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

Phase space geometry of collective spin systems: Scaling and Fractality

Miguel Gonzalez Instituto de Ciencias Nucleares, Universidad Nacional Autónoma de México, Apdo. Postal 70-543, C.P. 04510 Cd. Mx., Mexico    Miguel A. Bastarrachea-Magnani Departamento de Física, Universidad Autónoma Metropolitana-Iztapalapa, Av. Ferrocarril San Rafael Atlixco 186, C.P. 09310 Mexico City, Mexico    Jorge G. Hirsch Instituto de Ciencias Nucleares, Universidad Nacional Autónoma de México, Apdo. Postal 70-543, C.P. 04510 Cd. Mx., Mexico
Abstract

We examine the scaling of the inverse participation ratio of spin coherent states in the energy basis of three collective spin systems: a bounded harmonic oscillator, the Lipkin-Meshkov-Glick model, and the Quantum Kicked Top. The finite-size quantum probing provides detailed insights into the structure of the phase space, particularly the relationship between critical points in classical dynamics and their quantum counterparts in collective spin systems. We introduce a finite-size scaling mass exponent that makes it possible to identify conditions under which a power-law behavior emerges, allowing to assign a fractal dimension to a coherent state. For the Quantum Kicked Top, the fractal dimension of coherent states -when well-defined- exhibits three general behaviors: one related to the presence of critical points and two associated with regular and chaotic dynamics. The finite-size scaling analysis paves the way toward exploring collective spin systems relevant to quantum technologies within the quantum-classical framework.

I Introduction

Phase space methods are becoming a valuable tool for characterizing quantum systems with finite-dimensional Hilbert spaces  [1], particularly for systems with few degrees of freedom. They also play a central role in studying the quantum chaos of systems with a well-defined (semi)classical limit, as is the case for collective spin systems [2], where spin coherent states are the quantum states used to construct such phase space. It is then an essential task to characterize the behavior of the spin coherent states expressed in the eigenbasis of a particular collective spin system. Also, the spin coherent states are central in addressing several questions about quantum chaos in the context of the quantum-classical correspondence. To investigate the latter, one has to increase the dimension of the Hilbert space to make the relevant Planck constant tend to zero, and this is how the fundamental question of this work about the spin coherent states arises: how do the spin coherent states scale in a given eigenbasis, as a function of the dimension of the Hilbert space?

The present contribution is devoted to analyzing the geometry of the phase space of quantum collective spin models (or single J spin models [1]) given by the structure of the eigenvectors of the particular Hilbert space in the context of a classical-quantum correspondence. To do so, we study and classify the scaling of the general inverse participation ratio (IPR) of spin coherent states on different eigenbasis, according to their location on the Poincaré surface of section. First, we exhibit the importance of studying the behavior of the scaling, as it is not guaranteed that a power-law behavior develops to ensure a fractal analysis. We demonstrate the IPR scaling possesses three overall behaviors offering different information: the first one is related to its closeness to critical points; the second signals the regular or chaotic nature of the region where it is located; and the third one indicates specific dynamics around the critical point where the spin coherent state is located, which will be addressed in a future work [3].

The scaling of the moments of the distribution of some quantity can have a critical exponent associated with its fractal dimension or an infinite set of critical exponents associated with multifractality [4, 5]. In quantum mechanics, it characterizes the extension of a state in a given basis of the Hilbert space [6, 7]. There is a close relationship between the concepts of localization in the classical phase space and its quantum counterpart in the Hilbert space [8, 9, 10]. In the present contribution, these concepts are explored employing the scaling of the IPR and, when available, its fractal dimension.

The structure of this article is as follows. In Sec II, we define the tools for a finite-size scaling analysis, including the inverse participation ratio (IPR), its generalized version (IPRq\text{IPR}_{q}), and introduce the concept of the finite-size mass exponent, which allows for both a fractal analysis of the system and a quantitative characterization of the scaling behavior of spin coherent states. We summarize the quantum-classical correspondence of collective spin Hamiltonians via spin coherent states in Sec. III. In Sec. IV, we study the scaling of spin coherent states in the Dicke basis to benchmark the typical behavior of regular dynamics and show indicators of the non-power law behavior around fixed points. The following sections present the finite-size scaling analysis in paradigmatic models. In Sec. V, we study the scaling of the IPR in the eigenbasis of the Lipkin-Meshkov-Glick (LMG) model, which stands as a point of reference for only regular dynamics and unstable fixed points. In Sec. VI, we perform the same analysis applied to the Quantum Kicked Top (QKT) for spin coherent states in the Floquet basis, exhibiting that it has three different (mono)fractal regimes, as well as regions where the asymptotic limit has not been reached. No fractal study can be performed there, but the finite-size analysis still contains valuable information about phase space geometry. Finally, in Sec. VII, we present our conclusions.

II Inverse Participation Ratio and fractality

The inverse participation ratio IPR is a central tool to study the localization of a quantum state |ψ|\psi\rangle [11, 12, 13]. For a given basis {|ϕi}i=1N\{|\phi_{i}\rangle\}_{i=1}^{N} with finite dimension NN, the IPR takes values between 1/N1/N and 11, for a maximally delocalized state and a maximally localized state, respectively. Here, we employ the generalized inverse participation ratio IPRq\text{IPR}_{q}, defined as [14, 15]

IPRq:=iNpiq,\text{IPR}_{q}:=\sum_{i}^{N}p_{i}^{q}\,, (1)

where the sum runs over all the elements of the basis under consideration, and the coefficients pi=|ci|2=|ϕi|ψ|2p_{i}=|c_{i}|^{2}=|\langle\phi_{i}|\psi\rangle|^{2} are interpreted as a probability distribution due to the normalization condition ipi=1\sum_{i}p_{i}=1. This quantity depends on the particular state under consideration and the basis on which it is expressed. The case q=1q=1 is the state’s normalization condition, while q=2q=2 coincides with the usual definition of the IPR. The IPRq\text{IPR}_{q} in Eq. 1 allows us to obtain a quantitative picture of how the state |ψ|\psi\rangle distributes over the basis under consideration at different scales, controlled by the parameter qq, as a function of the basis dimension. The effect of qq in the range 0<q<10<q<1 is to amplify the contribution of the smallest probabilities pip_{i}: the closer qq is to zero, the larger the amplification. Instead, in the range q>1q>1, the effect emphasizes the intensity of the largest probabilities in the distribution. In the following, we will restrict to the interval q(0,4]q\in(0,4] to avoid numerical difficulties [16].

The scaling of IPRq\text{IPR}_{q} as a function of qq has become a customary tool for multifractal analysis in several contexts [5, 14] from turbulent flows [4] and strange attractors [17, 18], to a vast array of phenomena in quantum systems [19, 20, 21, 22, 23, 24, 25]. However, a necessary condition to perform the formal multifractal analysis of a quantum state is that, for each value of qq, the IPRq(N)\text{IPR}_{q}(N) behaves asymptotically as a power-law for large NN [26, 27, 28, 29, 30, 31]. In that case, the scaling of the IPRq\text{IPR}_{q} signals the absence of length scales [32], a concept related to criticality.

Although there are examples in which multifractality for a single state can be analytically described, as, for example, spin coherent states in the spin-z (or computational basis) basis [33], in most quantum systems, it has to be done numerically. This presents a challenge to study the scaling in quantum systems, not only due to computational limitations but also because, in general, there is no simple correspondence for eigenvectors calculated for different dimensions [34]. Only a few states can be easily identified at any dimension, e.g., the ground-state and the highest energy state of a time-independent Hamiltonian in a finite-dimensional Hilbert space. This difficulty increases when one considers time-dependent, periodic Hamiltonians, as there is neither a ground nor maximum energy state. There is a work in progress [35] in which it is proposed a novel way to directly estimate the IPR of a quantum state.

In the present work, we investigate the scaling of coherent states |z(Q,P)|z(Q,P)\rangle distributed in three different eigenbasis basis {|ϕi}i=1N\{|\phi_{i}\rangle\}_{i=1}^{N}, where N=2J+1N=2J+1. Here, the coefficients squared: |ϕi|z(Q,P)|2|\langle\phi_{i}|z(Q,P)\rangle|^{2} have a simple and valuable geometrical interpretation: they are the Husimi distribution of the Eigenvectors in phase space. The Husimi function can represent the structures observed in the Poincaré section in the semiclassical description, a phenomenon known as the principle of uniform semiclassical condensation of Wigner functions of eigenstates (PUSC) [36, 37]. In a mixed phase space, where regular and chaotic dynamics coexist, there are Husimi distribution of some eigenvectors that cover both regions in the classical phase space, whose study is still in progress [38, 37]. When JJ increases, approaching the classical limit, the coherent states remain centered at the same point in phase space, with its area diminishing as 1/J1/J, allowing the study of its scaling as function of JJ.

When the IPRq\text{IPR}_{q} scales as a power-law for large NN, one can employ the mass exponents as a tool to perform a multifractal analysis [5, 31]. They read

τq:=limNlnIPRqlnN\tau_{q}:=-\lim_{N\to\infty}\dfrac{\ln{\text{IPR}_{q}}}{\ln{N}}\, (2)

and are calculated by linearly fitting the ratio between lnIPR\ln\text{IPR} and lnN\ln{N}. Moreover, one can write the mass exponents as

τq=Dq(q1),\tau_{q}=D_{q}(q-1), (3)

where DqD_{q} is called is the generalized (Rényi) dimension [39, 40]. When DqD_{q} is constant, we talk about a monofractal with Hausdorff dimension D0D_{0}. Instead, when DqD_{q} is a function of qq, hence τq\tau_{q} being non-linear, we call the object multifractal [5]. For the particular case of q=2q=2, corresponding to the standard IPR it becomes τ2=D2\tau_{2}=D_{2}, so the mass exponent is equal to the generalized dimension, and D2D_{2} is also called correlation dimension [17, 18].

