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

Transition from eigenmodes to geometric modes characterized by the quantum SU(2) coupled oscillator model: a review

Open Access Open Access

Abstract

The quantum and classical dynamics of the SU(2) coupled oscillator model are systematically reviewed to provide the quantum eigenstates and stationary coherent states for characterizing laser transverse modes from the analogy with the quantum-classical connection. The integral formula for the representation of the stationary coherent states derived from the evolution of the time-dependent wave packet state is completely reviewed. Several calculated results for the stationary coherent states are illustratively presented to display the spatial distributions for the quantum-classical transition and the plentiful variations of phase singularities. The overall review is believed to provide a comprehensive insight into laser transverse modes characterized by the stationary coherent states of the SU(2) coupled oscillator model.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Plane waves are the most fundamental wave distributions, while structured waves which display spatial features essentially different from plane waves are naturally to appear in the wave equation with external potentials [1], such as atomic orbits [2,3], eigenstates of quantum dots [46], resonant modes in laser resonators [711], acoustic waves [12,13], Landau states in a magnetic field [1418], etc. From the characteristics of the interference of multiple plane waves, it can be shown that the structured waves generally contain phase singularities [19]. Since the probability current density swirls around phase singularities, the term of vortex is used to represent the phase singularity [20]. The distribution of phase singularities is intimately associated with the orbital angular momentum density (OAMD). Although vortices appeared in many early studies of optical and quantum waves, the systematic exploration of phase singularities was first given in 1974 by Nye and Berry in ultrasonic fields [21] and by Hirschfelder et al. in quantum wave functions [22,23]. The quantum theory has been extensively exploited to study the distribution of phase singularities not only in quantum physics but also in different branches of physics [2429]. Nowadays, the exploration of structured coherent waves with vortices becomes an emerging area of research.

For a spherical laser cavity, the eigenmodes are usually represented as Hermite-Gaussian (HG) modes with rectangular symmetry or Laguerre-Gaussian (LG) modes with circular symmetry. In addition to HG and LG modes, the generalized representation of eigenmodes can be expressed as the Hermite-Laguerre-Gaussian (HLG) modes by using the quantum SU(2) coupled oscillator model [30]. The SU(2) Lie algebra has been developed to investigate various quantum phenomena such as entanglement in coherent states [31], generation and evolution of vortex states [32,33], orbital magnetism in quantum dots [34], charge particles in external fields [35], and shell effects in nuclei and metallic clusters [36]. A systematic review for laser transverse modes characterized by the SU(2) coupled oscillator model can be found in Ref. [30]. In experimental aspects, the HLG modes can be generated by using a spherical laser cavity with astigmatism or by using an astigmatic mode converter (AMC) to transform HG modes.

Besides eigenmodes, the high-order transverse modes in the degenerate cavities often exhibit the so-called geometric modes with the intensities concentrated on the geometric rays. The domination of eigenmodes or geometric modes is mainly dependent on the transverse order N and the ratio ΔfTxfTy, where ΔfTx and ΔfTy are the transverse mode spacings in the two eigenaxes x and y, respectively. The laser cavity to be degenerate is determined by the condition of the ratio ΔfTxfTy to be the rational number p/q, where p and q are coprime integers. Experimental results revealed that the lasing transverse modes were usually dominated by the eigenmodes for N < p with N < q and dominated by the geometric modes for N >p with N > q. The geometric modes corresponding to HG eigenmodes are called the Lissajous geometric modes, the counterpart corresponding to LG eigenmodes are the trochoidal geometric modes. By nature, the geometric modes corresponding to HLG eigenmodes can be called the Lissajous-trochoidal geometric modes. In theoretical aspects, the quantum SU(2) coupled oscillator model has been used to derive the stationary coherent states that not only can display the wave intensities well localized on the classical trajectories but also can exhibit the continuous transformation between Lissajous and trochoid geometric curves [37,38]. Nevertheless, since the accidental degeneracies and the high-order Hermite polynomials are involved in computations, the spatial structures and phase singularities are rater difficult to evaluate. Recently, we have originally developed the integral method to accurately compute the stationary coherent states for manifesting the spatially phase singularities of the Lissajous geometric modes [3945,30,4648]. So far, the integral method has not been applied to the general case for the transition between the HLG eigenmode and counterpart geometric modes. It will be greatly useful to develop the generalized integral method from the quantum SU(2) coupled oscillator model for utterly understanding the laser transverse modes related to the transition between the HLG eigenmode and geometric modes.

In this paper, we systematically overview the quantum wavefunctions and classical trajectories for the SU(2) coupled oscillator. We further develop the integral method based on the Fourier transformation to represent the stationary coherent state of the SU(2) coupled oscillator as an integral evolution of the time-dependent wave packet state over a complete period. The developed integral formula is then exploited to analyze the spatial structures of the stationary coherent states for manifesting the transition from low-order eigenmodes to geometric modes. The calculated results for the spatial distributions can not only reveal the quantum-classical connection but also display the plentiful variations of phase singularities and edge dislocations. More importantly, the overall review and the developed model can help readers employ the integral formula to reconstruct the structured optical waves related to HLG modes and Lissajous-trochoidal geometric modes that are experimentally transformed by using AMC for the HG modes and Lissajous geometric modes, respectively.

2. SU(2) quantum oscillator model

The quantum SU(2) states are related to the attractive oscillator system, whereas the quantum SU(1,1) states are associated with the repulsive oscillator system. Since the resonant modes of the laser cavity are analogous to the states of the confined system, the quantum SU(2) states can be analogously employed to characterize laser transverse modes generated from cavities with various astigmatism. The generalized Hamiltonian related to the SU(2) interaction for the coupled oscillator systems can be given by [33,37,38]

$$\hat{H} = {\hat{H}_o} + \sum\limits_{i = 1}^3 {\hbar {\Omega _i}\;{{\hat{L}}_i}}, $$
where
$${\hat{H}_o} = ({a_1^\mathrm{\dagger }{a_1} + a_2^\mathrm{\dagger }{a_2} + 1} )\,\hbar {\omega _o}, $$
$${\hat{L}_1} = 1/2({a_1^\mathrm{\dagger }{a_2} + a_2^\mathrm{\dagger }{a_1}} ), $$
$${\hat{L}_2} ={-} i/2({a_1^\mathrm{\dagger }{a_2} - a_2^\mathrm{\dagger }{a_1}} ), $$
$${\hat{L}_3} = 1/2({a_1^\mathrm{\dagger }{a_1} - a_2^\mathrm{\dagger }{a_2}} ), $$
${\omega _o}$ is the natural frequency, ${a_i}$ and $a_i^\dagger$ are mode operators satisfying Bosonic algebra, and ${\Omega _i}$ are the coupling frequencies. The ${\hat{H}_o}$ operator is the two-dimensional (2D) isotropic oscillator. The ${\hat{L}_i}$ operators proposed by Schwinger [49] satisfy the usual angular-momentum commutation relations, i.e. Lie commutator algebra [${L_i}$, ${L_j}$] $= i{\varepsilon _{i,j,k}}\,{L_k}$, where the Levi-Civita tensor ${\varepsilon _{i,j,k}}$ is equal to +1 and –1 for even and odd permutations of its indices, respectively, and zero otherwise.

The classical counterpart of the system $\hat{H}$ in Eq. (1) can be expressed by using the dimensionless variables ${a_1} = {{({\tilde{x} + i\,{{\tilde{p}}_x}} )} / {\sqrt 2 }}$, $a_1^\dagger = {{({\tilde{x} - i\,{{\tilde{p}}_x}} )} / {\sqrt 2 }}$, ${a_2} = {{({\tilde{y} + i\,{{\tilde{p}}_y}} )} / {\sqrt 2 }}$, and $a_2^\dagger = {{({\tilde{y} - i\,{{\tilde{p}}_y}} )} / {\sqrt 2 }}$. The dimensionless variables $\tilde{x}$, $\tilde{y}$, ${\tilde{p}_x}$, and ${\tilde{p}_y}$ are given by $\tilde{x} = x\sqrt {\mu {\omega _o}/\hbar } $, $\tilde{y} = y\sqrt {\mu {\omega _o}/\hbar } $, ${\tilde{p}_x} = {p_x}\sqrt {1/\hbar \mu {\omega _o}} $, and ${\tilde{p}_y} = {p_y}\sqrt {1/\hbar \mu {\omega _o}} $, where μ is the mass of the particle, x and y are the position variables, and px and py are the momentum variables. In terms of $\tilde{x}$, $\tilde{y}$, ${\tilde{p}_x}$, and ${\tilde{p}_y}$, the classical Hamiltonian for $H$ in Eq. (1) is given by

$$\begin{aligned}{l} H &= \frac{{\hbar {\omega _o}}}{2}({{{\tilde{x}}^2} + \tilde{p}_x^2 + {{\tilde{y}}^2} + \tilde{p}_y^2} )+ \frac{{\hbar {\Omega _1}}}{2}({\tilde{x}\tilde{y} + {{\tilde{p}}_x}{{\tilde{p}}_y}} )\\ &\textrm{ } + \frac{{\hbar {\Omega _2}}}{2}({\tilde{x}{{\tilde{p}}_y} - \tilde{y}{{\tilde{p}}_x}} )+ \frac{{\hbar {\Omega _3}}}{4}({{{\tilde{x}}^2} + \tilde{p}_x^2 - {{\tilde{y}}^2} - \tilde{p}_y^2} )\end{aligned}. $$

