType

Article\SpecialTopicSPECIAL TOPIC: \YearYear \MonthMonth \Vol66 \No1 \DOI?? \ArtNo000000 \ReceiveDateMonth Day, Year \AcceptDateMonth Day, Year

A pseudoclassical theory for the wavepacket dynamics of the kicked rotor model

\AuthorMark

Zou Z X

\AuthorCitation

Zou Z X, Wang J

A pseudoclassical theory for the wavepacket dynamics of the kicked rotor model

Zhixing Zou    Jiao Wang Department of Physics and Key Laboratory of Low Dimensional Condensed Matter Physics (Department of Education of Fujian Province),
Xiamen University, Xiamen 361005, Fujian, China
Lanzhou Center for Theoretical Physics, Lanzhou University, Lanzhou 730000, Gansu, China
Abstract

In this study, we propose a generalized pseudoclassical theory for the kicked rotor model in an attempt to discern the footprints of the classical dynamics in the deep quantum regime. Compared with the previous pseudoclassical theory that applies only in the neighborhoods of the lowest two quantum resonances, the proposed theory is applicable in the neighborhoods of all quantum resonances in principle by considering the quantum effect of the free rotation at a quantum resonance. In particular, it is confirmed by simulations that the quantum wavepacket dynamics can be successfully forecasted based on the generalized pseudoclassical dynamics, offering an intriguing example where it is feasible to bridge the dynamics in the deep quantum regime to the classical dynamics. The application of the generalized pseudoclassical theory to the 𝒫𝒯\mathcal{PT}-symmetric kicked rotor is also discussed.

keywords:
Quantum-classical correspondence, Kicked rotor model, Pseudoclassical theory, Wavepacket dynamics
\PACS

05.45.Mt, 03.65.Sq, 03.65.-w

1 Introduction

Quantum-classical correspondence in a general system is the core issue of quantum chaos research or quantum chaology [1]. Based on the quantum-classical correspondence principle, the characteristics of a classically chaotic system are anticipated to manifest themselves in the corresponding quantum system in the semiclassical limit 0\hbar\to 0 (\hbar is the effective Planck constant). Indeed, after 4 decades of intensive research, the general quantum manifestations of classical chaos, such as spectral statistics and the morphologies of wavefunctions, have been well revealed [2]. However, for quantum systems in the deep quantum regime opposite to the semiclassical limit, thus far, the relation\Authorfootnote

between them and their classical counterparts have not been systematically addressed yet. The question arises: Are there any close connections between the classical chaotic dynamics and the quantum properties in the deep quantum regime? Without the guide of any physical law or principle, the answer is not immediate.

An interesting and illuminating example is the pseudoclassical dynamics [3, 4] found in the kicked rotor model [5], a paradigm of quantum chaos with a free rotor subjected to periodic stroboscopic external kicks. Its properties significantly depend on the dimensionless parameter α=T/I\alpha=\hbar T/I, where TT is the kicking period and II is the rotational inertia of the rotor. For a general value of α\alpha, the quantum kicked rotor follows the classical diffusive dynamics first and then the quantum dynamical localization eventually takes over  [5, 6]; however, by contrast, for α\alpha being a rational multiple of 4π4\pi, that is, α=4πr/s\alpha=4\pi r/s with coprime integers rr and ss, a quantum state usually ballistically spreads, which is named as “quantum resonance” [6, 7]. For low-order resonance with a small ss value, the system is in the deep quantum regime; however, surprisingly, Fishman et al. [3, 4] observed that when α\alpha is slightly detuned from the resonance condition by a nonzero δ\delta, the quantum motion can be interpreted according to a certain fictitious classical system that is different from but closely related to the original classical counterpart. The fictitious classical system is named the pseudoclassical system, and the limit δ0\delta\to 0 is the pseudoclassical limit. This result demonstrates a novel and unconventional aspect of quantum-classical correspondence.

Unfortunately, the so-developed pseudoclassical theory only directly applies near the lowest two resonances, that is, s=1s=1 and 22. When the neighborhood of high-order resonance is considered, because a global pseudoclassical phase-space approximation regarding a unique classical Hamiltonian is impossible, one has to resort to local pseudoclassical approximations of different Hamiltonians related to the quasienergy bands of the considered quantum resonance [8]. Nevertheless, the interaction between the quasienergy bands may ruin the local pseudoclassical approximations to make their predictions invalid [9]. Recently [10], we realized that owing to the periodicity of the phase space, in a system with a spherical or cylindrical phase space, the free rotation of a wavepacket at quantum resonance may lead to the simultaneous presence of multiple wavepackets, in clear contrast to the usual scenario where only one wavepacket is found throughout. Such a wavepacket-multiplying effect is a pure quantum interference effect and has not been considered earlier. Considering this, the pseudoclassical theory may find wide applications when being generalized by considering this effect. This conjecture has been confirmed with the kicked top model [10], another paradigm of quantum chaos with a spherical phase space, which is closely connected with the kicked rotor [11, 12].

The objective of this work is to establish a generalized pseudoclassical theory for the kicked rotor model with a cylindrical phase space with regard to its significant role in quantum chaos research. This attempt is successful again. Based on our theory, for the lowest two resonances of s=1s=1 and 22, the free rotation does not lead to the wavepacket multiplying, and this is why the previous pseudoclassical theory is valid in these two cases. However, for higher-order resonance, particularly that of an odd ss, the wavepacket multiplying can be shown to occur for certain, and the generalized theory is thus imperative. Our study might be a preliminary but positive attempt to address the quantum-classical correspondence issue in the deep quantum regime.

This article is organized as follows: Sec. II briefly describes the kicked rotor model and discusses in detail the pseudoclassical limit of the wavepacket dynamics. Sec. III compares the quantum evolution of a wavepacket with the prediction of the pseudoclassical theory via numerical simulations, convincingly demonstrating that the latter is effective. Sec. IV presents the extension to the 𝒫𝒯\mathcal{PT}-symmetric kicked rotor and confirmation by numerical simulations. Finally, Sec. V concludes this work.

2 The pseudoclassical limit of the kicked rotor

The Hamiltonian of the kicked rotor is

H=P22I+KIωTcos(ωθ)m=δ(τmT).\displaystyle H=\frac{P^{2}}{2I}+\frac{KI}{\omega T}\cos(\omega\theta)\sum_{m=-\infty}^{\infty}\delta(\tau-mT). (1)

Here, II, PP, and θ\theta are the rotational inertia, the angular momentum, and the conjugate angular coordinate of the rotor, respectively, whereas TT and KK are the kicking period and the dimensionless kicking strength, respectively. The integer parameter ω\omega is introduced for our aim here, which is unity in the conventional kicked rotor model. The angular coordinate θ\theta is imposed with a period of 2π2\pi, such that the phase space is a cylinder.

For the quantum kicked rotor, it is convenient to adopt the basis of the angular momentum eigenstates, {|n;<n<}\{|n\rangle;-\infty<n<\infty\}, where nn is an integer and |n|n\rangle meets P|n=n|nP|n\rangle=n\hbar|n\rangle. Because the Hamiltonian is periodic with time TT, the evolution of the rotor for time TT can be fulfilled by applying the Floquet operator

U=exp(iα2ν^2)exp(iKαωcos(ωθ))\displaystyle U=\exp\left(-i\frac{\alpha}{2}\hat{\nu}^{2}\right)\exp(-i\frac{K}{\alpha\omega}\cos(\omega\theta)) (2)

to its current state just before a kick, where αT/I\alpha\equiv\hbar T/I and ν^P/\hat{\nu}\equiv P/\hbar. As stated in the Introduction, the quantum dynamics depends qualitatively on the fact if α\alpha is a rational multiple of 4π4\pi. The case α=4πr/s\alpha=4\pi r/s with coprime integers rr and ss corresponds to quantum resonance, and the quantum state usually ballistically spreads. An exception is for s=2s=2 and ω\omega being odd so that U2=1U^{2}=1, suggesting that the quantum state does not change after time 2T2T, which is termed as “quantum antiresonance” [7, 13].

