Chinese Physics Letters, 2022, Vol. 39, No. 7, Article code 073101 Dynamics of Quantum State and Effective Hamiltonian with Vector Differential Form of Motion Method Long Xiong (熊龙)1, Wei-Feng Zhuang (庄伟峰)1, and Ming Gong (龚明)1,2* Affiliations 1CAS Key Lab of Quantum Information, School of Physics, University of Science and Technology of China, Hefei 230026, China 2Synergetic Innovation Center of Quantum Information and Quantum Physics, University of Science and Technology of China, Hefei 230026, China Received 8 April 2022; accepted manuscript online 31 May 2022; published online 29 June 2022 *Corresponding author. Email: gongm@ustc.edu.cn Citation Text: Xiong L, Zhuang W F, and Gong M 2022 Chin. Phys. Lett. 39 073101    Abstract Effective Hamiltonians in periodically driven systems have received widespread attention for realization of novel quantum phases, non-equilibrium phase transition, and Majorana mode. Recently, the study of effective Hamiltonian using various methods has gained great interest. We consider a vector differential equation of motion to derive the effective Hamiltonian for any periodically driven two-level system, and the dynamics of the spin vector are an evolution under the Bloch sphere. Here, we investigate the properties of this equation and show that a sudden change of the effective Hamiltonian is expected. Furthermore, we present several exact relations, whose expressions are independent of the different starting points. Moreover, we deduce the effective Hamiltonian from the high-frequency limit, which approximately equals the results in previous studies. Our results show that the vector differential equation of motion is not affected by a convergence problem, and thus, can be used to numerically investigate the effective models in any periodic modulating system. Finally, we anticipate that the proposed method can be applied to experimental platforms that require time-periodic modulation, such as ultracold atoms and optical lattices.
cpl-39-7-073101-fig1.png
cpl-39-7-073101-fig2.png
cpl-39-7-073101-fig3.png
cpl-39-7-073101-fig4.png
cpl-39-7-073101-fig5.png
DOI:10.1088/0256-307X/39/7/073101 © 2022 Chinese Physics Society Article Text Periodic modulation systems have become an important concern in ultracold atoms, atomic and molecular physics, and quantum optics recently because they can produce exotic phases[1–12] and spin-orbit coupling.[13–17] Thus, they have paved the way for research of some fundamental principles in modern physics, such as many-body localization,[18–22] time crystals, and ergodicity.[23–27] For an arbitrary time-depending Hamiltonian, $H(t+T) = H(t)$, where $T = 2\pi/\omega$ is the time period. We can define $H(t) = H_0 + \sum_{n} H_n e^{itn\omega}$. Then, in the high-frequency limit, the effective Floquet Hamiltonian, in analogy to the Bloch theorem in a spatially periodic system,[28] can be written as $$ H_{\rm eff} = H_0 + \sum_{n > 0} {[H_n, H_{-n}] \over n\omega} + \mathcal{O}(\omega^{-2}).~~ \tag {1} $$ Here, the high-frequency $\omega$ ensures that higher-order terms are negligible.[29–37] Various methods have been used to derive the above Hamiltonian. In the Magnus expansion method,[38–41] a necessary but not sufficient condition for convergence is given by $$ \int_0^T ||H(t)||_2\,dt \le \pi,~~ \tag {2} $$ where $||\cdot||_2$ is the norm-2 of the Hamiltonian. The same criteria are applied to the Floquet expansion in Eq. (1), regarding the equivalence between both methods in deriving the effective Hamiltonians. Due to the non-commuting relations between $H_n$ and $H_{-n}$, the above effective Hamiltonian can acquire some new terms excluded in the original Hamiltonian, which are essential for their novel applications. The modulating method provides a new way to explore and discover fundamental physics in the areas of many Majorana modes, Floquet topological insulator, and time crystal. Vector Differential Form of Motion Method. We study the