From Eq. (6), the classical equation of motion for the Hamiltonian H can be derived as

$$\frac{d}{{dt}}\left[ {\begin{array}{c} {{v_1}}\\ {{v_2}} \end{array}} \right] ={-} i\left[ {\begin{array}{cc} {{\omega_o} + ({\Omega _3}/2)}&{({\Omega _1} - i\,{\Omega _2})/2}\\ {({\Omega _1} + i\,{\Omega _2})/2}&{{\omega_o} - ({\Omega _3}/2)} \end{array}} \right]\;\left[ {\begin{array}{c} {{v_1}}\\ {{v_2}} \end{array}} \right], $$
where ${v_1} = (\tilde{x} + i\,{\tilde{p}_x})/\sqrt 2$ and ${v_2} = (\tilde{y} + i\,{\tilde{p}_y})/\sqrt 2$ . Equation (7) is not only the classical equation of motion for the generalized SU(2) coupled oscillators; interestingly, its form is the same as the Schrödinger equation for a 2-level system such as spinor states in a magnetic field [50]. The general solution for Eq. (7) can be solved to be
$$\left[ \begin{array}{l} {v_1}(t)\\ {v_2}(t) \end{array} \right] = \left[ {\begin{array}{cc} {{e^{ - i\alpha /2}}\cos (\beta /2)}&{ - {e^{ - i\alpha /2}}\sin (\beta /2)}\\ {{e^{i\alpha /2}}\sin (\beta /2)}&{{e^{i\alpha /2}}\cos (\beta /2)} \end{array}} \right]\;\left[ \begin{array}{l} {u_1}(t)\\ {u_2}(t) \end{array} \right], $$
where $\alpha = {\tan ^{ - 1}}({{{{\Omega _2}} / {{\Omega _1}}}} )$, $\beta = {\tan ^{ - 1}}({{\sqrt {\Omega _1^2 + \Omega _2^2} } / {{\Omega _3}}})$, ${u_1}(t) = \sqrt {{N_1}} \,{e^{\, - i\,({\omega _1}\,t - {\phi _{\,1}})}}$, ${u_2}(t) =$$\sqrt {{N_2}} \,{e^{\, - i\,({\omega _2}\,t - {\phi _{\,2}})}}$, ${\omega _1} = {\omega _o} + (\Omega /2)$, ${\omega _2} = {\omega _o} - (\Omega /2)$, and $\Omega = \sqrt {\Omega _1^2 + \Omega _2^2 + \Omega _3^2} $, where ${N_1}$, ${N_2}$ , ${\phi _{\,1}}$, and ${\phi _{\,2}}$ are related to the initial conditions.

Considering the frequency ratio ${\omega _1}/{\omega _2}$ to be a rational number $q/p$, the characteristic frequencies can be written as ${\omega _1} = q\omega$ and ${\omega _2} = p\omega$ with $\omega = 2{\omega _o}/(p + q)$ by using ${\omega _1} + {\omega _2} = 2{\omega _o}$ and ${\omega _1}/{\omega _2} = q/p$, where q and p are coprime integers,. Note that the integer q is always positive; however, the integer p can be either positive for ${\omega _o} > (\Omega /2)$ or negative for ${\omega _o} < (\Omega /2)$ . By using the relations of ${v_1} = (\tilde{x} + i\,{\tilde{p}_x})/\sqrt 2$ and ${v_2} = (\tilde{y} + i\,{\tilde{p}_y})/\sqrt 2$, the classical periodic orbit can be obtained from $\tilde{x} = \sqrt 2 Re ({v_1})$ and $\tilde{y} = \sqrt 2 Re ({v_2})$. Consequently, the parametric equations for the classical trajectory for ${\omega _1}/{\omega _2} = q/p$ can be found by using Eq. (8) and explicitly expressed as

$$\left[ \begin{array}{l} \tilde{x}(t)\\ \\ \tilde{y}(t) \end{array} \right] = \left[ \begin{array}{l} \sqrt {2{N_1}} \cos \left( {\frac{\beta }{2}} \right)\cos \left( {q\omega t + \frac{\alpha }{2} - {\phi_1}} \right) - \sqrt {2{N_2}} \sin \left( {\frac{\beta }{2}} \right)\cos \left( {p\omega t + \frac{\alpha }{2} - {\phi_2}} \right)\\ \sqrt {2{N_1}} \sin \left( {\frac{\beta }{2}} \right)\cos \left( {q\omega t - \frac{\alpha }{2} - {\phi_1}} \right) + \sqrt {2{N_2}} \cos \left( {\frac{\beta }{2}} \right)\cos \left( {p\omega t - \frac{\alpha }{2} - {\phi_2}} \right) \end{array} \right]. $$

Equation (9) indicates that the periodic orbits of the SU(2) coupled oscillators display a wide variety of famous curves that are determined by the values of α and β. For $\alpha = 0$, the orbits in Eq. (9) are Lissajous figures at an angle of β/2 with respect to the x-axis. For $\alpha = \pi /2$ and $\beta = \pi /2$, the orbits are hypotrochoidal or epitrochoidal curves, depending on whether the integer p is positive or negative [37]. Overall, the orbits in Eq. (9) can be called Lisssjous-trochoidal curves since these curves are associated with a continuous transformation between Lissajous figures and trochoidal curves by varying the values of the parameter α and β.

The quantum eigenstates of the coupled oscillator $\hat{H}$ in Eq. (1) can be analytically derived with the SU(2) algebra via defining a new pair of operators

$$\left[ \begin{array}{l} b_1^\dagger \\ b_2^\dagger \end{array} \right] = \left[ {\begin{array}{cc} {{e^{ - i\alpha /2}}\cos (\beta /2)}&{{e^{i\alpha /2}}\sin (\beta /2)}\\ { - {e^{ - i\alpha /2}}\sin (\beta /2)}&{{e^{i\alpha /2}}\cos (\beta /2)} \end{array}} \right]\;\left[ \begin{array}{l} a_1^\dagger \\ a_2^\dagger \end{array} \right], $$

In terms of $b_1^\dagger$ and $b_2^\dagger$, the coupled oscillator $\hat{H}$ can be untangled as

$$\hat{H} = ({b_1^\mathrm{\dagger }{b_1} + 1/2} )\;{\omega _1} + ({b_2^\mathrm{\dagger }{b_2} + 1/2} )\;{\omega _2}. $$

Consequently, the quantum eigenstates of the Hamiltonian $\hat{H}$ can be given by

$${|{{n_1},{n_2}} \rangle _{\hat{H}}} = \frac{{{{(b_1^\mathrm{\dagger })}^{{n_1}}}}}{{\sqrt {{n_1}!} }}\frac{{{{(b_2^\mathrm{\dagger })}^{{n_2}}}}}{{\sqrt {{n_2}!} }}{|{0,0} \rangle _{\hat{H}}}, $$
where the indices $({n_1},{n_2})$ are the quantum numbers. The eigenvalue corresponding to ${|{{n_1},{n_2}} \rangle _H}$ is given by
$${E_{{n_1},{n_2}}} = ({{n_1} + 1/2} )\,\hbar {\omega _1} + ({{n_2} + 1/2} )\,\hbar {\omega _2}. $$

