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

Bures and Sjöqvist Metrics over Thermal State Manifolds for Spin Qubits and Superconducting Flux Qubits

Carlo Cafaro1 and Paul M. Alsing2 1SUNY Polytechnic Institute, 12203 Albany, New York, USA 2Air Force Research Laboratory, Information Directorate, 13441 Rome, New York, USA
Abstract

The interplay among differential geometry, statistical physics, and quantum information science has been increasingly gaining theoretical interest in recent years.

In this paper, we present an explicit analysis of the Bures and Sjöqvist metrics over the manifolds of thermal states for specific spin qubit and the superconducting flux qubit Hamiltonian models. While the two metrics equally reduce to the Fubini-Study metric in the asymptotic limiting case of the inverse temperature approaching infinity for both Hamiltonian models, we observe that the two metrics are generally different when departing from the zero-temperature limit. In particular, we discuss this discrepancy in the case of the superconducting flux Hamiltonian model. We conclude the two metrics differ in the presence of a nonclassical behavior specified by the noncommutativity of neighboring mixed quantum states. Such a noncommutativity, in turn, is quantified by the two metrics in different manners. Finally, we briefly discuss possible observable consequences of this discrepancy between the two metrics when using them to predict critical and/or complex behavior of physical systems of interest in quantum information science.

pacs:
Quantum Computation (03.67.Lx), Quantum Information (03.67.Ac), Quantum Mechanics (03.65.-w), Riemannian Geometry (02.40.Ky), Statistical Mechanics (05.20.-y).

I Introduction

Geometry plays a special role in the description and, to a certain extent, in the understanding of various physical phenomena pettini07 ; karol06 . The concepts of length, area, and volume are ubiquitous in physics and their meaning can prove quite helpful in explaining physical phenomena from a more intuitive perspective cafaroprd22 ; cafaropre22 . The notions of “longer” and “shorter” are extensively used in virtually all disciplines cafarophysicaa22 . Indeed, geometric formulations of classical and quantum evolutions along with geometric descriptions of classical and quantum mechanical aspects of thermal phenomena are becoming increasingly important in science. Concepts, such as thermodynamic length, area law, and statistical volumes are omnipresent in geometric thermodynamics, general relativity, and statistical physics, respectively. The concept of entropy finds application in essentially any realm of science, from classical thermodynamics to quantum information science. The notions of “hotter” and “cooler” are widely used in many fields. Entropy can be used to provide measures of distinguishability of classical probability distributions, as well as pure and mixed quantum states. It can also be used to propose measures of complexity for classical motion, quantum evolution, and entropic motion on curved statistical manifolds underlying the entropic dynamics of physical systems for which only partial knowledge of relevant information can be obtained cafaroPhD ; cafaroCSF ; felice18 . Furthermore, entropy can also be used to express the degree of entanglement in a quantum state specifying a composite quantum system. For instance, concepts such as Shannon entropy, von Neumann entropy, and Umegaki relative entropy are ubiquitous in classical information science, quantum information theory, and information geometric formulations of mixed quantum state evolutions amari , respectively. In this paper, inspired by the increasing theoretical interest in the interplay among differential geometry, statistical physics, and quantum information science zanardiprl07 ; zanardi07 ; pessoa21 ; silva21 ; silva21B ; mera22 , we present an explicit analysis of the Bures bures69 ; uhlman76 ; hubner92 and Sjöqvist erik20 metrics over the manifolds of thermal states for the spin qubit and the superconducting flux qubit Hamiltonian models. From a chronological standpoint, the first physical application of the Sjöqvist interferometric metric occurs in the original paper by Sjöqvist himself in Ref. erik20 . Here, the author considered his newly proposed interferometric metric to quantify changes in behavior of a magnetic system in a thermal state under modifications of temperature and magnetic field intensity. Keeping the temperature constant while changing the externally applied magnetic field, the Sjöqvist interferometric metric was shown to be physically linked to the magnetic susceptibility. This quantity, in turn, quantifies how much a material will become magnetized when immersed in a magnetic field. A second application of the Sjöqvist interferometric metric happens in Refs. silva21 ; silva21B . Here, this metric is used to characterize the finite-temperature phase transitions in the framework of band insulators. In particular, the authors considered the massive Dirac Hamiltonian model for a band insulator in two spatial dimensions. The corresponding Sjöqvist interferometric metric was calculated and expressed in terms of two physical parameters, the temperature, and the hopping parameter. Furthermore, the Sjöqvist interferometric metric was physically regarded as an interferometric susceptibility. Interestingly, it was observed in Refs. silva21 ; silva21B a dramatic difference between the Sjöqvist interferometric metric and the Bures metric when studying topological phase transitions in these types of systems. Specifically, while the topological phase transition is captured for all temperatures in the case of the Sjöqvist interferometric metric, the topological phase transition is captured only at zero temperature in the case of the Bures metric. Clearly, the authors leave as an unsolved question the experimental observation of the singular behavior of the Sjöqvist interferometric metric in actual laboratory experiments in Refs. silva21 ; silva21B . A third interesting work is the one presented in Ref. mera22 . Here the authors focus on the zero-temperature aspects of certain quantum systems in their (pure) ground quantum state. They consider two systems. The first system is described by the XYXY anisotropic spin-1/21/2 chain with NN-sites on a circle in the presence of an external magnetic field. The second model is the Haldane model, a two-dimensional condensed-matter lattice model haldane88 . In the first case, the two parameters that determine both the Hamiltonian and the parameter manifold are the anisotropy degree and the magnetic field intensity. In the second case, instead, the two key parameters become the on-site energy and the phase in the model being considered. Expressing the so-called quantum metric in terms of these tunable parameters, they study the thermodynamical limit of this metric along critical submanifolds of the whole parameter manifold. They observe a singular (regular) behavior of the metric along normal (tangent) directions to the critical submanifolds. Therefore, they conclude that tangent directions to critical manifolds are special. Finally, the authors also point out that it would be interesting to understand how their findings generalize to the finite-temperature case where states become mixed. Interestingly, without reporting any explicit analysis, the authors state they expect the Bures and the Sjöqvist metrics to assume different functional forms.

In this paper, inspired by the previously mentioned relevance of comprehending the physical significance of choosing one metric over another one in such geometric characterizations of physical aspects of quantum systems, we report a complete and straightforward analysis of the link between the Sjöqvist interferometric metric and the Bures metric for two special classes of nondegenerate mixed quantum states. Specifically , focusing on manifolds of thermal states for the spin qubit and the superconducting flux qubit Hamiltonian models, we observe that while the two metrics both reduce to the Fubini-Study metric provost80 ; wootters81 ; braunstein94 in the zero-temperature asymptotic limiting case of the inverse temperature β=def(kBT)1\beta\overset{\text{def}}{=}\left(k_{B}T\right)^{-1} (with kBk_{B} being the Boltzmann constant) approaching infinity for both Hamiltonian models, the two metrics are generally different. Furthermore, we observe this different behavior in the case of the superconducting flux Hamiltonian model. More generally, we note that the two metrics seem to differ when a nonclassical behavior is present since in this case, the metrics quantify noncommutativity of neighboring mixed quantum states in different manners. Finally, we briefly discuss the possible observable consequences of this discrepancy between the two metrics when using them to predict critical and/or complex behavior of physical systems of interest. We acknowledge that despite the fact that most of the preliminary background results presented in this paper are partially known in the literature, they appear in a scattered fashion throughout several papers written by researchers working in distinct fields of physics who may not be necessary aware of each other’s findings. For this reason, we present here an explicit and unified comparative formulation of the Bures and Sjöqvist metrics for simple quantum systems in mixed states. In particular, as mentioned earlier, we illustrate our results on the examples of thermal state manifolds for spin qubits and superconducting flux qubits. These applications are original and, to the best of our knowledge, do not appear anywhere in the literature.

The layout of the rest of this paper is as follows. In Section II, we present with explicit derivations the expressions of the Bures metric in Hübner’s hubner92 and Zanardi’s zanardi07 forms. In Section III, we present an explicit derivation of the Sjöqvist interferometric metric erik20 between two neighboring nondegenerate density matrices. In Section IV, we present two Hamiltonian models. The first Hamiltonian model describes a spin-1/21/2 particle in a uniform and time-independent external magnetic field oriented along the zz-axis. The second Hamiltonian model, instead, specifies a superconducting flux qubit. Bringing these two systems in thermal equilibrium with a reservoir at finite and non-zero temperature TT, we construct the two corresponding parametric families of thermal states. In Section V, we present an explicit calculation of both the Sjöqvist and the Bures metrics for each one of the two distinct families of parametric thermal states. From our comparative analysis, we find that the two metric coincide for the first Hamiltonian model (electron in a constant magnetic field along the zz-direction), while they differ for the second Hamiltonian model (superconducting flux qubit). In Section VI, we discuss the effects that arise from the comparative analysis carried out in Section V concerning the Bures and Sjöqvist metrics for spin qubits and superconducting flux qubits Hamiltonian models introduced in Section IV. Finally, we conclude with our final remarks along with a summary of our main findings in Section VII.

II The Bures metric

In this section, we present two explicit calculations. In the first calculation, we carry out a detailed derivation of the Bures metric by following the original work presented by Hübner in Ref. hubner92 . In the second calculation, we recast the expression of the Bures metric obtained by Hübner in a way that is more suitable in the framework of geometric analyses on thermal states manifolds. Here, we follow the original work presented by Zanardi and collaborators in Ref. zanardi07 .

II.1 The explicit derivation of Hübner’s general expression

We begin by carrying out an explicit derivation of the Bures metric inspired by Hübner hubner92 . Recall that the squared Bures distance between two density matrices infinitesimally far apart is given by,

[dBures(ρρ+dρ)]2=22tr[ρ1/2(ρ+dρ)ρ1/2]1/2.\left[d_{\mathrm{Bures}}\left(\rho\text{, }\rho+d\rho\right)\right]^{2}=2-2\mathrm{tr}\left[\rho^{1/2}\left(\rho+d\rho\right)\rho^{1/2}\right]^{1/2}\text{.} (1)

To find a useful expression for [dBures(ρρ+dρ)]2\left[d_{\mathrm{Bures}}\left(\rho\text{, }\rho+d\rho\right)\right]^{2}, we follow the line of reasoning used by Hübner in Ref. hubner92 . Consider an Hermitian matrix A(t)A\left(t\right) with tt\in\mathbb{R} defined as

A(t)=def[ρ1/2(ρ+tdρ)ρ1/2]1/2,A\left(t\right)\overset{\text{def}}{=}\left[\rho^{1/2}\left(\rho+td\rho\right)\rho^{1/2}\right]^{1/2}\text{,} (2)

with A(t)A(t)=ρ1/2(ρ+tdρ)ρ1/2A\left(t\right)A\left(t\right)=\rho^{1/2}\left(\rho+td\rho\right)\rho^{1/2}. Note that A(0)=ρA\left(0\right)=\rho and, for later use, we assume ρ\rho to be invertible. At this point we observe that knowledge of the metric tensor gij(ρ)g_{ij}\left(\rho\right) at ρ\rho requires knowing [dBures(ρρ+tdρ)]2\left[d_{\mathrm{Bures}}\left(\rho\text{, }\rho+td\rho\right)\right]^{2} up to the second order in tt. Hübner’s ansatz (i.e., educated guess) is

[dBures(ρρ+tdρ)]2=t2gij(ρ)dρidρj,\left[d_{\mathrm{Bures}}\left(\rho\text{, }\rho+td\rho\right)\right]^{2}=t^{2}g_{ij}\left(\rho\right)d\rho^{i}d\rho^{j}\text{,} (3)

with {ρi}\left\{\rho^{i}\right\} denoting a given set of coordinates on the manifold of density matrices. From Eq. (3), we note that

gij(ρ)dρidρj=12(d2dt2[dBures(ρρ+tdρ)]2)t=0.g_{ij}\left(\rho\right)d\rho^{i}d\rho^{j}=\frac{1}{2}\left(\frac{d^{2}}{dt^{2}}\left[d_{\mathrm{Bures}}\left(\rho\text{, }\rho+td\rho\right)\right]^{2}\right)_{t=0}\text{.} (4)

Observe that using Eq. (2), the RHS of Eq. (4) becomes

12(d2dt2[dBures(ρρ+tdρ)]2)t=0\displaystyle\frac{1}{2}\left(\frac{d^{2}}{dt^{2}}\left[d_{\mathrm{Bures}}\left(\rho\text{, }\rho+td\rho\right)\right]^{2}\right)_{t=0} =12d2dt2{22tr[ρ1/2(ρ+tdρ)ρ1/2]1/2}t=0\displaystyle=\frac{1}{2}\frac{d^{2}}{dt^{2}}\left\{2-2\mathrm{tr}\left[\rho^{1/2}\left(\rho+td\rho\right)\rho^{1/2}\right]^{1/2}\right\}_{t=0}
=12d2dt2{22tr[A(t)]}t=0\displaystyle=\frac{1}{2}\frac{d^{2}}{dt^{2}}\left\{2-2\mathrm{tr}\left[A\left(t\right)\right]\right\}_{t=0}
=tr[A¨(t)]t=0,\displaystyle=-\mathrm{tr}\left[\ddot{A}\left(t\right)\right]_{t=0}\text{,} (5)

that is,

12(d2dt2[dBures(ρρ+tdρ)]2)t=0=tr[A¨(t)]t=0.\frac{1}{2}\left(\frac{d^{2}}{dt^{2}}\left[d_{\mathrm{Bures}}\left(\rho\text{, }\rho+td\rho\right)\right]^{2}\right)_{t=0}=-\mathrm{tr}\left[\ddot{A}\left(t\right)\right]_{t=0}\text{.} (6)

From Eqs. (4) and (6), we get

gij(ρ)dρidρj=tr[A¨(t)]t=0.g_{ij}\left(\rho\right)d\rho^{i}d\rho^{j}=-\mathrm{tr}\left[\ddot{A}\left(t\right)\right]_{t=0}\text{.} (7)

Differentiating two times the relation A(t)A(t)=ρ1/2(ρ+tdρ)ρ1/2A\left(t\right)A\left(t\right)=\rho^{1/2}\left(\rho+td\rho\right)\rho^{1/2}, setting t=0t=0 and, finally, assuming ρ\rho diagonalized in the form

ρ=iλi|ii|,\rho=\sum_{i}\lambda_{i}\left|i\right\rangle\left\langle i\right|\text{,} (8)

we have

{d2dt2[A(t)A(t)]}t=0={d2dt2[ρ1/2(ρ+tdρ)ρ1/2]}t=0.\left\{\frac{d^{2}}{dt^{2}}\left[A\left(t\right)A\left(t\right)\right]\right\}_{t=0}=\left\{\frac{d^{2}}{dt^{2}}\left[\rho^{1/2}\left(\rho+td\rho\right)\rho^{1/2}\right]\right\}_{t=0}\text{.} (9)

More explicitly, we notice that

ddt[A(t)A(t)]=A˙(t)A(t)+A(t)A˙(t)\frac{d}{dt}\left[A\left(t\right)A\left(t\right)\right]=\dot{A}\left(t\right)A\left(t\right)+A\left(t\right)\dot{A}\left(t\right) (10)

and,

ddt[ρ1/2(ρ+tdρ)ρ1/2]=ρ1/2dρρ1/2.\frac{d}{dt}\left[\rho^{1/2}\left(\rho+td\rho\right)\rho^{1/2}\right]=\rho^{1/2}d\rho\rho^{1/2}\text{.} (11)

Setting t=0t=0, from Eqs. (10) and (11) we obtain

A˙(0)A(0)+A(0)A˙(0)=ρ1/2dρρ1/2.\dot{A}\left(0\right)A\left(0\right)+A\left(0\right)\dot{A}\left(0\right)=\rho^{1/2}d\rho\rho^{1/2}\text{.} (12)

After the second differentiation of A(t)A(t)A\left(t\right)A\left(t\right) and ρ1/2(ρ+tdρ)ρ1/2\rho^{1/2}\left(\rho+td\rho\right)\rho^{1/2}, we get

A¨(0)A(0)+2 A˙(0)A˙(0)+A(0)A¨(0)=0.\ddot{A}\left(0\right)A\left(0\right)+2\text{ }\dot{A}\left(0\right)\dot{A}\left(0\right)+A\left(0\right)\ddot{A}\left(0\right)=0\text{.} (13)

Multiplying both sides of Eq. (13) by A1(0)A^{-1}\left(0\right) from the right and using the cyclicity of the trace operation, we get

tr[A¨(0)]=tr[A1(0)A˙(0)2].\mathrm{tr}\left[\ddot{A}\left(0\right)\right]=-\mathrm{tr}\left[A^{-1}\left(0\right)\dot{A}\left(0\right)^{2}\right]\text{.} (14)

From Eqs. (7) and (14), the Bures metric dsBures2ds_{\mathrm{Bures}}^{2} becomes

dsBures2(ρρ+dρ)\displaystyle ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right) =gij(ρ)dρidρj\displaystyle=g_{ij}\left(\rho\right)d\rho^{i}d\rho^{j}
=tr[A¨(t)]t=0\displaystyle=-\mathrm{tr}\left[\ddot{A}\left(t\right)\right]_{t=0}
=tr[A1(0)A˙(0)2]\displaystyle=\mathrm{tr}\left[A^{-1}\left(0\right)\dot{A}\left(0\right)^{2}\right]
=ii|A1(0)A˙(0)2|i,\displaystyle=\sum_{i}\left\langle i\left|A^{-1}\left(0\right)\dot{A}\left(0\right)^{2}\right|i\right\rangle\text{,} (15)

that is,

dsBures2(ρρ+dρ)=12i,k,l[i|A1(0)|kk|A˙(0)|ll|A˙(0)|i+i|A1(0)|ll|A˙(0)|kk|A˙(0)|i].ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right)=\frac{1}{2}\sum_{i,k,l}\left[\left\langle i\left|A^{-1}\left(0\right)\right|k\right\rangle\left\langle k\left|\dot{A}\left(0\right)\right|l\right\rangle\left\langle l\left|\dot{A}\left(0\right)\right|i\right\rangle+\left\langle i\left|A^{-1}\left(0\right)\right|l\right\rangle\left\langle l\left|\dot{A}\left(0\right)\right|k\right\rangle\left\langle k\left|\dot{A}\left(0\right)\right|i\right\rangle\right]\text{.} (16)

Observe that A(0)|k=ρ|k=λk|kA\left(0\right)\left|k\right\rangle=\rho\left|k\right\rangle=\lambda_{k}\left|k\right\rangle and, therefore, A1(0)|k=ρ1|k=λk1|kA^{-1}\left(0\right)\left|k\right\rangle=\rho^{-1}\left|k\right\rangle=\lambda_{k}^{-1}\left|k\right\rangle with λk0\lambda_{k}\neq 0 for any kk. We need to find an expression for i|A˙(0)|j\left\langle i\left|\dot{A}\left(0\right)\right|j\right\rangle. From Eq. (12), we get

i|A˙(0)A(0)+A(0)A˙(0)|j=i|ρ1/2dρρ1/2|j.\left\langle i\left|\dot{A}\left(0\right)A\left(0\right)+A\left(0\right)\dot{A}\left(0\right)\right|j\right\rangle=\left\langle i\left|\rho^{1/2}d\rho\rho^{1/2}\right|j\right\rangle\text{.} (17)

We note that

i|A˙(0)A(0)+A(0)A˙(0)|j\displaystyle\left\langle i\left|\dot{A}\left(0\right)A\left(0\right)+A\left(0\right)\dot{A}\left(0\right)\right|j\right\rangle =i|A˙(0)ρ+ρA˙(0)|j\displaystyle=\left\langle i\left|\dot{A}\left(0\right)\rho+\rho\dot{A}\left(0\right)\right|j\right\rangle
=kλki|A˙(0)|kk|j+kλki|kk|A˙(0)|j\displaystyle={\displaystyle\sum\limits_{k}}\lambda_{k}\left\langle i\left|\dot{A}\left(0\right)\right|k\right\rangle\left\langle k\left|j\right.\right\rangle+{\displaystyle\sum\limits_{k}}\lambda_{k}\left\langle i\left|k\right.\right\rangle\left\langle k\left|\dot{A}\left(0\right)\right|j\right\rangle
=λji|A˙(0)|j+λii|A˙(0)|j\displaystyle=\lambda_{j}\left\langle i\left|\dot{A}\left(0\right)\right|j\right\rangle+\lambda_{i}\left\langle i\left|\dot{A}\left(0\right)\right|j\right\rangle
=(λi+λj)i|A˙(0)|j,\displaystyle=\left(\lambda_{i}+\lambda_{j}\right)\left\langle i\left|\dot{A}\left(0\right)\right|j\right\rangle\text{,} (18)

that is,

i|A˙(0)A(0)+A(0)A˙(0)|j=(λi+λj)i|A˙(0)|j.\left\langle i\left|\dot{A}\left(0\right)A\left(0\right)+A\left(0\right)\dot{A}\left(0\right)\right|j\right\rangle=\left(\lambda_{i}+\lambda_{j}\right)\left\langle i\left|\dot{A}\left(0\right)\right|j\right\rangle\text{.} (19)

Moreover, we observe that

i|ρ1/2dρρ1/2|j\displaystyle\left\langle i\left|\rho^{1/2}d\rho\rho^{1/2}\right|j\right\rangle =i|(kλk|kk|)dρ(mλm|mm|)|\displaystyle=\left\langle i\left|\left(\sum_{k}\sqrt{\lambda_{k}}\left|k\right\rangle\left\langle k\right|\right)d\rho\left(\sum_{m}\sqrt{\lambda_{m}}\left|m\right\rangle\left\langle m\right|\right)\right|\right\rangle
=k,mλkλmi|kk|dρ|mm|j\displaystyle=\sum_{k,m}\sqrt{\lambda_{k}}\sqrt{\lambda_{m}}\left\langle i\left|k\right.\right\rangle\left\langle k\left|d\rho\right|m\right\rangle\left\langle m\left|j\right.\right\rangle
=k,mλkλmδikδmjk|dρ|m\displaystyle=\sum_{k,m}\sqrt{\lambda_{k}}\sqrt{\lambda_{m}}\delta_{ik}\delta_{mj}\left\langle k\left|d\rho\right|m\right\rangle
=λiλji|dρ|j,\displaystyle=\sqrt{\lambda_{i}}\sqrt{\lambda_{j}}\left\langle i\left|d\rho\right|j\right\rangle\text{,} (20)

that is,

i|ρ1/2dρρ1/2|j=λiλji|dρ|j.\left\langle i\left|\rho^{1/2}d\rho\rho^{1/2}\right|j\right\rangle=\sqrt{\lambda_{i}}\sqrt{\lambda_{j}}\left\langle i\left|d\rho\right|j\right\rangle\text{.} (21)

Using Eqs. (17), (19), and (21), we have

i|A˙(0)|j=λiλjλi+λji|dρ|j.\left\langle i\left|\dot{A}\left(0\right)\right|j\right\rangle=\frac{\sqrt{\lambda_{i}}\sqrt{\lambda_{j}}}{\lambda_{i}+\lambda_{j}}\left\langle i\left|d\rho\right|j\right\rangle\text{.} (22)

Finally, using Eq. (22), dsBures2ds_{\mathrm{Bures}}^{2} in Eq. (16) becomes

dsBures2(ρρ+dρ)=12k,lλl+λk(λl+λk)2|k|dρ|l|2,ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right)=\frac{1}{2}\sum_{k,l}\frac{\lambda_{l}+\lambda_{k}}{\left(\lambda_{l}+\lambda_{k}\right)^{2}}\left|\left\langle k\left|d\rho\right|l\right\rangle\right|^{2}\text{,} (23)

that is, relabelling the dummy indices (i.e., kik\rightarrow i and ljl\rightarrow j),

dsBures2(ρρ+dρ)=12i,j|i|dρ|j|2λi+λj.ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right)=\frac{1}{2}\sum_{i,j}\frac{\left|\left\langle i\left|d\rho\right|j\right\rangle\right|^{2}}{\lambda_{i}+\lambda_{j}}\text{.} (24)

The derivation of Eq. (24) ends our revisitation of Hübner’s original analysis presented in Ref. hubner92 . Note that in obtaining Eq. (24), there is no need to introduce any Hamiltonian H that might be responsible for the changes from ρ\rho to ρ+dρ\rho+d\rho. For this reason, the expression of the Bures metric in Eq. (24) is said to be general.

II.2 The explicit derivation of Zanardi’s general expression

In Ref. zanardi07 , Zanardi and collaborators provided an alternative expression of the Bures metric in Eq. (24). Interestingly, in this alternative expression, the Bures metric in Eq. (24) can be decomposed in its classical and nonclassical parts. To begin, in view of future geometric investigations in statistical physics, let us use a different notation and rewrite the Bures metric in Eq. (24) as

dsBures2(ρρ+dρ)=def12nm|m|dρ|n|2pm+pn,ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right)\overset{\text{def}}{=}\frac{1}{2}\sum_{n\text{, }m}\frac{\left|\left\langle m|d\rho|n\right\rangle\right|^{2}}{p_{m}+p_{n}}\text{,} (25)

with 1m1\leq m, nNn\leq N. Let us assume that the quantities ρ\rho and dρd\rho in Eq. (25) are given by,

ρ=defnpn|nn|,\rho\overset{\text{def}}{=}\sum_{n}p_{n}\left|n\right\rangle\left\langle n\right|\text{,} (26)

and

dρ=defn[dpn|nn|+pn|dnn|+pn|ndn|],d\rho\overset{\text{def}}{=}\sum_{n}\left[dp_{n}\left|n\right\rangle\left\langle n\right|+p_{n}\left|dn\right\rangle\left\langle n\right|+p_{n}\left|n\right\rangle\left\langle dn\right|\right]\text{,} (27)

respectively, with n|m=δn,m\left\langle n\left|m\right.\right\rangle=\delta_{n,m}. Let us use Eqs. (27) and (25) to find a more explicit expression for the Bures metric dsBures2(ρρ+dρ)ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right). Observe that the quantity i|dρ|j\left\langle i\left|d\rho\right|j\right\rangle can be recast as,

i|dρ|j\displaystyle\left\langle i\left|d\rho\right|j\right\rangle =i|(n[dpn|nn|+pn|dnn|+pn|ndn|])|j\displaystyle=\left\langle i\left|\left(\sum_{n}\left[dp_{n}\left|n\right\rangle\left\langle n\right|+p_{n}\left|dn\right\rangle\left\langle n\right|+p_{n}\left|n\right\rangle\left\langle dn\right|\right]\right)\right|j\right\rangle
=ndpni|nn|j+pni|dnn|j+pni|ndn|j\displaystyle=\sum_{n}dp_{n}\left\langle i|n\right\rangle\left\langle n|j\right\rangle+p_{n}\left\langle i|dn\right\rangle\left\langle n|j\right\rangle+p_{n}\left\langle i|n\right\rangle\left\langle dn|j\right\rangle
=ndpnδinδnj+pni|dnδnj+pnδindn|j\displaystyle=\sum_{n}dp_{n}\delta_{in}\delta_{nj}+p_{n}\left\langle i|dn\right\rangle\delta_{nj}+p_{n}\delta_{in}\left\langle dn|j\right\rangle
=dpiδij+pji|dj+pidi|j,\displaystyle=dp_{i}\delta_{ij}+p_{j}\left\langle i|dj\right\rangle+p_{i}\left\langle di|j\right\rangle\text{,} (28)

that is,

i|dρ|j=dpiδij+pji|dj+pidi|j.\left\langle i\left|d\rho\right|j\right\rangle=dp_{i}\delta_{ij}+p_{j}\left\langle i|dj\right\rangle+p_{i}\left\langle di|j\right\rangle\text{.} (29)

Note that the orthonormality condition i|j=δij\left\langle i|j\right\rangle=\delta_{ij} implies di|j+i|dj=0\left\langle di|j\right\rangle+\left\langle i|dj\right\rangle=0, that is

di|j=i|dj.\left\langle di|j\right\rangle=-\left\langle i|dj\right\rangle\text{.} (30)

Using Eq. (30), i|dρ|j\left\langle i\left|d\rho\right|j\right\rangle in Eq. (29) becomes

i|dρ|j=dpiδij+pji|djpii|dj=δijdpi+(pjpi)i|dj,\left\langle i\left|d\rho\right|j\right\rangle=dp_{i}\delta_{ij}+p_{j}\left\langle i|dj\right\rangle-p_{i}\left\langle i|dj\right\rangle=\delta_{ij}dp_{i}+\left(p_{j}-p_{i}\right)\left\langle i|dj\right\rangle\text{,}

that is,

i|dρ|j=δijdpi+(pjpi)i|dj.\left\langle i\left|d\rho\right|j\right\rangle=\delta_{ij}dp_{i}+\left(p_{j}-p_{i}\right)\left\langle i|dj\right\rangle\text{.} (31)