effective Hamiltonian for a periodically arbitrary Hamiltonian using the equation-of-motion method, and discuss the generic features and interesting results. Here, we consider a two-level model subjected to the Schrödinger equation:[42–46] $$ i \hbar \frac{\partial }{\partial t} \psi(t) = H(t) \psi(t), \quad H(t) = \sum_{i=1}^3 h_i \sigma^i = {\boldsymbol h}\cdot {\boldsymbol \sigma},~~ \tag {3} $$ where $\sigma^i$ (for $i=x, y, z$) are the three traceless Pauli matrices; $\psi(t)=[\psi_1(t),\psi_2(t)]^{\rm T}$ is the system wave function with two components and vector ${\boldsymbol h} = (h_1, h_2, h_3)$. The dynamics of the wave function can be described by the following unitary transformation: $$ \psi(t) = U(t, t_0) \psi(t_0),\quad iU_t(t,t_0) = H(t) U(t,t_0).~~ \tag {4} $$ Note that the unitary operator depends strongly on its initial value $t_0$. For this two-band model in Eq. (3), we assume[47] $$ U= \exp(-i \theta {\boldsymbol n}\cdot {\boldsymbol \sigma}) = \cos\theta - i {\boldsymbol n}\cdot {\boldsymbol \sigma} \sin\theta ,~~ \tag {5} $$ where ${\boldsymbol n}(t) = (n_1, n_2, n_3)$, $|{\boldsymbol n}| = 1$, and $\theta = \theta(t)$. Then, we derive the effective Hamiltonian using an equation of motion. By introducing Eq. (5) into the Schrodinger equation, we find $$\begin{alignat}{1} i(-\dot{\theta}-i\dot{\boldsymbol n}\cdot {\boldsymbol \sigma})\sin\theta + {\boldsymbol n} \cdot {\boldsymbol \sigma} (\cos\theta) \dot{\theta} ={\boldsymbol h} \cdot {\boldsymbol \sigma} U(t). ~~~~ \tag {6} \end{alignat} $$ The real and imaginary parts of the above equation give the following two equations: $$\begin{align} \dot{\theta} = {\boldsymbol n}\cdot {\boldsymbol h}, ~~ \dot{\boldsymbol n} = {\boldsymbol n}\times {\boldsymbol \omega},~~ \tag {7} \end{align} $$ where $$ {\boldsymbol \omega}=-{\boldsymbol h}-({\boldsymbol n}\times {\boldsymbol h}){\rm\cot}\theta .~~ \tag {8} $$ The above Bloch equation limits the amplitude of the spin, thus $|{\boldsymbol n}| = 1$. Both different equations above are subjected to the boundary conditions $U(t_0, t_0) = 1$ and $U(pT + t_0, t_0) = [U(T + t_0, t_0)]^p$, for $p \in \mathbb{Z}^+$. This constraint can be demonstrated by $$\begin{align} & {\boldsymbol n}(pT + t_0, t_0) = {\boldsymbol n}(T + t_0, t_0),~~ \tag {9} \\ & \theta(pT + t_0, t_0) = p \theta(T + t_0, t_0) {\rm mod}(2\pi). \end{align} $$ For different initial values, their expressions vary. The above equation of motion has a solution when $H(t) = H_0$ is independent of time. Thus, we can define ${\boldsymbol n} = {\boldsymbol h}_0 // ||{\boldsymbol h}_0||$ and $\theta = {\boldsymbol n} \cdot {\boldsymbol h} t$. Here, ${\boldsymbol n}\times {\boldsymbol h}_0 = 0$, then the singular point in the Cotangent function is negligible.
cpl-39-7-073101-fig1.png
Fig. 1. Equivalence between the two effective Hamiltonians $H_{{\rm eff}}^{i}$ with different starting times $t_i$ ($i$ = I, II). Both Hamiltonians are connected by a unitary transformation determined using $U(t_2, t_1)$.
For a periodically modulating system with period $T$, the above unitary operator can be expressed using an effective Floquet Hamiltonian as follows: $$ U(t_0+T, t_0) = \exp(-iH_{\rm eff}T).~~ \tag {10} $$ The effective Floquet Hamiltonian[48–58] depends on the initial time $t_0$ and the time period $T$. The equivalence between these Floquet Hamiltonians is schematically presented in Fig. 1. Since both effective Hamiltonians are subject to two starting times, $t_1$ and $t_2$, the evolution from $t_1$ to $t_2 + T$ may have two trajectories if $t_1 < t_2$, $$ U(t_2+T, t_2) U(t_2, t_1) = U(t_2+T, t_1+T)U(t_1+T, t_1).~~ \tag {11} $$ Due to the periodicity of the driving Hamiltonian, $U(t_2, t_1) = U(t_2+T, t_1+T)$, and results in the formula, $$ \exp(-iH_{\rm eff}^{{\rm II}}) = \exp[-i U(t_2, t_1)^† H_{\rm eff}^{{\rm I}} U(t_2, t_1)].~~ \tag {12} $$ Hence, the Floquet system described by the effective Hamiltonian at any initial time is equivalent.