Here, we generalize this definition to explore the scaling at finite values of NN. We define the finite mass exponent τ\tau as the slope of the lnIPR\ln\text{IPR} vs lnN\ln N between two consecutive points of the IPR curve.

τ:=ln(IPR(N2)/IPR(N1))ln(N2/N1).\displaystyle\tau:=\dfrac{\ln(\text{IPR}(N_{2})/\text{IPR}(N_{1}))}{\ln(N_{2}/N_{1})}\,. (4)

For spin coherent states, we have analyzed different values of qq, and concluded that all values provide the same information as q=2q=2, the case in which we concentrate in the present study.

Via the finite mass exponent, we identify and characterize changes in the behavior of IPR as a function of NN. It is worth noting that it is not aimed to measure the number of basis states needed to express the state of interest, as it does, e.g., the number of elements to fulfill the normalization condition IPR1=ipi=1\text{IPR}_{1}=\sum_{i}p_{i}=1. Instead, it provides information on the ratio between this number and the Hilbert space dimension. Furthermore, it has no restrictions to its functional behavior as it does the standard mass exponent τq\tau_{q}, whose curvature must be positive for all q>0q>0 [40, 41, 5]. It only serves as a quantitative measure to analyse the changes as the dimension of the Hilbert space varies.

Refer to caption
Figure 1: The overall behavior of the scaling of IPR with the dimension of the Hilbert space of spin coherent states in the eigenbasis studied in this work. The dashed vertical colored lines indicate approximately the dimension at which a particular spin coherent state acquires its asymptotic behavior.

The general behavior of the scaling of the IPR can be seen in Fig. 1. The gray solid curve represent a spin coherent states which picks up an asymptotic value for very low values of the dimension (all of the models analyzed have a dimension N=2J+1N=2J+1, and in most cases we will be investigating the scaling from J=50J=50 to J=5000J=5000). Once an asymptotic has been found, depending of their placement, all other spin coherent states will behave typically like the other four colors, with the vertical dashed lines indicating the individual values of Ln(N)Ln(N) where each one picks the same asymptotic values. In general, we will be studying two asymptotic values which correspond to regular and chaotic dynamics. How fast a spin coherent state reaches the asymptotic value strongly depends on its closeness to critical points.

III Classical correspondence and spin coherent states

The phase space of quantum collective spin models can be defined by employing Bloch, also called atomic, or spin coherent states [42], where an effective Planck constant goes as eff=1/J\hbar_{eff}=1/J. These are [43],

|z1(1+|z|2)JezJ^+|J,J,|z\rangle\equiv\dfrac{1}{(1+|z|^{2})^{J}}e^{z\hat{J}_{+}}|J,-J\rangle, (5)

where z=(Q+iP)/4(Q2+P2)z=(Q+iP)/\sqrt{4-(Q^{2}+P^{2})}. J^+=Jx^+iJy^\hat{J}_{+}=\hat{J_{x}}+i\hat{J_{y}} is the raising operator [44], J^α\hat{J}_{\alpha} are the components of the angular momentum vector, and they are generators of the group SU(2)\text{SU}(2) satisfying [J^α,J^β]=iεαβγJ^γ[\hat{J}_{\alpha},\hat{J}_{\beta}]=i\varepsilon_{\alpha\beta\gamma}\hat{J}_{\gamma}  [45]. JJ is the magnitude of the total angular momentum operator J^2\hat{J}^{2}. The Hilbert space of the system is spanned by 2J+12J+1 basis vectors, so it is natural to employ the Dicke basis |J,m|J,m\rangle, where J^2|J,m=J(J+1)|J,m\hat{J}^{2}|J,m\rangle=J(J+1)|J,m\rangle and J^z|J,m=m|J,m\hat{J}_{z}|J,m\rangle=m|J,m\rangle [46]. Here, we restrict the analysis to only integer values of JJ.

A corresponding classical model for a collective spin model can be obtained by taking the expectation value of the quantum Hamiltonian with respect to the coherent states in the thermodynamic limit JJ\rightarrow\infty  [2]

HlimJz|H^|zJ.H\equiv\lim_{J\to\infty}\dfrac{\langle z|\hat{H}|z\rangle}{J}\,. (6)

A classical vector is defined in the same fashion SJ^z/J=(Sx,Sy,Sz)\vec{S}\equiv\langle\hat{J}_{z}\rangle/J=(S_{x},S_{y},S_{z}). The coordinate transformations are Sx=Q1Q2+P24S_{x}=Q\sqrt{1-\frac{Q^{2}+P^{2}}{4}}, Sy=P1Q2+P24S_{y}=P\sqrt{1-\frac{Q^{2}+P^{2}}{4}} and Sz=Q2+P221S_{z}=\frac{Q^{2}+P^{2}}{2}-1; which correspond to a stereographic projection of the Bloch sphere to the plane with the convention that the North pole corresponds to the center and the South pole to the circumference. Coherent states make contact with classical dynamics as JJ increases and serve as a powerful tool to analyze specific regions of phase space by centering the coherent state at the point in phase space of interest [8, 47, 48, 31].

In the three following sections we explore the general properties of the scaling behavior of spin coherent states in three different bases. In the three systems, similar behavior of the IPR of the coherent states is observed as the dimension NN is increased: an initial non-power law behavior, depending on the proximity of their coordinates with a critical point, followed in most cases by a similar power-law. The first model describes simple rotations around the zz axis, the bounded Harmonic Oscillator, whose eigenstates form the Dicke basis. The second is the Lipkin-Meshkov-Glick model, which has regular dynamics. The third is the QKT, which allows the exploration of the transition to chaos.

The analysis of the scaling in the Dicke basis, being simple, with only one critical point, helps to establish the methodology employed to study the following models. The LMG model has more critical points and has regular classical dynamics because the total angular momentum and the energy are conserved quantities, with a time-independent interaction. Next, the QKT model extends the LMG model, introducing a time-periodic function in the interaction term. In this case, the energy is not conserved, and the classical system is usually described in the literature as having 1.5 degrees of freedom [49], exhibiting a transition from regular to chaotic behavior as a function of the kicking strength. In the three cases, the same overall qualitative behavior is found on the scaling of the IPR of the coherent states, with the addition in the regions of chaotic motion of a different power-law exponent.

IV Finite-size scaling for regular dynamics

Refer to caption
Figure 2: (a) Classical trajectories in phase-space of the eight initial conditions chosen as coordinates for the coherent states evolving under HHOH_{HO} (see text for details). (b) Log-Log plots of the IPRIPR of the same coherent states as a function of the dimension of the Hilbert space N=2J+1N=2J+1, JJ varies from 500 to 15,000 in steps of 500. (c) Linear-log plots of the τ\tau of the same coherent states as a function of the dimension (see text for details). The orange dashed line corresponds to the value of 0.5, and the bottom gray dashed line to the value of 0.

We start analyzing the IPR of coherent states in the Dicke basis |J,m|J,m\rangle, a natural representation given that coherent states are built as a rotation of the fiducial state that serves as the vacuum for creating Dicke states [43]. Here and throughout this work, we will be considering the entire Hilbert space, even though the three Hamiltonians analyzed commute with the Parity operator P^=eiπJ^z\hat{P}=e^{-i\pi\hat{J}_{z}}. The reason for this choice is more obvious when considering coherent states, which will significantly overlap with the origin of coordinates that correspond to the Dicke state |J,J|J,-J\rangle. For integer values of JJ, there will be J+1J+1(JJ) Dicke states with positive (negative) parity, and the Dicke state |J,m=0|J,m=0\rangle will always belong to the positive parity. This creates an unnecessary scaling factor to consider when analyzing other regions of phase space. The second reason is that it is known that a parity projected coherent state will be a superposition of two coherent states, as is exemplified in the LMG model, where the ground state is a superposition of two coherent states centered in each well, presented in the next section.

The Dicke basis spans the eigenvectors of the Hamiltonian H^HO=αJ^z\hat{H}_{HO}=\alpha\hat{J}_{z}, which describes rotations around the z-axis. Using the procedure described in III, its associated classical Hamiltonian is built, taking the expected value with respect to the coherent states. It is

HHO(Q,P)=α(Q2+P221).H_{HO}(Q,P)=\alpha\left(\frac{Q^{2}+P^{2}}{2}-1\right). (7)

It is the Hamiltonian of a (bounded) simple Harmonic oscillator in the canonical conjugate variables satisfying the Poisson bracket relation {Q,P}=1\{Q,P\}=1. There is only one critical point that corresponds to the ground state, a stable center, and no other relevant structure is present [44]. Classically, if an initial condition is placed at the center, it will remain there, and all other initial conditions will revolve around this fixed point. In Fig. 2 (a), the classical trajectories of eight initial conditions are depicted, with coordinates starting from Q=0Q=0 to Q=0.14Q=0.14 in increments of 0.02 for a representative value α=0.84\alpha=0.84.

