Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging...

92
Thermalization in Quantum Information Geometry University of Amsterdam, Institute of Physics Submitted in partial fulfillment of the requirements for the degree of Master of Science Daniel Chernowitz August 2017

Transcript of Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging...

Page 1: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Thermalization in Quantum Information

Geometry

University of Amsterdam, Institute of Physics

Submitted in partial fulfillment of the requirements for the degree

of Master of Science

Daniel Chernowitz

August 2017

Page 2: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Abstract

In this work we approach Thermalization of discrete quantum systems from ageometrical vantage point. The aim is statistical results showing most Hamilto-nians thermalize their subsystems. Fidelity, a concept in quantum informationtheory, relates distinguishability by measurements between states. We have nu-merically generated Hamiltonians according to the Haar measure and evolvedthe pure full system unitarily. We averaged the fidelity of subsystems betweeninitial times and later times, finding them typically distinguishable. Conversely,the fidelity between two later times is seen to approach one for almost all sys-tems. The inability to distinguish later times points to a steady state: themaximally mixed state. Moving on, we used Weingarten functions to inte-grate subsystem density matrices during evolution, over all possible Hamilto-nian eigenvectors in the unitary group. This resulted in the mean subsystemdensity matrix, in which the initial conditions are seen to compete with themaximally mixed state over time. We observe a transition from a ’known’ to’unknown’ phase after a time inversely proportional to the standard deviationin energies. However, the initial information is never completely lost. Also,polynomial functions of density matrices at any time can be integrated: weconsidered the variance over the unitary group and purity. In all cases, thesubsystem mixing was seen to increase with bath size. Finally, we consideredtransformation properties of the Quantum Geometric Tensor under abelian andnon-abelian gauge symmetries. The Bures metric, a natural generalization ofthe QGT, was found not to be gauge invariant in a degenerate spectrum.

1

Page 3: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Acknowledgements

It goes without saying that I could not have produced this work on my own. Mymain lifeline has been my preceptor, Vladimir Gritsev. This man has a mindfor physics I can never hope to rival. If every theory is a phone line, professorGritsev is the switchboard operator. No matter what problem I might have,he knows ten relevant articles. The only downside is that it can be hard tofigure out which ideas are bad (there were many), if he can find merit in any ofthem. All in all, it has been my honor to stand so close to the source. Spasibo,Vladimir.

Besides this, I would like to extend gratitude to Maris Ozols. If not for ourchance encounter, and him taking time to listen to my mathematical troubles,chapter 4 would have never come about. A suggestion can go a long way, andhe showed me what was possible and what wasn’t. Thank you as well.

2

Page 4: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Contents

1 Introduction to Thermalization 51.1 Isolated Quantum Systems . . . . . . . . . . . . . . . . . . . . . 51.2 Quantum Geometry . . . . . . . . . . . . . . . . . . . . . . . . . 7

2 Theory and Setup 82.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82.2 Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2.2.1 Formal Problem . . . . . . . . . . . . . . . . . . . . . . . 102.3 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.3.1 Density Matrices . . . . . . . . . . . . . . . . . . . . . . . 122.3.2 Tracing and Partial Tracing . . . . . . . . . . . . . . . . . 12

2.4 Measures of Mixing . . . . . . . . . . . . . . . . . . . . . . . . . . 152.4.1 Entanglement Entropy . . . . . . . . . . . . . . . . . . . . 152.4.2 Purity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2.5 Quantum Distinguishability . . . . . . . . . . . . . . . . . . . . . 162.6 Random Matrix Theory . . . . . . . . . . . . . . . . . . . . . . . 18

2.6.1 Generating Unitary Matrices . . . . . . . . . . . . . . . . 212.7 Rudimentary Group Theory . . . . . . . . . . . . . . . . . . . . . 22

3 Numerical Fidelity Simulations 243.1 Example Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . 253.2 Statistics of Fidelity . . . . . . . . . . . . . . . . . . . . . . . . . 28

3.2.1 Experiment Parameters . . . . . . . . . . . . . . . . . . . 293.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

4 Analytic Mixed State Averages 364.1 Analytic statistics of local density matrices . . . . . . . . . . . . 36

4.1.1 Average Reduced Density Matrix . . . . . . . . . . . . . . 374.1.2 Variance of the Reduced Density Matrix . . . . . . . . . . 41

4.2 Mean Subsystem Dynamical Purity . . . . . . . . . . . . . . . . . 50

5 Quantum Differential Geometry 555.1 Manifold of Quantum States . . . . . . . . . . . . . . . . . . . . . 55

5.1.1 Nondegenerate Case . . . . . . . . . . . . . . . . . . . . . 55

3

Page 5: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

5.1.2 Degenerate Case . . . . . . . . . . . . . . . . . . . . . . . 585.2 Linear Algebraic Formulation . . . . . . . . . . . . . . . . . . . . 60

5.2.1 Nondegenerate case . . . . . . . . . . . . . . . . . . . . . 605.2.2 Degenerate Case . . . . . . . . . . . . . . . . . . . . . . . 62

5.3 Curvature Operator . . . . . . . . . . . . . . . . . . . . . . . . . 645.4 Mixed States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

5.4.1 Bures Metric . . . . . . . . . . . . . . . . . . . . . . . . . 675.4.2 Gauge Invariance of the Bures Metric . . . . . . . . . . . 69

6 Conclusions and Outlook 736.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 736.2 Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

A Fidelity Plots 76

B Python Code 80

4

Page 6: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Chapter 1

Introduction toThermalization

Fill a glass of boiling water, and leave it alone in a room. After a while, thebrightest minds cannot distinguish how exactly it was poured. It will cool toroom temperature, and it might as well have been discharged freezing. There isno way to tell, because, after a while, all water looks the same. This, in essence,is the phenomenon called thermalization. It borrows its name from the thermalequilibrium attained with its surroundings. In nature thermalization is therule and there are few exceptions. It appears any interplay with surroundingswill thermalize a system. Moreover, in classical statistical mechanics, one ofthe celebrated results is that large collections of particles, even when isolated,maximize entropy and tend towards a thermal state. The same is even trueon the smallest scales. Many quantum systems are seen to decohere, dissipate,and even with vanishingly weak coupling to the environment, thermalize. Butthis poses a fundamental challenge: contrary to the classical theory, quantummechanics is inherently linear, and reversible -at the cost of an exponentiallyaugmented state space. The reversibility means there must be a one-to-onemapping from initial states to later states by any evolution operator. Linearityprecludes processes like classical chaos with a strong nonlinear dependence oflater effects on initial conditions. This means an isolated system should notbe capable of thermalization, or the trend to a single steady state from all ofmultiple, vastly diverse starting states. This begs a resolution.

1.1 Isolated Quantum Systems

One of the first to make this observation, and in 1929 a pioneer into quantumthermalization of isolated systems, was John von Neumann.[1]. His original workproved what is called the Quantum Ergodic Theorem, effectively contending thatany energy eigenstate of macroscopic many-body quantum systems is in a sensethermal. He did this by exploring what is now termed ’normal typicality’: For a

5

Page 7: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

reasonable set of commuting macroscopic observables, the distribution of theseobservables evaluated on a state, restricted to an energy shell, is close to that ofthe micro-canonical ensemble at the same energy. In later years this result washeavily criticized as being ”dynamically vacuous”, and after Von Neumann’sdeath in 1957 largely forgotten.[2]. In the 1970s Michael Berry reinvigoratedthe research, by considering the quantum analogs of classical systems that areregular (integrable) or irregular (ergodic) [3]. He found that this distinctioncarried over to the statistics of the wave function in the quantum case, andconjectured that for many body quantum systems, the distribution of eigenstatemomenta is Gaussian. He expected this to hold for classically chaotic systemsonly. Josh Deutsch continued this reasoning in the 1990s, leading up to Srednickipostulating that excited states of many-body wave functions appear to be planewaves with normally distributed momenta, and formulating what is known asthe Eigenstate Thermalization Hypothesis or ETH [4][5]. This is one of themost frequently cited frameworks to describe why some systems thermalize,and notably, others don’t. The ETH is a statement about observables. Itcontends that for any initial state, any instantaneous observable quantity inthat state will on average evolve irreversibly towards the expectation value inthe micro-canonical ensemble. The observable values then only depend on theglobal nature of the system, and the approximate energy level. The hypothesisis predicated on a set of conditions. These are that in the basis of increasingenergy eigenstates, the observable’s operator elements are close to a diagonalform varying smoothly with energy, up to a margin inversely proportional tothe Hilbert space dimension. Then additionally, the observable will not strayfar from this ensemble value at any later time. It holds for typical randomHamiltonians [6]. Since this original formulation, other have chipped away thecriticism initially imparted to Von Neumann, leading to his ultimate vindicationin 2010 by Goldstein et al [7].

The cases which do not thermalize fall mostly into two categories. They areeither classically integrable (perhaps non-interacting), or live in a low-energyregime and exhibit many-body localization (MBL), manifestly the converse ofplane-wave eigenstates. In order to envision MBL states, consider low-lyingsingle particle energy eigenstates that are strongly spatially local, living oncertain lattice sites, for instance. MBL is said to be observed when, after turningon interactions, the many-body wave functions remain local, in a slightly morediffuse sense. They are however not simply products of local single particlestates. An initial state centered on a certain position will retain traces of itsspatial distribution there for all times [8].

Alternatives to the ETH focus on the states themselves, and are usuallyformulated in terms of the occupation of energy eigenstates in the initial su-perposition. They argue that this occupation must be uncorrelated to theobservable’s matrix element, and thus the expectation value of this operatorconstitutes random sampling from its elements, recovering the microcanonicalensemble. However, for many initial states with specific, out-of-equilibrium ob-servable values, this lack of correlation is questionable [9].

If we do not wish to make reference to temperatures and energies, the former

6

Page 8: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

concept somewhat alien to finite discrete systems, we formally only describeequilibration [10]. In this case, the system moves away from its initial conditionto a steady, perhaps predictable state, which in general need not be given by anyparticular (micro)canonical ensemble. It has already been argued appreciably in[11] that we may expect statistically for a typical quantum system to equilibrateindependent of its microscopic initial conditions. In the next section, we willshift gears, and discuss methods, more than problems.

1.2 Quantum Geometry

There is a new emerging school of thought in the field of quantum condensedmatter and quantum statistical physics. Since the seminal paper by Aharonovand Bohm in 1959, increasingly geometric aspects have been found crucial fora full understanding of quantum mechanics [12][13][14]. In the 1980s, manyinsights were achieved. Provost and Vallee are credited with the idea of con-structing a metric tensor on a manifold of pure quantum ground states, pa-rameterized by some experimental dials [15]. After this, Berry again broke newground by observing his now eponymous geometric phase, explaining anew theaforementioned Aharonov Bohm effect [16]. Wilczek and Zee generalized this tonon-abelian phases [17], employing work from the mathematician Uhlmann [18].This opened the gates to the Bures metric [19][20], the natural augmentationof the tensor of Provost and Vallee to mixed state density matrices. This met-ric is directly related to the fidelity, a quantum-informatic measure of distanceinspired by distinguishability through measurements [21][22]. Currently, manyhave turned to the powerful mathematical techniques of geometry to describethermal ensembles and phenomena such as critical behavior. See [23] by Reza-khani et al. and [24] by Zanardi et al. In the first of this pair of papers, the aimis to imbue quantum states of a system with a Riemannian metric as it evolvesadiabatically in time through Hilbert space. In the second, over the manifoldof thermal Gibbs states at different temperatures. Also nonadiabatic responseshave been studied [25].

Inspired by the elegance and effectiveness in the chronology above, this workwill also make use of geometrical tools, to tackle the problem of thermalizationin isolated discrete quantum systems. We hope to make statistical argumentsevocative of Neumann’s, explained and expanded by Reimann [10]. In the lat-ter’s work, it is not proven that all systems thermalize all observables, as thereare obvious exceptions. However, over the full space of possible macroscopicsystems, one can constrain the proportion of these systems that deviates appre-ciably from microcanonical predictions, and the share of time they do so. Theseproportions turn out to scale inversely with a quantity exponential in systemsize. In other words, just as in the classical case, failure to thermalize is possi-ble, but among ’uniformly sampled systems’ exceedingly unlikely to observe inpractice.

7

Page 9: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Chapter 2

Theory and Setup

As a first step into the actual physics of thermalization, we will elaborate ontheory, formulas and formalisms that are useful later on. The criterion formaterial included in this chapter is that these results are all borrowed, in someregard standard, or basic, whereas the subsequent chapters contain more originalwork. The last two sections (2.6,2.7) are rather disconnected from the narrativeof the first five, but are nonetheless needed at a certain point.

2.1 Motivation

As mentioned in the introduction, quantum mechanical time evolution is uni-tary. This means no information can be lost, because we can always in principlereverse the evolution and obtain the state at any time prior, so long as we knowthe state at present and know the Hamiltonian H. This follows from the factthat any unitary operator has an inverse, in this case generated by −H. Imaginewe have a system, evolving according to:

U(t) := e−iHt (2.1)

An operator translating states in time. In this work we have set ~ = 1. Asa Gedanken experiment, start the system in either of two states, |ψ〉 or |φ〉.Assume | 〈ψ|φ〉 | < 1, then we may think of there being some distance in Hilbertspace H between them. We will later learn that | 〈ψ|φ〉 | is called the fidelitybetween these states, and it is a genuinely geometrical concept. Continuing, it isimmediate that this distance cannot grow over time. We work in the Schrodingerpicture, so time is kept on the states. Then:

|ψt〉 ≡ U(t) |ψ〉 , |φt〉 ≡ U(t) |φ〉 (2.2)

The distance at a later time is given by:

| 〈ψt|φt〉 | = | 〈ψ|U†(t)U(t)|φ〉 | = | 〈ψ|φ〉 | (2.3)

8

Page 10: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

The inner product of Hilbert space vectors is conserved by the action of aHermitian Hamiltonian. In the jargon of classical chaos theory, the Lyapunovexponent is zero [5].

However, in practice quantum mechanical systems do exhibit a kind of ther-malization. For many such systems, the properties at late times seem indepen-dent of most of the specifics of the initial conditions, save perhaps some globalquantities such as total energy [26][1][10][9][11]. The resolution, spearheadedby Von Neumann in the 1920s, lies in a shift of focus away from states, to op-erators. The quantum state offers the best possible description of the system,given that it cannot be fully known due to the Heisenberg uncertainty relation.However, a state is a mathematical object, and an experimenter does not dealwith the state directly. He or she performs measurements on the system, andthese are codified by operators evaluated on the states. Moreover, we deal onlywith those operators we choose to represent our preferred observables, the oneswe can access and in which we are interested. This restriction will allow us tounderstand the mechanism of losing information in quantum mechanics. Themoral will be that the information of course is not lost, it is just very inaccessi-ble for an experimenter with measurement apparatus that live in a certain kindof space, corresponding to operators that are natural only in a certain basis.

2.2 Setup

In order to formalize these claims, we first construct some standard machinery[27][11]. For the rest of this work, we may envision our full quantum systemto be partitioned into two parts. There is the subsystem, S, and the muchlarger environment, or bath, B. The intuition should be that an experimenterhas access to S, and not to B. However, he or she cannot prevent the two frominteracting. Moreover, from the perspective of the full system S+B, the divisionmay be highly arbitrary and unnatural. There is simply some rule designatingcertain degrees of freedom to correspond to S, and the rest to B. Think of a setof coupled quantum spins, of which a subset are selected for analysis. But alsosome specific Fourier modes or particles, or the inhabitants of some submanifoldof real space will do. See figure 2.1 for an illustration.

As both S and B are quantum, when isolated, they are in some quantumstates |ψS〉 ∈ HS and |ψB〉 ∈ HB , respectively. The full coupled Hilbert spaceis the tensor product of the constituent spaces:

H = HS ⊗HB (2.4)

And given any two bases of the constituents, one can immediately constructa basis of the full system:

HS = span(|φSj 〉), HB = span(|φBk 〉), H = span(|φSj 〉 ⊗ |φBk 〉) (2.5)

.

9

Page 11: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Figure 2.1: Division of the degrees of freedom (e.g. particles or spins on alattice) into a subsystem S of nS spins or particles and bath B of nB spins orparticles. A product state in the tensored Hilbert space is also noted.

2.2.1 Formal Problem

Turning now to the isolated subsystem, forget the coupling to the bath mo-mentarily. We take S to start at t = 0 in a well defined state |ψS〉, evolvingaccording to a unitary operator US(t) := e−iHSt = US , we will suppress theargument for now. Say, at time t = 0, we know some property of the system,an observable OS , describing some measurement we can perform. The intuitionshould be that this observable is local, in a sense that will become clear below.For definiteness assume |ψS〉 is an eigenstate of OS , so there is no uncertaintyof the outcome. Then our initial information, or condition, is the value x:

x = 〈ψS |OS |ψS〉 (2.6)

One could generalize to an optimal set of commuting operators OkS, ofwhich the expectation values xk form the maximal available initial information.We will illustrate the single operator case.

This approach means our initial information is given by observable values,not states, and losing it corresponds to the inability to perform measurementsthat output the value. When we are further along in time, the state is US |φS〉.We can find the expected instantaneous observable by taking the inner prod-uct 〈φS |U†SOSUS |φS〉, which may differ from the experiment outcome: assume

[HS , OS ] 6= 0. Instead, we are interested in recovering the initial conditionsfrom the state at this later time, so we will need a different operator. Theproposed form is evident: USOSU

†S , which corresponds to evolving the state

backwards to the initial time, measuring the state, and evolving forward again.It is manifestly Hermitian, so a well defined observable, and its action is thefollowing:

〈ψS |U†S USOSU†S US |ψS〉 = 〈ψS |OS |ψS〉 ≡ x (2.7)

10

Page 12: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Although this is theoretically possible, it is impractical. The operator’s formis very sensitive to the exact time of evaluation and value of HS . It is generallydifficult to construct. Still, there is no conceptual mechanism for informationloss. We must restore the bath.

In what follows, we will always assume the subsystem and bath at initial timeare in a product state. In other words: there is no initial entanglement. This willallow us to gauge to what extent the Hamiltonian generates it. The full systemat t = 0 is in the state |ψ〉 = |ψS〉 |ψB〉, unwritten tensor multiplication implied.In order to find our initial condition, we must also augment our observable:

O := OS ⊗ 1B (2.8)

This is our working definition of a local operator. It only depends on a few’close’ degrees of freedom: those contained in S. We call O local, when it istrivial on B: it does not act on B or depend on it. We can now formalize thenotion that there are ’preferred observables’: after partitioning S and B, we arerestricted to local operators. The ones accessible to us can measure a subset ofthe degrees of freedom fully, but have no access to the rest.

Within this framework, the initial condition on S is still well posed. Observe,in this larger Hilbert space, for any x,

〈ψ|O|ψ〉 = 〈ψS |OS |ψS〉 〈ψB |1B |ψB〉 = x · 1 = x (2.9)

Owing to the product nature of the state and operator, and normalizationof all states. We can still describe the initial information the same way.

Subsequently we turn on time evolution Udriven by a general H on S +B.At later time, the full system is in some state:

U |ψ〉 =∑jk

Cjk |φjS〉 |φkB〉 6= |ψS〉 |ψB〉 (2.10)

In general, the state will be a superposition over many products of com-ponent vectors, which cannot be written as a product state through any basistransformation. This constitutes entanglement between S and B. For moststates, the Schmidt number (minimal number of terms in the sum above) islarger than one [14]. From here, endeavoring to recover initial information, wemay attempt to reproduce the trick of equation 2.7. But we will fail on onecount: U OU† is not local. It too will be an intricate superposition of tensoredoperators, not trivial on the bath. We cannot reverse coupled time evolutionwith a local operator. There is one exception: for H = HS ⊗ 1B + 1S ⊗ HB ,U = US ⊗ UB , and all product forms are preserved. Without interaction, thereis no entanglement, and there can be no information loss or thermalization.

We get a sense that what we call local is subjective, a consequence of whatbasis we use to describe our system. Trivially, there is always a basis whichexhibits no thermalization, the basis of energy eigenstates. Their occupation isconserved in time, but the eigenstates do not in general respect the partitioninto S and B: they are not simple products of subsystem and bath states, but

11

Page 13: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

are entangled. They would be product states without interaction. The morethe energy eigenstates resemble our lab coordinates, the better behaved (moreaccessible and predictable) our system will appear to us. Some systems willentangle and thermalize more than others, and one wonders what is typical.Such questions deserve quantification, which we will attempt in later sections.