cpl-39-7-073101-fig2.png
Fig. 2. Classical interpretation of a vector trajectory on the Bloch sphere. The evolution of the quantum state is an evolution of a spin vector evolving in an effective magnetic field. One term is made by the Hamiltonian due to the Bloch evolution, and the other due to the periodic oscillation of the magnetic field.
Some calculations consider the unitary transformations to the Hamiltonian, in which the vector ${\boldsymbol n}$ and angle $\theta$ need corresponding changes. We denote the unitary transformation operator as follows: $$ \mathcal{U} = e^{-iS}, \quad S = \phi {\boldsymbol m}\cdot {\boldsymbol \sigma}.~~ \tag {13} $$ By this unitary transformation, we assume $$ \mathcal{U}^† U(t) \mathcal{U} = e^{-i\theta'(t) {\boldsymbol n}'(t)\cdot {\boldsymbol \sigma}}.~~ \tag {14} $$ Following the same procedure and after a few transformation, we find $$\begin{align} &\theta' = {\boldsymbol n}(t)\cdot{\boldsymbol h} = \theta, \\ &{\boldsymbol n}' = {\boldsymbol n}(t) + \sin(2 \phi) {\boldsymbol n}(t)\times {\boldsymbol m} +2 \sin^2\phi \,{\boldsymbol n}\times {\boldsymbol m}\times {\boldsymbol m}, \\ &{\boldsymbol \omega}' = 1+\sin(2 \phi) {\boldsymbol m} + 2 \sin^2 \phi \,{\boldsymbol m}\times {\boldsymbol m}.~~ \tag {15} \end{align} $$ Notably, these new variables still satisfy the equation of motion, i.e., $\dot{\boldsymbol n}'(t) = {\boldsymbol \omega}'\times {\boldsymbol n}'$. For a time-dependent unitary transformation, the expressions can still be solved similarly. Time-Evolution Process. To gain a deep understanding of the dynamics for effective Hamiltonian and corresponding wave functions, ${\boldsymbol n}$ and $\theta$ are presented using the two-band model: $$\begin{alignat}{1} H=\begin{pmatrix} h_z+ k\cos(\omega t) & g+k_1 \sin(\omega t +\phi)\\ g+k_1 \sin(\omega t +\phi) & -h_z-k\cos(\omega t) \end{pmatrix},~~~~~~~ \tag {16} \end{alignat} $$ where $h_z$, $g$, $k$, and $k_1$ are adjustable parameters. This model is relevant to qubits in various quantum systems, such as NV color centers, superconducting qubits, and trapped ions. For $k = k_1 = 0$, the system degenerates to the time-independent case. Then, the following expression suffices: $$\begin{align} \exp \Big(\int_{0}^{\rm T} i H d t\Big)&=\exp \Big(\int_{0}^{\rm T} i {\boldsymbol h} \cdot {\boldsymbol \sigma} d t\Big)\\ &=\exp (i {\boldsymbol h} \cdot {\boldsymbol \sigma} T) =\exp (-i H_{\mathrm{eff}} T)\\ &=\exp (-i {\boldsymbol n} \cdot {\boldsymbol \sigma} \theta).~~ \tag {17} \end{align} $$ From Eq. (7), when the system is time-independent, the angle $\theta(t)$ increases linearly and its value reaches $\pi$, and its evolution is described by the black line in Fig. 5. The component of vector ${\boldsymbol n}$ remains constant. Moreover, our numerical calculation confirms this phenomenon. However, due to the influence of the time-dependent term, $\theta(t)$ fluctuates periodically, as exemplified in Figs. 3(a)–3(d). For the periodic modulation, the angle $\theta$ can be directly deduced from the differential equation and it is neither $0$ nor $\pi$ when $t>0$. When $\theta$ approaches these two singular points, the terms $-{\boldsymbol n}\times ({\boldsymbol n}\times {\boldsymbol h}){\rm\cot}\theta $ will dominate since ${\rm\cot}\theta $ approaches infinity, which will quickly transform ${\boldsymbol n} \rightarrow -{\boldsymbol n}$. During the fluctuation of vector ${\boldsymbol n}$, $\theta$ shifts slightly, thus the unitary operator satisfies a smooth function of time. However, in weak modulation, as $\theta$ ranges from 0 to $\pi$, ${\boldsymbol n}(t)$ and ${\boldsymbol h}(t)$ are almost parallel to each other, and we have $$\begin{align} {\boldsymbol n} \times {\boldsymbol h}=C \cdot \sin \phi \approx C \cdot \phi \approx 0,~~ \tag {18} \end{align} $$ where $\varphi$ is the angle between ${\boldsymbol n}(t)$ and ${\boldsymbol h}(t)$. From Eq. (7), when $\theta(t)$ approaches $\pi$, we obtain $$\begin{align} \cot \theta \approx \frac{1}{\theta-\pi}.