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

Phase space formulation of the Abelian and non-Abelian quantum geometric tensor

Diego Gonzalez1,2, Daniel Gutiérrez-Ruiz1 and J. David Vergara1 1 Departamento de Física de Altas Energías, Instituto de Ciencias Nucleares, Universidad Nacional Autónoma de México, Apartado Postal 70-543, Ciudad de México, 04510, México 2 Departamento de Física, Cinvestav, Avenida Instituto Politécnico Nacional 2508, San Pedro Zacatenco, 07360, Gustavo A. Madero, Ciudad de México, México. dgonzalez@fis.cinvestav.mx daniel.gutierrez@correo.nucleares.unam.mx vergara@nucleares.unam.mx
Abstract

The geometry of the parameter space is encoded by the quantum geometric tensor, which captures fundamental information about quantum states and contains both the quantum metric tensor and the curvature of the Berry connection. We present a formulation of the Berry connection and the quantum geometric tensor in the framework of the phase space or Wigner function formalism. This formulation is obtained through the direct application of the Weyl correspondence to the geometric structure under consideration. In particular, we show that the quantum metric tensor can be computed using only the Wigner functions, which opens an alternative way to experimentally measure the components of this tensor. We also address the non-Abelian generalization and obtain the phase space formulation of the Wilczek-Zee connection and the non-Abelian quantum geometric tensor. In this case, the non-Abelian quantum metric tensor involves only the non-diagonal Wigner functions. Then, we verify our approach with examples and apply it to a system of NN coupled harmonic oscillators, showing that the associated Berry connection vanishes and obtaining the analytic expression for the quantum metric tensor. Our results indicate that the developed approach is well adapted to study the parameter space associated with quantum many-body systems.

, and

Keywords: Wigner formalism, Berry connection, Wilczek-Zee connection, Quantum geometric tensor

1 Introduction

The last decade has seen remarkable advances in understanding the geometric aspects of the parameter space in several systems of condensed matter physics [1, 2]. At the core of these developments is the quantum geometric tensor, which is a powerful tool to characterize the geometry of the eigenstates of Hamiltonians depending smoothly on external parameters. The symmetric part of this tensor gives rise to the quantum metric tensor on the system’s parameter manifold [3], whereas the antisymmetric part provides the curvature of the Berry connection, whose flux gives the (Berry) geometric phase [4]. In turn, these geometrical objects have been used independently to reveal the appearance of quantum phase transitions [5, 6, 7].

One of the obstacles to study the quantum geometric tensor in its original (Hamiltonian) formulation is the difficulty in determining analytically or perturbatively the wave function. Encouraged by this fact and previous works in the context of the holographic principle [8, 9, 10], in Ref [11] was proposed a path integral formulation of the quantum geometric tensor, which turns out to be advantageous in quantum field theories or systems where the exact solution is not attainable [12]. In the framework of the phase space or Wigner function formalism there is also a formulation of the Berry curvature [13]. This formulation sheds some light on the semiclassical relation between the Berry phase and the Hannay angle (see also Ref. [14] for an early work); however, it was motivated mainly by a conjecture made by the author that has not been proved [13].

The Wigner function formalism is not only a remarkable picture of quantum mechanics [15, 16], but also an appropriate framework for analyzing the correspondence between classical and quantum mechanics [17, 18]. This picture is based on the Wigner function [19], which is a quasidistribution in phase space that has found applications in several branches of physics [20]. Furthermore, a substantial advantage of the Wigner function is that it can be measured experimentally [21, 22]. This feature would be particularly important to allow, for instance, measurements of the quantum metric tensor. However, despite these attractive properties, there is no description of the quantum geometric tensor (or even of the quantum metric) in this formalism.

The purpose of this paper is to report a formulation of the Berry connection and the quantum geometric tensor in the framework of the Wigner function formalism. This formulation is achieved using the Weyl correspondence [23] and involves, in addition to the Wigner functions, certain complex functions in phase space variables, which allow us to express these geometrical structures of the parameter space as integrals over phase space. Moreover, in this work we also provide the phase space formulation of the generalized non-Abelian quantum geometric tensor [24] and the non-Abelian Berry connection or Wilczek-Zee connection [25], which are the corresponding geometrical structures that emerge in the case of a degenerate system. As in the Abelian case, the symmetric part of the non-Abelian quantum geometric tensor corresponds to a non-Abelian quantum metric tensor, while its antisymmetric part leads to a non-Abelian Berry curvature or Wilczek-Zee curvature. In particular, we shall see that the Abelian and non-Abelian quantum metric tensors can be computed using the Wigner functions purely, which opens a new line of attack on the problem of experimentally measuring the components of this tensor. To illustrate the use and validity of the new formulation, we present a well-known example of the literature, obtaining the expected results for the geometrical structures. As an application of this approach, we study the parameter space of a system of NN coupled harmonic oscillators, showing that the associated Berry connection vanishes and obtaining the analytic expression for the quantum metric tensor. Then, we consider the classical Hamiltonian corresponding to the above quantum system and obtain the classical analog [26] of the resulting quantum metric tensor. We find that, in the case where the transformation that diagonalizes the Hamiltonians does not depend on the system’s parameters, the resulting quantum metric tensor and its classical analog are the same, modulo the use of the Bohr-Sommerfeld quantization rule for action variables. We also apply this approach to a degenerate system composed of three coupled oscillators to determine the non-Abelian quantum metric tensor.

The outline of the paper is as follows. In Sec. 2, we present an overview of the fundamental structures underlying the parameter space and the Wigner-function formalism. In Sec. 3, we derive the phase space formulation of the Abelian and non-Abelian quantum geometric tensor, as well as the formulation of the Berry connection and the Wilczek-Zee connection. In Sec. 4, we compute the Berry connection and the quantum geometric tensor for the generalized harmonic oscillator, showing how the phase space formulation leads to familiar results. In Sec. 5, we apply the phase space approach to a system of NN coupled harmonic oscillators, which has been used in the study of entanglement entropy [27, 28]. In Sec. 6, we show how the phase space formulation works in the case of a degenerate system of three coupled oscillators. Finally, in Sec. 7 we present our conclusions.

2 Framework

In order to set up the notations and the main assumptions, we begin this paper by briefly reviewing some aspects of the local geometry of the parameter space and the Wigner function formalism.

2.1 Geometry of the parameter space

Consider a quantum system defined by a Hamiltonian 𝐇^(𝐪^,𝐩^;x)\hat{\bf{H}}(\hat{\bf{q}},\hat{\bf{p}};x), where 𝐪^={𝐪^a}\hat{\bf{q}}=\{\hat{\bf{q}}_{a}\} and 𝐩^={𝐩^a}\hat{\bf{p}}=\{\hat{\bf{p}}_{a}\} (a,b,=1,,Na,b,\dots\!=1,\dots,N) are respectively the position and momentum operators, and x(τ)={xi}x(\tau)=\{x^{i}\} (i,j,=1,,mi,j,\dots\!=1,\dots,m) is a set of mm real adiabatic parameters, i.e., slowly varying functions of time τ\tau. To simplify notation, x(τ)x(\tau) will be denoted as xx. If 𝐇^\hat{\bf{H}} has at least one orthonormal eigenvector |n(x)\lvert n(x)\rangle with nondegenerate eigenvalue En(x)E_{n}(x), then the Abelian quantum geometric tensor defined in the parameter space of the system is given by

Qij(n):=in|(1|nn|)|jn,\displaystyle Q^{(n)}_{ij}:=\langle\partial_{i}n\rvert\left(1-\lvert n\rangle\langle n\rvert\right)\lvert\partial_{j}n\rangle, (1)

where i=/xi\partial_{i}=\partial/\partial x^{i}. By construction, Eq (1) is invariant under the phase transformation [3]

|n(x)|n(x)=eαn(x)|n(x),\displaystyle\lvert n(x)\rangle\ \rightarrow\ \lvert n^{\prime}(x)\rangle={\rm e}^{\rmi\alpha_{n}(x)}\lvert n(x)\rangle, (2)

where αn\alpha_{n} is an arbitrary real function of the parameters, meaning that we can choose the basis |n\lvert n\rangle up to a U(1)U(1) gauge transformation (see also Ref. [29]). The relevance of this tensor lies in the fact that it provides the fundamental structures underlying the parameter space.

The (symmetric) real part of Eq. (1) gives the quantum metric tensor [3]

gij(n)=ReQij(n),\displaystyle g^{(n)}_{ij}={\rm Re}\,Q^{(n)}_{ij}, (3)

which defines the line element as dl2=gij(n)δxiδxjdl^{2}=g^{(n)}_{ij}\delta x^{i}\delta x^{j} and provides a distance between the two neighbor states |n(x)\lvert n(x)\rangle and |n(x+δx)\lvert n(x+\delta x)\rangle over the parameter space. Actually, the metric (3) stems from the leading term of the fidelity f=|n(x)|n(x+δx)|f=|\langle n(x)|n(x+\delta x)\rangle|, a measurement of the “similarity” between the states |n(x)\lvert n(x)\rangle and |n(x+δx)\lvert n(x+\delta x)\rangle.

The (antisymmetric) imaginary part of Eq. (1) yields the Berry curvature [30]

Fij(n)=2ImQij(n)=iAj(n)jAi(n),\displaystyle F^{(n)}_{ij}=-2\,{\rm Im}\,Q^{(n)}_{ij}=\partial_{i}A^{(n)}_{j}-\partial_{j}A^{(n)}_{i}, (4)

which defines the 2-form F(n)=12Fij(n)δxiδxjF^{(n)}=\frac{1}{2}F^{(n)}_{ij}\delta x^{i}\wedge\delta x^{j} in the parameter space and is associated with the (Abelian) Berry connection

Ai(n):=n|in.\displaystyle A^{(n)}_{i}:=\rmi\langle n|\partial_{i}n\rangle. (5)

It is worthy of noticing that Ai(n)(x)A^{(n)}_{i}(x) is real since n|in\langle n|\partial_{i}n\rangle is purely imaginary and that, under the gauge transformation (2), it changes as

Ai(n)Ai(n)=Ai(n)iαn,\displaystyle A^{(n)}_{i}\ \rightarrow\ A^{\prime(n)}_{i}=A^{(n)}_{i}-\partial_{i}\alpha_{n}, (6)

which is the transformation law for a genuine U(1)U(1) gauge connection. The integral of the 1-form A(n)=Ai(n)δxiA^{(n)}=A^{(n)}_{i}\delta x^{i} over a closed path CC in the parameter space or the flux of the 2-form F(n)F^{(n)} through a surface with boundary CC yields the gauge invariant Berry phase [30].

On the other hand, if the Hamiltonian 𝐇^\hat{\bf{H}} has a set of gng_{n} orthonormal eigenvectors |nI(x)\lvert n_{I}(x)\rangle (I,J,=1,2,,gnI,J,\dots\!=1,2,\dots,g_{n}) associated with the eigenvalue En(x)E_{n}(x), i.e., 𝐇^(x)|nI(x)=En(x)|nI(x)\hat{\bf{H}}(x)\lvert n_{I}(x)\rangle=E_{n}(x)\lvert n_{I}(x)\rangle, then the non-Abelian quantum geometric tensor defined in the parameter space of the system is given by [24]

QijIJ(n):=inI|(1K=1gn|nKnK|)|jnJ.\displaystyle Q^{(n)}_{ijIJ}:=\langle\partial_{i}n_{I}\rvert\left(1-\sum_{K=1}^{g_{n}}\lvert n_{K}\rangle\langle n_{K}\rvert\right)\lvert\partial_{j}n_{J}\rangle. (7)

It is easy to see that in the nondegenerate case gn=1g_{n}=1, the non-Abelian tensor (7) reduces to the Abelian one (1). Furthermore, under the unitary transformation

|nI(x)|nI(x)=J=1gn|nJ(x)UJI(x),\displaystyle\lvert n_{I}(x)\rangle\ \rightarrow\ \lvert n^{\prime}_{I}(x)\rangle=\sum_{J=1}^{g_{n}}\lvert n_{J}(x)\rangle U_{JI}(x), (8)

where UIJ(x)U_{IJ}(x) are the entries of a parameter-dependent unitary gn×gng_{n}\times g_{n} matrix U(x)U(x), the tensor (7) transforms covariantly

QijIJ(n)QijIJ(n)=K,L=1gnUKIQijKL(n)ULJ,\displaystyle Q^{(n)}_{ijIJ}\ \rightarrow\ Q^{\prime(n)}_{ijIJ}=\sum_{K,L=1}^{g_{n}}U^{*}_{KI}Q^{(n)}_{ijKL}U_{LJ}, (9)

where ‘’ stands for complex conjugation.