Substituting Eq. (10) into Eq. (12) and using the fact ${|{0,0} \rangle _{\hat{H}}} = {|{0,0} \rangle _{{{\hat{H}}_o}}}$, the eigenstates ${|{{n_1},{n_2}} \rangle _{\hat{H}}}$ can be explicitly expressed as a linear combination of the eigenstates of ${\hat{H}_o}$:

$${|{{n_1},{n_2}} \rangle _{\hat{H}}} = {e^{iN\alpha /2}}\sum\limits_{{m_1} = 0}^N {{e^{ - i\,{m_1}\,\alpha }}} \,d_{{m_1} - \frac{N}{2},\,{n_1} - \frac{N}{2}}^{\frac{N}{2}}(\beta )\,|{{m_1},{m_2}} \rangle {}_{{{\hat{H}}_o}}, $$
where $N = {n_1} + {n_2} = {m_1} + {m_2}$ and
$$\begin{aligned} d_{{m_1} - \frac{N}{2},\,{n_1} - \frac{N}{2}}^{\frac{N}{2}}(\beta )= &\sqrt {{m_\textrm{1}}\textrm{ }!(N - {m_1})\textrm{ }!{n_1}\textrm{ }!(N - {n_1})\textrm{ }!} \;\\ &{\kern 1cm}\textrm{ } \times \sum\limits_{\nu = \max [0,{m_1} - {n_1}]}^{\min [N - {n_1},{m_1}]} {\textrm{ }\frac{{{{( - 1\textrm{)}}^\nu }{{[\cos (\beta /2)]}^{\,N - {n_1} + {m_1} - 2\nu }}{{[\sin (\beta /2)]}^{\,{n_1} - {m_1} + 2\nu }}}}{{\nu \textrm{ }!(N - {n_1} - \nu )\textrm{ !}\,\textrm{(}{m_\textrm{1}}\textrm{ - }\nu \textrm{) !(}{n_\textrm{1}} - {m_1} + \nu \textrm{) !}}}} \end{aligned}, $$
are the Wigner d-matrix elements [51]. The real space representation for ${|{{n_1},{n_2}} \rangle _H}$ is given by [37]
$$\psi _{{n_{\,1}},{n_{\,2}}}^{(\hat{H})}(\tilde{x},\tilde{y}) = \left\langle {\tilde{x},\tilde{y}} \right.{|{{n_1},{n_2}} \rangle _{\hat{H}}} = \left[ {{e^{iN\alpha /2}}\sum\limits_{{m_{\;1}} = 0}^N {{e^{ - i\,{m_{\,1}}\,\alpha }}} \,d_{{m_{\,1}} - \frac{N}{2},\,{n_{\,1}} - \frac{N}{2}}^{\frac{N}{2}}(\beta )\,{\psi_{{m_{\;1}}}}(\tilde{x})\;{\psi_{N - {m_{\;1}}}}(\tilde{y})} \right]. $$
where ${\psi _n}$ is given by
$${\psi _n}(\tilde{x}) = \frac{1}{{\sqrt {{2^n}n\,!\sqrt \pi } }}{H_n}(\tilde{x})\;{e^{ - {{\tilde{x}}^2}/2}}, $$
and ${H_n}(\tilde{x})$ are the Hermite polynomials.

3. Quantum stationary coherent states manifesting classical periodic orbits

The quantum stationary coherent states manifesting the classical periodic orbits can be derived by using the Schrödinger coherent state that is obtained from the generating function for the Hermite polynomials given by [52]

$${e^{ - {\tau ^2} + 2\tau \,\tilde{x}}} = \sum\limits_{n = 0}^\infty {\;\frac{{{\tau ^n}}}{{n\,!}}{H_n}(\tilde{x})}. $$

Note that the generating function is valid for all complex values of τ. Setting $\tau = u/\sqrt 2$ and multiplying the term ${e^{ - (|u{|^2} + {{\tilde{x}}^2})/2}}$ on the both sides of Eq. (18), after some rearrangement and in terms of ${\psi _n}({\tilde{x}} )$, the Schrödinger coherent state can be given by

$$g(\tilde{x};u) = \frac{1}{{{\pi ^{1/4}}}}\exp \left[ { - \,\,\frac{{({{\tilde{x}}^2} - 2\sqrt 2 u\,\tilde{x} + {u^2} + |u{|^2})}}{2}} \right] = \;\sum\limits_{n = 0}^\infty {\;\frac{{{u^n}}}{{\sqrt {n\,!} }}{e^{ - |u{|^2}/2}}\;{\psi _n}(\tilde{x})}. $$
The Schrödinger coherent state corresponds to a decentered Gaussian wave packet that is a coherent superposition of all eigenmodes . Substituting into Eq. (19), the decentered Gaussian wave packet can be rewritten as
$$g(\tilde{x};u) = \frac{1}{{{\pi ^{1/4}}}}\,\;{e^{i\,\tilde{x}\sqrt 2 {\mathop{\rm Im}\nolimits} (u)}}\;{e^{ - iRe (u){\mathop{\rm Im}\nolimits} (u)}}\;exp \left\{ { - \;\frac{{{{[\tilde{x} - \sqrt 2 Re (u)]}^{\,2}}}}{2}} \right\}. $$
The intensity ${|{g(\tilde{x};u)} |^2}$ clearly represents a Gaussian function with the central peak at $\tilde{x} = \sqrt 2 Re (u)$.

From Eqs. (19) and (20), the time-dependent Gaussian wave packet state linking to the trajectory $\tilde{x} = \sqrt 2 Re ({v_1})$ and $\tilde{y} = \sqrt 2 Re ({v_2})$ can be expressed as the product of two Schrödinger coherent states:

$$\begin{aligned} g(\tilde{x},{v_1})g(\tilde{y},\;{v_2}) &= \frac{1}{{\sqrt \pi }}{e^{\frac{{ - ({{\tilde{x}}^2} - 2\sqrt {2\,} {v_1}\,\tilde{x} + v_1^2 + |{v_1}{|^2})}}{2}}}\;{e^{\frac{{ - ({{\tilde{y}}^2} - 2\sqrt 2 \;{v_2}\;\tilde{y} + v_2^2 + |{v_2}{|^2})}}{2}}}\\ &= {e^{ - (|{v_1}{|^2} + |{v_2}{|^2})/2}}\sum\limits_{{m_1} = 0}^\infty {\;\sum\limits_{{m_2} = 0}^\infty {\;\frac{{v_1^{{m_1}}}}{{\sqrt {{m_1}\,!} }}\;\frac{{v_2^{{m_2}}}}{{\sqrt {{m_2}\,!} }}\;{\psi _{{m_{\,1}}}}(\tilde{x}){\psi _{{m_{\;2}}}}(\tilde{y})} } \end{aligned}$$

In terms of $a_1^\dagger$ and $a_2^\dagger$, the wave packet state $g(\tilde{x},{v_1})g(\tilde{y},\;{v_2})$ can be written as

$$g(\tilde{x},{v_1})g(\tilde{y},\;{v_2}) = {e^{ - ({v_1}{|^2} + |{v_2}{|^2})/2}}\sum\limits_{{m_1} = 0}^\infty {\;\sum\limits_{{m_2} = 0}^\infty {\;\frac{{{{({v_1}a_1^\mathrm{\dagger })}^{{m_1}}}}}{{{m_1}\,!}}\;\frac{{{{({v_2}a_2^\mathrm{\dagger })}^{{m_2}}}}}{{{m_2}\,!}}{\psi _0}(\tilde{x}){\psi _0}(\tilde{y})} }. $$
Using the relationship between $({{v_1},\,{v_2}} )$ and $({{u_1},\,{u_2}} )$ in Eq. (8) with $|{u_1}{|^2} + |{u_2}{|^2} = |{v_1}{|^2} + |{v_2}{|^2}$ and the relationship between $({a_1^\dagger ,\,a_2^\dagger } )$ and $({b_1^\dagger ,\,b_2^\dagger } )$ in Eq. (10), the wave packet state $g(\tilde{x},{v_1})g(\tilde{y},\;{v_2})$ can be changed as
$$g(\tilde{x},{v_1})g(\tilde{y},\;{v_2}) = {e^{ - (|{u_{\,1\,}}{|^2} + |{u_{\,2}}{|^2})/2}}\textrm{ }\sum\limits_{{n_1} = 0}^\infty {\;\sum\limits_{{n_2} = 0}^\infty {\;\frac{{{{({u_1}b_1^\mathrm{\dagger })}^{{n_{\,1}}}}}}{{{n_1}\,!}}\;\frac{{{{({u_2}b_2^\mathrm{\dagger })}^{{n_{\,2}}}}}}{{{n_2}\,!}}{\psi _0}(\tilde{x}){\psi _0}(\tilde{y})} }. $$

Consequently, Eq. (12) can be used to express the wave packet state $g(\tilde{x},{v_1})g(\tilde{y},\;{v_2})$ as the coherent superposition of the quantum eigenstates $\psi _{{n_{\,1}},{n_{\,2}}}^{(\hat{H})}(\tilde{x},\tilde{y})$:

$$\begin{aligned} g(\tilde{x},{v_1})g(\tilde{y},\;{v_2}) &= \frac{1}{{\sqrt \pi }}{e^{\frac{{ - ({{\tilde{x}}^2} - 2\sqrt {2\,} {v_1}\,\tilde{x} + v_1^2 + |{v_1}{|^2})}}{2}}}\;{e^{\frac{{ - ({{\tilde{y}}^2} - 2\sqrt 2 \;{v_2}\;\tilde{y} + v_2^2 + |{v_2}{|^2})}}{2}}}\\ &\textrm{ } = {e^{ - (|{u_{\,1\,}}{|^2} + |{u_{\,2}}{|^2})/2}}\textrm{ }\sum\limits_{{n_1} = 0}^\infty {\;\sum\limits_{{n_2} = 0}^\infty {\;\frac{{{u_1}^{{n_{\,1}}}}}{{\sqrt {{n_1}\,!} }}\;\frac{{{u_2}^{{n_{\,2}}}}}{{\sqrt {{n_2}\,!} }}\psi _{{n_{\,1}},{n_{\,2}}}^{(\hat{H})}(\tilde{x},\tilde{y})} } \end{aligned}. $$