~~ \tag {19} \end{align} $$ Hence, $[{\boldsymbol n}(t) \times {\boldsymbol h}(t)] \cot [\theta(t)]$ always remains finite, leading to the continuous fluctuation of $n(t)$ and $\theta(t)$ from perturbation. As long as there is a time-dependent term, the system will be quasi-periodic, which is consistent with the Floquet system. However, for the strong modulation, the upper and lower limits of $\theta$ significantly deviate from these two bounds (0 and $\pi$). As discussed before, both bounds are attained only when the Hamiltonian is time-independent and ${\boldsymbol n} \times {\boldsymbol h} = 0$, regarding that the effective equation of motion is exact. Figure 5 shows that $\theta(t)$ translates with an interval of $2\pi$. The first boundary condition is verified from results in Figs. 3(b)–3(d). In these calculations, the results were presented with an initial time $t_0 = 0$. In addition, we see that ${\boldsymbol n}(t, t_0) \ne {\boldsymbol n}(t+p T, t_0)$ for $t_0 \ne 0$ and $p \in \mathbb{Z}^+$. This result can be understood from the basic feature that $U(t+p T, t_0) = U(t, t_0)[U(t_0+T, t_0)]^p$, in which the time-evolution operator $U(t, t_0)$ breaks the periodicity of ${\boldsymbol n}(t+p T, t_0)$ for $t_0 \ne 0$. The intersection points of the black line and all colored lines can be described by the function $$\begin{align} \theta(n T) =|2\,m \pi-n \theta(T)|,~~ \tag {20} \end{align} $$ where $m$ is a positive integer and $m$ does not equal $n$. Obviously, $\theta(n T)$ ranges from $0$ to $\pi$.
cpl-39-7-073101-fig3.png
Fig. 3. Two parameter sets were used. The first set is $k=0.7$, $k_1=0.6$ (red line), the second is $k=6\times10^{-9}$, $k_1=6\times10^{-9}$ (blue line). It is obvious that the greater the amplitude of the time term, the more intense the fluctuation of the curve.
Notably, through a period of time evolution, the wave function never attains its initial value. We illustrate this result by calculating the Loschmidt echo $L = \langle \psi(t_0)|U(t,t_0)| \psi(t_0)\rangle$, and plot the value of $\ln(1-L)$ against time in Figs. 3(e) and 3(f). In the weak modulation, the wave function infinitesimally approaches but never attains the initial value. Also, we demonstrate the dynamics of wave function in the Bloch sphere in Fig. 4. Suppose Bloch vectors ${\boldsymbol P} = (n_{x1},n_{x2},n_{x3})$, then we project the vector ${\boldsymbol n}(t)=[{\boldsymbol n}_1(t),{\boldsymbol n}_2(t),{\boldsymbol n}_3(t)]$ onto the Bloch sphere as $$\begin{align} &n_{x 1}=\sin b_{\theta} \cos b_{\phi}, \quad n_{x 2}=\sin b_{\theta} \sin b_{\phi},\\ &n_{x 3}=\cos b_{\theta},~~ \tag {21} \end{align} $$ with $$b_{\theta}=\operatorname{{\rm Arccos}}n_{3},~~ b_{\phi}=i \ln \Big(\frac{n_{1}-i n_{2}}{\sqrt{(n_{1})^{2}+(n_{2})^{2}}}\Big). $$ Apparently, $n_{x3}=n_{3}$. We find that the dynamics exhibit some complex trajectories, and this vector almost goes through all points in the whole Bloch sphere in the long time limit. This is a general result, independent of whether the modulating Hamiltonian is trivial or topological.