2.3 Methods

In the previous section, the idea was elucidated of only having a local, partialaccess to a global state. This begs the question: what is our best local de-scription of the subsystem S? What relates the information on S that can beobtained without access to B?

2.3.1 Density Matrices

For context, we remind the reader of the density matrix formalism. For anystate in bra-ket notation, we may assign a density matrix [14]:

|ψ〉 7→ ρ0 := |ψ〉 〈ψ| (2.11)

It is an operator, because it takes states to states -in this case a projectionoperator. Furthermore, it is Hermitian and positive semidefinite: Its eigenvaluesare nonnegative.

So far, we have been certain, in a classically probabilistic sense, of whatstate the system occupies: |ψ〉 with probability one. Therefore, ρ0 is a purestate, which is always the case when obtained as in 2.11. By contrast, if one isuncertain of the exact occupied state, one might wish to account for a discreteprobability distribution over different states. In order to be concrete, say a pjchance to be in state |ψj〉. The corresponding density matrix of such a mixedstate is the following:

ρ :=∑j

pj |ψj〉 〈ψj | (2.12)

It is equivalently an observable that measures the probability to be in eachof the states |ψj〉.

2.3.2 Tracing and Partial Tracing

We are already familiar with the trace, a linear mapping from operators toscalars. For any basis |φj〉 of H, and any operator Z ∈ HS, the Hilbert-Schmidt space of operators on H, we define:

Tr Z =∑j

〈φj |Z|φj〉 (2.13)

Although formulated in terms of a basis, the trace is a basis independentvalue. It can be thought of as an integral over the Hilbert space. As ρ is an

12

Page 14: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

operator also, we can take the trace Tr(ρ), and we will find that it is alwaysunit, as probabilities sum to 1. The following also holds due to 2.11 and thecyclic property of the trace:

〈ψ|Z|ψ〉 = Tr(ρ0Z) (2.14)

So there is an equivalent formalism to obtain expectations, which generalizesstraightforwardly to mixed states:

Tr(ρZ) = Tr

∑j

pj |ψj〉 〈ψj | Z

=∑j

pj Tr(|ψj〉 〈ψj | Z

)=∑j

pj 〈ψj |Z|ψj〉

(2.15)By linearity of the trace, and of expectations, this becomes the expectation

over the classical ensemble of expectations on quantum states: our best estimateof the observable.

A natural specification of the trace, is the partial trace: instead of integratingover the whole system, we only integrate out the bath (or subsystem, for thatmatter). This is done in practice by choosing a basis of B: |φjB〉, and summingsimilar to 2.13, but now it’s a mapping to a smaller Hilbert space:

TrB Z :=∑j

〈φJB |Z|φjB〉 (2.16)

Performing the partial trace on a pure density matrix ρ0 in HS is a naturalway to obtain a mixed density matrix on a constituent Hilbert-Schmidt space.Notably, the one not traced over.

TrB(ρ0) = ρS ∈ HSS (2.17)

This is our best description of S: a mixed state. The usefulness of the partialtrace stems from the following. For any product state |ψ〉 = |ψS〉 |ψB〉 in 2.11and operator of the form 2.8:

〈ψS |OS |ψS〉 = Tr(OSρS) = Tr(Oρ0) = 〈ψ|O|ψ〉 (2.18)

By linearity the last three expressions are valid for expectations over entan-gled states |ψ〉, without the ill-posed first equality in 2.18, as there is no |ψS〉.Note that also TrB(O) ∝ OS . In fact, a basis independent definition of partialtrace is that it is the unique linear operator satisfying:

trB(WS ⊗XB) = tr(XB)WS ∀XB ∈ HSB ,WS ∈ HSS (2.19)

For an illustration of trace and partial trace, we elicit the simplest nontrivialexample of a composite Hilbert space: when S and B are spin- 1

2 particles, orqubits. Then each lives in

HS = span(|↑〉 , |↓〉

)= HB (2.20)

13

Page 15: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Figure 2.2: Schematic of tracing ρ in a four-dimensional Hilbert space, resultingin a scalar.

Figure 2.3: Schematic of partially tracing 2 dimensions out of ρ, resulting in ρSin a two-dimensional Hilbert space.

A state in the full H will in general be a superposition of all combinationsof subsystem configurations.

|ψ〉 = C0 |↑↑〉+ C1 |↑↓〉+ C2 |↓↑〉+ C3 |↓↓〉 ∈ HS ⊗HB (2.21)

Then treating the bra-vectors as column coordinates, and ket-vectors asrows,

ρ = |ψ〉 〈ψ| = |C0|2 |↑↑〉 〈↑↑|+ C0C∗1 |↑↑〉 〈↑↓|+ . . . (2.22)

naturally reshapes to a matrix, see the figure 2.2. Identify a = C20 , b = C0C

∗1

etc. from 2.22 in the figures. The partial trace is also drawn for this example infigure 2.3. Note that although this graphic represents the operations in a specificbasis, the trace is basis independent, and the partial trace is independent of thebases of the constituent spaces.

We will work in the supposition that full systems are always pure. In theauthor’s view, there is no such thing as physical mixed states, which is insteada mathematical tool to express one’s lack of complete knowledge. Subsystemsmay be mixed, from an experimenters perspective, as full information on themrequires access to the bath. This view is evidenced by [28].

Combining the Schrodinger equation on the |ψ〉 ∼ ρ0 with the partial trace,we may construct the time evolution for density matrices.

d

dtρS(t) = TrB

(− i[H, ρ0(t)]

)(2.23)

This is called the Liouville-Von Neumann equation [27].

14

Page 16: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

2.4 Measures of Mixing

Although touched upon in the previous section, there is more to be said than thebinary division into conditions of mixed and pure. We envision a continuum ofvarying mixed states, and would like to delineate between cases with measures.We will define two.

2.4.1 Entanglement Entropy

The first is a well known concept, the entropy. One instinctively thinks of theamount of disarray. Many fields have their own definition of entropy, and in thisfield it is in a statistical sense, the uncertainty of a mixed state. Also known asthe Von Neumann or entanglement entropy [14]:

S(ρ) := −Tr(ρ ln ρ

)(2.24)

Indeed, it is weakly linked to the classical entropy that counts microstates.This is no secret. Von Neumann famously quipped to Claude Shannon, theformula’s inventor: ”You should call it entropy, for two reasons. In the firstplace your uncertainty function has been used in statistical mechanics underthat name, so it already has a name. In the second place, and more important,no one really knows what entropy really is, so in a debate you will always havethe advantage.”[29]

Involving a matrix logarithm, 2.24 is only defined through diagonalizationto the matrix’s eigenbasis. For a Hermitian operator Z ∈ HS with dimensiond, this always exists for some W ∈ U(d) -more on this in section 2.6- and inthat basis any function f is performed f(Z) = f(WDW †) = Wf(D)W †. Thefunction is simply evaluated element-wise on the entries of the diagonal D. Dueto the trace in 2.24, the expression is ultimately independent of the basis W inwhich the state is represented.

The entropy is zero if and only if ρ is pure. In its eigenbasis, (ρ0)j,k = δj,1δk,1and S(ρ0) is understood as the limit

S(ρ0) =∑j,k

((ρ0 ln ρ0)j,k

)δjk = −1 · 1 · ln(1)− (d− 1) · lim

x→0

(x ln(x)

)= 0 (2.25)

The maximally mixed state is ρ = 1/d, and S(1/d) = ln(d) also maximizesthe entropy. This entropy has more attractive features, but these will sufficefor our purposes. In the following chapters, we will sometimes wish to compareentropies in different size Hilbert spaces. For this, the normalized entropy isuseful:

S1(ρ) :=S(ρ)

S(1/d)=

Tr(ρ ln ρ

)ln(d)

(2.26)

Normalizing by the maximum entropy, S1(ρ) ≤ 1 for any space.

15

Page 17: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

2.4.2 Purity

Moving on to an inverted measure, the purity [30]. The emphasis is that itmeasures how pure the density matrix is, not how mixed or entropic.

γ(ρ) := Tr(ρ2)

(2.27)

The smaller the purity, the higher the degree of mixing. γ(ρ) = 1 is equiva-lent to ρ being pure. Our example in 2.11 gives ρ2

0 = |ψ〉 〈ψ|ψ〉 〈ψ| = |ψ〉 〈ψ| = ρ0

which has unit trace. An arbitrary mixed state defined in 2.12 has γ(ρ) =∑j p

2j ≤ 1, because squaring probabilities cannot increase them. The maxi-

mally mixed ρ = 1/d has:

γ(1/d) = 1/d (2.28)

So γ(ρ) ∈ [1/d, 1]. This function is computationally useful as it does notrequire diagonalization.

2.5 Quantum Distinguishability

Moving on, an underlying theme in the above has been distinguishability ofquantum states, specifically obtained under different initial conditions. A num-ber of useful tools shall be introduced here and in chapter 5 that formalizequantum distance. The first we encounter, a powerful concept, is the fidelity,which relates the resemblance between two states of the same system [14]. Men-tioned already at the beginning of the section, for pure states it is a simple innerproduct modulus.

F (|ψ〉 , |φ〉) := | 〈ψ|φ〉 | (2.29)

Some properties:

• Due to the absolute signs, the fidelity is always real, positive, and is sym-metric in its arguments.

• Due to normalization of Hilbert space vectors, F ∈ [0, 1].

• The fidelity is basis independent, as long as both states are in the samebasis.

• Lastly, a fidelity of 1 means |ψ〉 and |φ〉 are equal up to a U(1) gaugetransformation factor.

Though not directly evident, these properties carry over when we generalizeto mixed states [22], on which the fidelity is also defined, thusly:

F (ρ, η) := Tr√√

ηρ√η (2.30)

16

Page 18: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

In order to verify that the definitions agree in the limiting case, first set oneof the arguments pure: η = |φ〉 〈φ| = η2 [21]:

F (ρ, |φ〉 〈φ|) = Tr

√√|φ〉 〈φ|ρ

√|φ〉 〈φ| = Tr

√|φ〉 〈φ|ρ|φ〉 〈φ|

=√〈φ|ρ|φ〉Tr

√|φ〉 〈φ| =

√〈φ|ρ|φ〉

(2.31)

From here, also specify ρ = |ψ〉 〈ψ|:

F (|ψ〉 〈ψ| , |φ〉 〈φ|) =√〈φ|ψ〉 〈ψ|φ〉 = | 〈ψ|φ〉 | (2.32)

We recover expression 2.29. An additional property of the mixed state fi-delity ties it to the pure version as follows:

F (ρS , ηS) = max|ψ〉,|φ〉

| 〈ψ|φ〉 | (2.33)

Where the states |ψ〉 , |φ〉 are called purifications of ρS and ηS , i.e. states inan doubled Hilbert space such that tracing out one space reduces them to ρSand ηS :

|ψ〉 , |φ〉 ∈ HS ⊗HS , TrS(|ψ〉 〈ψ|

)= ρS , TrS

(|φ〉 〈φ|

)= ηS (2.34)

So the mixed fidelity is the maximal pure fidelity over possible purificationsof the arguments [14].

Why are we interested in the fidelity specifically? It is uniquely motivatedby the focus on operators posited above. Ideally, ’distance’ in the Hilbert spaceis directly related to distinguishability by measurements. In order for this tobe more formal, we refer to the theory of Positive Operator Value Measures(POVM) [30]. We recollect the fact that a quantum measurement has no pre-dictable outcome in general. Instead, any possible eigenstate or eigen-subspaceof the measurement-operator can result, with probability according to the Bornrule. A certain measurement M having a number of possible outcomes labelledby j induces a POVM: a set of operators Kj such that

∑j Kj = 1, and the

probability to find the jth outcome on state ρ is P (j) = Tr(ρKj). By linearitytotal probability is one. The experimenter deals with these probabilities, andtheir distribution is his or her tool to map the preceding state, by means ofquantum tomography.

The next ingredient is a result from classical probability theory: the Bhat-tacharyya coefficient, cos ∆. For two discrete classical probability distributionspj and qj, the overlap between the two statistical samples is given by:

cos ∆ =∑j

√pjqj (2.35)

17

Page 19: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Where ∆ is called the ’angle of divergence’. It is clear that cos ∆ = 1 ⇔qj = pj. The more distinct the distributions, the smaller the cosine, andthus the larger the angle of divergence. This angle is the most sensitive measureof the ability to distinguish the classical distributions.

As measurement on quantum states induces a classical probability distribu-tion, a natural question to ask, for two differing states, is the following: Whatdivergence between classical distributions resulting from these states is maxi-mal over all possible measurements? Equivalently, what is the minimum Bhat-tacharyya coefficient on the P (j)’s, taken over all POVMs? The answer wasproven to be the fidelity, 2.30, in [31].

F (ρ, η) = minKj

∑j

√Tr(ρKj) Tr(ηKj) (2.36)

Thus the fidelity directly relates an experimenters ability to distinguishstates, using the ideal measurement for that purpose. We will later use thefidelity as the preferred identifier of ’change’ to the quantum state. The largerF , the harder to tell the states apart, or the more similar they look.

In chapter 5, we will elaborate on differential concepts closely related tofidelity: the Quantum Geometric Tensor, and the Bures metric.

2.6 Random Matrix Theory

In discrete, finite Hilbert spaces, operators are represented by matrices. Whenwe discuss statistics of these operators, of quantum systems, the unavoidablelanguage is that of Random Matrix Theory (RMT). The idea is to treat a matrixas a random variable, living in a sample space, and following some probabilitydistribution, just like any scalar variable. Strongly coupled real world Hamil-tonian systems are often found to be well approximated by RMT, using onlyminimal knowledge of the actual inner workings of the models.

We will borrow from Eynard’s lecture notes [32].Random matrices may be viewed as statistical variables drawn from a Ran-

dom Matrix Ensemble (RME). The motivating example will be the ensemble ofHermitian matrices H ∈ E . This is because Hamiltonians are Hermitian, andwe will semantically equate a Hamiltonian to a specific quantum system.

In order to calculate probabilities in a continuous sample space, one needsa probability density function P(H), and a measure dH. The following is evi-dent notation for the probability to find H in some subset of the full space, ormanifold, E :

P (H ∈ Y ⊂ E) =

∫Y

P(H)dH (2.37)

And we will assume the measure and density are normalized:∫EP(H)dH = 1 (2.38)

18

Page 20: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

The spectral theorem guarantees a Hermitian matrix has a complete set oflinearly independent eigenvectors. A fact that will be used often in this work,all d× d Hermitian matrices H can be diagonalized as follows:

H = V ΛV †, V ∈ U(d), Λ = diag(E1, E2, . . . Ed) ∈ Rd (2.39)

Where Λ is a diagonal matrix holding eigenvalues of H, evocatively denotedE• for energy, and V is unitary matrix of size d carrying the eigenvectors of Hin its columns. In this case U(d) is what we call the symmetry group, of theRME, a Lie group. Intuitively, we wish for the particular V not to influence theprobability of H being drawn: thus it is the symmetric under the (left or right)action of U(d). Define the mapping (Λ, V ) 7→ H through 2.39, then (Λ, V ) isequiprobable to (Λ, V W ) and (Λ,WV ) for any W ∈ U(d). This reflects theearlier stated sentiment that there is no preferred basis.

Moving on, we examine the objects dH and P(H). The probability is con-ventionally written as a Boltzmann weight of a potential, which in turn is thetrace of a polynomial in H:

P(H) :=1

Ze−V(H) =

1

ZeTr(

∑dn=0 gnH

n) (2.40)

Z is the partition function, and normalizes the full probability as in 2.38.According to the Cayley-Hamilton theorem, the matrix powers Hndn=1 arelinearly dependent, thus so are their traces by linearity. This means we maycurtail the polynomial at order d. Observe that for V ∈ U(d), Tr(V ΛV †)n =Tr(V ΛnV †) = Tr(Λn). Combine this with linearity of the trace, and we find thatpotentials of this form are independent of the basis V , as desired: P(H) = P(Λ).To begin with, we usually take g2 = g, other gn = 0, in which case the potentialis Gaussian. This Gaussian distribution and unitary symmetry lend their namesto the Gaussian Unitary Ensemble or GUE, E . In fact, in this work, we will notbe much concerned with the potential, as it only influences the distribution ofenergies of the Hamiltonian. It has been treated here for sake of completeness.The interested reader may look into the saddle point method as one way ofreducing the task of integrating over this potential [32].

We now turn our attention to the most subtle component: the measure,dH. This measure simultaneously tracks the 1

2d(d + 1) independent real and12d(d − 1) imaginary components of a Hermitian matrix satisfying H† = H. Itturns out that a good model to reproduce physical behavior is to treat each ofthese 1

2d(d + 1) + 12d(d − 1) = d2 variables as real independent dimensions in

Euclidean space Rd2

, and see to what this means when they are arranged intoa matrix. This is called the Lebesgue measure:

dH =

d∏j=1

dHjj

∏j<k

dReHjkdImHjk (2.41)

In order to recover our basis invariance, we wish to change to the coordinatesin equation 2.39 and find a form:

19

Page 21: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

dH = d(V ΛV †) = J(Λ)dΛdV (2.42)

Where the Lebesgue measure splits into a product form of a measure onthe eigenvalues dΛ and an (invariant) measure dV on the eigenvectors. A co-ordinate transformation is accompanied by a Jacobian, which here is explicitlyindependent of V by assumption. The only choice for dV is the Haar measure,the unique uniform measure on a compact group, such as U(d), invariant underthe action of the group. It satisfies

∫U(d)

dV = 1, thus is a probability measure,

and d(WV ) = d(VW ) = dV , ∀W ∈ U(d) [33]. In order to find J(Λ), we equate2.41 and 2.42. From the latter:

d(V ΛV †) = dV ΛV † + V dΛV † + V ΛdV † (2.43)

As we assume the measure dH is independent of V , we may calculate itaround V = 1. This dramatically simplifies the equality. We also employ:

0 = d1 = d(V †V ) = dV †V + V †dV ⇔ dV † = −V †dV V † (2.44)

Where again for V = 1, dV † = −dV . Then:

dH = dV Λ1† + 1dΛ1† + 1ΛdV † = dΛ + [dV,Λ] (2.45)

From this expression, it is straightforward to read off the components dHjk.On the diagonal:

diag(dH) = dΛ⇔ dHjj = dEj (2.46)

Off the diagonal:

dH − diag(dH) = [dV,Λ]⇔ dHjk = (Ej − Ek)dVjk (2.47)

In expression 2.41, the imaginary and real parts multiply each other, result-ing in:

dH =

d∏j=1

dEj∏j<k

(Ej − Ek)2dReVjkdImVjk (2.48)

We now recognize the terms in 2.42:

dΛ =

d∏j=1

dEj , J(Λ) =∏j<k

(Ej − Ek)2, dV =∏j<k

dReVjkdImVjk (2.49)

The form of J(Λ) is the square of the so called Vandermonde determinantof the eigenvalues of H. It causes eigenvalue repulsion, a mutual interaction ofthe energies on top of their global, independent distribution determined by thepotential. This phenomenon often observed in strongly coupled and interacting

20

Page 22: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

systems, such as atomic nuclei of high atom number. All eigenstates are sensi-tive to the shape of all other eigenstates, which begets so called Wigner-Dysonenergy gap statistics, where the probability of finding a gap ∆E around zerovanishes: hence repulsion. Contrast this to Poisson statistics, symptomatic ofintegrable systems with many decoupled degrees of freedom. The composite,product structure of the energy eigenstates allows for many energy crossings,and the vanishing energy gaps are the most likely.

The moral of this section is that we may consider the distribution of theeigenvalues of H, scaling with P(Λ)J(Λ)dΛ, as an independent degree of free-dom from the distribution of its eigenvectors, according to the Haar measuredV . For instance, we will later integrate over the unitary group, leaving the dis-tribution over eigenvalues untouched. In this way, the GUE is like an ensembleof spheres. Equation 2.39 is then a decomposition into spherical coordinates,with an angular component V and radial Λ.

2.6.1 Generating Unitary Matrices

A crucial step in later simulations will prove to be the ability to sample unitarymatrices of arbitrary dimension d, uniformly according to the Haar measure, orfrom the Circular Unitary Ensemble [32]. An algorithm was explained by Ozolsin [34], following [35].

The starting point is the so called Ginibre Ensemble, which is computa-tionally available readily. One generates d2 independent identically distributedstandard normal complex random variables zjk.

f(zjk) =e−|zjk|

2

π(2.50)

These in turn can be generated by treating |zjk|2 as an exponentially dis-tributed random variable with parameter λ = 1. Then the complex phase ischosen uniformly at random from [0, 2π], and zjk is decomposed using polarcoordinates on C.