When the system is slightly detuned away from a quantum resonance, a pseudoclassical theory has been developed to address the quantum dynamics via a classical map, the so-called pseudoclassical limit [3, 4]. However, this theory only works at the lowest two resonances. In the following of this section, we attempt to extend this theory to the neighborhoods of higher-order resonances with α=4πr/s+δ\alpha=4\pi r/s+\delta, where δ\delta (incommensurate to π\pi) is a weak perturbation to the resonance condition. To perform a close comparison between the quantum and the classical dynamics, which is critical to our objective, we invoke the (squeezed) coherent state in the former, with regard to the advantages that the Husimi distribution of a coherent state has the minimum uncertainty in the phase space and its center point in the phase space represents exactly the classical counterpart of the coherent state. For convenience explained later, we use (p,θ)(p,\theta) to denote a classical state, or a point in the phase space, with pP(T/I)(δ/α)p\equiv P(T/I)(\delta/\alpha). The expression of the coherent state centered at (p,θ)(p,\theta) is

|p,θ=cnexp(δ2(npδ)2)exp(inθ)|n,\displaystyle|p,\theta\rangle=c\sum_{n}\exp\left(-\frac{\delta}{2}(n-{\frac{p}{\delta}})^{2}\right)\exp\left(-in\theta\right)|n\rangle, (3)

where cc is the normalization factor. When δ\delta is small, the uncertainty of pp and θ\theta is the same, δpδθδ/2\delta_{p}\approx\delta_{\theta}\approx\sqrt{\delta/2}. In the phase space, the Husimi distribution of the coherent state is a Gaussian function centered at (p,θ)(p,\theta). In the limit of δ0\delta\to 0, it collapses to the point (p,θ)(p,\theta).

Our task is to determine the one-step evolution for the classical state (p,θ)(p,\theta) by analogy according to the corresponding quantum evolution for the coherent state |p,θ|p,\theta\rangle. To this end, note that for α=4πr/s+δ\alpha=4\pi r/s+\delta, the Floquet operator can be rewritten as

U=exp(i2πrsν^2)exp(iδ2ν^2)exp(ikδωcosωθ),\displaystyle U=\exp\left(-i2\pi\frac{r}{s}\hat{\nu}^{2}\right)\exp\left(-i\frac{\delta}{2}\hat{\nu}^{2}\right)\exp(-i\frac{k}{\delta\omega}\cos\omega\theta), (4)

where k=Kδ/αk=K\delta/\alpha. It includes two parts, i.e., U=UfUδU=U_{f}U_{\delta}, with Uf=exp(i2πrsν^2)U_{f}=\exp(-i2\pi\frac{r}{s}\hat{\nu}^{2}) and Uδ=exp(iδ2ν^2)exp(ikδωcos(ωθ))U_{\delta}=\exp\left(-i\frac{\delta}{2}\hat{\nu}^{2}\right)\exp(-i\frac{k}{\delta\omega}\cos(\omega\theta)). The former, UfU_{f}, represents a pure free rotation, while the latter, UδU_{\delta}, represents kicked rotor dynamics (see Eq. (2)) with αδ\alpha\to\delta and KkK\to k. We thus break our task down into two steps. First, note that in the limit δ0\delta\to 0, the quantum operation UδU_{\delta} has a well-defined semiclassical limit. This can be observed more clearly by imagining δ\delta as a virtual Planck constant so that we can write down the classical Hamiltonian corresponding to UδU_{\delta} as

Hδ=p22+kωcos(ωθ)n=δ(τn).\displaystyle H_{\delta}=\frac{p^{2}}{2}+\frac{k}{\omega}\cos(\omega\theta)\sum_{n=-\infty}^{\infty}\delta(\tau-n). (5)

Based on HδH_{\delta}, the classical Poincaré map corresponding to UδU_{\delta}, denoted as δ\mathcal{M}_{\delta}, can be derived straightforwardly. To be concrete, in terms of pp and θ\theta, the map δ\mathcal{M}_{\delta} that evolves the state (p,θ)(p,\theta) to (p~,θ~)(\tilde{p},\tilde{\theta}) reads

δ:{p~=p+ksin(ωθ),θ~=θ+p~.\displaystyle\mathcal{M}_{\delta}:\begin{cases}\tilde{p}=p+k\sin(\omega\theta),\\ \tilde{\theta}=\theta+\tilde{p}.\end{cases} (6)

Note that the phase space portrait created by δ\mathcal{M}_{\delta} is periodic in both pp and θ\theta of period 2π/ω2\pi/\omega.

Due to this close relation between the quantum operation UδU_{\delta} and the classical map δ\mathcal{M}_{\delta} in the limit of δ0\delta\to 0, we assume that

|p~,θ~=Uδ|p,θ\displaystyle|\tilde{p},\tilde{\theta}\rangle=U_{\delta}|p,\theta\rangle (7)

as well when δ\delta is small, which is the only significant approximation we adopt for our theory. Equation (7) is the counterpart of the classical map δ:(p,θ)(p~,θ~)\mathcal{M}_{\delta}:(p,\theta)\to(\tilde{p},\tilde{\theta}) (see Eq. (6)), and thus, we finish the first step of our task.

Afterward, we need to derive Uf|p~,θ~U_{f}|\tilde{p},\tilde{\theta}\rangle and work out its classical counterpart. The detailed calculation of Uf|p~,θ~U_{f}|\tilde{p},\tilde{\theta}\rangle is given in Appendix A, resulting in

Uf|p~,θ~=l=0s1Gl|p~,θ~+2πrsl,\displaystyle U_{f}\left|\tilde{p},\tilde{\theta}\right\rangle=\sum_{l=0}^{s-1}G_{l}|\tilde{p},\tilde{\theta}+2\pi\frac{r}{s}l\rangle, (8)

where GlG_{l} is the Gaussian sum [14]

Gl=1sm=0s1exp(i2πrsm(ml)).\displaystyle G_{l}=\frac{1}{s}\sum_{m=0}^{s-1}\exp\left(-i2\pi\frac{r}{s}m(m-l)\right). (9)

The physical meaning of Eq. (8) is evident: The intermediate coherent state |p~,θ~|\tilde{p},\tilde{\theta}\rangle is mapped by UfU_{f} into ss coherent states whose centers are located along the line of p=p~p=\tilde{p} in the phase space. They are separated in θ\theta by 2πrs2\pi\frac{r}{s} (and its multiples) from each other, and each coherent state has an associated complex amplitude given by a Gaussian sum. This is a peculiar characteristic of the quantum rotor at resonance. Note that not all of these ss coherent states exist necessarily. A coherent state disappears if the associated amplitude GlG_{l} disappears. Suppose that there are 𝒩\mathcal{N} nonzero amplitudes, Eq. (8) can be rewritten as

Uf|p~,θ~=j=1𝒩Aj|p~,θ~+Δj.\displaystyle U_{f}|\tilde{p},\tilde{\theta}\rangle=\sum_{j=1}^{\mathcal{N}}A_{j}\left|\tilde{p},\tilde{\theta}+\Delta_{j}\right\rangle. (10)