cpl-39-7-073101-fig4.png
Fig. 4. Dynamics of a vector in a periodically driven system at $t = T$, $10T$, and $50T$. Since $\theta \ne \pi, 0$ in most cases, the quantum states cover the whole Bloch sphere after long periods. This feature is independent of whether the Hamiltonian is topological nontrivial or trivial.
cpl-39-7-073101-fig5.png
Fig. 5. Curves of $\theta(t)$ and $2 m\pi \pm \theta(t)$ shown at $nT$. The black line indicates the fluctuation of $\theta(t)$.
Applications. The vector differential equation-of-motion method can be used to overcome the divergence existing in most high-frequency series expansions and to study the dynamics of some physical models in low-frequency regimes. Here, we discuss an interesting application of this equation-of-motion method. We try to derive the effective Hamiltonian in the high-frequency limit from our equation. Consider the following vector $$ {\boldsymbol h}={\boldsymbol h}_0+\sum_{p \ge 1}{\boldsymbol h}_p e^{i p \omega t}+{\boldsymbol h}_{-p}e^{-ip\omega t}.~~ \tag {22} $$ The vector ${\boldsymbol n}$ is assumed as $$ {\boldsymbol n}={\boldsymbol n}_0+ \sum_{p \ge 1}{\boldsymbol n}_p e^{ip\omega t}+{\boldsymbol n}_{-p} e^{-ip\omega t}.~~ \tag {23} $$ Suppose $h_p$ for $p \ne 0$ is much smaller than $h_0$, i.e., $|{\boldsymbol n}_p| \ll {\boldsymbol n}_0$ and $|{\boldsymbol n}_p| \ll 1$. Moreover, ${\boldsymbol n}_0 = {\boldsymbol h}_0//||{\boldsymbol h}_0||$. By substituting these expressions into the equation, we find $$ \theta(T) = \Big({\boldsymbol n}_0 \cdot {\boldsymbol h}_0 + \sum_{p \ne 0} {\boldsymbol n}_p \cdot {\boldsymbol h}_{-p}\Big) {2\pi \over \omega}.~~ \tag {24} $$ Other terms, such as ${\boldsymbol n}_p \cdot {\boldsymbol h}_{p'}$, when $p + p' \ne 0$, disappear by integrating from $0$ to $T$. Here, we only consider this initial condition since all Floquet Hamiltonians with different initial times are equivalent. The high-frequency limit ensures that $\theta(T)$ considered small enough, so that the limit can be derived as $$\begin{align} \cot[\theta(T)] & = \frac{\omega}{2\pi {\boldsymbol n}_0 \cdot {\boldsymbol h}_0+i\mathcal{Q}} \\ & \approx \frac{\omega}{2\pi^2|{\boldsymbol h}_0| }-\frac{i\omega}{4\pi^2 |{\boldsymbol h}_0|^2 } \mathcal{Q},~~ \tag {25} \end{align} $$ where $$ \mathcal{Q} = \sum_{p \ge 1} {\boldsymbol n}_0 \cdot {\boldsymbol h}_{-p}+ {\boldsymbol n}_{-p} \cdot {\boldsymbol h}_0-{\boldsymbol n}_0 \cdot {\boldsymbol h}_{p}- {\boldsymbol n}_{p} \cdot {\boldsymbol h}_0.~~ \tag {26} $$ For the second vector differential equation of motion, we find $$\begin{align} \dot{\boldsymbol n}={}&\sum_{p \ge 1} ip\omega [\cos(p\omega t)({\boldsymbol n}_p-{\boldsymbol n}_{-p})\\ &-p\omega \sin(p\omega t)({\boldsymbol n}_p+{\boldsymbol n}_{-p})].~~ \tag {27} \end{align} $$ Since the imaginary terms do not exist, ${\boldsymbol n}_p={\boldsymbol n}_{-p}$. By substituting these results into the equation and keeping only the leading term, we obtain the formula $$\begin{align} \dot{\boldsymbol n}&=-{\boldsymbol n}_0 \times {\boldsymbol h}_0 - \cot(\theta) {\boldsymbol n}_0 \times ({\boldsymbol n}_0 \times {\boldsymbol h}_0) \\ &=-\sum_{p \ge 1} p \omega \sin(p\omega t)({\boldsymbol n}_p+{\boldsymbol n}_{-p}).~~ \tag {28} \end{align} $$ From this result, we find $$ p \omega |{\boldsymbol h}_0|{\boldsymbol n}_p=i \epsilon_{ijk} {\boldsymbol h}_p \cdot {\boldsymbol h}_{-p}.