Once these are obtained, one arranges them in a matrix Z ∈ Cd×d, which isdistributed according to the Ginibre ensemble, of which the joint distributionis:

fG(Z) =

d∏j,k=1

f(zjk) =1

πd2exp

− d∑j,k=1

|zjk|2 =

1

πd2exp

(−Tr(Z†Z)

)(2.51)

It is immediately clear that this distribution is invariant under left -alsoright- action of the unitary group V ∈ U(d):

fG(V Z) =1

πd2exp

(−Tr((V Z)†(V Z))

)= fG(Z) (2.52)

In the ensemble, the subset of Z that are singular is measure zero. This canbe seen as follows. After generating any number < d of columns, the chance to

21

Page 23: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

generate another column inside their span is vanishingly small. Thus we mayassume Z is invertible and admits a so called QR-decomposition.

Z = QR, Q ∈ U(d), R ∈ T ′(n) (2.53)

Here, T ′(n) is the set of set of invertible d×d upper-triangular complex ma-trices. The QR decomposition of a matrix can be computed readily by standardGram-Schmidt orthogonalization algorithms. In general, it is not unique. Forany diagonal D ∈ U(d),

QR = QDD†R = (QD)(D†R) = Q′R′ (2.54)

is also a decomposition of Z. In order to guarantee uniqueness, we addition-ally impose R ∈ T (n), the specification of T ′(n) to have real positive diagonalentries. This can be done by setting

D = diag

(r11

|r11|,r22

|r22|, . . . ,

rdd|rdd|

)(2.55)

Then R′ = D†R ∈ T (n) from 2.54 and the Q′ = QD is unique, and themapping Z 7→ (Q′, R′) is well defined.

It remains to be shown that our candidateQ′ is uniformly drawn according tothe Haar measure. But this is immediate from 2.52: V Z = V Q′R′ 7→ (V Q′, R′)is equally likely to be generated as Z = Q′R′ 7→ (Q′, R′), for any possible V .As group action is closed, U(d)Q′ = U(d): V Q′ is the entire group. Each Q′

is then equally likely. A more rigorous explanation can be found in the source,[34].

2.7 Rudimentary Group Theory

We shortly recall from [36] a number of rote applications of discrete group theory,specifically pertaining to the symmetric group Sq, necessitated by chapter 4.

The symmetric group Sq, is the group of permutations on q symbols. Anelement of this group, σ, is a mapping from a set of q distinct, ordered elementsI := i1, i2, . . . , iq to itself. This is symbolically denoted Sq 3 σ : I → I. Thereare q! such permutations, so that is the cardinality of the finite group.

The basic building blocks are cycles: (a1a2 . . . ar) is an r-cycle, with eachaj ∈ I. This cycle corresponds to the mapping aj 7→ aj+1 for j = 1, 2, . . . , r −1, and the last ar 7→ a1. The unused elements I \ aj are mapped to them-selves. Clearly, r ≤ q.

For disjoint aj and bj, the cycles (a1a2 . . . ar) and (b1b2 . . . bs) can becomposed, or multiplied -the group action- to (a1a2 . . . ar)(b1b2 . . . bs). This issimply the mapping defined by carrying out both constituent mappings, andthe multiplication is commutative. By contrast, for aj ∩ bj 6= ∅, the mul-tiplication is consecutive application of the cycles, starting with the rightmostwritten one. It is in general not commutative. For all aj = bk, the total map-ping is bk−1 7→ bk = aj 7→ aj+1 or bk−1 7→ aj+1. This generalizes readily to

22

Page 24: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

elements σ, τ ∈ Sq which are themselves composed of multiple cycles. They arecomposed as στ ∈ Sq. Any element can be reduced to a representation of aproduct of disjunct cycles: in which each symbol only features once among allthe constituent cycles. Moreover, this representation is unique. After all, eachsymbol is only mapped to and from once.

There is a unique element Id ∈ Sq, the identity, which takes all symbols inI to themselves, and σId = Idσ = σ ∀σ ∈ Sq. Furthermore, any element σhas a unique inverse σ−1, satisfying σσ−1 = σ−1σ = Id. For an isolated cycle,(a1a2 . . . ar), the inverse is (arar−1 . . . a1), as it takes aj 7→ aj+1 7→ aj , andsimilarly for ar. Disjunct cycles commute, so this generalizes to the inverse ofcomposite elements in cycle representation, by inverting the constituent cycles.

Once a representation of σ ∈ Sq has been reduced to a product of disjunctcycles σ = (a1,1a1,2 . . . a1,r1)(a2,1a2,2 . . . a2,r2) . . . (as,1as,2 . . . as,rs), we say it isin the conjugacy class r1, r2, . . . , rs. This is because any conjugation τστ−1

consists of cycles of the same length, and any such cycle in the class can beobtained from any other for appropriate choice of τ ∈ Sq.

We are finally equipped to derive the results in the main body of this work.

23

Page 25: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Chapter 3

Numerical FidelitySimulations

We now move to the first results of this thesis, numerical calculations. Theywill aid us in getting a grasp of the behavior of quantum systems. Followingthe setup of section 2.2, we imagine a number of degrees of freedom for thesubsystem S and bath B. Numerics call for concreteness, so for the remainderof this section we set the building block of all discussed Hilbert spaces to bequbits, or spin 1

2 -particles. This is customary in many quantum-informationcontexts, as the qubit is the smallest unit of quantum information. Also it givesthe Hilbert space the maximally composite structure, maximizing the possibleamount of tracing relative to total dimension. This is useful because simulationson classical computers of high-dimensional quantum systems quickly becomehard to handle. What follows may be redesigned with any Hilbert space, solong as its dimension isn’t prime. We assume the predictions generalize tosystems that are not composed of qubits, but e.g. qudits: particles of largerspin, or any other composite discrete Hilbert space for that matter.

The aim of this chapter is to observe thermalization, or at least equilibration,in most quantum (sub)systems, determined by their driving full system Hamil-tonian. The most important statistic will be the fidelity, defined in 2.30. Asmentioned, the fidelity quantifies measurable distinguishability between states.If a subsystem tangibly moves away from its initial state, the fidelity betweenthe initial state and the later-time state must be less than unity. Conversely, ifthe subsystem reaches a steady state after some time, the fidelity between thesubsystem at two times after this point should approach unity. The strategyis thus to generate a random Hamiltonian, check the fidelity between differenttimes, and repeat. With enough such Hamiltonians, a trend is expected toemerge.

24

Page 26: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Figure 3.1: Schematic of a number of spins, all starting spin up, and all mutualinteractions possible. Here, two of the spins are S, and the rest are B.

3.1 Example Systems

In order to obtain insight, an example of fidelity over time is plotted in the figures3.2 and 3.3. For a number of different sized baths (nB = 4, 5 or 6 spins) andsubsystems (nS = 1 or 2 spins) we have generated one interacting Hamiltonianwith eigenvectors V from the Haar measure, and eigenvalues according to aflat individual distribution on [−π, π], but following the mutual repulsion of theVandermonde determinant in equation 2.49. These systems start evolution in aproduct state, so the subsystem is pure, and allowed to evolve according to theSchrodinger equation. All units are such that ~ = 1. A schematic of the setupis included in figure 3.1.

What function do we display exactly? F takes two arguments, so we havechosen one of them fixed, in essence a slice of the full picture, in order to havea 2D graph. The expressions F0 := F

(ρS(0), ρS(t)

)and F1 := F

(ρS(1), ρS(t)

)are plotted over t. Of course, the time t = 1 is arbitrary, but empirically forthe energies used, it is enough to be considerably far from the initial state,corroborated by figure 3.2.

It is already apparent in figure 3.2 that the larger the subsystem, the moreit can deviate from the initial state. There is, in a sense, more ’room’ to getaway, as the Hilbert space is larger. Next, compare with a not-initial state infigure 3.3.

As expected this figure has value one at t = 1, the fidelity between a stateand itself is maximal. After this time, the value hovers close to the maximum:the subsystem doesn’t deviate far from its value at time 1. By the same token,we can plot the time dependent entropy S (from 2.24) and purity γ (from 2.27)for similar systems. See the figures 3.5 and 3.4.

Pursuant to theory, the entanglement entropy of a pure state is 0. After this,the entropy plateaus around a value correlated to the system size. In fact, themaximum entropy for a nS-spin subsystem is ln(dS) = ln(2nS ) = nS · ln 2, so itappears these systems quickly maximize disorder. This conclusion is echoed bythe purity.

Note again that the minimum purity, 1/dS = n−2S corresponds to maximum

entropy. As a test of theory, random noninteracting Hamiltonians H = HS ⊗1B + 1S ⊗HB , were constructed and used to evolve the system (not plotted).

25

Page 27: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Figure 3.2: Example: F(ρS(0), ρS(t)

)plotted over time t, for 6 random spin

systems, consisting of different size subsystem and bath. Units are such that~ = 1.

Figure 3.3: Example: F(ρS(1), ρS(t)

)plotted over time t, for 6 random spin

systems. At t = 1, the fidelity is taken between the state and itself, and is thusexactly unit for all systems.

26

Page 28: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Figure 3.4: Example: S(ρS(t)) plotted over time t, for 6 random spin systems.Initially in a product state, the subsystem entropy is zero. It quickly increases,approaching a value dependent on nS .

Figure 3.5: Example: γ(ρS(t)) plotted over time t, for 6 random spin systems.Initially in a product state, the subsystem purity is 1. It quickly moves away,approaching a value dependent on nS .

27

Page 29: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Here entropy remained zero for all times, as one would expect.A final note: each of the figures 3.2, 3.3, 3.4, 3.5 were created with different

sets of random Hamiltonians, they are not intended to be compared directly.

3.2 Statistics of Fidelity

Let us now be more specific and rigorous. The numerical experiments detailedbelow have a common structure. The system lives in the variable n-spin Hilbertspace HS ⊗ HB . A Hamiltonian is generated, with eigenvectors V uniformlydrawn from the Haar measure, and eigenvalues (energies) arbitrary. We start,for definiteness, in the product state |1〉 = |1S〉⊗|1B〉 of the computational basisin 2.5. This may be thought of as all spins pointing up, as in figure 3.1. It isconventionally the first basis vector, and is a product state as it can be writtentrivially as |(↑↑ . . .)S〉 ⊗ |(↑↑ . . .)B〉, using the all up states of the components.The composite system is evolved forward to other times |ψ(t)〉 = U |1〉 accordingto the Schrodinger equation, 2.1. When we wish to evaluate the fidelity on thesubsystem, the bath is traced out from the full system pure state, as in 2.17:

ρS(t) = TrB(|ψ(t)〉 〈ψ(t)|

)(3.1)

Done for two different times, this may be used as arguments in definition2.30. We will construct two main statistics of the fidelity. For the first, considerthe expression:

〈F0〉 := limT→∞

1

T

∫ T

0

dtF (ρS(0), ρS(t)) (3.2)

This is the all-time t average of the fidelity of the subsystem between time 0,and itself at time t. It tracks how far the subsystem deviates from its startingstate. As mentioned before and forecast by figure 3.2, we expect this value tobe less than one for almost any system. It is clear from the construction thatρS(0) = |1S〉 〈1S |, a pure state. Then equation 2.31 tells us the F0 is simply:

F0 ≡ F(ρS(0), ρS(t)

)=√〈1S |ρS(t)|1S〉 =

√ρS(t)1,1 (3.3)

In index notation, it is simply the square root of the first diagonal element.This can be substituted into 3.2 to expedite the computation. The secondinteresting formula is:

〈F∞〉 := limT→∞

1

T 2

∫ T

0

dt

∫ T

0

duF (ρS(u), ρS(t)) (3.4)

It is the all time average over t and u of the fidelity between the subsystemat times t and u. This relates to the degree of equilibration, how well wecan distinguish the state at two later times. By our reasoning and evaluationof figure 3.3 we expect this to approach one for almost any system, implyingthermalization.

28

Page 30: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Analogous to the calculation of 〈F0〉, one can average the normalized entan-glement entropy, defined in 2.26, over time:

〈S1〉 := limT→∞

1

T

∫ T

0

dtS1 (ρS(t)) (3.5)

In the simulation, the choice was made to generate a unitary matrix V to rep-resent the eigenvectors of H as in subsection 2.6.1. Concerning the eigenvaluesΛ: time always multiplies energy in the time-dependent Schrodinger equation.Owing to the nature of time in expressions 3.2 and 3.4 as a dummy variable-it is always integrated out- the results should be energy independent. We onlyformally preclude any degenerate energy gaps, i.e. Eh − Ej = Ek − El ⇔(h, j) = (k, l), which otherwise would cause resonance and amplified Poincarerecurrence. In the general case without symmetries, due to the fully couplednature of the world, this is a physically reasonable assumption. Then 〈F0〉 and〈F∞〉 are essentially only functions of the eigenvectors of the Hamiltonian, theenergy eigenstates, arranged as the columns of V . Under this rationale, wesimplified the computational burden by sampling the energies independently atrandom from a uniform [−π, π] distribution. This does not reflect the interde-pendence caught by the RMT-prescribed Vandermonde determinant in section2.6, but that should not systematically affect the final estimates.

3.2.1 Experiment Parameters

A sequence of simulations were run, varying the amount nB of bath, and nSof subsystem spins. Hilbert spaces consisting of n = nB + nS equals 3, 4, 5,6, 7 and 8 total spins, subsequently divided into a subsystem of nS equalling1 or 2 and a bath comprising the rest were considered. For each of these 12situations, 4000 random Hamiltonians were generated and on each Hamiltonian,the following quantities were approximated by a Monte Carlo integral over time:

• 〈F0〉

• 〈F∞〉

• The temporal variance in ’F∞’: σ2∞

• Time average entanglement entropy 〈S1〉

We might also have included σ20 , the variance in F0, but it did not prove

very insightful on first estimates and was very similar to σ2∞.

Sampling 150 random times uniformly from t ∈ [0, 1000] constituted anordered dataset of density matrices for one Hamiltonian. Calculating F0 as in3.3 for each of these times, and taking the average over this vector, served asa proxy for 〈F0〉. Similarly, calculating the fidelity between each subsequentdensity matrix in the set, and averaging that, is how we achieved an estimate of〈F∞〉. Taking the variance of the vector resulted in σ2

∞. Though this may notbe the best, unbiased estimator for the actual variance, it gives ample sense of

29

Page 31: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Figure 3.6: Distribution of 4000 random Hamiltonians: their time average fi-delity to the initial pure state, 〈F0〉, for a system of n = 3, 4, 5 spins, and nS = 2spins. The surface under each plot is normalized to one.

the spread of F∞ over time. Lastly, the entropy 〈S〉 was obtained analogouslyto 〈F0〉. collecting all of these averages formed a single data point. These datapoints are displayed tallied into selected histograms in the rest of the chapter.More may be found in the appendix.

3.3 Results

This section will predominantly consist of figures accompanied by clarification.The horizontal scale of the distributions of fidelity is heavily dependent on theamount of total spins. For a ’large system’ (6,7 or 8 total spins n), the spreadis found to be an order of magnitude smaller than with a ’small system’, (n of3, 4 or 5). For clarity, the two cases will be plotted separately, allowing naturalaxes.

The first result is 〈F0〉 (equation 3.2) for a small system of which two spinsare taken to constitute the subsystem. See figure 3.6. The rest of the two-spinsubsystem cases, as part of the larger system, is plotted in 3.7.

The analogous plots, using nS = 1, are displaced to appendix A, figures A.1,A.2. It is clear from these four figures that 〈F0〉 tends to converge on a valuedepending on nS , and the degree to which the distribution clusters increaseswith bath size. For ns = 2, 〈F0〉 −−−−−→

nB→∞12 , for ns = 1, 〈F0〉 −−−−−→

nB→∞1√2. This is

consistent with the following supposition. The state ρS(t 6= 0) is generally closeto the maximally mixed state 1S/dS = 1S ·2−nS . The larger the bath, the closer.

30

Page 32: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Figure 3.7: Distribution of 4000 random Hamiltonians: their time average fi-delity to the initial pure state, 〈F0〉, for a system of n = 6, 7, 8 spins, and nS = 2spins. The surface under each plot is normalized to one.

As statistically more information is moved to the bath, the uncertainty on thesubsystem grows, capping off at 1S/dS . Using formula 3.3, and observing thatthe first (indeed any) diagonal element of the maximally mixed state is 2−nS ,we have:

limnB→∞

〈F0〉 =√

2−nS (3.6)

This confirms the prediction voiced in the previous section.We move on to the next quantity: the distribution of 〈F∞〉 (equation 3.4).

The case of a two-spin subsystem is plotted in figures 3.8 and 3.9. For theone-spin subsystem, see appendix A, figures A.3 and A.4.

From these figures, it is clear that for a random system, the average fidelitybetween different times does indeed have a large chance to be close to one.Moreover, the clustering of the distribution increases with bath size, and withthe ratio of bath to system. This is a confirmation of the forecast that almost allsystems, for close to any initial conditions, will become indistinguishable in time:a case for thermalization being the norm. The initial state will likely comprisemany energy eigenstates and will be mixed randomly from the perspective of anysubsystem. Conversely, states of the full system for which a small subsystem isout of this ’equilibrium’ must be rare. In the field of quantum dissipation, in factan infinite bath size is necessary to observe thermalization. Caldeira and Leggettfirst formulated a bath consisting of an infinite number of harmonic oscillators.With these infinite degrees of freedom, Poincare recurrence, or the coalescing

31

Page 33: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Figure 3.8: Distribution of 4000 random Hamiltonians: their time average fi-delity between all times, 〈F∞〉, for a system of n = 3, 4, 5 spins, and nS = 2spins. The surface under each plot is normalized to one.

Figure 3.9: Distribution of 4000 random Hamiltonians: their time average fi-delity between all times, 〈F∞〉, for a system of n = 6, 7, 8 spins, and nS = 2spins. The surface under each plot is normalized to one.

32

Page 34: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Figure 3.10: Distribution of 4000 random Hamiltonians: their temporal varianceof the fidelity between different times, σ2

∞, for a system of n = 6, 7, 8 spins, andnS = 2 spins. The surface under each plot is normalized to one.

of the phases of the system to eventually reproduce the initial conditions, ispostponed ad infinitum [37].

Supplementing this result is the temporal variance σ2∞. In figure 3.10, it

is seen to be very small for a large system with a two-spin subsystem. Thesmall system variant is in appendix A, figure A.5. The small variance meansnot only is the average fidelity close to one, it also stays in the neighborhoodof one for most times. Moreover, there is a theoretical argument making thisthe only possibility. 〈F∞〉 lies near one, the maximum, for most systems. UsingMarkov’s inequality on 1 − F ≥ 0, it is inescapable that the fidelity must beclose to one at nearly all times. A large portion of time with low fidelity couldnot be compensated in this average.

These graphs can be somewhat opaque. The state of the system at differenttimes ’looks’ the same, resembling some steady state. Although hypothesizingabove about the maximally mixed state, we do not know for certain what state itis. The distribution of the time-averaged entropy sheds light on this definitively.See figure 3.11 for the time average normalized entanglement entropy 〈S1〉 from3.5, for a subsystem of nS = 2 spins, and all systems n. The case of nS = 1 isremoved to to appendix A, figure A.6

This confirms our suspicion induced by figure 3.4 that the average subsystemmoves toward the maximally mixed state, the state that maximizes entropy. Thedegree to which it adheres scales with the ratio of the bath to the subsystem. I.e.a larger bath increases the probability to find the subsystem close to 1S for arandom coupled time evolution. This effect may be called quantum decoherence:

33

Page 35: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Figure 3.11: Distribution of 4000 random Hamiltonians: their time averagednormalized entropy 〈S1〉, for n = 3, 4, 5, 6, 7, 8 and nS = 2. The entropy is scaledby the subsystem dimension, such that the maximum entropy is 1, correspondingto ρS ∝ 1.

due to coupling with the environment, the information of the mutual phasebetween subsystem states is lost [27].

It is interesting to note that the peaks of the distributions in figures 3.11and A.6 agree neatly with an analytic result by Sen [38]. Hilbert space, being aprojective space, is compact. It is possible to draw a vector from it uniformlyat random according to the Haar measure [14]. If one does this for a composited = dS · dB dimensional system, and traces out the dB degrees of freedom, theresulting density matrix will on average have the entanglement entropy givenby:

〈S(ρS)〉dS ,dB =

d∑j=dB+1

1

j

− dS − 1

2dB= Ψ(d+ 1)−Ψ(dB + 1)− dS − 1

2dB(3.7)

Here, Ψ(z) is the digamma function. The normalized (equation 2.26) valuesof this average entropy 〈S1(ρS)〉dS ,dB corresponding to the system dimensionsin figures 3.11 and A.6 are tabulated in Table 1. We learn that for most timesand systems, the state |ψ(t)〉 is essentially uniformly random in H, as far asentanglement entropy is concerned.