Here, for the jjth component coherent state, its amplitude AjA_{j} is a nonzero Gaussian sum GlG_{l}, and its position bias Δj\Delta_{j} is associated with the subscript of GlG_{l} by Δj=2πlr/s\Delta_{j}=2\pi lr/s mod 2π2\pi. Similarly, as in the limit δ0\delta\to 0, a coherent state reduces to a point in phase space, Eq. (10) can be interpreted pseudoclassically. That is, the intermediate classical state (p~,θ~)(\tilde{p},\tilde{\theta}) is mapped by the pseudoclassical counterpart of UfU_{f}, denoted as f\mathcal{M}_{f}, into a set of 𝒩\mathcal{N} states, and each of these is associated with a complex amplitude,

f:(p~,θ~){[(p~,θ~+Δj);Aj],j=1,,𝒩}.\displaystyle\mathcal{M}_{f}:(\tilde{p},\tilde{\theta})\to\{[(\tilde{p},\tilde{\theta}+\Delta_{j});A_{j}],j=1,\cdots,\mathcal{N}\}. (11)

This concludes the second step of our task.

Therefore, formally, the pseudoclassical map corresponding to the quantum evolution U|p,θU|p,\theta\rangle that we are seeking for, denoted as \mathcal{M}, can be expressed as =fδ\mathcal{M}=\mathcal{M}_{f}\mathcal{M}_{\delta}, that is,

:(p,θ){[(p~,θ~+Δj);Aj],j=1,,𝒩}.\displaystyle\mathcal{M}:(p,\theta)\to\{[(\tilde{p},\tilde{\theta}+\Delta_{j});A_{j}],j=1,\cdots,\mathcal{N}\}. (12)

The intermediate state (p~,θ~)(\tilde{p},\tilde{\theta}) is related to (p,θ)(p,\theta) by the map δ\mathcal{M_{\delta}} (see Eq. (6)). This is the core result that we have obtained. As shown in the next section, it does allow the prediction of the quantum dynamics in such a pseudoclassical way. Here, we emphasize that the amplitudes {Aj}\{A_{j}\} are crucial to this end. Specifically, |Aj|2|A_{j}|^{2} has to be taken as the weight of the jj\emph{}th state (p~,θ~+Δj)(\tilde{p},\tilde{\theta}+\Delta_{j}) that it is associated with to assess the expected value of a given observable. In addition, the phases encoded in these amplitudes have to be simultaneously considered to correctly trace the quantum evolution.

Some remarks are in order. First, for the lowest two resonances s=1s=1 and s=2s=2, 𝒩=1\mathcal{N}=1 and the amplitude of the only resultant state is unity. In addition, the map \mathcal{M} reduces to that given by the original pseudoclassical theory [3, 4], which has a seemingly pure classical form. For higher-order resonance, even though the map \mathcal{M} is substantially more complex, there are no quantum operations and parameters that are explicitly involved. Despite this, the map \mathcal{M} should be understood as a mix of the quantum and the classical dynamics according to our reasoning for deriving \mathcal{M}. It might be appropriate to regard the seemingly simple form for the lowest two resonances as a coincidence.

Second, for an odd ss, all ss Gaussian sums are nonzero [14] so that 𝒩=s\mathcal{N}=s; for an even ss, it can be shown that half of them must be zero so that 𝒩=s/2\mathcal{N}=s/2 (see Appendix B). In either case, the main challenge for the implementation of the pseudoclassical map \mathcal{M} lies in the rapid proliferation of the involved states when 𝒩2\mathcal{N}\geq 2. In general, their number exponentially increases, 𝒩t\sim\mathcal{N}^{t}, as the number of iterations tt (or the evolving time τ=tT\tau=tT). This suggests why quantum dynamics is more complicated from such a novel perspective. Thus, for a general case, it can be hoped that in practice, our pseudoclassical theory only works in a short time to predict quantum evolution, which can be attributed to the intrinsic complexity of quantum dynamics.

Third, however, for the following two cases equipped with an ss- or s/2s/2-fold translational symmetry in the θ\theta direction of the phase space, that is,
C1) ss is odd and ω=s\omega=s
C2) ss is even and ω=s/2\omega=s/2
the proliferation problem can be prevented because from the second iteration on (t2t\geq 2), two or more states can be mapped into one so that the total number of resultant states remains bounded by ss in case C1 and by s/2s/2 in case C2. Appendix B provides a detailed discussion.

In fact, note that for a general case, UfU_{f} and UδU_{\delta} do not commute. However, for case C1, they commute (see Appendix C), demonstrating that the operator for the tt steps of iteration, (UfUδ)t(U_{f}U_{\delta})^{t}, can be split into (UfUδ)t=UftUδt(U_{f}U_{\delta})^{t}=U_{f}^{t}U_{\delta}^{t}. As the state proliferation is exclusively caused by Uft=exp(i2πtrsν2)U_{f}^{t}=\exp(-i2\pi t\frac{r}{s}\nu^{2}), we can conclude that the number of final resultant coherent states cannot surpass ss for any time tt. Moreover, the corresponding pseudoclassical map can be expressed as t=ftδt\mathcal{M}^{t}=\mathcal{M}_{f}^{t}\mathcal{M}_{\delta}^{t}, based on which the computing of the pseudoclassical evolution can be substantially simplified.

In case C2, UfU_{f} and UδU_{\delta} do not generally commute. However, UfU^{\prime}_{f} and UδU^{\prime}_{\delta} commute, where Uf=Ufexp(iπrν)U^{\prime}_{f}=U_{f}\exp(i\pi r\nu) and Uδ=exp(iπrν)UδU^{\prime}_{\delta}=\exp(-i\pi r\nu)U_{\delta} (see Appendix C). Consequently, (UfUδ)t=UftUδt(U_{f}U_{\delta})^{t}={U^{\prime}_{f}}^{t}{U^{\prime}_{\delta}}^{t}, which has similar implications to case C1.

3 Verification of the pseudoclassical theory

The effectiveness of the pseudoclassical theory is checked by the comparison of its predictions on the wavepacket dynamics and that directly obtained with the quantum Floquet operator. Three representative examples are examined, of which one is for a general case with r=1r=1, s=4s=4, and ω=1\omega=1, where the number of coherent states doubles after every step of iteration. The other two instances, one with r=1r=1, s=3s=3, and ω=3\omega=3 and another with r=1r=1, s=4s=4, and ω=2\omega=2, are for cases C1 and C2, respectively, where the proliferation of the coherent states is suppressed. As shown in the following of this section, the pseudoclassical theory effectively works in all three cases.

Refer to caption
Figure 1: Husimi distribution for the quantum state at time t=0t=0 (a), t=1t=1 (b), t=2t=2 (c), and t=3t=3 (d) for a general case of r=1r=1, s=4s=4, and ω=1\omega=1. Here, δ=0.04\delta=0.04, k=0.5k=0.5, and the initial state is |p0,θ0=|0.5,0.5|p_{0},\theta_{0}\rangle=|0.5,0.5\rangle. The black pluses denote the resultant states by the pseudoclassical map [Eq. (14)].