~~ \tag {29} $$ With these expressions, we can obtain the following Floquet Hamiltonian: $$\begin{align} H_{\rm eff}&=\frac{{\boldsymbol n} \theta}{T}\cdot {\boldsymbol \sigma}\\ &=\frac{[{\boldsymbol n}_0+\sum_{p \ge 1}({\boldsymbol n}_p+{\boldsymbol n}_{-p})]({\boldsymbol n}_0 \cdot {\boldsymbol h}_0) \frac{2\pi}{\omega}}{T}\cdot {\boldsymbol \sigma} \\ &={\boldsymbol n}_0+\sum_{p \ge 1}({\boldsymbol n}_p+{\boldsymbol n}_{-p})|{\boldsymbol h}_0|\cdot {\boldsymbol \sigma} \\ &={\boldsymbol h}_0 \cdot {\boldsymbol \sigma} + \sum_{p \ge 1} \frac{[{\boldsymbol h}_p \cdot {\boldsymbol \sigma},{\boldsymbol h}_{-p} \cdot {\boldsymbol \sigma}]}{p\omega},~~ \tag {30} \end{align} $$ which is identical to the first and second terms in Eq. (1) obtained in the literature.[28] Therefore, we are able to directly obtain the effective Floquet Hamiltonian from our vector differential equation-of-motion method. Furthermore, this method can be generalized to Hamiltonians with greater orders. Noticeably, one representative work[59] found that a higher non-commuting term can produce the long-range hopping more easily in a periodical driving system. Specifically, the scheme uses a 1D periodic p-wave superconductor and provides good guidance for effectively generating long-range hopping in experiments. Thus, the generalization of our proposed method can be used to obtain higher-order non-commuting terms, giving a different perspective to the study of many Majorana mode problems. Discussion and Conclusion. We have presented the vector differential equation of motion to study any periodically driven two-level system and deduced the effective Floquet Hamiltonian. Notably, evolution can be regarded as a spin vector evolving in an effective magnetic field. Furthermore, we discuss an interesting application of this method in the high-frequency limit of the effective Floquet Hamiltonian, corresponding to the results of the second-order Magnus. Finally, we expect that our method can provide some insights into degenerate structures and multi-level systems and that its generalization can give a different perspective to the study of many Majorana modes. Acknowledgment. This work was supported by the National Natural Science Foundation of China (Grant No. 11774328).
References Tomography of Band Insulators from Quench DynamicsRealizing the Harper Hamiltonian with Laser-Assisted Tunneling in Optical LatticesRealization of the Hofstadter Hamiltonian with Ultracold Atoms in Optical LatticesFloquet topological states in shaking optical latticesShaping topological properties of the band structures in a shaken optical latticeTopological insulator and particle pumping in a one-dimensional shaken optical latticeFloquet-Bloch Theory and Topology in Periodically Driven LatticesExperimental reconstruction of the Berry curvature in a Floquet Bloch bandTunable Chern insulator with optimally shaken latticesTowards large-Chern-number topological phases by periodic quenchingPhotonic Floquet topological insulatorsFloquet topological insulator in semiconductor quantum wellsMeasuring the Chern number of Hofstadter bands with ultracold bosonic atomsExperimental realization of the topological Haldane model with ultracold fermionsSpin–orbit coupling in quantum gasesTopological Fulde–Ferrell–Larkin–Ovchinnikov states in spin–orbit-coupled Fermi gasesRealization of two-dimensional spin-orbit coupling for Bose-Einstein