By analogy with Eqs. (3) and (4), the corresponding (symmetric) non-Abelian quantum metric tensor and the (antisymmetric) non-Abelian Berry curvature or Wilczek-Zee curvature are given by

gijIJ(n)=12(QijIJ(n)+QijJI(n)),\displaystyle g^{(n)}_{ijIJ}=\frac{1}{2}(Q^{(n)}_{ijIJ}+Q^{(n)*}_{ijJI}), (10)

and

FijIJ(n)=(QijIJ(n)QijJI(n)),\displaystyle F^{(n)}_{ijIJ}=\rmi(Q^{(n)}_{ijIJ}-Q^{(n)*}_{ijJI}), (11)

respectively. Moreover, the associated non-Abelian Berry connection or Wilczek-Zee connection is defined by [25, 31]

AiIJ(n):=nI|inJ.\displaystyle A^{(n)}_{iIJ}:=\rmi\langle n_{I}|\partial_{i}n_{J}\rangle. (12)

This connection provides the entries of a gn×gng_{n}\times g_{n} Hermitian matrix for each ii and, under the transformation (8), transforms as a proper gauge potential

AiIJ(n)AiIJ(n)=K,L=1gnUKIAiKL(n)ULJ+K=1gnUKIiUKJ.\displaystyle A^{(n)}_{iIJ}\ \rightarrow\ A^{\prime(n)}_{iIJ}=\sum_{K,L=1}^{g_{n}}U^{*}_{KI}A^{(n)}_{iKL}U_{LJ}+\rmi\sum_{K=1}^{g_{n}}U^{*}_{KI}\partial_{i}U_{KJ}. (13)

2.2 Wigner-function formalism

Given an operator 𝐐^(𝐪^,𝐩^;x)\hat{\bf{Q}}(\hat{\bf{q}},\hat{\bf{p}};x), the Weyl correspondence associates a quantum phase space function 𝒬(q,p;x)\mathcal{Q}(q,p;x) defined as [23]

𝒬(q,p;x)=dNyeipyq+y2|𝐐^(𝐪^,𝐩^;x)|qy2,\displaystyle\mathcal{Q}(q,p;x)=\int_{-\infty}^{\infty}{\rm d}^{N}y\,{\rm e}^{-\frac{{\rm i}p\cdot y}{\hbar}}\langle q+\tfrac{y}{2}\rvert\hat{\bf{Q}}(\hat{\bf{q}},\hat{\bf{p}};x)\lvert q-\tfrac{y}{2}\rangle, (14)

which is known as the Weyl transform of 𝐐^\hat{\bf{Q}}. Here, the variables q={qa}q=\{q_{a}\} and p={pa}p=\{p_{a}\} are respectively the eigenvalues of the operators 𝐪^\hat{\bf{q}} and 𝐩^\hat{\bf{p}}, and we have written dNydy1dyN{\rm d}^{N}\!y\equiv{\rm d}y_{1}\dotsi{\rm d}y_{N}, and pya=1Npayap\cdot y\equiv\sum_{a=1}^{N}p_{a}y_{a}. An important property of the Weyl correspondence is that it allows to write the trace of the product of two operators, 𝐐^\hat{\bf{Q}} and 𝐎^\hat{\bf{O}}, as

Tr(𝐐^𝐎^)=1(2π)NdNqdNp𝒬𝒪,\displaystyle{\rm Tr}(\hat{\bf{Q}}\hat{\bf{O}})=\frac{1}{(2\pi\hbar)^{N}}\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,\mathcal{Q}\mathcal{O}, (15)

where 𝒬\mathcal{Q} and 𝒪\mathcal{O} are the Weyl transform of these operators.

The Wigner function Wn(q,p;x)W_{n}(q,p;x), which is the main tool of this formalism, is defined as the function corresponding to the density operator ρ^(x)\hat{\rho}(x). More precisely, for a pure state, ρ^n(x)=|n(x)n(x)|\hat{\rho}_{n}(x)=\lvert n(x)\rangle\langle n(x)\rvert, it is given by [19]

Wn(q,p;x)=1(2π)NdNyeipyψn(q+y2;x)ψn(qy2;x),\displaystyle W_{n}(q,p;x)=\frac{1}{(2\pi\hbar)^{N}}\int_{-\infty}^{\infty}{\rm d}^{N}y\,{\rm e}^{-\frac{{\rm i}p\cdot y}{\hbar}}\psi_{n}(q+\tfrac{y}{2};x)\psi^{*}_{n}(q-\tfrac{y}{2};x), (16)

where ψn(q+y2;x)=q+y2|n(x)\psi_{n}(q+\tfrac{y}{2};x)=\langle q+\tfrac{y}{2}|n(x)\rangle and ψn(qy2;x)=n(x)|qy2\psi^{*}_{n}(q-\tfrac{y}{2};x)=\langle n(x)|q-\tfrac{y}{2}\rangle. Thus, the Wigner function provides a phase space representation of the quantum state |n(x)\lvert n(x)\rangle.

From Eq. (15) with 𝐐^ρ^n\hat{\bf{Q}}\rightarrow\hat{\rho}_{n}, it is straightforward to see that the expectation value of an operator 𝐎^\hat{\bf{O}} can be written as

𝐎^n=Tr(ρ^n𝐎^)=dNqdNpWn𝒪.\displaystyle\langle\hat{\bf{O}}\rangle_{n}={\rm Tr}(\hat{\rho}_{n}\,\hat{\bf{O}})=\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,W_{n}\,\mathcal{O}. (17)

This expression is reminiscent of the average of the function 𝒪\mathcal{O} (to which 𝐎^\hat{\bf{O}} corresponds) with respect to the phase space “quasiprobability” distribution WnW_{n}, which is real but not always positive. Notice that setting 𝐎^\hat{\bf{O}} equal to the identity operator 𝟙^\hat{\mathds{1}} in Eq. (17) and using the fact that the Weyl transform of 𝟙^\hat{\mathds{1}} is 11, we have

dNqdNpWn=1,\displaystyle\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,W_{n}=1, (18)

which shows that WnW_{n} is normalized in phase space. For later purposes, we also write down the following expression derived by taking 𝐎^ρ^n\hat{\bf{O}}\rightarrow\hat{\rho}_{n} in Eq. (17):

dNqdNpWn2=1(2π)N.\displaystyle\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,W_{n}^{2}=\frac{1}{(2\pi\hbar)^{N}}. (19)

For more details on the Wigner function, we refer the reader to Ref. [32].

3 Phase space formulation of the geometry of the parameter space

This section is devoted to recasting the geometrical structures involved in the parameter space into the Wigner function formalism; both the Abelian and non-Abelian cases are addressed. The approach we will follow here takes advantage of the Weyl correspondence and is different from that used in Ref. [13], where the Abelian and non-Abelian Berry curvatures are considered. In essence, the idea is very straightforward. It relies on the observation that these geometrical structures can be expressed as the trace of quantum operators, which will allow us to apply Eq. (15), and then link with the phase space formalism.

3.1 Abelian case

Let us first consider the Abelian Berry connection. We start by writing Eq. (5) as

Ai(n)=Tr𝐀^i(n),\displaystyle A_{i}^{(n)}={\rm Tr}\,\hat{\bf{A}}^{(n)}_{i}, (20)

where 𝐀^i(n)\hat{\bf{A}}^{(n)}_{i} is a quantum operator defined by

𝐀^i(n):=|inn|.\displaystyle\hat{\bf{A}}^{(n)}_{i}:=\rmi\lvert\partial_{i}n\rangle\langle n\rvert. (21)

Note that 𝐀^i(n)\hat{\bf{A}}^{(n)}_{i} is non-Hermitian, and then its Weyl transform

𝒜i(n)(q,p;x)\displaystyle\mathcal{A}_{i}^{(n)}(q,p;x) =\displaystyle= dNyeipyq+y2|inn|qy2\displaystyle\rmi\int_{-\infty}^{\infty}{\rm d}^{N}y\,{\rm e}^{-\frac{{\rm i}p\cdot y}{\hbar}}\langle q+\tfrac{y}{2}|\partial_{i}n\rangle\langle n|q-\tfrac{y}{2}\rangle (22)
=\displaystyle= dNyeipyiψn(q+y2;x)ψn(qy2;x),\displaystyle\rmi\int_{-\infty}^{\infty}{\rm d}^{N}y\,{\rm e}^{-\frac{{\rm i}p\cdot y}{\hbar}}\,\partial_{i}\psi_{n}(q\!+\!\tfrac{y}{2};x)\,\psi^{*}_{n}(q\!-\!\tfrac{y}{2};x),

is a complex function in phase space variables. Actually, by using Eqs. (16) and (22), it is not hard to show that

Im(𝒜i(n))=(2π)N2iWn.\displaystyle{\rm Im}(\mathcal{A}_{i}^{(n)})=\frac{(2\pi\hbar)^{N}}{2}\partial_{i}W_{n}. (23)

Having Eq. (20), we can now apply Eq. (15) to relate the phase space function 𝒜i(n)\mathcal{A}_{i}^{(n)} with the Berry connection. Indeed, taking 𝐐^𝟙^\hat{\bf{Q}}\rightarrow\hat{\mathds{1}} and 𝐎^𝐀^i(n)\hat{\bf{O}}\rightarrow\hat{\bf{A}}^{(n)}_{i} in Eq. (15) and using Eq. (20), we have

Ai(n)=1(2π)NdNqdNp𝒜i(n).\displaystyle A^{(n)}_{i}=\frac{1}{(2\pi\hbar)^{N}}\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,\mathcal{A}_{i}^{(n)}. (24)

This equation provides an expression of the Berry connection in the Wigner-function formalism. It can be checked that, under the gauge transformation (2), the function 𝒜i(n)\mathcal{A}_{i}^{(n)} transforms according to

𝒜i(n)𝒜(n)i=𝒜i(n)(2π)NWniαn,\displaystyle\mathcal{A}_{i}^{(n)}\ \rightarrow\ \mathcal{A}{{}^{\prime}}_{i}^{(n)}=\mathcal{A}_{i}^{(n)}-(2\pi\hbar)^{N}W_{n}\,\partial_{i}\alpha_{n}, (25)

from which, together with Eq (18), it follows that the connection (24) satisfies the transformation law (6), as expected. Notice that the transformation law for 𝒜i(n)\mathcal{A}_{i}^{(n)} is similar to the one of a U(1)U(1) gauge connection, but not identical because of the presence of WnW_{n} in Eq. (25). This is a consequence of the fact that, while the gauge transformation (2) is restricted to the parameter space, 𝒜i(n)(q,p;x)\mathcal{A}_{i}^{(n)}(q,p;x) is a mixed structure of the parameter space and the physical phase space. Of course, it is possible to consider a less restrictive and more general gauge transformation (see Ref. [29], for instance). It is also worth noting that only the real part of the phase space function 𝒜i(n)\mathcal{A}_{i}^{(n)} contributes to the Berry connection (24), which is a consequence of Eq. (23) and the fact that

dNqdNpiWn=0.\displaystyle\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,\partial_{i}W_{n}=0. (26)

To recast the Abelian quantum metric tensor into the Wigner-function formalism, here we follow the same procedure as in the case of the Berry connection. We observe that the quantum geometric tensor (1) can be written as

Qij(n)=Tr𝐐^ij(n),\displaystyle Q^{(n)}_{ij}={\rm Tr}\,\hat{\bf{Q}}^{(n)}_{ij}, (27)

where 𝐐^ij(n)\hat{\bf{Q}}^{(n)}_{ij} is an operator defined by

𝐐^ij(n):=(𝐀^i(n)𝐀^i(n))𝐀^j(n),\displaystyle\hat{\bf{Q}}^{(n)}_{ij}:=(\hat{\bf{A}}^{(n)\dagger}_{i}-\hat{\bf{A}}^{(n)}_{i})\hat{\bf{A}}^{(n)}_{j}, (28)

with 𝐀^i(n)\hat{\bf{A}}^{(n)}_{i} given by Eq.  (21). Therefore, noting that 𝐀^i(n)𝐀^i(n)=iρ^n\hat{\bf{A}}^{(n)\dagger}_{i}-\hat{\bf{A}}^{(n)}_{i}=-\rmi\partial_{i}\hat{\rho}_{n}, from Eq. (15) with 𝐐^iρ^n\hat{\bf{Q}}\rightarrow-\rmi\partial_{i}\hat{\rho}_{n} and 𝐎^𝐀^j(n)\hat{\bf{O}}\rightarrow\hat{\bf{A}}^{(n)}_{j} it follows that

Qij(n)=dNqdNpiWn𝒜j(n),\displaystyle Q^{(n)}_{ij}=-\rmi\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,\partial_{i}W_{n}\,\mathcal{A}_{j}^{(n)}, (29)

which, after using Eq. (23), becomes

Qij(n)=2(2π)NdNqdNpIm(𝒜i(n))𝒜j(n).\displaystyle Q^{(n)}_{ij}=-\frac{2\rmi}{(2\pi\hbar)^{N}}\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,{\rm Im}(\mathcal{A}_{i}^{(n)})\,\mathcal{A}_{j}^{(n)}. (30)

This is a formulation of the quantum metric tensor within the phase space formalism. Notice that we only need the knowledge of 𝒜i(n)\mathcal{A}_{i}^{(n)} in order to compute the quantum geometric tensor (30). This shows that the new phase space functions 𝒜i(n)\mathcal{A}_{i}^{(n)}, introduced here for calculating the Berry connection, actually encode all the relevant information embodied in the parameter space. On the other hand, it can be checked, by using Eq. (25) together with

dNqdNpWniWn=0,\displaystyle\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,W_{n}\partial_{i}W_{n}=0, (31)

which follows from Eq. (19), that Eq. (30) is gauge invariant under the transformation (2).