To perform a similar analysis in the quantum regime, the behavior of ln(IPR)\ln(\text{IPR}) vs. ln(N)\ln(N) is shown in Fig. 2 (b), for the same coordinates, now employed to define the coherent states, and with the same color code. The IPR of the first coherent state corresponds to the first pink horizontal line at the top of Fig. 2 (b). This is a consequence of the fact that, for any dimension, the coherent state with coordinates (0,0)(0,0) corresponds to the Dicke state |z(0,0)=|J,J|z(0,0)\rangle=|J,-J\rangle. This behavior is associated with the coherent state that does not scale; in other words, its power law has a null exponent. This will happen for any coherent state which only takes a constant amount of elements of the basis into consideration, as a function of NN. The second curve corresponds to the coordinate (Q,P)=(0.02,0)(Q,P)=(0.02,0). It can be seen that, from Ln(N)=7Ln(N)=7 to Ln(N)=9Ln(N)=9, the IPR exhibits a curved, a non-power law behavior, and transitions to a linear, power-law one, for larger values of NN. As this coherent state |z(0.02,0)|z(0.02,0)\rangle is described by a small fraction of elements of the Dicke basis, it requires a large dimension to express the power law behavior. The closer a coherent state is to the center, the more non-linear behavior the IPR curve will show in the log-log plot. The remaining coherent states very quickly pick up the same power-law behavior as a function of NN.

To get a more quantitative description of the previous analysis, in Fig. 2 (c), we show the finite mass exponent τ\tau of the eight coherent states as a function of the dimension. In Fig. 2 (b), we connected each point with linear segments. The slopes of each segment are the values of the finite mass exponent. When the IPR shows a power-law behavior in the log-log plot, τ\tau is the exponent of this power-law. If this is not the case, then the finite mass exponent only represents the change in the proportion of the Dicke states that the coherent state is using when the dimension is increased. As mentioned in the previous paragraph, the power-law of the first coherent state is exactly zero, and its τ\tau curve is located at the bottom. From top to bottom, the second curve corresponds to the coherent state |z(0.02,0)|z(0.02,0)\rangle. It is the last one to start picking up the same power-law of the other six coherent states: for lnN=ln30001=10.3\ln{N}=\ln{30001}=10.3 the value of τ\tau has almost reached the limiting value of 0.5.

Refer to caption
Figure 3: (a) Log-log curve of the lnIPRq\ln{\text{IPR}_{q}} vs ln(N)\ln{\left(N\right)} of the coherent state |z(0.02,0)|z(0.02,0)\rangle. (b) Log-log curve of the lnIPRq\ln{\text{IPR}_{q}} vs ln(N)\ln{\left(N\right)} of the coherent state |z(0.14,0)|z(0.14,0)\rangle. In (a) and (b), the values of qq go from q=0.1q=0.1 to q=4.0q=4.0 in steps of 0.1 from top to bottom. (c) τq\tau_{q} vs qq of the coherent state |z(0.14,0)|z(0.14,0)\rangle. The orange dashed line corresponds to a mono-fractal state, which scales as a Gaussian distribution [31], and the black dashed horizontal line corresponds to a quantum state, which does not scale with the dimension NN.

Up to this point, we have only studied the power-law behavior captured by the finite mass exponent for the case of IPRq=2\text{IPR}_{q=2}. A multifractal characterization requires a similar analysis for other parameter values qq. We explore this in the interval for q[0.1,4]q\in[0.1,4]. Employing values of qq closer to zero or higher than 4 requires expanding the numerical precision and does not provide relevant additional information in the present context. In Fig. 3(b) we show the IPRq\text{IPR}_{q} for the last initial condition, with coordinates (Q,P)=(0.14,0)(Q,P)=(0.14,0). When IPRq\text{IPR}_{q} exhibits a power law behavior, its slope is proportional to q1q-1, as shown in Eq. 3. The curves with negative slope correspond to the values with q<1q<1, and the curves with positive slope to the values with q>1q>1; the last two kinds of slopes are separated by the horizontal line with zero slope, q=1q=1, which is the normalization condition. The IPRq\text{IPR}_{q} of the second coherent state is shown in Fig. 3 (a). In the range of dimensions employed, from J=500,N=1001J=500,N=1001 to J=15000,N=30001J=15000,N=30001, for all values of qq there is no linear relation between Ln(IPRq)Ln(\text{IPR}_{q}) and Ln(N)Ln(N), no power law behavior. It implies that it is not possible to perform a formal MFA in this state. From observing the other coherent states, we conclude that the closer the coherent state is to the critical point, the greater the dimensions needed to perform the MFA.

To summarize, in Fig. 3 (c) the slopes τq\tau_{q} of all the lines in Fig. 3 (b) are plotted against qq. The linear relation with qq is clearly visible, allowing us to employ Eq. 3 to derive the fractal dimension D0=12D_{0}=\frac{1}{2}. It is a clear example of monofractal scaling. This line is representative of nearly all other initial conditions. The exceptions are the first one, which does not scale with the dimension and is equivalent to the horizontal black dashed line, and the second one, where the MFA cannot be performed for the range of dimensions explored in this work. The most significant learning from this section is the usefulness of the finite mass exponent τ\tau. It indicates when the dimension is large enough to perform a multifractal analysis (MFA): in the region where a power-law scaling of the IPRq\text{IPR}_{q} is observed for all the values of qq of interest. In the case of the IPR of the coherent states in the Dicke basis, the second coherent state |z(0.02,0)|z(0.02,0)\rangle does not exhibit a power-law in the range of dimensions from N=1001N=1001 to N=30001N=30001. It has been shown analytically that a coherent state with a Gaussian distribution in a particular basis, associated with regular classical dynamics, in the limit of N1N\gg 1, has a D0=1/2D_{0}=1/2 [31]. In numerical studies, the challenge is to employ a dimension large enough to exhibit this power-law scaling for a given coherent state. The finite mass exponent helps determine when a formal MFA analysis can be performed in each case.

It is worth mentioning that there is more information on the dependence of the finite mass exponent with NN than that MFA can capture. When it is not possible to perform an MFA, the absence of a power-law provides valuable information about the location of the coherent state in phase space. We show elsewhere [3] how the dynamics of a coherent state located close to a fixed point leaves a fingerprint in the scaling of the IPR, even in the absence of a power-law scaling and without the need to go to the limit of N1N\gg 1.

V Finite size scaling in the LMG model

Refer to caption
Figure 4: (a) Contour plot of the semiclassical energy function of the LMG model for α=0.84\alpha=0.84 and k=2k=-2. (b) Classical energy cut from Fig. 4 (a) at P=0P=0. The coordinates of 250 coherent states are shown as each colored dot. The yellow cross indicates the unstable fixed point, the green one indicates the stable minimum, and the cyan indicates the separatrix edge where there is an ESQPT. The gray dots are shown only for reference. (c) Log-Log plot of the IPR as a function of NN for the 250 coherent states in the LMG basis. The dimension goes from N=201N=201 to N=11001N=11001 in increments of J=100J=100.

In what follows, we present a detailed study of the scaling of the IPR for selected coherent states |z(Q,P)|z(Q,P)\rangle in Eq. 5, representing points in the phase space associated with the dynamics of the Lipkin-Meshkov-Glick Hamiltonian [50, 51, 52] (LMG), a model from nuclear physics and quantum magnetism that have become a customary tool to study collective qubit systems in several fields including such as cold atoms [53, 54, 55] or cavity QED setups [56], whose critical properties and classical limit have been extensively explored [57, 58, 59, 60, 61]. The LMG model corresponds to the non-kicked limit of the QKT, which will be analyzed later. It describes the continuous motion of an angular momentum vector with rotation and torsion. Its Hamiltonian is (=1\hbar=1):

H^LMG=αJ^z+k2JJ^x2.\hat{H}_{LMG}=\alpha\hat{J}_{z}+\dfrac{k}{2J}\hat{J}_{x}^{2}.

Its semiclassical limit, obtained via spin coherent states, reads

HLMG=αSz+k2Sx2.H_{LMG}=\alpha S_{z}+\frac{k}{2}S_{x}^{2}. (8)

In what follows, we set the values α=0.84\alpha=0.84 and k=2k=-2. The classical dynamics is regular, and for the selected parameters, three critical points emerge: two stable minima and one unstable saddle point with a positive Lyapunov exponent  [62]. In Fig. 4 (a), the corresponding classical energy surfaces of the LMG model are shown, consisting of a double well and a local maximum. Fixed points are crucial in our analysis because the scaling we investigate gets us closer to the semiclassical limit. The coherent states provide a mean-field description of these critical points, which becomes exact in the thermodynamic limit NN\rightarrow\infty [2, 1, 44, 63].

Refer to caption
Figure 5: Finite Size Scaling in the LMG Model. (a) Log-Log curves of the IPR vs NN eleven coherent states with coordinates P=0P=0 and QQ from 1.04 to 1.14 in steps of 0.01. (b) (linear-linear) Finite mass exponent as a function of ln(N)\ln{\left(N\right)} for the same eleven coherent states. The coherent state closest to the stable fixed point (right minimum in Fig. 4) is marked with the green cross. (c) Same as (a) but for eleven coherent states with coordinates P=0P=0 and QQ from 1.48 to 1.58 in steps of 0.01. (d) Same in (b) but for the same eleven coherent states in (c). The coherent state closest to the separatrix edge is marked with the cyan cross. (e) Same as (a) but for ten coherent states with coordinates P=0P=0 and QQ from -0.05 to 0.05 in steps of 0.01. (f) Same as (b) but for the same eleven coherent states in (e). The coherent states next to the unstable fixed point are marked with the yellow cross |z(±0.01,0)|z(\pm 0.01,0)\rangle. Due to the symmetry of the location of the coherent states, the pink curves in both panels are exactly under the green ones. The diagonalization was done from ln(N)=ln(2(100)+1)\ln{\left(N\right)}=\ln{\left(2(100)+1\right)} = 5.30 to ln(N)=log2(10000)+1\ln{\left(N\right)}=\log{2(10000)+1} = 9.90 in steps of J=100J=100. The dashed horizontal orange line in (b), (d), and (e) corresponds to the value of 0.5.