34

Page 36: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Table 1: 〈S1(ρS)〉dS ,dB for spin systemsn = 3 n = 4 n = 5 n = 5 n = 6 n = 7

nS = 1 0.74 0.87 0.93 0.97 0.98 0.99nS = 2 0.34 0.67 0.83 0.92 0.96 0.98

The results above results agree qualitatively, in particular, with less geomet-ric calculations such as those by Reimann [10]. However, it is difficult to makequantitative statements, due to the arbitrary magnitude of constants like themacroscopic measurement accuracy, which has no real analog in the language ofthis work. Also [11] draws similar conclusions, both in results and in philosophyof the nature of our ignorance.

The simulations above were carried out in Python. The main scripts, ac-companied by usage instructions, are to be found in appendix B.

35

Page 37: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Chapter 4

Analytic Mixed StateAverages

In this chapter, we will use geometrical, analytic arguments to quantify ther-malization of the typical quantum system.

4.1 Analytic statistics of local density matrices

The physical setup is the same as in the previous chapters. We construct aHamiltonian with unspecified eigenvalues Eii, i ∈ 1, 2, . . . , d, and eigenvec-tors defined by the unitary matrix V , drawn from the Haar measure on the uni-tary group U(d). Considering qubits again, one may take d = 2(nS+nB) = dimH,as it represents the dimension of nS system qubits (spins) and nB bath qubits,which will be coupled together by the Hamiltonian. Then dB = 2nB , dS = 2nS

are the dimensions of the component Hilbert spaces. However, these results aremore general, and will hold for any dS , dB sized component systems totallingd = dS · dB full system dimensions. In a diagonalized form, from section 2.6:

H = V ΛV †, V ∈ U(d), Λ = diag(E1, E2, . . . , Ed) (4.1)

Then, following the standard construction, we have H evolve the coupledsystem and bath, starting in a product state |1〉 = |1S〉 ⊗ |1B〉 at t = 0. Thisconstitutes one of our assumptions, that there is no initial entanglement betweenthe artificially partitioned S and B. We have chosen the basis of the full Hilbertspace H = HS ⊗ HB such that our (random) initial state is indeed the firstbasis vector, in turn defined to be the tensor product of the component firstbasis vectors. The full basis is |k〉. We may always do this by means of localunitary basis transformations, which can then be absorbed into V . By default:

|ψ(t)〉 = e−iHt |1〉 , ρS(t) = TrB(|ψ(t)〉 〈ψ(t)|) (4.2)

36

Page 38: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

We cannot at present perform the integrals from chapter 3, for instancethe fidelity between ρ at different times, due to the diagonalization necessaryin the definition of the matrix root. There are identities to integrate over theunitary group, as long as the integrand is of a certain symmetric form, as inthe Harish-Chandra-Itzykson-Zuber-formula [39]. Then we must ask slightlydifferent questions.

4.1.1 Average Reduced Density Matrix

We will take a different route from the previous, proving a related result. Weaim to find the elements of ρS(t), averaged uniformly over the possible V . Westart by explicitly writing the components of |ψ(t)〉 :=

∑k ψk(t) |k〉, which are

found by ψk(t) = 〈k|ψ(t)〉

|ψ(t)〉 = e−iHt |1〉 = e−iV ΛV †t |1〉 = V e−iΛtV † |1〉 (4.3)

ψk(t) = 〈k|V e−iΛtV † |1〉 =

d∑j=1

Vkje−iEjt(V †)j1 (4.4)

The components of the full density matrix ρ(t) = |ψ(t)〉 〈ψ(t)| are given byρ(t)kl = 〈k|ψ(t)〉 〈ψ(t)|l〉 = ψk(t)ψ∗l (t)

Now it must be observed that each index is actually a multi-index j '(jS , jB), owing to the tensor product structure of the Hilbert space. By formula2.4, the full basis is formed by combinations of the component bases. In orderto calculate the reduced density matrix, one must contract only the index of thebath from 1 to dB . We will make the multi-indexed nature explicit only whereneeded, on the outer sides of ρ(t). In components:

ρS(t)kSgS =

dB∑kB=1

d∑j,m=1

V(kS ,kB)je−iEjtV †j1V1me

iEmtV †m(gS ,kB) (4.5)

This expression reveals that the time and energy and time dependence is justa multiplicative factor on each term of the sum. We invoke an identity applicableusing Weingarten functions, following the techniques of Benoit Collins [40].Products of individual elements of V may be integrated over U(d). Withoutgoing into too much detail of the derivation of this identity, which involvesrepresentation theory, we will simply state the result:

∫U(d)

dV Vi1,j1 . . . Viq,jqV†j′1,i′1. . . V †j′q,i′q =∑

σ,τ∈Sq

δi1,i′σ(1) . . . δiq,i′σ(q)

δj1,j′σ(1) . . . δjq,j′σ(q)

Wg(d, στ−1)(4.6)

In this expression, Sq is the symmetric group on q symbols, and Wg(d, σ)is on of a fixed set of Weingarten functions on d dimensions and permutation

37

Page 39: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

σ. Which particular function of the set is used depends only on the conjugacyclass (cycle lengths) of σ. All are quotients of polynomials in d. Thankfully,they has been tabulated by Brouwer and Beenakker for arbitrary d and forpermutations up to q = 5, in [41], so all need be done is to sum over σ, τ inthe symmetric group, and multiply the Kronecker deltas by the Weingartenfunction corresponding to the class of στ−1. For the usage of permutationcycles, conjugation, inverses, and classes we refer to section 2.7, and for morebackground, to any introductory text on group theory, e.g. [36].

In general, not all delta functions will be satisfied, and they will kill manyterms in the sum of 4.5. In our case, for the average ρS(t), it is clear that q = 2,so the possible permutations are τ, σ ∈ Id, (12). We can also populate theindices i1, i2, j1, j2, i

′1, i′2, j′1, j′2 in 4.6. Reordering the scalar, indexed terms, our

expression now takes the form:

∫U(d)

dV ρS(t) =

∫U(d)

dV

dB∑kB=1

d∑j,m=1

V(kS ,kB)jV1mV†j1V

†m(gS ,kB)e

i(Em−Ej)t =

∑σ,τ∈S2

dB∑kB=1

d∑j,m=1

δi1,i′σ(1)δi2,i′σ(2)

δj1,j′σ(1)δj2,j′σ(2)

Wg(d, στ−1)ei(Em−Ej)t

(4.7)

Where we imply the following index equalities:

(i1, i2) =((kS , kB), 1

); (i′1, i

′2) =

(1, (gS , kB)

)(j1, j2) = (j,m); (j′1, j

′2) = (j,m)

(4.8)

There are four combinations of σ, τ in the sum in 4.7. We may write:∫U(d)

dV ρS(t) =∑

σ,τ∈S2

Rσ,τ (4.9)

The R•,•’s involve two contributing Weingarten functions for S2. These are,symbolically dependent on cycle length as the 2nd argument:

Wg(d, 12) =1

d2 − 1, Wg(d, 2) =

−1

d(d2 − 1)(4.10)

For σ = τ = Id, στ−1 ∼ 12, the contribution Rσ,τ becomes:

RId,Id =

dB∑kB=1

d∑j,m=1

δ(kS ,kB),1δ1,(gS ,kB)δj,jδm,mWg(d, 12)ei(Em−Ej)t

=χ(t)

d2 − 1δkS ,1δgS ,1

(4.11)

38

Page 40: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Figure 4.1: The function χ(t), gen-erated with d = 5 random, mutuallyrepulsive energies Ej drawn from[−π, π].

Figure 4.2: The function χ(t), gen-erated with d = 100 such energies.The time t = σ−1

E is indicated onthe horizontal axis.

The first deltas kill all terms over kB except the first, as only multi-index(kS , kB) = (1, 1) = 1 as a single index. Here we have also defined function χ(t)to carry the time and energy dependence:

χ(t) :=

d∑j,m=1

ei(Em−Ej)t = d+∑j<m

2 cos ((Em − Ej)t) (4.12)

Observe that χ(t = 0) = d2 and empirically, the function quickly drops toits time average d, in time t = O(σ−1

E ), around the inverse standard deviationin energies. See the figures 4.1 and that beside it, for an example.

The χ(t) plotted are constructed with energies uniform in [−π, π], but gener-ated by the techniques of RMT, thus with mutual energy repulsion. For a smallnumber of dimensions, i.e. d = 5, the harmonic character of χ(t) is still preva-lent, Poincare recurrence is rapid, and the function is erratic. Increasing thenumber to d = 100, we see that χ(t) is more predictable, and seems ’damped’due to the strong eventual decoherence of the phases, causing destructive in-terference. This d may seem like a large number, but 7 qubits already have a27 = 128-dimensional Hilbert space. These are not experimentally intractablesystems. From some (admittedly subjective) trials, it appears the neat decayingbehavior depends on the eigenvalue-repulsion, under which rapidly oscillatingcosines are suppressed. For independent random energies from the same sup-port, the damped character of the function is less pronounced, though to makethis suspicion rigorous, more study is needed.

The next term is for (σ, τ) =(Id, (12)

), στ−1 ∼ 2:

RId,(12) =

dB∑kB=1

d∑j,m=1

δ(kS ,kB),1δ1,(gS ,kB)δj,mδm,jWg(d, 2)ei(Em−Ej)t

=−d

d(d2 − 1)δkS ,1δgS ,1

(4.13)

39

Page 41: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Here, m = j is forced by the delta function, leaving the exponent equal tounity for all times. Moving on, (σ, τ) =

((12), Id

):

R(12),Id =

dB∑kB=1

d∑j,m=1

δ(kS ,kB),(gS ,kB)δ1,1δj,jδm,mWg(d, 2)ei(Em−Ej)t

=−dBχ(t)

d(d2 − 1)δkS ,gS

(4.14)

Here, the delta in the multi-indices is satisfied for every dummy kB , con-tributing a factor dB , but only on diagonal entries in the system’s density matrixkS = gS . The last combination, (σ, τ) =

((12), (12)

), στ−1 ∼ 12:

R(12),(12) =

dB∑kB=1

d∑j,m=1

δ(kS ,kB),(gS ,kB)δ1,1δj,mδm,jWg(d, 12)ei(Em−Ej)t

=dB · dd2 − 1

δkS ,gS

(4.15)

Summing the four terms R•,• in 4.9 we obtain the full integral.

ρS(t) :=

∫U(d)

dV ρS(t) =(χ(t)

d2 − 1+

−dd(d2 − 1)

)δkS ,1δgS ,1 +

(−dBχ(t)

d(d2 − 1)+

dBd

d2 − 1

)δkS ,gS =(

χ(t)− 1

d2 − 1

)|1S〉 〈1S |+

(dB(d2 − χ(t))

d(d2 − 1)

)1S

(4.16)

Elements off the main diagonal are zero, and on the diagonal, the first ele-ment (kS , gS) = (1, 1) is unique.

First a consistency check: The integrand ρS has trace 1 at all times, andthe measure is normalized such that

∫U(d)

dV = 1. Furthermore integrating and

tracing are linear and must commute. So 1 =∫U(d)

1dV =∫U(d)

Tr(ρS)dV =

tr(∫U(d)

ρSdV ) = Tr(ρ(t)

). Let us verify:

Tr(ρS(t)) =

(χ(t)− 1

d2 − 1

)+ dS

(dB(d2 − χ(t))

d(d2 − 1)

)=χ(t)− 1 + d2 − χ(t)

d2 − 1= 1

(4.17)Indeed, using the fact that dS · dB = d. Next, we consider the behavior for

t = 0, or averaged over all time. As mentioned above, χ(0) = d2. Substitutingthis in 4.16, the diagonal terms vanish, and the first element term becomes unit,as we would expect: at time t = 0, no matter the Hamiltonian, the unevolvedsystem is pure, in the first basis state, ρS(0) = ρS(0) = |1S〉 〈1S |. Conversely,

40

Page 42: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

averaged over time, the cosines in 4.12 vanish, and 〈χ〉t = d. As ρS(t) is linearin χ(t), the time average 〈ρS(t)〉t is obtained by substituting χ(t) 7→ d.

〈ρS(t)〉t =

(d− 1

d2 − 1

)|1S〉 〈1S |+

(dB(d2 − d)

d(d2 − 1)

)1S =(

1

d+ 1

)|1S〉 〈1S |+

(dBd+ 1

)1S

(4.18)

Which is close to the maximally mixed state with equal entries on each diag-onal. We conjecture that χ(t) spends most time close to its time average. Theninformation of the initial conditions persist in the anomalous extra occupationof the first element forever. One may think of a phase transition of sorts be-tween t = 0 and t → ∞, moving from a ’known’ to ’unknown’ phase, with theshift occurring around σ−1

E .From here interestingly, using d = dB ·dS , it is evident that taking dB →∞,

leaving dS fixed, we indeed obtain the maximally mixed state for the time andsystem averaged density matrix.

limdB→∞

〈ρS(t)〉t = limdB→∞

[(1

dBdS + 1

)|1S〉 〈1S |+

(dB

dBdS + 1

)1S

]=

1

dS1S

(4.19)

The density matrix is a mathematical object, it cannot be observed directly.In order to obtain a more physical result, one would like to evaluate a localoperator OS on ρS , as in equation 2.18. By linearity of expectation (trace andintegral commute), this is straightforward:

OS(t) :=

∫U(d)

dV Tr(OSρS(t)

)= Tr

(OS

∫U(d)

dV ρS(t)

)= Tr

(OS ρS(t)

)=

(χ(t)− 1

d2 − 1

)〈1S |OS |1S〉+

(dB(d2 − χ(t))

d(d2 − 1)

)Tr(OS)

(4.20)

Where we have substituted 4.16. Integrating over all systems has left lit-tle structure of the operator, a contribution from the initial condition, and acontribution from the diagonal, mixed ensemble average.

4.1.2 Variance of the Reduced Density Matrix

The next logical step is to calculate the variance of the matrix elements ofρ(t) over the unitary group. This will lend credence to the average. For a smallvariance, we may assume most systems behave like their average. Unfortunately,this is expression quartic in V as well as in V †, so q = 4, and the number of

41

Page 43: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

terms contributing to the integral scales as (q!)2, exploding to 576 combinatorialpairs σ, τ ∈ S4. We will touch upon the important steps of the calculation, mostof the manipulation was performed by computer.

The variance for a complex quantity is given by:

var (ρS(t)) :=

∫U(d)

dV∣∣ρS(t)

∣∣2 − ∣∣∣∣∣∫U(d)

dV ρS(t)

∣∣∣∣∣2

(4.21)

In this expression, modulus squared is taken element-wise, a Hadamardpower, not as a matrix, for which the circle serves as a reminder. The sec-ond term is simply the Hadamard square of ρS(t). The first, we elaborate:

∫U(d)

dV∣∣ρS(t)

∣∣2 =

∫U(d)

dV

∣∣∣∣∣∣dB∑kB=1

d∑j,m=1

V(kS ,kB)jV1mV†j1V

†m(gS ,kB)e

i(Em−Ej)t

∣∣∣∣∣∣2

=

∫U(d)

dV

dB∑kB ,gB=1

d∑j,m,l,h=1

V(kS ,kB)jV1mV∗(kS ,gB)lV

∗1h×

V †j1V†m(gS ,kB)V

Tl1 V

Th(gS ,gB)e

i(Em+El−Ej−Eh)t

=

∫U(d)

dV

dB∑kB ,gB=1

d∑j,m,l,h=1

V(kS ,kB)jV1mV1lV(gS ,gB)h×

V †j1V†m(gS ,kB)V

†l(kS ,gB)V

†h1e

i(Em+El−Ej−Eh)t

(4.22)

Now invoking Collins’ identity in 4.6, a peculiar pattern emerges: The fullrange indices j,m, l, h decouple from the split ones kS , kB , gS , gB . Definingmulti-indices I := (i1, i2, i3, i4) and similar for I ′, J, J ′:

I = ((kS , kB), 1, 1, (gS , gB)) ; I ′ = (1, (gS , kB), (kS , gB), 1)J = (j,m, l, h); J ′ = (j,m, l, h)

(4.23)

This means σ only takes split indices (or index 1) in I to themselves in I ′. τonly takes full indices in J to other full indices in J ′. The sums over these setsindices can be then performed independently.

42

Page 44: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

∫U(d)

dV∣∣ρS(t)

∣∣2 =

∑σ,τ∈S4

dB∑kB ,gB=1

d∑j,m,l,h=1

δI,σ(I′)δJ,τ(J′)ei(Em+El−Ej−Eh)t

Wg(d, στ−1) =

∑σ,τ∈S4

dB∑kB ,gB=1

δI,σ(I′)

d∑j,m,l,h=1

δJ,τ(J′)ei(Em+El−Ej−Eh)t

Wg(d, στ−1)

(4.24)

In this expression, the obvious shorthand is δI,σ(I′) :=∏4z=1 δiz,i′σ(z)

This allows the following simplification: We will create a table of contribu-tions Rσ and Qτ due to σ, τ ∈ S4 respectively.

Rσ :=

dB∑kB ,gB=1

δI,σ(I′), Qτ :=

d∑j,m,l,h=1

δJ,τ(J′)ei(Em+El−Ej−Eh)t (4.25)

Each combination from each table must be multiplied by the correspondingWeingarten function, and forms one term of the final expression. From 4.24 and4.25, we substitute:∫

U(d)

dV∣∣ρS(t)

∣∣2 =∑

σ,τ∈S4

RσQτWg(d, στ−1) (4.26)

See the expressions for Rσ and Rτ in Table 2.

Table 2: |ρS(t)|2 Sum Factors

σ, τ ∈ S4 Rσ Qτ σ, τ ∈ S4 Rσ QτId |1S〉 〈1S | χ2(t) (12) dB |1S〉 〈1S | dχ(t)

(123) dB |1S〉 〈1S | χ(t) (13) |kS〉 〈1S | dχ(t)(132) |kS〉 〈1S | χ(t) (14) |1S〉 〈1S | ξ−(t)(124) dB |1S〉 〈1S | χ(t) (23) |1S〉 〈1S | ξ+(t)(142) |1S〉 〈gS | χ(t) (24) |1S〉 〈gS | dχ(t)(134) |kS〉 〈1S | χ(t) (34) dB |1S〉 〈1S | dχ(t)(143) dB |1S〉 〈1S | χ(t) (1234) dB |1S〉 〈1S | d(234) |1S〉 〈gS | χ(t) (1243) d2

B1S d(243) dB |1S〉 〈1S | χ(t) (1324) |kS〉 〈1S | d

(12)(34) d2B1S d2 (1342) dB |kS〉 〈gS | d

(13)(24) dB |kS〉 〈gS | d2 (1423) |1S〉 〈gS | d(14)(23) |1S〉 〈1S | ζ(t) (1432) dB |1S〉 〈1S | d

We will elaborate on the derivation of the various terms. In the following,we use the observation that δ2

•,• = δ•,•, and will move back and forth betweenindex and bra-ket notation when needed. Let us start with the column of Rσ.

43

Page 45: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

For σ = Id, the identity, we have I = I ′, or from 4.23:

(kS , kB) = 1, 1 = (gS , kB), 1 = (kS , gB), (gS , gB) = 1 (4.27)

Which kills all the summands except for kB = 1 = gB , and elements exceptgS = 1 = kS .

RId =

dB∑kB ,gB=1

δ(kS ,kB),1δ1,(gS ,kB)δ1,(kS ,gB)δ(gS ,gB),1 = |1S〉 〈1S | (4.28)

Permutations sending (1, 4) 7→ (1, 4) have the same assignment (as theypermute (2, 3) 7→ (2, 3) as well). So for σ ∈ (14), (23), (14)(23), Rσ = RId.

For σ = (12), one finds:

(kS , kB) = (gS , kB), 1 = 1, 1 = (kS , gB), (gS , gB) = 1 (4.29)

This is satisfied for all kB in the sum, contributing a factor dB . All otherindices are set to one and equation 4.25 simplifies to R(12) = dB |1S〉 〈1S |. Thesame assignments hold for σ ∈ (124), (123), (1234). When σ ∈ (34), (143), (243), (1432),a factor dB comes from free summation over gB , and Rσ = dB |1S〉 〈1S | too.