Separating Eq. (30) into its real and imaginary parts, we obtain the expression for Abelian quantum metric tensor

gij(n)=2(2π)NdNqdNpIm(𝒜i(n))Im(𝒜j(n)),\displaystyle g^{(n)}_{ij}=\frac{2}{(2\pi\hbar)^{N}}\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,{\rm Im}(\mathcal{A}_{i}^{(n)})\,{\rm Im}(\mathcal{A}_{j}^{(n)}), (32)

which, after using Eq. (23), can be written as

gij(n)=(2π)N2dNqdNpiWnjWn,\displaystyle g^{(n)}_{ij}=\frac{(2\pi\hbar)^{N}}{2}\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,\partial_{i}W_{n}\,\partial_{j}W_{n}, (33)

and the expression for the Abelian Berry curvature tensor

Fij(n)=4(2π)NdNqdNpIm(𝒜i(n))Re(𝒜j(n)).\displaystyle F^{(n)}_{ij}=\frac{4}{(2\pi\hbar)^{N}}\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,{\rm Im}(\mathcal{A}_{i}^{(n)})\,{\rm Re}(\mathcal{A}_{j}^{(n)}). (34)

Some remarks are in order. First, note that in contrast to the standard approach to determine the Abelian quantum metric tensor (namely, Eq. (3)), the phase space approach based on Eq. (33) does not require the knowledge of wave functions. This is because the Wigner functions can also be obtained as the solutions of suitable functional equations in phase space, bypassing the use of wave functions [16]. Second, note that the expression (34) for the Abelian Berry curvature is different from the one conjectured in Ref. [13], although both expressions must yield identical results. In fact, while Eq. (34) only involves the phase space function 𝒜i(n)\mathcal{A}_{i}^{(n)}, the expression proposed in [13], which uses Eq. (17) and takes elements of the curvature associated with the classical Hannay angle [33], requires the knowledge of variables q={qa}q=\{q_{a}\} and p={pa}p=\{p_{a}\} in terms of the angle-action variables. In this regard, it is worth recalling that the angle-action variables are restricted to classical integrable systems, and hence the expression of Ref. [13] can be applied only to those quantum systems whose classical counterpart is integrable, in contrast to Eq (34) which does not have that limitation. Furthermore, we would also like to point out that Eqs. (24) and (32)–(34) for the Abelian geometrical structures have not, to our knowledge, been reported in the literature. They are new and can be useful not only to gain more insight in the nature of the parameter space of quantum systems, but also for practical applications in a variety of quantum models. Third, notice that in Eq. (20) we could also consider the operator 𝐁^i(n):=(β1𝐀^i(n)+β2𝐀^i(n))/(β1+β2)\hat{\bf{B}}^{(n)}_{i}:=(\beta_{1}\hat{\bf{A}}^{(n)}_{i}+\beta_{2}\hat{\bf{A}}^{(n)\dagger}_{i})/(\beta_{1}+\beta_{2}) with β1\beta_{1} and β2\beta_{2} being arbitrary complex numbers, because of Ai(n)=Tr𝐁^i(n)A_{i}^{(n)}={\rm Tr}\,\hat{\bf{B}}^{(n)}_{i}. Nevertheless, a more complicated choice could be inconvenient for practical applications and does not give any further insight into the nature of the parameter space.

3.2 Non-Abelian case

Here, we extend the phase space description of the parameter space to include the non-Abelian geometrical structures. Following the procedure considered for the Abelian case, we begin by expressing the Wilczek-Zee connection as

AiIJ(n)=Tr𝐀^iIJ(n),\displaystyle A^{(n)}_{iIJ}={\rm Tr}\,\hat{\bf{A}}^{(n)}_{iIJ}, (35)

where now the associated quantum operator takes the form

𝐀^iIJ(n):=|inJnI|.\displaystyle\hat{\bf{A}}^{(n)}_{iIJ}:=\rmi\lvert\partial_{i}n_{J}\rangle\langle n_{I}\rvert. (36)

The Weyl transform for this operator is given by

𝒜iIJ(n)(q,p;x)\displaystyle\mathcal{A}_{iIJ}^{(n)}(q,p;x) =\displaystyle= dNyeipyq+y2|inJnI|qy2\displaystyle\rmi\int_{-\infty}^{\infty}{\rm d}^{N}y\,{\rm e}^{-\frac{{\rm i}p\cdot y}{\hbar}}\langle q+\tfrac{y}{2}|\partial_{i}n_{J}\rangle\langle n_{I}|q-\tfrac{y}{2}\rangle (37)
=\displaystyle= dNyeipyiψnJ(q+y2;x)ψnI(qy2;x),\displaystyle\rmi\int_{-\infty}^{\infty}{\rm d}^{N}y\,{\rm e}^{-\frac{{\rm i}p\cdot y}{\hbar}}\partial_{i}\psi_{nJ}(q+\tfrac{y}{2};x)\psi^{*}_{nI}(q-\tfrac{y}{2};x),

and, as in the Abelian case, it gives rise to a complex function in phase space variables, since 𝒜iIJ(n)\mathcal{A}^{(n)}_{iIJ} is a non-Hermitian operator.

By combining Eq. (35) and Eq. (15) with 𝐐^𝟙^\hat{\bf{Q}}\rightarrow\hat{\mathds{1}} and 𝐎^𝐀^iIJ(n)\hat{\bf{O}}\rightarrow\hat{\bf{A}}^{(n)}_{iIJ}, it is direct to see that the expression for the Wilczek-Zee connection in phase space formalism is

AiIJ(n)=1(2π)NdNqdNp𝒜iIJ(n),\displaystyle A^{(n)}_{iIJ}=\frac{1}{(2\pi\hbar)^{N}}\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,\mathcal{A}_{iIJ}^{(n)}, (38)

which is the natural generalization of the (Abelian) Berry connection (24). It is worth noting that Eq. (38) gives the entries of a gn×gng_{n}\times g_{n} Hermitian matrix for each ii, as it should be. This can be proved as follows. From Eq. (37) we have

𝒜iIJ(n)=𝒜iJI(n)+(2π)NiWnJI,\displaystyle\mathcal{A}_{iIJ}^{(n)}=\mathcal{A}_{iJI}^{(n)*}+\rmi(2\pi\hbar)^{N}\,\partial_{i}W_{nJI}, (39)

where WnIJW_{nIJ} are the non-diagonal Wigner functions

WnIJ:=1(2π)NdNyeipyψnI(q+y2;x)ψnJ(qy2;x).\displaystyle W_{nIJ}:=\frac{1}{(2\pi\hbar)^{N}}\int_{-\infty}^{\infty}{\rm d}^{N}y\,{\rm e}^{-\frac{{\rm i}p\cdot y}{\hbar}}\psi_{nI}(q+\tfrac{y}{2};x)\psi^{*}_{nJ}(q-\tfrac{y}{2};x). (40)

Then, plugging Eq. (39) into Eq. (38) and using the fact that

dNqdNpWnIJ=δIJ,\displaystyle\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,W_{nIJ}=\delta_{IJ}, (41)

it is straightforward to see that AiIJ(n)=AiJI(n)A^{(n)}_{iIJ}=A^{(n)*}_{iJI}.

Let us now prove that the connection (38) obeys the gauge transformation law (13) for the Wilczek-Zee connection. Under the unitary transformation (8), the phase space function 𝒜iIJ(n)\mathcal{A}_{iIJ}^{(n)} changes as

𝒜iIJ(n)𝒜iIJ(n)=K,L=1gn[UKI𝒜iKL(n)ULJ+(2π)NUKIWnLKiULJ],\displaystyle\mathcal{A}^{(n)}_{iIJ}\ \rightarrow\ \mathcal{A}^{\prime(n)}_{iIJ}=\sum_{K,L=1}^{g_{n}}\left[U^{*}_{KI}\mathcal{A}^{(n)}_{iKL}U_{LJ}+\rmi(2\pi\hbar)^{N}\,U^{*}_{KI}\,W_{nLK}\,\partial_{i}U_{LJ}\right], (42)

which generalizes Eq. (25). Note that while the second term on the right side of Eq. (25) is purely imaginary, the corresponding term in Eq. (42) is complex in general. Now, taking into account Eq (42), Eq (38) immediately implies

AiIJ(n)AiIJ(n)=K,L=1gn[UKIAiKL(n)ULJ+UKIiULJdNqdNpWnLK],\displaystyle A^{(n)}_{iIJ}\ \rightarrow\ A^{\prime(n)}_{iIJ}=\sum_{K,L=1}^{g_{n}}\left[U^{*}_{KI}A^{(n)}_{iKL}U_{LJ}+\rmi\,U^{*}_{KI}\partial_{i}U_{LJ}\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,W_{nLK}\,\right], (43)

which, after using Eq. (41), becomes the transformation law (13).

We now turn to the phase space formulation of the non-Abelian quantum geometric tensor. We find that Eq. (7) can be written as

QijIJ(n)=Tr𝐐^ijIJ(n),\displaystyle Q^{(n)}_{ijIJ}={\rm Tr}\,\hat{\bf{Q}}^{(n)}_{ijIJ}, (44)

where the operator 𝐐^ijIJ(n)\hat{\bf{Q}}^{(n)}_{ijIJ} is defined as

𝐐^ijIJ(n):=i𝔓^n𝐀^jIJ(n).\displaystyle\hat{\bf{Q}}^{(n)}_{ijIJ}:=-\rmi\,\partial_{i}\hat{\mathfrak{P}}_{n}\,\hat{\bf{A}}^{(n)}_{jIJ}. (45)

with 𝔓^n:=I=1gn|nInI|\hat{\mathfrak{P}}_{n}:=\sum_{I=1}^{g_{n}}\lvert n_{I}\rangle\langle n_{I}\rvert the projection operator. Then, taking 𝐐^i𝔓^n\hat{\bf{Q}}\rightarrow-\rmi\,\partial_{i}\hat{\mathfrak{P}}_{n} and 𝐎^𝐀^jIJ(n)\hat{\bf{O}}\rightarrow\hat{\bf{A}}^{(n)}_{jIJ} in Eq. (15), it is clear that Eq (44) is equivalent to

QijIJ(n)=dNqdNpK=1gniWnKK𝒜jIJ(n),\displaystyle Q^{(n)}_{ijIJ}=-\rmi\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,\sum_{K=1}^{g_{n}}\partial_{i}W_{nKK}\,\mathcal{A}_{jIJ}^{(n)}, (46)

Let us write this equation directly in terms of the phase space function 𝒜iIJ(n)\mathcal{A}_{iIJ}^{(n)}. Using Eqs. (37) and (40), it can be demonstrated that

Im(𝒜iKK(n))=(2π)N2iWnKK,\displaystyle{\rm Im}(\mathcal{A}_{iKK}^{(n)})=\frac{(2\pi\hbar)^{N}}{2}\partial_{i}W_{nKK}, (47)

which is the analog of Eq. (23). Thus, substituting Eq. (47) into Eq. (46), we get

QijIJ(n)=2(2π)NdNqdNpK=1gnIm(𝒜iKK(n))𝒜jIJ(n),\displaystyle Q^{(n)}_{ijIJ}=\frac{-2\rmi}{(2\pi\hbar)^{N}}\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,\sum_{K=1}^{g_{n}}{\rm Im}(\mathcal{A}_{iKK}^{(n)})\,\mathcal{A}_{jIJ}^{(n)}, (48)

which is an expression for the non-Abelian quantum geometric tensor in the phase space formalism. This equation tells us that, as in the Abelian case, all we need to know in order to compute the non-Abelian quantum geometric tensor is the phase space function 𝒜iIJ(n)\mathcal{A}_{iIJ}^{(n)}.

Now we are in a position to find the expressions for the non-Abelian quantum metric tensor and Wilczek-Zee curvature. From Eqs. (10) and (48), the non-Abelian quantum metric tensor turns out to be

gijIJ(n)=(2π)NdNqdNpK=1gnIm(𝒜iKK(n))(𝒜jIJ(n)𝒜jJI(n)),\displaystyle g^{(n)}_{ijIJ}=\frac{-\rmi}{(2\pi\hbar)^{N}}\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\sum_{K=1}^{g_{n}}{\rm Im}(\mathcal{A}_{iKK}^{(n)})(\mathcal{A}_{jIJ}^{(n)}-\mathcal{A}_{jJI}^{(n)*}), (49)

which, with the help of Eqs. (39) and (47), takes the form

gijIJ(n)=(2π)N2dNqdNpK=1gniWnKKjWnIJ,\displaystyle g^{(n)}_{ijIJ}=\frac{(2\pi\hbar)^{N}}{2}\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\sum_{K=1}^{g_{n}}\partial_{i}W_{nKK}\,\partial_{j}W_{nIJ}, (50)

whereas from Eqs. (11) and (48), the Wilczek-Zee curvature is recasted as

FijIJ(n)=2(2π)NdNqdNpK=1gnIm(𝒜iKK(n))(𝒜jIJ(n)+𝒜jJI(n)).\displaystyle F^{(n)}_{ijIJ}=\frac{2}{(2\pi\hbar)^{N}}\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\sum_{K=1}^{g_{n}}{\rm Im}(\mathcal{A}_{iKK}^{(n)})(\mathcal{A}_{jIJ}^{(n)}+\mathcal{A}_{jJI}^{(n)*}). (51)

We remark here that, as in the Abelian case, the expression (51) for the Wilczek-Zee curvature is different from the one proposed in Ref. [13]. In fact, whereas Eq. (51) only depends on the phase space function 𝒜iIJ(n)\mathcal{A}_{iIJ}^{(n)}, the expression of Ref. [13] requires the non-diagonal Wigner functions and the variables q={qa}q=\{q_{a}\} and p={pa}p=\{p_{a}\} in terms of the angle-action variables, meaning that it is restricted to quantum systems that have a classical integrable counterpar. It is also worth mentioning that the phase space formulation obtained in this subsection for the non-Abelian geometrical structures is new and can be used to deal with a wide variety of degenerate quantum systems.