Making use of Eq. (31), we can now find an explicit expression of the quantity |m|dρ|n|2\left|\left\langle m|d\rho|n\right\rangle\right|^{2} in Eq. (25). Indeed, from Eq. (31) we have

|m|dρ|n|2\displaystyle\left|\left\langle m|d\rho|n\right\rangle\right|^{2} =m|dρ|nm|dρ|n\displaystyle=\left\langle m|d\rho|n\right\rangle\left\langle m|d\rho|n\right\rangle^{\ast}
=m|dρ|nn|dρ|m\displaystyle=\left\langle m|d\rho|n\right\rangle\left\langle n|d\rho|m\right\rangle
=[δmndpm+(pnpm)m|dn][δnmdpn+(pmpn)n|dm]\displaystyle=\left[\delta_{mn}dp_{m}+\left(p_{n}-p_{m}\right)\left\langle m|dn\right\rangle\right]\left[\delta_{nm}dp_{n}+\left(p_{m}-p_{n}\right)\left\langle n|dm\right\rangle\right]
=δmndpmδnmdpn+δmndpm(pmpn)n|dm+(pnpm)m|dnδnmdpn+\displaystyle=\delta_{mn}dp_{m}\delta_{nm}dp_{n}+\delta_{mn}dp_{m}\left(p_{m}-p_{n}\right)\left\langle n|dm\right\rangle+\left(p_{n}-p_{m}\right)\left\langle m|dn\right\rangle\delta_{nm}dp_{n}+
+(pnpm)m|dn(pmpn)n|dm\displaystyle+\left(p_{n}-p_{m}\right)\left\langle m|dn\right\rangle\left(p_{m}-p_{n}\right)\left\langle n|dm\right\rangle
=δnmdpn2(pnpm)2m|dnn|dm,\displaystyle=\delta_{nm}dp_{n}^{2}-\left(p_{n}-p_{m}\right)^{2}\left\langle m|dn\right\rangle\left\langle n|dm\right\rangle\text{,} (32)

that is,

|m|dρ|n|2=δnmdpn2(pnpm)2m|dnn|dm.\left|\left\langle m|d\rho|n\right\rangle\right|^{2}=\delta_{nm}dp_{n}^{2}-\left(p_{n}-p_{m}\right)^{2}\left\langle m|dn\right\rangle\left\langle n|dm\right\rangle\text{.} (33)

Using Eq. (30), Eq. (33) reduces to

|m|dρ|n|2\displaystyle\left|\left\langle m|d\rho|n\right\rangle\right|^{2} =δnmdpn2+(pnpm)2dm|nn|dm\displaystyle=\delta_{nm}dp_{n}^{2}+\left(p_{n}-p_{m}\right)^{2}\left\langle dm|n\right\rangle\left\langle n|dm\right\rangle
=δnmdpn2+(pnpm)2|n|dm|2,\displaystyle=\delta_{nm}dp_{n}^{2}+\left(p_{n}-p_{m}\right)^{2}\left|\left\langle n|dm\right\rangle\right|^{2}\text{,} (34)

that is,

|m|dρ|n|2=δnmdpn2+(pnpm)2|n|dm|2.\left|\left\langle m|d\rho|n\right\rangle\right|^{2}=\delta_{nm}dp_{n}^{2}+\left(p_{n}-p_{m}\right)^{2}\left|\left\langle n|dm\right\rangle\right|^{2}\text{.} (35)

Finally, substituting Eq. (35) into Eq. (25), dsBures2(ρρ+dρ)ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right) becomes

dsBures2(ρρ+dρ)\displaystyle ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right) =12nmδnmdpn2+(pnpm)2|n|dm|2pm+pn\displaystyle=\frac{1}{2}\sum_{n\text{, }m}\frac{\delta_{nm}dp_{n}^{2}+\left(p_{n}-p_{m}\right)^{2}\left|\left\langle n|dm\right\rangle\right|^{2}}{p_{m}+p_{n}}
=14ndpn2pn+12nm|n|dm|2(pnpm)2pm+pn,\displaystyle=\frac{1}{4}\sum_{n}\frac{dp_{n}^{2}}{p_{n}}+\frac{1}{2}\sum_{n\neq m}\left|\left\langle n|dm\right\rangle\right|^{2}\frac{\left(p_{n}-p_{m}\right)^{2}}{p_{m}+p_{n}}\text{,} (36)

that is,

dsBures2(ρρ+dρ)=14ndpn2pn+12nm|n|dm|2(pnpm)2pm+pn.ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right)=\frac{1}{4}\sum_{n}\frac{dp_{n}^{2}}{p_{n}}+\frac{1}{2}\sum_{n\neq m}\left|\left\langle n|dm\right\rangle\right|^{2}\frac{\left(p_{n}-p_{m}\right)^{2}}{p_{m}+p_{n}}\text{.} (37)

Eq. (37) is the explicit expression of dsBures2(ρρ+dρ)ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right) we were searching for. As a side remark, note that if both |n\left|n\right\rangle and |mker(ρ)\left|m\right\rangle\in\ker\left(\rho\right), we have that n|dρ|m=0\left\langle n|d\rho|m\right\rangle=0. Indeed, from ρ|m=|0\rho\left|m\right\rangle=\left|0\right\rangle, we have dρ|m+ρ|dm=|0d\rho\left|m\right\rangle+\rho\left|dm\right\rangle=\left|0\right\rangle. Therefore, we have n|dρ|m+n|ρ|dm=n|0\left\langle n|d\rho|m\right\rangle+\left\langle n|\rho|dm\right\rangle=\left\langle n|0\right\rangle, that is, n|dρ|m=0\left\langle n|d\rho|m\right\rangle=0 since n|0=0\left\langle n|0\right\rangle=0 and n|ρ|dm=0\left\langle n|\rho|dm\right\rangle=0. The expression of the Bures metric in Eq. (37) can be regarded as given by two contribution, a classical and a nonclassical term. The first term in dsBures2(ρρ+dρ)ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right) in Eq. (37) is the classical one and is represented by the classical Fisher-Rao information metric between the two probability distributions {pn}1nN\left\{p_{n}\right\}_{1\leq n\leq N} and {pn+dpn}1nN\left\{p_{n}+dp_{n}\right\}_{1\leq n\leq N}. The second term is the nonclassical one and emerges from the noncommutativity of the density matrices ρ\rho and ρ+dρ\rho+d\rho (i.e., [ρρ+dρ]=[ρdρ]0\left[\rho\text{, }\rho+d\rho\right]=\left[\rho\text{, }d\rho\right]\neq 0, in general). When [ρρ+dρ]=0\left[\rho\text{, }\rho+d\rho\right]=0, the problem becomes classical and the Bures metric reduces to the classical Fisher-Rao metric.

II.3 The explicit derivation of Zanardi’s expression for thermal states

In what follows, we specialize on the functional form of dsBures2(ρρ+dρ)ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right) in Eq. (37) for thermal states. Specifically, let us focus on mixed quantum states ρ(βλ)\rho\left(\beta\text{, }\lambda\right) of the form,

ρ(βλ)=defeβH(λ)𝒵=eβH(λ)tr(eβH(λ)),\rho\left(\beta\text{, }\lambda\right)\overset{\text{def}}{=}\frac{e^{-\beta\mathrm{H}\left(\lambda\right)}}{\mathcal{Z}}=\frac{e^{-\beta\mathrm{H}\left(\lambda\right)}}{\mathrm{tr}\left(e^{-\beta\mathrm{H}\left(\lambda\right)}\right)}\text{,} (38)

with 𝒵=deftr(eβH(λ))\mathcal{Z}\overset{\text{def}}{=}\mathrm{tr}\left(e^{-\beta\mathrm{H}\left(\lambda\right)}\right) denoting the partition function of the system. The Hamiltonian H\mathrm{H} in Eq. (38) depends on a set of parameters {λ}\left\{\lambda\right\} and is such that H|n=En|n\mathrm{H}\left|n\right\rangle=E_{n}\left|n\right\rangle or, equivalently

H=nEn|nn|,\mathrm{H}=\sum_{n}E_{n}\left|n\right\rangle\left\langle n\right|\text{,} (39)

with 1nN1\leq n\leq N. Using the spectral decomposition of H\mathrm{H} in Eq. (39), ρ(βλ)\rho\left(\beta\text{, }\lambda\right) in Eq. (38) can be recast as

ρ(βλ)=npn|nn|=neβEn𝒵|nn|,\rho\left(\beta\text{, }\lambda\right)=\sum_{n}p_{n}\left|n\right\rangle\left\langle n\right|=\sum_{n}\frac{e^{-\beta E_{n}}}{\mathcal{Z}}\left|n\right\rangle\left\langle n\right|\text{,} (40)

with pn=defeβEn/𝒵p_{n}\overset{\text{def}}{=}e^{-\beta E_{n}}/\mathcal{Z}. Note that the λ\lambda-dependence of ρ\rho in Eq. (40) appears, in general, in both En=En(λ)E_{n}=E_{n}\left(\lambda\right) and |n=|n(λ)\left|n\right\rangle=\left|n\left(\lambda\right)\right\rangle. Before presenting the general case where both β\beta and the set of {λ}\left\{\lambda\right\} can change, we focus on the sub-case where β\beta is kept constant while {λ}\left\{\lambda\right\} is allowed to change.

II.3.1 Case: β\beta-constant and λ\lambda-nonconstant

In what follows, assuming that β\beta is fixed, we wish to find the expression of dsBures2(ρρ+dρ)ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right) in Eq. (37) when ρ\rho is given as in Eq. (40).  Clearly, we note that the two key quantities that we need to find are dpi2dp_{i}^{2} and i|dj\left\langle i|dj\right\rangle. Let us start with the latter. From H|j=Ej|j\mathrm{H}\left|j\right\rangle=E_{j}\left|j\right\rangle, we have dH|j+H|dj=dEj|j+Ej|djd\mathrm{H}\left|j\right\rangle+\mathrm{H}\left|dj\right\rangle=dE_{j}\left|j\right\rangle+E_{j}\left|dj\right\rangle. Assuming iji\neq j, we have

i|dH|j+i|H|dj=i|dEj|j+i|Ej|dj=dEjδij+Eji|dj=Eji|dj,\left\langle i\right|d\mathrm{H}\left|j\right\rangle+\left\langle i\right|\mathrm{H}\left|dj\right\rangle=\left\langle i\right|dE_{j}\left|j\right\rangle+\left\langle i\right|E_{j}\left|dj\right\rangle=dE_{j}\delta_{ij}+E_{j}\left\langle i|dj\right\rangle=E_{j}\left\langle i|dj\right\rangle\text{,} (41)

that is,

i|dH|j+i|H|dj=Eji|dj.\left\langle i\right|d\mathrm{H}\left|j\right\rangle+\left\langle i\right|\mathrm{H}\left|dj\right\rangle=E_{j}\left\langle i|dj\right\rangle\text{.} (42)

Observe that,

i|H|dj=dj|H|i=dj|H|i=Eidj|i=Eii|dj.\left\langle i\right|\mathrm{H}\left|dj\right\rangle=\left\langle dj\right|\mathrm{H}^{\dagger}\left|i\right\rangle^{\ast}=\left\langle dj\right|\mathrm{H}\left|i\right\rangle^{\ast}=E_{i}\left\langle dj|i\right\rangle^{\ast}=E_{i}\left\langle i|dj\right\rangle\text{.} (43)

Substituting Eq. (43) into Eq. (42), we get

i|dH|j+Eii|dj=Eji|dj,\left\langle i\right|d\mathrm{H}\left|j\right\rangle+E_{i}\left\langle i|dj\right\rangle=E_{j}\left\langle i|dj\right\rangle\text{,} (44)

that is,

i|dj=i|dH|jEjEi.\left\langle i|dj\right\rangle=\frac{\left\langle i\right|d\mathrm{H}\left|j\right\rangle}{E_{j}-E_{i}}\text{.} (45)

Eq. (45) is the first piece of relevant information we were looking for. Let us not focus on calculating dpi2dp_{i}^{2}. From pi=defeβEi/𝒵p_{i}\overset{\text{def}}{=}e^{-\beta E_{i}}/\mathcal{Z}, we get

dpi\displaystyle dp_{i} =d(eβEi𝒵)\displaystyle=d\left(\frac{e^{-\beta E_{i}}}{\mathcal{Z}}\right)
=1𝒵d(eβEi)+eβEid(1𝒵)\displaystyle=\frac{1}{\mathcal{Z}}d\left(e^{-\beta E_{i}}\right)+e^{-\beta E_{i}}d\left(\frac{1}{\mathcal{Z}}\right)
=1𝒵ddEi(eβEi)dEi+eβEi(1𝒵2d𝒵)\displaystyle=\frac{1}{\mathcal{Z}}\frac{d}{dE_{i}}\left(e^{-\beta E_{i}}\right)dE_{i}+e^{-\beta E_{i}}\left(-\frac{1}{\mathcal{Z}^{2}}d\mathcal{Z}\right)
=βeβEi𝒵dEieβEi𝒵d𝒵𝒵\displaystyle=-\beta\frac{e^{-\beta E_{i}}}{\mathcal{Z}}dE_{i}-\frac{e^{-\beta E_{i}}}{\mathcal{Z}}\frac{d\mathcal{Z}}{\mathcal{Z}}
=βeβEi𝒵dEieβEi𝒵j(d𝒵dEjdEj𝒵),\displaystyle=-\beta\frac{e^{-\beta E_{i}}}{\mathcal{Z}}dE_{i}-\frac{e^{-\beta E_{i}}}{\mathcal{Z}}\sum_{j}\left(\frac{d\mathcal{Z}}{dE_{j}}\frac{dE_{j}}{\mathcal{Z}}\right)\text{,} (46)

that is,

dpi=βeβEi𝒵dEieβEi𝒵j(d𝒵dEjdEj𝒵).dp_{i}=-\beta\frac{e^{-\beta E_{i}}}{\mathcal{Z}}dE_{i}-\frac{e^{-\beta E_{i}}}{\mathcal{Z}}\sum_{j}\left(\frac{d\mathcal{Z}}{dE_{j}}\frac{dE_{j}}{\mathcal{Z}}\right)\text{.} (47)

At this point, note that

jd𝒵dEjdEj𝒵\displaystyle\sum_{j}\frac{d\mathcal{Z}}{dE_{j}}\frac{dE_{j}}{\mathcal{Z}} =jddEj(keβEk)dEj𝒵\displaystyle=\sum_{j}\frac{d}{dE_{j}}\left(\sum_{k}e^{-\beta E_{k}}\right)\frac{dE_{j}}{\mathcal{Z}}
=βjeβEj𝒵dEj\displaystyle=-\beta\sum_{j}\frac{e^{-\beta E_{j}}}{\mathcal{Z}}dE_{j}
=βjpjdEj.\displaystyle=-\beta\sum_{j}p_{j}dE_{j}\text{.} (48)

Substituting Eq. (48) into Eq. (47), we get

dpi=βpidEi+βpijpjdEj=βpi[dEijpjdEj].dp_{i}=-\beta p_{i}dE_{i}+\beta p_{i}\sum_{j}p_{j}dE_{j}=-\beta p_{i}\left[dE_{i}-\sum_{j}p_{j}dE_{j}\right]\text{.} (49)

Eq. (49) is the second piece of relevant information we were looking for. We can now calculate dsBures2(ρρ+dρ)ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right) in Eq. (37) by means of Eqs. (45) and (49). We obtain,

14idpi2pi\displaystyle\frac{1}{4}\sum_{i}\frac{dp_{i}^{2}}{p_{i}} =14iβ2pi2pi[dEijpjdEj]2\displaystyle=\frac{1}{4}\sum_{i}\beta^{2}\frac{p_{i}^{2}}{p_{i}}\left[dE_{i}-\sum_{j}p_{j}dE_{j}\right]^{2}
=β24ipi[dEidEβ]2\displaystyle=\frac{\beta^{2}}{4}\sum_{i}p_{i}\left[dE_{i}-\left\langle dE\right\rangle_{\beta}\right]^{2}
=β24(dE2βdEβ2)\displaystyle=\frac{\beta^{2}}{4}\left(\left\langle dE^{2}\right\rangle_{\beta}-\left\langle dE\right\rangle_{\beta}^{2}\right)
=β24(dHd2βdHdβ2),\displaystyle=\frac{\beta^{2}}{4}\left(\left\langle d\mathrm{H}_{d}^{2}\right\rangle_{\beta}-\left\langle d\mathrm{H}_{d}\right\rangle_{\beta}^{2}\right)\text{,} (50)

that is,

14idpi2pi=β24(dHd2βdHdβ2).\frac{1}{4}\sum_{i}\frac{dp_{i}^{2}}{p_{i}}=\frac{\beta^{2}}{4}\left(\left\langle d\mathrm{H}_{d}^{2}\right\rangle_{\beta}-\left\langle d\mathrm{H}_{d}\right\rangle_{\beta}^{2}\right)\text{.} (51)

The quantity dHdd\mathrm{H}_{d} in Eq. (51) is defined as

dHd=defjdEj|jj|,d\mathrm{H}_{d}\overset{\text{def}}{=}\sum_{j}dE_{j}\left|j\right\rangle\left\langle j\right|\text{,} (52)

and is different from dHd\mathrm{H}. For clarity, we also observe that

dHdβ=defipidEi, and dHd2β=defipidEi2.\left\langle d\mathrm{H}_{d}\right\rangle_{\beta}\overset{\text{def}}{=}\sum_{i}p_{i}dE_{i}\text{, and }\left\langle d\mathrm{H}_{d}^{2}\right\rangle_{\beta}\overset{\text{def}}{=}\sum_{i}p_{i}dE_{i}^{2}\text{.} (53)

Finally, using Eqs. (45) and (51), we get

dsBures2(ρρ+dρ)=β24(dHd2βdHdβ2)+12nm|n|dH|mEnEm|2(eβEneβEm)2𝒵(eβEn+eβEm).ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right)=\frac{\beta^{2}}{4}\left(\left\langle d\mathrm{H}_{d}^{2}\right\rangle_{\beta}-\left\langle d\mathrm{H}_{d}\right\rangle_{\beta}^{2}\right)+\frac{1}{2}\sum_{n\neq m}\left|\frac{\left\langle n|d\mathrm{H}|m\right\rangle}{E_{n}-E_{m}}\right|^{2}\frac{\left(e^{-\beta E_{n}}-e^{-\beta E_{m}}\right)^{2}}{\mathcal{Z}\left(e^{-\beta E_{n}}+e^{-\beta E_{m}}\right)}\text{.} (54)

The Bures metric dsBures2(ρρ+dρ)ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right) in Eq. (54) is the Bures metric in Eq. (37) between two mixed thermal states ρ(βλ)\rho\left(\beta\text{, }\lambda\right) and (ρ+dρ)(βλ)\left(\rho+d\rho\right)\left(\beta\text{, }\lambda\right) when only changes in λ\lambda are permitted.

II.3.2 Case: β\beta-nonconstant and λ\lambda-nonconstant

In what follows, we consider the general case where both β\beta and the set of {λ}\left\{\lambda\right\} can change. The sub-case where β\beta changes while the set of {λ}\left\{\lambda\right\} is kept constant is then obtained as a special case. For simplicity, let us assume we have two parameters, β\beta and a single parameter λ\lambda that we denote with hh (a magnetic field intensity, for instance). In this two-dimensional parametric case, we generally have that

dsBures2(βh)=(dβdh)(gββgβhghβghh)(dβdh)=gββdβ2+ghhdh2+2gβhdβdh,ds_{\mathrm{Bures}}^{2}\left(\beta\text{, }h\right)=\left(\begin{array}[c]{cc}d\beta&dh\end{array}\right)\left(\begin{array}[c]{cc}g_{\beta\beta}&g_{\beta h}\\ g_{h\beta}&g_{hh}\end{array}\right)\left(\begin{array}[c]{c}d\beta\\ dh\end{array}\right)=g_{\beta\beta}d\beta^{2}+g_{hh}dh^{2}+2g_{\beta h}d\beta dh\text{,} (55)

where we used the fact that ghβ=gβhg_{h\beta}=g_{\beta h}. From Eq. (55), we note that

dsBures2(βh)={gββ(βh)dβ2, if h=const.ghh(βh)dh2, if β=const.gββ(βh)dβ2+ghh(βh)dh2+2gβh(βh)dβdh, if βconst. and hconst..ds_{\mathrm{Bures}}^{2}\left(\beta\text{, }h\right)=\left\{\begin{array}[c]{c}g_{\beta\beta}\left(\beta\text{, }h\right)d\beta^{2}\text{, if }h=\text{{const}.}\\ g_{hh}\left(\beta\text{, }h\right)dh^{2}\text{, if }\beta=\text{{const}.}\\ g_{\beta\beta}\left(\beta\text{, }h\right)d\beta^{2}+g_{hh}\left(\beta\text{, }h\right)dh^{2}+2g_{\beta h}\left(\beta\text{, }h\right)d\beta dh\text{, if }\beta\neq\text{{const}. and }h\neq\text{{const}.}\end{array}\right.\text{.} (56)

Recalling dsBures2(ρρ+dρ)ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right) in Eq. (37), we start by calculating the expression of dpndp_{n} with pn=pn(hβ)=defeβEn/𝒵p_{n}=p_{n}\left(h\text{, }\beta\right)\overset{\text{def}}{=}e^{-\beta E_{n}}/\mathcal{Z}. We observe that dpndp_{n} can be written as,

dpn=pnhdh+pnβdβ,dp_{n}=\frac{\partial p_{n}}{\partial h}dh+\frac{\partial p_{n}}{\partial\beta}d\beta\text{,} (57)

where pn/β\partial p_{n}/\partial\beta is given by,

pnβ\displaystyle\frac{\partial p_{n}}{\partial\beta} =β(eβEn𝒵)\displaystyle=\frac{\partial}{\partial\beta}\left(\frac{e^{-\beta E_{n}}}{\mathcal{Z}}\right)
=1𝒵β(eβEn)+eβEnβ(1𝒵)\displaystyle=\frac{1}{\mathcal{Z}}\frac{\partial}{\partial\beta}\left(e^{-\beta E_{n}}\right)+e^{-\beta E_{n}}\frac{\partial}{\partial\beta}\left(\frac{1}{\mathcal{Z}}\right)
=En𝒵eβEn+eβEn𝒵(1𝒵)𝒵β\displaystyle=-\frac{E_{n}}{\mathcal{Z}}e^{-\beta E_{n}}+e^{-\beta E_{n}}\frac{\partial}{\partial\mathcal{Z}}\left(\frac{1}{\mathcal{Z}}\right)\frac{\partial\mathcal{Z}}{\partial\beta}
=pnEneβEn𝒵1𝒵𝒵β\displaystyle=-p_{n}E_{n}-\frac{e^{-\beta E_{n}}}{\mathcal{Z}}\frac{1}{\mathcal{Z}}\frac{\partial\mathcal{Z}}{\partial\beta}
=pnEnpnln𝒵β\displaystyle=-p_{n}E_{n}-p_{n}\frac{\partial\ln\mathcal{Z}}{\partial\beta}
=pnEn+pn1𝒵nEneβEn\displaystyle=-p_{n}E_{n}+p_{n}\frac{1}{\mathcal{Z}}\sum_{n}E_{n}e^{-\beta E_{n}}
=pnEn+pnnpnEn\displaystyle=-p_{n}E_{n}+p_{n}\sum_{n}p_{n}E_{n}
=pnEn+pnH,\displaystyle=-p_{n}E_{n}+p_{n}\left\langle\mathrm{H}\right\rangle\text{,} (58)

that is,

pnβdβ=pn[HEn]dβ.\frac{\partial p_{n}}{\partial\beta}d\beta=p_{n}\left[\left\langle\mathrm{H}\right\rangle-E_{n}\right]d\beta\text{.} (59)

Note that the expectation value H\left\langle\mathrm{H}\right\rangle in Eq. (59) is defined as H\left\langle\mathrm{H}\right\rangle =defnpnEn\overset{\text{def}}{=}\sum_{n}p_{n}E_{n}. From Eq. (49), we also have

pnhdh=βpn[EnhjpjEjh]dh=βpn[(hH)dhEn]dh,\frac{\partial p_{n}}{\partial h}dh=-\beta p_{n}\left[\frac{\partial E_{n}}{\partial h}-\sum_{j}p_{j}\frac{\partial E_{j}}{\partial h}\right]dh=\beta p_{n}\left[\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle-\partial_{h}E_{n}\right]dh\text{,} (60)

where (hH)d\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle is defined as

(hH)d=defjpjhEj.\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle\overset{\text{def}}{=}\sum_{j}p_{j}\partial_{h}E_{j}\text{.} (61)

Using Eqs. (57), (59), and (60), we wish to calculate the term (1/4)ndpn2/pn\left(1/4\right)\sum_{n}dp_{n}^{2}/p_{n} in dsBures2(ρρ+dρ)ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right) in Eq. (37). Let us begin by observing that

dpn2=(pnhdh+pnβdβ)2=(hpndh)2+(βpndβ)2+2βpnhpndβdh.dp_{n}^{2}=\left(\frac{\partial p_{n}}{\partial h}dh+\frac{\partial p_{n}}{\partial\beta}d\beta\right)^{2}=\left(\partial_{h}p_{n}dh\right)^{2}+\left(\partial_{\beta}p_{n}d\beta\right)^{2}+2\partial_{\beta}p_{n}\partial_{h}p_{n}d\beta dh\text{.} (62)

Therefore, we get

14ndpn2pn\displaystyle\frac{1}{4}\sum_{n}\frac{dp_{n}^{2}}{p_{n}} =14n(hpndh)2+(βpndβ)2+2βpnhpndβdhpn\displaystyle=\frac{1}{4}\sum_{n}\frac{\left(\partial_{h}p_{n}dh\right)^{2}+\left(\partial_{\beta}p_{n}d\beta\right)^{2}+2\partial_{\beta}p_{n}\partial_{h}p_{n}d\beta dh}{p_{n}}
=14n(hpn)2pndh2+14n(βpn)2pndβ2+14n2βpnhpnpndβdh.\displaystyle=\frac{1}{4}\sum_{n}\frac{\left(\partial_{h}p_{n}\right)^{2}}{p_{n}}dh^{2}+\frac{1}{4}\sum_{n}\frac{\left(\partial_{\beta}p_{n}\right)^{2}}{p_{n}}d\beta^{2}+\frac{1}{4}\sum_{n}\frac{2\partial_{\beta}p_{n}\partial_{h}p_{n}}{p_{n}}d\beta dh\text{.} (63)

First, note that

14n(βpn)2pndβ2=14[H2H2]dβ2,\frac{1}{4}\sum_{n}\frac{\left(\partial_{\beta}p_{n}\right)^{2}}{p_{n}}d\beta^{2}=\frac{1}{4}\left[\left\langle\mathrm{H}^{2}\right\rangle-\left\langle\mathrm{H}\right\rangle^{2}\right]d\beta^{2}\text{,} (64)

where H\left\langle\mathrm{H}\right\rangle and H2\left\langle\mathrm{H}^{2}\right\rangle are defined as

H=defipiEi, and H2=defipiEi2,\left\langle\mathrm{H}\right\rangle\overset{\text{def}}{=}\sum_{i}p_{i}E_{i}\text{, and }\left\langle\mathrm{H}^{2}\right\rangle\overset{\text{def}}{=}\sum_{i}p_{i}E_{i}^{2}\text{,} (65)

respectively. Indeed, using Eq. (59), we have

n(βpn)2pn\displaystyle\sum_{n}\frac{\left(\partial_{\beta}p_{n}\right)^{2}}{p_{n}} =npn2[HEn]2pn\displaystyle=\sum_{n}\frac{p_{n}^{2}\left[\left\langle\mathrm{H}\right\rangle-E_{n}\right]^{2}}{p_{n}}
=npn2H2+pn2En22Hpn2Enpn\displaystyle=\sum_{n}\frac{p_{n}^{2}\left\langle\mathrm{H}\right\rangle^{2}+p_{n}^{2}E_{n}^{2}-2\left\langle\mathrm{H}\right\rangle p_{n}^{2}E_{n}}{p_{n}}
=H2+H22H2\displaystyle=\left\langle\mathrm{H}\right\rangle^{2}+\left\langle\mathrm{H}^{2}\right\rangle-2\left\langle\mathrm{H}\right\rangle^{2}
=H2H2.\displaystyle=\left\langle\mathrm{H}^{2}\right\rangle-\left\langle\mathrm{H}\right\rangle^{2}\text{.} (66)