From Eqs. (8) and (9), the classical trajectory can be found to only rely on the relative phase between ϕ1 and ϕ2. Therefore, it is concisely convenient to express the parameters u1 and u2 as

$$\left[ {\begin{array}{c} {{u_1}}\\ {{u_2}} \end{array}} \right] = \left[ {\begin{array}{c} {\sqrt {{N_1}} \;{e^{ - iq(\theta + \phi /2)}}}\\ {\sqrt {{N_2}} \;{e^{ - ip(\theta - \phi /2)}}} \end{array}} \right], $$
where the variable $\theta = \omega \,t$ is in the range of $0 \le \theta \le 2\pi$ for representing a periodic orbit and the phase factor ϕ is related to the orbital structure. By using Eqs. (8) and (25), the parameters v1 and v2 can be expressed as
$$\left[ \begin{array}{l} {v_1}\\ {v_2} \end{array} \right] = \left[ {\begin{array}{cc} {{e^{ - i\alpha /2}}\cos (\beta /2)}&{ - {e^{ - i\alpha /2}}\sin (\beta /2)}\\ {{e^{i\alpha /2}}\sin (\beta /2)}&{{e^{i\alpha /2}}\cos (\beta /2)} \end{array}} \right]\;\left[ {\begin{array}{c} {\sqrt {{N_1}} \;{e^{ - iq(\theta + \phi /2)}}}\\ {\sqrt {{N_2}} \;{e^{ - ip(\theta - \phi /2)}}} \end{array}} \right]. $$

Substituting Eq. (25) into Eq. (24), the summation terms in the wave packet state $g(\tilde{x},{v_1})g(\tilde{y},\;{v_2})$ can be rearranged as

$$\begin{aligned} g(\tilde{x},{v_1})g(\tilde{y},\;{v_2}) &= \frac{1}{{\sqrt \pi }}{e^{\frac{{ - ({{\tilde{x}}^2} - 2\sqrt {2\,} {v_1}\,\tilde{x} + v_1^2 + |{v_1}{|^2})}}{2}}}\;{e^{\frac{{ - ({{\tilde{y}}^2} - 2\sqrt 2 \;{v_2}\;\tilde{y} + v_2^2 + |{v_2}{|^2})}}{2}}}\\ &\textrm{ } = {e^{ - \frac{{{N_1} + {N_2}}}{2}}}\,\sum\limits_{{n_1} = 0}^\infty \; \sum\limits_{{n_2} = 0}^\infty {\frac{{{N_1}^{{n_1}/2}}}{{\sqrt {{n_1}\,!} }}\frac{{{N_2}^{{n_2}/2}}}{{\sqrt {{n_2}\,!} }}\psi _{{n_1},{n_2}}^{(\hat{H})}(\tilde{x},\tilde{y})\;} {e^{ - i(q{n_1} + p{n_2})\theta }}\;{e^{ - i(q{n_1} - p{n_2})\phi /2}} \end{aligned}. $$

From ${\omega _1} = q\omega$, ${\omega _2} = p\omega$ and Eq. (13), the eigenvalues can be found to mainly depend on the factor $q{n_1} + p{n_2}$. For a given $({N_1},{N_2})$, the eigenstates $\psi _{{n_1},{n_2}}^{(\hat{H})}(\tilde{x},\tilde{y})$ with $({n_1},{n_2})$ meeting $q{n_1} + p{n_2} = q{N_1} + p{N_2}$ can be therefore confirmed to be degenerate. As a consequence, the time-dependent wave packet $g(\tilde{x},{v_1})g(\tilde{y},\;{v_2})$ in Eq. (27) can be expressed as the superposition of the stationary coherent states with the eigenvalues to be related to $q{n_1} + p{n_2}$.

The representation for the stationary coherent states can be obtained by using the Fourier transform. The kernel of the Fourier transform is given by

$$\frac{1}{{2\pi }}\int_{\;0}^{2\pi } {{e^{i(n - n^{\prime})\theta }}d\theta } = {\delta _{n,n^{\prime}}}$$

Applying Eq. (28) to Eq. (27), the stationary coherent state for a given $({N_1},{N_2})$ can be derived from

$$\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi ) = \frac{1}{{2\pi }}\int_0^{2\pi } {g(\tilde{x},{v_1})g(\tilde{y},\;{v_2})\;{e^{i\,(q{N_1} + p{N_2}\,)\theta }}d\theta }. $$

For a given $({N_1},{N_2})$, the group for the degenerate eigenstates $\psi _{{n_1},{n_2}}^{(\hat{H})}(\tilde{x},\tilde{y})$ can be expressed as ${n_1} = {N_1} - pK$ and ${n_2} = {N_2} + qK$, where K is an arbitrary integer. The stationary coherent state $\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )$ in Eq. (29) can therefore be derived as

$$\begin{array}{l} \Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )\\ = \frac{1}{{2\pi }}\int_0^{2\pi } {\frac{1}{{\sqrt \pi }}{e^{\frac{{ - ({{\tilde{x}}^2} - 2\sqrt {2\,} {v_1}\,\tilde{x} + v_1^2 + |{v_1}{|^2})}}{2}}}\;{e^{\frac{{ - ({{\tilde{y}}^2} - 2\sqrt 2 \;{v_2}\;\tilde{y} + v_2^2 + |{v_2}{|^2})}}{2}}}{e^{i\,(q{N_1} + p{N_2}\,)\theta }}d\theta } \\ = \;{e^{\frac{{ - i(q{N_1} - p{N_2})\,\phi }}{2}}}{e^{\frac{{ - ({N_1} + {N_2})}}{2}}}\,\sum\limits_{K ={-} [{N_2}/q]}^{[{N_1}/p]} \; \frac{{{N_1}^{\frac{{({N_1} - pK)}}{2}}\,{N_2}^{\frac{{({N_2} + qK)}}{2}}}}{{\sqrt {({N_1} - pK)\,!} \sqrt {({N_2} + qK)\,!} }}{e^{i\,\,q\,p\,K\,\phi }}\;\psi _{{N_1} - pK,{N_2} + qK}^{(\hat{H})}(\tilde{x},\tilde{y})\; \end{array}. $$

Equation (30) reveals that the total number of the degenerate eigenstates $\psi _{{N_1} - pK,{N_2} + qK}^{(\hat{H})}(\tilde{x},\tilde{y})$ for a given $({N_1},{N_2},\phi )$ in the stationary coherent state $\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )$ is ${N_t} = [{N_1}/p] + [{N_2}/q] + 1$. For ${N_t} > > 1$, the spatial distributions ${|{\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )} |^2}$ can be confirmed to be well localized on the classical orbits. The summation expression in Eq. (30) can be associated with the representation of the SU(2) coherent state that has been used to analyze quantum wave functions corresponding to classical orbits [37,38]. Furthermore, Eq. (30) also reveals how the stationary coherent state to be generated from the integral evolution of the time-dependent wave packet state. The integral evolution was originally proposed for numerically analyzing the nonintegrable system [53,54]. We are the first group to develop the integral evolution for dealing with the 2D and higher-dimensional coupled oscillator systems. More importantly, we first exploited the SU(2) coupled oscillator model to manifest the laser transverse modes covering the HLG and geometric modes [3945,30,4648].