Fig. 4 (b) displays a cut of the energy surface HLMGH_{LMG} for P=0P=0. Each dot represents the coordinates of one of 250 coherent states selected along a segment from (Q,P)=(0.5,0)(Q,P)=(-0.5,0) to nearly the edge of the phase space (Q,P)=(1.99,0)(Q,P)=(1.99,0). In Fig. 4 (c), we plot ln(IPR)\ln(\text{IPR}) versus ln(N)\ln(N) with the same color code for each coherent state in Fig. 4 (b). By looking at their asymptotic behavior, most of these coherent states show a clear linear behavior, with their IPR scaling with the power-law N1/2N^{-1/2}, with a finite mass exponent τ=0.5\tau=0.5. As mentioned above, this slope has been associated before with a coherent state that scales as a Gaussian when distributed over the eigenstates of a Hamiltonian [31]. While it is not guaranteed that the components of an atomic coherent state will have a Gaussian profile when expressed on the LMG basis, we have confirmed that this is the case for all the particular coherent states we investigated.

There is a set of coherent states whose values do not change as fast as a function of NN. Their curves are those at the top in Fig. 4 (c). This set encloses those coherent states centered closer to the fixed points and the separatrix, denoted by the three different crosses in Fig. 4 (b)]. They seem to require a greater dimension to display the same power-law behavior of N1/2N^{-1/2}. In these cases, the asymptotic limit has not been reached for these values of NN, which are very large, so there is not a distinguishable power-law behavior of the IPR, a mass exponent τq\tau_{q} cannot be obtained, making not possible to perform a formal multifractal analysis in this range of dimensions. The first curve corresponds to the coherent state, which is closest to the right minimum in the energy surface [Fig. 4 (a)] indicated by the yellow cross. For large NN, it becomes very close to the exact ground state. Notice that the exact quantum ground state combines the right and left coherent states at the minima due to symmetry. This state has IPR2\text{IPR}\sim 2 and remains essentially constant, with a null slope, scaling as N0N^{0}. Inspired by Ref. [64], we perform a detailed scaling analysis around the two critical points and the separatrix edge to better understand these coherent states.

In Fig. 5 (a), we show the log-log curves of the IPR vs. lnN\ln{N} and in panel (b), the finite mass exponent of eleven coherent states around the stable fixed point. Most of these coherent states overlap significantly with many eigenvectors with energies close to the minimum, except for the one associated with a single eigenvector, which corresponds to the ground state. The IPR curves are very smooth as a function of the dimension. They have qualitatively the same behavior as those found in the Dicke basis. The coherent state located near the right minimum |z(1.08,0)|z(1.08,0)\rangle is the state that has not reached its asymptotic power law at the dimension considered here N=20001N=20001. We believe that if the coherent state is not located exactly at the fixed point for sufficiently high dimensions, it will acquire the same power-law of IPRN0.5\text{IPR}\approx N^{-0.5}.

In figure 5 (c) and (d), we analyze another eleven coherent states in detail, but now focusing on the region of the separatrix edge, where there is an Excited State Quantum Phase Transition (ESQPT) in the LMG model [65, 66]. In this case, we get similar results as in previous sections: some coherent states have reached their asymptotic values with slope 1/21/2, while others would require larger dimensions to arrive at this value. A detailed study of the connection of the ESQPT and the specific behavior of the scaling of the IPR is beyond the scope of the present article.

In Fig. 5 (e) and (f), we show the scaling of the IPR and its corresponding finite mass exponent τ\tau of five coherent states to the left of the unstable fixed point, located at (Q,P)=(0,0)(Q,P)=(0,0), and five to the right. Due to the symmetry of their location, only the green curves can be seen. In this case, we observe exactly the same qualitative behavior of the IPR of the coherent states located around a critical point as in the previous section. The closer the coherent state is to a critical point, the closest being these two: |z(±0.01,0)|z(\pm 0.01,0)\rangle, the larger the dimension needed to exhibit the power-law IPRN0.5\text{IPR}\approx N^{-0.5}. The slope between two consecutive points in Fig. 5 (e) is the finite mass exponent, whose dependence on Ln(N)Ln(N) is shown in Fig. 5 (f). It exhibits strong fluctuations around the limiting value of τ=0.5\tau=0.5. These sudden changes are associated with the fact that any coherent state with coordinates close to (Q,P)(0,0)(Q,P)\approx(0,0) has very few elements in the Dicke basis, and also in the H^LMG\hat{H}_{LMG} basis, which makes the IPR to have significant changes as NN is increased. Although in the thermodynamic limit NN\to\infty, there will be only one eigenvector at the critical point energy, for any finite value of NN, the exact location and energy of this eigenvector usually fluctuate around the energy of the coherent state |z(0,0)|z(0,0)\rangle. This is a crucial difference between the critical points analyzed in the previous section because the center of the phase space in the LMG model is now an unstable fixed point. A detailed analysis relating the fluctuations of the IPR quantified by τ\tau and the stable or unstable fixed points will be investigated elsewhere [3].

Finally, Fig. 6 shows the MFA analysis for one representative coherent state expressed in the LMG basis. It gives the same information as the Dicke basis: the coherent state is located in a regular region. All other initial conditions analyzed in which a MFA can be performed with the conditions mentioned before give exactly this same figure.

Refer to caption
Figure 6: τq\tau_{q} vs qq of the coherent state |z(0.5,0)|z(-0.5,0)\rangle for the LMG model. The orange dashed line corresponds to a mono-fractal state, and the black dashed horizontal line corresponds to a quantum state that does not scale.

VI Scaling in the Quantum Kicked Top

The quantum kicked-top model is one of the simplest quantum systems that allows the study of the transition from regular to chaotic dynamics as the kicking strength increases  [46, 67, 68, 10]. Its advantage lies in its algebraic simplicity, finite dimension, and well-defined semiclassical limit. The quantum kicked top model has been experimentally realized using the combined electronic and nuclear spin of a single atom[69], where deep in the quantum regime, a good correspondence between the quantum dynamics and classical phase space structures was observed. The dynamics of a mean-field quantum kicked top can be simulated by performing weak collective measurement on a large ensemble of two-level quantum systems and applying global rotations conditioned on the measurement outcome [70]. The presence of chaos can be observed by measuring the out-of-time-order correlation functions [71, 72]. Effective thermalization [73], and ergodic dynamics have been observed in a small quantum system consisting of only three superconducting qubits [74, 75]. The fixed-point analysis and calculation of the Lyapunov exponents show correspondence to the random matrix theory in the chaotic regime [76].

The Hamiltonian of the QKT includes a zz-axis rotation of a quantum angular momentum vector J^\hat{\vec{J}} and kicked torsions around the xx-axis, given by (\hbar=1)  [46]

H^(α,k,t)=αJ^z+k2JJ^x2n=0δ(tnτ),\hat{H}(\alpha,k,t)=\alpha\hat{J}_{z}+\frac{k}{2J}\hat{J}_{x}^{2}\sum_{n=0}^{\infty}\delta(t-n\tau), (9)

The zz-rotation is parametrized by α\alpha, and periodical torsions are modeled by a train of delta functions with strength controlled by k2J\frac{k}{2J}.

VI.1 Floquet formalism

Because the Hamiltonian is periodic in time, one can apply the Floquet theorem [77] to calculate a state’s temporal evolution at times tn=nt_{n}=n, considering a unitary time scale τ=1\tau=1. The time-evolved state is given by |ψ(tn)=F^n|ψ(0)|\psi(t_{n})\rangle=\hat{F}^{n}|\psi(0)\rangle\,, where the Floquet operator F^\hat{F} is

F^(α,k)=exp[iαJ^z]exp[ik2JJ^x2].\hat{F}(\alpha,k)=\exp\left[-i\alpha\hat{J}_{z}\right]\exp\left[\frac{-ik}{2J}\hat{J}_{x}^{2}\right]\,. (10)

The Floquet eigenvectors |ϕj|\phi_{j}\rangle\, fulfill the equation F^|ϕj=eiϕj|ϕj\hat{F}|\phi_{j}\rangle=e^{i\phi_{j}}|\phi_{j}\rangle\,, where the eigenvalues ϕj\phi_{j} (mod 2π2\pi) are called quasienergies [46]. Due to the unitary character of the Floquet operator, quasienergies are complex, with unitary modulus, and distributed on the unitary circle in the complex plane. Depending on the chosen branch, they can be listed in any interval of size 2π2\pi, like [π,π][-\pi,\pi] or [0,2π][0,2\pi]. Because the Floquet operator commutes with the parity operator P^=eiπJ^z\hat{P}=e^{-i\pi\hat{J}_{z}}, [F^,P^]=0[\hat{F},\hat{P}]=0, it does not change the parity of well-defined parity initial state. For instance, the Dicke basis and the eigenvectors of F^\hat{F} have well-defined parity, while the atomic coherent states do not. The model has several geometric symmetries, classical and quantum, affecting its dynamics [67, 78]. In Ref. [79], the authors describe most of the classical symmetries of the QKT and related p-spin models. The overall impact of these symmetries depends heavily on the particular phenomena one is investigating. Also, depending on the parameters chosen, one may encounter other symmetries as the ones described in  [80]. When chaos is fully developed in the whole phase space, the quasienergy spectrum of the Floquet operator is very well characterized by the Circular Orthogonal ensemble (COE), as described in  [67], due to its anti-unitary time reversal invariance.