Second, observe that

14n(hpn)2pndh2=14β2{[(hH)d]2(hH)d2}dh2,\frac{1}{4}\sum_{n}\frac{\left(\partial_{h}p_{n}\right)^{2}}{p_{n}}dh^{2}=\frac{1}{4}\beta^{2}\left\{\left\langle\left[\left(\partial_{h}\mathrm{H}\right)_{d}\right]^{2}\right\rangle-\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle^{2}\right\}dh^{2}\text{,} (67)

where (hH)d\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle is given in Eq. (61) and [(hH)d]2\left\langle\left[\left(\partial_{h}\mathrm{H}\right)_{d}\right]^{2}\right\rangle is defined as

[(hH)d]2=defipi(hEi)2.\left\langle\left[\left(\partial_{h}\mathrm{H}\right)_{d}\right]^{2}\right\rangle\overset{\text{def}}{=}\sum_{i}p_{i}\left(\partial_{h}E_{i}\right)^{2}\text{.} (68)

Indeed, using Eq. (60), we have

n(hpn)2pn\displaystyle\sum_{n}\frac{\left(\partial_{h}p_{n}\right)^{2}}{p_{n}} =n(βpn)2[(hH)dhEn]2pn\displaystyle=\sum_{n}\frac{\left(\beta p_{n}\right)^{2}\left[\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle-\partial_{h}E_{n}\right]^{2}}{p_{n}}
=β2n[pn(hH)d2+pn(hEn)22pn(hH)dhEn]\displaystyle=\beta^{2}\sum_{n}\left[p_{n}\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle^{2}+p_{n}\left(\partial_{h}E_{n}\right)^{2}-2p_{n}\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle\partial_{h}E_{n}\right]
=β2{(hH)d2+[(hH)d]22(hH)d2}\displaystyle=\beta^{2}\left\{\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle^{2}+\left\langle\left[\left(\partial_{h}\mathrm{H}\right)_{d}\right]^{2}\right\rangle-2\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle^{2}\right\}
=β2{[(hH)d]2(hH)d2}.\displaystyle=\beta^{2}\left\{\left\langle\left[\left(\partial_{h}\mathrm{H}\right)_{d}\right]^{2}\right\rangle-\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle^{2}\right\}\text{.} (69)

Third, we note that

14n2βpnhpnpndβdh=142β[H(hH)dH(hH)d]dβdh.\frac{1}{4}\sum_{n}\frac{2\partial_{\beta}p_{n}\partial_{h}p_{n}}{p_{n}}d\beta dh=\frac{1}{4}2\beta\left[\left\langle\mathrm{H}\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle-\left\langle\mathrm{H}\right\rangle\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle\right]d\beta dh\text{.} (70)

Indeed, using Eqs. (59) and (60), we get

n2βpnhpnpn\displaystyle\sum_{n}\frac{2\partial_{\beta}p_{n}\partial_{h}p_{n}}{p_{n}} =n2pn[HEn]βpn[(hH)dhEn]pn\displaystyle=\sum_{n}\frac{2p_{n}\left[\left\langle\mathrm{H}\right\rangle-E_{n}\right]\beta p_{n}\left[\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle-\partial_{h}E_{n}\right]}{p_{n}}
=n2β[HEn][(hH)dhEn]pn\displaystyle=\sum_{n}2\beta\left[\left\langle\mathrm{H}\right\rangle-E_{n}\right]\left[\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle-\partial_{h}E_{n}\right]p_{n}
=n2β[H(hH)dHhEnEn(hH)d+EnhEn]pn\displaystyle=\sum_{n}2\beta\left[\left\langle\mathrm{H}\right\rangle\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle-\left\langle\mathrm{H}\right\rangle\partial_{h}E_{n}-E_{n}\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle+E_{n}\partial_{h}E_{n}\right]p_{n}
=2β[H(hH)dH(hH)dH(hH)d+H(hH)d]\displaystyle=2\beta\left[\left\langle\mathrm{H}\right\rangle\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle-\left\langle\mathrm{H}\right\rangle\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle-\left\langle\mathrm{H}\right\rangle\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle+\left\langle\mathrm{H}\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle\right]
=2β[H(hH)dH(hH)d],\displaystyle=2\beta\left[\left\langle\mathrm{H}\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle-\left\langle\mathrm{H}\right\rangle\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle\right]\text{,} (71)

where H(hH)d\left\langle\mathrm{H}\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle is defined as

H(hH)d=defipiEihEi.\left\langle\mathrm{H}\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle\overset{\text{def}}{=}\sum_{i}p_{i}E_{i}\partial_{h}E_{i}\text{.} (72)

Finally, employing Eqs. (64), (67), and (70), the most general expression of dsBures2(ρρ+dρ)ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right) in Eq. (37) between two mixed thermal states ρ(βh)\rho\left(\beta\text{, }h\right) and (ρ+dρ)(βh)\left(\rho+d\rho\right)\left(\beta\text{, }h\right) when either changes in the parameter β\beta or hh are allotted becomes

dsBures2(ρρ+dρ)\displaystyle ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right) =14[H2H2]dβ2\displaystyle=\frac{1}{4}\left[\left\langle\mathrm{H}^{2}\right\rangle-\left\langle\mathrm{H}\right\rangle^{2}\right]d\beta^{2}
+14{β2{[(hH)d]2(hH)d2}+2nm|n|hH|mEnEm|2(eβEneβEm)2𝒵(eβEn+eβEm)}dh2+\displaystyle+\frac{1}{4}\left\{\beta^{2}\left\{\left\langle\left[\left(\partial_{h}\mathrm{H}\right)_{d}\right]^{2}\right\rangle-\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle^{2}\right\}+2\sum_{n\neq m}\left|\frac{\left\langle n|\partial_{h}\mathrm{H}|m\right\rangle}{E_{n}-E_{m}}\right|^{2}\frac{\left(e^{-\beta E_{n}}-e^{-\beta E_{m}}\right)^{2}}{\mathcal{Z}\left(e^{-\beta E_{n}}+e^{-\beta E_{m}}\right)}\right\}dh^{2}+
+14{2β[H(hH)dH(hH)d]}dβdh.\displaystyle+\frac{1}{4}\left\{2\beta\left[\left\langle\mathrm{H}\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle-\left\langle\mathrm{H}\right\rangle\left\langle\left(\partial_{h}\mathrm{H}\right)_{d}\right\rangle\right]\right\}d\beta dh\text{.} (73)

Note that dsBures2(ρρ+dρ)ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right) is the sum of two contributions, the classical Fisher-Rao information metric contribution and the non-classical metric contribution expressed in the summation term in the right-hand-side of Eq. (73). For later convenience, we also remark that the quadratic term |n|hH|m|2\left|\left\langle n|\partial_{h}\mathrm{H}|m\right\rangle\right|^{2} in the summation term in the right-hand-side of Eq. (73) is invariant under change of sign of the Hamiltonian of the system. Clearly, from Eq. (73) we find that dsBures2(ρρ+dρ)=(1/4)[H2H2]dβ2ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right)=(1/4)\left[\left\langle\mathrm{H}^{2}\right\rangle-\left\langle\mathrm{H}\right\rangle^{2}\right]d\beta^{2} when h=h=const. and only β\beta can change. If β=\beta=const., dsBures2(ρρ+dρ)ds_{\mathrm{Bures}}^{2}\left(\rho\text{, }\rho+d\rho\right) in Eq. (73) reduces to Eq. (54). The explicit derivation of Eq. (73) ends our calculation of the Bures metric between neighboring thermal states undergoing temperature and/or magnetic field intensity changes as originally presented by Zanardi and collaborators in Ref. zanardi07 .

III The Sjöqvist metric

In this section, we introduce the Sjöqvist metric erik20 for nondegerante mixed states with an explicit derivation. Assume to consider two rank-NN neighboring nondegenerate density operators ρ(t)\rho\left(t\right) and ρ(t+dt)\rho\left(t+dt\right) linked by means of a smooth path tρ(t)t\mapsto\rho\left(t\right) specifying the evolution of a given quantum system. The nondegeneracy property implies that the phase of the eigenvectors represents the gauge freedom in the spectral decomposition of the density operators. As a consequence, there exists a one-to-one correspondence between the set of two orthogonal rays {eiϕk(t)|ek(t):0ϕk(t)<2π}1kN\left\{e^{i\phi_{k}\left(t\right)}\left|e_{k}\left(t\right)\right\rangle:0\leq\phi_{k}\left(t\right)<2\pi\right\}_{1\leq k\leq N} that specify the spectral decomposition along the path tρ(t)t\mapsto\rho\left(t\right) and the rank-NN nondegenerate density operator ρ(t)\rho\left(t\right). Obviously, if some nonzero eigenvalue of ρ(t)\rho\left(t\right) is degenerate, this correspondence would no longer exist. We present next the explicit derivation of the Sjöqvist metric.

III.1 The explicit derivation

Consider two neighboring states ρ(t)\rho\left(t\right) and ρ(t+dt)\rho\left(t+dt\right) with spectral decompositions given by (t)={pk(t)eifk(t)|nk(t)}1kN\mathcal{B}\left(t\right)=\left\{\sqrt{p_{k}\left(t\right)}e^{if_{k}\left(t\right)}\left|n_{k}\left(t\right)\right\rangle\right\}_{1\leq k\leq N} and (t+dt)={pk(t+dt)eifk(t+dt)|nk(t+dt)}1kN\mathcal{B}\left(t+dt\right)=\left\{\sqrt{p_{k}\left(t+dt\right)}e^{if_{k}\left(t+dt\right)}\left|n_{k}\left(t+dt\right)\right\rangle\right\}_{1\leq k\leq N}, respectively. The quantity NN denotes the rank of the nondegenerate density operator ρ(t)\rho\left(t\right). Consider the infinitesimal distance d2(tt+dt)d^{2}\left(t\text{, }t+dt\right) between ρ(t)\rho\left(t\right) and ρ(t+dt)\rho\left(t+dt\right) defined as

d2(tt+dt)=defk=1Npk(t)eifk(t)|nk(t)pk(t+dt)eifk(t+dt)|nk(t+dt)2.d^{2}\left(t\text{, }t+dt\right)\overset{\text{def}}{=}\sum_{k=1}^{N}\left\|\sqrt{p_{k}\left(t\right)}e^{if_{k}\left(t\right)}\left|n_{k}\left(t\right)\right\rangle-\sqrt{p_{k}\left(t+dt\right)}e^{if_{k}\left(t+dt\right)}\left|n_{k}\left(t+dt\right)\right\rangle\right\|^{2}\text{.} (74)

The Sjöqvist metric is defined as the minimum of d2(tt+dt)d^{2}\left(t\text{, }t+dt\right) in Eq. (74). Note that the squared norm term pk(t)eifk(t)|nk(t)pk(t+dt)eifk(t+dt)|nk(t+dt)2\left\|\sqrt{p_{k}\left(t\right)}e^{if_{k}\left(t\right)}\left|n_{k}\left(t\right)\right\rangle-\sqrt{p_{k}\left(t+dt\right)}e^{if_{k}\left(t+dt\right)}\left|n_{k}\left(t+dt\right)\right\rangle\right\|^{2} can be written as

(pk(t)eifk(t)nk(t)|pk(t+dt)eifk(t+dt)nk(t+dt)|)\displaystyle\left(\sqrt{p_{k}\left(t\right)}e^{-if_{k}\left(t\right)}\left\langle n_{k}\left(t\right)\right|-\sqrt{p_{k}\left(t+dt\right)}e^{-if_{k}\left(t+dt\right)}\left\langle n_{k}\left(t+dt\right)\right|\right)
(pk(t)eifk(t)|nk(t)pk(t+dt)eifk(t+dt)|nk(t+dt))\displaystyle\left(\sqrt{p_{k}\left(t\right)}e^{if_{k}\left(t\right)}\left|n_{k}\left(t\right)\right\rangle-\sqrt{p_{k}\left(t+dt\right)}e^{if_{k}\left(t+dt\right)}\left|n_{k}\left(t+dt\right)\right\rangle\right)
=pk(t)+pk(t+dt)pk(t)pk(t+dt)ei[fk(t+dt)fk(t)]nk(t)|nk(t+dt)+\displaystyle=p_{k}\left(t\right)+p_{k}\left(t+dt\right)-\sqrt{p_{k}\left(t\right)p_{k}\left(t+dt\right)}e^{i\left[f_{k}\left(t+dt\right)-f_{k}\left(t\right)\right]}\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle+
pk(t)pk(t+dt)ei[fk(t+dt)fk(t)]nk(t+dt)|nk(t),\displaystyle-\sqrt{p_{k}\left(t\right)p_{k}\left(t+dt\right)}e^{-i\left[f_{k}\left(t+dt\right)-f_{k}\left(t\right)\right]}\left\langle n_{k}\left(t+dt\right)|n_{k}\left(t\right)\right\rangle\text{,} (75)

with ei[fk(t+dt)fk(t)]nk(t)|nk(t+dt)+ei[fk(t+dt)fk(t)]nk(t+dt)|nk(t)e^{i\left[f_{k}\left(t+dt\right)-f_{k}\left(t\right)\right]}\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle+e^{-i\left[f_{k}\left(t+dt\right)-f_{k}\left(t\right)\right]}\left\langle n_{k}\left(t+dt\right)|n_{k}\left(t\right)\right\rangle equal to

2Re{ei[fk(t+dt)fk(t)]nk(t)|nk(t+dt)}\displaystyle 2\operatorname{Re}\left\{e^{i\left[f_{k}\left(t+dt\right)-f_{k}\left(t\right)\right]}\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle\right\}
=2Re{ei[fk(t+dt)fk(t)]|nk(t)|nk(t+dt)|eiarg(nk(t)|nk(t+dt))}\displaystyle=2\operatorname{Re}\left\{e^{i\left[f_{k}\left(t+dt\right)-f_{k}\left(t\right)\right]}\left|\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle\right|e^{i\arg\left(\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle\right)}\right\}
=2|nk(t)|nk(t+dt)|Re{ei[fk(t+dt)fk(t)]eiarg(nk(t)|nk(t+dt))}\displaystyle=2\left|\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle\right|\operatorname{Re}\left\{e^{i\left[f_{k}\left(t+dt\right)-f_{k}\left(t\right)\right]}e^{i\arg\left(\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle\right)}\right\}
=2|nk(t)|nk(t+dt)|cos[fk(t+dt)fk(t)+arg(nk(t)|nk(t+dt))].\displaystyle=2\left|\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle\right|\cos\left[f_{k}\left(t+dt\right)-f_{k}\left(t\right)+\arg\left(\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle\right)\right]\text{.} (76)

Note that fk(t+dt)=fk(t)+f˙k(t)dt+O(dt2)f_{k}\left(t+dt\right)=f_{k}\left(t\right)+\dot{f}_{k}\left(t\right)dt+O\left(dt^{2}\right) and |nk(t+dt)=|nk(t)+|n˙k(t)dt+O(dt2)\left|n_{k}\left(t+dt\right)\right\rangle=\left|n_{k}\left(t\right)\right\rangle+\left|\dot{n}_{k}\left(t\right)\right\rangle dt+O\left(dt^{2}\right). Therefore, we have

cos[fk(t+dt)fk(t)+arg(nk(t)|nk(t+dt))]=cos{f˙k(t)dt+arg[1+nk(t)|n˙k(t)dt]+O(dt2)}.\cos\left[f_{k}\left(t+dt\right)-f_{k}\left(t\right)+\arg\left(\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle\right)\right]=\cos\left\{\dot{f}_{k}\left(t\right)dt+\arg\left[1+\left\langle n_{k}\left(t\right)|\dot{n}_{k}\left(t\right)\right\rangle dt\right]+O\left(dt^{2}\right)\right\}\text{.} (77)

Setting f˙k(t)dt+arg[1+nk(t)|n˙k(t)dt]+O(dt2)=defλk(tt+dt)\dot{f}_{k}\left(t\right)dt+\arg\left[1+\left\langle n_{k}\left(t\right)|\dot{n}_{k}\left(t\right)\right\rangle dt\right]+O\left(dt^{2}\right)\overset{\text{def}}{=}\lambda_{k}\left(t\text{, }t+dt\right), the infinitesimal distance d2(tt+dt)d^{2}\left(t\text{, }t+dt\right) becomes

d2(tt+dt)=22k=1Npk(t)pk(t+dt)|nk(t)|nk(t+dt)|cos[λk(tt+dt)].d^{2}\left(t\text{, }t+dt\right)=2-2\sum_{k=1}^{N}\sqrt{p_{k}\left(t\right)p_{k}\left(t+dt\right)}\left|\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle\right|\cos\left[\lambda_{k}\left(t\text{, }t+dt\right)\right]\text{.} (78)

Then, the Sjöqvist metric dsSjo¨qvist2ds_{\mathrm{Sj\ddot{o}qvist}}^{2} is the minimum of d2(tt+dt)d^{2}\left(t\text{, }t+dt\right), dmin2(tt+dt)d_{\min}^{2}\left(t\text{, }t+dt\right), and is obtained when λk(tt+dt)\lambda_{k}\left(t\text{, }t+dt\right) equals zero for any 1kN1\leq k\leq N. Its expression is given by,

dsSjo¨qvist2=22k=1Npk(t)pk(t+dt)|nk(t)|nk(t+dt)|.ds_{\mathrm{Sj\ddot{o}qvist}}^{2}=2-2\sum_{k=1}^{N}\sqrt{p_{k}\left(t\right)p_{k}\left(t+dt\right)}\left|\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle\right|\text{.} (79)

It is worthwhile emphasizing that the minimum of d2(tt+dt)d^{2}\left(t\text{, }t+dt\right) is achieved by selecting phases {fk(t)fk(t+dt)}\left\{f_{k}\left(t\right)\text{, }f_{k}\left(t+dt\right)\right\} such that

f˙k(t)dt+arg[1+nk(t)|n˙k(t)dt+O(dt2)]=0\dot{f}_{k}\left(t\right)dt+\arg\left[1+\left\langle n_{k}\left(t\right)\left|\dot{n}_{k}\left(t\right)\right.\right\rangle dt+O\left(dt^{2}\right)\right]=0 (80)

Observing that enk(t)|n˙k(t)dt=1+nk(t)|n˙k(t)dt+O(dt2)e^{\left\langle n_{k}\left(t\right)\left|\dot{n}_{k}\left(t\right)\right.\right\rangle dt}=1+\left\langle n_{k}\left(t\right)\left|\dot{n}_{k}\left(t\right)\right.\right\rangle dt+O\left(dt^{2}\right) is such that arg[enk(t)|n˙k(t)dt]=ink(t)|n˙k(t)dt\arg\left[e^{\left\langle n_{k}\left(t\right)\left|\dot{n}_{k}\left(t\right)\right.\right\rangle dt}\right]=-i\left\langle n_{k}\left(t\right)\left|\dot{n}_{k}\left(t\right)\right.\right\rangle dt, Eq. (80) can be rewritten to the first order in dtdt as

f˙k(t)ink(t)|n˙k(t)=0.\dot{f}_{k}\left(t\right)-i\left\langle n_{k}\left(t\right)\left|\dot{n}_{k}\left(t\right)\right.\right\rangle=0\text{.} (81)

Eq. (81) denotes the parallel transport condition ψk(t)|ψk(t)=0\left\langle\psi_{k}\left(t\right)\left|\psi_{k}\left(t\right)\right.\right\rangle=0 where |ψk(t)=defeifk(t)|nk(t)\left|\psi_{k}\left(t\right)\right\rangle\overset{\text{def}}{=}e^{if_{k}\left(t\right)}\left|n_{k}\left(t\right)\right\rangle is associated with individual pure state paths in the chosen ensemble that defines the mixed state ρ(t)\rho\left(t\right) aharonov87 . To find a more useful expression of dsSjo¨qvist2ds_{\mathrm{Sj\ddot{o}qvist}}^{2}, let us start by observing that,

pk(t)pk(t+dt)=pk(t)1+dpk(t)pk(t)=pk+12p˙kdt18p˙k2pkdt2+O(dt2).\sqrt{p_{k}\left(t\right)p_{k}\left(t+dt\right)}=p_{k}\left(t\right)\sqrt{1+\frac{dp_{k}\left(t\right)}{p_{k}\left(t\right)}}=p_{k}+\frac{1}{2}\dot{p}_{k}dt-\frac{1}{8}\frac{\dot{p}_{k}^{2}}{p_{k}}dt^{2}+O\left(dt^{2}\right)\text{.} (82)

Furthermore, to the second order in dtdt, the state |nk(t+dt)\left|n_{k}\left(t+dt\right)\right\rangle can be written as

|nk(t+dt)=|nk(t)+|n˙k(t)dt+12|n¨k(t)dt2+O(dt2).\left|n_{k}\left(t+dt\right)\right\rangle=\left|n_{k}\left(t\right)\right\rangle+\left|\dot{n}_{k}\left(t\right)\right\rangle dt+\frac{1}{2}\left|\ddot{n}_{k}\left(t\right)\right\rangle dt^{2}+O\left(dt^{2}\right)\text{.} (83)

Therefore, to the second order in dtdt, the quantum overlap nk(t)|nk(t+dt)\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle becomes

nk(t)|nk(t+dt)=nk(t)|nk(t)+nk(t)|n˙k(t)dt+12nk(t)|n¨k(t)dt2+O(dt2)\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle=\left\langle n_{k}\left(t\right)|n_{k}\left(t\right)\right\rangle+\left\langle n_{k}\left(t\right)|\dot{n}_{k}\left(t\right)\right\rangle dt+\frac{1}{2}\left\langle n_{k}\left(t\right)|\ddot{n}_{k}\left(t\right)\right\rangle dt^{2}+O\left(dt^{2}\right) (84)

Let us focus now on calculating |nk(t)|nk(t+dt)|=|nk(t)|nk(t+dt)|2\left|\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle\right|=\sqrt{\left|\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle\right|^{2}}, where

|nk(t)|nk(t+dt)|2=nk(t)|nk(t+dt)nk(t+dt)|nk(t).\left|\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle\right|^{2}=\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle\left\langle n_{k}\left(t+dt\right)|n_{k}\left(t\right)\right\rangle\text{.} (85)

Using Eq. (83), Eq. (85) becomes

nk(t)|nk(t+dt)nk(t+dt)|nk(t)\displaystyle\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle\left\langle n_{k}\left(t+dt\right)|n_{k}\left(t\right)\right\rangle [nk(t)|nk(t)+nk(t)|n˙k(t)dt+12nk(t)|n¨k(t)dt2]\displaystyle\approx\left[\left\langle n_{k}\left(t\right)|n_{k}\left(t\right)\right\rangle+\left\langle n_{k}\left(t\right)|\dot{n}_{k}\left(t\right)\right\rangle dt+\frac{1}{2}\left\langle n_{k}\left(t\right)|\ddot{n}_{k}\left(t\right)\right\rangle dt^{2}\right]\cdot
[nk(t)|nk(t)+n˙k(t)|nk(t)dt+12n¨k(t)|nk(t)dt2]\displaystyle\cdot\left[\left\langle n_{k}\left(t\right)|n_{k}\left(t\right)\right\rangle+\left\langle\dot{n}_{k}\left(t\right)|n_{k}\left(t\right)\right\rangle dt+\frac{1}{2}\left\langle\ddot{n}_{k}\left(t\right)|n_{k}\left(t\right)\right\rangle dt^{2}\right]
=[1+nk|n˙kdt+12nk|n¨kdt2][1+n˙k|nkdt+12n¨k|nkdt2]\displaystyle=\left[1+\left\langle n_{k}|\dot{n}_{k}\right\rangle dt+\frac{1}{2}\left\langle n_{k}|\ddot{n}_{k}\right\rangle dt^{2}\right]\left[1+\left\langle\dot{n}_{k}|n_{k}\right\rangle dt+\frac{1}{2}\left\langle\ddot{n}_{k}|n_{k}\right\rangle dt^{2}\right]
1+n˙k|nkdt+12n¨k|nkdt2+nk|n˙kdt+nk|n˙kn˙k|nkdt2+12nk|n¨kdt2\displaystyle\approx 1+\left\langle\dot{n}_{k}|n_{k}\right\rangle dt+\frac{1}{2}\left\langle\ddot{n}_{k}|n_{k}\right\rangle dt^{2}+\left\langle n_{k}|\dot{n}_{k}\right\rangle dt+\left\langle n_{k}|\dot{n}_{k}\right\rangle\left\langle\dot{n}_{k}|n_{k}\right\rangle dt^{2}+\frac{1}{2}\left\langle n_{k}|\ddot{n}_{k}\right\rangle dt^{2}
=1+[n˙k|nk+nk|n˙k]dt+nk|n˙kn˙k|nkdt2+12[nk|n¨k+n¨k|nk]dt2\displaystyle=1+\left[\left\langle\dot{n}_{k}|n_{k}\right\rangle+\left\langle n_{k}|\dot{n}_{k}\right\rangle\right]dt+\left\langle n_{k}|\dot{n}_{k}\right\rangle\left\langle\dot{n}_{k}|n_{k}\right\rangle dt^{2}+\frac{1}{2}\left[\left\langle n_{k}|\ddot{n}_{k}\right\rangle+\left\langle\ddot{n}_{k}|n_{k}\right\rangle\right]dt^{2}
=1+nk|n˙kn˙k|nkdt2n˙k|n˙kdt2,\displaystyle=1+\left\langle n_{k}|\dot{n}_{k}\right\rangle\left\langle\dot{n}_{k}|n_{k}\right\rangle dt^{2}-\left\langle\dot{n}_{k}|\dot{n}_{k}\right\rangle dt^{2}\text{,} (86)

that is,

nk(t)|nk(t+dt)nk(t+dt)|nk(t)=1+nk|n˙kn˙k|nkdt2n˙k|n˙kdt2+O(dt2),\left\langle n_{k}\left(t\right)|n_{k}\left(t+dt\right)\right\rangle\left\langle n_{k}\left(t+dt\right)|n_{k}\left(t\right)\right\rangle=1+\left\langle n_{k}|\dot{n}_{k}\right\rangle\left\langle\dot{n}_{k}|n_{k}\right\rangle dt^{2}-\left\langle\dot{n}_{k}|\dot{n}_{k}\right\rangle dt^{2}+O\left(dt^{2}\right)\text{,} (87)

since nk|nk=1\left\langle n_{k}|n_{k}\right\rangle=1 implies n˙k|nk+nk|n˙k=0\left\langle\dot{n}_{k}|n_{k}\right\rangle+\left\langle n_{k}|\dot{n}_{k}\right\rangle=0 and nk|n¨k+n¨k|nk=2n˙k|n˙k\left\langle n_{k}|\ddot{n}_{k}\right\rangle+\left\langle\ddot{n}_{k}|n_{k}\right\rangle=-2\left\langle\dot{n}_{k}|\dot{n}_{k}\right\rangle. Finally, using Eqs. (82) and (87) along with noting that kp˙k=0\sum_{k}\dot{p}_{k}=0, the Sjöqvist metric dsSjo¨qvist2ds_{\mathrm{Sj\ddot{o}qvist}}^{2} in Eq. (79) becomes

dsSjo¨qvist2\displaystyle ds_{\mathrm{Sj\ddot{o}qvist}}^{2} 22k=1N(pk+12p˙kdt18p˙k2pkdt2)(1+12nk|n˙kn˙k|nkdt212n˙k|n˙kdt2)\displaystyle\approx 2-2\sum_{k=1}^{N}\left(p_{k}+\frac{1}{2}\dot{p}_{k}dt-\frac{1}{8}\frac{\dot{p}_{k}^{2}}{p_{k}}dt^{2}\right)\left(1+\frac{1}{2}\left\langle n_{k}|\dot{n}_{k}\right\rangle\left\langle\dot{n}_{k}|n_{k}\right\rangle dt^{2}-\frac{1}{2}\left\langle\dot{n}_{k}|\dot{n}_{k}\right\rangle dt^{2}\right)
22k=1Npkk=1Npknk|n˙kn˙k|nkdt2+k=1Npkn˙k|n˙kdt2+14k=1Np˙k2pkdt2,\displaystyle\approx 2-2\sum_{k=1}^{N}p_{k}-\sum_{k=1}^{N}p_{k}\left\langle n_{k}|\dot{n}_{k}\right\rangle\left\langle\dot{n}_{k}|n_{k}\right\rangle dt^{2}+\sum_{k=1}^{N}p_{k}\left\langle\dot{n}_{k}|\dot{n}_{k}\right\rangle dt^{2}+\frac{1}{4}\sum_{k=1}^{N}\frac{\dot{p}_{k}^{2}}{p_{k}}dt^{2}\text{,} (88)

that is,

dsSjo¨qvist2\displaystyle ds_{\mathrm{Sj\ddot{o}qvist}}^{2} 14k=1Np˙k2pkdt2+k=1Npk[n˙k|n˙knk|n˙kn˙k|nk]dt2\displaystyle\approx\frac{1}{4}\sum_{k=1}^{N}\frac{\dot{p}_{k}^{2}}{p_{k}}dt^{2}+\sum_{k=1}^{N}p_{k}\left[\left\langle\dot{n}_{k}|\dot{n}_{k}\right\rangle-\left\langle n_{k}|\dot{n}_{k}\right\rangle\left\langle\dot{n}_{k}|n_{k}\right\rangle\right]dt^{2}
14k=1Np˙k2pkdt2+k=1Npk[n˙k|(I|nknk|)|n˙k]dt2,\displaystyle\approx\frac{1}{4}\sum_{k=1}^{N}\frac{\dot{p}_{k}^{2}}{p_{k}}dt^{2}+\sum_{k=1}^{N}p_{k}\left[\left\langle\dot{n}_{k}|\left(\mathrm{I}-\left|n_{k}\right\rangle\left\langle n_{k}\right|\right)|\dot{n}_{k}\right\rangle\right]dt^{2}\text{,} (89)

where I in Eq. (89) denotes the identity operator on the NN-dimensional Hilbert space. Finally, neglecting terms that are smaller than O(dt2)O\left(dt^{2}\right) in Eq. (79) and defining dsk2=def[n˙k|(I|nknk|)|n˙k]dt2ds_{k}^{2}\overset{\text{def}}{=}\left[\left\langle\dot{n}_{k}|\left(\mathrm{I}-\left|n_{k}\right\rangle\left\langle n_{k}\right|\right)|\dot{n}_{k}\right\rangle\right]dt^{2}, the expression of the Sjöqvist metric will be formally taken to be

dsSjo¨qvist2=def14k=1Np˙k2pkdt2+k=1Npkdsk2.ds_{\mathrm{Sj\ddot{o}qvist}}^{2}\overset{\text{def}}{=}\frac{1}{4}\sum_{k=1}^{N}\frac{\dot{p}_{k}^{2}}{p_{k}}dt^{2}+\sum_{k=1}^{N}p_{k}ds_{k}^{2}\text{.} (90)

The derivation of Eq. (90) concludes our explicit calculation of the Sjöqvist metric for nondegerante mixed states. Interestingly, note that dsk2=defn˙k|(I|nknk|)|n˙kdt2ds_{k}^{2}\overset{\text{def}}{=}\left\langle\dot{n}_{k}\left|\left(\mathrm{I}-\left|n_{k}\right\rangle\left\langle n_{k}\right|\right)\right|\dot{n}_{k}\right\rangle dt^{2} in Eq. (90) can be written as dsk2=nk|nkds_{k}^{2}=\left\langle\nabla n_{k}\left|\nabla n_{k}\right.\right\rangle with |nk=defP(k)|n˙k\left|\nabla n_{k}\right\rangle\overset{\text{def}}{=}\mathrm{P}_{\bot}^{\left(k\right)}\left|\dot{n}_{k}\right\rangle being the covariant derivative of |nk\left|n_{k}\right\rangle and P(k)=defI|nknk|\mathrm{P}_{\bot}^{\left(k\right)}\overset{\text{def}}{=}\mathrm{I}-\left|n_{k}\right\rangle\left\langle n_{k}\right| denoting the projector onto states perpendicular to |nk\left|n_{k}\right\rangle. In analogy to the Bures metric case (see the comment right below Eq. (73)), we stress for later convenience that the quadratic term dsk2ds_{k}^{2} does not change under change of sign of the Hamiltonian of the system. The expression of the Sjöqvist metric in Eq. (90) can be viewed as expressed by two contributions, a classical and a nonclassical term. The first term in Eq. (90) is the classical one and is represented by the classical Fisher-Rao information metric between the two probability distributions {pk}1kN\left\{p_{k}\right\}_{1\leq k\leq N} and {pk+dpk}1kN\left\{p_{k}+dp_{k}\right\}_{1\leq k\leq N}. The second term is the nonclassical one and is represented by a weighted average of pure state Fubini-Study metrics along directions specified by state vectors {|nk}1kN\left\{\left|n_{k}\right\rangle\right\}_{1\leq k\leq N}. We are now ready to introduce our Hamiltonian models.