4. Results and discussion

First of all, we demonstrate the case with $\alpha = 0$ and $\beta = 0$ that corresponds to the 2D anisotropic harmonic oscillator. Figure 1 shows the calculated patterns of the stationary coherent states $\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )$ in Eq. (30) with the parameters of $(q,p) = (5,4)$, and $\phi = 0$ for different orders of $({N_1},{N_2}) = (Mp,Mq)$ with the factor M = 1, 2, 3, 8, and 10. As discussed above, the physical meaning of the parameter M can be related to the total number ${N_t}$ of eigenmodes participating in the coherent state $\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )$ as ${N_t} = 2M + 1$. On the other hand, the case of $(q,p) = (5,4)$ is used here just for illustrative example. Consequently, the factor M is used here to display the transition of the wave patterns ${|{\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )} |^2}$ from low-level to high-level states. The wave pattern for M = 1 can be seen to be similar to the HG eigenmodes of the 2D anisotropic harmonic oscillator. To be more specific, when the indices $({N_1},{N_2})$ are small and satisfy ${N_1} < p$ and ${N_2} < q$, the stationary coherent states $\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )$ represent pure single eigenstates and their wave patterns are independent of the phase factor ϕ since no degenerate states are in the low-level energies. With increasing the factor M, it can be seen that more and more degenerate eigenstates participate in the stationary coherent states $\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )$ to make the wave patterns gradually change from the widespread distributions of HG eigenstates to the concerned structures corresponding to classical orbits. As expected, the wave patterns are localized on the Lissajous figures that are the classical periodic orbits for the cases with $\alpha = 0$ and $\beta = 0$.

 figure: Fig. 1.

Fig. 1. Calculated patterns of the stationary coherent states ${|{\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )} |^2}$ in Eq. (30) with the parameters of $(q,p) = (5,4)$, $\alpha = 0$, $\beta = 0$, and $\phi = 0$ for different orders of $({N_1},{N_2}) = (Mp,Mq)$ with the factor M = 1, 2, 3, 8, and 10.

Download Full Size | PDF

Next, we discuss the case with $\alpha = \pi /2$ and $\beta = \pi /2$ that corresponds to the problem of a charged particle moving in a 2D isotropic harmonic oscillator potential in the xy plane and subjected to a constant external magnetic field along the z axis. Figure 2 shows the calculated results of the stationary coherent states $\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )$ with the parameters of $(q,p) = (8,1)$, and $\phi = 0$ for different orders of $({N_1},{N_2}) = (Mp,Mq)$ with the factor M = 1, 2, 3, 8, and 10. The wave pattern for M = 1 can be seen to be similar to the LG eigenmodes in the 2D isotropic harmonic oscillator. As shown in Fig. 1 for the case with $\alpha = 0$ and $\beta = 0$, the stationary coherent state $\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )$ gradually transits from the extensively spatial characteristics of LG eigenstates to the localized structures with intensities concentrated on classical orbits with increasing the index order $({N_1},{N_2})$. The wave patterns are localized on the trochoidal curves that are the classical periodic trajectories for the cases with $\alpha = \pi /2$ and $\beta = \pi /2$.

 figure: Fig. 2.

Fig. 2. Calculated results of the stationary coherent states ${|{\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )} |^2}$ with the parameters of $(q,p) = (8,1)$, $\alpha = \pi /2$, $\beta = \pi /2$, and $\phi = 0$ for different orders of $({N_1},{N_2}) = (Mp,Mq)$ with the factor M = 1, 2, 3, 8, and 10.

Download Full Size | PDF

This result indicates that in addition to the order $({N_1},{N_2})$, the mean orbital angular momentum $\left\langle {{L_z}} \right\rangle$ depends on the coupling parameters α and β. The mean orbital angular momentum $\left\langle {{L_z}} \right\rangle$ can reach maximum for the cases with $\alpha = \pi /2$ and $\beta = \pi /2$, as discussed in Fig. 2. On the other hand, the mean orbital angular momentum $\left\langle {{L_z}} \right\rangle$ is zero for the cases with $\alpha = 0$ or $\beta = 0$.

Figure 3 shows the variation of the wave patterns ${|{\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )} |^2}$ with the parameter β for a fixed $\alpha = \pi /2$ for the low-level states without degeneracy, where $(q,p) = (9,8)$ and $({N_1},{N_2}) = (2,3)$, $\phi = 0$. The overall feature of the wave patterns is clear to change gradually from rectangular HG mode to circular LG mode with increasing β from 0 to π/2. The second row of Fig. 3 shows the numerical results for the phase structures calculated by $\Theta (\tilde{x},\tilde{y}) = {\tan ^{ - 1}}\{ {\mathop{\rm Im}\nolimits} [\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )]/Re [\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )]\}$. It can be seen that a 2D array of isolated singularities appears when the parameter β increases from zero. The isolated singularities finally evolve into a single vortex when the parameter β is up to $\pi /2$. Figure 4 shows the dependence of the wave patterns ${|{\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )} |^2}$ on the parameter β for a fixed $\alpha = \pi /2$ for the high-level states with accidental degeneracies, where the values of other parameters used in the calculation are $(q,p) = (4,1)$ and $({N_1},{N_2}) = (6,16)$, and $\phi ={-} \pi$. The localized structures of the wave patterns ${|{\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )} |^2}$ can be seen to transform continuously from Lissajous to trochoidal orbits with increasing β from 0 to π /2. At the second row of Fig. 4, numerical results for the phase structures $\Theta (\tilde{x},\tilde{y})$ are shown to manifest the formation and variation of isolated singularities.

 figure: Fig. 3.

Fig. 3. First row: variation of the wave patterns ${|{\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )} |^2}$ with the parameter β for a fixed $\alpha = \pi /2$ for the low-level states without degeneracy, where $(q,p) = (9,8)$ and $({N_1},{N_2}) = (2,3)$, and $\phi = 0$. Second row: numerical results for the phase structures.

Download Full Size | PDF

 figure: Fig. 4.

Fig. 4. First row: dependence of the wave patterns ${|{\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )} |^2}$ on the parameter β for a fixed $\alpha = \pi /2$ for the high-level states with accidental degeneracies, where the values of other parameters used in the calculation are $(q,p) = (4,1)$ and $({N_1},{N_2}) = (6,16)$, and $\phi ={-} \pi$. Second row: numerical results for the phase structures.

Download Full Size | PDF

The stationary coherent states in the SU(2) coupled oscillator can be used to analyze the structured optical waves that can be experimentally generated from laser transverse modes with astigmatic transformation. Abramochkin and Volostnikov [55] originally used an AMC formed by a matched pair cylindrical lenses to generate the so-called HLG beams. The HLG beam, a continuous evolution between Hermite-Gaussian and Laguerre-Gaussian beams, can successively be realized by rotating the AMC about the optical axis by an angle η from 0 to π/4. The HLG modes display plentiful evolutions of point dislocations and edge dislocations. The AMC was later used to generate the light fields with nonzero orbital angular momentum [56]. Here we demonstrate that the stationary coherent state ${\Psi _{{N_{\,1}},{N_{\,2}}}}(\tilde{x},\tilde{y})$ with a fixed $\alpha = \pi /2$ for the indices $({N_1},{N_2})$ satisfying ${N_1} \le p$ and ${N_2} \le q$ can be used to analyze the HLG modes which are generated by using an AMC to transform the Hermite-Gaussian mode. Figure 5 shows the case with $({N_1},{N_2}) = (15,5)$ for demonstration, where the experimental patterns are obtained by using a diode-pumped solid-state laser with off-axis pumping [57,58] to generate the Hermite-Gaussian mode that is subsequently converted by an AMC with different angle η. The theoretical patterns ${|{{\Psi _{{N_{\,1}},{N_{\,2}}}}(\tilde{x},\tilde{y})} |^2}$ are calculated with $q = 7$ and $p = 16$ for different β. The one-to-one correspondence can be seen to be given by $\beta \textrm{ = 2}\eta$. As mentioned above, the theoretical wave patterns are independent of the phase factor ϕ because they are pure eigenstates without degeneracy. The third row of Fig. 5 shows numerical results for the phase structures. The formation and variation of isolated singularities can be clearly seen to vary from rectangular to circular distributions.

 figure: Fig. 5.

Fig. 5. First row: experimental transverse patterns obtained by using a diode-pumped solid-state laser with an external AMC with different angle η. Second row: theoretical patterns ${|{\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )} |^2}$ calculated with $({N_1},{N_2}) = (15,5)$, $q = 7$ and $p = 16$ for different β = 2η. Third row: numerical results for the phase structures.

Download Full Size | PDF

In addition to HG modes, it has been experimentally observed that the tiny symmetry breaking can lead to the three-dimensional (3D) high-order laser modes with the transverse patterns to be well concentrated on the 2D Lissajous figures. The transverse wave functions of these laser modes can be described by the high-level stationary coherent state ${\Psi _{{N_{\,1}},{N_{\,2}}}}(\tilde{x},\tilde{y})$ with $\alpha = \pi /2$ and $\beta = 0$. As the generation of HLG modes by using an AMC, the transverse Lissajous pattern can be continuously transformed to the trochoidal pattern by passing through an AMC with the angle η changing from 0 to π /4. The one-to-one correspondence between the transformed laser patterns and the stationary coherent state ${\Psi _{{N_{\,1}},{N_{\,2}}}}(\tilde{x},\tilde{y})$ can be obtained with $\beta \textrm{ = 2}\eta$. Figure 6 shows the case with $(q,p) = (4,1)$ for demonstration, where the theoretical patterns $|{{\Psi _{{N_{\,1}},{N_{\,2}}}}(\tilde{x},\tilde{y})} |$ are calculated with $({N_1},{N_2}) = (30,80)$, and $\phi ={-} \pi$. Numerical results for the phase structures are also shown at the third row of Fig. 6. Once again, the theoretical patterns are in excellent agreement with the experimental results.

 figure: Fig. 6.