The following sections are devoted to exemplify and apply this phase space approach to the parameter space.

4 Illustrative example

Let us take the archetypal system of a generalized harmonic oscillator, which is defined by the quantum Hamiltonian

𝐇^=12[X𝐪^2+Y(𝐪^𝐩^+𝐩^𝐪^)+Z𝐩^2],\displaystyle\hat{\bf{H}}=\frac{1}{2}\left[X\hat{\bf{q}}^{2}+Y(\hat{\bf{q}}\hat{\bf{p}}+\hat{\bf{p}}\hat{\bf{q}})+Z\hat{\bf{p}}^{2}\right], (52)

where x={xi}=(X,Y,Z)x=\{x^{i}\}=(X,Y,Z) (i,j,=1,2,3i,j,\dots\!=1,2,3) are the adiabatic parameters. The corresponding normalized wave functions are well known and take the form

ψn(q;x)=(ωZ)1/4χn(qωZ)exp(iYq22Z),\displaystyle\psi_{n}(q;x)=\left(\frac{\omega}{Z\hbar}\right)^{1/4}\chi_{n}\left(q\sqrt{\frac{\omega}{Z\hbar}}\right)\exp\left(-\frac{iYq^{2}}{2Z\hbar}\right), (53)

where nn are non-negative integers, ω=XZY2\omega=\sqrt{XZ-Y^{2}} is the parameter-dependent frequency which requires XZY2>0XZ-Y^{2}>0, and χn(ξ)=(2nn!π)1/2eξ2/2Hn(ξ)\chi_{n}(\xi)=\left(2^{n}n!\sqrt{\pi}\right)^{-1/2}{\rm e}^{-\xi^{2}/2}H_{n}(\xi) are the Hermite functions, with Hn(ξ)=(1)neξ2dndξneξ2H_{n}(\xi)=(-1)^{n}{\rm e}^{\xi^{2}}\frac{d^{n}}{d\xi^{n}}{\rm e}^{-\xi^{2}} being the Hermite polynomials.

Our first task is to obtain the associated phase space function 𝒜i(n)\mathcal{A}_{i}^{(n)}. Plugging Eq. (53) into Eq. (22) and using the relations

dχndξ=n2χn1n+12χn+1,\displaystyle\frac{d\chi_{n}}{d\xi}=\sqrt{\frac{n}{2}}\ \chi_{n-1}-\sqrt{\frac{n+1}{2}}\ \chi_{n+1}, (54a)
ξχn=n2χn1+n+12χn+1,\displaystyle\xi\,\chi_{n}=\sqrt{\frac{n}{2}}\ \chi_{n-1}+\sqrt{\frac{n+1}{2}}\ \chi_{n+1}, (54b)

it is straightforward to see that 𝒜i(n)\mathcal{A}_{i}^{(n)} is given by

𝒜i(n)=πZ2ω{i(ωZ)Ξn()+i(YZ)[Ξn(+)+(2n+1)Wn]},\displaystyle\mathcal{A}_{i}^{(n)}=\frac{\pi\hbar Z}{2\omega}\left\{\rmi\partial_{i}\left(\frac{\omega}{Z}\right)\Xi^{(-)}_{n}+\partial_{i}\left(\frac{Y}{Z}\right)\left[\Xi^{(+)}_{n}+(2n+1)W_{n}\right]\right\}, (55)

where WnW_{n} is the corresponding Wigner function [13],

Wn=(1)nπeλ/2Ln(λ),(λ:=4H/ω),\displaystyle W_{n}=\frac{(-1)^{n}}{\pi\hbar}{\rm e}^{-\lambda/2}L_{n}(\lambda),\qquad(\lambda:=4H/\hbar\omega), (56)

with H=(1/2)(Xq2+2Ypq+Zp2)H=(1/2)\left(Xq^{2}+2Ypq+Zp^{2}\right) and Ln(λ)L_{n}(\lambda) the Laguerre polynomials. Furthermore, we have defined the phase space functions

Ξn(±):=n(n1)fn2,n±(n+1)(n+2)fn+2,n,\displaystyle\Xi^{(\pm)}_{n}:=\sqrt{n(n-1)}f_{n-2,n}\pm\sqrt{(n+1)(n+2)}f_{n+2,n}, (57)

where

fn±2,n=12πdyeipyψn±2(q+y2;x)ψn(qy2;x),\displaystyle f_{n\pm 2,n}=\frac{1}{2\pi\hbar}\int_{-\infty}^{\infty}{\rm d}y\,{\rm e}^{-\frac{{\rm i}p\cdot y}{\hbar}}\psi_{n\pm 2}(q+\tfrac{y}{2};x)\psi^{*}_{n}(q-\tfrac{y}{2};x), (58)

for n20n-2\geq 0, are the non-diagonal Wigner functions, which explicitly read

fn2,n\displaystyle f_{n-2,n} =\displaystyle= (1)n+12(PωQ)2n(n1)π2ωeλ2Ln2(2)(λ),\displaystyle\frac{(-1)^{n+1}2\left(P-\rmi\omega Q\right)^{2}}{\sqrt{n(n-1)}\,\pi\hbar^{2}\omega}{\rm e}^{-\frac{\lambda}{2}}L^{(2)}_{n-2}(\lambda), (59a)
fn+2,n\displaystyle f_{n+2,n} =\displaystyle= (1)n+12(P+ωQ)2(n+1)(n+2)π2ωeλ2Ln(2)(λ).\displaystyle\frac{(-1)^{n+1}2\left(P+\rmi\omega Q\right)^{2}}{\sqrt{(n+1)(n+2)}\,\pi\hbar^{2}\omega}{\rm e}^{-\frac{\lambda}{2}}L^{(2)}_{n}(\lambda). (59b)

Here, Ln(α)(λ)L^{(\alpha)}_{n}(\lambda) are the associated Laguerre polynomials (Ln(0)(λ)=Ln(λ)L^{(0)}_{n}(\lambda)=L_{n}(\lambda)) and we have used the transformation

Q=1Zq,P=Z(p+YZq).\displaystyle Q=\frac{1}{\sqrt{Z}}q,\qquad P=\sqrt{Z}\left(p+\frac{Y}{Z}q\right). (60)

With this at hand, Eq. (24) can now be readily applied to Eq. (55). By doing so, and using Eq. (18) together with the fact that [16]

dqdpfn±2,n=0,\displaystyle\int_{-\infty}^{\infty}{\rm d}q\,{\rm d}p\,f_{n\pm 2,n}=0, (61)

we arrive at the Berry connection

A1(n)(x)=0,A2(n)(x)=(n+12)12ω,A3(n)(x)=(n+12)Y2Zω.\displaystyle A_{1}^{(n)}(x)=0,\qquad A_{2}^{(n)}(x)=\left(n+\frac{1}{2}\right)\frac{1}{2\omega},\qquad A_{3}^{(n)}(x)=-\left(n+\frac{1}{2}\right)\frac{Y}{2Z\omega}. (62)

This result coincides with the usual expression found in the literature (see Ref [26], for instance). Notice that as a consequence of Eq. (61), only the term involving WnW_{n} in Eq. (55) contributes to the non-zero Berry connection (62). This means that for the standard Harmonic oscillator (Y=0Y=0), in whose case Eq. (55) reduces to 𝒜i(n)=πZ2ωΞn()iXZ\mathcal{A}_{i}^{(n)}=\frac{\rmi\pi\hbar Z}{2\omega}\Xi^{(-)}_{n}\partial_{i}\sqrt{\frac{X}{Z}}, the resulting Berry connection is zero, as expected. Regarding the Berry curvature, it can be computed directly from Eq. (4) with Eq. (62) and the result is the well-known expression

F12(n)(x)=(n+12)Z4ω3,F13(n)(x)=(n+12)Y4ω3,\displaystyle F_{12}^{(n)}(x)=-\left(n+\frac{1}{2}\right)\frac{Z}{4\omega^{3}},\qquad F_{13}^{(n)}(x)=\left(n+\frac{1}{2}\right)\frac{Y}{4\omega^{3}},
F23(n)(x)=(n+12)X4ω3.\displaystyle F_{23}^{(n)}(x)=-\left(n+\frac{1}{2}\right)\frac{X}{4\omega^{3}}. (63)

We now focus on the computation of the quantum metric tensor, which can be done either from Eq. (32) with the phase space function (55) or from Eq. (33) with the Wigner function (56). Indeed, taking the imaginary part of Eq. (55) and using the relation

Ln(2)Ln2(2)=Ln2Lnλ,\displaystyle L_{n}^{(2)}-L_{n-2}^{(2)}=L_{n}-2\frac{\partial L_{n}}{\partial\lambda}, (64)

or directly calculating the partial derivatives of Eq (56) with respect to the parameters, it follows that

(iWn)q,p\displaystyle(\partial_{i}W_{n})_{q,p} =\displaystyle= 1πIm(𝒜i(n))\displaystyle\frac{1}{\pi\hbar}{\rm Im}(\mathcal{A}_{i}^{(n)}) (65)
=\displaystyle= (1)n2πeλ/2(Ln2Lnλ)(iλ)q,p.\displaystyle-\frac{(-1)^{n}}{2\pi\hbar}{\rm e}^{-\lambda/2}\,\left(L_{n}-2\frac{\partial L_{n}}{\partial\lambda}\right)(\partial_{i}\lambda)_{q,p}.

Then, substituting Eq. (65) into Eq. (33), and taking into account

0dλeλλ2(Ln(λ)2Ln(λ)λ)2=2(n2+n+1),\displaystyle\int_{0}^{\infty}{\rm d}\lambda\,{\rm e}^{-\lambda}\,\lambda^{2}\,\left(L_{n}(\lambda)-2\frac{\partial L_{n}(\lambda)}{\partial\lambda}\right)^{2}=2(n^{2}+n+1), (66)

we arrive at

gij(n)(x)=n2+n+132ω4(Z22YZ2Y2XZ2YZ4XZ2XY2Y2XZ2XYX2),\displaystyle g^{(n)}_{ij}(x)=\frac{n^{2}+n+1}{32\omega^{4}}\left(\begin{array}[]{ccc}Z^{2}&-2YZ&2Y^{2}-XZ\\ -2YZ&4XZ&-2XY\\ 2Y^{2}-XZ&-2XY&X^{2}\end{array}\right), (70)

which is the same expression for the quantum metric tensor as that obtained in Ref. [26] by using Eq. (3) with the wave function (53). This corroborates that the phase space formulation of the quantum metric tensor, namely Eq. (33), yields the right results.

5 NN coupled harmonic oscillators

We now want to go further and extend our analysis of the parameter space to the case of NN coupled harmonic oscillators. Specifically, the system under consideration is described by the Hamiltonian

𝐇^=12a=1N𝐩^a2+12a,b=1NKab𝐪^a𝐪^b,\displaystyle\hat{\bf{H}}=\frac{1}{2}\sum_{a=1}^{N}\hat{\bf{p}}_{a}^{2}+\frac{1}{2}\sum_{a,b=1}^{N}K_{ab}\hat{\bf{q}}_{a}\hat{\bf{q}}_{b}, (71)

where Kab(=Kba)K_{ab}(=K_{ba}) are the entries of an N×NN\times N symmetric and positive-definite matrix K(x)K(x), which is constructed out of the adiabatic parameters x={xi}x=\{x^{i}\} (i,j,=1,,ni,j,\dots\!=1,\dots,n). This system has been employed to study the properties of entanglement entropy [27, 28], with remarkable consequences.