IV The Hamiltonian Models

In this section, we present two Hamiltonian models. The first Hamiltonian model specifies a spin-1/21/2 particle in a uniform and time-independent external magnetic field oriented along the zz-axis. The second Hamiltonian model, instead, describes a superconducting flux qubit. Finally, we construct the two corresponding parametric families of thermal states by bringing these two systems in thermal equilibrium with a reservoir at finite and non-zero temperature TT.

Refer to caption
Figure 1: Schematic depiction of a spin qubit (a) and a superconducting flux qubit (b) in thermal equilibrium with a reservoir at non-zero temperature T. The spin qubit in (a) has opposite orientations of the spin along the quantization axis as its two states. The superconducting flux qubit in (b), instead, has circulating currents of opposite sign as its two states.

IV.1 Spin-1/2 qubit Hamiltonian

Consider a spin-1/21/2 particle represented by an electron of mm, charge e-e with e0e\geq 0 immersed in an external magnetic field B(t)\vec{B}\left(t\right). From a quantum-mechanical perspective, the Hamiltonian of this system can be described the Hermitian operator H(t)\left(t\right) given by H(t)=defμB(t)\mathrm{H}\left(t\right)\overset{\text{def}}{=}-\vec{\mu}\mathbf{\cdot}\vec{B}\left(t\right) sakurai , with μ\vec{\mu} denoting the electron magnetic moment operator. The quantity μ\vec{\mu} is defined as μ=def(e/m)s\vec{\mu}\overset{\text{def}}{=}-\left(e/m\right)\vec{s} with s=def(/2)σ\vec{s}\overset{\text{def}}{=}\left(\hslash/2\right)\vec{\sigma} being the spin operator. Clearly, =defh/(2π)\hslash\overset{\text{def}}{=}h/(2\pi) is the reduced Planck constant and σ=def(σxσyσz)\vec{\sigma}\overset{\text{def}}{=}\left(\sigma_{x}\text{, }\sigma_{y}\text{, }\sigma_{z}\right) is the usual Pauli spin vector operator. Assuming a time-independent magnetic field along the zz-direction given by B(t)=B0z^\vec{B}\left(t\right)=B_{0}\hat{z} and introducing the frequency ω=def(e/m)B0\omega\overset{\text{def}}{=}(e/m)B_{0}, the spin-1/21/2 qubit (SQ) Hamiltonian becomes

HSQ(ω)=defω2σz,\mathrm{H}_{\mathrm{SQ}}\left(\omega\right)\overset{\text{def}}{=}\frac{\hslash\omega}{2}\sigma_{z}\text{,} (91)

where σz=def||||\sigma_{z}\overset{\text{def}}{=}\left|\uparrow\right\rangle\left\langle\uparrow\right|-\left|\downarrow\right\rangle\left\langle\downarrow\right| with |\left|\uparrow\right\rangle and |\left|\downarrow\right\rangle denoting the spin-up and the spin-down quantum states, respectively. Observe that with the sign convention used for HSQ(ω)\mathrm{H}_{\mathrm{SQ}}\left(\omega\right) in Eq. (91), we have that |\left|\downarrow\right\rangle (|\left|\uparrow\right\rangle) denotes the ground (excited) state of the system with energy ω/2-\hslash\omega/2 (+ω/2+\hslash\omega/2).

IV.2 Superconducting flux qubit Hamiltonian

It is known that a qubit is a two-level (or, a two-state) quantum system and, moreover, it is possible to realize the two levels in a number of ways. For example, the two-levels can be regarded as the spin-up and spin-down of an electron, or as the vertical and horizontal polarization of a single photon. Interestingly, the two-levels of a qubit can be also realized as the supercurrent flowing in an anti-clockwise and clockwise directions in a superconducting loop clarke08 ; devoret13 . A flux qubit is a superconducting loop interrupted by one or three Josephson junctions (i.e., a dissipationless device with a nonlinear inductance). An arbitrary flux qubit can be described as a superposition of two persistent current basis states. The two quantum states are total magnetic flux Φ\Phi pointing up |\left|\uparrow\right\rangle and Φ\Phi pointing down |\left\langle\downarrow\right|. Alternatively, as previously mentioned, the two-levels of the quantum system can be described as the supercurrent IqI_{q} circulating in the loop anti-clockwise and IqI_{q} circulating clockwise. The Hamiltonian of a superconducting flux qubit (SFQ) in persistent current basis {||}\left\{\left|\uparrow\right\rangle\text{, }\left\langle\downarrow\right|\right\} is given by chiorescu03 ; pekola07 ; paauw09 ; pekola16 ,

HSFQ(Δϵ)=def2(Δσx+ϵσz).\mathrm{H}_{\mathrm{SFQ}}\left(\Delta\text{, }\epsilon\right)\overset{\text{def}}{=}-\frac{\hslash}{2}\left(\Delta\sigma_{x}+\epsilon\sigma_{z}\right)\text{.} (92)

In Eq. (92), =defh/(2π)\hslash\overset{\text{def}}{=}h/\left(2\pi\right) is the reduced Planck constant, while σx\sigma_{x} and σz\sigma_{z} are Pauli matrices. Furthermore, ϵ=def2Iq(ΦeΦ02)\hslash\epsilon\overset{\text{def}}{=}2I_{q}\left(\Phi_{e}-\frac{\Phi_{0}}{2}\right) is the magnetic energy bias defined in terms of the supercurrent IqI_{q}, the externally applied magnetic flux Φe\Phi_{e}, and the magnetic flux quantum Φ0=defh/(2e)\Phi_{0}\overset{\text{def}}{=}h/\left(2e\right) with ee being the absolute value of the electron charge. Finally, Δ\hslash\Delta is the energy gap at the degeneracy point specified by the relation Φe=Φ0/2\Phi_{e}=\Phi_{0}/2 (i.e., ϵ=0\epsilon=0) and represents the minimum splitting of the energy levels of the ground state |g\left|g\right\rangle and the first excited state |e\left|e\right\rangle of the superconducting qubit. At the gap, the coherence properties of the qubit are optimal. Away from the degeneracy point, ϵ0\epsilon\neq 0 and the energy-level splitting becomes ν=defϵ2+Δ2\hslash\nu\overset{\text{def}}{=}\hslash\sqrt{\epsilon^{2}+\Delta^{2}}, with ν\nu being the transition angular frequency of the qubit. The energy level splitting Δ\hslash\Delta depends on the critical current of the three Josephson junctions and their capacitance paauw09 . For flux qubits one has ΔEC/EJ\Delta\sim E_{C}/E_{J} with ECE_{C} and EJE_{J} denoting the Cooper pair charging energy and the Josephson coupling energy pekola16 , respectively. In summary, a flux qubit can be represented by a double-well potential whose shape (symmetrical versus asymmetrical) can be tuned with the externally applied magnetic flux Φe\Phi_{e}. When Φe=Φ0/2\Phi_{e}=\Phi_{0}/2, the double-well is symmetric, the energy eigenstates (i.e., ground state and first excited states |g\left|g\right\rangle and |e\left|e\right\rangle, respectively) are symmetric (i.e., |g=def[|+|]/2\left|g\right\rangle\overset{\text{def}}{=}\left[\left|\uparrow\right\rangle+\left|\downarrow\right\rangle\right]/\sqrt{2}) and antisymmetric (i.e., |e=def[||]/2\left|e\right\rangle\overset{\text{def}}{=}\left[\left|\uparrow\right\rangle-\left|\downarrow\right\rangle\right]/\sqrt{2}) superpositions of the two states |\left|\uparrow\right\rangle and |\left|\downarrow\right\rangle and, finally, the splitting of the energy levels of |g\left|g\right\rangle and |e\left|e\right\rangle is Δ\Delta. Instead, when ΦeΦ0/2\Phi_{e}\neq\Phi_{0}/2, the double-well is not symmetric, the energy eigenstates are arbitrary superpositions of the basis states |\left|\uparrow\right\rangle and |\left|\downarrow\right\rangle (i.e., α|±β|\alpha\left|\uparrow\right\rangle\pm\beta\left|\downarrow\right\rangle with |α|2+|β|2=1\left|\alpha\right|^{2}+\left|\beta\right|^{2}=1) and, finally, the energy gap becomes ν=defϵ2+Δ2\hslash\nu\overset{\text{def}}{=}\hslash\sqrt{\epsilon^{2}+\Delta^{2}}. For more details on the theory underlying superconducting flux qubits, we refer to Ref. clarke08 .

The transition from (isolated) physical systems specified by pure states evolving according to the Hamiltonians in Eqs. (91) and (92) to the same (open) physical systems described by mixed quantum states can be explained as follows. Assume a quantum system specified by an Hamiltonian H\mathrm{H} is in thermal equilibrium with a reservoir at non-zero temperature TT. Then, following the principles of quantum statistical mechanics huang87 , the system has temperature TT and its state is described by a thermal state strocchi08 specified by a density matrix ρ\rho given by,

ρ=defeβHtr(eβH).\rho\overset{\text{def}}{=}\frac{e^{-\beta\mathrm{H}}}{\mathrm{tr}\left(e^{-\beta\mathrm{H}}\right)}\text{.} (93)

In Eq. (93), β=def(kBT)1\beta\overset{\text{def}}{=}\left(k_{B}T\right)^{-1} denotes the so-called inverse temperature, while kBk_{B} is the Boltzmann constant. In what follows, we shall consider two families of mixed quantum thermal states given by

ρSQ(βω)=defeβHSQ(ω)tr(eβHSQ(ω)) and, ρSFQ(βϵ)=defeβHSFQ(ϵ)tr(eβHSFQ(ϵ)).\rho_{\mathrm{SQ}}\left(\beta\text{, }\omega\right)\overset{\text{def}}{=}\frac{e^{-\beta\mathrm{H}_{\mathrm{SQ}}\left(\omega\right)}}{\mathrm{tr}\left(e^{-\beta\mathrm{H}_{\mathrm{SQ}}\left(\omega\right)}\right)}\text{ and, }\rho_{\mathrm{SFQ}}\left(\beta\text{, }\epsilon\right)\overset{\text{def}}{=}\frac{e^{-\beta\mathrm{H}_{\mathrm{SFQ}}\left(\epsilon\right)}}{\mathrm{tr}\left(e^{-\beta\mathrm{H}_{\mathrm{SFQ}}\left(\epsilon\right)}\right)}\text{.} (94)

Note that in ρSFQ(βϵ)\rho_{\mathrm{SFQ}}\left(\beta\text{, }\epsilon\right) in Eq. (94), we assume that the parameter Δ\Delta is fixed. For a work on how to tune the energy gap Δ\Delta in a flux qubit from an experimental standpoint, we refer to Ref. paauw09 . In Fig. 11, we present a schematic depiction of of a spin qubit and a superconducting flux qubit in thermal equilibrium with a reservoir at non-zero temperature TT.

V Applications

In this section, we calculate both the Sjöqvist and the Bures metrics for each one of the two distinct families of parametric thermal states mentioned in the previous section. From our comparative investigation, we find that the two metric coincide for the first Hamiltonian model (electron in a constant magnetic field along the zz-direction), while they differ for the second Hamiltonian model (superconducting flux qubit).

V.1 Spin qubits

Let us consider a system with an Hamiltonian described by HSQ(ω)=def(ω/2)σz\mathrm{H}_{\mathrm{SQ}}\left(\omega\right)\overset{\text{def}}{=}\left(\hslash\omega/2\right)\sigma_{z} in Eq. (91). Observe that HSQ(ω)\mathrm{H}_{\mathrm{SQ}}\left(\omega\right) can be recast as

HSQ(ω)=n=01En|nn|=ω2|00|ω2|11|,\mathrm{H}_{\mathrm{SQ}}\left(\omega\right)=\sum_{n=0}^{1}E_{n}\left|n\right\rangle\left\langle n\right|=\frac{\hslash\omega}{2}\left|0\right\rangle\left\langle 0\right|-\frac{\hslash\omega}{2}\left|1\right\rangle\left\langle 1\right|\text{,} (95)

where E0=defω/2E_{0}\overset{\text{def}}{=}\hslash\omega/2, E1=defω/2E_{1}\overset{\text{def}}{=}-\hslash\omega/2, and {|n}=def{|0=||1=|}\left\{\left|n\right\rangle\right\}\overset{\text{def}}{=}\left\{\left|0\right\rangle=\left|\uparrow\right\rangle\text{, }\left|1\right\rangle=\left|\downarrow\right\rangle\right\}. For clarity, note that |1=|\left|1\right\rangle=\left|\downarrow\right\rangle (|0=|\left|0\right\rangle=\left|\uparrow\right\rangle) denotes here the ground (excited) state corresponding to the lowest (highest) energy level with E1=defω/2E_{1}\overset{\text{def}}{=}-\hslash\omega/2 (E0=defω/2E_{0}\overset{\text{def}}{=}\hslash\omega/2). Observe that the thermal state ρSQ\rho_{\mathrm{SQ}} emerging from the Hamiltonian HSQ\mathrm{H}_{\mathrm{SQ}} in Eq. (95) can be written as

ρSQ=ρSQ(βω)=defeβHSQ(ω)tr(eβHSQ(ω)).\rho_{\mathrm{SQ}}=\rho_{\mathrm{SQ}}\left(\beta\text{, }\omega\right)\overset{\text{def}}{=}\frac{e^{-\beta\mathrm{H}_{\mathrm{SQ}}\left(\omega\right)}}{\mathrm{tr}\left(e^{-\beta\mathrm{H}_{\mathrm{SQ}}\left(\omega\right)}\right)}\text{.} (96)

The thermal state ρSQ(βω)\rho_{\mathrm{SQ}}\left(\beta\text{, }\omega\right) in Eq. (96) can be rewritten as,

ρSQ(βω)\displaystyle\rho_{\mathrm{SQ}}\left(\beta\text{, }\omega\right) =eβω2σztr(eβω2σz)\displaystyle=\frac{e^{-\beta\frac{\hslash\omega}{2}\sigma_{z}}}{\mathrm{tr}\left(e^{-\beta\frac{\hslash\omega}{2}\sigma_{z}}\right)}
=(eβω200eβω2)eβω2+eβω2\displaystyle=\frac{\left(\begin{array}[c]{cc}e^{-\beta\frac{\hslash\omega}{2}}&0\\ 0&e^{\beta\frac{\hslash\omega}{2}}\end{array}\right)}{e^{-\beta\frac{\hslash\omega}{2}}+e^{\beta\frac{\hslash\omega}{2}}} (99)
=(cosh(βω2)sinh(βω2)00cosh(βω2)+sinh(βω2))2cosh(βω2)\displaystyle=\frac{\left(\begin{array}[c]{cc}\cosh\left(\beta\frac{\hslash\omega}{2}\right)-\sinh\left(\beta\frac{\hslash\omega}{2}\right)&0\\ 0&\cosh\left(\beta\frac{\hslash\omega}{2}\right)+\sinh\left(\beta\frac{\hslash\omega}{2}\right)\end{array}\right)}{2\cosh\left(\beta\frac{\hslash\omega}{2}\right)} (102)
=12(1tanh(βω2)001+tanh(βω2))\displaystyle=\frac{1}{2}\left(\begin{array}[c]{cc}1-\tanh\left(\beta\frac{\hslash\omega}{2}\right)&0\\ 0&1+\tanh\left(\beta\frac{\hslash\omega}{2}\right)\end{array}\right) (105)
=12[Itanh(βω2)σz],\displaystyle=\frac{1}{2}\left[\mathrm{I}-\tanh\left(\beta\frac{\hslash\omega}{2}\right)\sigma_{z}\right]\text{,} (106)

that is,

ρSQ(βω)=12[Itanh(βω2)σz].\rho_{\mathrm{SQ}}\left(\beta\text{, }\omega\right)=\frac{1}{2}\left[\mathrm{I}-\tanh\left(\beta\frac{\hslash\omega}{2}\right)\sigma_{z}\right]\text{.} (107)

In what follows, we shall use ρSQ(βω)\rho_{\mathrm{SQ}}\left(\beta\text{, }\omega\right) in Eq. (107) to calculate the Bures and the Sjöqvist metrics.

V.1.1 The Bures metric

We begin by noticing that dsBures2ds_{\mathrm{Bures}}^{2} in Eq. (73) becomes in our case

dsBures2\displaystyle ds_{\mathrm{Bures}}^{2} =14[H2H2]dβ2\displaystyle=\frac{1}{4}\left[\left\langle\mathrm{H}^{2}\right\rangle-\left\langle\mathrm{H}\right\rangle^{2}\right]d\beta^{2}
+14{β2{[(ωH)d]2(ωH)d2}+2nm|n|ωH|mEnEm|2(eβEneβEm)2𝒵(eβEn+eβEm)}dω2+\displaystyle+\frac{1}{4}\left\{\beta^{2}\left\{\left\langle\left[\left(\partial_{\omega}\mathrm{H}\right)_{d}\right]^{2}\right\rangle-\left\langle\left(\partial_{\omega}\mathrm{H}\right)_{d}\right\rangle^{2}\right\}+2\sum_{n\neq m}\left|\frac{\left\langle n|\partial_{\omega}\mathrm{H}|m\right\rangle}{E_{n}-E_{m}}\right|^{2}\frac{\left(e^{-\beta E_{n}}-e^{-\beta E_{m}}\right)^{2}}{\mathcal{Z}\cdot\left(e^{-\beta E_{n}}+e^{-\beta E_{m}}\right)}\right\}d\omega^{2}+
+14{2β[H(ωH)dH(ωH)d]}dβdω,\displaystyle+\frac{1}{4}\left\{2\beta\left[\left\langle\mathrm{H}\left(\partial_{\omega}\mathrm{H}\right)_{d}\right\rangle-\left\langle\mathrm{H}\right\rangle\left\langle\left(\partial_{\omega}\mathrm{H}\right)_{d}\right\rangle\right]\right\}d\beta d\omega\text{,} (108)

where, for simplicity of notation, we denote HSQ(ω)\mathrm{H}_{\mathrm{SQ}}\left(\omega\right) in Eq. (95) with H\mathrm{H}. To calculate dsBures2ds_{\mathrm{Bures}}^{2} in Eq. (73), we perform three distinct calculations. Specifically, we compute the metric tensor components gββg_{\beta\beta}, 2gβω2g_{\beta\omega}, and gωωg_{\omega\omega} defined as

gββ(βω)=def14[H2H2]2gβω(βω)=def14{2β[H(ωH)dH(ωH)d]}g_{\beta\beta}\left(\beta\text{, }\omega\right)\overset{\text{def}}{=}\frac{1}{4}\left[\left\langle\mathrm{H}^{2}\right\rangle-\left\langle\mathrm{H}\right\rangle^{2}\right]\text{, }2g_{\beta\omega}\left(\beta\text{, }\omega\right)\overset{\text{def}}{=}\frac{1}{4}\left\{2\beta\left[\left\langle\mathrm{H}\left(\partial_{\omega}\mathrm{H}\right)_{d}\right\rangle-\left\langle\mathrm{H}\right\rangle\left\langle\left(\partial_{\omega}\mathrm{H}\right)_{d}\right\rangle\right]\right\}\text{, } (109)

and,

gωω(βω)=def14{β2{[(ωH)d]2(ωH)d2}+2nm|n|ωH|mEnEm|2(eβEneβEm)2𝒵(eβEn+eβEm)},g_{\omega\omega}\left(\beta\text{, }\omega\right)\overset{\text{def}}{=}\frac{1}{4}\left\{\beta^{2}\left\{\left\langle\left[\left(\partial_{\omega}\mathrm{H}\right)_{d}\right]^{2}\right\rangle-\left\langle\left(\partial_{\omega}\mathrm{H}\right)_{d}\right\rangle^{2}\right\}+2\sum_{n\neq m}\left|\frac{\left\langle n|\partial_{\omega}\mathrm{H}|m\right\rangle}{E_{n}-E_{m}}\right|^{2}\frac{\left(e^{-\beta E_{n}}-e^{-\beta E_{m}}\right)^{2}}{\mathcal{Z}\cdot\left(e^{-\beta E_{n}}+e^{-\beta E_{m}}\right)}\right\}\text{,} (110)

respectively.

First sub-calculation

Let us begin with calculating (1/4)[H2H2]dβ2(1/4)\left[\left\langle\mathrm{H}^{2}\right\rangle-\left\langle\mathrm{H}\right\rangle^{2}\right]d\beta^{2}. Observe that the expectation value H2\left\langle\mathrm{H}^{2}\right\rangle of H2\mathrm{H}^{2} is given by,

H2\displaystyle\left\langle\mathrm{H}^{2}\right\rangle =tr(H2ρ)=i=01piEi=p0E0+p1E1\displaystyle=\mathrm{tr}\left(\mathrm{H}^{2}\rho\right)=\sum_{i=0}^{1}p_{i}E_{i}=p_{0}E_{0}+p_{1}E_{1}
=eβE0𝒵(ω2)2+eβE1𝒵(ω2)2\displaystyle=\frac{e^{-\beta E_{0}}}{\mathcal{Z}}\left(\frac{\hslash\omega}{2}\right)^{2}+\frac{e^{-\beta E_{1}}}{\mathcal{Z}}\left(-\frac{\hslash\omega}{2}\right)^{2}
=2ω24(eβω2𝒵+eβω2𝒵)\displaystyle=\frac{\hslash^{2}\omega^{2}}{4}\left(\frac{e^{-\beta\frac{\hslash\omega}{2}}}{\mathcal{Z}}+\frac{e^{\beta\frac{\hslash\omega}{2}}}{\mathcal{Z}}\right)
=2ω24,\displaystyle=\frac{\hslash^{2}\omega^{2}}{4}\text{,} (111)

that is,

H2=2ω24,\left\langle\mathrm{H}^{2}\right\rangle=\frac{\hslash^{2}\omega^{2}}{4}\text{,} (112)

where the partition function is 𝒵=defeβω2+eβω2=2cosh(βω2)\mathcal{Z}\overset{\text{def}}{=}e^{-\beta\frac{\hslash\omega}{2}}+e^{\beta\frac{\hslash\omega}{2}}=2\cosh\left(\beta\frac{\hslash\omega}{2}\right). Furthermore, we note that the expectation value H\left\langle\mathrm{H}\right\rangle of the Hamiltonian is

H\displaystyle\left\langle\mathrm{H}\right\rangle =tr(ρH)=i=01piEi=p0E0+p1E1=eβE0𝒵ω2eβE1𝒵ω2\displaystyle=\mathrm{tr}\left(\rho\mathrm{H}\right)=\sum_{i=0}^{1}p_{i}E_{i}=p_{0}E_{0}+p_{1}E_{1}=\frac{e^{-\beta E_{0}}}{\mathcal{Z}}\frac{\hslash\omega}{2}-\frac{e^{-\beta E_{1}}}{\mathcal{Z}}\frac{\hslash\omega}{2}
=eβE0eβE1𝒵ω2=eβω2eβω2eβω2+eβω2ω2=2sinh(βω2)2cosh(βω2)ω2=ω2tanh(βω2),\displaystyle=\frac{e^{-\beta E_{0}}-e^{-\beta E_{1}}}{\mathcal{Z}}\frac{\hslash\omega}{2}=\frac{e^{-\beta\frac{\hslash\omega}{2}}-e^{\beta\frac{\hslash\omega}{2}}}{e^{-\beta\frac{\hslash\omega}{2}}+e^{\beta\frac{\hslash\omega}{2}}}\frac{\hslash\omega}{2}=-\frac{2\sinh\left(\beta\frac{\hslash\omega}{2}\right)}{2\cosh\left(\beta\frac{\hslash\omega}{2}\right)}\frac{\hslash\omega}{2}=-\frac{\hslash\omega}{2}\tanh\left(\beta\frac{\hslash\omega}{2}\right)\text{,} (113)

that is,

H=ω2tanh(βω2).\left\langle\mathrm{H}\right\rangle=-\frac{\hslash\omega}{2}\tanh\left(\beta\frac{\hslash\omega}{2}\right)\text{.} (114)

Therefore, using Eqs. (112) and (114), (1/4)[H2H2]dβ2(1/4)\left[\left\langle\mathrm{H}^{2}\right\rangle-\left\langle\mathrm{H}\right\rangle^{2}\right]d\beta^{2} becomes

gββ(βω)dβ2=def14[H2H2]dβ2=216ω2[1tanh2(βω2)]dβ2.g_{\beta\beta}\left(\beta\text{, }\omega\right)d\beta^{2}\overset{\text{def}}{=}\frac{1}{4}\left[\left\langle\mathrm{H}^{2}\right\rangle-\left\langle\mathrm{H}\right\rangle^{2}\right]d\beta^{2}=\frac{\hslash^{2}}{16}\omega^{2}\left[1-\tanh^{2}\left(\beta\frac{\hslash\omega}{2}\right)\right]d\beta^{2}\text{.} (115)

For completeness, we remark that 1tanh2[β(ω/2)]1-\tanh^{2}\left[\beta\left(\hslash\omega/2\right)\right] in Eq. (115) can also be expressed as 11/cosh2[β(ω/2)]\cosh^{2}\left[\beta\left(\hslash\omega/2\right)\right]. The calculation of gββ(βω)g_{\beta\beta}\left(\beta\text{, }\omega\right) in Eq. (115) ends our first sub-calculation.