To conduct a close comparison between the quantum wavepacket evolution and its pseudoclassical counterpart, it is appropriate to visualize the quantum evolution in the phase space with the Husimi distribution [15]. For a given quantum state |ψ|\psi\rangle, at the given point (p,θ)(p,\theta) in the phase space, the Husimi distribution (p,θ)\mathcal{H}(p,\theta) is defined as the expectation value of the density matrix ρ=|ψψ|\rho=|\psi\rangle\langle\psi| with respect to the corresponding coherent state |p,θ|p,\theta\rangle. That is,

(p,θ)=p,θ|ρ|p,θ.\displaystyle\mathcal{H}(p,\theta)=\langle p,\theta|\rho|p,\theta\rangle. (13)

First, for the first representative example of a general case where r=1r=1, s=4s=4, and ω=1\omega=1, based on Eq. (9), we have two nonzero GlG_{l} and as such 𝒩=2\mathcal{N}=2. The pseudoclassical map is

:(p,θ){(p~,θ~);A1,(p~,θ~+π);A2,\displaystyle\mathcal{M}:(p,\theta)\rightarrow\begin{cases}(\tilde{p},\tilde{\theta});~{}~{}A_{1},\\ (\tilde{p},\tilde{\theta}+\pi);~{}~{}A_{2},\\ \end{cases} (14)

with the complex amplitudes A1=1+i2A_{1}=\frac{1+i}{2} and A2=1i2A_{2}=\frac{1-i}{2}. In addition, based on Eq. (6), p~=p+ksinθ\tilde{p}=p+k\sin\theta and θ~=θ+p~\tilde{\theta}=\theta+\tilde{p} in this case.

Refer to caption
Figure 2: The expected value of (pp0)2(p-p_{0})^{2} as a function of time for the pseudoclassical dynamics (black pluses) and the quantum dynamics with δ=103\delta=10^{-3} (red squares), 10210^{-2} (blue circles), and 10110^{-1} (orange triangles), respectively, for the general case of r=1r=1, s=4s=4, and ω=1\omega=1. Here, k=0.5k=0.5, and for the initial state, p0=θ0=0.5p_{0}=\theta_{0}=0.5.

On the quantum aspect, for a given initial coherent state |ψ0=|p0,θ0|\psi_{0}\rangle=|p_{0},\theta_{0}\rangle, the state after tt iterations, |ψt=Ut|ψ0|\psi_{t}\rangle=U^{t}|\psi_{0}\rangle, can be numerically determined by repeatedly applying the evolution operator UU. The contour plot of the corresponding Husimi distribution of |ψt|\psi_{t}\rangle for t=0t=0 to 3 is presented in Fig. 1, where the wavepacket proliferation owing to quantum resonance can be clearly observed. On the pseudoclassical aspect, using the corresponding initial state (p0,θ0)(p_{0},\theta_{0}), we can acquire 𝒩t\mathcal{N}^{t} states after tt iterations by the pseudoclassical map. The positions of these states are indicated in Fig. 1 as well. By comparison, we can observe that they can indeed well-capture the skeleton of the quantum state.

In addition to the skeleton, more information on the quantum state is encoded in the complex amplitudes of the pseudoclassical states. To determine if this information is sufficient to forecast the expected value of a given observable, the angular momentum diffusion behavior is considered as an example, quantified by (pp0)2\langle(p-p_{0})^{2}\rangle, which is of particular interest in the quantum kicked rotor research. Fig. 2 compares the results by the quantum and the pseudoclassical dynamics. We may expect that in the limit of δ0\delta\to 0, the agreement between them should be progressively enhanced, which is well corroborated.

Note that despite its success, as shown in Fig. 1 and Fig. 2, owing to the issue of state proliferation, the implementation of the pseudoclassical theory would be prohibitively costly for a long evolution time. For instance, in the case discussed, the involved states would have reached up to 10910^{9} for t=30t=30.

However, for a system that has the translational symmetry in θ\theta as in cases C1 and C2, the state proliferation challenge can be overcome. As two examples of case C1 and C2, we consider r=1r=1, s=3s=3, and ω=3\omega=3 and r=1r=1, s=4s=4, and ω=2\omega=2, respectively.

Refer to caption
Figure 3: The same as Fig. 1 but for case C1 with r=1r=1, s=3s=3, and ω=3\omega=3.

For the former, 𝒩=3\mathcal{N}=3 and the pseudoclassical map reads as

:(p,θ){(p~,θ~);A1,(p~,θ~+2π3);A2,(p~,θ~+4π3);A3,\displaystyle\mathcal{M}:(p,\theta)\rightarrow\begin{cases}(\tilde{p},\tilde{\theta});~{}~{}A_{1},\\ (\tilde{p},\tilde{\theta}+\frac{2\pi}{3});~{}~{}A_{2},\\ (\tilde{p},\tilde{\theta}+\frac{4\pi}{3});~{}~{}A_{3},\\ \end{cases} (15)

where A1=3i3A_{1}=-\frac{\sqrt{3}i}{3}, A2=A3=3+3i6A_{2}=A_{3}=\frac{3+\sqrt{3}i}{6}, and (p~,θ~)(\tilde{p},\tilde{\theta}) is generated from (p,θ)(p,\theta) according to Eq. (6). This suggests that, after each step, a point will be mapped into three of the same weight but of two different phases.

Refer to caption
Figure 4: The same as Fig. 1 but for case C2 with r=1r=1, s=4s=4, and ω=2\omega=2.

The results of the Husimi distribution of the quantum state at t=0t=0 to 3 and the pseudoclassical map are shown in Fig. 3. It can be observed that at t=1t=1, the initial quantum coherent state is mapped into three, and their centers perfectly overlap with the three resulting pseudoclassical states. At t=2t=2, of the nine expected pseudoclassical states, three pairs cancel each other so that only three survive. This is completely supported by quantum evolution. Finally, at t=3t=3, the three remaining pseudoclassical states merge into one rather than split into nine, successfully further predicting quantum evolution. For the quantum evolution, it can be straightforwardly shown that Uf3=1U_{f}^{3}=1, suggesting that the number of involved coherent states must reduce to one after every three steps.

For the latter case of r=1r=1, s=4s=4, and ω=2\omega=2, the pseudoclassical map is the same as Eq. (14) but with p~=p+ksin(2θ)\tilde{p}=p+k\sin(2\theta) and θ~=θ+p~\tilde{\theta}=\theta+\tilde{p} instead. As A1,2=1±i2A_{1,2}=\frac{1\pm i}{2}, a point will be mapped into two with the same weight but different phases. Fig. 4 compares the quantum and pseudoclassical dynamics, and good agreement between them is the same as in the previous case. Note that for this case, although Uf4=1U_{f}^{4}=1, Uf21U_{f}^{2}\neq 1. However, Uf2=1{U^{\prime}}_{f}^{2}=1 such that (UfUδ)2=Uδ2(U_{f}U_{\delta})^{2}={U^{\prime}}_{\delta}^{2}, which explains why after every two steps, the number of pseudoclassical states becomes one.

In these two cases, as the state proliferation problem is well suppressed owing to translational symmetry, the pseudoclassical evolution can be conveniently conducted up to a much longer time than in the general case. For these two cases, Figs. 5(a) and 5(b) compare the time dependence of (pp0)2\langle(p-p_{0})^{2}\rangle calculated with the quantum and the pseudoclassical dynamics over a wide time range. In particular, kk is fixed, but perturbation δ\delta is changed to observe how the quantum results depend on it. Indeed, as expected, as δ\delta decreases, the quantum result tends to approach the pseudoclassical result. In addition, for the diffusion time tdifft_{\text{diff}}, which is empirically defined as the time when the quantum result deviates from the pseudoclassical result by 15%15\% from below, it follows the scaling tdiffδ2t_{\text{diff}}\sim\delta^{-2} [see Figs.  5(c) and 5(d)], the same as in the conventional semiclassical limit of the kicked rotor model if δ\delta is recognized with the effective Planck constant [16]. All these results consistently support the effectiveness of our pseudoclassical theory.