condensatesA quantum Newton's cradleObservation of many-body localization of interacting fermions in a quasirandom optical latticeDicke quantum phase transition with a superfluid gas in an optical cavityObservation of the Efimovian expansion in scale-invariant Fermi gasesQuantum thermalization through entanglement in an isolated many-body systemFloquet time crystal in the Lipkin-Meshkov-Glick modelEquilibration and order in quantum Floquet matterObservation of a discrete time crystalFloquet Time CrystalsDiscrete Time Crystals: Rigidity, Criticality, and RealizationsThe Magnus expansion and some of its applicationsHigh-frequency Floquet theory: Test of the applicability of the golden ruleHigh-frequency approximation for periodically driven quantum systems from a Floquet-space perspectiveFloquet-Bloch theory of high-harmonic generation in periodic structuresUniversal high-frequency behavior of periodically driven systems: from dynamical stabilization to Floquet engineeringEffective Hamiltonians for Rapidly Driven Many-Body Lattice Systems: Induced Exchange Interactions and Density-Dependent HoppingsEffective time-independent description of optical lattices with periodic drivingHigh-frequency Floquet-theory content of wave-packet dynamicsFloquet theory: exponential perturbative treatmentEquivalence of the Floquet–Magnus and Fer Expansions to Investigate the Dynamics of a Spin System in the Three-Level SystemEffective Hamiltonians for periodically driven systemsEffective Theory of Floquet Topological TransitionsPeriodically Driven Quantum Systems: Effective Hamiltonians and Engineered Gauge FieldsEffective Hamiltonians in Floquet theory of magic angle spinning using van Vleck transformationQuasienergy and Floquet states in a time-periodic driven two-level systemTwo-level quantum dynamics, integrability, and unitary NOT gatesTime evolution of two-level systems driven by periodic fieldsQuasi-Periodic Solutions for Two-Level SystemsFloquet spectrum for two-level systems in quasiperiodic time-dependent fieldsSolution of the Schrödinger Equation with a Hamiltonian Periodic in TimeFloquet scattering theory of quantum pumpsStochastic resonanceFloquet theory and applicationsAnalytical approach for the Floquet theory of delay differential equationsFloquet theorem with open systems and its applicationsGeneralized Floquet Theory: Application to Dynamical Systems with Memory and Bloch’s Theorem for Nonlocal PotentialsSteady States and Quasienergies of a Quantum-Mechanical System in an Oscillating FieldPeriodic Thermodynamics of Isolated Quantum SystemsGenerating many Majorana modes via periodic driving: A superconductor model
[1] Hauke P, Lewenstein M, and Eckardt A 2014 Phys. Rev. Lett. 113 045303
[2] Miyake H, Siviloglou G A, Kennedy C J, Burton W C, and Ketterle W 2013 Phys. Rev. Lett. 111 185302
[3] Aidelsburger M, Atala M, Lohse M, Barreiro J T, Paredes B, and Bloch I 2013 Phys. Rev. Lett. 111 185301
[4] Zheng W and Zhai H 2014 Phys. Rev. A 89 061603
[5] Zhang S L and Zhou Q 2014 Phys. Rev. A 90 051601
[6] Mei F, You J B, Zhang D W, Yang X C, Fazio R, Zhu S L, and Kwek L C 2014 Phys. Rev. A 90 063638
[7] Gómez-León A and Platero G 2013 Phys. Rev. Lett. 110 200403
[8] Fläschner N, Rem B, Tarnowski M, Vogel D, Lühmann D S, Sengstock K, and Weitenberg C 2016 Science 352 1091
[9] Verdeny A and Mintert F 2015 Phys. Rev. A 92 063615
[10] Xiong T S, Gong J, and An J H 2016 Phys. Rev. B 93 184306