Second sub-calculation

Let us focus on the second term in Eq. (109). We start by noting that H(ωH)d\left\langle\mathrm{H}\left(\partial_{\omega}\mathrm{H}\right)_{d}\right\rangle is given by

H(ωH)d\displaystyle\left\langle\mathrm{H}\left(\partial_{\omega}\mathrm{H}\right)_{d}\right\rangle =i=01piEiωEi=p0E0ωE0+p1E1ωE1\displaystyle=\sum_{i=0}^{1}p_{i}E_{i}\partial_{\omega}E_{i}=p_{0}E_{0}\partial_{\omega}E_{0}+p_{1}E_{1}\partial_{\omega}E_{1}
=p0ω2ω(ω2)+p1(ω2)ω(ω2)\displaystyle=p_{0}\frac{\hslash\omega}{2}\partial_{\omega}\left(\frac{\hslash\omega}{2}\right)+p_{1}\left(-\frac{\hslash\omega}{2}\right)\partial_{\omega}\left(-\frac{\hslash\omega}{2}\right)
=24ωp0+24ωp1=24ω(p0+p1)=24ω,\displaystyle=\frac{\hslash^{2}}{4}\omega p_{0}+\frac{\hslash^{2}}{4}\omega p_{1}=\frac{\hslash^{2}}{4}\omega\left(p_{0}+p_{1}\right)=\frac{\hslash^{2}}{4}\omega\text{,} (116)

that is,

H(ωH)d=24ω.\left\langle\mathrm{H}\left(\partial_{\omega}\mathrm{H}\right)_{d}\right\rangle=\frac{\hslash^{2}}{4}\omega\text{.} (117)

Moreover, (ωH)d\left\langle\left(\partial_{\omega}\mathrm{H}\right)_{d}\right\rangle can be expressed as

(ωH)d\displaystyle\left\langle\left(\partial_{\omega}\mathrm{H}\right)_{d}\right\rangle =i=01piωEi=p0ωE0+p1ωE1=2p02p1\displaystyle=\sum_{i=0}^{1}p_{i}\partial_{\omega}E_{i}=p_{0}\partial_{\omega}E_{0}+p_{1}\partial_{\omega}E_{1}=\frac{\hslash}{2}p_{0}-\frac{\hslash}{2}p_{1}
=2eβE0eβE1𝒵=2eβω2eβω2eβω2+eβω2=22sinh(βω2)2cosh(βω2)\displaystyle=\frac{\hslash}{2}\frac{e^{-\beta E_{0}}-e^{-\beta E_{1}}}{\mathcal{Z}}=\frac{\hslash}{2}\frac{e^{-\beta\frac{\hslash\omega}{2}}-e^{\beta\frac{\hslash\omega}{2}}}{e^{-\beta\frac{\hslash\omega}{2}}+e^{\beta\frac{\hslash\omega}{2}}}=-\frac{\hslash}{2}\frac{2\sinh\left(\beta\frac{\hslash\omega}{2}\right)}{2\cosh\left(\beta\frac{\hslash\omega}{2}\right)}
=2tanh(βω2),\displaystyle=-\frac{\hslash}{2}\tanh\left(\beta\frac{\hslash\omega}{2}\right)\text{,} (118)

that is,

(ωH)d=2tanh(βω2).\left\langle\left(\partial_{\omega}\mathrm{H}\right)_{d}\right\rangle=-\frac{\hslash}{2}\tanh\left(\beta\frac{\hslash\omega}{2}\right)\text{.} (119)

Therefore, using Eqs. (114), (117), and (119), we obtain

2gβω(βω)dβdω=def14{2β[H(ωH)dH(ωH)d]}dβdω=28βω[1tanh2(βω2)]dβdω.2g_{\beta\omega}\left(\beta\text{, }\omega\right)d\beta d\omega\overset{\text{def}}{=}\frac{1}{4}\left\{2\beta\left[\left\langle\mathrm{H}\left(\partial_{\omega}\mathrm{H}\right)_{d}\right\rangle-\left\langle\mathrm{H}\right\rangle\left\langle\left(\partial_{\omega}\mathrm{H}\right)_{d}\right\rangle\right]\right\}d\beta d\omega=\frac{\hslash^{2}}{8}\beta\omega\left[1-\tanh^{2}\left(\beta\frac{\hslash\omega}{2}\right)\right]d\beta d\omega\text{.} (120)

The calculation of 2gβω(βω)2g_{\beta\omega}\left(\beta\text{, }\omega\right) in Eq. (120) ends our second sub-calculation.

Third sub-calculation

Let us now calculate the term in Eq. (110). Recall from Eq. (119) that (ωH)d=2tanh(βω2)\left\langle\left(\partial_{\omega}\mathrm{H}\right)_{d}\right\rangle=-\frac{\hslash}{2}\tanh\left(\beta\frac{\hslash\omega}{2}\right). Therefore, we have

(ωH)d2=24tanh2(βω2).\left\langle\left(\partial_{\omega}\mathrm{H}\right)_{d}\right\rangle^{2}=\frac{\hslash^{2}}{4}\tanh^{2}\left(\beta\frac{\hslash\omega}{2}\right)\text{.} (121)

Moreover, we note that [(ωH)d]2\left\langle\left[\left(\partial_{\omega}\mathrm{H}\right)_{d}\right]^{2}\right\rangle can be rewritten as

[(ωH)d]2\displaystyle\left\langle\left[\left(\partial_{\omega}\mathrm{H}\right)_{d}\right]^{2}\right\rangle =i=01pi(ωEi)2=p0(ωE0)2+p1(ωE1)2\displaystyle=\sum_{i=0}^{1}p_{i}\left(\partial_{\omega}E_{i}\right)^{2}=p_{0}\left(\partial_{\omega}E_{0}\right)^{2}+p_{1}\left(\partial_{\omega}E_{1}\right)^{2}
=p0[ω(ω2)]2+p1[ω(ω2)]2\displaystyle=p_{0}\left[\partial_{\omega}\left(-\frac{\hslash\omega}{2}\right)\right]^{2}+p_{1}\left[\partial_{\omega}\left(\frac{\hslash\omega}{2}\right)\right]^{2}
=24p0+24p1=24(p0+p1)=24,\displaystyle=\frac{\hslash^{2}}{4}p_{0}+\frac{\hslash^{2}}{4}p_{1}=\frac{\hslash^{2}}{4}\left(p_{0}+p_{1}\right)=\frac{\hslash^{2}}{4}\text{,} (122)

that is,

[(ωH)d]2=24.\left\langle\left[\left(\partial_{\omega}\mathrm{H}\right)_{d}\right]^{2}\right\rangle=\frac{\hslash^{2}}{4}\text{.} (123)

Finally, note that

2nm|n|ωH|mEnEm|2(eβEneβEm)2𝒵(eβEn+eβEm)\displaystyle 2\sum_{n\neq m}\left|\frac{\left\langle n|\partial_{\omega}\mathrm{H}|m\right\rangle}{E_{n}-E_{m}}\right|^{2}\frac{\left(e^{-\beta E_{n}}-e^{-\beta E_{m}}\right)^{2}}{\mathcal{Z}\cdot\left(e^{-\beta E_{n}}+e^{-\beta E_{m}}\right)} =2𝒵|0|ωH|1E0E1|2eβE0eβE1eβE0+eβE1+2𝒵|1|ωH|0E1E0|2eβE1eβE0eβE1+eβE0\displaystyle=\frac{2}{\mathcal{Z}}\left|\frac{\left\langle 0|\partial_{\omega}\mathrm{H}|1\right\rangle}{E_{0}-E_{1}}\right|^{2}\frac{e^{-\beta E_{0}}-e^{-\beta E_{1}}}{e^{-\beta E_{0}}+e^{-\beta E_{1}}}+\frac{2}{\mathcal{Z}}\left|\frac{\left\langle 1|\partial_{\omega}\mathrm{H}|0\right\rangle}{E_{1}-E_{0}}\right|^{2}\frac{e^{-\beta E_{1}}-e^{-\beta E_{0}}}{e^{-\beta E_{1}}+e^{-\beta E_{0}}}
=0,\displaystyle=0\text{,} (124)

since 0|ωH|1=1|ωH|0=0\left\langle 0|\partial_{\omega}\mathrm{H}|1\right\rangle=\left\langle 1|\partial_{\omega}\mathrm{H}|0\right\rangle=0 as a consequence of the fact that H=HSQ(ω)\mathrm{H}=\mathrm{H}_{\mathrm{SQ}}\left(\omega\right) in Eq. (91) is diagonal. Therefore, using Eqs. (121), (123), and (124), we finally get that Eq. (110) becomes

gωω(βω)dω2=216β2[1tanh2(βω2)]dω2.g_{\omega\omega}\left(\beta\text{, }\omega\right)d\omega^{2}=\frac{\hslash^{2}}{16}\beta^{2}\left[1-\tanh^{2}\left(\beta\frac{\hslash\omega}{2}\right)\right]d\omega^{2}\text{.} (125)

The calculation of gωω(βω)g_{\omega\omega}\left(\beta\text{, }\omega\right) in Eq. (125) ends our third sub-calculation.

In conclusion, exploiting Eqs. (115), (120), and (125), the Bures metric dsBures2=gββ(βω)dβ2+gωω(βω)dω2+2gβω(βω)dβdωds_{\mathrm{Bures}}^{2}=g_{\beta\beta}\left(\beta\text{, }\omega\right)d\beta^{2}+g_{\omega\omega}\left(\beta\text{, }\omega\right)d\omega^{2}+2g_{\beta\omega}\left(\beta\text{, }\omega\right)d\beta d\omega in Eq. (108) becomes

dsBures2=2ω216[1tanh2(βω2)]dβ2+2β216[1tanh2(βω2)]dω2+2βω8[1tanh2(βω2)]dβdω.ds_{\mathrm{Bures}}^{2}=\frac{\hslash^{2}\omega^{2}}{16}\left[1-\tanh^{2}\left(\beta\frac{\hslash\omega}{2}\right)\right]d\beta^{2}+\frac{\hslash^{2}\beta^{2}}{16}\left[1-\tanh^{2}\left(\beta\frac{\hslash\omega}{2}\right)\right]d\omega^{2}+\frac{\hslash^{2}\beta\omega}{8}\left[1-\tanh^{2}\left(\beta\frac{\hslash\omega}{2}\right)\right]d\beta d\omega\text{.} (126)

Using Einstein’s summation convention, dsBures2=gij(Bures)(βω)dθidθjds_{\mathrm{Bures}}^{2}=g_{ij}^{\left(\mathrm{Bures}\right)}\left(\beta\text{, }\omega\right)d\theta^{i}d\theta^{j} with θ1=defβ\theta^{1}\overset{\text{def}}{=}\beta and θ2=defω\theta^{2}\overset{\text{def}}{=}\omega. Finally, using Eq. (126), the Bures metric metric tensor gij(Bures)(βω)g_{ij}^{\left(\mathrm{Bures}\right)}\left(\beta\text{, }\omega\right) becomes

gij(Bures)(βω)=216[1tanh2(βω2)](ω2βωβωβ2),g_{ij}^{\left(\mathrm{Bures}\right)}\left(\beta\text{, }\omega\right)=\frac{\hslash^{2}}{16}\left[1-\tanh^{2}\left(\beta\frac{\hslash\omega}{2}\right)\right]\left(\begin{array}[c]{cc}\omega^{2}&\beta\omega\\ \beta\omega&\beta^{2}\end{array}\right)\text{,} (127)

with 1i1\leq i, j2j\leq 2. Note that gij(Bures)(βω)g_{ij}^{\left(\mathrm{Bures}\right)}\left(\beta\text{, }\omega\right) in Eq. (127) equals the classical Fisher-Rao metric since there is no non-classical contribution in this case. The derivation of gij(Bures)(βω)g_{ij}^{\left(\mathrm{Bures}\right)}\left(\beta\text{, }\omega\right) in Eq. (127) ends our calculation of the Bures metric tensor for spin qubits. Interestingly, we observe that setting kB=1k_{B}=1, β=t1\beta=t^{-1}, and ωz=t\omega_{z}=t, our Eq. (127) reduces to the last relation obtained by Zanardi and collaborators in Ref. zanardi07 .

V.1.2 The Sjöqvist metric

Given the expression of ρSQ(βω)\rho_{\mathrm{SQ}}\left(\beta\text{, }\omega\right) in Eq. (107), we can proceed with the calculation of the Sjöqvist metric given by

dsSjo¨qvist2=14k=01dpk2pk+k=01pkdnk|(I|nknk|)|dnk.ds_{\mathrm{Sj\ddot{o}qvist}}^{2}=\frac{1}{4}\sum_{k=0}^{1}\frac{dp_{k}^{2}}{p_{k}}+\sum_{k=0}^{1}p_{k}\left\langle dn_{k}|\left(\mathrm{I}-\left|n_{k}\right\rangle\left\langle n_{k}\right|\right)|dn_{k}\right\rangle\text{.} (128)

In our case, we note that the probabilities p0p_{0} and p1p_{1} are given by

p0=p0(βω)=def1tanh(βω2)2, and p1=p1(βω)=def1+tanh(βω2)2,p_{0}=p_{0}\left(\beta\text{, }\omega\right)\overset{\text{def}}{=}\frac{1-\tanh\left(\beta\frac{\hslash\omega}{2}\right)}{2}\text{, and }p_{1}=p_{1}\left(\beta\text{, }\omega\right)\overset{\text{def}}{=}\frac{1+\tanh\left(\beta\frac{\hslash\omega}{2}\right)}{2}\text{,} (129)

respectively. Furthermore, the states |n0\left|n_{0}\right\rangle and |n1\left|n_{1}\right\rangle are

|n0=def|0, and |n1=def|1.\left|n_{0}\right\rangle\overset{\text{def}}{=}\left|0\right\rangle\text{, and }\left|n_{1}\right\rangle\overset{\text{def}}{=}\left|1\right\rangle\text{.} (130)

Observe that since nk=nk(βω)n_{k}=n_{k}\left(\beta\text{, }\omega\right), we have that dnk=defnkβdβ+nkωdωdn_{k}\overset{\text{def}}{=}\frac{\partial n_{k}}{\partial\beta}d\beta+\frac{\partial n_{k}}{\partial\omega}d\omega. In our case, we get from Eq. (130) that |dnk=|0\left|dn_{k}\right\rangle=\left|0\right\rangle. From Eq. (128), dsSjo¨qvist2ds_{\mathrm{Sj\ddot{o}qvist}}^{2} reduces to

dsSjo¨qvist2=14k=01dpk2pk=14(dp02p0+dp12p1),ds_{\mathrm{Sj\ddot{o}qvist}}^{2}=\frac{1}{4}\sum_{k=0}^{1}\frac{dp_{k}^{2}}{p_{k}}=\frac{1}{4}\left(\frac{dp_{0}^{2}}{p_{0}}+\frac{dp_{1}^{2}}{p_{1}}\right)\text{,} (131)

where the differentials dp0dp_{0} and dp1dp_{1} are given by

dp0=defp0βdβ+p0ωdω, and dp1=defp1βdβ+p1ωdω,dp_{0}\overset{\text{def}}{=}\frac{\partial p_{0}}{\partial\beta}d\beta+\frac{\partial p_{0}}{\partial\omega}d\omega\text{, and }dp_{1}\overset{\text{def}}{=}\frac{\partial p_{1}}{\partial\beta}d\beta+\frac{\partial p_{1}}{\partial\omega}d\omega\text{,} (132)

respectively. Therefore, substituting Eq. (132) into Eq. (131), we get

dsSjo¨qvist2\displaystyle ds_{\mathrm{Sj\ddot{o}qvist}}^{2} =14(βp0dβ+ωp0dω)2p0+14(βp1dβ+ωp1dω)2p1\displaystyle=\frac{1}{4}\frac{\left(\partial_{\beta}p_{0}d\beta+\partial_{\omega}p_{0}d\omega\right)^{2}}{p_{0}}+\frac{1}{4}\frac{\left(\partial_{\beta}p_{1}d\beta+\partial_{\omega}p_{1}d\omega\right)^{2}}{p_{1}}
=(βp0)2dβ2+(ωp0)2dω2+2βp0ωp0dβdω4p0+(βp1)2dβ2+(ωp1)2dω2+2βp1ωp1dβdω4p1\displaystyle=\frac{\left(\partial_{\beta}p_{0}\right)^{2}d\beta^{2}+\left(\partial_{\omega}p_{0}\right)^{2}d\omega^{2}+2\partial_{\beta}p_{0}\partial_{\omega}p_{0}d\beta d\omega}{4p_{0}}+\frac{\left(\partial_{\beta}p_{1}\right)^{2}d\beta^{2}+\left(\partial_{\omega}p_{1}\right)^{2}d\omega^{2}+2\partial_{\beta}p_{1}\partial_{\omega}p_{1}d\beta d\omega}{4p_{1}}
=[(βp0)24p0+(βp1)24p1]dβ2+[(ωp0)24p0+(ωp1)24p1]dω2+[2βp0ωp04p0+2βp1ωp14p1]dβdω,\displaystyle=\left[\frac{\left(\partial_{\beta}p_{0}\right)^{2}}{4p_{0}}+\frac{\left(\partial_{\beta}p_{1}\right)^{2}}{4p_{1}}\right]d\beta^{2}+\left[\frac{\left(\partial_{\omega}p_{0}\right)^{2}}{4p_{0}}+\frac{\left(\partial_{\omega}p_{1}\right)^{2}}{4p_{1}}\right]d\omega^{2}+\left[\frac{2\partial_{\beta}p_{0}\partial_{\omega}p_{0}}{4p_{0}}+\frac{2\partial_{\beta}p_{1}\partial_{\omega}p_{1}}{4p_{1}}\right]d\beta d\omega\text{,} (133)

that is,

dsSjo¨qvist2=[(βp0)24p0+(βp1)24p1]dβ2+[(ωp0)24p0+(ωp1)24p1]dω2+[2βp0ωp04p0+2βp1ωp14p1]dβdω.ds_{\mathrm{Sj\ddot{o}qvist}}^{2}=\left[\frac{\left(\partial_{\beta}p_{0}\right)^{2}}{4p_{0}}+\frac{\left(\partial_{\beta}p_{1}\right)^{2}}{4p_{1}}\right]d\beta^{2}+\left[\frac{\left(\partial_{\omega}p_{0}\right)^{2}}{4p_{0}}+\frac{\left(\partial_{\omega}p_{1}\right)^{2}}{4p_{1}}\right]d\omega^{2}+\left[\frac{2\partial_{\beta}p_{0}\partial_{\omega}p_{0}}{4p_{0}}+\frac{2\partial_{\beta}p_{1}\partial_{\omega}p_{1}}{4p_{1}}\right]d\beta d\omega\text{.} (134)

From Eq. (129), we observe that

βp0\displaystyle\partial_{\beta}p_{0} =ω4[1tanh2(βω2)]ωp0=β4[1tanh2(βω2)],\displaystyle=-\frac{\hslash\omega}{4}\left[1-\tanh^{2}\left(\beta\frac{\hslash\omega}{2}\right)\right]\text{, }\partial_{\omega}p_{0}=-\frac{\hslash\beta}{4}\left[1-\tanh^{2}\left(\beta\frac{\hslash\omega}{2}\right)\right]\text{,}
βp1\displaystyle\partial_{\beta}p_{1} =ω4[1tanh2(βω2)]ωp1=β4[1tanh2(βω2)].\displaystyle=\frac{\hslash\omega}{4}\left[1-\tanh^{2}\left(\beta\frac{\hslash\omega}{2}\right)\right]\text{, }\partial_{\omega}p_{1}=\frac{\hslash\beta}{4}\left[1-\tanh^{2}\left(\beta\frac{\hslash\omega}{2}\right)\right]\text{.} (135)

Finally, substituting Eq. (135) into Eq. (134), we obtain

dsSjo¨qvist2=2ω216[1tanh2(βω2)]dβ2+2β216[1tanh2(βω2)]dω2+2βω8[1tanh2(βω2)]dβdω.ds_{\mathrm{Sj\ddot{o}qvist}}^{2}=\frac{\hslash^{2}\omega^{2}}{16}\left[1-\tanh^{2}\left(\beta\frac{\hslash\omega}{2}\right)\right]d\beta^{2}+\frac{\hslash^{2}\beta^{2}}{16}\left[1-\tanh^{2}\left(\beta\frac{\hslash\omega}{2}\right)\right]d\omega^{2}+\frac{\hslash^{2}\beta\omega}{8}\left[1-\tanh^{2}\left(\beta\frac{\hslash\omega}{2}\right)\right]d\beta d\omega\text{.} (136)

Using Einstein’s summation convention, dsSjo¨qvist2=gij(Sjo¨qvist)(βω)dθidθjds_{\mathrm{Sj\ddot{o}qvist}}^{2}=g_{ij}^{\left(\mathrm{Sj\ddot{o}qvist}\right)}\left(\beta\text{, }\omega\right)d\theta^{i}d\theta^{j} with θ1=defβ\theta^{1}\overset{\text{def}}{=}\beta and θ2=defω\theta^{2}\overset{\text{def}}{=}\omega. Finally, using Eq. (136), the Sjöqvist metric metric tensor becomes

gij(Sjo¨qvist)(βω)=216[1tanh2(βω2)](ω2βωβωβ2),g_{ij}^{\left(\mathrm{Sj\ddot{o}qvist}\right)}\left(\beta\text{, }\omega\right)=\frac{\hslash^{2}}{16}\left[1-\tanh^{2}\left(\beta\frac{\hslash\omega}{2}\right)\right]\left(\begin{array}[c]{cc}\omega^{2}&\beta\omega\\ \beta\omega&\beta^{2}\end{array}\right)\text{,} (137)

with 1i1\leq i, j2j\leq 2. Note that gij(Sjo¨qvist)(βω)g_{ij}^{\left(\mathrm{Sj\ddot{o}qvist}\right)}\left(\beta\text{, }\omega\right) in Eq. (127) is equal to the classical Fisher-Rao metric since the non-classical contribution is absent in this case. The derivation of gij(Sjo¨qvist)(βω)g_{ij}^{\left(\mathrm{Sj\ddot{o}qvist}\right)}\left(\beta\text{, }\omega\right) in Eq. (137) ends our calculation of the Sjöqvist metric tensor for spin qubits.

Recalling the general expressions of the Bures and Sjöqvist metrics in Eqs. (37) and (90) and, moreover, from our first set of explicit calculations, a few remarks are in order. First, both metrics have a classical and a non-classical contribution. Second, the classical Fisher-Rao metric contribution is related to changes dpn=βpndβ+hpndhdp_{n}=\partial_{\beta}p_{n}d\beta+\partial_{h}p_{n}dh in the probabilities pn(βh)eβEn(h)p_{n}\left(\beta\text{, }h\right)\propto e^{-\beta E_{n}\left(h\right)} with {En(h)}\left\{E_{n}\left(h\right)\right\} being the eigenvalues of the Hamiltonian. Finally, the non-classical contribution in the two metrics is linked to changes |dn=h|ndh=|hndh\left|dn\right\rangle=\partial_{h}\left|n\right\rangle dh=\left|\partial_{h}n\right\rangle dh in the eigenvectors {|n(h)}\left\{\left|n\left(h\right)\right\rangle\right\} of the Hamiltonian. In our first Hamiltonian model, Hσz\propto\sigma_{z} is diagonal and, thus, its eigenvectors do not depend on any parameter. Therefore, we found that both the Bures and Sjöqvist metrics reduce to the classical Fisher-Rao metric. However, one expects that if  H is not proportional to the Pauli matrix operator σz\sigma_{z}, non-classical contributions do not vanish any longer and the two metrics may yield different quantum (i.e., non-classical) metric contributions. Indeed, if one considers a spin qubit Hamiltonian specified by a magnetic field with an orientation that is not constrained to be along the zz-axis, the Bures and Sjöqvist metrics happen to be different. In particular, for a time-independent and uniform magnetic field given by B=Bxx^+Bzz^\vec{B}=B_{x}\hat{x}+B_{z}\hat{z}, the spin qubit Hamiltonian becomes HSQ(ωxωz)=def(/2)(ωxσx+ωzσz){}_{\mathrm{SQ}}\left(\omega_{x}\text{, }\omega_{z}\right)\overset{\text{def}}{=}\left(\hslash/2\right)(\omega_{x}\sigma_{x}+\omega_{z}\sigma_{z}). Assuming ωx\omega_{x}-fixed0\neq 0, tuning only the parameters β\beta and ωz\omega_{z}, and repeating our metric calculations, it can be shown that the Bures and Sjöqvist metric tensor components gijBures(βωz)g_{ij}^{\mathrm{Bures}}\left(\beta\text{, }\omega_{z}\right) and gijSjo¨qvist(βωz)g_{ij}^{\mathrm{Sj\ddot{o}qvist}}\left(\beta\text{, }\omega_{z}\right) are

gijBures(βωz)=216[1tanh2(βωx2+ωz22)](ωx2+ωz2βωzβωzβ2ωz2ωx2+ωz2+42ωx2(ωx2+ωz2)2tanh2(βωx2+ωz22)1tanh2(βωx2+ωz22)),g_{ij}^{\mathrm{Bures}}\left(\beta\text{, }\omega_{z}\right)=\frac{\hslash^{2}}{16}\left[1-\tanh^{2}\left(\beta\frac{\hslash\sqrt{\omega_{x}^{2}+\omega_{z}^{2}}}{2}\right)\right]\left(\begin{array}[c]{cc}\omega_{x}^{2}+\omega_{z}^{2}&\beta\omega_{z}\\ \beta\omega_{z}&\beta^{2}\frac{\omega_{z}^{2}}{\omega_{x}^{2}+\omega_{z}^{2}}+\frac{4}{\hslash^{2}}\frac{\omega_{x}^{2}}{\left(\omega_{x}^{2}+\omega_{z}^{2}\right)^{2}}\frac{\tanh^{2}\left(\beta\frac{\hslash\sqrt{\omega_{x}^{2}+\omega_{z}^{2}}}{2}\right)}{1-\tanh^{2}\left(\beta\frac{\hslash\sqrt{\omega_{x}^{2}+\omega_{z}^{2}}}{2}\right)}\end{array}\right)\text{,} (138)

and,

gijSjo¨qvist(βωz)=216[1tanh2(βωx2+ωz22)](ωx2+ωz2βωzβωzβ2ωz2ωx2+ωz2+42ωx2(ωx2+ωz2)211tanh2(βωx2+ωz22)),g_{ij}^{\mathrm{Sj\ddot{o}qvist}}\left(\beta\text{, }\omega_{z}\right)=\frac{\hslash^{2}}{16}\left[1-\tanh^{2}\left(\beta\frac{\hslash\sqrt{\omega_{x}^{2}+\omega_{z}^{2}}}{2}\right)\right]\left(\begin{array}[c]{cc}\omega_{x}^{2}+\omega_{z}^{2}&\beta\omega_{z}\\ \beta\omega_{z}&\beta^{2}\frac{\omega_{z}^{2}}{\omega_{x}^{2}+\omega_{z}^{2}}+\frac{4}{\hslash^{2}}\frac{\omega_{x}^{2}}{\left(\omega_{x}^{2}+\omega_{z}^{2}\right)^{2}}\frac{1}{1-\tanh^{2}\left(\beta\frac{\hslash\sqrt{\omega_{x}^{2}+\omega_{z}^{2}}}{2}\right)}\end{array}\right)\text{,} (139)

respectively. For completeness, we remark that useful calculation techniques to arrive at expressions as in Eqs. (138) and (139) will be performed in the next subsection where HSQ(ωxωz){}_{\mathrm{SQ}}\left(\omega_{x}\text{, }\omega_{z}\right) will be replaced by the superconducting flux qubit Hamiltonian HSFQ(Δϵ)=def(/2)(Δσx+ϵσz)\mathrm{H}_{\mathrm{SFQ}}\left(\Delta\text{, }\epsilon\right)\overset{\text{def}}{=}\left(-\hslash/2\right)\left(\Delta\sigma_{x}+\epsilon\sigma_{z}\right). Returning to our considerations, recall that for any xx\in\mathbb{R}, we have

tanh2(x)1tanh2(x)=sinh2(x), and 11tanh2(x)=cosh2(x).\frac{\tanh^{2}\left(x\right)}{1-\tanh^{2}\left(x\right)}=\sinh^{2}\left(x\right)\text{, and }\frac{1}{1-\tanh^{2}\left(x\right)}=\cosh^{2}\left(x\right)\text{.} (140)

Then, using Eqs. (138) and (139), we obtain

0gωzωznc, Bures(βωz)gωzωzncSjo¨qvist(βωz)=sinh2(βωx2+ωz22)cosh2(βωx2+ωz22)=tanh2(βωx2+ωz22)1,0\leq\frac{g_{\omega_{z}\omega_{z}}^{\mathrm{nc}\text{, {Bures}}}\left(\beta\text{, }\omega_{z}\right)}{g_{\omega_{z}\omega_{z}}^{\mathrm{nc}\text{, }\mathrm{Sj\ddot{o}qvist}}\left(\beta\text{, }\omega_{z}\right)}=\frac{\sinh^{2}\left(\beta\frac{\hslash\sqrt{\omega_{x}^{2}+\omega_{z}^{2}}}{2}\right)}{\cosh^{2}\left(\beta\frac{\hslash\sqrt{\omega_{x}^{2}+\omega_{z}^{2}}}{2}\right)}=\tanh^{2}\left(\beta\frac{\hslash\sqrt{\omega_{x}^{2}+\omega_{z}^{2}}}{2}\right)\leq 1\text{,} (141)

with gωzωznc, Bures(βωz)g_{\omega_{z}\omega_{z}}^{\mathrm{nc}\text{, {Bures}}}\left(\beta\text{, }\omega_{z}\right) and gωzωzncSjo¨qvist(βωz)g_{\omega_{z}\omega_{z}}^{\mathrm{nc}\text{, }\mathrm{Sj\ddot{o}qvist}}\left(\beta\text{, }\omega_{z}\right) denoting the non-classical contributions in the Bures and Sjöqvist metric cases, respectively. From Eqs. (138) and (139), we conclude that the introduction of a nonvanishing component of the magnetic field along the xx-direction introduces a visible non-commutative probabilistic structure in the quantum mechanics of the system characterized by a non-classical scenario with [ρρ+dρ]0\left[\rho\text{, }\rho+d\rho\right]\neq 0). In such a case, the Bures and the Sjöqvist metrics exhibit a different behavior as evident from their nonclassical metric tensor components (i.e., gωzωznc(βωz)g_{\omega_{z}\omega_{z}}^{\mathrm{nc}}\left(\beta\text{, }\omega_{z}\right)) in Eq. (141).