Refer to caption
Figure 5: (a) Time dependence of (pp0)2\langle(p-p_{0})^{2}\rangle for r=1r=1, s=3s=3, and ω=3\omega=3. Red, green, blue, and cyan dashed curves are for the quantum results with δ=102\delta=10^{-2}, 102.410^{-2.4}, 102.810^{-2.8}, and 103.210^{-3.2}, respectively, for the initial state |ψ(0)=|0|\psi(0)\rangle=|0\rangle. The black curve is for the ensemble average of the pseudoclassical results for 10610^{6} initial states uniformly distributed on the line of p=0p=0 as the classical counterpart of |0|0\rangle. (b) Same as (a) but for r=1r=1, s=4s=4, and ω=2\omega=2. Panels (c) and (d) are for the δ\delta dependence of the diffusion time tdifft_{\text{diff}} for the quantum system presented in (a) and (b), respectively. Here, in all the simulations, k=2k=2.

4 Application to the 𝒫𝒯\mathcal{PT}-symmetric kicked rotor

Quantum mechanical Hamiltonians that are 𝒫𝒯\mathcal{PT}-symmetric but not Hermitian has recently been a frontier subject  [17, 20, 21, 22, 18, 23, 19]. A Hamiltonian HH is regarded as 𝒫𝒯\mathcal{PT}-symmetric if [H,𝒫𝒯]=0[H,\mathcal{PT}]=0, where the parity operator, 𝒫\mathcal{P}, is a unitary operator that satisfies 𝒫2=1\mathcal{P}^{2}=1 and the time-reversal operator, 𝒯\mathcal{T}, is an antiunitary operator that satisfies 𝒯2=±1\mathcal{T}^{2}=\pm 1. As a result, 𝒫𝒯\mathcal{PT} is an antiunitary operator as well. Surprisingly, as observed in some previous works and highlighted in Ref. [17], it is feasible for a 𝒫𝒯\mathcal{PT}-symmetric Hamiltonian to have a real spectrum, despite the fact that it can be non-Hermitian. Moreover, as the gain (or loss) parameter λ\lambda that controls the degree of non-Hermiticity changes, a spontaneous 𝒫𝒯\mathcal{PT} symmetry breaking may take place.

Refer to caption
Figure 6: The same as Fig. 1 but for the 𝒫𝒯\mathcal{PT}-symmetric kicked model with λ=0.2\lambda=0.2. (The results shown in Fig. 1 are equivalent to the case of λ=0\lambda=0.) The black pluses stand for the positions of the resultant states generated by the pseudoclassical map NH\mathcal{M}^{\mathrm{NH}} [see Eq. (19)].

In this section, we attempt to use our pseudoclassical theory to the 𝒫𝒯\mathcal{PT}-symmetric kicked rotor model whose Hamiltonian is [18]

HNH=P22I+KIωT(cosωθ+iλsinωθ)m=δ(τmT),\displaystyle H^{\mathrm{NH}}=\frac{P^{2}}{2I}+\frac{KI}{\omega T}(\cos\omega\theta+i\lambda\sin\omega\theta)\sum_{m=-\infty}^{\infty}\delta(\tau-mT), (16)

where λ0\lambda\geq 0 is the non-Hermitian parameter that controls the strength of the imaginary part of the kicking potential. For λ=0\lambda=0, it reduces to the conventional Hermitian kicked rotor. When αT/I=4πrs\alpha\equiv\hbar T/I=4\pi\frac{r}{s} with rr and ss two coprime integers, similar to the conventional kicked rotor, the quasienergy spectrum is absolutely continuous and composed of ss quasienergy bands, which is also named quantum resonance [18]. In the following, we will focus on the perturbed case where α=4πrs+δ\alpha=\frac{4\pi r}{s}+\delta, for which the Floquet operator can be written as

UNH\displaystyle U^{\mathrm{NH}} =\displaystyle= exp(i2πrsν^2)exp(iδ2ν^2)\displaystyle\exp\left(-i2\pi\frac{r}{s}\hat{\nu}^{2}\right)\exp\left(-i\frac{\delta}{2}\hat{\nu}^{2}\right) (17)
×exp(ikδcos(ωθ))exp(kλδsin(ωθ)).\displaystyle\times\exp(-i\frac{k}{\delta}\cos(\omega\theta))\exp(\frac{k\lambda}{\delta}\sin(\omega\theta)).

It includes two parts: UNH=UfUδNHU^{\mathrm{NH}}=U_{f}U_{\delta}^{\mathrm{NH}} with Uf=exp(i2πrsν^2)U_{f}=\exp\left(-i2\pi\frac{r}{s}\hat{\nu}^{2}\right) and UδNH=exp(iδ2ν^2)exp(ikδcos(ωθ))exp(kλδsin(ωθ))U_{\delta}^{\mathrm{NH}}=\exp\left(-i\frac{\delta}{2}\hat{\nu}^{2}\right)\exp(-i\frac{k}{\delta}\cos(\omega\theta))\\ \exp(\frac{k\lambda}{\delta}\sin(\omega\theta)). The difference lies in the last term of UδNHU_{\delta}^{\mathrm{NH}}. For λ=0\lambda=0, this term is identity, UδNHU_{\delta}^{\mathrm{NH}} reduces to UδU_{\delta} and UNHU^{\mathrm{NH}} reduces to UU.

For λ0\lambda\neq 0 and in the limit of δ0\delta\to 0, when the last term of UδNHU_{\delta}^{\mathrm{NH}} acts on a quantum state expressed as the superposition of multiple coherent states, only the component coherent state |pj,θj|p_{j},\theta_{j}\rangle that maximizes sin(ωθ)\sin(\omega\theta) (with θj\theta_{j}) significantly contributes to the result; the contributions of other components can be ignored. In such a sense, one role that the last term of UδNHU_{\delta}^{\mathrm{NH}} plays is a “selector”. Formally, the pseudoclassical counterpart of this role can be denoted as s\mathcal{M}_{s}. It chooses the component state (pj,θj)(p_{j},\theta_{j}) from others as the first step of the pseudoclassical dynamics.

Refer to caption
Figure 7: Expected value of the momentum as a function of time for the pseudoclassical dynamics (black pluses) and the quantum dynamics with δ=103\delta=10^{-3} (red squares), 10210^{-2} (blue circles), and 10110^{-1} (orange triangles), respectively, of the 𝒫𝒯\mathcal{PT}-symmetric kicked model with λ=0.01\lambda=0.01. Here, r=1r=1, s=4s=4, ω=1\omega=1, k=0.5k=0.5, and for the initial state, p0=θ0=0p_{0}=\theta_{0}=0.

The last term of UδNHU_{\delta}^{\mathrm{NH}} also plays a key role in retrieving the pseudoclassical counterpart, denoted as δNH\mathcal{M}_{\delta}^{\mathrm{NH}}, of UδNHU_{\delta}^{\mathrm{NH}}. Based on the generalized canonical structure theory [20], in the limit of δ0\delta\to 0, the motion of the center of a given coherent state UδNHU_{\delta}^{\mathrm{NH}} acts on is governed by the following equations:

p˙\displaystyle\dot{p} =\displaystyle= ksinωθm=δ(τm),\displaystyle k\sin\omega\theta\sum_{m=-\infty}^{\infty}\delta(\tau-m),
θ˙\displaystyle\dot{\theta} =\displaystyle= p+kλcosωθm=δ(τm).\displaystyle p+k\lambda\cos\omega\theta\sum_{m=-\infty}^{\infty}\delta(\tau-m). (18)

Thus, the pseudoclassical operation δNH\mathcal{M}_{\delta}^{\mathrm{NH}} thus represents the integration of these two functions up to a unit time. For λ=0\lambda=0, it reduces to δ\mathcal{M}_{\delta} given by Eq. (6).