It is convenient to start by introducing the linear transformation

𝐐^a=b=1NUab𝐪^b,𝐏^a=b=1NUab𝐩^b,\displaystyle\hat{\bf{Q}}_{a}=\sum_{b=1}^{N}U_{ab}\hat{\bf{q}}_{b},\qquad\hat{\bf{P}}_{a}=\sum_{b=1}^{N}U_{ab}\hat{\bf{p}}_{b}, (72)

where UabU_{ab} are the entries of an N×NN\times N orthogonal matrix UU such that K=UTΩ2UK=U^{T}\Omega^{2}U, with Ω=diag(ω1,,ωN)\Omega={\rm diag}(\omega_{1},\dots,\omega_{N}) being a diagonal matrix whose elements ωa\omega_{a} are the frequencies of the system. Using Eq. (72), the Hamiltonian (71) becomes

𝐇^=12a=1N(𝐏^a2+ωa2𝐐^a2),\displaystyle\hat{\bf{H}}=\frac{1}{2}\sum_{a=1}^{N}\left(\hat{\bf{P}}_{a}^{2}+\omega_{a}^{2}\hat{\bf{Q}}_{a}^{2}\right), (73)

which enables us to write the associated normalized wave function as

ψn1,,nN(q1,,qN;x)=a=1Nψna(Qa;x),\displaystyle\psi_{n_{1},\dots,n_{N}}(q_{1},\dots,q_{N};x)=\prod_{a=1}^{N}\psi_{n_{a}}(Q_{a};x), (74)

where ψna(Qa;x)\psi_{n_{a}}(Q_{a};x) is the wave function of the aath uncoupled oscillator with quantum number nan_{a}:

ψna(Qa;x)=(ωa)1/4χna(Qaωa).\displaystyle\psi_{n_{a}}(Q_{a};x)=\left(\frac{\omega_{a}}{\hbar}\right)^{1/4}\chi_{n_{a}}\left(Q_{a}\sqrt{\frac{\omega_{a}}{\hbar}}\right). (75)

Here we restrict ourselves to the case where the eigenvalues

En1,,nN=a=1N(na+12)ωa,\displaystyle E_{n_{1},\dots,n_{N}}=\sum_{a=1}^{N}\left(n_{a}+\frac{1}{2}\right)\hbar\omega_{a}, (76)

are nondegenerate, which ensures that we stay in the Abelian setting.

With the wave function at hand, we can apply Eq. (22) to get the corresponding phase space function 𝒜i(n1,,nN)\mathcal{A}_{i}^{(n_{1},\cdots,n_{N})}. Indeed, taking into account Eqs. (54a) and (54b), and making the change of variables

Ya=b=1NUabyb,\displaystyle Y_{a}=\sum_{b=1}^{N}U_{ab}\,y_{b}, (77)

it is relatively straightforward to show that Eq. (22) with Eq (74) leads to

𝒜i(n1,,nN)\displaystyle\mathcal{A}_{i}^{(n_{1},\dots,n_{N})} =\displaystyle= (2π)N(14a=1NiωaωaΞnab=1,baNWnb\displaystyle\rmi(2\pi\hbar)^{N}\Bigg{(}\frac{1}{4}\sum_{a=1}^{N}\frac{\partial_{i}\omega_{a}}{\omega_{a}}\,\Xi_{n_{a}}\prod_{\begin{subarray}{c}b=1,\\ b\neq a\end{subarray}}^{N}W_{n_{b}} (78)
+a,b,c=1NωaωciUabUcbΘna(+)Θnc()d=1,da,cNWnd),\displaystyle+\sum_{a,b,c=1}^{N}\sqrt{\frac{\omega_{a}}{\omega_{c}}}\,\partial_{i}U_{ab}\,U_{cb}\,\Theta^{(+)}_{n_{a}}\Theta^{(-)}_{n_{c}}\prod_{\begin{subarray}{c}d=1,\\ d\neq a,c\end{subarray}}^{N}W_{n_{d}}\Bigg{)},

where WnaW_{n_{a}} are phase space functions given by

Wna=(1)naπeλa2Lna(λa),(λa:=4Ha/ωa),\displaystyle W_{n_{a}}=\frac{(-1)^{n_{a}}}{\pi\hbar}{\rm e}^{-\frac{\lambda_{a}}{2}}L_{n_{a}}(\lambda_{a}),\qquad(\lambda_{a}:=4H_{a}/\hbar\omega_{a}), (79)

with Ha=(1/2)(Pa2+ωa2Qa2)H_{a}=(1/2)(P_{a}^{2}+\omega_{a}^{2}Q_{a}^{2}) and so that

Wn1,,nN(q1,,qN,p1,,qN;x)=a=1NWna(Qa,Pa;x),\displaystyle W_{n_{1},\dots,n_{N}}(q_{1},\dots,q_{N},p_{1},\dots,q_{N};x)=\prod_{a=1}^{N}W_{n_{a}}(Q_{a},P_{a};x), (80)

is the Wigner function of the system. Furthermore, the functions Ξna\Xi_{n_{a}} are defined by

Ξna:=na(na1)fna2,na(na+1)(na+2)fna+2,na,\displaystyle\Xi_{n_{a}}:=\sqrt{n_{a}(n_{a}-1)}f_{n_{a}-2,n_{a}}-\sqrt{(n_{a}+1)(n_{a}+2)}f_{n_{a}+2,n_{a}}, (81)

with the non-diagonal Wigner functions

fna±2,na:=12πdYaeiPaYaψna±2(Qa+Ya2;x)×ψna(QaYa2;x),\displaystyle f_{n_{a}\pm 2,n_{a}}:=\frac{1}{2\pi\hbar}\int_{-\infty}^{\infty}{\rm d}Y_{a}\,{\rm e}^{-\frac{{\rm i}P_{a}Y_{a}}{\hbar}}\psi_{n_{a}\pm 2}(Q_{a}+\tfrac{Y_{a}}{2};x)\times\psi_{n_{a}}(Q_{a}-\tfrac{Y_{a}}{2};x), (82)

which explicitly read

fna2,na=(1)na+12(PaωaQa)2na(na1)π2ωaeλa2Lna2(2)(λa),\displaystyle f_{n_{a}-2,n_{a}}=\frac{(-1)^{n_{a}+1}2\,(P_{a}-\rmi\omega_{a}Q_{a})^{2}}{\sqrt{n_{a}(n_{a}-1)}\,\pi\hbar^{2}\omega_{a}}{\rm e}^{\frac{-\lambda_{a}}{2}}L^{(2)}_{n_{a}-2}(\lambda_{a}), (83a)
fna+2,na=(1)na+12(Pa+ωaQa)2(na+1)(na+2)π2ωaeλa2Lna(2)(λa),\displaystyle f_{n_{a}+2,n_{a}}=\frac{(-1)^{n_{a}+1}2\,(P_{a}+\rmi\omega_{a}Q_{a})^{2}}{\sqrt{(n_{a}+1)(n_{a}+2)}\,\pi\hbar^{2}\omega_{a}}{\rm e}^{\frac{-\lambda_{a}}{2}}L^{(2)}_{n_{a}}(\lambda_{a}), (83b)

whereas the functions Θna(±)\Theta^{(\pm)}_{n_{a}} have the form

Θna(±)=na2fna1,na±na+12fna+1,na,\displaystyle\Theta^{(\pm)}_{n_{a}}=\sqrt{\frac{n_{a}}{2}}\,f_{n_{a}-1,n_{a}}\pm\sqrt{\frac{n_{a}+1}{2}}f_{n_{a}+1,n_{a}}, (84)

where fna±1,naf_{n_{a}\pm 1,n_{a}} have the same definition as in Eq. (82), but with na±2n_{a}\pm 2 replaced by na±1n_{a}\pm 1, and explicitly read

fna1,na=(1)na+121/2(Pa+ωaQa)na1/2π3/2ωa1/2eλa2Lna1(1)(λa),\displaystyle f_{n_{a}-1,n_{a}}=\frac{(-1)^{n_{a}+1}2^{1/2}(\rmi P_{a}+\omega_{a}Q_{a})}{n_{a}^{1/2}\pi\hbar^{3/2}\omega_{a}^{1/2}}{\rm e}^{\frac{-\lambda_{a}}{2}}L^{(1)}_{n_{a}-1}(\lambda_{a}), (85a)
fna+1,na=(1)na+121/2(PaωaQa)(na+1)1/2π3/2ωa1/2eλa2Lna(1)(λa).\displaystyle f_{n_{a}+1,n_{a}}=\frac{(-1)^{n_{a}+1}2^{1/2}(\rmi P_{a}-\omega_{a}Q_{a})}{(n_{a}+1)^{1/2}\,\pi\hbar^{3/2}\omega_{a}^{1/2}}{\rm e}^{\frac{-\lambda_{a}}{2}}L^{(1)}_{n_{a}}(\lambda_{a}). (85b)

It should be noted that for systems where the matrix UU does not depend on the parameters (iUab=0\partial_{i}U_{ab}=0), the terms of the second line of Eq. (78) vanish and hence the function 𝒜i(n1,,nN)\mathcal{A}_{i}^{(n_{1},\dots,n_{N})} reduces to 𝒜i(n1,,nN)=(2π)N4a=1NiωaωaΞnab=1,baNWnb\mathcal{A}_{i}^{(n_{1},\dots,n_{N})}=\frac{\rmi(2\pi\hbar)^{N}}{4}\sum_{a=1}^{N}\frac{\partial_{i}\omega_{a}}{\omega_{a}}\,\Xi_{n_{a}}\prod_{\begin{subarray}{c}b=1,\\ b\neq a\end{subarray}}^{N}W_{n_{b}}. This is the case, for instance, of the system of two symmetric coupled oscillators, which is described by the Hamiltonian

𝐇^=12[𝐩^12+𝐩^22+k(𝐪^12+𝐪^22)+k(𝐪^1𝐪^2)2],\displaystyle\hat{\bf{H}}=\frac{1}{2}\left[\hat{\bf{p}}_{1}^{2}+\hat{\bf{p}}_{2}^{2}+k(\hat{\bf{q}}_{1}^{2}+\hat{\bf{q}}_{2}^{2})+k^{\prime}(\hat{\bf{q}}_{1}-\hat{\bf{q}}_{2})^{2}\right], (86)

with the adiabatic parameters {xi}=(k,k)\{x^{i}\}=(k,k^{\prime}), and is of interest in the study of quantum entanglement [28, 34] and circuit complexity in quantum field theory [35]. Indeed, for such a system the entries of the matrix UU are given by U11=U12=U21=U22=1/2U_{11}=U_{12}=U_{21}=-U_{22}=1/\sqrt{2} and the function 𝒜i(n1,n2)\mathcal{A}_{i}^{(n_{1},n_{2})} takes the simple form 𝒜i(n1,n2)=(π)2(iω1ω1Ξn1Wn2+iω2ω2Ξn2Wn1)\mathcal{A}_{i}^{(n_{1},n_{2})}=\rmi(\pi\hbar)^{2}\left(\frac{\partial_{i}\omega_{1}}{\omega_{1}}\Xi_{n_{1}}W_{n_{2}}+\frac{\partial_{i}\omega_{2}}{\omega_{2}}\Xi_{n_{2}}W_{n_{1}}\right) where ω1=k\omega_{1}=\sqrt{k} and ω2=k+2k\omega_{2}=\sqrt{k+2k^{\prime}}.

Continuing with the general analysis, the associated Berry connection can be obtained directly from Eq. (24) with Eq. (78) and it turns out to be

Ai(n1,,nN)(x)=0,\displaystyle A_{i}^{(n_{1},\dots,n_{N})}(x)=0, (87)

where we have used the change of variables (72) and

dQadPaWna=1,\displaystyle\int_{-\infty}^{\infty}{\rm d}Q_{a}\,{\rm d}P_{a}\,W_{n_{a}}=1, (88a)
dQadPafna±1,na=0,\displaystyle\int_{-\infty}^{\infty}{\rm d}Q_{a}\,{\rm d}P_{a}\,f_{n_{a}\pm 1,n_{a}}=0, (88b)
dQadPafna±2,na=0.\displaystyle\int_{-\infty}^{\infty}{\rm d}Q_{a}\,{\rm d}P_{a}\,f_{n_{a}\pm 2,n_{a}}=0. (88c)

Clearly, Eq. (87) entails the vanishing of the corresponding Berry curvature

Fij(n1,,nN)(x)=0.\displaystyle F_{ij}^{(n_{1},\dots,n_{N})}(x)=0. (89)

This result is nontrivial and indicates that, for those systems described by a Hamiltonian of the form (71), the (Abelian) Berry phase is zero for arbitrary quantum numbers n1,,nNn_{1},\dots,n_{N} and nondegenerate eigenvalues.