V.2 Superconducting flux qubits

Let us consider a system with an Hamiltonian described by HSFQ(Δϵ)=def(/2)(Δσx+ϵσz)\mathrm{H}_{\mathrm{SFQ}}\left(\Delta\text{, }\epsilon\right)\overset{\text{def}}{=}\left(-\hslash/2\right)\left(\Delta\sigma_{x}+\epsilon\sigma_{z}\right) in Eq. (92). The thermal state ρSFQ(βϵ)\rho_{\mathrm{SFQ}}\left(\beta\text{, }\epsilon\right) corresponding to HSFQ(Δϵ)\mathrm{H}_{\mathrm{SFQ}}\left(\Delta\text{, }\epsilon\right) with Δ\Delta assumed to be constant is given by

ρSFQ(βϵ)=defeβHSFQ(Δϵ)tr(eβHSFQ(Δϵ)).\rho_{\mathrm{SFQ}}\left(\beta\text{, }\epsilon\right)\overset{\text{def}}{=}\frac{e^{-\beta\mathrm{H}_{\mathrm{SFQ}}\left(\Delta\text{, }\epsilon\right)}}{\mathrm{tr}\left(e^{-\beta\mathrm{H}_{\mathrm{SFQ}}\left(\Delta\text{, }\epsilon\right)}\right)}\text{.} (142)

Observe that HSFQ(Δϵ)\mathrm{H}_{\mathrm{SFQ}}\left(\Delta\text{, }\epsilon\right) is diagonalizable and can be recast as HSFQ=MHSFQHSFQ(diagonal)MHSFQ1\mathrm{H}_{\mathrm{SFQ}}=M_{\mathrm{H}_{\mathrm{SFQ}}}\mathrm{H}_{\mathrm{SFQ}}^{\left(\mathrm{diagonal}\right)}M_{\mathrm{H}_{\mathrm{SFQ}}}^{-1} where MHSFQM_{\mathrm{H}_{\mathrm{SFQ}}} and MHSFQ1M_{\mathrm{H}_{\mathrm{SFQ}}}^{-1} are the eigenvector matrix and its inverse, respectively. Therefore, after some algebra, ρSFQ(βϵ)\rho_{\mathrm{SFQ}}\left(\beta\text{, }\epsilon\right) in Eq. (142) can be rewritten as

ρSFQ(βϵ)\displaystyle\rho_{\mathrm{SFQ}}\left(\beta\text{, }\epsilon\right) =eβHSFQ(Δϵ)tr(eβHSFQ(Δϵ))=eβMHSFQHSFQ(diagonal)MHSFQ1tr(eβMHSFQHSFQ(diagonal)MHSFQ1)\displaystyle=\frac{e^{-\beta\mathrm{H}_{\mathrm{SFQ}}\left(\Delta\text{, }\epsilon\right)}}{\mathrm{tr}(e^{-\beta\mathrm{H}_{\mathrm{SFQ}}\left(\Delta\text{, }\epsilon\right)})}=\frac{e^{-\beta M_{\mathrm{H}_{\mathrm{SFQ}}}\mathrm{H}_{\mathrm{SFQ}}^{\left(\mathrm{diagonal}\right)}M_{\mathrm{H}_{\mathrm{SFQ}}}^{-1}}}{\mathrm{tr}(e^{-\beta M_{\mathrm{H}_{\mathrm{SFQ}}}\mathrm{H}_{\mathrm{SFQ}}^{\left(\mathrm{diagonal}\right)}M_{\mathrm{H}_{\mathrm{SFQ}}}^{-1}})}
=MHSFQeβHSFQ(diagonal)MHSFQ1tr(MHSFQeβHSFQ(diagonal)MHSFQ1)\displaystyle=\frac{M_{\mathrm{H}_{\mathrm{SFQ}}}e^{-\beta\mathrm{H}_{\mathrm{SFQ}}^{\left(\mathrm{diagonal}\right)}}M_{\mathrm{H}_{\mathrm{SFQ}}}^{-1}}{\mathrm{tr}(M_{\mathrm{H}_{\mathrm{SFQ}}}e^{-\beta\mathrm{H}_{\mathrm{SFQ}}^{\left(\mathrm{diagonal}\right)}}M_{\mathrm{H}_{\mathrm{SFQ}}}^{-1})}
=MHSFQeβHSFQ(diagonal)tr(eβHSFQ(diagonal))MHSFQ1,\displaystyle=M_{\mathrm{H}_{\mathrm{SFQ}}}\frac{e^{-\beta\mathrm{H}_{\mathrm{SFQ}}^{\left(\mathrm{diagonal}\right)}}}{\mathrm{tr}(e^{-\beta\mathrm{H}_{\mathrm{SFQ}}^{\left(\mathrm{diagonal}\right)}})}M_{\mathrm{H}_{\mathrm{SFQ}}}^{-1}\text{,} (143)

that is,

ρSFQ(βϵ)=MHSFQeβHSFQ(diagonal)tr(eβHSFQ(diagonal))MHSFQ1=MHSFQρSFQ(diagonal)(βϵ)MHSFQ1.\rho_{\mathrm{SFQ}}\left(\beta\text{, }\epsilon\right)=M_{\mathrm{H}_{\mathrm{SFQ}}}\frac{e^{-\beta\mathrm{H}_{\mathrm{SFQ}}^{\left(\mathrm{diagonal}\right)}}}{\mathrm{tr}(e^{-\beta\mathrm{H}_{\mathrm{SFQ}}^{\left(\mathrm{diagonal}\right)}})}M_{\mathrm{H}_{\mathrm{SFQ}}}^{-1}=M_{\mathrm{H}_{\mathrm{SFQ}}}\rho_{\mathrm{SFQ}}^{\left(\mathrm{diagonal}\right)}\left(\beta\text{, }\epsilon\right)M_{\mathrm{H}_{\mathrm{SFQ}}}^{-1}\text{.} (144)

The quantity HSFQ(diagonal)\mathrm{H}_{\mathrm{SFQ}}^{\left(\mathrm{diagonal}\right)} in Eq. (144) is defined as,

HSFQ(diagonal)=defE0|n1n1|+E1|n0n0|.\mathrm{H}_{\mathrm{SFQ}}^{\left(\mathrm{diagonal}\right)}\overset{\text{def}}{=}E_{0}\left|n_{1}\right\rangle\left\langle n_{1}\right|+E_{1}\left|n_{0}\right\rangle\left\langle n_{0}\right|\text{.} (145)

The the eigenvalues E0E_{0} and E1E_{1} are given by E0=def(/2)νE_{0}\overset{\text{def}}{=}-\left(\hslash/2\right)\nu and E1=def+(/2)νE_{1}\overset{\text{def}}{=}+\left(\hslash/2\right)\nu, respectively, with ν=defΔ2+ϵ2\nu\overset{\text{def}}{=}\sqrt{\Delta^{2}+\epsilon^{2}}. For later use, it is convenient to introduce the notation E~0=defE1\tilde{E}_{0}\overset{\text{def}}{=}E_{1} and E~1=defE0\tilde{E}_{1}\overset{\text{def}}{=}E_{0} so that HSFQ(diagonal)=defE~0|n0n0|+E~1|n1n1|\mathrm{H}_{\mathrm{SFQ}}^{\left(\mathrm{diagonal}\right)}\overset{\text{def}}{=}\tilde{E}_{0}\left|n_{0}\right\rangle\left\langle n_{0}\right|+\tilde{E}_{1}\left|n_{1}\right\rangle\left\langle n_{1}\right|. The two orthonormal eigenvectors corresponding to E0E_{0} and E1E_{1} are |n1\left|n_{1}\right\rangle and |n0\left|n_{0}\right\rangle, respectively. They are given by

|n0=def12(ϵϵ2+Δ2ϵ2+Δ2ϵϵ2+Δ2Δϵ2+Δ2ϵϵ2+Δ2) and, |n1=def12(ϵ+ϵ2+Δ2ϵ2+Δ2+ϵϵ2+Δ2Δϵ2+Δ2+ϵϵ2+Δ2),\left|n_{0}\right\rangle\overset{\text{def}}{=}\frac{1}{\sqrt{2}}\left(\begin{array}[c]{c}\frac{\epsilon-\sqrt{\epsilon^{2}+\Delta^{2}}}{\sqrt{\epsilon^{2}+\Delta^{2}-\epsilon\sqrt{\epsilon^{2}+\Delta^{2}}}}\\ \frac{\Delta}{\sqrt{\epsilon^{2}+\Delta^{2}-\epsilon\sqrt{\epsilon^{2}+\Delta^{2}}}}\end{array}\right)\text{ and, }\left|n_{1}\right\rangle\overset{\text{def}}{=}\frac{1}{\sqrt{2}}\left(\begin{array}[c]{c}\frac{\epsilon+\sqrt{\epsilon^{2}+\Delta^{2}}}{\sqrt{\epsilon^{2}+\Delta^{2}+\epsilon\sqrt{\epsilon^{2}+\Delta^{2}}}}\\ \frac{\Delta}{\sqrt{\epsilon^{2}+\Delta^{2}+\epsilon\sqrt{\epsilon^{2}+\Delta^{2}}}}\end{array}\right)\text{,} (146)

respectively. A suitable choice for the eigenvector matrix MHSFQM_{\mathrm{H}_{\mathrm{SFQ}}} and its inverse MHSFQ1M_{\mathrm{H}_{\mathrm{SFQ}}}^{-1} in Eq. (144) can be expressed as

MHSFQ=def(ϵ+νΔϵνΔ11) and, MHSFQ1=def(Δ2ννϵ2νΔ2νν+ϵ2ν),M_{\mathrm{H}_{\mathrm{SFQ}}}\overset{\text{def}}{=}\left(\begin{array}[c]{cc}\frac{\epsilon+\nu}{\Delta}&\frac{\epsilon-\nu}{\Delta}\\ 1&1\end{array}\right)\text{ and, }M_{\mathrm{H}_{\mathrm{SFQ}}}^{-1}\overset{\text{def}}{=}\left(\begin{array}[c]{cc}\frac{\Delta}{2\nu}&\frac{\nu-\epsilon}{2\nu}\\ -\frac{\Delta}{2\nu}&\frac{\nu+\epsilon}{2\nu}\end{array}\right)\text{,} (147)

respectively. Using Eqs. (145) and (147), ρSFQ(βϵ)\rho_{\mathrm{SFQ}}\left(\beta\text{, }\epsilon\right) in Eq. (144) becomes

ρSFQ(βϵ)=12(1+ϵϵ2+Δ2tanh(βϵ2+Δ22)Δϵ2+Δ2tanh(βϵ2+Δ22)Δϵ2+Δ2tanh(βϵ2+Δ22)1ϵϵ2+Δ2tanh(βϵ2+Δ22)),\rho_{\mathrm{SFQ}}\left(\beta\text{, }\epsilon\right)=\frac{1}{2}\left(\begin{array}[c]{cc}1+\frac{\epsilon}{\sqrt{\epsilon^{2}+\Delta^{2}}}\tanh\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)&\frac{\Delta}{\sqrt{\epsilon^{2}+\Delta^{2}}}\tanh\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\\ \frac{\Delta}{\sqrt{\epsilon^{2}+\Delta^{2}}}\tanh\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)&1-\frac{\epsilon}{\sqrt{\epsilon^{2}+\Delta^{2}}}\tanh\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\end{array}\right)\text{,} (148)

that is,

ρSFQ(βϵ)=12[I+(Δϵ2+Δ2σx+ϵϵ2+Δ2σz)tanh(βϵ2+Δ22)].\rho_{\mathrm{SFQ}}\left(\beta\text{, }\epsilon\right)=\frac{1}{2}\left[\mathrm{I}+\left(\frac{\Delta}{\sqrt{\epsilon^{2}+\Delta^{2}}}\sigma_{x}+\frac{\epsilon}{\sqrt{\epsilon^{2}+\Delta^{2}}}\sigma_{z}\right)\tanh\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]\text{.} (149)

For completeness, we note here that the spectral decomposition of ρSFQ(βϵ)=MHSFQρSFQ(diagonal)(βϵ)MHSFQ1\rho_{\mathrm{SFQ}}\left(\beta\text{, }\epsilon\right)=M_{\mathrm{H}_{\mathrm{SFQ}}}\rho_{\mathrm{SFQ}}^{\left(\mathrm{diagonal}\right)}\left(\beta\text{, }\epsilon\right)M_{\mathrm{H}_{\mathrm{SFQ}}}^{-1} in Eq. (149) is given by ρSFQ=defp0|n0n0|+p1|n1n1|\rho_{\mathrm{SFQ}}\overset{\text{def}}{=}p_{0}\left|n_{0}\right\rangle\left\langle n_{0}\right|+p_{1}\left|n_{1}\right\rangle\left\langle n_{1}\right|. The probabilities p0p_{0} and p1p_{1} are

p0=defeβE~0𝒵=12[1tanh(βϵ2+Δ22)], and p1=defeβE~1𝒵=12[1+tanh(βϵ2+Δ22)],p_{0}\overset{\text{def}}{=}\frac{e^{-\beta\tilde{E}_{0}}}{\mathcal{Z}}=\frac{1}{2}\left[1-\tanh\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]\text{, and }p_{1}\overset{\text{def}}{=}\frac{e^{-\beta\tilde{E}_{1}}}{\mathcal{Z}}=\frac{1}{2}\left[1+\tanh\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]\text{,} (150)

respectively, with 𝒵=defeβE0+eβE1=eβE~0+eβE~1=2cosh(βϵ2+Δ22)\mathcal{Z}\overset{\text{def}}{\mathcal{=}}e^{-\beta E_{0}}+e^{-\beta E_{1}}=e^{-\beta\tilde{E}_{0}}+e^{-\beta\tilde{E}_{1}}=2\cosh(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}) denoting the partition function of the system. In what follows, we shall use ρSFQ(βϵ)\rho_{\mathrm{SFQ}}\left(\beta\text{, }\epsilon\right) in Eq. (149) to calculate the Bures and the Sjöqvist metrics.

V.2.1 The Bures metric

For simplicity of notation, we replace HSFQ\mathrm{H}_{\mathrm{SFQ}} with H\mathrm{H} in the forthcoming calculation. We begin by noting that, in our case, the general expression of the Bures metric dsBures2ds_{\mathrm{Bures}}^{2} in Eq. (73) becomes

dsBures2\displaystyle ds_{\mathrm{Bures}}^{2} =14[H2H2]dβ2\displaystyle=\frac{1}{4}\left[\left\langle\mathrm{H}^{2}\right\rangle-\left\langle\mathrm{H}\right\rangle^{2}\right]d\beta^{2}
+14{β2{[(ϵH)d]2(ϵH)d2}+2nm|n|ϵH|mE~nE~m|2(eβE~neβE~m)2𝒵(eβE~n+eβE~m)}dϵ2+\displaystyle+\frac{1}{4}\left\{\beta^{2}\left\{\left\langle\left[\left(\partial_{\epsilon}\mathrm{H}\right)_{d}\right]^{2}\right\rangle-\left\langle\left(\partial_{\epsilon}\mathrm{H}\right)_{d}\right\rangle^{2}\right\}+2\sum_{n\neq m}\left|\frac{\left\langle n|\partial_{\epsilon}\mathrm{H}|m\right\rangle}{\tilde{E}_{n}-\tilde{E}_{m}}\right|^{2}\frac{\left(e^{-\beta\tilde{E}_{n}}-e^{-\beta\tilde{E}_{m}}\right)^{2}}{\mathcal{Z}\cdot\left(e^{-\beta\tilde{E}_{n}}+e^{-\beta\tilde{E}_{m}}\right)}\right\}d\epsilon^{2}+
+14{2β[H(ϵH)dH(ϵH)d]}dβdϵ.\displaystyle+\frac{1}{4}\left\{2\beta\left[\left\langle\mathrm{H}\left(\partial_{\epsilon}\mathrm{H}\right)_{d}\right\rangle-\left\langle\mathrm{H}\right\rangle\left\langle\left(\partial_{\epsilon}\mathrm{H}\right)_{d}\right\rangle\right]\right\}d\beta d\epsilon\text{.} (151)

As previously pointed out in this manuscript, dsBures2ds_{\mathrm{Bures}}^{2} is the sum of two contributions, the classical Fisher-Rao information metric contribution and the non-classical metric contribution described in the summation term in the right-hand-side of Eq. (151). In what follows, we shall the that the presence of nonvanishing terms |n|ϵH|m|2\left|\left\langle n|\partial_{\epsilon}\mathrm{H}|m\right\rangle\right|^{2} leads to the existence of a non-classical contribution in dsBures2ds_{\mathrm{Bures}}^{2}. Following our previous line of reasoning, we partition our calculation in three parts. In particular, since dsBures2=gββ(βϵ)dβ2+gϵϵ(βϵ)dϵ2+2gβϵ(βϵ)dβdϵds_{\mathrm{Bures}}^{2}=g_{\beta\beta}\left(\beta\text{, }\epsilon\right)d\beta^{2}+g_{\epsilon\epsilon}\left(\beta\text{, }\epsilon\right)d\epsilon^{2}+2g_{\beta\epsilon}\left(\beta\text{, }\epsilon\right)d\beta d\epsilon, we focus on computing gββ(βϵ)g_{\beta\beta}\left(\beta\text{, }\epsilon\right), 2gβϵ(βϵ)2g_{\beta\epsilon}\left(\beta\text{, }\epsilon\right), and gϵϵ(βϵ)g_{\epsilon\epsilon}\left(\beta\text{, }\epsilon\right).

V.2.2 First sub-calculation

We proceed here with the first sub-calculation. We recall that gββ(βϵ)dβ2=(1/4)[H2H2]dβ2g_{\beta\beta}\left(\beta\text{, }\epsilon\right)d\beta^{2}=(1/4)\left[\left\langle\mathrm{H}^{2}\right\rangle-\left\langle\mathrm{H}\right\rangle^{2}\right]d\beta^{2}. Note that H2\left\langle\mathrm{H}^{2}\right\rangle and H2\left\langle\mathrm{H}\right\rangle^{2} are given by,

H2=tr(H2ρ)=24(ϵ2+Δ2),\left\langle\mathrm{H}^{2}\right\rangle=\mathrm{tr}\left(\mathrm{H}^{2}\rho\right)=\frac{\hslash^{2}}{4}\left(\epsilon^{2}+\Delta^{2}\right)\text{,} (152)

and,

H2=[tr(Hρ)]2=24(ϵ2+Δ2)tanh2(βϵ2+Δ22),\left\langle\mathrm{H}\right\rangle^{2}=\left[\mathrm{tr}\left(\mathrm{H}\rho\right)\right]^{2}=\frac{\hslash^{2}}{4}\left(\epsilon^{2}+\Delta^{2}\right)\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\text{,} (153)

respectively. Therefore, using Eqs. (152) and (153), gββ(βϵ)dβ2g_{\beta\beta}\left(\beta\text{, }\epsilon\right)d\beta^{2} becomes

gββ(βϵ)dβ2=216(ϵ2+Δ2)[1tanh2(βϵ2+Δ22)]dβ2.g_{\beta\beta}\left(\beta\text{, }\epsilon\right)d\beta^{2}=\frac{\hslash^{2}}{16}\left(\epsilon^{2}+\Delta^{2}\right)\left[1-\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]d\beta^{2}\text{.} (154)

Our fist sub-calculation ends with the derivation of Eq. (154).

V.2.3 Second sub-calculation

In our second calculation, we focus on calculating the term 2gβϵ(βϵ)dβdϵ2g_{\beta\epsilon}\left(\beta\text{, }\epsilon\right)d\beta d\epsilon defined as

2gβϵ(βϵ)dβdϵ=def14{2β[H(ϵH)dH(ϵH)d]}dβdϵ.2g_{\beta\epsilon}\left(\beta\text{, }\epsilon\right)d\beta d\epsilon\overset{\text{def}}{=}\frac{1}{4}\left\{2\beta\left[\left\langle\mathrm{H}\left(\partial_{\epsilon}\mathrm{H}\right)_{d}\right\rangle-\left\langle\mathrm{H}\right\rangle\left\langle\left(\partial_{\epsilon}\mathrm{H}\right)_{d}\right\rangle\right]\right\}d\beta d\epsilon\text{.} (155)

Note that H(ϵH)d\left\langle\mathrm{H}\left(\partial_{\epsilon}\mathrm{H}\right)_{d}\right\rangle can be recast as

H(ϵH)d\displaystyle\left\langle\mathrm{H}\left(\partial_{\epsilon}\mathrm{H}\right)_{d}\right\rangle =i=01piE~iϵE~i=p0E~0ϵE~0+p1E~1ϵE~1\displaystyle=\sum_{i=0}^{1}p_{i}\tilde{E}_{i}\partial_{\epsilon}\tilde{E}_{i}=p_{0}\tilde{E}_{0}\partial_{\epsilon}\tilde{E}_{0}+p_{1}\tilde{E}_{1}\partial_{\epsilon}\tilde{E}_{1}
=p0(2ϵ2+Δ2)ϵ(2ϵ2+Δ2)+p1(2ϵ2+Δ2)ϵ(2ϵ2+Δ2)\displaystyle=p_{0}\left(\frac{\hslash}{2}\sqrt{\epsilon^{2}+\Delta^{2}}\right)\partial_{\epsilon}\left(\frac{\hslash}{2}\sqrt{\epsilon^{2}+\Delta^{2}}\right)+p_{1}\left(-\frac{\hslash}{2}\sqrt{\epsilon^{2}+\Delta^{2}}\right)\partial_{\epsilon}\left(-\frac{\hslash}{2}\sqrt{\epsilon^{2}+\Delta^{2}}\right)
=(p0+p1)(2ϵ2+Δ2)ϵ(2ϵ2+Δ2)\displaystyle=\left(p_{0}+p_{1}\right)\left(\frac{\hslash}{2}\sqrt{\epsilon^{2}+\Delta^{2}}\right)\partial_{\epsilon}\left(\frac{\hslash}{2}\sqrt{\epsilon^{2}+\Delta^{2}}\right)
=(2ϵ2+Δ2)ϵ(2ϵ2+Δ2)\displaystyle=\left(\frac{\hslash}{2}\sqrt{\epsilon^{2}+\Delta^{2}}\right)\partial_{\epsilon}\left(\frac{\hslash}{2}\sqrt{\epsilon^{2}+\Delta^{2}}\right)
=24ϵ,\displaystyle=\frac{\hslash^{2}}{4}\epsilon\text{,} (156)

that is,

H(ϵH)d=24ϵ.\left\langle\mathrm{H}\left(\partial_{\epsilon}\mathrm{H}\right)_{d}\right\rangle=\frac{\hslash^{2}}{4}\epsilon\text{.} (157)

We also note that the expectation value H\left\langle\mathrm{H}\right\rangle of the Hamiltonian H\mathrm{H} equals

H=2ϵ2+Δ2tanh(βϵ2+Δ22).\left\langle\mathrm{H}\right\rangle=-\frac{\hslash}{2}\sqrt{\epsilon^{2}+\Delta^{2}}\tanh\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\text{.} (158)

Finally, the quantity (ϵH)d\left\langle\left(\partial_{\epsilon}\mathrm{H}\right)_{d}\right\rangle can be rewritten as

(ϵH)d\displaystyle\left\langle\left(\partial_{\epsilon}\mathrm{H}\right)_{d}\right\rangle =i=01piϵE~i=p0ϵE~0+p1ϵE~1\displaystyle=\sum_{i=0}^{1}p_{i}\partial_{\epsilon}\tilde{E}_{i}=p_{0}\partial_{\epsilon}\tilde{E}_{0}+p_{1}\partial_{\epsilon}\tilde{E}_{1}
=p0ϵ(2ϵ2+Δ2)+p1ϵ(2ϵ2+Δ2)\displaystyle=p_{0}\partial_{\epsilon}\left(\frac{\hslash}{2}\sqrt{\epsilon^{2}+\Delta^{2}}\right)+p_{1}\partial_{\epsilon}\left(-\frac{\hslash}{2}\sqrt{\epsilon^{2}+\Delta^{2}}\right)
=eβϵ2+Δ22𝒵(2)ϵϵ2+Δ2+eβϵ2+Δ22𝒵(2)ϵϵ2+Δ2\displaystyle=\frac{e^{-\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}}}{\mathcal{Z}}\left(\frac{\hslash}{2}\right)\frac{\epsilon}{\sqrt{\epsilon^{2}+\Delta^{2}}}+\frac{e^{\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}}}{\mathcal{Z}}\left(-\frac{\hslash}{2}\right)\frac{\epsilon}{\sqrt{\epsilon^{2}+\Delta^{2}}}
=22sinh(βϵ2+Δ22)2cosh(βϵ2+Δ22)ϵϵ2+Δ2\displaystyle=-\frac{\hslash}{2}\frac{2\sinh\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)}{2\cosh\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)}\frac{\epsilon}{\sqrt{\epsilon^{2}+\Delta^{2}}}
=2ϵϵ2+Δ2tanh(βϵ2+Δ22),\displaystyle=-\frac{\hslash}{2}\frac{\epsilon}{\sqrt{\epsilon^{2}+\Delta^{2}}}\tanh\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\text{,} (159)

that is,

(ϵH)d=2ϵϵ2+Δ2tanh(βϵ2+Δ22).\left\langle\left(\partial_{\epsilon}\mathrm{H}\right)_{d}\right\rangle=-\frac{\hslash}{2}\frac{\epsilon}{\sqrt{\epsilon^{2}+\Delta^{2}}}\tanh\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\text{.} (160)

Finally, using Eqs. (157), (158), and (160), 2gβϵ(βϵ)dβdϵ2g_{\beta\epsilon}\left(\beta\text{, }\epsilon\right)d\beta d\epsilon in Eq. (155) becomes

2gβϵ(βϵ)dβdϵ\displaystyle 2g_{\beta\epsilon}\left(\beta\text{, }\epsilon\right)d\beta d\epsilon =14{2β[24ϵ+2ϵ2+Δ2tanh(βϵ2+Δ22)(2ϵϵ2+Δ2tanh(βϵ2+Δ22))]}dβdϵ\displaystyle=\frac{1}{4}\left\{2\beta\left[\begin{array}[c]{c}\frac{\hslash^{2}}{4}\epsilon+\frac{\hslash}{2}\sqrt{\epsilon^{2}+\Delta^{2}}\tanh\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\cdot\\ \left(-\frac{\hslash}{2}\frac{\epsilon}{\sqrt{\epsilon^{2}+\Delta^{2}}}\tanh\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right)\end{array}\right]\right\}d\beta d\epsilon (163)
=14{2β[24ϵ24ϵtanh2(βϵ2+Δ22)]}dβdϵ\displaystyle=\frac{1}{4}\left\{2\beta\left[\frac{\hslash^{2}}{4}\epsilon-\frac{\hslash^{2}}{4}\epsilon\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]\right\}d\beta d\epsilon
=28βϵ[1tanh2(βϵ2+Δ22)]dβdϵ,\displaystyle=\frac{\hslash^{2}}{8}\beta\epsilon\left[1-\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]d\beta d\epsilon\text{,} (164)

that is,

2gβϵ(βϵ)dβdϵ=28βϵ[1tanh2(βϵ2+Δ22)]dβdϵ.2g_{\beta\epsilon}\left(\beta\text{, }\epsilon\right)d\beta d\epsilon=\frac{\hslash^{2}}{8}\beta\epsilon\left[1-\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]d\beta d\epsilon\text{.} (165)

Our second sub-calculation ends with the derivation of Eq. (165).