Formally, the pseudoclassical map for the 𝒫𝒯\mathcal{PT}-symmetric kicked rotor can be written as

NH=fδNHs.\displaystyle\mathcal{M}^{\mathrm{NH}}=\mathcal{M}_{f}\mathcal{M}^{\mathrm{NH}}_{\delta}\mathcal{M}_{s}. (19)

To test its effectiveness, a general case of the 𝒫𝒯\mathcal{PT}-symmetric kicked rotor is simulated with r=1r=1, s=4s=4, and ω=1\omega=1, the same as in Figs. 1 and 2. Figs. 6 and 7 illustrate the results. Fig. 6 displays the Husimi distribution of the quantum state evolved from an initial coherent state. Owing to the gain (or loss) operation of the selecting operator exp(kλδsinθ)\exp(\frac{k\lambda}{\delta}\sin\theta), only two component coherent states appear for t2t\geq 2, in clear contrast to the conventional kicked rotor corresponding to λ=0\lambda=0 (see Fig. 1 for comparison). Namely, the state proliferation problem in the latter is effectively suppressed here by the selection operator. Meanwhile, we can see that the positions of the two component coherent states are well predicted by the pseudoclassical map.

An interesting feature of the 𝒫𝒯\mathcal{PT}-symmetric kicked rotor is that it can generate the directed current [18, 19]. Here we study this property with α\alpha being slightly perturbed from the quantum resonance condition. Fig. 7 compares the expected values of the momentum, p\langle p\rangle, for quantum and pseudoclassical dynamics. It can be observed that as δ\delta decreases, they do converge and pt\langle p\rangle\sim t, which implies that the directed current also exists in the deep quantum regime near quantum resonances and interestingly, it has a pseudoclassical explanation.

5 Summary

In this study, by considering the quantum effect of the free rotation at quantum resonances, a generalized pseudoclassical theory is designed for the kicked rotor model. Its effectiveness suggests that, even in the deep quantum regime, quantum dynamics may have a close connection to classical dynamics. With regard to this, one may wonder if it is imperative to extend the conventional quantum chaos study from the semiclassical regime [1] to the deep quantum regime. In this context, it depends on how general the pseudoclassical theory could be, which warrants further investigation, except the kicked top and the kicked rotor model, if the pseudoclassical theory can be extended to other Floquet systems.

Extensive experiments have been conducted on the kicked rotor model with cold atoms [24] owing to its paradigmatic role in illustrating quantum chaos. It would be enticing to examine the effects of the pseudoclassical dynamics near higher-order quantum resonances. To this end, it is necessary to adapt the pseudoclassical theory to the kicked particle model first, which is in progress.

\Acknowledgements

This work is supported by the National Natural Science Foundation of China (Grants No. 12075198, No. 12247106, and No. 12247101).

\InterestConflict

The authors declare that they have no conflict of interest.

References

  • [1] M. Berry, Physica Scripta 40, 335 (1989).
  • [2] F. Haake, S. Gnutzmann, and M. Kuś, Quantum Signatures of Chaos (4th ed.) (Springer, 2018).
  • [3] S. Fishman, I. Guarneri, and L. Rebuzzini, Phys. Rev. Lett. 89, 084101 (2002).
  • [4] S. Fishman, I. Guarneri, and L. Rebuzzini, J. Stat. Phys. 110, 911 (2003).
  • [5] G. Casati, B. V. Chirikov, F. M. Izraelev, and J. Ford, Lect. Notes Phys. 93 (1979) 334.
  • [6] G. Casati and B. V. Chirikov, Quantum Chaos: Between Order and Disorder (Cambridge University Press, Cambridge, England, 1995).
  • [7] F. M. Izrailev, and D. L. Shepelyanskii, Theor. Math. Phys. 43, 553 (1980).
  • [8] L. Rebuzzini, I. Guarneri, and R. Artuso, Phys. Rev. A 79, 033614 (2009).
  • [9] H. Wang, J. Wang, I. Guarneri, G. Casati, and J. Gong, Phys. Rev. E 88, 052919 (2013).
  • [10] Z. Zou and J. Wang, Entropy 24, 1092 (2022).
  • [11] F. Haake, M. Kuś, and R. Scharf, Z. Physik B - Condensed Matter  65, 381-395 (1987).
  • [12] F. Haake and D. L. Shepelyansky, Europhys. Lett. 5, 671 (1988).
  • [13] I. Dana, E. Eisenberg, and N. Shnerb, Phys. Rev. E 54, 5948 (1996).
  • [14] B. C. Berndt, R. J. Evans, K. S. Williams, Gauss and Jacobi Sums (Wiley, 1998).
  • [15] K. Takahashi and N. Saitôn, Phys. Rev. Lett. 55, 645 (1985).
  • [16] F. M. Izrailev, and D. L. Shepelyanskii, Physics Reports 196, 299-392 (1990).
  • [17] C. M. Bender and S. Boettcher Phys. Rev. Lett. 80, 5243 (1998).
  • [18] S. Longhi, Phys. Rev. A 95, 012125 (2017).
  • [19] W. Zhao, J. Wang, X. Wang, and P. Tong, Phys. Rev. E 99, 042201 (2019).
  • [20] E. M. Graefe, M. Höning, and H. J. Korsch, J. Phys. A: Math. Theor. 43, 075306 (2010).
  • [21] C. T. West, T. Kottos, and T. Prosen, Phys. Rev. Lett. 104, 054102 (2010).
  • [22] C. M. Bender, J. Phys.: Conf. Ser. 631 012002 (2015).
  • [23] R. El-Ganainy, K. G. Makris, M. Khajavikhan, et al, Nat. Phys., 14: 11-19 (2018).
  • [24] J.-C. Garreau, C. R. Phys. 18, 31 (2017).

Appendix A Derivation of Equation (8)

Regarding the eigenstates {|n}\{|n\rangle\} of the operator ν^\hat{\nu}, the coherent state |p~,θ~|{\tilde{p},\tilde{\theta}}\rangle can be expressed as |p~,θ~=ncn|n|{\tilde{p},\tilde{\theta}}\rangle=\sum_{n}c_{n}|n\rangle. Applying the operator exp(i2πrsν^2)\exp\left(-i\frac{2\pi r}{s}\hat{\nu}^{2}\right) to both sides,

exp(i2πrsν^2)|p~,θ~=k=0s1exp(i2πrsk2)mod(n,s)=kcn|n.\displaystyle\exp\left(-i\frac{2\pi r}{s}\hat{\nu}^{2}\right)|{\tilde{p},\tilde{\theta}}\rangle=\sum_{k=0}^{s-1}\exp\left(-i\frac{2\pi r}{s}k^{2}\right)\sum_{\mod(n,s)=k}c_{n}|n\rangle. (20)