VI.2 Classical correspondence in the QKT

In the classical limit, the system becomes a classical top fixed by its tip, free to rotate in the zz direction and experiencing periodical kicks along the xx axis [46, 47].

H=αSz+k2Sx2n=+δ(tnτ),H=\alpha S_{z}+\dfrac{k}{2}S_{x}^{2}\sum_{n=-\infty}^{+\infty}\delta(t-n\tau)\,, (11)

After each kick, the dynamics on the Poincaré section are given by the Poincaré map, which results from solving the classical equation of motion before and after the kicks [81]. One then gets from the classical Hamiltonian in Eq. 11, a nonlinear area-preserving map, the Poincaré map Sn+1=FSn\vec{S}_{n+1}=\,F\vec{S}_{n} with FF given by

F=(cos(α)sin(α)cos(kSx)sin(α)sin(kSx)sin(α)cos(α)cos(kSx)cos(α)sin(kSx)0sin(kSx)cos(kSx))F=\begin{pmatrix}\cos{(\alpha)}&-\sin{(\alpha)}\cos{(kS_{x})}&\sin{(\alpha)}\sin{(kS_{x})}\\ \sin{(\alpha)}&\cos{(\alpha)}\cos{(kS_{x})}&-\cos{(\alpha)}\sin{(kS_{x})}\\ 0&\sin{(kS_{x})}&\cos{(kS_{x})}\end{pmatrix} (12)

VI.3 Quantum chaos in the QKT

The quantum model is periodic in the kicking strength for integer and semi-integer JJ integer [80]. Still, in either case, when chaos is fully developed in the phase space, both parity sectors of the Floquet operator used in this work always belong to the COE ensemble  [67]. Throughout this work, we keep fixed the value α=0.84\alpha=0.84 and study the transition from regularity to chaos as a function of kk. To quantify quantum chaos, a useful signature is the average r\langle r\rangle of the spectral parameter ri=min(1/δI,δI)r_{i}=\text{min}(1/\delta_{I},\delta_{I}) with δi=(ϕi+1ϕi)/(ϕiϕi1)\delta_{i}=(\phi_{i+1}-\phi_{i})/(\phi_{i}-\phi_{i-1}) [47]. The rr factor is calculated over the full quasienergy spectrum {ϕi}\{\phi_{i}\}, separated by parity and sorted in ascending order. In Fig. 7, the value of r\langle r\rangle is shown as a function of the kicking strength kk, for a fixed value of α\alpha and J=2000J=2000. The orange dashed line in Fig. 7 at r=0.386r=0.386 indicates the value associated with Poisson statistics, hence regular dynamics, and the golden dashed line at 0.53 corresponds to a Wigner-Dyson statistics of the spacings, associated with chaotic dynamics. It can be seen that the QKT has essentially regular dynamics for k<2k<2 and chaotic dynamics for k>7k>7. Notice that there are subtle differences between each parity sector.

Refer to caption
Figure 7: Mean value of rr as a function of kk for both parity sectors in the QKT. The dimension of the Floquet operator is N=4001N=4001, and each parity sector is of dimension N+=2001N_{+}=2001 and N=2000N_{-}=2000. The orange dashed horizontal line marks Poissonian statistics, and the golden dashed horizontal line marks Wigner-Dyson statistics.
Refer to caption
Figure 8: Log-Log plots of the IPRIPR vs NN of 31 coherent states in the QKT for four values of the kicking strength: (a) k=0.5k=0.5, (b) k=1.3k=1.3, (c) k=2.5k=2.5, and k=30k=30. For all initial conditions, P=1P=1 and QQ vary from 0.2 to 0.8 in steps of 0.02. The insets show the classical Poincaré section, with each initial condition sharing the color with the IPR of the corresponding coherent state.

VI.4 Finite size scaling in the QKT model

There is a fundamental difference between the QKT and the LMG and rotor models. The description of the QKT model is stroboscopic in the classical and quantum regimes. So, the quantum procedure expresses the coherent states on the Floquet basis, not on a Hamiltonian basis. On the other hand, it has been shown in the literature that there is indeed a quantum-classical correspondence between the eigenvectors of the time-periodic description of the QKT model with the Floquet operator and the predictions of Random Matrix Theory in the chaotic regime, with the corresponding classical structures generated by the Poincaré map in the classical regime  [36, 82, 83, 48].

While some results are similar to those of the previous sections, the scaling analysis presents additional challenges. At variance from the previous systems, whose dynamics are regular and have very few critical points, in the QKT model, there are many critical points. They are described by the Poincaré-Birkhoff theorem: when a resonant Tori breaks, it leaves behind a new set of fixed points  [84, 85]), which are the classical structures surviving in phase space after a system is brought from regular to chaotic dynamics [83]. When the kicking strength kk is non-null, there are infinitely many critical points, and the scaling analysis could become cumbersome because, as it was shown in the previous sections, the proximity of critical points requires higher dimensions to rich the linear power law behavior where the MFA can be performed. When it is possible, in what follows, it is shown that, in the chaotic regime and for very large dimensions, there is a new exponent for the power-law scaling of the IPR, which captures the chaotic motion. In Fig. 8 it is shown the scaling of the IPR of thirty-one coherent states for four values of the kicking strength, calculated from J=50J=50 to J=5000J=5000 in steps of 50. The insets present the associated Poincaré sections.

VI.5 Generic regular power-law behavior of the IPRIPR

Because the k=0.5k=0.5 and k=1.3k=1.3 cases exhibit dominant regular dynamics, we anticipate regions where one obtains scaling results similar to those of the HO and LMG Hamiltonians. In Fig. 8 (a) and (b) it is shown the IPR curves for thirty-one coherent states for two different values of the kicking strength k=0.5k=0.5, and k=1.3k=1.3. Most curves have already reached the same asymptotic power-law of IPRN0.5\text{IPR}\approx N^{0.5}. In panel Fig. 8 (b), there are several curves colored in purple with an initial non-power law behavior, which will be treated in the following subsection. These coherent states are located across a six-cycle, and as we have already shown in the previous section, their initial non-power-law is due to their closeness to a critical point, in this case, the six-cycle [3]. All the other curves follow the same power law.

We select the coherent state |z(0.2,1)|z(0.2,1)\rangle as representative among all other coherent states for the cases of k=0.5k=0.5 and k=1.3k=1.3, except the purple ones mentioned in the previous paragraph. Its scaling resembles those of the HO and LMG Hamiltonians. In Fig. 9 (a) it is shown the lnIPRq\ln{\text{IPR}_{q}} vs lnN\ln{N} of this coherent state for forty values of qq, from 0.1 to 4.0. Notice that these curves refer to a single coherent state. We are not averaging from many states as is a customary procedure in the MFA [47]. As there is a linear power law scaling, their slopes, the finite mass exponent, are plotted in Fig. 9 (b). Each brown dot corresponds to the slope of the linear fitting from each curve of Fig. 9 (a). The other three dashed lines correspond to the three limiting cases explained in Sec. III. The golden dashed line indicates not only the maximally distributed state in a given basis but also corresponds to any state that grows exactly proportionally to the increase of the dimension, as it will be numerically demonstrated in the next section. The third black dashed line corresponds to a state that does not scale with the dimension, which, in our analysis, corresponds to a coherent state located at a critical point inside a regular region. The second orange dashed line corresponds to the analytical τq\tau_{q} of state having a Gaussian distribution in a given basis  [31]. All of the coherent states analyzed so far, which are not in the immediate vicinity of a critical point, are subject to a formal MFA, and the result is that they acquire the same exact relationship as a Gaussian distribution. Appendix A and B present the finite mass exponents calculated for all the initial conditions. As mentioned above, more information about the phase space dynamics can be extracted from the IPR scaling, which will be presented in a future work [3].

Refer to caption
Figure 9: (a) Log-log curve of the lnIPRq\ln{\text{IPR}_{q}} vs ln(N)\ln{\left(N\right)} of a typical coherent state placed inside a regular part of the phase space of the QKT. From top to bottom, the values of qq go from q=0.1q=0.1 to q=4.0q=4.0 in steps of 0.1. (b) τq\tau_{q} vs qq of a coherent state placed inside a regular part of the phase space.

VI.6 Non-generic behavior of the IPRIPR in mixed phase space

The mixed case k=2.5k=2.5 develops chaos in phase space, as seen in the Poincaré section in the insert of Fig. 8(C). In Fig. 10, we show the finite mass exponent for three initial conditions in light tones. They display strong fluctuations, which do not get smoother when the dimension is increased. Even taking a moving average, shown in darker tones, the fluctuations remain, making it very difficult to extract information from the scaling of the coherent states. The finite mass exponent calculated for these initial conditions is presented in Appendix C.