Fig. 6. First row: experimental transverse patterns obtained by using a diode-pumped solid-state laser with an external AMC with different angle η. Second row: theoretical patterns ${|{\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )} |^2}$ calculated with $(q,p) = (4,1)$, $({N_1},{N_2}) = (30,80)$, and $\phi ={-} \pi$ for different β = 2η. Third row: numerical results for the phase structures.

Download Full Size | PDF

5. Conclusion

In summary, we have systematically reviewed the SU(2) coupled oscillator model to deliver the quantum eigenstates and stationary coherent states for characterizing laser transverse modes in an analogous way based on the quantum-classical connection. We further overviewed the integral formula for the representation of the stationary coherent states derived from the evolution of the time-dependent wave packet state. Several examples for the stationary coherent states are illustratively presented to display the spatial distributions and phase singularities for the quantum-classical transition. It is believed that the generalized integral method derived from the quantum SU(2) coupled oscillator model can provide a comprehensive understanding for the laser transverse modes related to the transition between the HLG eigenmode and geometric modes. Finally, it is worthwhile to mention that the quantum SU(2) states are related to the attractive oscillator system, whereas the quantum SU(1,1) states are associated with the repulsive oscillator system. Since the transverse modes of the laser cavity are analogous to the states of the confined system, all experimentally observed laser modes so far are linked to the SU(2) states instead of SU(1,1) states.

Funding

Ministry of Science and Technology, Taiwan (109-2119-M-009-015-MY3).

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. M. Brack and R. K. Bhaduri, Semiclassical Physics (Addison Wesley, 1997).

2. J. A. Yeazell, M. Mallalieu, J. Parker, and C. R. Stroud, “Classical periodic motion atomic-electron wave packets,” Phys. Rev. A 40(9), 5040–5043 (1989). [CrossRef]  

3. C. Lena, D. Delande, and J. C. Gay, “Wave Functions of Atomic Elliptic States,” Europhys. Lett. 15(7), 697–702 (1991). [CrossRef]  

4. B. L. Johnson and G. Kirczenow, “Enhanced dynamical symmetries and quantum degeneracies in mesoscopic quantum dots: role of the symmetries of closed classical orbits,” Europhys. Lett. 51(4), 367–373 (2000). [CrossRef]  

5. A. V. Madhav and T. Chakraborty, “Electronic properties of anisotropic quantum dots in a magnetic field,” Phys. Rev. B 49(12), 8163–8168 (1994). [CrossRef]  

6. S. Avetisyan, P. Pietilainen, and T. Chakraborty, “Strong enhancement of Rashba spin-orbit coupling with increasing anisotropy in the Fock-Darwin states of a quantum dot,” Phys. Rev. B 85(15), 153301 (2012). [CrossRef]  

7. K. F. Huang, Y. F. Chen, H. C. Lai, and Y. P. Lan, “Observation of the wave function of a quantum billiard from the transverse patterns of vertical cavity surface emitting lasers,” Phys. Rev. Lett. 89(22), 224102 (2002). [CrossRef]  

8. Y. F. Chen, T. H. Lu, K. W. Su, and K. F. Huang, “Devil’s staircase in three dimensional coherent waves localized on Lissajous parametric surfaces,” Phys. Rev. Lett. 96(21), 213902 (2006). [CrossRef]  

9. T. H. Lu, Y. C. Lin, Y. F. Chen, and K. F. Huang, “Three-dimensional coherent optical waves localized on trochoidal parametric surfaces,” Phys. Rev. Lett. 101(23), 233901 (2008). [CrossRef]  

10. Y. F. Chen, J. C. Tung, P. Y. Chiang, H. C. Liang, and K. F. Huang, “Exploring the effect of fractional degeneracy and the emergence of ray-wave duality in solid-state lasers with off-axis pumping,” Phys. Rev. A 88(1), 013827 (2013). [CrossRef]  

11. J. C. Tung, P. H. Tuan, H. C. Liang, K. F. Huang, and Y. F. Chen, “Fractal frequency spectrum in laser resonators and three-dimensional geometric topology of optical coherent waves,” Phys. Rev. A 94(2), 023811 (2016). [CrossRef]  

12. P. H. Tuan, C. P. Wen, Y. T. Yu, P. Y. Chiang, H. C. Liang, K. F. Huang, and Y. F. Chen, “Exploring the resonant vibration of thin plates: Reconstruction of Chladni patterns and determination of resonant wave numbers,” J. Acoust. Soc. Am. 137(4), 2113–2123 (2015). [CrossRef]  

13. P. H. Tuan, J. C. Tung, H. C. Liang, P. Y. Chiang, K. F. Huang, and Y. F. Chen, “Resolving the formation of modern Chladni figures,” Europhys. Lett. 111(6), 64004 (2015). [CrossRef]  

14. M. Moshinsky, J. Patera, and P. Winternitz, “Canonical transformation and accidental degeneracy. III. A unified approach to the problem,” J. Math. Phys. 16(1), 82 (1975). [CrossRef]  

15. M. Moshinsky and C. Quesne, “Does accidental degeneracy imply a symmetry group?” Ann. Phys. 148(2), 462–488 (1983). [CrossRef]  

16. M. Moshinsky, N. Mendez, E. Murow, and J. W. B. Hughes, “Accidental degeneracies in the zeeman effect and the symmetry groups,” Ann. Phys. 155(2), 231–268 (1984). [CrossRef]  

17. C. Quesne, “Accidental degeneracies and symmetry group of the harmonic oscillator in a strong magnetic field,” J. Phys. A: Math. Gen. 19(7), 1127–1139 (1986). [CrossRef]  

18. E. Drigho-Filho, Ş. Kuru, J. Negro, and L.M. Nieto, “Superintegrability of the Fock–Darwin system,” Ann. Phys. 383, 101–119 (2017). [CrossRef]  

19. J. Masajada and B. Dubik, “Optical vortex generation by three plane wave interference,” Opt. Commun. 198(1-3), 21–27 (2001). [CrossRef]  

20. M. S. Soskin and M. V. Vasnetsov, “Singular optics,” Prog. Opt. 42, 219–276 (2001). [CrossRef]  

21. J. F. Nye and M. V. Berry, “Dislocations in wave trains,” Proc. Roy. Soc. A 336, 165 (1974). [CrossRef]  

22. J. O. Hirschfelder, A. C. Christoph, and W. E. Palke, “Quantum mechanical streamlines. I. Square potential barrier,” J. Chem. Phys. 61(12), 5435–5455 (1974). [CrossRef]  

23. J. O. Hirschfelder, C. J. Goebel, and L. W. Bruch, “Quantized vortices around wavefunction nodes. II,” J. Chem. Phys. 61(12), 5456–5459 (1974). [CrossRef]  

24. I. Bialynicki-Birula, Z. Bialynicka-Birula, and C. Sliwa, “Motion of vortex lines in quantum mechanics motion of vortex lines in quantum mechanics,” Phys. Rev. A 61(3), 032110 (2000). [CrossRef]  

25. L. J. Allen, H. M. L. Faulkner, M. P. Oxley, and D. Paganin, “Phase retrieval and aberration correction in the presence of vortices in high-resolution transmission electron microscopy,” Ultramicroscopy 88(2), 85–97 (2001). [CrossRef]  

26. K. Y. Bliokh, Y. P. Bliokh, S. Savel’ev, and F. Nori, “Semiclassical dynamics of electron wave packet states with phase vortices,” Phys. Rev. Lett. 99(19), 190404 (2007). [CrossRef]  

27. M. Uchida and A. Tonomura, “Generation of electron beams carrying orbital angular momentum,” Nature 464(7289), 737–739 (2010). [CrossRef]  

28. J. Verbeeck, H. Tian, and P. Schattschneider, “Production and application of electron vortex beams,” Nature 467(7313), 301–304 (2010). [CrossRef]  

29. B. J. McMorran, A. Agrawal, I. M. Anderson, A. A. Herzing, H. J. Lezec, J. J. McClelland, and J. Unguris, “Electron vortex beams with high quanta of orbital angular momentum,” Science 331(6014), 192–195 (2011). [CrossRef]  

30. Y. F. Chen, C. C. Lee, C. H. Wang, and M. X. Hsieh, “Laser transverse modes of spherical resonators: a review,” Chin. Opt. Lett. 18(9), 091404 (2020). [CrossRef]  