Note that two coherent states separating in θ\theta by ϕ\phi can be related by the translation operator exp(iν^ϕ)\exp(-i\hat{\nu}\phi), we have

|p~,θ~+2πrsl\displaystyle|\tilde{p},\tilde{\theta}+\frac{2\pi r}{s}l\rangle =\displaystyle= exp(iν^2πrsl)|p~,θ~\displaystyle\exp\left(-i\hat{\nu}\frac{2\pi r}{s}l\right)|\tilde{p},\tilde{\theta}\rangle (21)
=\displaystyle= k=0s1exp(i2πkrsl)mod(n,s)=kcn|n\displaystyle\sum_{k=0}^{s-1}\exp\left(-i\frac{2\pi kr}{s}l\right)\sum_{\mod(n,s)=k}c_{n}|n\rangle

by setting ϕ=2πlr/s\phi=2\pi lr/s, where ll is an integer. Thereafter, multiplying both sides with exp(i2πλlr/s)\exp(i2\pi\lambda lr/s), where λ\lambda is an integer, 0λs10\leq\lambda\leq s-1, and taking summation over ll from l=0l=0 to s1s-1,

l=0s11sexp(i2πrsλl)|pc,θc+2πrsl=mod(n,s)=λcn|n.\displaystyle\sum_{l=0}^{s-1}\frac{1}{s}\exp\left(i\frac{2\pi r}{s}\lambda l\right)|p_{c},\theta_{c}+\frac{2\pi r}{s}l\rangle=\sum_{\mod(n,s)=\lambda}c_{n}|n\rangle. (22)

Finally, by replacing λ\lambda with kk and substituting this equation into Eq.  (20), we get Eq. (8).

Appendix B Translational symmetry of cases C1 and C2

To prevent the proliferation problem of the pseudoclassical dynamics, one way is to introduce the translational symmetry into the phase space so that two or more state points will be mapped into one by f\mathcal{M}_{f}. This can be fulfilled by setting the proper integer value of ω\omega.

First, for case C1 where ss is odd, based on the analytical results by Ref. [14], all ss Gaussian sums are nonzero so that 𝒩=s\mathcal{N}=s and any two neighboring state points resulted by acting f\mathcal{M}_{f} to a given state are separated in θ\theta by 2π/s2\pi/s. Hence, to ensure the resultant state points overlap at the next step, we can set ω=s\omega=s. It ensures that the number of states is up-bounded by ss throughout.

For case C2, where ss is even, note that the Gaussian sum GlG_{l} can be rewritten as

Gl\displaystyle G_{l} =\displaystyle= 1sk=0s1exp(i2πrsk(kl))\displaystyle\frac{1}{s}\sum_{k=0}^{s-1}\exp\left(-i\frac{2\pi r}{s}k(k-l)\right) (23)
=\displaystyle= 1sk=0s/21exp(i2πrsk(kl))+exp(i2πrs(k+s2)(k+s2l))\displaystyle\frac{1}{s}\sum_{k=0}^{s/2-1}\exp\left(-i\frac{2\pi r}{s}k(k-l)\right)+\exp\left(-i\frac{2\pi r}{s}(k+\frac{s}{2})(k+\frac{s}{2}-l)\right)
=\displaystyle= 1sk=0s/21exp(i2πrsk(kl))(1+exp(i2πr(s4l2))).\displaystyle\frac{1}{s}\sum_{k=0}^{s/2-1}\exp\left(-i\frac{2\pi r}{s}k(k-l)\right)\left(1+\exp\left(-i2\pi r\left(\frac{s}{4}-\frac{l}{2}\right)\right)\right).

It is evident that the last term, 1+exp(i2πr(s4l2))1+\exp\left(-i2\pi r\left(\frac{s}{4}-\frac{l}{2}\right)\right), is zero for all odd ll when mod(s,4)=0(s,4)=0 and is zero for all even ll when mod(s,4)=2(s,4)=2. On the one hand, this implies that half of all ss Gaussian sums are zero, whereas the other half are nonzero based on Ref. [14], so that 𝒩=s/2\mathcal{N}=s/2. On the other hand, this suggests that of all s/2s/2 state points resulted by acting f\mathcal{M}_{f} to a given state, any two neighboring points are separated in θ\theta by 4π/s4\pi/s. Thus, similar to case C1, to ensure the evolving state points overlap at the following steps, we can set ω=𝒩=s/2\omega=\mathcal{N}=s/2. This guarantees that the number of states is up-bounded by 𝒩=s/2\mathcal{N}=s/2 throughout.

Appendix C Commutation relation of UfU_{f} and UδU_{\delta}

In the representation of ν^\hat{\nu}, the elements of UfU_{f} and UδU_{\delta} are, respectively,

(Uf)m,m\displaystyle(U_{f})_{m,m^{\prime}} =\displaystyle= exp(i2πrsm2)δm,m,\displaystyle\exp\left(-i\frac{2\pi r}{s}m^{2}\right)\delta_{m,m^{\prime}}, (24)
(Uδ)m,m\displaystyle(U_{\delta})_{m,m^{\prime}} =\displaystyle= 12πexp(im22δ)02πexp(i(mm)θ)exp(ikωδcos(ωθ))𝑑θ,\displaystyle\frac{1}{2\pi}\exp\left(-i\frac{m^{2}}{2}\delta\right)\int_{0}^{2\pi}\exp\left(-i(m-m^{\prime})\theta\right)\exp\left(-i\frac{k}{\omega\delta}\cos\left(\omega\theta\right)\right)d\theta, (25)

following which the matrix element for the commutator (UfUδUδUf)m,m(U_{f}U_{\delta}-U_{\delta}U_{f})_{m,m^{\prime}} is

(UfUδUδUf)m,m\displaystyle(U_{f}U_{\delta}-U_{\delta}U_{f})_{m,m^{\prime}} =\displaystyle= n(Uf)m,n(Uδ)n,mn(Uδ)m,n(Uf)n,m\displaystyle\sum_{n}(U_{f})_{m,n}(U_{\delta})_{n,m^{\prime}}-\sum_{n}(U_{\delta})_{m,n}(U_{f})_{n,m^{\prime}} (26)
=\displaystyle= ((Uf)mm(Uf)m,m)(Uδ)m,m\displaystyle((U_{f})_{mm}-(U_{f})_{m^{\prime},m^{\prime}})(U_{\delta})_{m,m^{\prime}}
=\displaystyle= (exp(i2πrsm2)exp(i2πrsm2))exp(im22δ)2π02πexp(i(mm)θ)exp(ikωδcos(ωθ))𝑑θ.\displaystyle\left(\exp\left(-i\frac{2\pi r}{s}m^{2}\right)-\exp\left(-i\frac{2\pi r}{s}m^{\prime 2}\right)\right)\frac{\exp\left(-i\frac{m^{2}}{2}\delta\right)}{2\pi}\int_{0}^{2\pi}\exp\left(-i(m-m^{\prime})\theta\right)\exp\left(-i\frac{k}{\omega\delta}\cos\left(\omega\theta\right)\right)d\theta.