Refer to caption
Figure 10: The finite mass exponent of three coherent states with coordinates P=1P=1 and Q=0.2,0.38,0.8Q=0.2,0.38,0.8 (red, magenta, and blue respectively) inside mixed phase space of the QKT (see Fig. 8). The light tone curves are the real values of τ\tau, and the darker tones are moving averages calculated in intervals of five values. The horizontal dashed lines indicate the asymptotic values for τ\tau with chaotic (golden) and regular(orange) phase space.

VI.7 Generic chaotic power-law behavior of the IPRIPR

From the Poincaré section in the insert of Fig. 8(D), the case k=30k=30 can be considered as fully chaotic, despite the possible presence of small regular regions  [81]. In this case, it is possible to perform the MFA for large values of NN, and the result is a new exponent τ=1\tau=1 for the power-law, which we identify with the chaotic behavior. It can be observed in Fig, 11 (a), where the finite mass exponent of the coherent state |z(0.2,1)|z(0.2,1)\rangle is shown as a function of NN for k=30k=30. The slope for each consecutive pair of points in Fig. 8(D) is presented with light tone lines, and the darker tone is a moving average of τ\tau calculated in intervals of five values. The horizontal dashed line indicates the asymptotic value, which is close to 1 and is representative of most of the coherent states for this kicking strength.

In Fig. 11 (b), we show the IPRq\text{IPR}_{q} of the same coherent state |z(0.2,1)|z(0.2,1)\rangle, for values of q(0.1,4)q\in(0.1,4), as a typical example of the other thirty initial conditions. All the lines exhibit a power law scaling, allowing for the MFA. The fractal dimension of all these states is D0=1D_{0}=1. In Fig. 11 (c), we have performed the MFA from N=2001N=2001 to N=10001N=10001 for the chaotic case. Notice that the numerical value for the mass exponent is very close to one. This happens for the other thirty initial conditions, their Dq1D_{q}\approx 1.

Refer to caption
Figure 11: (a) Typical finite mass exponent of the coherent state |z(0.2,1)|z(0.2,1)\rangle for almost chaotic phase space (see Fig. 8). The light tone curve is the real value of τ\tau, and the darker tone is a moving average calculated in intervals of five values. The horizontal dashed line indicates the asymptotic value for a typical initial condition inside a chaotic phase space. (b) Log-log curve of the lnIPRq\ln{\text{IPR}_{q}} vs ln(N)\ln{\left(N\right)} of the coherent state |z(0.2,1)|z(0.2,1)\rangle. From top to bottom, the values of qq go from q=0.1q=0.1 to q=4.0q=4.0 in steps of 0.1. (c) τq\tau_{q} vs. qq of a typical coherent state placed inside a chaotic part of the phase space.

VII Discussion and Conclusion

We have presented a detailed study of the scaling of the inverse participation ratio (IPR) for representative spin coherent states in three different bases, the bounded Harmonic Oscillator, the Lipkin-Meshokov-Glick model, and the Quantum Kicked Top, and related it to its corresponding classical dynamics, including regular, mixed, and fully chaotic regions in phase space.

We have exhibited that the scaling of the coherent states does not always display a power-law scaling, where a multifractal analysis (MFA) can be performed. To identify the dimensions where MFA is a valuable tool, we introduced a NN dependent finite mass exponent τ\tau, which coincides with the standard mass exponent τq\tau_{q} when, on average, acquires a constant value. This tool has proved to help probe phase space structures, such as regularity, chaos, and fixed points, but it also provides insight into situations where the power-law behavior required to perform a multifractal analysis cannot be asymptotically reached and identify conditions where it emerges.

Coherent states centered at regions of phase space associated with regularity exhibit, in general, power-law scaling of the IPR as a function of the dimension NN, which has the same finite mass exponent τ=1/2\tau=1/2 for all values qq of the mass exponent τq\tau_{q}. These states have monofractal behavior, with scaling IPRqJ12\text{IPR}_{q}\approx J^{\frac{1}{2}}. How fast this asymptotic behavior is reached depends on the closeness of the coherent state to critical points.

The finite mass exponent shows an important consequence of critical points in phase space, i.e., that scaling coherent states close to them needs larger values of NN to reach the asymptotic power-law behavior required to perform MFA. This was verified for the integrable rotor associated with a bounded harmonic oscillator, for the LMG model, and the regular regions of the QKT. At the limit of large NN, the fraction of eigenstates needed to describe any of these coherent states goes to zero, showing that the localization of the coherent states in phase space, whose area scales as 1/J1/J, goes in parallel with their localization in the Hilbert space. It was also shown that coherent states provide a very accurate description of the eigenstates at fixed points in phase space, implying that their IPR is close to 1 and independent of the dimension of the Hilbert space. The scaling is IPRJ0IPR\approx J^{0}, and the mass exponent is null in these cases. Although the calculations presented cover two orders of magnitude of values of JJ, from 50 to 5000, for some states very close to the fixed points, reaching the power-law asymptotic region was not possible.

On the other hand, it was shown that coherent states in the fully chaotic region, where the Poincaré section has no visible structure, the IPR scales as IPRJ1\text{IPR}\approx J^{1}, with mass exponent 1, and a monofractal behavior. One should underline that this scaling does not imply that any coherent state in the chaotic region has components in all the eigenbasis. It means that the fraction of eigenstates needed to describe each coherent state remains constant as the dimension of the Hilbert space increases. It was also shown that the scaling of coherent states in the chaotic region of the mixed-phase space has strong fluctuations. In many cases, it was not possible to reach an asymptotic behavior numerically. Again, in these cases, a formal multifractal analysis could be performed.

Studying the scaling of the inverse participation ratio and the finite mass exponents confirms previous results for collective spins systems [8, 47, 48, 31] but also allows the analysis of individual coherent states, complementary to techniques that rely on averaging over phase space. We expect these results to pave the way to study the mixed phase space and finite-size features of collective spin systems and other algebraic systems with a well-defined classical limit, which are now relevant in the context of quantum technologies.

Acknowledgements.
We wish to acknowledge S. Lerma-Hernández for his useful comments on these topics. We also acknowledge the support of the Computation Center—ICN, particularly Enrique Palacios, Luciano Díaz, and Eduardo Murrieta. This research was partially funded by DGAPA- UNAM Project No. IN109523.

Appendix A Finite mass exponent for k=0.5k=0.5

In Fig. 12, we show all the finite mass exponent calculated for the thirty-one initial conditions with coordinates P=1P=1 and Q[0.2,0.8]Q\in[0.2,0.8] for the QKT. Notice that the value of all the τ\tau’s converge quickly to the expected value obtained in the previous sections, allowing the MFA to be performed. In the end, near ln(N)=9\ln(N)=9, there is a disturbance in the value of the finite mass exponent for some initial conditions. It claims for a more detailed analysis, which is in process and will be presented in future article [3].

Refer to caption
Figure 12: The finite mass exponent of the thirty-one initial conditions with coordinates P=1P=1 and Q[0.2,0.8]Q\in[0.2,0.8] of the QKT.

Appendix B Finite mass exponent for k=1.3k=1.3

All the other mass exponents for the case of k=1.3k=1.3 of the QKT are presented. Although there are regular dynamics in all the phase space, there are several new critical points compared with the LMG model due to the breaking of integrability induced by the kicking potential. This new geometry of the phase space is also captured by the scaling of the IPR. For the initial conditions analyzed in the main text (VI) where it was possible to perform the MFA, in Figs. 13 (a) and 13 (b), the finite mass exponent are shown, exhibiting the result for regular phase space D0=12D_{0}=\frac{1}{2}. In Fig. 13 (c), we show the initial conditions where the MFA cannot be performed due to the closeness of those coherent states to the center of that ellipse (see Fig. 8), which belongs to a six-cycle.

Refer to caption
Figure 13: (a) Finite mass exponent for the coherent states with coordinates at (a) P=1P=1 and Q[0.2,0.4]Q\in[0.2,0.4] (b) P=1P=1 and Q[0.42,0.52]Q\in[0.42,0.52], and (c) P=1P=1 and Q[0.54,0.8]Q\in[0.54,0.8] of the QKT (same color code as the Fig. 8). The horizontal dashed line indicates the asymptotic values for τ\tau inside regular phase space.

Appendix C Closer look to mixed phase space for k=2.5k=2.5

Here, we show the IPR of another four initial conditions placed inside the chaotic part of the mixed phase space of the QKT. This time, we scaled the system from J=50J=50 up to J=10050J=10050 (dimension N=20101N=20101) in steps of ΔJ=50\Delta J=50 in order to see if the finite mass exponent would stabilize its value so the MFA could be performed. We show the scaling in Fig. 14 (a). The finite mass exponent gives the result shown in Fig. 14 (b). As can be seen in this figure, their value fluctuates very erratically. To be able to see in greater detail around what value those finite mass exponents are fluctuating, in Fig. 14 (c)-(f), we have calculated the cumulative sum as a function of the dimension of the values shown in Fig. 14 (b). Three curves tend to fluctuate around the value one, which we have related to the chaotic dynamics. For the initial condition plotted in green one there is not a clear limit.

Refer to caption
Figure 14: (a) Scaling of IPR for four initial conditions completely inside of the chaotic part of mixed phase space of the QKT (see Fig. 8). The golden dashed line is a straight line with a slope of one, just for reference. (b) Finite mass exponent for four initial conditions completely inside the chaotic part of mixed phase space (see Fig. 8). The horizontal golden(orange) dashed line corresponds to the asymptotic value found for chaotic(regular) dynamics. (c)-(f) Cumulative finite mass exponent for four initial conditions completely inside of the chaotic part of mixed phase space (see Fig. 8). The horizontal golden(orange) dashed line corresponds to the asymptotic value found for chaotic(regular) dynamics.