Now we move to the calculation of the Abelian quantum metric tensor. The derivative of the Wigner function (80) with respect to the parameters leads to

(iWn1,n2)q,p\displaystyle(\partial_{i}W_{n_{1},n_{2}})_{q,p} =2(2π)NIm(𝒜i(n1,n2))\displaystyle=\frac{2}{(2\pi\hbar)^{N}}{\rm Im}(\mathcal{A}_{i}^{(n_{1},n_{2})}) (90)
=a=1N(1)na+12πeλa/2[Lna(λa)2λaLna(λa)](iλa)q,pb=1,baNWnb,\displaystyle=\sum_{a=1}^{N}\frac{(-1)^{n_{a}+1}}{2\pi\hbar}{\rm e}^{-\lambda_{a}/2}\bigg{[}L_{n_{a}}(\lambda_{a})\!-\!2\frac{\partial}{\partial\lambda_{a}}L_{n_{a}}(\lambda_{a})\bigg{]}(\partial_{i}\lambda_{a})_{q,p}\prod_{\begin{subarray}{c}b=1,\\ b\neq a\end{subarray}}^{N}W_{n_{b}},

where

(iλa)q,p=2ωa[(Pa2ωa2Qa2)iωaωa2b,c=1N(PaPc+ωa2QaQc)UcbiUab].\displaystyle(\partial_{i}\lambda_{a})_{q,p}=-\frac{2}{\hbar\omega_{a}}\bigg{[}(P_{a}^{2}\!-\!\omega_{a}^{2}Q_{a}^{2})\frac{\partial_{i}\omega_{a}}{\omega_{a}}-2\sum_{b,c=1}^{N}\left(P_{a}P_{c}+\omega_{a}^{2}Q_{a}Q_{c}\right)U_{cb}\partial_{i}U_{ab}\bigg{]}. (91)

After substituting this result into Eq. (33), and bearing in mind that

0dλaeλa(Lna(λa))2=1,\displaystyle\int_{0}^{\infty}{\rm d}\lambda_{a}{\rm e}^{-\lambda_{a}}(L_{n_{a}}(\lambda_{a}))^{2}=1, (92a)
0dλaeλaλa2(Lna(λa)2Lna(λa)λa)2=2(na2+na+1),\displaystyle\int_{0}^{\infty}{\rm d}\lambda_{a}{\rm e}^{-\lambda_{a}}\lambda_{a}^{2}\left(L_{n_{a}}(\lambda_{a})-2\frac{\partial L_{n_{a}}(\lambda_{a})}{\partial\lambda_{a}}\right)^{2}=2(n_{a}^{2}+n_{a}+1), (92b)
0dλaeλaλa(Lna(λa))2=2na+1,\displaystyle\int_{0}^{\infty}{\rm d}\lambda_{a}\,{\rm e}^{-\lambda_{a}}\,\lambda_{a}(L_{n_{a}}\,(\lambda_{a}))^{2}=2n_{a}+1, (92c)
0dλaeλaλa(Lna(λa)2Lna(λa)λa)2=2na+1,\displaystyle\int_{0}^{\infty}{\rm d}\lambda_{a}\,{\rm e}^{-\lambda_{a}}\lambda_{a}\left(L_{n_{a}}(\lambda_{a})-2\frac{\partial L_{n_{a}}(\lambda_{a})}{\partial\lambda_{a}}\right)^{2}=2n_{a}+1, (92d)
0dλaeλaλaLna(λa)(Lna(λa)2Lna(λa)λa)=1,\displaystyle\int_{0}^{\infty}{\rm d}\lambda_{a}\,{\rm e}^{-\lambda_{a}}\lambda_{a}\,L_{n_{a}}(\lambda_{a})\left(L_{n_{a}}(\lambda_{a})-2\frac{\partial L_{n_{a}}(\lambda_{a})}{\partial\lambda_{a}}\right)=1, (92e)

we arrive at the components of the Abelian quantum metric tensor associated to the family of NN coupled harmonic oscillators described by the Hamiltonian (71):

gij(n1,,nN)(x)\displaystyle g_{ij}^{(n_{1},\dots,n_{N})}(x) =\displaystyle= 18a=1N(na2+na+1)iωajωaωa214a,b=1NiUabjUab\displaystyle\frac{1}{8}\sum_{a=1}^{N}(n_{a}^{2}+n_{a}+1)\frac{\partial_{i}\omega_{a}\partial_{j}\omega_{a}}{\omega_{a}^{2}}-\frac{1}{4}\sum_{a,b=1}^{N}\partial_{i}U_{ab}\,\partial_{j}U_{ab} (93)
+12a,b,c,d=1N(na+12)(nb+12)(ωaωb+ωbωa)UacUadiUbcjUbd.\displaystyle+\frac{1}{2}\sum_{a,b,c,d=1}^{N}\left(n_{a}+\frac{1}{2}\right)\!\left(n_{b}+\frac{1}{2}\right)\!\left(\frac{\omega_{a}}{\omega_{b}}\!+\!\frac{\omega_{b}}{\omega_{a}}\right)U_{ac}U_{ad}\partial_{i}U_{bc}\partial_{j}U_{bd}.

This result is new and highly nontrivial. It reveals that the quantum metric tensor, for the ground state or any excited state, manifests a singular behavior at those points of the parameter space where at least one frequency vanishes. For the ground state of the system, these singularities might correspond to quantum phase transitions [1], whereas for excited states, they might be an indicator of excited-state quantum phase transitions, which can be regarded as extensions of quantum phase transitions [36]. Furthermore, given its simplicity, this metric appears to be suitable for analyzing its large NN behavior, in order to gain a better understanding of the the quantum metric tensor in quantum field theory. On the other hand, notice that for systems where the matrix UU does not depend on the parameters, only the terms involving the derivatives of the frequencies in the quantum metric tensor (93) survive. Indeed, it is direct to see that for the particular case of the two symmetric coupled harmonic oscillators (86), where N=2N=2 and iUab=0\partial_{i}U_{ab}=0, the metric tensor (93) reduces to

gij(n1,n2)(x)=(n12+n1+1)iω1jω18ω12+(n22+n2+1)iω2jω28ω22,\displaystyle g^{(n_{1},n_{2})}_{ij}(x)=(n_{1}^{2}+n_{1}+1)\frac{\partial_{i}\omega_{1}\partial_{j}\omega_{1}}{8\omega_{1}^{2}}+(n_{2}^{2}+n_{2}+1)\frac{\partial_{i}\omega_{2}\partial_{j}\omega_{2}}{8\omega_{2}^{2}}, (94)

with ω1=k\omega_{1}=\sqrt{k} and ω2=k+2k\omega_{2}=\sqrt{k+2k^{\prime}}. This expression shows that the metric is composed of two contributions representing the two uncoupled oscillators, and that its parameter structure is not modified by the quantum numbers. For the particular case n1=0n_{1}=0 and n2=0n_{2}=0, Eq. (94) reduces to the quantum metric tensor obtained in Ref. [37] through the path integral approach.

5.1 Classical analog of the quantum metric tensor

We would like now to compare and contrast the quantum metric tensor (93) with its classical counterpart. It is worth recalling that for a classical integrable system described by a Hamiltonian H(q,p;x)H(q,p;x) with adiabatic parameters xx, the classical analog of the quantum metric tensor is a metric that provides a measure of the distance between two points in phase space with infinitesimally different parameters. This metric is defined as [26]

gij(I,x)=GiGjClassGiClassGjClass,\displaystyle g_{ij}(I,x)=\left<G_{i}G_{j}\right>_{\rm Class}-\left<G_{i}\right>_{\rm Class}\left<G_{j}\right>_{\rm Class}, (95)

where GiδxiG_{i}\delta x^{i} is the generator of the infinitesimal canonical transformation (q(x),p(x))(q(x+δx),p(x+δx))(q(x),p(x))\rightarrow(q(x+\delta x),p(x+\delta x)), which in terms of the angle-action variables (φ,I)(\varphi,I) reads

Gi(φ,I;x):=a=1Npa(iqa)φ,I(iS)φ,I,\displaystyle G_{i}(\varphi,I;x):=\sum_{a=1}^{N}p_{a}(\partial_{i}q_{a})_{\varphi,I}-(\partial_{i}S)_{\varphi,I}, (96)

with pa=pa(φ,I;x)p_{a}=p_{a}(\varphi,I;x), qa=qa(φ,I;x)q^{a}=q^{a}(\varphi,I;x), and SS the generating function of the canonical transformation (q,p)(φ,I)(q,p)\rightarrow(\varphi,I). Furthermore, f(φ,I;x)Class=1(2π)n𝑑φf(φ,I;x)\left<f(\varphi,I;x)\right>_{\rm Class}=\frac{1}{(2\pi)^{n}}\oint d\varphi f(\varphi,I;x), with 𝑑φ=a=1n02π𝑑φa\oint d\varphi=\prod_{a=1}^{n}\int_{0}^{2\pi}d\varphi_{a}, is the average of f(φ,I;x)f(\varphi,I;x) over the (fast) angle variables.

The classical analog of the quantum Hamiltonian (71) will be taken to be

H=12a=1Npa2+12a,b=1NKabqaqb,\displaystyle H=\frac{1}{2}\sum_{a=1}^{N}p_{a}^{2}+\frac{1}{2}\sum_{a,b=1}^{N}K_{ab}q_{a}q_{b}, (97)

which can be written as H=(1/2)a=1N(Pa2+ωa2Qa2)H=(1/2)\sum_{a=1}^{N}(P_{a}^{2}+\omega_{a}^{2}Q_{a}^{2}), after using the linear canonical transformation analogous to that of Eq. (72), i.e., Qa=b=1NUabqbQ_{a}=\sum_{b=1}^{N}U_{ab}q_{b} and Pa=b=1NUabpbP_{a}=\sum_{b=1}^{N}U_{ab}p_{b}, with the generator

F=a,b=1NUabPaqb.\displaystyle F=\sum_{a,b=1}^{N}U_{ab}P_{a}q_{b}. (98)

Notice that the matrix UU is the same as in Eq. (72). In turn, the transformation from the variables (Q,P)(Q,P) to the action-angle variables (φ,I)(\varphi,I) is

Qa=(2Iaωa)1/2sinφa,Pa=(2ωaIa)1/2cosφa,\displaystyle Q_{a}=\left(\frac{2I_{a}}{\omega_{a}}\right)^{1/2}\sin\varphi_{a},\qquad P_{a}=\left(2\omega_{a}I_{a}\right)^{1/2}\cos\varphi_{a}, (99)

and its associated generating function is given by

S=a=1NIa(φa+sinφacosφa).\displaystyle S^{\prime}=\sum_{a=1}^{N}I_{a}\left(\varphi_{a}+\sin\varphi_{a}\cos\varphi_{a}\right). (100)

It is not hard to show that, after the canonical transformations (q,p)(Q,P)(φ,I)(q,p)\rightarrow(Q,P)\rightarrow(\varphi,I), the generating function (96) takes the form

Gi=a=1NPa(iQa)φ,I(iS)φ,I(iF)q,P.\displaystyle G_{i}=\sum_{a=1}^{N}P_{a}(\partial_{i}Q_{a})_{\varphi,I}-(\partial_{i}S^{\prime})_{\varphi,I}-(\partial_{i}F)_{q,P}. (101)

Then, plugging Eq. (101) into Eq. (95) and performing the integrals over the variables φa\varphi_{a}, the classical analog of the quantum metric tensor is then found to be

gij(I;x)=18a=1NIa2iωajωaωa2+12a,b,c,d=1NIaIb(ωaωb+ωbωa)UacUadiUbcjUbd.\displaystyle g_{ij}(I;x)=\frac{1}{8}\sum_{a=1}^{N}I_{a}^{2}\,\frac{\partial_{i}\omega_{a}\partial_{j}\omega_{a}}{\omega_{a}^{2}}+\frac{1}{2}\sum_{a,b,c,d=1}^{N}I_{a}I_{b}\left(\frac{\omega_{a}}{\omega_{b}}+\frac{\omega_{b}}{\omega_{a}}\right)U_{ac}U_{ad}\partial_{i}U_{bc}\partial_{j}U_{bd}. (102)

In order to compare Eqs. (93) and (102), it is reasonable to use the Bohr-Sommerfeld quantization rule Ia=(na+1/2)I_{a}=(n_{a}+1/2)\hbar and the identifications Ia2=(na2+na+1)2I_{a}^{2}=(n_{a}^{2}+n_{a}+1)\hbar^{2}. Under this consideration, the first thing we notice is that the classical metric (102) and the quantum metric (93) are identical except for the term 14a,b=1NiUabjUab\frac{1}{4}\sum_{a,b=1}^{N}\partial_{i}U_{ab}\,\partial_{j}U_{ab} of Eq. (93). Actually, we can see that the relation between the two metrics is

gij(n1,,nN)(x)=12(gij(I;x)24a,b=1NiUabjUab).\displaystyle g_{ij}^{(n_{1},\dots,n_{N})}(x)=\frac{1}{\hbar^{2}}\bigg{(}g_{ij}(I;x)-\frac{\hbar^{2}}{4}\sum_{a,b=1}^{N}\partial_{i}U_{ab}\partial_{j}U_{ab}\bigg{)}. (103)

Clearly, for those systems in which ijUab=0\partial_{i}\partial_{j}U_{ab}=0, the second term on the right hand side of Eq. (103) vanishes, and then both metrics produce the same parameter structure, satisfying the semiclassical relation established in Ref. [37]. The origin of the extra term can be attributed to the fact that the quantum metric tensor (93) has corrections of order 2\hbar^{2} to the classical metric (102).

5.2 Example: Linearly coupled harmonic oscillators

Using Eq. (93) it is quite easy to obtain the quantum metric tensor associated to a quantum system described by a Hamiltonian of the form (71). Let us now illustrate this with a system for which the corresponding matrix UU depends on the parameters. We consider the set of two coupled harmonic oscillators described by the Hamiltonian

𝐇^=12(𝐩^12+𝐩^22+A𝐪^12+B𝐪^22+C𝐪^1𝐪^2),\displaystyle\hat{\bf{H}}=\frac{1}{2}\left(\hat{\bf{p}}_{1}^{2}+\hat{\bf{p}}_{2}^{2}+A\hat{\bf{q}}_{1}^{2}+B\hat{\bf{q}}_{2}^{2}+C\hat{\bf{q}}_{1}\hat{\bf{q}}_{2}\right), (104)

where x={xi}=(A,B,C)x=\{x^{i}\}=(A,B,C) (i,j,=1,2,3i,j,\dots\!=1,2,3) are the adiabatic parameters, which are assumed to satisfy ABA\neq B. Although this model has been discussed extensively in the literature (see Refs. [38, 39, 40], for instance), little research has focused on its parameter space [37]. Moreover, its associated quantum metric tensor is known only for the ground state [37]. Here, as a direct application of Eq. (93), we shall obtain this geometrical structure for arbitrary quantum numbers.