For σ = (13), see the assignment:

(kS , kB) = (kS , gB), 1 = (gS , kB), 1 = 1, (gS , gB) = 1 (4.30)

Here, the condition on kS is trivial, the ket vector component in |ρS(t)|2. Allother indices are forced to 1: R(13) = |kS〉 〈1S |. Also σ ∈ (132), (134), (1324)permute the 1st index to the 3rd, giving Rσ = R(13).

Conversely, σ ∈ (24), (142), (234), (1423) all permute the 4th index to the 2nd

and the rest of the indices are set to unity by equation 4.23. This means the gSindex is free, or the bra vector component: Rσ = |1S〉 〈gS |.

The next permutation not yet treated is σ = (1243). Here we obtain theassignment:

(kS , kB) = (gS , kB), 1 = 1, 1 = 1, (gS , gB) = (kS , gB) (4.31)

This restricts to diagonal elements kS = gS , but both dummy bath indicesgB , kB are free, resulting in R(1243) = d2

B1S = R(12)(34) by analogous reasoning.The final Rσ comes from σ ∈ (1342), (13)(24):

(kS , kB) = (kS , gB), 1 = 1, 1 = 1, (gS , gB) = (gS , kB) (4.32)

In this case, all bra and ket elements survive, only a δgB ,kB kills one dummysum, leaving a factor dB : R(1342) = dB |kS〉 〈gS | = R(13)(24).

44

Page 46: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

We now move on to the column of Qτ defined in 4.25. The first to consideris τ = Id. From 4.23, we see immediately that there are only tautologicalconditions on the indices j,m, l, h. Then the sum factorizes:

QId =

d∑j,m,l,h=1

δJ,Jei(Em+El−Ej−Eh)t

=

d∑j,m=1

ei(Em−Ej)t

d∑l,h=1

ei(El−Eh)t

= χ2(t)

(4.33)

Where we are reminded of the definition of χ(t) in 4.12.Continuing to τ = (12), we have the identification j = m, and l, h are trivial.

Then after annihilating the Kronecker delta against the summation over m, Emcancels Ej in the exponent of 4.25:

Q(12) =

d∑j,m,l,h=1

δj,mei(Em+El−Ej−Eh)t =

d∑j,l,h=1

ei(El−Eh)t = dχ(t) (4.34)

We clarify that this is not a differential, but a multiplication by dimensiond. The same results from τ ∈ (13), (24), (34), any permutation taking m∨ l↔j ∨ h.

Turning to τ = (14), the assignment is j = h after killing the Kroneckerdelta. This leaves us with a newly defined function:

Q(14) =

d∑j,m,l,h=1

δj,hei(Em+El−Ej−Eh)t = ξ−(t) :=

d∑j,l,m=1

ei(Em+El−2Ej)t

(4.35)We similarly define, for τ = (23),m = l:

Q(23) =

d∑j,m,l,h=1

δm,lei(Em+El−Ej−Eh)t = ξ+(t) :=

d∑j,l,m=1

ei(2Em−Eh−Ej)t

(4.36)We note that, up until now, all elements have been real, and by construction

the variance is real. However, in general ξ∓(t) are complex, and only com-binations such as ξ− + ξ+ are real, where we can pair each term e• with itsconjugate e−•. A good consistency check will be that the final answer does notfeature asymmetric combinations of these functions. On to the first three-cycle,τ = (123). Here, as with all three-cycles, 3 out of 4 indices are equated, such asj = m = l, and the 4th, h is left free. In all these cases, two of the three equatedindices kill their energies in the exponent, i.e. Em +El−Ej = Em, and the 4th

has opposite sign. All three-cycles contribute a factor:

45

Page 47: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Q(123) =

d∑j,m,l,h=1

δj,mδm,lδl,jei(Em+El−Ej−Eh)t =

d∑j,h=1

ei(Ej−Eh)t ≡ χ(t)

(4.37)Analogously, all four-cycles equate all indices j = m = l = h. Three sum-

mations kill four Kronecker-deltas, as the last is redundant (cyclic).

Q(1234) =

d∑j,m,l,h=1

δj,mδm,lδl,hδh,jei(Em+El−Ej−Eh)t =

d∑j=1

e0 = d (4.38)

They all contribute Qτ = d. Continuing to the double two-cycle (12)(34),j = m, l = h, only 2 summations suffice to cancel the exponent:

Q(12)(34) =

d∑j,m,l,h=1

δj,mδl,hei(Em+El−Ej−Eh)t =

d∑j,l=1

e0 = d2 (4.39)

The same is true for the cross-assignment, τ = (13)(24). The final termis Q(14)(23), where j = h,m = l. This doubles the frequency of the exponent,defining:

Q(14)(23) =

d∑j,m,l,h=1

δj,hδm,lei(Em+El−Ej−Eh)t =

d∑j,m=1

ei(2Em−2Ej)t

= ζ(t) := d+∑j<m

2 cos (2(Em − Ej)t)(4.40)

Finally, see table 3 for the numerators of the appropriate Weingarten func-tions, from [41]. All Wg(d, σ) for σ in a particular Sq have the same denomina-tor.

Wg(d, σ)∣∣∣σ∈S4

=A4(d, σ)

B4(d),

B4(d) = d2(d2 − 1)(d2 − 4)(d2 − 9) =

3∏z=0

(d+ z)(d− z)(4.41)

Table 3: Wg(d, σ) functionsσ conjugacy class A4(d, σ)

1, 1, 1, 1 d4 − 8d2 + 62, 1, 1 −d3 + 4d2, 2 d2 + 63, 1 2d2 − 34 −5d

46

Page 48: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

We are now equipped to perform the sum in 4.26, using tables 2 and 3.Simplifying the expression is arduous, but straightforward. From equation 4.41,it is clear that all terms of the sum will have the same denominator. We firstcollect the numerator, and state it here without the preceding simplification.

∑σ,τ∈S4

RσQτA4(d, στ−1) = |1S〉 〈1S |

[(ξ−(t) + ξ+(t) + ζ(t) + χ2(t)− 4χ(t)

(d4 − 2d3 − 7d2 + 8d+ 12

)− 2dB

d5 − 2d4 − 9d3 + 18d2−

χ(t)(d5 − 2d4 − 5d3 + 2d2 + 4d+ 24

)+(ξ−(t) + ξ+(t) + ζ(t) + χ2(t)

)(d3 − 4d2 + d+ 6

)]

−(|kS〉 〈1S |+ |1S〉 〈gS |

)[d5 − 2d4 − 9d3 + 18d2 − χ(t)

(d5 − 2d4 − 5d3 + 2d2 + 4d+ 24

)+(ξ−(t) + ξ+(t) + ζ(t) + χ2(t)

)(d3 − 4d2 + d+ 6

)]+(d2B1S + dB |kS〉 〈gS |

)·[

d6 − d5 − 11d4 + 9d3 + 18d2 − 2χ(t)(d4 − 2d3 − 7d2 + 8d+ 12

)+

(ξ−(t) + ξ+(t) + ζ(t) + χ2(t)

)(d2 − 5d+ 6

)](4.42)

The polynomials in d here can be factorized neatly, almost all are productsof terms of the form (d± z), z ∈ 0, 1, 2, 3, as is denominator B4(d).

∑σ,τ∈S4

RσQτA4(d, στ−1) = |1S〉 〈1S |

[(Ξ(t)− 4χ(t)

)(d+ 1)(d− 2)(d+ 2)(d− 3)

]

−(|kS〉 〈1S |+ |1S〉 〈gS |+ 2dB |1S〉 〈1S |

)[d2(d− 2)(d− 3)(d+ 3)

− χ(t)(d− 2)(d+ 2)(d− 3)(d2 + d+ 2) + Ξ(t)(d+ 1)(d− 2)(d− 3)

]

+(d2B1S + dB |kS〉 〈gS |

)[d2(d+ 1)(d− 2)(d− 3)(d+ 3)

− 2χ(t)(d+ 1)(d− 2)(d+ 2)(d− 3) + Ξ(t)(d− 2)(d− 3)

](4.43)

Here we have also substituted the placeholder:

47

Page 49: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Ξ(t) := ξ−(t) + ξ+(t) + ζ(t) + χ2(t) (4.44)

We can divide by B4(d) in 4.41 to find the average of ρS squared.

∫U(d)

dV∣∣ρS(t)

∣∣2 = |1S〉 〈1S |

[Ξ(t)− 4χ(t)

d2(d− 1)(d+ 3)

]

−(|kS〉 〈1S |+ |1S〉 〈gS |+ 2dB |1S〉 〈1S |

)[ 1

(d+ 1)(d− 1)(d+ 2)

− χ(t)(d2 + d+ 2)

d2(d+ 1)(d− 1)(d+ 3)+

Ξ(t)

d2(d− 1)(d+ 2)(d+ 3)

]

+(d2B1S + dB |kS〉 〈gS |

)[ 1

(d− 1)(d+ 2)

− 2χ(t)

d2(d− 1)(d+ 3)+

Ξ(t)

d2(d− 1)(d+ 1)(d+ 2)(d+ 3)

]

(4.45)

From this expression, we must subtract the squared average to find thevariance, 4.21. From 4.16,

∣∣∣∣∣∫U(d)

dV ρS(t)

∣∣∣∣∣2

=

∣∣∣∣(χ(t)− 1

d2 − 1

)|1S〉 〈1S |+

(dB(d2 − χ(t))

d(d2 − 1)

)1S

∣∣∣∣2 =[(χ(t)− 1

d2 − 1

)2

+ 2

(χ(t)− 1

d2 − 1

)(dB(d2 − χ(t))

d(d2 − 1)

)]|1S〉 〈1S |+

(dB(d2 − χ(t))

d(d2 − 1)

)2

1S =[d(χ(t)− 1

)2+ 2dB

(χ(t)− 1

)(d2 − χ(t)

)d(d2 − 1)2

]|1S〉 〈1S |+

(dB(d2 − χ(t))

d(d2 − 1)

)2

1S

(4.46)

Symbolically, the structure of this variance is:

Var(ρS) =

〈1S | 〈gS ||1S〉 ♣♥♥♠♦ ♥♠ . . . ♥♠ ♥♠

♥♠ ♥♦ . . . ♥ ♥...

.... . .

......

|kS〉 ♥♠ ♥ . . . ♥♦ ♥♥♠ ♥ . . . ♥ ♥♦

(4.47)

Composition of symbols simply means addition. The legend is the following:On the (1, 1) position:

48

Page 50: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

♣ ∼=Ξ(t)− 4χ(t)

d2(d− 1)(d+ 3)− χ2(t)− 2χ(t) + 1

(d2 − 1)2− 2dB

[d2 + χ2(t)− d2χ(t)− χ(t)

d(d2 − 1)2+[

1

(d+ 1)(d− 1)(d+ 2)− χ(t)(d2 + d+ 2)

d2(d+ 1)(d− 1)(d+ 3)+

Ξ(t)

d2(d− 1)(d+ 2)(d+ 3)

](4.48)

The first row and column feature:

♠ ∼= −1

(d+ 1)(d− 1)(d+ 2)+

χ(t)(d2 + d+ 2)

d2(d+ 1)(d− 1)(d+ 3)− Ξ(t)

d2(d− 1)(d+ 2)(d+ 3)(4.49)

On the diagonal, we have:

♦ ∼= d2B

[1

(d− 1)(d+ 2)− 2χ(t)

d2(d− 1)(d+ 3)+

Ξ(t)

d2(d− 1)(d+ 1)(d+ 2)(d+ 3)

− d4 − 2d2χ(t) + χ2(t)

d2(d2 − 1)2

](4.50)

Finally, every element of the variance has a contribution:

♥ ∼= dB

[1

(d− 1)(d+ 2)− 2χ(t)

d2(d− 1)(d+ 3)+

Ξ(t)

d2(d− 1)(d+ 1)(d+ 2)(d+ 3)

](4.51)

It is a straightforward exercise to check that the variance among the initialstates vanishes, by plugging in t = 0 into all dependent functions. For this weuse Ξ(0) = d4 + 2d3 + d2 and χ(0) = d2. After this, it is interesting to considerthe time average of the terms. First we will average of the time dependentterms. As before 〈χ(t)〉t = d. Its average square is not d2, though.

〈χ2(t)〉t =

⟨d+∑j<m

2 cos((Em − Ej)t)

2⟩t

=

⟨d2 + 4d

∑j<m

cos((Em − Ej)t) + 4∑

j<m,l<h

cos((Em − Ej)t) cos((El − Eh)t)

⟩t

= d2 + 4 · 0 + 4∑j<m

⟨cos2 ((Em − Ej)t)

⟩t

= d2 + 4 · d2 − d

2· 1

2= 2d2 − d

(4.52)

49

Page 51: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

In the last line, we used the fact that only the terms with (j, l) = (m,h)don’t average to zero, instead giving the average of the cosine squared, whichis half the cosine squared plus sine squared. Note there are (d2 − d)/2 termsj < m.

The following are straightforward, for only zero-frequency terms contribute:

〈ξ+(t)〉t = 〈ξ−(t)〉t = 〈ζ(t)〉t = d (4.53)

Then:

〈Ξ(t)〉t = 〈χ2(t) + ξ+(t) + ξ−(t) + ζ(t)〉t = 2d(d+ 1) (4.54)

Seen by 4.52, 4.44, and linearity of expectations. Substituting these into4.48 - 4.51, we find, the time average elements and their scaling limits:

〈♣〉t =−3d2 + d− 2 + 2dB(2d3 − 7d+ 1)

d(d− 1)(d+ 3)(d+ 1)2−−−→d→∞

4dBd2

〈♠〉t =d− 1

d(d+ 3)(d+ 1)−−−→d→∞

1

d2

〈♦〉t = −d2B

d3 + d2 + 6d+ 12

d(d− 1)(d+ 2)(d+ 3)(d+ 1)2−−−→d→∞

−d2B

d3

〈♥〉t =dB

d(d+ 3)−−−→d→∞

dBd2

(4.55)

One might be troubled by the negative expectation for ♦ on the diagonal,but it will be compensated by the larger, positive value on all elements ♥. Thevariance on the diagonal is thus expected to be lower than elsewhere. Moreover,the variance tends to the zero matrix as we increase the bath size, confirming in-tuition. The same is true when we increase the subsystem dimension. Although,enlarging the subsystem simply spreads occupation over a larger number of el-ements. The resulting lowered variance should not come as a surprise. Thevariance tends to zero faster than the average itself. The standard deviation, asthe square root, has comparable scaling behavior with the average.

Tentative numerical Monte-Carlo estimates of the average 4.16 and variance4.47 appear to corroborate these analytic results, though convergence is slow,and checking for many different times is laborious. For that reason, we suggestrigorous numerical validation as a good point of future study.

4.2 Mean Subsystem Dynamical Purity

The last quantity we will consider in this chapter, though a pleasantly insightfulone, is the mean subsystem purity, defined in 2.27. Like the entropy, it quantifiesdisorder, but unlike entropy, it is not calculated through diagonalization of thedensity matrix. The derivation is reminiscent of the standard deviation: it is

50

Page 52: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

also fourth order in elements of V . From 4.5 we contract the inner and outersubsystem indices in order to find Tr(ρ2

S(t))

γ(t) :=

∫U(d)

dV tr(ρ2S(t)) =

∫U(d)

dV

dB∑kB ,gB=1

dS∑kS ,gS=1

d∑j,l,m,h=1(

V(kS ,kB)jV1lV†j1V

†l(gS ,kB)e

i(El−Ej)tV(gS ,gB)mV1hV†m1V

†h(kS ,gB)e

i(Eh−Em)t

)=

∫U(d)

dV

dB∑kB ,gB=1

dS∑kS ,gS=1

d∑j,l,m,h

V(kS ,kB)jV1lV(gS ,gB)mV1h

V †j1V†l(gS ,kB)V

†m1V

†h(kS ,gB)e

i(El−Ej+Eh−Em)t

(4.56)

From this expression, we again compose the indices needed for Collins’ inte-gral, 4.6. Analogous to 4.23 observe:

I = ((kS , kB), 1, (gS , gB), 1) ; I ′ = (1, (gS , kB), 1, (kS , gB))J = (j,m, l, h); J ′ = (j,m, l, h)

(4.57)

Again the multi-indices decouple from the full indices. We can constructnew Rσ, Qτ , which are similar to those used in the variance, although for scalarpurity, these are all scalar expressions.

Rσ :=

dB∑kB ,gB=1

dS∑kS ,gS=1

δI,σ(I′), (4.58)

Qτ :=

d∑j,m,l,h=1

δJ,τ(J′)ei(El+Eh−Ej−Em)t (4.59)

For the results, see table 4:

Table 4: Tr(ρ2S(t)) Sum Factors

σ, τ ∈ S4 Rσ Qτ σ, τ ∈ S4 Rσ QτId 1 χ2(t) (12) dB dχ(t)

(123) dB χ(t) (13) 1 ξ+(t)(132) dS χ(t) (14) dS dχ(t)(124) dB χ(t) (23) dS dχ(t)(142) dS χ(t) (24) 1 ξ−(t)(134) dB χ(t) (34) dB dχ(t)(143) dS χ(t) (1234) d · dB d(234) dB χ(t) (1243) dB d(243) dS χ(t) (1324) dS d

(12)(34) d · dB d2 (1342) dB d(13)(24) 1 ζ(t) (1423) dS d(14)(23) d · dS d2 (1432) d · dS d

51

Page 53: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Due to the difference in signs in the exponent (there is no complex conjugatein 4.56 as there is in 4.22), the Qτ column has a slightly different order thanin table 2. Besides this, the derivation is exactly the same as in the previoussection.

We now treat the case of Rσ. A look at 4.57 tells us any permutation taking(1, 3) 7→ (1, 3) forces all indices to 1, and the sum in 4.58 is trivially unit, forσ ∈ Id, (13), (24), (13)(24). Any that maps 1 7→ 2 ∨ 3 7→ 4, will leave a singlefree bath index, producing Rσ = dB after summation, for σ ∈ (12), (34), (123),(124), (134), (234), (1243), (1342). Conversely, for a single free subsystem index,1 7→ 4 ∨ 3 7→ 1: Rσ = dS when σ ∈ (14), (23), (132), (142), (143), (243), (1324),(1423). Lastly, mapping 1 7→ 4∧3 7→ 2 leaves both subsystem indices undeter-mined, and equates the bath indices to each other, killing one summation. ThenRσ = d2

SdB = d · dS for σ ∈ (14)(23), (1432). Likewise, Rσ = dSd2B = d · dB

when the bath indices are both free, and subsystem indices are equated, inσ ∈ (12)(34), (1234). Remembering 4.41 and 4.56, and table 3, find the nu-merator:

∑σ,τ∈S4

RσQτA4(d, στ−1) = (Ξ(t)− 4χ(t))[d4 − 2d3 − 7d2 + 8d+ 12

]+ (dS + dB)

·[d7 − d6 − 13d5 + 13d4 + 36d3 − 36d2 − (Ξ(t)− 4χ(t))

(d3 − 3d2 − 4d+ 12

) ]= Γ(t)(d+ 1)(d− 2)(d+ 2)(d− 3) + (dS + dB)

·[d2(d− 1)(d− 2)(d+ 2)(d− 3)(d+ 3)− Γ(t)(d− 3)(d+ 2)(d− 2)[

](4.60)

In the last line, we have defined another function Γ(t) := Ξ(t) − 4χ(t) tohold the time and energy dependence.

Γ(t) :=

d+ 2∑j<k

cos ((Ej − Ek)t)

2

+ 2∑j,k,l

cos ((2Ej − Ek − El)t)

+ 2∑j<k

(cos (2(Ej − Ek)t)− 4 cos ((Ej − Ek)t)

)− 3d

(4.61)

See figure 4.3 and that beside it for typical forms of this function.Though larger in absolute sense, Γ(t) drops like χ(t). Commencing at its

maximal value, when all phases of the cosines coincide at zero, it equilibratesaround a value approximately the maximal value’s square root, when the cosinesdecohere. This transition again takes place after a duration empirically close toσ−1E and can be thought of as moving from a ’known’ to ’unknown’ phase. For

the function to be ’well behaved’, i.e. have a damped character, considerablyfewer dimensions are needed than with χ(t).

52

Page 54: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Figure 4.3: The function Γ(t), gen-erated with d = 5 random, mutuallyrepulsive energies Ej drawn from[−π, π].

Figure 4.4: The function Γ(t), gen-erated with d = 20 such energies.The time t = σ−1

E is indicated onthe horizontal axis.