Note that exp(ikωδcos(ωθ))\exp\left(-i\frac{k}{\omega\delta}\cos\left(\omega\theta\right)\right) is periodic with the period 2π/ω2\pi/\omega; thus, the integration term in the above equation can be rewritten as

02πexp(i(mm)θ)exp(ikωδcos(ωθ))𝑑θ\displaystyle\int_{0}^{2\pi}\exp\left(-i(m-m^{\prime})\theta\right)\exp\left(-i\frac{k}{\omega\delta}\cos\left(\omega\theta\right)\right)d\theta =\displaystyle= 02πω(u=0ω1exp(i(mm)(θ+2πuω)))exp(ikωδcos(ωθ))𝑑θ\displaystyle\int_{0}^{\frac{2\pi}{\omega}}\left(\sum_{u=0}^{\omega-1}\exp\left(-i(m-m^{\prime})\left(\theta+\frac{2\pi u}{\omega}\right)\right)\right)\exp\left(-i\frac{k}{\omega\delta}\cos\left(\omega\theta\right)\right)d\theta (27)
=\displaystyle= 02πω(exp(i(mm)θ)u=0ω1exp(i2π(mm)uω))exp(ikωδcos(ωθ))𝑑θ\displaystyle\int_{0}^{\frac{2\pi}{\omega}}\left(\exp\left(-i(m-m^{\prime})\theta\right)\sum_{u=0}^{\omega-1}\exp\left(-i\frac{2\pi(m-m^{\prime})u}{\omega}\right)\right)\exp\left(-i\frac{k}{\omega\delta}\cos\left(\omega\theta\right)\right)d\theta
=\displaystyle= ωδmod(mm,ω),002πωexp(i(mm)θ)exp(ikωδcos(ωθ))𝑑θ.\displaystyle\omega\delta_{\mathrm{mod}(m-m^{\prime},\omega),0}\int_{0}^{\frac{2\pi}{\omega}}\exp\left(-i(m-m^{\prime})\theta\right)\exp\left(-i\frac{k}{\omega\delta}\cos\left(\omega\theta\right)\right)d\theta.

Substituting it into Eq. (26),

(UfUδUδUf)m,m\displaystyle(U_{f}U_{\delta}-U_{\delta}U_{f})_{m,m^{\prime}} =\displaystyle= (exp(i2πrsm2)exp(i2πrsm2))×ω2πexp(im22δ)\displaystyle\left(\exp\left(-i\frac{2\pi r}{s}m^{2}\right)-\exp\left(-i\frac{2\pi r}{s}m^{\prime 2}\right)\right)\times\frac{\omega}{2\pi}\exp\left(-i\frac{m^{2}}{2}\delta\right) (28)
δmod(mm,ω),002πωexp(i(mm)θ)exp(ikωδcos(ωθ))𝑑θ.\displaystyle\delta_{\mathrm{mod}(m-m^{\prime},\omega),0}\int_{0}^{\frac{2\pi}{\omega}}\exp\left(-i(m-m^{\prime})\theta\right)\exp\left(-i\frac{k}{\omega\delta}\cos\left(\omega\theta\right)\right)d\theta.

For case C1 where ss is odd, as ω=s\omega=s, the delta function in Eq. (28) suggests that mm=nsm-m^{\prime}=ns (nn is an integer). Substituting it into Eq. (26), it becomes

(UfUδUδUf)m,m=0\displaystyle(U_{f}U_{\delta}-U_{\delta}U_{f})_{m,m^{\prime}}=0 (29)

because the difference term in Eq. (26) is

(exp(i2πrsm2)exp(i2πrsm2))\displaystyle\left(\exp\left(-i\frac{2\pi r}{s}m^{2}\right)-\exp\left(-i\frac{2\pi r}{s}m^{\prime 2}\right)\right) =\displaystyle= exp(i2πrsm2)(1exp(i2πr(n2s2mn)))\displaystyle\exp\left(-i\frac{2\pi r}{s}m^{2}\right)\left(1-\exp\left(-i2\pi r(n^{2}s-2mn)\right)\right) (30)
=\displaystyle= 0.\displaystyle 0.

However, for case C2 where ss is even and ω=s2\omega=\frac{s}{2}, the difference term in Eq. (26) does not equal zero in general, and thus, UfUδUδUf0U_{f}U_{\delta}-U_{\delta}U_{f}\neq 0. However, if we split the Floquet operator UU as U=UfUδU=U^{\prime}_{f}U^{\prime}_{\delta} with Uf=Ufexp(iπrν)U^{\prime}_{f}=U_{f}\exp(i\pi r\nu) and Uδ=exp(iπrν)UδU^{\prime}_{\delta}=\exp(-i\pi r\nu)U_{\delta}, as shown in the following, UfU^{\prime}_{f} and UδU^{\prime}_{\delta} do commutate.

The matrix element for UfU^{\prime}_{f} and UδU^{\prime}_{\delta} is, respectively,

(Uf)m,m=(Uf)m,mexp(iπrm)and(Uδ)m,m=exp(iπrm)(Uδ)m,m,\displaystyle(U^{\prime}_{f})_{m,m^{\prime}}=(U_{f})_{m,m^{\prime}}\exp(i\pi rm^{\prime}){\rm~{}~{}and~{}~{}}(U^{\prime}_{\delta})_{m,m^{\prime}}=\exp(-i\pi rm)(U_{\delta})_{m,m^{\prime}}, (31)

and thus,

(UfUδUδUf)m,m\displaystyle(U^{\prime}_{f}U^{\prime}_{\delta}-U^{\prime}_{\delta}U^{\prime}_{f})_{m,m^{\prime}} =\displaystyle= n(Uf)m,n(Uδ)n,mn(Uδ)m,n(Uf)n,m\displaystyle\sum_{n}(U^{\prime}_{f})_{m,n}(U^{\prime}_{\delta})_{n,m^{\prime}}-\sum_{n}(U^{\prime}_{\delta})_{m,n}(U^{\prime}_{f})_{n,m^{\prime}} (32)
=\displaystyle= ((Uf)mm(Uf)m,m)(Uδ)m,m\displaystyle((U^{\prime}_{f})_{mm}-(U^{\prime}_{f})_{m^{\prime},m^{\prime}})(U^{\prime}_{\delta})_{m,m^{\prime}}
=\displaystyle= (exp(i2πrsm2+iπrm)exp(i2πrsm2+iπrm))12πexp(im22δiπrm)\displaystyle\left(\exp\left(-i\frac{2\pi r}{s}m^{2}+i\pi rm\right)-\exp\left(-i\frac{2\pi r}{s}m^{\prime 2}+i\pi rm^{\prime}\right)\right)\frac{1}{2\pi}\exp\left(-i\frac{m^{2}}{2}\delta-i\pi rm\right)
×02πexp(i(mm)θ)exp(ikωδcos(ωθ))dθ.\displaystyle\times\int_{0}^{2\pi}\exp\left(-i(m-m^{\prime})\theta\right)\exp\left(-i\frac{k}{\omega\delta}\cos\left(\omega\theta\right)\right)d\theta.

Note that here the integration term is the same as Eq. (26); however, the delta function suggests mm=ns2m-m^{\prime}=n\frac{s}{2} instead (nn is an integer). Similarly, substituting it into Eq. (32), we obtain

(UfUδUδUf)m,m=0,\displaystyle(U^{\prime}_{f}U^{\prime}_{\delta}-U^{\prime}_{\delta}U^{\prime}_{f})_{m,m^{\prime}}=0, (33)

because the difference term in Eq. (32) is

exp(i2πrsm2+iπrm)exp(i2πrsm2+iπrm)\displaystyle\exp\left(-i\frac{2\pi r}{s}m^{2}+i\pi rm\right)-\exp\left(-i\frac{2\pi r}{s}m^{\prime 2}+i\pi rm^{\prime}\right) =\displaystyle= exp(i2πrsm2+iπrm)(1exp(i2πr(mnn(n1)s4)))\displaystyle\exp\left(-i\frac{2\pi r}{s}m^{2}+i\pi rm\right)\left(1-\exp\left(-i2\pi r\left(mn-\frac{n(n-1)s}{4}\right)\right)\right) (34)
=\displaystyle= 0\displaystyle 0

for any even ss.