By using the linear transformation (72) with the parameter-dependent matrix UU given by

U(x)=(cosαsinαsinαcosα),\displaystyle U(x)=\begin{pmatrix}\cos\alpha&-\sin\alpha\\ \sin\alpha&\cos\alpha\end{pmatrix}, (105)

where tanα=ϵ|ϵ|ϵ2+1ϵ\tan\alpha=\frac{\epsilon}{|\epsilon|}\sqrt{\epsilon^{2}+1}-\epsilon with ϵ=BAC\epsilon=\frac{B-A}{C}, the Hamiltonian (104) can be written as Eq. (73) with N=2N=2 and the frequencies

ω1=AC2tanαandω2=B+C2tanα.\displaystyle\omega_{1}=\sqrt{A-\frac{C}{2}\tan\alpha}\qquad{\rm and}\qquad\omega_{2}=\sqrt{B+\frac{C}{2}\tan\alpha}. (106)

Having Eqs. (105) and (106), the computation of the quantum metric tensor (93) is a simple exercise in differentiation. We obtain the metric

gij(n1,n2)(x)\displaystyle g_{ij}^{(n_{1},n_{2})}(x) =\displaystyle= (n12+n1+1)iω1jω18ω12+(n22+n2+1)iω2jω28ω22\displaystyle(n_{1}^{2}+n_{1}+1)\frac{\partial_{i}\omega_{1}\partial_{j}\omega_{1}}{8\omega_{1}^{2}}+(n_{2}^{2}+n_{2}+1)\frac{\partial_{i}\omega_{2}\partial_{j}\omega_{2}}{8\omega_{2}^{2}} (107)
+[(n1+12)(n2+12)(ω1ω2+ω2ω1)12]iαjα,\displaystyle+\left[\left(n_{1}+\frac{1}{2}\right)\left(n_{2}+\frac{1}{2}\right)\left(\frac{\omega_{1}}{\omega_{2}}+\frac{\omega_{2}}{\omega_{1}}\right)-\frac{1}{2}\right]\partial_{i}\alpha\partial_{j}\alpha,

which has the determinant

det[gij(n1,n2)(x)]=(n12+n1+1)(n22+n2+1)4096ω14ω24(ω12ω22)2[(n1+12)(n2+12)(ω1ω2+ω2ω1)12].\displaystyle\det[g^{(n_{1},n_{2})}_{ij}(x)]=\frac{(n_{1}^{2}+n_{1}+1)(n_{2}^{2}+n_{2}+1)}{4096\omega_{1}^{4}\omega_{2}^{4}(\omega_{1}^{2}-\omega_{2}^{2})^{2}}\left[\left(n_{1}+\frac{1}{2}\right)\!\left(n_{2}+\frac{1}{2}\right)\!\left(\frac{\omega_{1}}{\omega_{2}}+\frac{\omega_{2}}{\omega_{1}}\right)\!-\!\frac{1}{2}\right]\!.

It is not hard to see that for the particular case of n1=0n_{1}=0 and n2=0n_{2}=0, the metric (107) reduces to the one obtained in Ref. [37] by using the path integral formalism. This result then corroborates the validity and usefulness of Eq. (93).

6 Non-Abelian quantum metric tensor for three coupled oscillators

Let us consider a quantum mechanical system composed of three coupled oscillators and described by the Hamiltonian

𝐇^=12{a=13(𝐩^a2+k𝐪^a2)+k[(𝐪^1𝐪^2)2+(𝐪^2𝐪^3)2+(𝐪^3𝐪^1)2]},\displaystyle\hat{\bf{H}}=\frac{1}{2}\left\{\sum_{a=1}^{3}\left(\hat{\bf{p}}_{a}^{2}+k\hat{\bf{q}}_{a}^{2}\right)+k^{\prime}\big{[}(\hat{\bf{q}}_{1}-\hat{\bf{q}}_{2})^{2}+(\hat{\bf{q}}_{2}-\hat{\bf{q}}_{3})^{2}+(\hat{\bf{q}}_{3}-\hat{\bf{q}}_{1})^{2}\big{]}\right\}, (109)

where x={xi}=(k,k)x=\{x^{i}\}=(k,k^{\prime}) with i,j,=1,2i,j,\dots\!=1,2 are the adiabatic parameters. Note that because this Hamiltonian belongs to the class of quadratic Hamiltonians given by Eq. (71), we can apply Eq. (93) to easily obtain the Abelian quantum metric tensor associated with its nondegenerate ground state. Nevertheless, the aim here is to obtain the Non-Abelian quantum metric tensor for a particular set of degenerate states by means of Eq. (50).

Bearing in mind the linear transformation (72) with N=3N=3 and the parameter-independent matrix

U=(13131312012162316),\displaystyle U=\begin{pmatrix}\frac{1}{\sqrt{3}}&\frac{1}{\sqrt{3}}&\frac{1}{\sqrt{3}}\\ -\frac{1}{\sqrt{2}}&0&\frac{1}{\sqrt{2}}\\ \frac{1}{\sqrt{6}}&-\sqrt{\frac{2}{3}}&\frac{1}{\sqrt{6}}\end{pmatrix}, (110)

the Hamiltonian (109) can be put in the form

𝐇^=12a=13(𝐏^a2+ωa2𝐐^a2),\displaystyle\hat{\bf{H}}=\frac{1}{2}\sum_{a=1}^{3}\left(\hat{\bf{P}}_{a}^{2}+\omega_{a}^{2}\hat{\bf{Q}}_{a}^{2}\right), (111)

with the frequencies ω1=k\omega_{1}=\sqrt{k} and ω2=ω3=k+3k\omega_{2}=\omega_{3}=\sqrt{k+3k^{\prime}}. Consequently, the normalized wave functions of the system can be written as

ψn1,n2,n3(q1,q2,q3;x)=ψn1(Q1;x)ψn2(Q2;x)ψn3(Q3;x),\displaystyle\psi_{n_{1},n_{2},n_{3}}(q_{1},q_{2},q_{3};x)=\psi_{n_{1}}(Q_{1};x)\psi_{n_{2}}(Q_{2};x)\psi_{n_{3}}(Q_{3};x), (112)

where ψna(Qa;x)\psi_{n_{a}}(Q_{a};x) is the wave function of the aath uncoupled oscillator with quantum number na=0,1,2,n_{a}=0,1,2,\dots and has the same expression as in Eq. (75). The energy eigenvalues, which depend on three quantum numbers n1n_{1}, n2n_{2} and n3n_{3}, are then given by

En1,n2,n3=(n1+12)ω1+(n2+n3+1)ω2.\displaystyle E_{n_{1},n_{2},n_{3}}=\left(n_{1}+\frac{1}{2}\right)\hbar\omega_{1}+\left(n_{2}+n_{3}+1\right)\hbar\omega_{2}. (113)

Here, we consider the wave functions ψ0,0,1\psi_{0,0,1} and ψ0,1,0\psi_{0,1,0}, which have the same energy

E1:=12ω1+2ω2,\displaystyle E_{1}:=\frac{1}{2}\hbar\omega_{1}+2\hbar\omega_{2}, (114)

and then constitute a degenerate set (g1=2g_{1}=2). Notice that this degeneracy is a consequence of the fact that the Hamiltonian (111) is invariant under the interchange of Q1Q_{1} and Q2Q_{2}.

By introducing the notation {ψ(1)I}:=(ψ0,0,1,ψ0,1,0)\{\psi_{(1)I}\}:=(\psi_{0,0,1},\psi_{0,1,0}) with I,J=1,2I,J=1,2, the associated non-diagonal Wigner functions are obtained by using Eq. (40), which takes the form

W(1)IJ=1(2π)3d3yeipyψ(1)I(q+y2;x)ψ(1)J(qy2;x),\displaystyle W_{(1)IJ}=\frac{1}{(2\pi\hbar)^{3}}\int_{-\infty}^{\infty}{\rm d}^{3}y\,{\rm e}^{-\frac{{\rm i}p\cdot y}{\hbar}}\psi_{(1)I}(q+\tfrac{y}{2};x)\psi_{(1)J}(q-\tfrac{y}{2};x), (115)

and leads to

W(1)11=1(π)3(λ31)eλ1+λ2+λ32,\displaystyle W_{(1)11}=\frac{1}{(\pi\hbar)^{3}}(\lambda_{3}-1){\rm e}^{-\frac{\lambda_{1}+\lambda_{2}+\lambda_{3}}{2}}, (116a)
W(1)12=2π34ω2(P2ω2Q2)(P3+ω2Q3)eλ1+λ2+λ32,\displaystyle W_{(1)12}=\frac{2}{\pi^{3}\hbar^{4}\omega_{2}}(P_{2}-\rmi\omega_{2}Q_{2})(P_{3}+\rmi\omega_{2}Q_{3}){\rm e}^{-\frac{\lambda_{1}+\lambda_{2}+\lambda_{3}}{2}}, (116b)
W(1)21=2π34ω2(P2+ω2Q2)(P3ω2Q3)eλ1+λ2+λ32,\displaystyle W_{(1)21}=\frac{2}{\pi^{3}\hbar^{4}\omega_{2}}(P_{2}+\rmi\omega_{2}Q_{2})(P_{3}-\rmi\omega_{2}Q_{3}){\rm e}^{-\frac{\lambda_{1}+\lambda_{2}+\lambda_{3}}{2}}, (116c)
W(1)22=1(π)3(λ21)eλ1+λ2+λ32,\displaystyle W_{(1)22}=\frac{1}{(\pi\hbar)^{3}}(\lambda_{2}-1){\rm e}^{-\frac{\lambda_{1}+\lambda_{2}+\lambda_{3}}{2}}, (116d)

where λa:=4Ha/ωa\lambda_{a}:=4H_{a}/\hbar\omega_{a} with Ha=(1/2)(Pa2+ωa2Qa2)H_{a}=(1/2)(P_{a}^{2}+\omega_{a}^{2}Q_{a}^{2}).

Then, plugging the expressions for W(1)IJW_{(1)IJ} into Eq. (50), the components of the non-Abelian quantum metric tensor turn out to be

gij11(1)(x)=\displaystyle g^{(1)}_{ij11}(x)= gij22(1)(x)=132(1ω14+4ω2412ω2412ω2436ω24),\displaystyle g^{(1)}_{ij22}(x)=\frac{1}{32}\begin{pmatrix}\frac{1}{\omega_{1}^{4}}+\frac{4}{\omega_{2}^{4}}&\frac{12}{\omega_{2}^{4}}\\ \frac{12}{\omega_{2}^{4}}&\frac{36}{\omega_{2}^{4}}\end{pmatrix},
gij12(1)(x)=\displaystyle g^{(1)}_{ij12}(x)= gij21(1)(x)=(0000),\displaystyle g^{(1)}_{ij21}(x)=\begin{pmatrix}0&0\\ 0&0\end{pmatrix}, (117)

which are exactly the same that can be obtained directly from Eq. (10). Notice that the metric components gij11(1)(x)g^{(1)}_{ij11}(x) and gij22(1)(x)g^{(1)}_{ij22}(x) diverge at the points of the parameter space where the frequencies of the system approach to zero. Besides, note also that gij11(1)(x)g^{(1)}_{ij11}(x) and gij22(1)(x)g^{(1)}_{ij22}(x) can be thought of as invertible matrices, with determinant given by det[gij11(1)(x)]=det[gij22(1)(x)]=9/256ω14ω24\det[g^{(1)}_{ij11}(x)]=\det[g^{(1)}_{ij22}(x)]=9/256\omega_{1}^{4}\omega_{2}^{4}. Finally, it is not hard to show that the associated Wilczek-Zee connection and its corresponding curvature are zero. This indicates that non-Abelian quantum metric tensor may contain some valuable information that cannot be extracted from the Wilczek-Zee curvature.

7 Conclusion

In this paper, we have proposed a phase space formulation of the Berry connection, the Wilczek-Zee connection, and the Abelian and non-Abelian quantum geometric tensor, from which the corresponding quantum metrics and curvatures can be obtained in a unified fashion. In the new formulation, these geometrical structures are expressed as integrals over phase space of complex functions, which in principle could be used to obtain additional information about the parameter space. We have shown that the Abelian and non-Abelian quantum metric tensor can be formulated in the phase space formalism by using only parameter derivatives of the diagonal and non-diagonal Wigner functions, respectively. We have also obtained a phase space formulation of the Berry and Wilczek-Zee curvatures, which is different from the one proposed in Ref [13]. Indeed, the formulation of Ref [13] involves the angle-action variables and therefore, can be applied only to quantum systems whose classical counterpart is integrable. Our formulation, in contrast, uses the new phase space functions 𝒜i(n)\mathcal{A}_{i}^{(n)} (Abelian case) and 𝒜iIJ(n)\mathcal{A}_{iIJ}^{(n)} (non-Abelian case) and does not have that limitation.