At present the origin of the terms (d±z) in expressions such as 4.60 is unclearto the author, however their appearance is fortuitous: they cancel neatly withthe denominator. Dividing 4.60 by B4(d) from 4.41, we find the mean (uniformlyover systems) time dependent purity, for a system of d dimensions, subsystemand bath of dS and dB dimensions:

γ(t) =Γ(t)

d2(d− 1)(d+ 3)

(1− dS + dB

d+ 1

)+dS + dBd+ 1

(4.62)

This function is symmetric in S and B. This makes sense. We are remindedthat the nonzero eigenvalues of the reduced density matrices ρS and ρB areidentical, a fact derived from the full system pure state’s Schmidt decomposi-tion. The purity is only a function of the eigenvalues: γ(WρSW

†) = γ(ρS) forany basis transformation W ∈ U(dS). Also the entanglement entropy of bothsubsystem and bath are equal [14].

At t = 0, Γ(t) = d2(d−1)(d+3), so γ(0) = 1, as all systems start out pure byassumption. Furthermore, for all times, when dB = d or dS = d, the other beingunity (so dBdS = d), there is no tracing, and no loss of information. Indeed,also then γ(t) = 1. Finally, the time average value of Γ(t) is 〈Γ(t)〉t = 2d(d−1),substituting this linearly gives us the time average, mean subsystem purity:

〈γ(t)〉t =2

d(d+ 3)

(1− dS + dB

d+ 1

)+dS + dBd+ 1

(4.63)

Compare this to the average purity obtained by taking a random vector froma d = dB · dS dimensional Hilbert space, and tracing out the dB dimensions ofthe bath [42]:

〈γ〉dS ,dB =dS + dBd+ 1

(4.64)

More interesting statistics of systems distributed according to these so calledtrace measures can be found in e.g. [14].

53

Page 55: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

The minimum purity is found by minimizing dS+dBd+1 over dS , achieved for

dS = dB =√d. So the most uncertainty occurs for an equal partition. This may

be interpreted as a compromise between a larger bath taking more information,and a larger subsystem allowing for more possible states over which to spreadprobability. Finally, substituting dB = dS into 4.63, we find the time averageminimal purity and its limit:

〈γ(t)〉t∣∣∣∣dB=dS

=2(d5S + 3d3

S + d2S − 2dS + 1

)d2S (d2

S + 1) (d2S + 3)

−−−−→dS→∞

2

dS(4.65)

This scaling behaviour in the large dimension limit is twice minimum purityoverall, 1/dS , see 2.28. The results in 4.19 would suggest that the mixingtends to the the minimum purity, however this is for a fixed dS . Indeed inthat case, taking dB → ∞ in 4.63 reduces it to 〈γ(t)〉t → 1/dS . There is nocontradiction here: for a larger subsystem, 2/dS is soon smaller than 1/dS of asmall subsystem.

Using the same techniques, other polynomial functions of the reduced densitymatrix can be averaged over all systems. We have simply treated the mostobvious candidates.

54

Page 56: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Chapter 5

Quantum DifferentialGeometry

One promising concept, and a natural one to the field of quantum geometry, isto treat Hilbert space as a Riemannian manifold. This entails imbuing the spacewith a metric, and defining differential separation between states. Polkovnikovand Kolodrubetz [25] use a metric approach to thermalization. Understandingthe behavior of the metric -and its shortcomings- in the the thermodynamiclimit is important when taking the geometric line of attack to this problem.Lightly motivated by that theme, we find that the transformation properties ofthese quantum geometric objects are interesting in their own right. The focus ofthis chapter will be predominantly the transformation properties of geometricobjects under gauge symmetries.

This chapter is somewhat disconnected from the narrative of the thesis. Forinstance, it is not dependent on the bath/subsystem division ubiquitous above.Nonetheless, the author spent considerable time studying quantum metrics, soit is at least instructive to investigate.

5.1 Manifold of Quantum States

Consider, in all generality, a family of Hamiltonians H(λ), acting on a d-dimensional Hilbert space H spanned by the energy spectral basis |ψj(λ)〉.Here λµ ∈ M is a vector of parameters controlled by the experimenter, influ-encing the instantaneous spectrum. The dependence on λ will only be madeexplicit when it serves the clarity. What is λ physically? Think of actual dialssuch as magnetic field or position of some barrier.

5.1.1 Nondegenerate Case

For now, we take the system in an energy eigenstate |ψn〉 such as the groundstate. The case of the nth excited state is analogous, as additionally, we assume

55

Page 57: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

there are no level crossings anywhere on M, i.e. for:

H |ψj〉 = Ej |ψj〉 , Ej = Ek ⇔ |ψj〉 = |ψk〉 (5.1)

So this state can unambiguously be translated aroundM to the same level.Later, we will generalize to degenerate states and then density matrices.

We wish to define the Quantum Geometric Tensor (QGT) from [15] to nav-igate M in a gauge invariant way, on the state |ψn〉 ∈ H. It relates an innerproduct on M, the ’overlap’ between the rate of change of the state for an in-finitesimal step in parameter λµ, and that in parameter λν . Following Provostand Vallee, we begin with the instinctive ansatz of the differential:

||ψn(λ+ dλ)− ψn(λ)|| =(∂µ 〈ψn| dλµ

)(∂ν |ψn〉 dλν

)(5.2)

Summation convention implied, and using the shorthand operators ∂µ :=∂λµ . This leads us to investigate the metric-like expression below. Taking thedifferential operators inside the bras and kets:

Qµν := 〈∂µψn|∂νψn〉 (5.3)

We also suppress of the explicit dependence of |ψn〉 on λ. In the nondegen-erate case, the gauge symmetry of the Hilbert space is the U(1) invariance ofglobal phase, as all states technically live in a projective Hilbert space CPd−1, andphysical observables are unaffected by which point of the ray (c |ψn〉, c ∈ C)the state occupies [14]. On our manifold, this symmetry amounts to invarianceof the system under multiplication of a phase, which need not be the same atall points λ. For some α :M→ R:

|ψn〉 7→ |ψ′n〉 := eiα(λ) |ψn〉 (5.4)

i.e. these two kets define the same point on the manifold of rays. For anyobservable O:

〈ψ′n| O |ψ′n〉 = 〈ψn| e−iαOeiα |ψn〉 = 〈ψn| O |ψn〉 (5.5)

The local expectation value is the same. However, the tensor is not fullylocal but involves adjacent points. It transforms under this transformation as:

Qµν 7→ Q′µν = 〈∂µψn|∂νψn〉 − i∂µα 〈ψn|∂νψn〉+ i∂να 〈∂µψn|ψn〉+ ∂µα∂να 〈ψn|ψn〉= Qµν + ∂µαβµ + ∂ναβ

∗ν + ∂µα∂να

(5.6)

Using Leibniz repeatedly in the first step, and invoking the normalization of|ψn〉 in the second. We have defined βµ := −i 〈ψn|∂µψn〉. We gather that theexponentials cancel, but the exponents don’t.

Another consequence of the normalization is the following identity:

1 ≡ 〈ψn|ψn〉 ⇔ ∂µ 〈ψn|ψn〉 = 0 = 〈∂µψn|ψn〉+ 〈ψn|∂µψn〉 (5.7)

56

Page 58: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

But also:

〈∂µψn|ψn〉∗ = 〈ψn|∂µψn〉 (5.8)

So it follows that

〈ψn|∂µψn〉 = −〈ψn|∂µψn〉∗ ⇔ βµ ∈ R (5.9)

Note that we don’t assume differential expressions such as |∂νψn〉 normalized.In order to make the tensor gauge invariant, we observe that under 5.4, βµtransforms as:

βµ 7→ β′µ = βµ + ∂µα 〈ψn|ψn〉 = βµ + ∂µα (5.10)

So a gauge invariant expression denoted by curly Qµν , that will not besensitive to the particular choice of phase, is given by

Qµν := Qµν − β∗µβν = 〈∂µψn|∂νψn〉 − (−i2) 〈∂µψn|ψn〉 〈ψn|∂νψn〉 (5.11)

Where the star (∗) serves to remind that we conjugate the bracket in thedefinition of βµ. A way to look at this gauge invariant result, is taking the innerproduct in Qµν in the subspace in which we have projected out the instantaneousstate |ψn〉:

Qµν = 〈∂µψn|(1− |ψn〉 〈ψn|

)|∂νψn〉 (5.12)

Now we will find an expression for this tensor in terms of H. In order to doso, we first invoke the spectral theorem, writing the terms |∂µψn〉 in the basis|ψj〉.

|∂µψn〉 =∑j

cjµ |ψj〉 :=∑j

〈ψj |∂µψn〉 |ψj〉 (5.13)

Inserting this into 5.12, we find:

Qµν =∑j,k

(cjµ)∗ 〈ψj | (1− |ψn〉 〈ψn|) ckν |ψk〉 =∑j,k 6=n

(cjµ)∗ckνδjk =∑j 6=n

(cjµ)∗cjν

(5.14)Armed with the expression 5.14, we calculate the coefficients cjµ in terms of

energy. Our starting point is to differentiate the time-independent Schrodingerequation with respect to λµ, and take the inner product with a 〈ψj | , j 6= n.Essentially we are copying the Feynman-Hellman theorem [43].

∂µ (H |ψn〉) = ∂µ (En |ψn〉)∂µH |ψn〉+H |∂µψn〉 = ∂µEn |ψn〉+ En |∂µψn〉 ⇒

〈ψj | ∂µH |ψn〉+ 〈ψj |H |∂µψn〉 = 〈ψj | ∂µEn |ψn〉+ 〈ψj |En |∂µψn〉〈ψj | ∂µH |ψn〉+ 〈ψj |Ej |∂µψn〉 = 〈ψj |En |∂µψn〉

(5.15)

57

Page 59: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

In the last line we used the hermicity of H and the fact that the scalar ∂µEnis effectively diagonal between orthonormal eigenstates, killing one term. Theresult is the following identity:

cjµ ≡ 〈ψj |∂µψn〉 =〈ψj | ∂µH |ψn〉En − Ej

(5.16)

From 5.14:

Qµν =∑j 6=n

〈ψn| ∂µH |ψj〉 〈ψj | ∂νH |ψn〉(En − Ej)2

(5.17)

In 5.12 we observe that exchanging µ ↔ ν amounts to swapping the braand ket, or Qµν = Q∗νµ is Hermitian in its indices. Because we will use thisexpression as a metric in the contraction Qµνλµλν with the symmetric λµλν ,we may also use the symmetrized part of the metric:

gµν := Re(Qµν) =1

2(Qµν +Qνµ) =

1

2

(Qµν +Q∗µν

)(5.18)

This is as far as the nondegenerate case takes us. However, as alludedto at the start of the chapter, this object is employed in the study of ther-malization. In the thermodynamic limit, applicable to theories such as theEigenstate Thermalization Hypothesis, the energy difference between closebystates can become arbitrarily small, from a macroscopic point of view. Roughly,(En − Ej) ≈ e−S ≈ 1/d, so nearby energy differences scale inversely with theHilbert space dimension. In contrast to subsection 2.4.1, here with S we meanthe extensive thermodynamic entropy [25]. This poses a problem in equation thedenominator of 5.17 (combined with the fact that the numerator scales inverselyat a lower pace with d). In short: the tensor blows up in the large-dimensionallimit. We regard this as a consequence of the effective degeneracy of the energystates on an energy shell En ∈ [E,E + δE], for instance defined by the micro-canonical ensemble. This motivates us to investigate the behavior of the QGTin the case of a degenerate, but still finite, system.

5.1.2 Degenerate Case

In the following, we will consider the case of calculating gµν on a state |ψn〉 ∈HEn , the subspace of eigenvectors corresponding to energy En. For the sake ofdefiniteness, assume a k-fold degenerate level, with m− 1 linearly independentenergy eigenvectors existing with energies < En. Thus with any partial ordering,basis vectors m through (m + k − 1) have energy En. There are no crossingsof levels on M. I.e. the somewhat strong assumption is that this degeneracy isindependent of λ.

The Hamiltonian offers us no preferred basis with which to span HEn . Thusthe choice of basis, which can be made locally in M, is in essence a gaugechoice, and each choice is related to each other by a U(k) rotation. We haveseen above when the global phase of the state was a gauge choice. The subtlety

58

Page 60: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

is, contrary to the nondegenerate case, there can be two different global statesof the system |φ1(λ)〉 , |φ2(λ)〉 ∈ HEn at the same point λ, such that some otherobservable O 6= H could distinguish between the two: 〈φ1| O |φ1〉 6= 〈φ2| O |φ2〉.Just take orthogonal basis states 〈φ1|φ2〉 = 0 and define O := |φ1〉 〈φ1|. Ois Hermitian so it is an observable, and both states are eigenstates of O, but〈φ1| O |φ1〉 = 1, 〈φ2| O |φ2〉 = 0. This is by construction not the case with theU(1) phase, for within one energy eigenspace there can be no two normalizedorthogonal states.

Having mentioned this caveat to calling U(k) a gauge symmetry, we make anobservation. Define |φl(λ)〉, l ∈ 1, . . . , k a local orthonormal basis of HEn .This allows us to write an arbitrary eigenstate with energy En as:

|ψn〉 =∑l

cl |φl〉 ,∑l

|cl|2 = 1 (5.19)

Let us start from the time-independent Schrodinger equation again:

H |ψn〉 = En |ψn〉

∂µ

(H∑l

cl |φl〉

)= ∂µ

(En∑l

cl |φl〉

)∑l

(∂µHcl |φl〉+H∂µcl |φl〉+Hcl |∂µφl〉) =∑l

(∂µEncl |φl〉+ En∂µcl |φl〉+ Encl |∂µφl〉)

(5.20)

Now we take the inner product with a single basis state 〈φa|, and move termsaround to facilitate notation.

〈φa|H∂µca |φa〉+∑l

(〈φa| ∂µHcl |φl〉+ 〈φa|Hcl |∂µφl〉

)=

〈φa|(En∂µca + ∂µEnca

)|φa〉+

∑l

〈φa|Encl |∂µφl〉(5.21)

Using hermicity of H, we cancel the first term on the LHS of 5.21 with thefirst term on the RHS, as well as the second part of the LHS sum against thesum on the RHS.

∑l

〈φa| ∂µHcl |φl〉 = 〈φa| ∂µEnca |φa〉

ca =∑l

cl〈φa| ∂µH |φl〉

∂µEn

δla =〈φa| ∂µH |φl〉

∂µEn

(5.22)

59

Page 61: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Where the last step is inferred as it must hold for arbitrary cl. In otherwords, the upshot is that the derivative of H does not appear to mix the de-generate states. This is a consequence of the assumption that the degeneracy isconstant on M.

We wish to build up the machinery as we did in the previous section. Inthe degenerate case, however, the gauge freedom isn’t generated by an abelianphase α(λ) introduced in 5.4, but a non-abelian matrix valued phase A. Thedegenerate case cannot be completed in the manner shown in 5.11, there is noway to define a ’matrix βµ’ that can be added to the ansatz, to make it gaugeinvariant. Observe that 〈ψn|∂µA∂νA|ψn〉 6= 〈ψn|∂µA|ψn〉 〈ψn|∂νA|ψn〉 if A is anoncommuting operator. The framework of projectors in equation 5.12 is moreuseful. We will transition to a different language.

5.2 Linear Algebraic Formulation

At this point we elaborate on an alternative formulation and derivation. In thespirit of Quantum Mechanics as a form of Linear Algebra, define a ket-vectoras a normalized column vector in Cd:

|ψ〉 ∼

c1c2...cd

,∑j

|cj |2 = 1 (5.23)

These hold the coefficients of each linearly independent basis vector of somebasis of H. In the nondegenerate case, we naturally take the ordered basis ofenergy eigenstates. The degenerate case admits a partial ordering. Operators Oare again d× d normal matrices (satisfying O†O = OO†) [30], and a bra vectoris a complex conjugate row vector.

〈ψ| = |ψ〉† =(c∗1 c∗2 . . . c∗N

),∑j

|c∗j |2 = 1 (5.24)

Which is evidently still normalized. The derivative |∂µφ〉 is the element-wisederivative of the entries of the vectors. Now we consider the construction of theQGT.

5.2.1 Nondegenerate case

Let us briefly repeat the nondegenerate case. Here, all energy levels are distinct,and our state initially occupies one of them, the nth one, with energy En. Thisstate is equivalent to a vector |ψn〉 with cj = c · δjn, for some c ∈ C satisfying|c|2 = 1. Though the elements outside the nth place are zero by construction,their derivatives ∂µ need not be.

60

Page 62: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

|ψn〉 =

0...0c0...0

nth position (5.25)

The description is not unique. An additional global (in H) phase may beapplied locally (in M):

|ψn〉 7→ (u(λ) · 1) |ψn〉 = |ψ′n〉 (5.26)

For eiα(λ) = u(λ) ∈ U(1) an arbitrary element of the symmetry group of thewave vector. In this formalism, it is multiplied by the d×d identity: multiplyingall components by the same phase has no effect on the physics. Then, following[44] we wish for this u to also not feature in our expression for Qµν .

Qµν := 〈∂µψ′n|∂νψ′n〉 = ∂µ(〈ψn|u†)∂ν(u |ψn〉) (5.27)

I.e. for the tensor to only depend on the coefficients cj , and their derivatives,in combinations such as ∂•c∗j∂

•cj so that also the particular U(1) phase chosen inthe individual basis vectors cancels out. In order to find a suitable definition, weexpand the expression in 5.27 and substitute the generator: u = eiα, u† = e−iα.

〈∂µψ′n|∂νψ′n〉 = (〈∂µψn|u† + 〈ψn| ∂µu†)(∂νu |ψn〉+ u |∂νψn〉)= (〈∂µψn| e−iα − 〈ψn| i∂µαe−iα)(i∂ναe

iα |ψn〉+ eiα |∂νψn〉)= (〈∂µψn| − 〈ψn| i∂µα)(i∂να |ψn〉+ |∂νψn〉)

(5.28)

From this expression, it is clear that the effect of the phase u, or its generatorα, lives only in the subspace spanned by the nth basis vector. If we modify ourexpression of the QGT, to project out this subspace inside the inner product,the phase will no longer feature. Define:

P 1n := 1− |ψn〉 〈ψn| =

1 0 · · ·0 1 · · ·...

.... . .

· · · 0 0· · · 0 0. . .

......

10

1...

.... . .

0 0 · · ·0 0 · · ·

. . ....

...· · · 1 0· · · 0 1

(5.29)

61

Page 63: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

The nomenclature references one (1) dimension being projected out, namelythe nth. Then if we define our new QGT as follows:

Qµν := 〈∂µΨn|P1|∂νΨn〉 =∑j 6=n

∂µc∗j∂νcj (5.30)

Where the cj are specifically the elements in |ψn〉, we have recovered thegauge invariant form.

5.2.2 Degenerate Case

In this language, the generalization to the degenerate case is straightforward.The occupied wave vector, which we will still dub |ψn〉, is now subtle. Thepositions corresponding to the dimensions of the degenerate subspace HE0

arethe only ones that are occupied, the rest vanishing. Formally: cj = 0 ∀j /∈m,m+ 1, . . . ,m+ k − 1, and the remaining cl, l ∈ m,m+ 1, . . . ,m+ k − 1satisfy the normalization condition

∑l |cl|2 = 1. Conventionally, l will label the

degenerate dimensions. Then our state has the form:

|ψn〉 =

0...0cmcm+1

...cm+k−1

0...0

(5.31)

The particular distribution of the cl depends on the choice of basis used tospan HE0

, and a priori the system (defined by H) has no preferred basis. Wewish for also Qµν to reflect this, and not depend on the basis. We must howeverstart from some arbitrary basis of all of H, defined locally at each λ. A subsetof this, denoted |φl〉, are the basis vectors that span the subspace HEn , sothat |ψn〉 =

∑l cl |φl〉.

Then any other k-dimensional basis |φ′l〉 is obtained from this one by meansof a unitary basis transformation W ∈ U(k). With intuitive abuse of summedindex h: |φ′j〉 =

∑hWjh |φh〉. To have this work on the entire basis, we must

augment our transformation, in a manner that leaves all the other dimensionsinvariant. To this end, W is nestled in a larger matrix U ∈ U(d), so that the1st row of W is in the mth row of U , and likewise for columns. Outside thedimensions m, . . . ,m + k − 1, U is the identity matrix. This way, U actsnontrivially only upon the degenerate subspace. The resulting block diagonalstructure is the following:

62

Page 64: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

U =

1 0 · · ·0 1 · · ·...