V.2.4 Third sub-calculation

In what follows, we focus on the calculation of gϵϵ(βϵ)dϵ2g_{\epsilon\epsilon}\left(\beta\text{, }\epsilon\right)d\epsilon^{2},

gϵϵ(βϵ)dϵ2=def14{β2{[(ϵH)d]2(ϵH)d2}+2nm|n|ϵH|mE~nE~m|2(eβE~neβE~m)2𝒵(eβE~n+eβE~m)}dϵ2.\ g_{\epsilon\epsilon}\left(\beta\text{, }\epsilon\right)d\epsilon^{2}\ \overset{\text{def}}{=}\frac{1}{4}\left\{\beta^{2}\left\{\left\langle\left[\left(\partial_{\epsilon}\mathrm{H}\right)_{d}\right]^{2}\right\rangle-\left\langle\left(\partial_{\epsilon}\mathrm{H}\right)_{d}\right\rangle^{2}\right\}+2\sum_{n\neq m}\left|\frac{\left\langle n|\partial_{\epsilon}\mathrm{H}|m\right\rangle}{\tilde{E}_{n}-\tilde{E}_{m}}\right|^{2}\frac{\left(e^{-\beta\tilde{E}_{n}}-e^{-\beta\tilde{E}_{m}}\right)^{2}}{\mathcal{Z}\cdot\left(e^{-\beta\tilde{E}_{n}}+e^{-\beta\tilde{E}_{m}}\right)}\right\}d\epsilon^{2}\text{.} (166)

Let us recall that (ϵH)d\left\langle\left(\partial_{\epsilon}\mathrm{H}\right)_{d}\right\rangle is given in Eq. (160). Therefore, we get

(ϵH)d2=24ϵ2ϵ2+Δ2tanh2(βϵ2+Δ22).\left\langle\left(\partial_{\epsilon}\mathrm{H}\right)_{d}\right\rangle^{2}=\frac{\hslash^{2}}{4}\frac{\epsilon^{2}}{\epsilon^{2}+\Delta^{2}}\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\text{.} (167)

Moreover, (ϵH)d2\left\langle\left(\partial_{\epsilon}\mathrm{H}\right)_{d}^{2}\right\rangle is given by

(ϵH)d2\displaystyle\left\langle\left(\partial_{\epsilon}\mathrm{H}\right)_{d}^{2}\right\rangle =i=01pi(ϵE~i)2=p0(ϵE~0)2+p1(ϵE~1)2\displaystyle=\sum_{i=0}^{1}p_{i}\left(\partial_{\epsilon}\tilde{E}_{i}\right)^{2}=p_{0}\left(\partial_{\epsilon}\tilde{E}_{0}\right)^{2}+p_{1}\left(\partial_{\epsilon}\tilde{E}_{1}\right)^{2}
=p024ϵ2ϵ2+Δ2+p124ϵ2ϵ2+Δ2\displaystyle=p_{0}\frac{\hslash^{2}}{4}\frac{\epsilon^{2}}{\epsilon^{2}+\Delta^{2}}+p_{1}\frac{\hslash^{2}}{4}\frac{\epsilon^{2}}{\epsilon^{2}+\Delta^{2}}
=(p0+p1)24ϵ2ϵ2+Δ2\displaystyle=\left(p_{0}+p_{1}\right)\frac{\hslash^{2}}{4}\frac{\epsilon^{2}}{\epsilon^{2}+\Delta^{2}}
=24ϵ2ϵ2+Δ2,\displaystyle=\frac{\hslash^{2}}{4}\frac{\epsilon^{2}}{\epsilon^{2}+\Delta^{2}}\text{,} (168)

that is,

(ϵH)d2=24ϵ2ϵ2+Δ2.\left\langle\left(\partial_{\epsilon}\mathrm{H}\right)_{d}^{2}\right\rangle=\frac{\hslash^{2}}{4}\frac{\epsilon^{2}}{\epsilon^{2}+\Delta^{2}}\text{.} (169)

Finally, let us focus on the term in Eq. (166) given by

2nm|n|ϵH|mE~nE~m|2(eβE~neβE~m)2𝒵(eβE~n+eβE~m)\displaystyle 2\sum_{n\neq m}\left|\frac{\left\langle n|\partial_{\epsilon}\mathrm{H}|m\right\rangle}{\tilde{E}_{n}-\tilde{E}_{m}}\right|^{2}\frac{\left(e^{-\beta\tilde{E}_{n}}-e^{-\beta\tilde{E}_{m}}\right)^{2}}{\mathcal{Z}\cdot\left(e^{-\beta\tilde{E}_{n}}+e^{-\beta\tilde{E}_{m}}\right)} =2𝒵|n0|ϵH|n1E~0E~1|2(eβE~0eβE~1)2(eβE~0+eβE~1)+\displaystyle=\frac{2}{\mathcal{Z}}\left|\frac{\left\langle n_{0}|\partial_{\epsilon}\mathrm{H}|n_{1}\right\rangle}{\tilde{E}_{0}-\tilde{E}_{1}}\right|^{2}\frac{\left(e^{-\beta\tilde{E}_{0}}-e^{-\beta\tilde{E}_{1}}\right)^{2}}{\left(e^{-\beta\tilde{E}_{0}}+e^{-\beta\tilde{E}_{1}}\right)}+
+2𝒵|n1|ϵH|n0E~1E~0|2(eβE~1eβE~0)2(eβE~0+eβE~1)\displaystyle+\frac{2}{\mathcal{Z}}\left|\frac{\left\langle n_{1}|\partial_{\epsilon}\mathrm{H}|n_{0}\right\rangle}{\tilde{E}_{1}-\tilde{E}_{0}}\right|^{2}\frac{\left(e^{-\beta\tilde{E}_{1}}-e^{-\beta\tilde{E}_{0}}\right)^{2}}{\left(e^{-\beta\tilde{E}_{0}}+e^{-\beta\tilde{E}_{1}}\right)}
=2𝒵(eβE~0eβE~1)2(eβE~0+eβE~1)(|n0|ϵH|n1|2+|n1|ϵH|n0|2)|E~0E~1|2\displaystyle=\frac{2}{\mathcal{Z}}\frac{\left(e^{-\beta\tilde{E}_{0}}-e^{-\beta\tilde{E}_{1}}\right)^{2}}{\left(e^{-\beta\tilde{E}_{0}}+e^{-\beta\tilde{E}_{1}}\right)}\frac{\left(\left|\left\langle n_{0}|\partial_{\epsilon}\mathrm{H}|n_{1}\right\rangle\right|^{2}+\left|\left\langle n_{1}|\partial_{\epsilon}\mathrm{H}|n_{0}\right\rangle\right|^{2}\right)}{\left|\tilde{E}_{0}-\tilde{E}_{1}\right|^{2}}
=2(eβE~0eβE~1)2(eβE~0+eβE~1)224Δ2Δ2+ϵ2+24Δ2Δ2+ϵ22(ϵ2+Δ2)\displaystyle=2\frac{\left(e^{-\beta\tilde{E}_{0}}-e^{-\beta\tilde{E}_{1}}\right)^{2}}{\left(e^{-\beta\tilde{E}_{0}}+e^{-\beta\tilde{E}_{1}}\right)^{2}}\frac{\frac{\hslash^{2}}{4}\frac{\Delta^{2}}{\Delta^{2}+\epsilon^{2}}+\frac{\hslash^{2}}{4}\frac{\Delta^{2}}{\Delta^{2}+\epsilon^{2}}}{\hslash^{2}\left(\epsilon^{2}+\Delta^{2}\right)}
=Δ2(Δ2+ϵ2)2tanh2(βϵ2+Δ22),\displaystyle=\frac{\Delta^{2}}{\left(\Delta^{2}+\epsilon^{2}\right)^{2}}\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\text{,} (170)

that is,

2nm|n|ϵH|mE~nE~m|2(eβE~neβE~m)2𝒵(eβE~n+eβE~m)dϵ2=Δ2(Δ2+ϵ2)2tanh2(βϵ2+Δ22)dϵ2.2\sum_{n\neq m}\left|\frac{\left\langle n|\partial_{\epsilon}\mathrm{H}|m\right\rangle}{\tilde{E}_{n}-\tilde{E}_{m}}\right|^{2}\frac{\left(e^{-\beta\tilde{E}_{n}}-e^{-\beta\tilde{E}_{m}}\right)^{2}}{\mathcal{Z}\cdot\left(e^{-\beta\tilde{E}_{n}}+e^{-\beta\tilde{E}_{m}}\right)}d\epsilon^{2}=\frac{\Delta^{2}}{\left(\Delta^{2}+\epsilon^{2}\right)^{2}}\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)d\epsilon^{2}\text{.} (171)

For clarity, note that ϵH\partial_{\epsilon}\mathrm{H} in Eq. (171) equals (/2)σz\left(-\hslash/2\right)\sigma_{z} in the standard computational basis {|0|1}\left\{\left|0\right\rangle\text{, }\left|1\right\rangle\right\}. Therefore, combining Eqs. (167), (169), and (171) we get that gϵϵ(βϵ)dϵ2g_{\epsilon\epsilon}\left(\beta\text{, }\epsilon\right)d\epsilon^{2} in Eq. (166) equals

gϵϵ(βϵ)dϵ2=216{4Δ22(Δ2+ϵ2)2tanh2(βϵ2+Δ22)+β2ϵ2ϵ2+Δ2[1tanh2(βϵ2+Δ22)]}dϵ2.g_{\epsilon\epsilon}\left(\beta\text{, }\epsilon\right)d\epsilon^{2}=\frac{\hslash^{2}}{16}\left\{\frac{4\Delta^{2}}{\hslash^{2}\left(\Delta^{2}+\epsilon^{2}\right)^{2}}\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)+\frac{\beta^{2}\epsilon^{2}}{\epsilon^{2}+\Delta^{2}}\left[1-\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]\right\}d\epsilon^{2}\text{.} (172)

Then, using Eqs. (154), (165), and (172), dsBures2ds_{\mathrm{Bures}}^{2} in Eq. (151) becomes

dsBures2\displaystyle ds_{\mathrm{Bures}}^{2} =216(ϵ2+Δ2)[1tanh2(βϵ2+Δ22)]dβ2+\displaystyle=\frac{\hslash^{2}}{16}\left(\epsilon^{2}+\Delta^{2}\right)\left[1-\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]d\beta^{2}+
+28βϵ[1tanh2(βϵ2+Δ22)]dβdϵ+\displaystyle+\frac{\hslash^{2}}{8}\beta\epsilon\left[1-\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]d\beta d\epsilon+
+216{4Δ22(Δ2+ϵ2)2tanh2(βϵ2+Δ22)+β2ϵ2ϵ2+Δ2[1tanh2(βϵ2+Δ22)]}dϵ2.\displaystyle+\frac{\hslash^{2}}{16}\left\{\frac{4\Delta^{2}}{\hslash^{2}\left(\Delta^{2}+\epsilon^{2}\right)^{2}}\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)+\frac{\beta^{2}\epsilon^{2}}{\epsilon^{2}+\Delta^{2}}\left[1-\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]\right\}d\epsilon^{2}\text{.} (173)

Finally, using Eq. (173), the Bures metric tensor in the case of a superconducting flux qubit becomes

gij(Bures)(βϵ)=216[1tanh2(βϵ2+Δ22)](ϵ2+Δ2βϵβϵβ2ϵ2ϵ2+Δ2+4Δ22(Δ2+ϵ2)2tanh2(βϵ2+Δ22)1tanh2(βϵ2+Δ22)),g_{ij}^{\left(\mathrm{Bures}\right)}\left(\beta\text{, }\epsilon\right)=\frac{\hslash^{2}}{16}\left[1-\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]\left(\begin{array}[c]{cc}\epsilon^{2}+\Delta^{2}&\beta\epsilon\\ \beta\epsilon&\frac{\beta^{2}\epsilon^{2}}{\epsilon^{2}+\Delta^{2}}+\frac{4\Delta^{2}}{\hslash^{2}\left(\Delta^{2}+\epsilon^{2}\right)^{2}}\frac{\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)}{1-\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)}\end{array}\right)\text{,} (174)

with 1i1\leq i, j2j\leq 2. The derivation of Eqs. (173) and (174) completes our calculation of the Bures metric structure for a superconducting flux qubit.

V.2.5 The Sjöqvist metric

Let us observe that the Sjöqvist metric in Eq. (90) can be rewritten in our case as

dsSjo¨qvist2=def14k=01dpk2pk+k=01pkdsk2,ds_{\mathrm{Sj\ddot{o}qvist}}^{2}\overset{\text{def}}{=}\frac{1}{4}\sum_{k=0}^{1}\frac{dp_{k}^{2}}{p_{k}}+\sum_{k=0}^{1}p_{k}ds_{k}^{2}\text{,} (175)

where dsk2=def[dnk|(I|nknk|)|dnk]ds_{k}^{2}\overset{\text{def}}{=}\left[\left\langle dn_{k}|\left(\mathrm{I}-\left|n_{k}\right\rangle\left\langle n_{k}\right|\right)|dn_{k}\right\rangle\right] and nk|nk=δkk\left\langle n_{k}\left|n_{k^{\prime}}\right.\right\rangle=\delta_{kk^{\prime}}. From Eq. (146), the states |dn0\left|dn_{0}\right\rangle and |dn1\left|dn_{1}\right\rangle become

|dn0=def12(12Δ2ϵ2+Δ2ϵ2+Δ2ϵ(ϵ2+Δ2ϵϵ2+Δ2)3/2dϵ12Δ2ϵ2+Δ2(ϵ2+Δ2ϵ)2(ϵ2+Δ2ϵϵ2+Δ2)3/2dϵ), and |dn1=def12(12Δ2ϵ2+Δ2ϵ+ϵ2+Δ2(ϵ2+Δ2+ϵϵ2+Δ2)3/2dϵ12Δ2ϵ2+Δ2(ϵ+ϵ2+Δ2)2(ϵ2+Δ2+ϵϵ2+Δ2)3/2dϵ),\left|dn_{0}\right\rangle\overset{\text{def}}{=}\frac{1}{\sqrt{2}}\left(\begin{array}[c]{c}\frac{1}{2}\frac{\Delta^{2}}{\sqrt{\epsilon^{2}+\Delta^{2}}}\frac{\sqrt{\epsilon^{2}+\Delta^{2}}-\epsilon}{\left(\epsilon^{2}+\Delta^{2}-\epsilon\sqrt{\epsilon^{2}+\Delta^{2}}\right)^{3/2}}d\epsilon\\ \frac{1}{2}\frac{\Delta^{2}}{\sqrt{\epsilon^{2}+\Delta^{2}}}\frac{\left(\sqrt{\epsilon^{2}+\Delta^{2}}-\epsilon\right)^{2}}{\left(\epsilon^{2}+\Delta^{2}-\epsilon\sqrt{\epsilon^{2}+\Delta^{2}}\right)^{3/2}}d\epsilon\end{array}\right)\text{, and }\left|dn_{1}\right\rangle\overset{\text{def}}{=}\frac{1}{\sqrt{2}}\left(\begin{array}[c]{c}\frac{1}{2}\frac{\Delta^{2}}{\sqrt{\epsilon^{2}+\Delta^{2}}}\frac{\epsilon+\sqrt{\epsilon^{2}+\Delta^{2}}}{\left(\epsilon^{2}+\Delta^{2}+\epsilon\sqrt{\epsilon^{2}+\Delta^{2}}\right)^{3/2}}d\epsilon\\ -\frac{1}{2}\frac{\Delta^{2}}{\sqrt{\epsilon^{2}+\Delta^{2}}}\frac{\left(\epsilon+\sqrt{\epsilon^{2}+\Delta^{2}}\right)^{2}}{\left(\epsilon^{2}+\Delta^{2}+\epsilon\sqrt{\epsilon^{2}+\Delta^{2}}\right)^{3/2}}d\epsilon\end{array}\right)\text{,} (176)

respectively. Eqs. (146) and (176) will be used to calculate the nonclassical contribution that appears in the Sjöqvist metric in Eq. (175). In what follows, however, let us consider the classical contribution [dsSjo¨qvist2](classical)\left[ds_{\mathrm{Sj\ddot{o}qvist}}^{2}\right]^{\left(\mathrm{classical}\right)} in Eq. (175). We note that [dsSjo¨qvist2](classical)\left[ds_{\mathrm{Sj\ddot{o}qvist}}^{2}\right]^{\left(\mathrm{classical}\right)} equals

14k=01dpk2pk\displaystyle\frac{1}{4}\sum_{k=0}^{1}\frac{dp_{k}^{2}}{p_{k}} =14dp02p0+14dp12p1\displaystyle=\frac{1}{4}\frac{dp_{0}^{2}}{p_{0}}+\frac{1}{4}\frac{dp_{1}^{2}}{p_{1}}
=[(βp0)24p0+(βp1)24p1]dβ2+[(ϵp0)24p0+(ϵp1)24p1]dϵ2+[2βp0ϵp04p0+2βp1ϵp14p1]dβdϵ.\displaystyle=\left[\frac{\left(\partial_{\beta}p_{0}\right)^{2}}{4p_{0}}+\frac{\left(\partial_{\beta}p_{1}\right)^{2}}{4p_{1}}\right]d\beta^{2}+\left[\frac{\left(\partial_{\epsilon}p_{0}\right)^{2}}{4p_{0}}+\frac{\left(\partial_{\epsilon}p_{1}\right)^{2}}{4p_{1}}\right]d\epsilon^{2}+\left[\frac{2\partial_{\beta}p_{0}\partial_{\epsilon}p_{0}}{4p_{0}}+\frac{2\partial_{\beta}p_{1}\partial_{\epsilon}p_{1}}{4p_{1}}\right]d\beta d\epsilon\text{.} (177)

Using Eq. (150), [dsSjo¨qvist2](classical)\left[ds_{\mathrm{Sj\ddot{o}qvist}}^{2}\right]^{\left(\mathrm{classical}\right)} in Eq. (177) reduces to

[dsSjo¨qvist2](classical)\displaystyle\left[ds_{\mathrm{Sj\ddot{o}qvist}}^{2}\right]^{\left(\mathrm{classical}\right)} =14k=01dpk2pk\displaystyle=\frac{1}{4}\sum_{k=0}^{1}\frac{dp_{k}^{2}}{p_{k}}
=216(ϵ2+Δ2)[1tanh2(βϵ2+Δ22)]dβ2+\displaystyle=\frac{\hslash^{2}}{16}\left(\epsilon^{2}+\Delta^{2}\right)\left[1-\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]d\beta^{2}+
+216β2ϵ2ϵ2+Δ2[1tanh2(βϵ2+Δ22)]dϵ2+\displaystyle+\frac{\hslash^{2}}{16}\frac{\beta^{2}\epsilon^{2}}{\epsilon^{2}+\Delta^{2}}\left[1-\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]d\epsilon^{2}+
+28βϵ[1tanh2(βϵ2+Δ22)]dβdϵ.\displaystyle+\frac{\hslash^{2}}{8}\beta\epsilon\left[1-\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]d\beta d\epsilon\text{.} (178)

We can now return our focus on the nonclassical contribution [dsSjo¨qvist2](quantum)\left[ds_{\mathrm{Sj\ddot{o}qvist}}^{2}\right]^{\left(\text{{quantum}}\right)} that specifies the Sjöqvist metric. We have

[dsSjo¨qvist2](quantum)\displaystyle\left[ds_{\mathrm{Sj\ddot{o}qvist}}^{2}\right]^{\left(\text{{quantum}}\right)} =k=01pkdnk|(I|nknk|)|dnk\displaystyle=\sum_{k=0}^{1}p_{k}\left\langle dn_{k}|\left(\mathrm{I}-\left|n_{k}\right\rangle\left\langle n_{k}\right|\right)|dn_{k}\right\rangle
=p0dn0|dn0p0|dn0|n0|2+p1dn1|dn1p1|dn1|n1|2.\displaystyle=p_{0}\left\langle dn_{0}|dn_{0}\right\rangle-p_{0}\left|\left\langle dn_{0}|n_{0}\right\rangle\right|^{2}+p_{1}\left\langle dn_{1}|dn_{1}\right\rangle-p_{1}\left|\left\langle dn_{1}|n_{1}\right\rangle\right|^{2}\text{.} (179)

A simple check allows us to verify that dn0|n0=0\left\langle dn_{0}|n_{0}\right\rangle=0 and dn1|n1=0\left\langle dn_{1}|n_{1}\right\rangle=0. Therefore, [dsSjo¨qvist2](quantum)\left[ds_{\mathrm{Sj\ddot{o}qvist}}^{2}\right]^{\left(\text{{quantum}}\right)} becomes

[dsSjo¨qvist2](quantum)\displaystyle\left[ds_{\mathrm{Sj\ddot{o}qvist}}^{2}\right]^{\left(\text{{quantum}}\right)} =p0dn0|dn0+p1dn1|dn1\displaystyle=p_{0}\left\langle dn_{0}|dn_{0}\right\rangle+p_{1}\left\langle dn_{1}|dn_{1}\right\rangle
=p018Δ2Δ2+ϵ2(ϵΔ2+ϵ2)22Δ2+2ϵ22ϵΔ2+ϵ2(Δ2+ϵ2ϵΔ2+ϵ2)3dϵ2+\displaystyle=p_{0}\frac{1}{8}\frac{\Delta^{2}}{\Delta^{2}+\epsilon^{2}}\left(\epsilon-\sqrt{\Delta^{2}+\epsilon^{2}}\right)^{2}\frac{2\Delta^{2}+2\epsilon^{2}-2\epsilon\sqrt{\Delta^{2}+\epsilon^{2}}}{\left(\Delta^{2}+\epsilon^{2}-\epsilon\sqrt{\Delta^{2}+\epsilon^{2}}\right)^{3}}d\epsilon^{2}+
+p118Δ2Δ2+ϵ2(ϵ+Δ2+ϵ2)22Δ2+2ϵ2+2ϵΔ2+ϵ2(Δ2+ϵ2+ϵΔ2+ϵ2)3dϵ2\displaystyle+p_{1}\frac{1}{8}\frac{\Delta^{2}}{\Delta^{2}+\epsilon^{2}}\left(\epsilon+\sqrt{\Delta^{2}+\epsilon^{2}}\right)^{2}\frac{2\Delta^{2}+2\epsilon^{2}+2\epsilon\sqrt{\Delta^{2}+\epsilon^{2}}}{\left(\Delta^{2}+\epsilon^{2}+\epsilon\sqrt{\Delta^{2}+\epsilon^{2}}\right)^{3}}d\epsilon^{2}
=14Δ2(Δ2+ϵ2)2dϵ2\displaystyle=\frac{1}{4}\frac{\Delta^{2}}{\left(\Delta^{2}+\epsilon^{2}\right)^{2}}d\epsilon^{2}
=2164Δ22(Δ2+ϵ2)2dϵ2,\displaystyle=\frac{\hslash^{2}}{16}\frac{4\Delta^{2}}{\hslash^{2}\left(\Delta^{2}+\epsilon^{2}\right)^{2}}d\epsilon^{2}\text{,} (180)

that is,

[dsSjo¨qvist2](quantum)=2164Δ22(Δ2+ϵ2)2dϵ2.\left[ds_{\mathrm{Sj\ddot{o}qvist}}^{2}\right]^{\left(\text{{quantum}}\right)}=\frac{\hslash^{2}}{16}\frac{4\Delta^{2}}{\hslash^{2}\left(\Delta^{2}+\epsilon^{2}\right)^{2}}d\epsilon^{2}\text{.} (181)

Finally, combining Eqs. (178) and (181), the Sjöqvist metric dsSjo¨qvist2ds_{\mathrm{Sj\ddot{o}qvist}}^{2} becomes

dsSjo¨qvist2\displaystyle ds_{\mathrm{Sj\ddot{o}qvist}}^{2} =216(ϵ2+Δ2)[1tanh2(βϵ2+Δ22)]dβ2+\displaystyle=\frac{\hslash^{2}}{16}\left(\epsilon^{2}+\Delta^{2}\right)\left[1-\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]d\beta^{2}+
+28βϵ[1tanh2(βϵ2+Δ22)]dβdϵ+\displaystyle+\frac{\hslash^{2}}{8}\beta\epsilon\left[1-\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]d\beta d\epsilon+
+216{4Δ22(Δ2+ϵ2)2+β2ϵ2ϵ2+Δ2[1tanh2(βϵ2+Δ22)]}dϵ2.\displaystyle+\frac{\hslash^{2}}{16}\left\{\frac{4\Delta^{2}}{\hslash^{2}\left(\Delta^{2}+\epsilon^{2}\right)^{2}}+\frac{\beta^{2}\epsilon^{2}}{\epsilon^{2}+\Delta^{2}}\left[1-\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]\right\}d\epsilon^{2}\text{.} (182)

The metric tensor gij(Sjo¨qvist)(βϵ)g_{ij}^{\left(\mathrm{Sj\ddot{o}qvist}\right)}\left(\beta\text{, }\epsilon\right) from Eq. (182) is given by

gij(Sjo¨qvist)(βϵ)=216[1tanh2(βϵ2+Δ22)](ϵ2+Δ2βϵβϵβ2ϵ2ϵ2+Δ2+4Δ22(Δ2+ϵ2)211tanh2(βϵ2+Δ22)),g_{ij}^{\left(\mathrm{Sj\ddot{o}qvist}\right)}\left(\beta\text{, }\epsilon\right)=\frac{\hslash^{2}}{16}\left[1-\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]\left(\begin{array}[c]{cc}\epsilon^{2}+\Delta^{2}&\beta\epsilon\\ \beta\epsilon&\frac{\beta^{2}\epsilon^{2}}{\epsilon^{2}+\Delta^{2}}+\frac{4\Delta^{2}}{\hslash^{2}\left(\Delta^{2}+\epsilon^{2}\right)^{2}}\frac{1}{1-\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)}\end{array}\right)\text{,} (183)

with 1i1\leq i, j2j\leq 2. The derivation of Eqs. (182) and (183) completes our calculation of the Sjöqvist metric structure for superconducting flux qubits.

VI Considerations from the comparative analysis

In this section, we discuss the outcomes of the comparative analysis carried out in Section V concerning the Bures and Sjöqvist metrics for spin qubits and superconducting flux qubits Hamiltonian models presented in Section IV.