31. D. Ran, C. Hu, and Z. Yang, “Entanglement transfer from two-mode continuous variable SU(2) cat states to discrete qubits systems in Jaynes-Cummings Dimers,” Sci. Rep. 6(1), 32089 (2016). [CrossRef]  

32. J. Banerji and G. S. Agarwal, “Non-linear wave packet dynamics of coherent states of various symmetry groups,” Opt. Express 5(10), 220 (1999). [CrossRef]  

33. G S Agarwal and J Banerji, “Entanglement by linear SU (2) transformations: generation and evolution of quantum vortex states,” J. Phys. A: Math. Gen. 39(37), 11503–11519 (2006). [CrossRef]  

34. R. G. Nazmitdinov, “Magnetic field and symmetry effects in small quantum dots,” Phys. Part. Nucl. 40(1), 71–92 (2009). [CrossRef]  

35. P. K. Bhaduri, S. Li, K. Tanaka, and J. C. Waddington, “Quantum gaps and classical orbits in a rotating two-dimensional harmonic oscillator,” J. Phys. A: Math. Gen. 27(15), L553–L558 (1994). [CrossRef]  

36. W. D. Heiss and R. G. Nazmitdinov, “Classical analysis of phenomenological potentials for metallic clusters,” Phys. Rev. Lett. 73(9), 1235–1238 (1994). [CrossRef]  

37. Y. F. Chen, “Geometry of classical periodic orbits and quantum coherent states in coupled oscillators with SU(2) transformations,” Phys. Rev. A 83(3), 032124 (2011). [CrossRef]  

38. Y. F. Chen, J. C. Tung, P. H. Tuan, and K. F. Huang, “Symmetry Breaking Induced Geometric Surfaces with Topological Curves in Quantum and Classical Dynamics of the SU (2) Coupled Oscillators,” Ann. Phys. 529(10), 1600253 (2017). [CrossRef]  

39. Y. F. Chen, Y. H. Hsieh, and K. F. Huang, “Originating an integral formula and using the quantum Fourier transform to decompose the Hermite-Laguerre-Gaussian modes into elliptical orbital modes,” OSA Continuum 1(2), 744–754 (2018). [CrossRef]  

40. Y. F. Chen, Y. H. Hsieh, J. C. Tung, H. C. Liang, and K. F. Huang, “Manifesting the connection between topological structures of quantum stationary coherent states and bundles of classical Lissajous orbits,” Europhys. Lett. 122(3), 30002 (2018). [CrossRef]  

41. Y. F. Chen, M. X. Hsieh, Y. H. Lai, Y. H. Hsieh, C. W. Tu, H. C. Liang, and K. F. Huang, “Exploring elliptical vortex beams with the spatial damping effect,” Opt. Lett. 44(4), 959–962 (2019). [CrossRef]  

42. Y. F. Chen, Y. H. Lai, M. X. Hsieh, Y. H. Hsieh, C. W. Tu, H. C. Liang, and K. F. Huang, “Wave representation for asymmetric elliptic vortex beams generated from the astigmatic mode converter,” Opt. Lett. 44(8), 2028–2031 (2019). [CrossRef]  

43. Y. F. Chen, S. C. Li, Y. H. Hsieh, J. C. Tung, H. C. Liang, and K. F. Huang, “Laser wave-packet representation to unify eigenmodes and geometric modes in spherical cavities,” Opt. Lett. 44(11), 2649–2652 (2019). [CrossRef]  

44. Y. F. Chen, C. J. Tung, M. X. Hsieh, Y. H. Hsieh, H. C. Liang, and K. F. Huang, “Generalized wave-packet formulation with ray-wave connections for geometric modes in degenerate astigmatic laser resonators,” Opt. Lett. 44(21), 5366–5369 (2019). [CrossRef]  

45. Y. F. Chen, C. J. Tung, M. X. Hsieh, Y. H. Hsieh, H. C. Liang, and K. F. Huang, “Origin of continuous curves and dotted spots in laser transverse modes with geometric structures,” Opt. Lett. 44(24), 5989–5992 (2019). [CrossRef]  

46. Y. F. Chen, C. H. Wang, X. L. Zheng, and M. X. Hsieh, “Laser transverse modes with ray-wave duality: a review,” Appl. Sci. 11(19), 8913 (2021). [CrossRef]  

47. X. L. Zheng, M. X. Hsieh, and Y. F. Chen, “Quantifying the emergence of structured laser beams relevant to Lissajous parametric surfaces,” Opt. Lett. 47(10), 2518–2521 (2022). [CrossRef]  

48. Y. T. Yu, M. X. Hsieh, H. C. Liang, and Y. F. Chen, “Characterizing propagation-dependent spatial entanglement for structured laser beams generated by an astigmatic mode converter,” Opt. Lett. 47(13), 3223–3226 (2022). [CrossRef]  

49. J. Schwinger, in Quantum Theory of Angular Momentum, L. C. Biedenharn and H. van Dam, eds., (Academic Press, 1965).

50. N. Zettili, Quantum Mechanics: Concepts and Applications, 2nd edn. (John Wiley & Sons, 2009).

51. E. P. Wigner, Group Theory, (Academic Press, 1959).

52. N. N. Lebedev, Special Functions & Their Applications (Dover, 1972).

53. M. J. Davis and E. J. Heller, “Multidimensional wave functions from classical trajectories,” J. Chem. Phys. 75(8), 3916–3924 (1981). [CrossRef]  

54. N. DeLeon, M. J. Davis, and E. J. Heller, “Quantum manifestations of classical resonance zones,” J. Chem. Phys. 80(2), 794 (1984). [CrossRef]  

55. E. G. Abramochkin and V. G. Volostnikov, “Beam transformations and nontransformed beams,” Opt. Commun. 83(1-2), 123 (1991). [CrossRef]  

56. M. W. Beijersbergen, L. Allen, H. E. L. O. van der Veen, and J. P. Woerdman, “Astigmatic laser mode converters and transfer of orbital angular momentum,” Opt. Commun. 96(1-3), 123–132 (1993). [CrossRef]  

57. Y. F. Chen, T. M. Huang, C. F. Kao, C. L. Wang, and S. C. Wang, “Generation of Hermite Gaussian modes in fiber-coupled laser-diode end-pumped lasers,” IEEE J. Quantum Electron. 33(6), 1025–1031 (1997). [CrossRef]  

58. H. Laabs and B. Ozygus, “Excitation of Hermite-Gaussian modes in end-pumped solid-state lasers via off-axis pumping,” Opt. Laser Technol. 28(3), 213–214 (1996). [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. Calculated patterns of the stationary coherent states ${|{\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )} |^2}$ in Eq. (30) with the parameters of $(q,p) = (5,4)$, $\alpha = 0$, $\beta = 0$, and $\phi = 0$ for different orders of $({N_1},{N_2}) = (Mp,Mq)$ with the factor M = 1, 2, 3, 8, and 10.
Fig. 2.
Fig. 2. Calculated results of the stationary coherent states ${|{\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )} |^2}$ with the parameters of $(q,p) = (8,1)$, $\alpha = \pi /2$, $\beta = \pi /2$, and $\phi = 0$ for different orders of $({N_1},{N_2}) = (Mp,Mq)$ with the factor M = 1, 2, 3, 8, and 10.
Fig. 3.
Fig. 3. First row: variation of the wave patterns ${|{\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )} |^2}$ with the parameter β for a fixed $\alpha = \pi /2$ for the low-level states without degeneracy, where $(q,p) = (9,8)$ and $({N_1},{N_2}) = (2,3)$, and $\phi = 0$. Second row: numerical results for the phase structures.
Fig. 4.
Fig. 4. First row: dependence of the wave patterns ${|{\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )} |^2}$ on the parameter β for a fixed $\alpha = \pi /2$ for the high-level states with accidental degeneracies, where the values of other parameters used in the calculation are $(q,p) = (4,1)$ and $({N_1},{N_2}) = (6,16)$, and $\phi ={-} \pi$. Second row: numerical results for the phase structures.
Fig. 5.
Fig. 5. First row: experimental transverse patterns obtained by using a diode-pumped solid-state laser with an external AMC with different angle η. Second row: theoretical patterns ${|{\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )} |^2}$ calculated with $({N_1},{N_2}) = (15,5)$, $q = 7$ and $p = 16$ for different β = 2η. Third row: numerical results for the phase structures.
Fig. 6.
Fig. 6. First row: experimental transverse patterns obtained by using a diode-pumped solid-state laser with an external AMC with different angle η. Second row: theoretical patterns ${|{\Psi _{{N_1},{N_2}}^{(q,p)}(\tilde{x},\tilde{y},\phi )} |^2}$ calculated with $(q,p) = (4,1)$, $({N_1},{N_2}) = (30,80)$, and $\phi ={-} \pi$ for different β = 2η. Third row: numerical results for the phase structures.