.... . .

· · · 0 0· · · 0 0. . .

...... W

...

.... . .

0 0 · · ·0 0 · · ·

. . ....

...· · · 1 0· · · 0 1

(5.32)

This is a generic element of the gauge symmetry group of our system. Wemay express the state equally well as:

|ψ′n〉 = U |ψn〉 (5.33)

So long as U , which may vary from point to point in M, is always of theform in 5.32, it it is simply a matter of representation, and in particular thechoice should not effect the value of Qµν .

The attentive reader may point out that we must also consider a scalarglobal phase u as in the nondegenerate 5.26 multiplying the entire expression5.32. However, it will appear, fixing this U symmetry results in somethingmanifestly invariant under the global phase u as well.

Now we again consider the ansatz

Qµν := 〈∂µψ′n|∂νψ′n〉 = (〈∂µψn|U† + 〈ψn| ∂µU†)(∂νU |ψn〉+ U |∂νψn〉) (5.34)

And mimic the derivation above, although we must be more careful due tothe noncommuting nature of U , not a multiple of the identity. We may stillsubstitute U in exponential form, generated by a d × d Hermitian matrix Athusly: U = eiA. In our particular case, we make use of the Y generatingW = eiY in k dimensions, to find the form:

A =

0 · · ·...

. . .

· · · 0. . .

... Y

...

. . .

0 · · ·

. . ....

· · · 0

(5.35)

From the definition, (eiA)† = ei∗A† = e−iA, ∂µe

iA = i∂µA·eiA, and ∂µ(U†) =(∂µU)†, because λ is real. Substituting these identities:

Qµν = (〈∂µψn| e−iA − 〈ψn| ie−iA∂µA†)(i∂νAeiA |ψn〉+ eiA |∂νψn〉) (5.36)

63

Page 65: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

The crucial observation is that the elements of A outside the k columnsand rows of HE0 are always zero, or constant, for any λ: at any point in M,the degeneracy only allows us to mix these k basis vectors. This means thatthe same is true of the derivative ∂µA. This matrix has the same form as A,nonvanishing elements only in the diagonal block between rows and columns mand m + k − 1. For a Qµν independent of A, we must find a projector in thebracket that will exactly kill the terms involving A. A little thought assuresus that the smallest amount of information we can project out, is exactly thesecolumns and rows. We posit the projector P kn :

P kn := 1−∑l

|φl〉 〈φl| =

1 · · ·...

. . .

· · · 0. . .

...1 ∅

1

.... . .

0 · · ·

. . ....

· · · 1

(5.37)

In 5.37 the d×d identity matrix is modified with k zeroes (∅) on the diagonalelements m through (m+ k − 1). The following are immediate:

P knU = P kn = UP kn = (P kn )2, P kn∂µA = ∅ = ∂µAPkn (5.38)

Then, modifying 5.34, we propose:

Qµν := 〈∂µψ′n|Pk|∂νψ′n〉= (〈∂µψn| e−iA − 〈ψn| ie−iA∂µA†)Pk(i∂νAe

iA |ψn〉+ eiA |∂νψn〉)= (〈∂µψn|Pk − 〈ψn| ie−iA∅)(i∅eiA |ψn〉+ Pk |∂νψn〉)

= 〈∂µψn|Pk|∂νψn〉 =∑

j<m,j≥m+k

∂µc∗j∂νcj

(5.39)

Which is again manifestly gauge invariant, also for global QM phase u. Inconclusion we gather that for the QGT evaluated at some energy En, the entiresubspace corresponding to that energy must be projected out, be it one or k-dimensional.

5.3 Curvature Operator

In this section, we take a detour to calculate the general form of the Berry curva-ture, after Michael Berry [16], valid for both the degenerate and nondegeneratecases. Following [45], we define this curvature as:

64

Page 66: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Fµν := 2 · Im(Qµν) = Qµν −Qµν (5.40)

We wish to show how this curvature is formed by the relevant connection,abelian or not. To this end, we assume we are in a state with energy En definedby the initial condition ~b ∈ CN with |~b| = 1. In the nondegenerate case, the

elements of ~b satisfy bj = δn,j . In the degenerate case, the nonzero components

of~b are spread over the kn basis vectors that spanHEn . Then, initially, our state(dropping the index n) is given by |ψ〉 =

∑j bj |φj〉, and trivially bj = 〈φj |ψ〉.

It is important to note that in this section the bj give the initial distributionamong basis vectors, but they do not vary with λ.

The abelian Berry connection operator, or non-abelian Wilczek-Zee [17] con-nection living in Hilbert-Schmidt space, are jointly defined as follows:

Aµ :=∑j,k

|φj〉 〈φj |∂µφk〉 〈φk| , Ajkµ = 〈φj |∂µφk〉 (5.41)

Here, summation is only over the occupied (degenerate) subspace, otherelements are zero. This construction is sensible, for the relation:

〈ψ|∂µψ〉 =

∑j

b∗j 〈φj |

∂µ

(∑k

bk |φk〉

)

=∑jk

〈ψ|φj〉 〈φj |∂µφk〉 〈φk|ψ〉

= 〈ψ|Aµ |ψ〉

(5.42)

In this spirit, the full Fµν can be expressed as an observable in terms of Aµ.From 5.40:

Fµν = 〈∂µψ|

(1−

∑l

|φl〉 〈φl|

)|∂νψ〉 − 〈∂νψ|

(1−

∑l

|φl〉 〈φl|

)|∂µψ〉

= 〈∂µψ|∂νψ〉 − 〈∂µψ|∂νψ〉 −∑l

(〈∂µψ|φl〉 〈φl|∂νψ〉 − 〈∂νψ|φl〉 〈φl|∂µψ〉)

(5.43)

Where again the summation is over the (degenerate) occupied subspace only.Evaluating each antisymmetric half of this equation separately:

65

Page 67: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

〈∂µψ|∂νψ〉 − 〈∂µψ|∂νψ〉

=∑jk

(b∗j 〈∂µφj | bk |∂νφk〉 − b∗j 〈∂νφj | bk |∂µφk〉

)

= 〈ψ|

∑jk

|φj〉 [〈∂µφj |∂νφk〉 − 〈∂νφj |∂µφk〉] 〈φk|

|ψ〉= 〈ψ|

∑jk

|φj〉 [〈∂µφj |∂νφk〉+ 〈φj |∂µ∂νφk〉 − 〈∂νφj |∂µφk〉 − 〈φj |∂ν∂µφk〉] 〈φk|

|ψ〉= 〈ψ| (∂µAν − ∂νAµ) |ψ〉

(5.44)

We anticipated the product rule in the 3rd equality, and used the commutingattribute of derivatives. Next,

−∑l

(〈∂µψ|φl〉 〈φl|∂νψ〉 − 〈∂νψ|φl〉 〈φl|∂µψ〉)

= 〈ψ|

∑jlk

|φj〉 [〈∂νφj |φl〉 〈φl|∂µφk〉 − 〈∂µφj |φl〉 〈φl|∂νφk〉] 〈φk|

|ψ〉= 〈ψ|

(∑jlmk

|φj〉 〈∂νφj |φl〉 〈φl|φm〉 〈φm|∂µφk〉 〈φk| −

|φj〉 〈∂µφj |φl〉 〈φl|φm〉 〈φm|∂νφk〉 〈φk|

)|ψ〉 = 〈ψ| (A†νAµ −A†µAν) |ψ〉

(5.45)Where we insert a summed over delta function in the second equality. Ex-

panding on this theme for the final step, we note that the local bases are alwayschosen such that

〈φj |φk〉 = δjk ⇔ ∂µ 〈φj |φk〉 = 〈∂µφj |φk〉+ 〈φj |∂µφk〉 = 0 (5.46)

In terms of the connection, Ajkµ = −Akjµ , or Aµ = −A†µ, is skew-Hermitian.Combining 5.43, 5.44 & 5.45, we find:

Fµν = 〈ψ| (∂µAν − ∂νAµ + [Aµ, Aν ]) |ψ〉 (5.47)

This is the correct, gauge covariant expression [46]. In the nondegeneratecase the commutator vanishes. The occupied subspace is then one-dimensional,and the connection is effectively scalar (Ajkµ is a matrix with a single nonzeroentry, on the diagonal).

There are other scenario’s to map out, such as pure states that are superpo-sitions of energy eigenstates, which in turn may or may not be degenerate. We

66

Page 68: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

will skip these, and move on to mixed states and their treatment as a Rieman-nian manifold.

5.4 Mixed States

Recall our discussion of density matrices ρ from subsection 2.3.1. In the fol-lowing, for convenience, we assume them diagonal in the energy eigenbasis, atleast where this is nondegenerate. This is not a strong assumption. Off diagonalterms resulting from superpositions of energy eigenstates with distinct energiesEj , Ek, will oscillate with a frequency (Ej−Ek)/2π, setting ~ = 1. Over time, itis impossible in practise to keep track of the exact phase, and the instantaneousvalue will be lost: the states decohere. This situation, with random, unknownrelative phases between the energy states, is physically indistinguishable fromthe diagonal ensemble. In a sense, ρ(t) precesses periodically around its timeaverage 〈ρ〉t, which is diagonal.

Inside degenerate subspaces, kp × kp matrix sub-blocks may appear on thediagonal, with kp the degeneracy of the pth energy level. On mixed states, thereis a natural generalization of the QGT, or specifically, its real part gµν .

5.4.1 Bures Metric

In this subsection, we will introduce a metric on the manifold of mixed statedensity matrices. It is called the Bures metric, after Donald Bures [20].

First we introduce the Bures distance dB , directly related to the fidelity [21],introduced in 2.30.

d2B(ρ, η) := 2− 2F (ρ, η) (5.48)

A metric results by taking the Bures distance between infinitesimally sepa-rated states ρ, η = ρ+ dρ. Then, the infinitesimal version becomes:

d2B(ρ, ρ+ dρ) =

1

2Tr(dρG), Gρ+ ρG = dρ (5.49)

Involving the Hermitian one-form only defined implicitly through the secondequation.

Mirroring the sections above, we assume there is some mapping from λ 7→ρ(λ) through the Hamiltonian. Then expand the Taylor series:

ρ+ dρ = ρ+ ∂τρdλτ +

1

2∂µ∂νρdλ

µdλν + . . . (5.50)

We will see that the first order term cancels due to the normalization con-dition on the basis states. Then the first nonvanishing dependence on dλ isquadratic, and we curtail the higher order terms. What is left is of the form ofa metric. We cite from [14]:

67

Page 69: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

d2B(ρ, ρ+ dρ) =

1

2Tr(∂µρLν)dλµdλν ,

ρLν + Lνρ

2= ∂νρ (5.51)

Here, generalizing G, Lν is the symmetric logarithmic derivative.In order to recover the link to the old metric gµν , we must take ρ = |ψ〉 〈ψ|

pure. We first use expression 2.29 for the fidelity on pure states, and notationallytake the differential d |ψ〉 inside the ket:

F(|ψ〉 , |ψ + dψ〉

)= | 〈ψ|ψ + dψ〉 | =

√〈ψ|ψ + dψ〉 〈ψ + dψ|ψ〉 (5.52)

With 〈ψ|ψ〉 = 1, we find:

F(|ψ〉 , |ψ + dψ〉

)=√

1 + 〈ψ|dψ〉+ 〈dψ|ψ〉+ 〈dψ|ψ〉 〈ψ|dψ〉 (5.53)

We now specify:

d |ψ〉 = |∂τψ〉 dλτ +1

2|∂µ∂νψ〉 dλµdλν + . . . (5.54)

Inserting 5.54 into 5.53, and keeping terms up to quadratic in dλ:

F(|ψ〉 , |ψ + dψ〉

)=

[1 +

(〈ψ|∂τψ〉+ 〈∂τψ|ψ〉

)dλτ+

(1

2〈ψ|∂µ∂νψ〉+

1

2〈∂µ∂νψ|ψ〉+ 〈∂µψ|ψ〉 〈ψ|∂νψ〉

)dλµdλν

] 12

=

[1 +

(〈∂µψ|ψ〉 〈ψ|∂νψ〉 −

1

2〈∂µψ|∂νψ〉 −

1

2〈∂νψ|∂µψ〉

)dλµdλν

] 12

(5.55)

In the last line, we used the following identities:

〈ψ|∂τψ〉+ 〈∂τψ|ψ〉 = ∂τ 〈ψ|ψ〉 = ∂τ1 = 0 (5.56)

and:

〈ψ|∂µ∂νψ〉+ 〈∂µ∂νψ|ψ〉+ 〈∂µψ|∂νψ〉+ 〈∂νψ|∂µψ〉 = ∂µ∂ν 〈ψ|ψ〉 = 0 (5.57)

From 5.55, we may un-symmetrize the negative terms using the symmetricdλµdλν , to obtain:

68

Page 70: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

F(|ψ〉 , |ψ + dψ〉

)=

[1 +

(〈∂µψ|ψ〉 〈ψ|∂νψ〉 − 〈∂µψ|∂νψ〉

)dλµdλν

] 12

=

[1 + 〈∂µψ|

(|ψ〉 〈ψ| − 1

)|∂νψ〉 dλµdλν

] 12

=

[1− gµνdλµdλν

] 12

≈ 1− 1

2gµνdλ

µdλν

(5.58)

Using 5.12 and 5.18, and the Maclaurin series of√

1− x = 1− 12x+ . . . Then

from 5.48, the relation is immediate:

d2B(|ψ〉 , |ψ + dψ〉) ≈ gµνdλµdλν (5.59)

This is one of the attractive qualities of the Bures metric. It reduces nicely tothe quantum geometric tensor in the special case of pure states, i.e. it is Fubini-Study adjusted. Also, it is Fisher adjusted, coinciding with the Fisher-Rao met-ric of classic probability distributions when restricted to diagonal ensembles.On non-diagonal density matrices, it additionally takes into account the purelyquantum phenomenon of noncommutativity between ρ and ρ+dρ [19]. Further-more, as long as both arguments are in the same one, it is independent of thebasis, through the that property of the fidelity.

5.4.2 Gauge Invariance of the Bures Metric

When constructing density matrices, all elements are the product of a bra and aket vector. Any abelian global phase is annihilated this way, so any function ofdensity matrices is automatically invariant under quantum U(1) global phase.The same can not be said for larger gauge symmetries.

Let us assume the spectrum is multiply degenerate, consisting of L energylevels En for n ∈ 1, 2, . . . , L. These levels have degeneracies kn, forcing∑n kn = d = dim(H). Construct ρ in an energy eigenbasis, which of course is

not unique. An element of the total gauge symmetry is given by a product ofnon-abelian unitary group elements:

U =∏n

Un, Un ∈ U(kn) (5.60)

In matrix notation, in this basis, it is clear that each Un only works on thebasis vectors |φl〉 spanning HEn , so the factors commute.

69

Page 71: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

U =

(U1) 0 · · · 0

0 1 · · · 0...

.... . .

0 0 1

. . . 0 · · · 00 (U2) · · · 0...

.... . .

0 0 1

· · ·

1 0 · · · 00 1 · · · 0...

.... . .

0 0 (UL)

=

(U1) 0 · · · 0

0 (U2) · · · 0...

.... . .

0 0 (UL)

(5.61)

The reader reminded that the elements (Un) above are kn × kn block sub-matrices.

As before, we wish for the Bures metric to be gauge invariant under thislarger gauge symmetry. which is equivalent to demanding this of the fidelitybetween infinitesimally separated mixed states. A local (in M) gauge transfor-mation is given by |φl(λ)〉 7→ U(λ) |φl(λ)〉 (we will again suppress the argument).In the language of density matrices, this says the gauge transformation is givenby ρ 7→ UρU†. We now consider the ramifications for the fidelity. Inspired by[19], we take equation 2.30:

F(ρ, ρ+ δρ) = Tr√√

ρ(ρ+ δρ)√ρ

7→ Tr

√√UρU†(UρU† + δ(UρU†))

√UρU†

= Tr√U√ρU†(UρU† + δUρU† + UδρU† + UρδU†)U

√ρU†

= TrU√√

ρ(ρ+ U†δUρ+ δρ− ρU†δU)√ρU†

= Tr√√

ρ(ρ+ δρ+ [U†δU, ρ])√ρ

(5.62)

Using the fact that for any diagonalizable matrix A, we have (U†√AU)2 =

U†AU ⇔√U†AU = U†

√AU , thus we may bring the unitary conjugations

outside the square roots. We are only interested in positive roots of positiveoperators. Then the cyclic property of the trace allows us to kill the outerunitary elements, and finally we use δ(U†U) = δ1 = 0 ⇔ δU†U = −U†δU . Itis evident that the commutator term in 5.62 is gauge dependent, and we wouldlike to remove it. It is also immediate, that in the case where U is a scalar globalphase, then so is its variation, thus the fidelity is gauge invariant once more.

Intuitively, one might like to integrate out the gauge freedom of the choiceof basis of each separate degenerate subspace. Unfortunately, integrating overthe unitary group often results in vanishing elements, by symmetry, if it can be

70

Page 72: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

performed at all. Alternatively then, we decide to coarse-grain the Hilbert space,in order to have every column or row of the density matrix represent an entireenergy eigenspace. One proposal by the author is to pre-process the densitymatrix this way. In the same basis, we introduce the sub-block structure of ρ,each sub-block living in the tensor product of at most two energy eigenspaces:

ρ =

R11 R12 · · · R1L

R21 R22 · · · R2L

......

. . .

RL1 RL2 RLL

(5.63)

Where Rjk is a kj × kk submatrix of ρ, and Rjk = R†kj . Consider the actionof U in 5.61 on ρ in this form:

ρ 7→ UρU† =

U1R11U

†1 U1R12U

†2 · · · U1R1LU

†L

U2R21U†1 U2R22U

†2 · · · U2R2LU

†L

......

. . .

ULRL1U†1 ULRL2U

†2 ULRLLU

†L

(5.64)

One of the defining qualities of a unitary transformation is that it preservesthe norm of any vector it acts on. If it acts on a matrix from the left, it preservesthe norm of the columns, and from the right of the rows. The off-diagonal blocksof 5.64, are conjugated by different, arbitrary unitaries, so the only obviousconserved quantity within this subspace is the total sub-block matrix norm.

||Rjk|| :=√

Tr(RjkR

†jk

)=√

Tr(RjkRkj

)(5.65)

This is analogous to taking the norm of each column in a single row vector,and taking the norm of that row. The norm is manifestly invariant under thegauge symmetry:

||UjRjkU†k || =√

Tr(UjRjkU

†kUkRkjU

†j

)=√

Tr(RjkR

†jk

)≡ ||Rjk|| (5.66)

The blocks along the diagonal are square, and are conjugated by the sameunitary matrix left and right. All their eigenvalues are conserved under thesymmetry. A coarse-grained density matrix, with only one dimension for eachenergy eigenspace, is proposed as follows:

ρ 7→ ρC =

TrR11 ||R12|| · · · ||R1L||||R21|| TrR22 · · · ||R2L||

......

. . .

||RL1|| ||RL2|| TrRLL

(5.67)

A number of remarks: This transformation is only well posed in the energyeigenbasis, or more generally in a basis in which the gauge symmetry is defined

71

Page 73: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

like that in 5.61. Much information about the relative phases between energyeigenstates is lost. For instance, ρC is always real and positive. This does affordit the desirable quality of uniqueness: there is only one coarse-grained versionof any density matrix. Furthermore:

• ρC ∈ RL×L.

• ρC is symmetric, and with reality then Hermitian: ρC = ρTC = ρ†C

• One quickly verifies that Tr ρC = Tr ρ = 1.

• In order to be a well defined density matrix it must additionally be pos-itive semidefinite: z†ρCz ≥ 0 ∀z ∈ CL. This is equivalent to it havingnonnegative eigenvalues [14]. Although the author has not been able toprove this final attribute rigourously, it is conjectured and verified in nu-merical tests of 1000 of randomly generated density matrices ρ, reducedby any combination of degeneracies. None of the calculated coarse-grainedmatrices are found to have negative eigenvalues.

• Most importantly, it is manifestly invariant under the action of gaugegroup U on ρ.

In general, the purity γ(ρC) 6= γ(ρ), will not be preserved, but is expectedto increase, as the (nonzero) eigenvalues change. This is obvious in the case ofa diagonal ρ. Perhaps calculating the fidelity F (ρC , ρC + dρC) is more robustagainst this gauge group, although this brings other obvious complications tolife. Unfortunately, we will not go so far as to use the QGT or Bures metric totackle thermalization ourselves within the scope of this work. For that, we referto Polkovnikov and Kolodrubetz [25].