References

  • Koczor et al. [2019a] B. Koczor, R. Zeier, and S. J. Glaser, Continuous phase spaces and the time evolution of spins: star products and spin-weighted spherical harmonics, Journal of Physics A: Mathematical and Theoretical 52, 055302 (2019a).
  • Zhang and Batista [2021] H. Zhang and C. D. Batista, Classical spin dynamics based on SU(n)\mathrm{SU}(n) coherent states, Phys. Rev. B 104, 104409 (2021).
  • Miguel et al. [2025] M. Miguel, M. A. Bastarrachea-Magnani, and J. G. Hirsch,  (2025), in preparation.
  • Mandelbrot [1974] B. B. Mandelbrot, Intermittent turbulence in self-similar cascades: Divergence of high moments and dimension of the carrier, J. Fluid Mech. 62, 331 (1974).
  • Ott [2002] E. Ott, Chaos in Dynamical Systems, 2nd ed. (Cambridge University Press, 2002).
  • Mirlin [2000] A. D. Mirlin, Statistics of energy levels and eigenfunctions in disordered systems, Physics Reports 326, 259 (2000).
  • Nakayama and Yakubo [2013] T. Nakayama and K. Yakubo, Fractal Concepts in Condensed Matter Physics, 1st ed. (Springer Berlin, 2013).
  • Bastarrachea-Magnani et al. [2016] M. A. Bastarrachea-Magnani, B. López-del Carpio, J. Chávez-Carlos, S. Lerma-Hernández, and J. G. Hirsch, Delocalization and quantum chaos in atom-field systems, Phys. Rev. E 93, 022215 (2016).
  • Fritz Haake [2018] M. K. Fritz Haake, Sven Gnutzmann, Quantum Signatures of Chaos (Springer-Verlag, Berlin, 2018).
  • Santhanam et al. [2022] M. Santhanam, S. Paul, and J. B. Kannan, Quantum kicked rotor and its variants: Chaos, localization and beyond, Physics Reports 956, 1 (2022), quantum kicked rotor and its variants: Chaos, localization and beyond.
  • Kramer and MacKinnon [1993a] B. Kramer and A. MacKinnon, Localization: theory and experiment, Reports on Progress in Physics 56, 1469 (1993a).
  • Bell and Dean [1970] R. J. Bell and P. Dean, Atomic vibrations in vitreous silica, Discuss. Faraday Soc. 50, 55 (1970).
  • Thouless [1974] D. Thouless, Electrons in disordered systems and the theory of localization, Physics Reports 13, 93 (1974).
  • Evers and Mirlin [2008a] F. Evers and A. D. Mirlin, Anderson transitions, Rev. Mod. Phys. 80, 1355 (2008a).
  • Evers and Mirlin [2008b] F. Evers and A. D. Mirlin, Anderson transitions, Rev. Mod. Phys. 80, 1355 (2008b).
  • Chen et al. [2004] H. Chen, X. Sun, H. Chen, Z. Wu, and B. Wang, Some problems in multifractal spectrum computation using a statistical method, New Journal of Physics 6, 84 (2004).
  • Grassberger and Procaccia [1983] P. Grassberger and I. Procaccia, Characterization of strange attractors, Phys. Rev. Lett. 50, 346 (1983).
  • Grassberger and Procaccia [1984] P. Grassberger and I. Procaccia, Dimensions and entropies of strange attractors from a fluctuating dynamics approach, Physica D: Nonlinear Phenomena 13, 34 (1984).
  • Mirlin et al. [2006] A. D. Mirlin, Y. V. Fyodorov, A. Mildenberger, and F. Evers, Exact relations between multifractal exponents at the Anderson transition, Phys. Rev. Lett. 97, 046803 (2006).
  • Rodriguez et al. [2011] A. Rodriguez, L. J. Vasquez, K. Slevin, and R. A. Römer, Multifractal finite-size scaling and universality at the Anderson transition, Phys. Rev. B 84, 134209 (2011).
  • Roy et al. [2018a] S. Roy, I. M. Khaymovich, A. Das, and R. Moessner, Multifractality without fine-tuning in a Floquet quasiperiodic chain, SciPost Phys. 4, 025 (2018a).
  • Lindinger et al. [2019a] J. Lindinger, A. Buchleitner, and A. Rodríguez, Many-body multifractality throughout bosonic superfluid and Mott insulator phases, Phys. Rev. Lett. 122, 106603 (2019a).
  • Solórzano et al. [2021] A. Solórzano, L. F. Santos, and E. J. Torres-Herrera, Multifractality and self-averaging at the many-body localization transition, Phys. Rev. Research 3, L032030 (2021).
  • Sarkar et al. [2021] M. Sarkar, R. Ghosh, A. Sen, and K. Sengupta, Mobility edge and multifractality in a periodically driven Aubry-André model, Phys. Rev. B 103, 184309 (2021).
  • Sarkar et al. [2022] M. Sarkar, R. Ghosh, A. Sen, and K. Sengupta, Signatures of multifractality in a periodically driven interacting Aubry-André model, Phys. Rev. B 105, 024301 (2022).
  • Kramer and MacKinnon [1993b] B. Kramer and A. MacKinnon, Localization: theory and experiment, Reports on Progress in Physics 56, 1469 (1993b).
  • Chhabra and Jensen [1989] A. Chhabra and R. V. Jensen, Direct determination of the f(α\alpha) singularity spectrum, Phys. Rev. Lett. 62, 1327 (1989).
  • Lindinger et al. [2019b] J. Lindinger, A. Buchleitner, and A. Rodríguez, Many-body multifractality throughout bosonic superfluid and mott insulator phases, Phys. Rev. Lett. 122, 106603 (2019b).
  • Jeżewski [1998] W. Jeżewski, Multifractal and critical properties of the ising model, Phys. Rev. B 57, 10240 (1998).
  • Roy et al. [2018b] S. Roy, I. M. Khaymovich, A. Das, and R. Moessner, Multifractality without fine-tuning in a Floquet quasiperiodic chain, SciPost Phys. 4, 025 (2018b).
  • Bastarrachea-Magnani et al. [2024] M. A. Bastarrachea-Magnani, D. Villaseñor, J. Chávez-Carlos, S. Lerma-Hernández, L. F. Santos, and J. G. Hirsch, Quantum multifractality as a probe of phase space in the dicke model, Phys. Rev. E 109, 034202 (2024).
  • Vasquez et al. [2008] L. J. Vasquez, A. Rodriguez, and R. A. Römer, Multifractal analysis of the metal-insulator transition in the three-dimensional anderson model. i. symmetry relation under typical averaging, Phys. Rev. B 78, 195106 (2008).
  • Atas and Bogomolny [2012] Y. Y. Atas and E. Bogomolny, Multifractality of eigenfunctions in spin chains, Phys. Rev. E 86, 021104 (2012).
  • Beugeling et al. [2018] W. Beugeling, A. Bäcker, R. Moessner, and M. Haque, Statistical properties of eigenstate amplitudes in complex quantum systems, Phys. Rev. E 98, 022204 (2018).
  • Liu et al. [2024] Y. Liu, P. Sierant, P. Stornati, M. Lewenstein, and M. Płodzień, Quantum algorithms for inverse participation ratio estimation in multi-qubit and multi-qudit systems (2024), arXiv:2405.03338 [quant-ph] .
  • Robnik [1998] M. Robnik, Recent advances in quantum chaos of generic systems, Synergetics, edited by Axel Hutt and Hermann Haken (Springer US, 2020) pp , 133 (1998).
  • Wang and Robnik [2023a] Q. Wang and M. Robnik, Power-law decay of the fraction of the mixed eigenstates in kicked top model with mixed-type classical phase space, Phys. Rev. E 108, 054217 (2023a).
  • Bäcker et al. [2005] A. Bäcker, R. Ketzmerick, and A. G. Monastra, Flooding of chaotic eigenstates into regular phase space islands, Phys. Rev. Lett. 94, 054102 (2005).
  • Hentschel and Procaccia [1983] H. Hentschel and I. Procaccia, The infinite number of generalized dimensions of fractals and strange attractors, Phys. D: Nonlinear Phenom. 8, 435 (1983).
  • Halsey et al. [1986] T. C. Halsey, M. H. Jensen, L. P. Kadanoff, I. Procaccia, and B. I. Shraiman, Fractal measures and their singularities: The characterization of strange sets, Phys. Rev. A 33, 1141 (1986).
  • Janssen [1994] M. Janssen, Multifractal analysis of broadly-distributed observables at criticality, Int. J. Mod. Phys. B 08, 943 (1994).
  • Perelomov [1977] A. M. Perelomov, Generalized coherent states and some of their applications, Soviet Physics Uspekhi 20, 703 (1977).
  • Radcliffe [1971] J. M. Radcliffe, Some properties of coherent spin states, J. Phys. A: Gen. Phys 4 313 (1971).
  • Corps and Relaño [2022] A. L. Corps and A. Relaño, Dynamical and excited-state quantum phase transitions in collective systems, Phys. Rev. B 106, 024311 (2022).
  • Koczor et al. [2019b] B. Koczor, R. Zeier, and S. J. Glaser, Continuous phase spaces and the time evolution of spins: star products and spin-weighted spherical harmonics, Journal of Physics A: Mathematical and Theoretical 52, 055302 (2019b).