Equations (30)

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

H ^ = H ^ o + i = 1 3 Ω i L ^ i ,
H ^ o = ( a 1 a 1 + a 2 a 2 + 1 ) ω o ,
L ^ 1 = 1 / 2 ( a 1 a 2 + a 2 a 1 ) ,
L ^ 2 = i / 2 ( a 1 a 2 a 2 a 1 ) ,
L ^ 3 = 1 / 2 ( a 1 a 1 a 2 a 2 ) ,
l H = ω o 2 ( x ~ 2 + p ~ x 2 + y ~ 2 + p ~ y 2 ) + Ω 1 2 ( x ~ y ~ + p ~ x p ~ y )   + Ω 2 2 ( x ~ p ~ y y ~ p ~ x ) + Ω 3 4 ( x ~ 2 + p ~ x 2 y ~ 2 p ~ y 2 ) .
d d t [ v 1 v 2 ] = i [ ω o + ( Ω 3 / 2 ) ( Ω 1 i Ω 2 ) / 2 ( Ω 1 + i Ω 2 ) / 2 ω o ( Ω 3 / 2 ) ] [ v 1 v 2 ] ,
[ v 1 ( t ) v 2 ( t ) ] = [ e i α / 2 cos ( β / 2 ) e i α / 2 sin ( β / 2 ) e i α / 2 sin ( β / 2 ) e i α / 2 cos ( β / 2 ) ] [ u 1 ( t ) u 2 ( t ) ] ,
[ x ~ ( t ) y ~ ( t ) ] = [ 2 N 1 cos ( β 2 ) cos ( q ω t + α 2 ϕ 1 ) 2 N 2 sin ( β 2 ) cos ( p ω t + α 2 ϕ 2 ) 2 N 1 sin ( β 2 ) cos ( q ω t α 2 ϕ 1 ) + 2 N 2 cos ( β 2 ) cos ( p ω t α 2 ϕ 2 ) ] .
[ b 1 b 2 ] = [ e i α / 2 cos ( β / 2 ) e i α / 2 sin ( β / 2 ) e i α / 2 sin ( β / 2 ) e i α / 2 cos ( β / 2 ) ] [ a 1 a 2 ] ,
H ^ = ( b 1 b 1 + 1 / 2 ) ω 1 + ( b 2 b 2 + 1 / 2 ) ω 2 .
| n 1 , n 2 H ^ = ( b 1 ) n 1 n 1 ! ( b 2 ) n 2 n 2 ! | 0 , 0 H ^ ,
E n 1 , n 2 = ( n 1 + 1 / 2 ) ω 1 + ( n 2 + 1 / 2 ) ω 2 .
| n 1 , n 2 H ^ = e i N α / 2 m 1 = 0 N e i m 1 α d m 1 N 2 , n 1 N 2 N 2 ( β ) | m 1 , m 2 H ^ o ,
d m 1 N 2 , n 1 N 2 N 2 ( β ) = m 1   ! ( N m 1 )   ! n 1   ! ( N n 1 )   !   × ν = max [ 0 , m 1 n 1 ] min [ N n 1 , m 1 ]   ( 1 ) ν [ cos ( β / 2 ) ] N n 1 + m 1 2 ν [ sin ( β / 2 ) ] n 1 m 1 + 2 ν ν   ! ( N n 1 ν )  ! ( m 1  -  ν ) !( n 1 m 1 + ν ) ! ,
ψ n 1 , n 2 ( H ^ ) ( x ~ , y ~ ) = x ~ , y ~ | n 1 , n 2 H ^ = [ e i N α / 2 m 1 = 0 N e i m 1 α d m 1 N 2 , n 1 N 2 N 2 ( β ) ψ m 1 ( x ~ ) ψ N m 1 ( y ~ ) ] .
ψ n ( x ~ ) = 1 2 n n ! π H n ( x ~ ) e x ~ 2 / 2 ,
e τ 2 + 2 τ x ~ = n = 0 τ n n ! H n ( x ~ ) .
g ( x ~ ; u ) = 1 π 1 / 4 exp [ ( x ~ 2 2 2 u x ~ + u 2 + | u | 2 ) 2 ] = n = 0 u n n ! e | u | 2 / 2 ψ n ( x ~ ) .
g ( x ~ ; u ) = 1 π 1 / 4 e i x ~ 2 Im ( u ) e i R e ( u ) Im ( u ) e x p { [ x ~ 2 R e ( u ) ] 2 2 } .
g ( x ~ , v 1 ) g ( y ~ , v 2 ) = 1 π e ( x ~ 2 2 2 v 1 x ~ + v 1 2 + | v 1 | 2 ) 2 e ( y ~ 2 2 2 v 2 y ~ + v 2 2 + | v 2 | 2 ) 2 = e ( | v 1 | 2 + | v 2 | 2 ) / 2 m 1 = 0 m 2 = 0 v 1 m 1 m 1 ! v 2 m 2 m 2 ! ψ m 1 ( x ~ ) ψ m 2 ( y ~ )
g ( x ~ , v 1 ) g ( y ~ , v 2 ) = e ( v 1 | 2 + | v 2 | 2 ) / 2 m 1 = 0 m 2 = 0 ( v 1 a 1 ) m 1 m 1 ! ( v 2 a 2 ) m 2 m 2 ! ψ 0 ( x ~ ) ψ 0 ( y ~ ) .
g ( x ~ , v 1 ) g ( y ~ , v 2 ) = e ( | u 1 | 2 + | u 2 | 2 ) / 2   n 1 = 0 n 2 = 0 ( u 1 b 1 ) n 1 n 1 ! ( u 2 b 2 ) n 2 n 2 ! ψ 0 ( x ~ ) ψ 0 ( y ~ ) .
g ( x ~ , v 1 ) g ( y ~ , v 2 ) = 1 π e ( x ~ 2 2 2 v 1 x ~ + v 1 2 + | v 1 | 2 ) 2 e ( y ~ 2 2 2 v 2 y ~ + v 2 2 + | v 2 | 2 ) 2   = e ( | u 1 | 2 + | u 2 | 2 ) / 2   n 1 = 0 n 2 = 0 u 1 n 1 n 1 ! u 2 n 2 n 2 ! ψ n 1 , n 2 ( H ^ ) ( x ~ , y ~ ) .
[ u 1 u 2 ] = [ N 1 e i q ( θ + ϕ / 2 ) N 2 e i p ( θ ϕ / 2 ) ] ,
[ v 1 v 2 ] = [ e i α / 2 cos ( β / 2 ) e i α / 2 sin ( β / 2 ) e i α / 2 sin ( β / 2 ) e i α / 2 cos ( β / 2 ) ] [ N 1 e i q ( θ + ϕ / 2 ) N 2 e i p ( θ ϕ / 2 ) ] .
g ( x ~ , v 1 ) g ( y ~ , v 2 ) = 1 π e ( x ~ 2 2 2 v 1 x ~ + v 1 2 + | v 1 | 2 ) 2 e ( y ~ 2 2 2 v 2 y ~ + v 2 2 + | v 2 | 2 ) 2   = e N 1 + N 2 2 n 1 = 0 n 2 = 0 N 1 n 1 / 2 n 1 ! N 2 n 2 / 2 n 2 ! ψ n 1 , n 2 ( H ^ ) ( x ~ , y ~ ) e i ( q n 1 + p n 2 ) θ e i ( q n 1 p n 2 ) ϕ / 2 .
1 2 π 0 2 π e i ( n n ) θ d θ = δ n , n
Ψ N 1 , N 2 ( q , p ) ( x ~ , y ~ , ϕ ) = 1 2 π 0 2 π g ( x ~ , v 1 ) g ( y ~ , v 2 ) e i ( q N 1 + p N 2 ) θ d θ .
Ψ N 1 , N 2 ( q , p ) ( x ~ , y ~ , ϕ ) = 1 2 π 0 2 π 1 π e ( x ~ 2 2 2 v 1 x ~ + v 1 2 + | v 1 | 2 ) 2 e ( y ~ 2 2 2 v 2 y ~ + v 2 2 + | v 2 | 2 ) 2 e i ( q N 1 + p N 2 ) θ d θ = e i ( q N 1 p N 2 ) ϕ 2 e ( N 1 + N 2 ) 2 K = [ N 2 / q ] [ N 1 / p ] N 1 ( N 1 p K ) 2 N 2 ( N 2 + q K ) 2 ( N 1 p K ) ! ( N 2 + q K ) ! e i q p K ϕ ψ N 1 p K , N 2 + q K ( H ^ ) ( x ~ , y ~ ) .
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.