[11] Rechtsman M C, Zeuner J M, Plotnik Y, Lumer Y, Podolsky D, Dreisow F, Nolte S, Segev M, and Szameit A 2013 Nature 496 196
[12] Lindner N H, Refael G, and Galitski V 2011 Nat. Phys. 7 490
[13] Aidelsburger M, Lohse M, Schweizer C, Atala M, Barreiro J T, Nascimbene S, Cooper N, Bloch I, and Goldman N 2015 Nat. Phys. 11 162
[14] Jotzu G, Messer M, Desbuquois R, Lebrat M, Uehlinger T, Greif D, and Esslinger T 2014 Nature 515 237
[15] Galitski V and Spielman I B 2013 Nature 494 49
[16] Zhang W and Yi W 2013 Nat. Commun. 4 2711
[17] Wu Z, Zhang L, Sun W, Xu X T, Wang B Z, Ji S C, Deng Y, Chen S, Liu X J, and Pan J W 2016 Science 354 83
[18] Kinoshita T, Wenger T, and Weiss D S 2006 Nature 440 900
[19] Schreiber M, Hodgman S S, Bordia P, Lüschen H P, Fischer M H, Vosk R, Altman E, Schneider U, and Bloch I 2015 Science 349 842
[20] Baumann K, Guerlin C, Brennecke F, and Esslinger T 2010 Nature 464 1301
[21] Deng S, Shi Z Y, Diao P, Yu Q, Zhai H, Qi R, and Wu H 2016 Science 353 371
[22] Kaufman A M, Tai M E, Lukin A, Rispoli M, Schittko R, Preiss P M, and Greiner M 2016 Science 353 794
[23] Russomanno A, Iemini F, Dalmonte M, and Fazio R 2017 Phys. Rev. B 95 214307
[24] Moessner R and Sondhi S 2017 Nat. Phys. 13 424
[25] Zhang J, Hess P, Kyprianidis A et al. 2017 Nature 543 217
[26] Else D V, Bauer B, and Nayak C 2016 Phys. Rev. Lett. 117 090402
[27] Yao N Y, Potter A C, Potirniche I D, and Vishwanath A 2017 Phys. Rev. Lett. 118 030401
[28] Blanes S, Casas F, Oteo J A, and Ros J 2009 Phys. Rep. 470 151
[29] Lefebvre R 1999 Int. J. Quantum Chem. 72 261
[30] Eckardt A and Anisimovas E 2015 New J. Phys. 17 093039
[31] Faisal F H M and Kamiński J Z 1997 Phys. Rev. A 56 748
[32] Bukov M, D'Alessio L, and Polkovnikov A 2015 Adv. Phys. 64 139
[33] Itin A P and Katsnelson M I 2015 Phys. Rev. Lett. 115 075301
[34] Hemmerich A 2010 Phys. Rev. A 81 063626
[35] Wells J, Simbotin I, and Gavrila M 1997 Phys. Rev. A 56 3961
[36] Casas F, Oteo J, and Ros J 2001 J. Phys. A 34 3379
[37] Mananga E S 2017 J. Phys. Chem. A 121 6063
[38] Rahav S, Gilary I, and Fishman S 2003 Phys. Rev. A 68 013820
[39] Kundu A, Fertig H, and Seradjeh B 2014 Phys. Rev. Lett. 113 236803
[40] Goldman N and Dalibard J 2014 Phys. Rev. X 4 031027
[41] Ramesh R and Krishnan M S 2001 J. Chem. Phys. 114 5967
[42] Zhao X G 1994 Phys. Rev. B 49 16753
[43] Angelo R M and Wreszinski W F 2005 Phys. Rev. A 72 034105
[44] Barata J A C A and Cortez D A 2002 Phys. Lett. A 301 350
[45] Gentile G 2003 Commun. Math. Phys. 242 221
[46] Blekher P, Jauslin H, and Lebowitz J 1992 J. Stat. Phys. 68 271
[47]Scott W R 2012 Group Theory (Massachusetts: Courier)
[48]Zel'Dovich Y B 1967 Sov. Phys.-JETP 24 1006
[49]Ritus V 1967 Sov. Phys.-JETP 24 1041
[50] Shirley J H 1965 Phys. Rev. 138 B979
[51] Moskalets M and Büttiker M 2002 Phys. Rev. B 66 205320
[52] Gammaitoni L, Hänggi P, Jung P, and Marchesoni F 1998 Rev. Mod. Phys. 70 223
[53] Barone S R, Narcowich M A, and Narcowich F J 1977 Phys. Rev. A 15 1109
[54] Simmendinger C, Wunderlin A, and Pelster A 1999 Phys. Rev. E 59 5344
[55] Dai C, Shi Z, and Yi X 2016 Phys. Rev. A 93 032121
[56] Traversa F L, Di Ventra M, and Bonani F 2013 Phys. Rev. Lett. 110 170602
[57] Sambe H 1973 Phys. Rev. A 7 2203
[58] Lazarides A, Das A, and Moessner R 2014 Phys. Rev. Lett. 112 150401
[59] Tong Q J, An J H, Gong J, Luo H G, and Oh C 2013 Phys. Rev. B 87 201109