  • F. et al. [1987] H. F., K. M., and S. R., Classical and quantum chaos for a kicked top, Z. Phys. B Condensed Matter 65, 381–395 (1987).
  • Wang and Robnik [2021] Q. Wang and M. Robnik, Multifractality in quasienergy space of coherent states as a signature of quantum chaos, Entropy 2310.3390/e23101347 (2021).
  • Wang and Robnik [2023b] Q. Wang and M. Robnik, Statistics of phase space localization measures and quantum chaos in the kicked top model, Phys. Rev. E 107, 054213 (2023b).
  • Korsch et al. [1997] H. J. Korsch, B. Mirbach, and B. Schellhaaß, Semiclassical analysis of tunnelling splittings in periodically driven quantum systems, Journal of Physics A: Mathematical and General 30, 1659 (1997).
  • Lipkin et al. [1965] H. J. Lipkin, N. Meshkov, and A. Glick, Validity of many-body approximation methods for a solvable model: (i). exact solutions and perturbation theory, Nuclear Physics 62, 188 (1965).
  • Meshkov et al. [1965] N. Meshkov, A. J. Glick, and H. J. Lipkin, Validity of many-body approximation methods for a solvable model: (ii). linearization procedures, Nuclear Physics 62, 199 (1965).
  • Glick et al. [1965] A. J. Glick, H. J. Lipkin, and N. Meshkov, Validity of many-body approximation methods for a solvable model: (iii). diagram summations, Nuclear Physics 62, 211 (1965).
  • Chen et al. [2009] G. Chen, J. Q. Liang, and S. Jia, Interaction-induced lipkin-meshkov-glick model in a bose-einstein condensate inside an optical cavity, Opt. Express 17, 19682 (2009).
  • Sauerwein et al. [2023] N. Sauerwein, F. Orsi, P. Uhrich, S. Bandyopadhyay, F. Mattiotti, T. Cantat-Moltrecht, G. Pupillo, P. Hauke, and J.-P. Brantut, Engineering random spin models with atoms in a high-finesse cavity, Nature Physics 19, 1128 (2023).
  • Muniz et al. [2020] J. A. Muniz, D. Barberena, R. J. Lewis-Swan, D. J. Young, J. R. K. Cline, A. M. Rey, and J. K. Thompson, Exploring dynamical phase transitions with cold atoms in an optical  cavity, Nature 580, 602 (2020).
  • Morrison and Parkins [2008] S. Morrison and A. S. Parkins, Collective spin systems in dispersive optical cavity qed: Quantum phase transitions and entanglement, Phys. Rev. A 77, 043810 (2008).
  • Dusuel and Vidal [2004] S. Dusuel and J. Vidal, Finite-size scaling exponents of the Lipkin-meshkov-glick model, Phys. Rev. Lett. 93, 237204 (2004).
  • Castaños et al. [2006] O. Castaños, R. López-Peña, J. G. Hirsch, and E. López-Moreno, Classical and quantum phase transitions in the lipkin-meshkov-glick model, Phys. Rev. B 74, 104118 (2006).
  • Heiss [2006] W. D. Heiss, On the thermodynamic limit of the lipkin model, Journal of Physics A: Mathematical and General 39, 10081 (2006).
  • Ribeiro et al. [2007] P. Ribeiro, J. Vidal, and R. Mosseri, Thermodynamical limit of the lipkin-meshkov-glick model, Phys. Rev. Lett. 99, 050402 (2007).
  • Ribeiro et al. [2008] P. Ribeiro, J. Vidal, and R. Mosseri, Exact spectrum of the lipkin-meshkov-glick model in the thermodynamic limit and finite-size corrections, Phys. Rev. E 78, 021106 (2008).
  • Pilatowsky-Cameo et al. [2020] S. Pilatowsky-Cameo, J. Chávez-Carlos, M. A. Bastarrachea-Magnani, P. Stránský, S. Lerma-Hernández, L. F. Santos, and J. G. Hirsch, Positive quantum lyapunov exponents in experimental systems with a regular classical limit, Phys. Rev. E 101, 010202 (2020).
  • Muñoz Arias et al. [2023] M. H. Muñoz Arias, I. H. Deutsch, and P. M. Poggi, Phase-space geometry and optimal state preparation in quantum metrology with collective spins, PRX Quantum 4, 020314 (2023).
  • Benettin et al. [1986] G. Benettin, D. Casati, L. Galgani, A. Giorgilli, and L. Sironi, Apparent fractal dimensions in conservative dynamical systems, Physics Letters A 118, 325 (1986).
  • Cejnar et al. [2021] P. Cejnar, P. Stránský, M. Macek, and M. Kloc, Excited-state quantum phase transitions, Journal of Physics A: Mathematical and Theoretical 54, 133001 (2021).
  • Engelhardt et al. [2015] G. Engelhardt, V. M. Bastidas, W. Kopylov, and T. Brandes, Excited-state quantum phase transitions and periodic dynamics, Phys. Rev. A 91, 013631 (2015).
  • Kuś et al. [1987] M. Kuś, R. Scharf, and F. Haake, Symmetry versus degree of level repulsion for kicked quantum systems, Z. Phys. B Condensed Matter 66, 129-134 (1987).
  • Haake et al. [1987] F. Haake, M. Kuś, and R. Scharf, Quantum chaos for kicked spins, in Fundamentals of Quantum Optics II, edited by F. Ehlotzky (Springer Berlin Heidelberg, Berlin, Heidelberg, 1987) pp. 3–13.
  • Chaudhury et al. [2009] S. Chaudhury, A. Smith, B. E. Anderson, S. Ghose, and P. S. Jessen, Quantum signatures of chaos in a kicked top, Nature 461, 768 (2009).
  • Muñoz Arias et al. [2020] M. H. Muñoz Arias, P. M. Poggi, P. S. Jessen, and I. H. Deutsch, Simulating nonlinear dynamics of collective spins via quantum measurement and feedback, Phys. Rev. Lett. 124, 110503 (2020).
  • Swingle et al. [2016] B. Swingle, G. Bentsen, M. Schleier-Smith, and P. Hayden, Measuring the scrambling of quantum information, Phys. Rev. A 94, 040302 (2016).
  • Rozenbaum et al. [2017] E. B. Rozenbaum, S. Ganeshan, and V. Galitski, Lyapunov exponent and out-of-time-ordered correlator’s growth rate in a chaotic system, Phys. Rev. Lett. 118, 086801 (2017).
  • Altland and Haake [2012] A. Altland and F. Haake, Quantum chaos and effective thermalization, Phys. Rev. Lett. 108, 073601 (2012).
  • Neill et al. [2016] C. Neill, P. Roushan, M. Fang, Y. Chen, M. Kolodrubetz, Z. Chen, A. Megrant, R. Barends, B. Campbell, B. Chiaro, A. Dunsworth, E. Jeffrey, J. Kelly, J. Mutus, P. J. J. O’Malley, C. Quintana, D. Sank, A. Vainsencher, J. Wenner, T. C. White, A. Polkovnikov, and J. M. Martinis, Ergodic dynamics and thermalization in an isolated quantum system, Nature Physics 12, 1037 (2016).
  • Madhok et al. [2018] V. Madhok, S. Dogra, and A. Lakshminarayan, Quantum correlations as probes of chaos and ergodicity, Optics Communications 420, 189 (2018).
  • Ray et al. [2016] S. Ray, A. Ghosh, and S. Sinha, Quantum signature of chaos and thermalization in the kicked dicke model, Phys. Rev. E 94, 032103 (2016).
  • Sambe [1973] H. Sambe, Steady states and quasienergies of a quantum-mechanical system in an oscillating field, Phys. Rev. A 7, 2203 (1973).
  • Kus et al. [1988a] M. Kus, J. Mostowski, and F. Haake, Universality of eigenvector statistics of kicked tops of different symmetries, Journal of Physics A: Mathematical and General 21, L1073 (1988a).
  • Muñoz Arias et al. [2021] M. H. Muñoz Arias, P. M. Poggi, and I. H. Deutsch, Nonlinear dynamics and quantum chaos of a family of kicked pp-spin models, Phys. Rev. E 103, 052212 (2021).
  • Anand et al. [2024] A. Anand, J. Davis, and S. Ghose, Quantum recurrences in the kicked top, Phys. Rev. Res. 6, 023120 (2024).
  • Constantoudis and Theodorakopoulos [1997] V. Constantoudis and N. Theodorakopoulos, Lyapunov exponent, stretching numbers, and islands of stability of the kicked top, Phys. Rev. E 56, 5189 (1997).
  • Kus et al. [1988b] M. Kus, J. Mostowski, and F. Haake, Universality of eigenvector statistics of kicked tops of different symmetries, Journal of Physics A: Mathematical and General 21, L1073 (1988b).
  • D’Ariano et al. [1992] G. M. D’Ariano, L. R. Evangelista, and M. Saraceno, Classical and quantum structures in the kicked-top model, Phys. Rev. A 45, 3646 (1992).
  • del Castillo-Negrete et al. [1997] D. del Castillo-Negrete, J. Greene, and P. Morrison, Renormalization and transition to chaos in area preserving nontwist maps, Physica D: Nonlinear Phenomena 100, 311 (1997).
  • Muñoz Arias et al. [2022] M. H. Muñoz Arias, K. Chinni, and P. M. Poggi, Floquet time crystals in driven spin systems with all-to-all pp-body interactions, Phys. Rev. Res. 4, 023018 (2022).