VI.1 The asymptotic limit of β\beta approaching infinity

We begin by discussing the asymptotic limit of β\beta approaching infinity. In  the case of a spin qubit with Hamiltonian HSQ(ω){}_{\mathrm{SQ}}\left(\omega\right) in Eq. (91), the density matrix ρSQ(βω)\rho_{\mathrm{SQ}}\left(\beta\text{, }\omega\right) in Eq. (107) approaches ρSQβ(ω)=def|11|\rho_{\mathrm{SQ}}^{\beta\rightarrow\infty}\left(\omega\right)\overset{\text{def}}{=}\left|1\right\rangle\left\langle 1\right| as β\beta\rightarrow\infty. Observe that |1\left|1\right\rangle denotes here the ground state, the state of lowest energy ω/2-\hslash\omega/2. Since ρSQβ(ω)\rho_{\mathrm{SQ}}^{\beta\rightarrow\infty}\left(\omega\right) is a constant in ω\omega, the Bures and Sjöqvist metrics in Eqs. (126) and (136), respectively, both vanish in this limiting scenario. In this regard, the case of the superconducting flux qubit specified by the Hamiltonian HSFQ(Δϵ){}_{\mathrm{SFQ}}\left(\Delta\text{, }\epsilon\right) in Eq. (92) is more interesting. Indeed, in this case the density matrix ρSFQ(βϵ)\rho_{\mathrm{SFQ}}\left(\beta\text{, }\epsilon\right) in Eq. (149) approaches ρSFQβ(ϵ)\rho_{\mathrm{SFQ}}^{\beta\rightarrow\infty}\left(\epsilon\right) when β\beta approaches infinity. The quantity ρSFQβ(ϵ)\rho_{\mathrm{SFQ}}^{\beta\rightarrow\infty}\left(\epsilon\right) represents a pure (ground) state of lowest energy (/2)Δ2+ϵ2\left(-\hslash/2\right)\sqrt{\Delta^{2}+\epsilon^{2}} and is given by

ρSFQβ(ϵ)=12(1+ϵϵ2+Δ2Δϵ2+Δ2Δϵ2+Δ21ϵϵ2+Δ2),\rho_{\mathrm{SFQ}}^{\beta\rightarrow\infty}\left(\epsilon\right)=\frac{1}{2}\left(\begin{array}[c]{cc}1+\frac{\epsilon}{\sqrt{\epsilon^{2}+\Delta^{2}}}&\frac{\Delta}{\sqrt{\epsilon^{2}+\Delta^{2}}}\\ \frac{\Delta}{\sqrt{\epsilon^{2}+\Delta^{2}}}&1-\frac{\epsilon}{\sqrt{\epsilon^{2}+\Delta^{2}}}\end{array}\right)\text{,} (184)

with ρSFQβ(ϵ)=(ρSFQβ(ϵ))2\rho_{\mathrm{SFQ}}^{\beta\rightarrow\infty}\left(\epsilon\right)=(\rho_{\mathrm{SFQ}}^{\beta\rightarrow\infty}\left(\epsilon\right))^{2} and tr(ρSFQβ(ϵ))=1\left(\rho_{\mathrm{SFQ}}^{\beta\rightarrow\infty}\left(\epsilon\right)\right)=1. Furthermore, when β\beta\rightarrow\infty, the Bures and Sjöqvist metrics in Eqs. (173) and (182), respectively, reduce to the same expression

dsBures2β14Δ2(Δ2+ϵ2)2dϵ2, and dsSjo¨qvist2β14Δ2(Δ2+ϵ2)2dϵ2.ds_{\mathrm{Bures}}^{2}\overset{\beta\rightarrow\infty}{\rightarrow}\frac{1}{4}\frac{\Delta^{2}}{\left(\Delta^{2}+\epsilon^{2}\right)^{2}}d\epsilon^{2}\text{, and }ds_{\mathrm{Sj\ddot{o}qvist}}^{2}\overset{\beta\rightarrow\infty}{\rightarrow}\frac{1}{4}\frac{\Delta^{2}}{\left(\Delta^{2}+\epsilon^{2}\right)^{2}}d\epsilon^{2}\text{.} (185)

The limiting expressions assumed by the Bures and Sjöqvist metrics in Eq. (185) are, modulo an unimportant constant factor, the Fubini-Study metric dsFS2ds_{\mathrm{FS}}^{2} for pure states. Indeed, we have

dsFS2=deftr[(ρSFQβ(ϵ)ϵ)2]dϵ2=12Δ2(Δ2+ϵ2)2dϵ2.ds_{\mathrm{FS}}^{2}\overset{\text{def}}{=}\mathrm{tr}\left[\left(\frac{\partial\rho_{\mathrm{SFQ}}^{\beta\rightarrow\infty}\left(\epsilon\right)}{\partial\epsilon}\right)^{2}\right]d\epsilon^{2}=\frac{1}{2}\frac{\Delta^{2}}{\left(\Delta^{2}+\epsilon^{2}\right)^{2}}d\epsilon^{2}\text{.} (186)

In the next subsection, we discuss the discrepancy in the Bures (Eqs. (126) and (173)) and Sjöqvist (Eqs. (136) and (182)) metrics emerging from the different nature of the nonclassical contributions in the two metrics.

VI.2 The metrics discrepancy

Refer to caption
Figure 2: Plots of the metric discrepancy Δgϵϵnc(βϵ)\Delta g_{\epsilon\epsilon}^{\mathrm{nc}}\left(\beta\text{, }\epsilon\right) in Eq. (188) versus β\beta for ϵ=1\epsilon=1 (dashed), ϵ=1.25\epsilon=1.25 (thin solid), and ϵ=1.5\epsilon=1.5 (thick solid). For simplicity, we set =1\hslash=1 and Δ=1\Delta=1 in each plot. Observe that Δgϵϵnc(βϵ)\Delta g_{\epsilon\epsilon}^{\mathrm{nc}}\left(\beta\text{, }\epsilon\right) is nonvanishing because of the difference between the nonclassical parts gϵϵnc(βϵ)g_{\epsilon\epsilon}^{\mathrm{nc}}\left(\beta\text{, }\epsilon\right) of the metric tensor component gϵϵ(βϵ)=defgϵϵc(βϵ)+gϵϵnc(βϵ)g_{\epsilon\epsilon}\left(\beta\text{, }\epsilon\right)\overset{\text{def}}{=}g_{\epsilon\epsilon}^{\mathrm{c}}\left(\beta\text{, }\epsilon\right)+g_{\epsilon\epsilon}^{\mathrm{nc}}\left(\beta\text{, }\epsilon\right) in the Bures and Sjöqvist metric scenarios. Finally, note that the metric discrepancy vanishes in the zero-temperature limit, i.e., in the limit of β\beta approaching infinity.

We begin by noting that in the case of the spin qubit Hamiltonian model in Eq. (91), there is no discrepancy since the Bures and the Sjöqvist metrics in Eqs. (126) and (136), respectively, coincide. Indeed, in this case, both metrics reduce to the classical Fisher-Rao information metric. The nonclassical/quantum terms vanish in both metrics due to the commutativity of ρSQ(βω)\rho_{\mathrm{SQ}}\left(\beta\text{, }\omega\right) and (ρSQ+dρSQ)(βω)\left(\rho_{\mathrm{SQ}}+d\rho_{\mathrm{SQ}}\right)\left(\beta\text{, }\omega\right), with ρSQ(βω)\rho_{\mathrm{SQ}}\left(\beta\text{, }\omega\right) in Eq. (107). In the case of the superconducting flux qubit Hamiltonian model in Eq. (92), instead, the nonclassical/quantum terms vanish in neither the Bures nor the Sjöqvist metrics due to the non-commutativity of ρSFQ(βϵ)\rho_{\mathrm{SFQ}}\left(\beta\text{, }\epsilon\right) and (ρSFQ+dρSFQ)(βϵ)\left(\rho_{\mathrm{SFQ}}+d\rho_{\mathrm{SFQ}}\right)\left(\beta\text{, }\epsilon\right), with ρSQ(βϵ)\rho_{\mathrm{SQ}}\left(\beta\text{, }\epsilon\right) in Eq. (149). However, these nonclassical contributions differ in the two metrics and this leads to a discrepancy in the Bures and Sjöqvist metrics in Eqs. (173) and (182), respectively. More specifically, we have

dsSjo¨qvist2dsBures2=Δgϵϵnc(βϵ)dϵ20,ds_{\mathrm{Sj\ddot{o}qvist}}^{2}-ds_{\mathrm{Bures}}^{2}=\Delta g_{\epsilon\epsilon}^{\mathrm{nc}}\left(\beta\text{, }\epsilon\right)d\epsilon^{2}\geq 0\text{,} (187)

for any β\beta and ϵ\epsilon. Note that Δgϵϵnc(βϵ)=defgϵϵnc, Sjo¨qvist(βϵ)gϵϵnc, Bures(βϵ)\Delta g_{\epsilon\epsilon}^{\mathrm{nc}}\left(\beta\text{, }\epsilon\right)\overset{\text{def}}{=}g_{\epsilon\epsilon}^{\mathrm{nc,}\text{ }\mathrm{Sj\ddot{o}qvist}}\left(\beta\text{, }\epsilon\right)-g_{\epsilon\epsilon}^{\mathrm{nc,}\text{ }\mathrm{Bures}}\left(\beta\text{, }\epsilon\right) is the difference between the nonclassical (nc) contributions in the metric components gϵϵ(βϵ)=defgϵϵc(βϵ)+gϵϵnc(βϵ)g_{\epsilon\epsilon}\left(\beta\text{, }\epsilon\right)\overset{\text{def}}{=}g_{\epsilon\epsilon}^{\mathrm{c}}\left(\beta\text{, }\epsilon\right)+g_{\epsilon\epsilon}^{\mathrm{nc}}\left(\beta\text{, }\epsilon\right) and is given by

Δgϵϵnc(βϵ)=def14Δ2(Δ2+ϵ2)2[1tanh2(βϵ2+Δ22)],\Delta g_{\epsilon\epsilon}^{\mathrm{nc}}\left(\beta\text{, }\epsilon\right)\overset{\text{def}}{=}\frac{1}{4}\frac{\Delta^{2}}{\left(\Delta^{2}+\epsilon^{2}\right)^{2}}\left[1-\tanh^{2}\left(\beta\hslash\frac{\sqrt{\epsilon^{2}+\Delta^{2}}}{2}\right)\right]\text{,} (188)

with 00\leq tanh2(x)1\tanh^{2}\left(x\right)\leq 1 for any xx\in\mathbb{R}. To be crystal clear, it is useful to view the metric tensor gij(βϵ)g_{ij}\left(\beta\text{, }\epsilon\right) with 1i1\leq i, j2j\leq 2 (i.e., θ1=defβ\theta^{1}\overset{\text{def}}{=}\beta and θ2=defϵ\theta^{2}\overset{\text{def}}{=}\epsilon) recast as

gij(βϵ)=(gββ(βϵ)gβϵ(βϵ)gϵβ(βϵ)gϵϵc(βϵ)+gϵϵnc(βϵ)).g_{ij}\left(\beta\text{, }\epsilon\right)=\left(\begin{array}[c]{cc}g_{\beta\beta}\left(\beta\text{, }\epsilon\right)&g_{\beta\epsilon}\left(\beta\text{, }\epsilon\right)\\ g_{\epsilon\beta}\left(\beta\text{, }\epsilon\right)&g_{\epsilon\epsilon}^{\mathrm{c}}\left(\beta\text{, }\epsilon\right)+g_{\epsilon\epsilon}^{\mathrm{nc}}\left(\beta\text{, }\epsilon\right)\end{array}\right)\text{.} (189)

The discrepancy between the Bures and Sjöqvist metrics arises only because gϵϵnc, Sjo¨qvist(βϵ)gϵϵnc, Bures(βϵ)g_{\epsilon\epsilon}^{\mathrm{nc,}\text{ }\mathrm{Sj\ddot{o}qvist}}\left(\beta\text{, }\epsilon\right)\neq g_{\epsilon\epsilon}^{\mathrm{nc,}\text{ }\mathrm{Bures}}\left(\beta\text{, }\epsilon\right). However, the metric discrepancy Δgϵϵnc(βϵ)\Delta g_{\epsilon\epsilon}^{\mathrm{nc}}\left(\beta\text{, }\epsilon\right) in Eq. (188) vanishes in the asymptotic limit of β\beta approaching infinity. In Fig. 11, we plot the discrepancy between the Bures and the Sjöqvist metrics for the superconducting flux qubit Hamiltonian model. In Table I, instead, we summarize the links between the Bures and the Sjöqvist metrics.

Description of quantum states Quantum states Bures metric Sjöqvist metric
Pure ρ=ρ2\rho=\rho^{2} Fubini-Study metric Fubini-Study metric
Mixed, classical scenario ρρ2\rho\neq\rho^{2}, [ρρ+dρ]=0\left[\rho\text{, }\rho+d\rho\right]=0 Fisher-Rao metric Fisher-Rao metric
Mixed, nonclassical scenario ρρ2\rho\neq\rho^{2}, [ρρ+dρ]0\left[\rho\text{, }\rho+d\rho\right]\neq 0 dsBures2dsSjo¨qvist2ds_{\mathrm{Bures}}^{2}\neq ds_{\mathrm{Sj\ddot{o}qvist}}^{2} dsSjo¨qvist2dsBures2ds_{\mathrm{Sj\ddot{o}qvist}}^{2}\neq ds_{\mathrm{Bures}}^{2}
Table 1: Relation between the Bures and the Sjöqvist metrics. The Bures and the Sjöqvist metrics are identical when considering pure quantum states (ρ=ρ2)\left(\rho=\rho^{2}\right) or mixed quantum states (ρρ2)\left(\rho\neq\rho^{2}\right) for which the non-commutative probabilistic structure underlying quantum theory is not visible (i.e., in the classical scenario with [ρρ+dρ]=0\left[\rho\text{, }\rho+d\rho\right]=0). In particular, in the former and latter cases, they becomes the Fubini-Study and the Fisher-Rao information metrics, respectively. However, the Bures and the Sjöqvist metrics are generally distinct when considering mixed quantum states (ρρ2)\left(\rho\neq\rho^{2}\right) for which the non-commutative probabilistic structure of quantum mechanics is visible (i.e., in the nonclassical scenario with [ρρ+dρ]0\left[\rho\text{, }\rho+d\rho\right]\neq 0).

VII Conclusive Remarks

In this paper, building on our recent scientific effort in Ref. cafaroprd22 , we presented an explicit analysis of the Bures and Sjöqvist metrics over the manifolds of thermal states for the spin qubit (Eq. (91)) and the superconducting flux qubit Hamiltonian (Eq. (92)) models. We observed that while both metrics (Eqs. (126) and (136)) reduce to the Fubini-Study metric in the (zero-temperature) asymptotic limiting case of the inverse temperature β\beta approaching infinity for both Hamiltonian models, the two metrics are generally different. We observed this different behavior in the case of the superconducting flux Hamiltonian model. In general, we note that the two metrics (Eqs. (173) and (182)) seem to differ when nonclassical behavior is present since they quantify noncommutativity of neighboring mixed quantum states in different manners (Eqs. (187) and (188)). In summary, we reach (see Table I) the conclusion that for pure quantum states (ρ=ρ2)\left(\rho=\rho^{2}\right) and for mixed quantum states (ρρ2)\left(\rho\neq\rho^{2}\right) for which the non-commutative probabilistic structure underlying quantum theory is not visible (i.e., in the classical scenario with [ρρ+dρ]=0\left[\rho\text{, }\rho+d\rho\right]=0), the Bures and the Sjöqvist metrics are the same (Eqs. (127) and (137)). Indeed, in the former and latter cases, they reduce to the Fubini-Study and Fisher-Rao information metrics, respectively. Instead, when investigating mixed quantum states (ρρ2)\left(\rho\neq\rho^{2}\right) for which the non-commutative probabilistic structure of quantum mechanics is visible (i.e., in the non-classical scenario with [ρρ+dρ]0\left[\rho\text{, }\rho+d\rho\right]\neq 0), the Bures and the Sjöqvist metrics exhibit a different behavior (Eqs. (138) and (139); Eqs. (174) and (183)).

Our main conclusions can be outlined as follows:

  1. [i]

    We presented an explicit derivation of Bures metric for arbitrary density matrices in Hübner’s form (Eq. (24)) and in Zanardi’s form (Eq. (37)). Moreover, we presented a clear derivation of Zanardi’s form of the Bures metric suitable for the special class of thermal states (Eq. (73)). Finally, we reported an explicit derivation of the Sjöqvist metric for nondegenerate density matrices (Eq. (90)).

  2. [ii]

    Using our explicit derivations outlined in [i], we performed detailed analytical calculations yielding the expressions of the Bures (Eqs. (126) and (173)) and Sjöqvist (Eqs. (136) and (182)) metrics on manifolds of thermal states (Eqs. (107) and (149)) that correspond to a spin qubit (Eq. (91)) and a superconducting flux qubit (Eq. (92)) Hamiltonian models.

  3. [iii]

    In the absence of nonclassical features in which the neighboring density matrices ρ\rho and dρd\rho commute, the Bures and the Sjöqvist metrics lead to and identical metric expression exemplified by the classical Fisher-Rao metric tensor. We have explicitly verified this similarity in the case of a manifold of thermal states for spin qubits in the presence of a constant magnetic field along the quantization zz-axis.

  4. [iv]

    In general, the Bures and the Sjöqvist metrics are expected to yield different expressions. Indeed, the Bures and Sjöqvist metrics seem to quantify the noncommutativity of neighboring mixed states ρ\rho and dρd\rho in different manners, in general. We have explicitly verified this difference in the case of a manifold of thermal states for superconducting flux qubits (see Fig. 22).

  5. [v]

    In the asymptotic limit of β\beta\rightarrow\infty, both Bures and Sjöqvist metric tensors reduce to the same limiting value (Eq. (185)) specified by, modulo an unimportant constant factor, the Fubini-Study metric tensor (Eq. (186)) for the zero-temperature pure states.

  6. [vi]

    In the superconducting flux qubit Hamiltonian model considered here, we observe that the difference dsSjo¨qvist2dsBures2ds_{\mathrm{Sj\ddot{o}qvist}}^{2}-ds_{\mathrm{Bures}}^{2} is a positive quantity that depends solely on the difference between the nonclassical nature of the metric tensor component gϵϵ(βϵ)g_{\epsilon\epsilon}\left(\beta\text{, }\epsilon\right). Indeed, we have shown that dsSjo¨qvist2dsBures2=Δgϵϵnc(βϵ)dϵ20ds_{\mathrm{Sj\ddot{o}qvist}}^{2}-ds_{\mathrm{Bures}}^{2}=\Delta g_{\epsilon\epsilon}^{\mathrm{nc}}\left(\beta\text{, }\epsilon\right)d\epsilon^{2}\geq 0 (Eqs. (187) and (188)).

  7. [vii]

    The existence of nonclassical contributions in the Bures and Sjöqvist metrics is related to the presence of non-vanishing quadratic terms like |n|hH|m|2\left|\left\langle n\left|\partial_{h}\mathrm{H}\right|m\right\rangle\right|^{2} and |dn|n|2\left|\left\langle dn\left|n\right.\right\rangle\right|^{2}, respectively. The former term is related to modulus squared of suitable quantum overlaps defined in terms of parametric variations in the Hamiltonian of the system. The latter term, instead, is specified by the modulus squared of suitable quantum overlaps characterized by parametric variations of the eigenstates of the Hamiltonian of the system. It is not unreasonable to expect a formal connection between these two types of terms causing the noncommutativity between ρ\rho and ρ+dρ\rho+d\rho (see, for instance, Eq. (15.30) in Ref. karol06 ) and find a deeper relation between the Bures and Sjöqvist metrics for the class of thermal quantum states. Indeed, for a more quantitative discussion on the link between these two terms, see Ref. alsing23 .

  8. [viii]

    The differential dρ(βh)=defβρdβ+hρdhd\rho\left(\beta\text{, }h\right)\overset{\text{def}}{=}\partial_{\beta}\rho d\beta+\partial_{h}\rho dh depends both on eigenvalues and eigenvectors parametric variations. However, the noncommutativity between ρ\rho and dρd\rho is related to that part of dρd\rho that emerges from the eigenvectors parametric variations. These changes, in turn, can be related to the existence of a nonvanishing commutator between the Hamiltonian of the system and the density matrix specifying the thermal state. Indeed, in the two main examples studied in this paper, we have [HSQ(ω)ρSQ(βω)]=0\left[\mathrm{H}_{\mathrm{SQ}}\left(\omega\right)\text{, }\rho_{\mathrm{SQ}}\left(\beta\text{, }\omega\right)\right]=0 and [HSFQ(Δϵ)ρSFQ(βϵ)]0\left[\mathrm{H}_{\mathrm{SFQ}}\left(\Delta\text{, }\epsilon\right)\text{, }\rho_{\mathrm{SFQ}}\left(\beta\text{, }\epsilon\right)\right]\neq 0, respectively. In the former case, unlike the latter case, there is no contribution to dρd\rho arising from a variation in the eigenvectors of the Hamiltonian.

For the set of pure states, the scenario is rather unambiguous. The Fubini–Study metric represents the only natural option for a measure that characterizes “random states”. Alternatively, for mixed-state density matrices, the geometry of the state space is more complicated karol06 ; brody19 . There is a collection of distinct metrics that can be used, each of them with different physical inspiration, benefits and disadvantages that can rest on the peculiar application one  might be interested in examining. Specifically, both simple and complicated geometric quantities (i.e., for instance, path, path length, volume, curvature, and complexity) seem to depend on the measure selected on the space of mixed states that specify the quantum system being investigated karol99 ; cafaroprd22 . Therefore, our work in this paper can be particularly important in offering an explicit comparative study between the (emerging) Sjöqvist interferometric geometry and the (established) Bures geometry for mixed quantum states. Gladly, the importance of the existence of this kind of comparative investigation was lately emphasized in Refs. mera22 and cafaroprd22 too.

From a mathematics standpoint, it would be interesting to formally prove (or, alternatively, disprove with an explicit counterexample) the monotonicity petz96a ; petz99 of the Sjöqvist metric in an arbitrarily finite-dimensional space of mixed quantum states. From a physics perspective that relies on empirical evidence, instead, it would be very important to understand what the physical significance of employing either metric is mera22 ; cafaroprd22 .

In conclusion, despite its relatively limited scope, we hope this work will inspire either young or senior scientists to pursue ways to deepen our understanding (both mathematically and physically) of this fascinating connection among information geometry, statistical physics, and quantum mechanics cafaropre20 ; gassner21 ; hasegawa21 ; miller20 ; cc ; saito20 ; ito20 .

Acknowledgements.
C.C. is grateful to the United States Air Force Research Laboratory (AFRL) Summer Faculty Fellowship Program for providing support for this work. C. C. acknowledges helpful discussions with Orlando Luongo, Cosmo Lupo, Stefano Mancini, and Hernando Quevedo. P.M.A. acknowledges support from the Air Force Office of Scientific Research (AFOSR). Any opinions, findings and conclusions or recommendations expressed in this material are those of the author(s) and do not necessarily reflect the views of the Air Force Research Laboratory (AFRL).

References

  • (1) M. Pettini, Geometry and Topology in Hamiltonian Dynamics and Statistical Mechanics, Springer-Verlag New York (2007).
  • (2) I. Bengtsson and K. Zyczkowski, Geometry of Quantum States, Cambridge University Press (2006).
  • (3) C. Cafaro and P. M. Alsing, Complexity of pure and mixed qubit geodesic paths on curved manifolds, Phys. Rev. D106, 096004 (2022).
  • (4) C. Cafaro, S. Ray, and P. M. Alsing, Complexity and efficiency of minimum entropy production probability paths from quantum dynamical evolutions, Phys. Rev. E105, 034143 (2022).
  • (5) C. Cafaro, O. Luongo, S. Mancini, and H. Quevedo, Thermodynamic length, geometric efficiency and Legendre invariance, Physica A590, 126740 (2022).
  • (6) C. Cafaro, The Information Geometry of Chaos, PhD Thesis, State University of New York at Albany, Albany-NY, USA (2008). Available online at arXiv: math-ph/1601.07935 (2016).
  • (7) C. Cafaro, Works on an information geometrodynamical approach to chaos, Chaos, Solitons & Fractals 41, 886 (2009).
  • (8) D. Felice, C. Cafaro, and S. Mancini, Information geometric methods for complexity, Chaos 28, 032101 (2018).
  • (9) S. Amari and H. Nagaoka, Methods of Information Geometry, Oxford University Press (2000).
  • (10) P. Zanardi, P. Giorda, and M. Cozzini, Information-theoretic differential geometry of quantum phase transitions, Phys. Rev. Lett. 99, 100603 (2007).
  • (11) P. Zanardi, L. Campos Venuti, and P. Giorda, Bures metric over thermal manifolds and quantum criticality, Phys. Rev. A76, 062318 (2007).
  • (12) P. Pessoa and C. Cafaro, Information geometry for Fermi-Dirac and Bose-Einstein quantum statistics, Physica A576, 126061 (2021).
  • (13) H. Silva, B. Mera, and N. Paunkovic, Interferometric geometry from symmetry-broken Uhlmann gauge group with applications to topological phase transitions, Phys. Rev. B103, 085127 (2021).
  • (14) H. V. da Silva, Quantum information geometry and applications, MS Thesis in Engineering Physics, IT Lisboa (2021).
  • (15) B. Mera, N. Paunkovic, S. T. Amin, and V. R. Vieira, Information geometry of quantum critical submanifolds: Relevant, marginal, and irrelevant operators, Phys. Rev. B106, 155101 (2022).
  • (16) D. Bures, An extension of Kakutani’s theorem on infinite product measures to the tensor product of semifinite ω\omega^{\ast}-algebras, Trans. Amer. Math. Soc. 135, 199 (1969).
  • (17) A. Uhlmann, The “transition probability” in the state space of a \ast-algebra, Rep. Math. Phys. 9, 273 (1976).
  • (18) M. Hübner, Explicit computation of the Bures distance for density matrices, Phys. Lett. A163, 239 (1992).
  • (19) E. Sjöqvist, Geometry along evolution of mixed quantum states, Phys. Rev. Research 2, 013344 (2020).
  • (20) F. D. M. Haldane, Model for a quantum Hall effect without Landau levels: Condensed-matter realization of the “Parity Anomaly”, Phys. Rev. Lett. 61, 2015 (1988).
  • (21) J. P. Provost and G. Vallee, Riemannian structure on manifolds of quantum states, Commun. Math. Phys. 76, 289 (1980).
  • (22) W. K. Wootters, Statistical distance and Hilbert space, Phys. Rev. D23, 357 (1981).
  • (23) S. L. Braunstein and C. M. Caves, Statistical distance and the geometry of quantum states, Phys. Rev. Lett. 72, 3439 (1994).
  • (24) Y. Aharonov and J. Anandan, Phase change during a cyclic quantum evolution, Phys. Rev. Lett. 58, 1593 (1987).
  • (25) J. J. Sakurai and J. Napolitano, Modern Quantum Mechanics, Cambridge University Press (2017).
  • (26) J. Clarke and F. K. Wilhelm, Superconducting quantum bits, Nature 453, 1031 (2008).
  • (27) M. H. Devoret and R. J. Schoelkopf, Superconducting circuits for quantum information: An outlook, Science 339, 1169 (2013).
  • (28) I. Chirescu, Y. Nakamura, C. J. P. M. Harmans, and J. E. Mooij, Coherent quantum dynamics of a superconducting flux qubit, Science 299, 1869  (2003).
  • (29) A. O. Niskanen, Y. Nakamura, and J. P. Pekola, Information entropic superconducting microcooler, Phys. Rev. B76, 174523 (2007).
  • (30) F. G. Paauw, A. Fedorov, C. J. P. M. Harmans, and J. E. Mooij, Tuning the gap of a superconducting flux qubit, Phy. Rev. Lett. 102, 090501 (2009).
  • (31) B. Karimi and J. P. Pekola, Otto refrigerator based on a superconducting qubit: Classical and quantum performance, Phys. Rev. B94, 184503 (2016).
  • (32) K. Huang, Statistical Mechanics, John Wiley & Sons, Inc. (1987).
  • (33) F. Strocchi, Thermal states, Lect. Notes Phys. 732, 139 (2008).
  • (34) P. M. Alsing, C. Cafaro, O. Luongo, C. Lupo, S. Mancini, and H. Quevedo, Comparing metrics for mixed quantum states: Sjöqvist and Bures, to be published (2023).
  • (35) D. C. Brody and B. Longstaff, Evolution speed of open quantum dynamics, Phys. Rev. Research 1, 033127  (2019).
  • (36) K. Zyczkowski, Volume of the set of separable states. II, Phys, Rev. A60, 3496 (1999).
  • (37) D. Petz, Monotone metrics on matrix spaces, Lin. Algebra Appl. 244, 81 (1996).
  • (38) D. Petz and Cs. Sudar, Extending the Fisher metric to density matrices, in Geometry in Present Days Science, eds. O. E. Barndorff-Nielsen and E. B. Vendel, World Scientific, pp. 21-34 (1999).
  • (39) C. Cafaro and P. M. Alsing, Information geometry aspects of minimum entropy production paths from quantum mechanical evolutions, Phys. Rev. E101, 022110 (2020).
  • (40) S. Gassner, C. Cafaro, S. A. Ali, and P. M. Alsing, Information geometric aspects of probability paths with minimum entropy production for quantum state evolution, Int. J. Geom. Meth. Mod. Phys. 18, 2150127 (2021).
  • (41) T. Van Vu and Y. Hasegawa, Geometrical bounds of the irreversibility in Markovian systems, Phys. Rev. Lett. 126, 010601 (2021).
  • (42) H. J. D. Miller and M. Mehboudi, Geometry of work fluctuations versus efficiency in microscopic thermal machines, Phys. Rev. Lett. 125, 260602 (2020).
  • (43) C. Cafaro, D. Felice, and P. M. Alsing, Quantum Groverian geodesic paths with gravitational and thermal analogies, Eur. Phys. J. Plus 135, 900 (2020).
  • (44) K. Brandner and K. Saito, Thermodynamic geometry of microscopic heat engines, Phys. Rev. Lett. 124, 040602 (2020).
  • (45) S. Ito, M. Oizumi, and S. Amari, Unified framework for the entropy production and the stochastic interaction based on information geometry, Phys. Rev. Research 2, 033048 (2020).