We have illustrated the developed formulation by computing the Berry connection and the quantum geometric tensor for the generalized harmonic oscillator, obtaining the expected results. As an application of the formulation, we have derived the expression for the Abelian quantum metric tensor associated with a system of NN coupled harmonic oscillators. The resulting quantum metric exhibits a singular behavior at the points of the parameter space where there may be a frequency that vanishes. These singularities might be indicators of the presence of quantum phase transitions; however, this is still a matter of on-going research. We have also obtained the classical analog of this Abelian metric and found that both metrics have the same parameter structure, except for an extra term that does not involve singularities and that emerges when the transformation that diagonalizes the Hamiltonians depends on the parameters. This result suggests that the classical analog of the quantum metric tensor may help to get a first insight into quantum phase transitions. Finally, we have considered a system of three coupled harmonic oscillators and applied the phase space approach to compute the associated non-Abelian quantum metric tensor, finding that the resulting metric has singularities at the points of the parameter space where the frequencies vanish.

The phase space approach to the parameter space has several advantages over the standard formulation. First, it is suitable to perform a semiclassical analysis of the geometrical structures involved in the parameter space, because phase space is directly adopted as in many classical concepts. In this line of thought, since the study of chaos is more appropriately carried out in phase space than in configuration space, this approach could be well adapted to seek indicators of quantum chaos in the parameter space framework [41, 42]. Also, this approach seems to be a convenient way to study quadratic Hamiltonians, since it allows a systematic treatment of the system, leading to general results. Furthermore, this approach provides a picture in phase space of what is happening in the parameter space through the associated phase space functions 𝒜i(n)\mathcal{A}_{i}^{(n)} (Abelian case) and 𝒜iIJ(n)\mathcal{A}_{iIJ}^{(n)} (non-Abelian case), which encode all the information on the underlying parameter space and serve as the fundamental building blocks for the Abelian and non-Abelian geometrical structures. In addition, and remarkably, in this approach the expression for the Abelian and non-Abelian quantum metric tensor only requires the knowledge of the Wigner functions, which are much closer to classical intuition than the wave functions and can be experimentally measured. In this sense, our procedure also offers an alternative to address the issue of experimentally determining the quantum metric tensor.

The present approach can be used to tackle a variety of conceptual and applied issues. For instance, it can be applied to compute the quantum metric tensor associated with the inverted oscillator, which has been used to study quantum chaos [43] and whose Wigner function is known [44, 45]. Also, as a first step towards understanding the quantum metric tensor in quantum field theory, it would be worth analyzing the large NN behavior of the quantum metric tensor obtained in Sec. 5 for a lattice of coupled oscillators [35]. Another important scenario where this approach could be helpful is in analyzing the parameter space associated with coherent states, which have a very convenient representation in the Wigner formalism and have played a central role in quantum optics [46]. Apart from possible applications for pure states, the developed approach may be extended to achieve a phase space formulation of the (non-Abelian) Uhlmann’s connection for mixed quantum states [47, 48]. Such a formulation would take advantage of a linearity of mixed states in phase space [32]. Among other applications, the present approach to the parameter space can also be applied to other phase space representations of quantum mechanics, which have their worthiness depending on the kind of application. In particular, it can be implemented in the Husimi representation wherein the state of the system is described by the Husimi function, which is closer to the classical concepts than the Wigner function [49, 50] and, for Bloch coherent states, has an appealing interpretation in connection with the Fubini–Study distance [51].

Daniel Gutiérrez-Ruiz is supported with a CONACyT Ph.D. scholarship (No. 332577). Diego Gonzalez was partially supported by a DGAPA-UNAM postdoctoral fellowship and by Consejo Nacional de Ciencia y Tecnología (CONACyT), México, Grant No. A1-S-7701. This work was partially supported by DGAPA-PAPIIT Grant No. IN103919.

Appendix A Alternative derivation of Eq. (32).

First, let us express Eq. (3) in a more suitable form. Taking the second partial derivatives of the density operator ρ^n(x)=|nn|\hat{\rho}_{n}(x)=\lvert n\rangle\langle n\rvert with respect to the parameters, we have

ijρ^n=|ijnn|+|nijn|+|injn|+|jnin|,\displaystyle\partial_{i}\partial_{j}\hat{\rho}_{n}=\lvert\partial_{i}\partial_{j}n\rangle\langle n\rvert+\lvert n\rangle\langle\partial_{i}\partial_{j}n\rvert+\lvert\partial_{i}n\rangle\langle\partial_{j}n\rvert+\lvert\partial_{j}n\rangle\langle\partial_{i}n\rvert, (118)

In turn, the normalization condition n|n=1\langle n|n\rangle=1 implies ijn|n+n|ijn+in|jn+jn|in=0\langle\partial_{i}\partial_{j}n|n\rangle+\langle n|\partial_{i}\partial_{j}n\rangle+\langle\partial_{i}n|\partial_{j}n\rangle+\langle\partial_{j}n|\partial_{i}n\rangle=0, and hence

n|ijρ^n|n=in|jnjn|in+in|nn|jnx+jn|nn|in.\displaystyle\langle n\rvert\partial_{i}\partial_{j}\hat{\rho}_{n}\lvert n\rangle=-\langle\partial_{i}n|\partial_{j}n\rangle-\langle\partial_{j}n|\partial_{i}n\rangle+\langle\partial_{i}n|n\rangle\langle n|\partial_{j}n\rangle x+\langle\partial_{j}n|n\rangle\langle n|\partial_{i}n\rangle. (119)

Plugging this result back into Eq. (3), the quantum metric is expressed by the expectation value of ijρ^n\partial_{i}\partial_{j}\hat{\rho}_{n}, namely

gij(n)(x)=12ijρ^nn.\displaystyle g^{(n)}_{ij}(x)=-\frac{1}{2}\langle\partial_{i}\partial_{j}\hat{\rho}_{n}\rangle_{n}. (120)

This simple fact, together with Eqs. (16) and (17), now provides the connection between the quantum metric tensor and the Wigner function. Actually, from Eqs. (17) and (120), we have

gij(n)(x)=12dNqdNpWn(ijρ~n)q,p.\displaystyle g^{(n)}_{ij}(x)=-\frac{1}{2}\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,W_{n}(\partial_{i}\partial_{j}\tilde{\rho}_{n})_{q,p}. (121)

Then, with the help of Eq. (16), Eq. (121) becomes

gij(n)(x)=(2π)N2dNqdNpWn(ijWn)q,p,\displaystyle g^{(n)}_{ij}(x)=-\frac{(2\pi\hbar)^{N}}{2}\int_{-\infty}^{\infty}{\rm d}^{N}q\,{\rm d}^{N}p\,W_{n}(\partial_{i}\partial_{j}W_{n})_{q,p}, (122)

which is an equivalent and alternative form of the quantum metric tensor in the Wigner function formalism. Finally, integrating Eq. (122) by parts and using Eq. (31), we obtain Eq. (32).

References

References

  • [1] Gu S J 2010 Int. J. Mod. Phys. B 24 4371–4458
  • [2] Carollo A, Valenti D and Spagnolo B 2020 Phys. Rep. 838 1 – 72
  • [3] Provost J P and Vallee G 1980 Commun. Math. Phys. 76 289–301
  • [4] Berry M V 1984 Proc. R. Soc. A 392 45–57
  • [5] Chen G, Li J and Liang J Q 2006 Phys. Rev. A 74(5) 054101
  • [6] Campos Venuti L and Zanardi P 2007 Phys. Rev. Lett. 99(9) 095701
  • [7] Zanardi P, Giorda P and Cozzini M 2007 Phys. Rev. Lett. 99(10) 100603
  • [8] Miyaji M, Numasawa T, Shiba N, Takayanagi T and Watanabe K 2015 Phys. Rev. Lett. 115(26) 261602
  • [9] Bak D 2016 Phy. Lett. B 756 200 – 204 ISSN 0370-2693
  • [10] Trivella A 2017 Class. Quantum Grav. 34 105003
  • [11] Alvarez-Jimenez J, Dector A and Vergara J D 2017 J. High Energy Phys. 2017 44
  • [12] Alvarez-Jimenez J and Vergara J D 2019 Int. J. Quantum Inf. 17 1950017
  • [13] Chruściński D 2006 Open Sys. Inf. Dyn. 13 67–74
  • [14] Pati A K 1998 Ann. Phys. (N.Y.) 270 178 – 197 ISSN 0003-4916
  • [15] Kim Y S and Noz M E 1991 Phase Space Picture of Quantum Mechanics (World Scientific, Singapore)
  • [16] Zachos K, Fairlie D B and Curtright T L 2005 Quantum Mechanics in Phase Space (World Scientific, Singapore)
  • [17] Kenfack A and Życzkowski K 2004 , J. Opt. B 6 396–404
  • [18] Bernardini A E and Bertolami O 2019 J. Phys. Conf. Ser. 1275 012032
  • [19] Wigner E 1932 Phys. Rev. 40(5) 749–759
  • [20] Weinbub J and Ferry D K 2018 Appl. Phys. Rev. 5 041104
  • [21] Smithey D T, Beck M, Raymer M G and Faridani A 1993 Phys. Rev. Lett. 70(9) 1244–1247
  • [22] Dunn T J, Walmsley I A and Mukamel S 1995 Phys. Rev. Lett. 74(6) 884–887
  • [23] Weyl H 1931 The Theory of Groups and Quantum Mechanics (Dover, New York) ISBN 978
  • [24] Ma Y Q, Chen S, Fan H and Liu W M 2010 Phys. Rev. B 81(24) 245129
  • [25] Wilczek F and Zee A 1984 Phys. Rev. Lett. 52(24) 2111–2114
  • [26] Gonzalez D, Gutiérrez-Ruiz D and Vergara J D 2019 Phys. Rev. E 99(3) 032144
  • [27] Bombelli L, Koul R K, Lee J and Sorkin R D 1986 Phys. Rev. D 34(2) 373–383
  • [28] Srednicki M 1993 Phys. Rev. Lett. 71(5) 666–669
  • [29] Alvarez-Jimenez J and Vergara J D 2016 J. Mod. Phys. 7 1627–1634
  • [30] Berry M V 1985 J. Phys. A: Math. Gen. 18 15
  • [31] Wilczek F and Shapere A 1989 Geometric Phases in Physics (World Scientific, Singapore) ISBN 9789971506216
  • [32] Case W B 2008 Am. J. Phys. 76 937–946
  • [33] Hannay J H 1985 J. Phys. A: Math. Gen. 18 221
  • [34] Chandran S M and Shankaranarayanan S 2019 Phys. Rev. D 99(4) 045010
  • [35] Jefferson R A and Myers R C 2017 J. High Energy Phy. 2017 107 ISSN 1029-8479
  • [36] Caprio M, Cejnar P and Iachello F 2008 Ann. Phys. (N.Y.) 323 1106 – 1135
  • [37] Alvarez-Jimenez J, Gonzalez D, Gutiérrez-Ruiz D and Vergara J D Ann. Phys. (Berlin) 1900215
  • [38] Kim Y S and Noz M E 2005 J. Opt. B: Quantum Semiclass. Opt. 7 S458–S467
  • [39] Paz J P and Roncaglia A J 2008 Phys. Rev. Lett. 100(22) 220401
  • [40] Makarov D N 2018 Phys. Rev. E 97 042203
  • [41] Chávez-Carlos J, López-del Carpio B, Bastarrachea-Magnani M A, Stránský P, Lerma-Hernández S, Santos L F and Hirsch J G 2019 Phys. Rev. Lett. 122(2) 024101
  • [42] Pandey M, Claeys P W, Campbell D K, Polkovnikov A and Sels D 2020 (Preprint 2004.05043)
  • [43] Ali T, Bhattacharyya A, Haque S S, Kim E H, Moynihan N and Murugan J 2020 Phys. Rev. D 101(2) 026021
  • [44] Wolf K B 2010 Rev. Mex. Fis. E 56 83 – 91
  • [45] Heim D, Schleich W, Alsing P, Dahl J and Varro S 2013 Phys. Lett. A 377 1822 – 1825 ISSN 0375-9601
  • [46] Schleich W P 2001 Quantum States in Phase Space (John Wiley & Sons) ISBN 9783527294350
  • [47] Uhlmann A 1986 Rep. Math. Phys. 24 229 – 240 ISSN 0034-4877
  • [48] Uhlmann A 1989 Ann. Phys. (Berlin) 501 63–69
  • [49] Harriman J E and Casida M E 1993 Int. J. Quantum Chem. 45 263–294
  • [50] Toscano F, Kenfack A, Carvalho A R R, Rost J M and de Almeida A M O 2008 Proc. R. Soc. A 464 1503–1524 ISSN 13645021
  • [51] Bengtsson I and Życzkowski K 2017 Geometry of Quantum States: An Introduction to Quantum Entanglement (Cambridge University Press, Cambridge, England) ISBN 978-1-107-02625-4