72

Page 74: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Chapter 6

Conclusions and Outlook

6.1 Summary

In summary, we have expounded the theory needed to interpret information lossin quantum mechanics. We introduced a number of useful geometric notionsand expressions for quantifying distance between states, notably the fidelity,Quantum Geometric Tensor (QGT), and Bures metric. Other formulas assessignorance of the reduced subsystem state, these are the purity and entropy. Wehave considered a number of properties of the QGT, such as its behavior underabelian, and later non-abelian gauge symmetries. These can come about in thecase of a degenerate spectrum. The mixed state generalization of the QGT is theBures metric, and it was found not to be well behaved under non-abelian gaugetransformation. A step towards resolving this may have been proposed withthe coarse-grained density matrix. Moving to simulations, we shed light on thetypical behavior of a quantum system with respect to its subsystems. Typical,in the sense of being drawn uniformly from the possible sets of eigenvectors, a’uniform ensemble’. We saw that subsystems with less degrees of freedom thantheir baths very quickly are locally described best by a state approaching thecompletely mixed one, an effect increasing with bath size. This meant that formost times, no local experiment can tell the subsystem apart, one time fromanother. In order to reach this conclusion, we evaluated the average fidelitybetween the subsystem at one time and others. The fidelity is uniquely drivenby distinguishability through measurements, and is a natural measure in thissense. This is a strong case for equilibration, a step towards thermalization:at later times, measurements on the subsystem will be constant. Conversely, aproduct initial state is typically distinguishable from a later time state undercoupled evolution. This indicates that such an unentangled system will evolveaway from its statistically exceptional initial product state, and seldom return.Moreover, we saw that the temporal variance from the mean was typically verylow, indicating that the averages give a good estimate of the instantaneous sit-uation. Finally, the time averaged entanglement entropy appeared to approach

73

Page 75: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

the maximum, as bath size was increased: the typical system maximally mixesits subsystems.

Continuing to analytics, we found it possible to integrate functions of sub-system reduced density matrices over all possible Hamiltonian eigenbases in theunitary group. This yielded an insight into the ’descent into disorder’ of theaverage system from the uniform ensemble. The resulting functions feature onlyHilbert space dimensions, and a single, albeit intricate, function of time and thesystem’s spectrum. The first result is the average density matrix itself, at anytime of its evolution, ρ(t). Starting in a product state, in this expression we sawthe initial condition compete with the maximally mixed state. For all times,traces of the initial information persisted, until the bath was taken infinite, andstatistically it was diluted so far the initial information becomes impossible torecover. After this, an expression was derived for the matrix-valued varianceover the unitary group of the density matrix. Its elements decay at least at thesame rate as the amplitude of the average density matrix elements themselves.Finally, the scalar purity was averaged, γ(t), showing the degree of mixing overtime, and the typical transition from ’well known’ to ’poorly known’ phase ofthe subsystem. It was observed that even at infinite times, for the extremal ratioof subsystem to bath, the system tends to be twice as ’pure’ as the maximallymixed state.

The underlying philosophy in this work is that the quantum system has noknowledge of- or investment in- the experimenter’s sense of locality. The parti-tion into subsystem and bath is ours, and highly artificial. In a mathematicalsense, system doesn’t have any preference a priori, which degrees of freedomto couple more or less. In fact, entanglement is only defined after we choosea basis in which to form superpositions. We proceed to judge locality by itslikeness to our prescription of a good, spatial basis, reflecting our own spatiallylocalized nature. Another legitimate basis might be that of Fourier modes, butthere are infinitely (though compactly) many. One goal of this report was tomake basis-independent statements about thermalization and equilibration. Inthis endeavor, one necessarily obtains results so general and symmetric they nolonger describe our world closely, from which they aim to detach themselves.

6.2 Outlook

In order to reinstate reality, an important next step is to revisit the simulationsand calculations with a nonuniform distribution of eigenvectors over the unitarygroup. Think of increasing correlation between adjacency of particles or spins,and their coupling strength. An integral weighing the systems by their Hamilto-nian eigenbasis’s distance to some preferred one is straightforward to construct,unfortunately not to evaluate. The following is a potential not invariant underthe choice of basis V :

P(V ) = C exp(−λ · Tr

([V,Ω][V,Ω]†

)), 1 6= Ω ∈ U(d) (6.1)

Here Ω is some preferred basis, λ controls the strength of the weighting, and

74

Page 76: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

C is a constant insuring∫U(d)

dV PV = 1. As the traced expression is of the

form AA†, it is positive semidefinite, giving nonnegative trace. It vanishes forV being a basis commuting with Ω, maximizing the probability. An obviouschoice to promote the computational basis is a diagonal Ω, though Ω must benondegenerate so it was proposed to take elements Ωjj := e2πij . This can befurther generalized by adding potential terms of the form Tr

(f(Ω)

)for some

matrix function f . Numerically, this is accessible, but to repeat the analyticintegrals of section 4, the exact integrands are powers of elements of V . The bestthat can be done is a perturbative expansion in the sum defining the exponent.

Concerning the end of chapter 5, this story is admittedly unfinished. Oneidea, coarse-graining the density matrix, was proposed to remedy the Buresmetric’s transformation properties, but by no means have we made conclusivearguments that it is the best route. Our main result is pointing out the problem.

Returning to Random Matrix Theory, the elegance of this framework allowedus to decouple the statistics of eigenvalues and eigenvectors in treating systemsand their prevalence. We chose to focus on the eigenvectors, as they ultimatelydetermine what can happen in a system, where the eigenvalues determine mostlyhow fast it happens. That is not to say all the interesting behaviour is inthe vectors and there is still much to be said in closing this gap. Picking upafter sections 3 and 4, one can also impose a global distribution P(λ) on theenergies, on top of the Vandermonde determinant resulting from a coordinatetransformation. This makes statements on the expected equilibration timesmore rigorous. Besides this, one can think of restricting initial conditions to acertain energy shell or band, as is often done in work mimicking Von Neumann[10], and attempt to recover the microcanonical ensemble more closely. Finally,when taking infinite size, or thermodynamic limits, the concept of temperaturecan be added, allowing for Gibbs ensembles, involving other kinds of entropy,etc. A further wealth of rich and interesting physics may be discovered.

75

Page 77: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Appendix A

Fidelity Plots

Here we have collected the plots deemed superfluous to the narrative in chapter3.

Figure A.1: Distribution of 4000 random Hamiltonians: Their time averagefidelity to the initial pure state, 〈F0〉, for a system of n = 3, 4, 5 spins, andnS = 1 spin. Surface under each plot is normalized to one.

76

Page 78: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Figure A.2: Distribution of 4000 random Hamiltonians: Their time averagefidelity to the initial pure state, 〈F0〉, for a system of n = 6, 7, 8 spins, andnS = 1 spin. Surface under each plot is normalized to one.

Figure A.3: Distribution of 4000 random Hamiltonians: Their time averagefidelity between all times, 〈F∞〉, for a system of n = 3, 4, 5 spins, and nS = 2spins. The surface under each plot is normalized to one.

77

Page 79: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Figure A.4: Distribution of 4000 random Hamiltonians: Their time averagefidelity between all times, 〈F∞〉, for a system of n = 6, 7, 8 spins, and nS = 2spins. The surface under each plot is normalized to one.

Figure A.5: Distribution of 4000 random Hamiltonians: Their temporal varianceof the fidelity between different times, σ2

∞, for a system of n = 3, 4, 5 spins, andnS = 2 spins. The surface under each plot is normalized to one.

78

Page 80: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Figure A.6: Distribution of 4000 random Hamiltonians: Their time averagednormalized entropy 〈S1〉, for n = 3, 4, 5, 6, 7, 8 and nS = 1. The entropy is scaledby the subsystem dimension, such that the maximum entropy is 1, correspondingto ρS ∝ 1.

79

Page 81: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Appendix B

Python Code

Preamble: These commands are needed in most of the functions, and should beimported at the beginning of the script.

from numpy import arange, imag as im, cumsum

from numpy import zeros, kron, dot, random as rand, mean

from numpy import exp, log, diag, linalg as lina, arccos

from numpy import array as arr, real as re, linspace, var

from numpy import median, trace as tr, ...

... count_nonzero as hamming

from numpy import cos, sin, eye, pi, reshape as resh

from scipy.linalg import logm, fractional_matrix_power as ...

...frapow

from scipy.linalg import expm, triu, det, norm, svd, sqrtm

from random import sample, seed

from matplotlib import pyplot as plt

from scipy.special import erf, digamma

from scipy.optimize import root

from copy import deepcopy

from scipy.stats.stats import pearsonr

from time import clock

from itertools import permutations

from math import factorial

The function ’compactify’ coarse-grains the density matrix ’rho’, accordingto a vector of degeneracies ’D’.

∑Lj=1Dj = d, the Hilbert space dimension. See

subsection 5.4.2.

80

Page 82: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

def compactify(rho,D):

d=len(D)

E = [0]

E.extend(cumsum(D))

rho2 = zeros((d,d))

for n in range(d):

for m in range(n+1,d):

sub = rho[E[n]:E[n+1],E[m]:E[m+1]]

rho2[n,m]=(re(tr(dot(sub,sub.T.conj()))))**0.5

rho2[m,n]=deepcopy(rho2[n,m])

for n in range(d):

rho2[n,n]=re(tr(rho[E[n]:E[n+1],E[n]:E[n+1]]))

return rho2

The function ’genrho’ generates a random density matrix of dimension ’dim’[14]:

def genrho(dim):

A = 1j*rand.normal(0,1,(dim,dim))+...

...rand.normal(0,1,(dim,dim))

rho = dot(A,A.T.conj())/tr(dot(A,A.T.conj()))

rho -= 1j*im(diag(diag(rho)))

return rho

The function ’rtr’ takes a square matrix ’rho’, and traces out the ’bath’rightmost qubits in the computational basis.

81

Page 83: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

def rtr(rho,bath):

m = 2**bath

n = int(len(rho)/m)

rho2 = zeros((n,n),dtype=complex)

for a in range(n):

for b in range(n):

for c in range(m):

rho2[a][b] += rho[m*a+c][m*b+c]

return rho2

The function ’ltr’ takes a square matrix ’rho’, and traces out the ’subs’leftmost qubits in the computational basis.

def ltr(rho,subs):

m = 2**subs

n = int(len(rho)/m)

rho2 = zeros((n,n),dtype=complex)

for a in range(n):

for b in range(n):

for c in range(m):

rho2[a][b] += rho[a+n*c][b+n*c]

return rho2

The function ’unita’ generates a uniformly random unitary matrix of size’d’, according to the method explained in 2.6.1.

def unita(dim):

R = (rand.exponential(1,(dim,dim)))**0.5

phi = (rand.random((dim,dim)))*2*pi

Z = R*(cos(phi)+1j*sin(phi))

q,r = lina.qr(Z)

lam = eye(dim)*(Z/abs(Z))

Q = dot(q,lam)

return Q

The function ’puretrace’ takes a Hilbert space vector ’psi’, turns it into a

82

Page 84: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

density matrix, and traces out ’trace’ dimensions from the right in the compu-tational basis.

def puretrace(psi,trace):

A = resh(psi,(int(len(psi)/trace),trace))

rho = dot(A,A.T.conj())

return rho

The function ’genham’ generated a random Hermitian matrix of dimensions’size’ times ’size, with eigenvalues between -pi and pi. The eigenvalues aremutually repulsive, and the eigenvectors are from the Haar measure, as in section2.6.

def genham(size):

Hh = 1j*logm(unita(size))

Hu = triu(Hh,1)

H = re(diag(diag(Hh)))+Hu+Hu.T.conj()

return H

The function ’evol’ creates the data used in the histograms in chapter 3. Itgenerates ’runs’ Hamiltonians (4000 was used) for ’subs’ subsystem spins (1 and2 were used). Output are the mean values 〈F0〉, 〈F∞〉, σ0, variances σ2

∞, andσ2

0 , normalized entanglement entropy S1, and the minima over time mint F0 andmint F∞ in vectors. See section 2.2.

83

Page 85: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

def evol(runs,subs):

compareinitial = []

compareall = []

entro = []

samplesize = 150

for sites in range(3,9):

clo = clock()

init = [[],[],[]]

ra = [[],[],[]]

ent = [[],[],[],]

for I in range(runs):

B = [[],[],[]]

V = unita(2**sites)

ei = 2*pi*rand.random(2**sites)-pi

rho1 = puretrace(dot(V,exp(-1j*ei*rand.random()*...

...1000)*V[0].conj()),2**(sites-subs))

for t in range(samplesize):

rho2 = deepcopy(rho1)

rho1 = puretrace(dot(V,exp(-1j*ei*...

...rand.random()*1000)*V[0].conj()),2**...

...(sites-subs))

B[0].append(re(rho1[0][0]**0.5))

B[1].append(re(tr(sqrtm(dot(dot(sqrtm(rho2),...

...rho1),sqrtm(rho2))))))

B[2].append(-re(tr(dot(rho1,logm(rho1))))/...

...(subs*log(2)))

init[0].append(mean(B[0]))

init[1].append(var(B[0]))

init[2].append(float(min(B[0])))

ent[0].append(mean(B[2]))

ent[1].append(var(B[2]))

ent[2].append(float(min(B[2])))

ra[0].append(mean(B[1]))

ra[1].append(var(B[1]))

ra[2].append(float(min(B[1])))

compareinitial.append(init)

compareall.append(ra)

entro.append(ent)

print (’seconds elapsed: ’, int(clock()-clo), ’ for ...

...sites: ’, sites)

return compareinitial,compareall,entro

84

Page 86: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

The function ’conjclass’ takes a vector ’A’, consisting of the numbers 0 toq-1, as the result of the mapping from [0,1,...,q-1] by the permutation σ ∈ Sq,as a proxy for σ. It returns the cycle lengths of the conjugacy class of σ. Seesubsection 2.7.

def conjclass(A):

c = []

a = []

for n in range(len(A)):

if n not in a:

a.append(n)

t = 1

m = deepcopy(n)

while A[m] != n:

m= A[m]

a.append(m)

t += 1

c.append(t)

return sorted(c,reverse=True)

the function ’finvert’ refers to the list ’permutations([0,1,2,3])’ as a proxyfor the elements of S4, and returns the list entry number of the inverse ofpermutation ’c’.

def finvert(c):

L = list(permutations([0,1,2,3]))

A1 = arr(L[c])

for d in range(24):

B1 = arr(L[d])

C1= A1[B1]

if all(C1==arr([0,1,2,3])):

f = deepcopy(d)

return f

The function ’wintegrate2’ takes a collection of data in ’data’, collected intables 2 or 4, and prints in the console, term by term, the result of the unitaryintegrals in chapter 4. If ’pri’6= 1, the output of printed terms is suppressed.

85

Page 87: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

def wintegrate2(data,pri):

siglabel,tauilabel,siglegend,tauilegend = data

l = len([j for i in siglegend for j in i])

h = deepcopy(l)

dim=1

while dim < h:

h /=dim

dim+=1

if factorial(dim) != l:

print(’incorrect legend length’)

lens = len(siglabel)

lent = len(tauilabel)

lenw = dim + int(dim>3)+int(dim>4)

L = list(permutations(list(range(dim))))

final = zeros(lens*lent*lenw)

for s in range(l):

svec=[]

for a in range(lens):

svec.append(int(s in siglegend[a]))

for t in range(l):

tvec = []

for a in range(lent):

tvec.append(int(t in tauilegend[a]))

wvec = zeros(lenw)

SIG = arr(L[s])

TAUI = arr(L[t])

c = conjclass(SIG[TAUI])

typ = sum(c)-len(c)+int(c[0]>2)+int(c[0]*sum(c)>16)

wvec[typ]=1

final += kron(kron(svec,tvec),wvec)

if pri == 1:

Wdenom = [’d’,’d(d^2-1)’,’d(d^2-1)(d^2-4)’,’d^2...

...(d^2-1)(d^2-4)(d^2-9)’,’d^2(d^2-1)(d^2-4)...

...(d^2-9)(d^2-16)’]

Wenum = [[’1’],[’d’,’-1’],[’d^2-2’,’-d’,’2’],...

...[’d^4-8d^2+6’,’-d^3+4d’,’d^2+6’,’2d^2-3’,’-5d’],...

...[’d^5-20d^3+78d’,’-d^4+14d^2-24’,’d^3-2d’,...

...’2d^3-18d’,’-2d^2-24’,’-5d^2+24’,’14d’]]

print(’denominator =’,Wdenom[dim-1],’ ...

...’,hamming(final),’nonzero terms coming up!’)

for a in range(lens):

for b in range(lent):

for w in range(lenw):

n = lent*lenw*a+lenw*b+w

if final[n] != 0:

input(’next?’)

print(final[n],’*’,siglabel[a],’*’,...

...tauilabel[b],’*’,Wenum[dim-1][w])

return final

86

Page 88: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

The function ’storage4’ can be called without arguments to serve as the datain ’wintegrate2’ for the calculation of the mean dynamical subsystem purity γ(t).

def storage4():

siglabel = [’1’,’B’,’dB’,’A’,’dA’]

tauilabel = [’d’,’d^2’,’X’,’X^2’,’dX’,’e2’,’e3’,’e-3’]

siglegend = [[0,5,14,16],[1,3,6,8,10,11,13,15],[7,9],...

...[2,4,12,17,19,20,21,22],[18,23]]

tauilegend = [[9,10,13,17,18,22],[7,23],[3,4,8,11,12,...

...15,19,20],[0],[1,2,6,21],[16],[14],[5]]

return siglabel, tauilabel, siglegend, tauilegend

The function ’storagevar’ can be called without arguments to serve as thedata in ’wintegrate2’ for the calculation of the average ρ2.

def storagevar():

siglabel = [’(1,1)’,’(1,1)*B’,’(i,1)’,’(1,j)’,’(i,j)*B’,...

...’(i,i)*B^2’]

tauilabel = [’d’,’d^2’,’X’,’X^2’,’dX’,’e2’,’e3’,’e-3’]

siglegend = [[0,2,21,23],[1,4,6,8,9,11,18,20],...

...[12,14,17,19],[3,5,15,22],[13,16],[7,10]]

tauilegend = [[9,10,13,17,18,22],[7,16],...

...[3,4,8,11,12,15,19,20],[0],[1,5,6,14],[23],[21],[2]]

return siglabel, tauilabel, siglegend, tauilegend

87

Page 89: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

Bibliography

[1] J. Von Neumann, “Proof of the ergodic theorem and the h-theorem inquantum mechanics,” The European Physical Journal H, vol. 35, pp. 201–237, Nov 2010.

[2] S. Goldstein, J. L. Lebowitz, R. Tumulka, and N. Zanghı, “Long-time be-havior of macroscopic quantum systems,” The European Physical JournalH, vol. 35, pp. 173–200, Nov 2010.

[3] M. V. Berry, “Regular and irregular semiclassical wavefunctions,” Journalof Physics A: Mathematical and General, vol. 10, no. 12, 1977.

[4] J. M. Deutsch, “Quantum statistical mechanics in a closed system,” Phys.Rev. A, vol. 43, pp. 2046–2049, Feb 1991.

[5] M. Srednicki, “Chaos and quantum thermalization,” Phys. Rev. E, vol. 50,pp. 888–901, Aug 1994.

[6] M. Srednicki, “The approach to thermal equilibrium in quantized chaoticsystems,” Journal of Physics A: Mathematical and General, vol. 32, no. 7,p. 1163, 1999.

[7] S. Goldstein, J. L. Lebowitz, C. Mastrodonato, R. Tumulka, and N. Zanghı,“Normal typically and von neumann’s quantum ergodic theorem,” Proceed-ings: Mathematical, Physical and Engineering Sciences, vol. 466, no. 2123,pp. 3203–3224, 2010.

[8] D. A. Huse, R. Nandkishore, V. Oganesyan, A. Pal, and S. L. Sondhi, “Lo-calization protected quantum order,” Phys. Rev., vol. B88, no. 1, p. 014206,2013.

[9] M. Rigol and M. Srednicki, “Alternatives to eigenstate thermalization,”Phys. Rev. Lett., vol. 108, p. 110601, Mar 2012.

[10] P. Reimann, “Generalization of von neumann’s approach to thermaliza-tion,” Phys. Rev. Lett., vol. 115, p. 010403, Jul 2015.

[11] N. Linden, S. Popescu, A. J. Short, and A. Winter, “Quantum mechanicalevolution towards thermal equilibrium,” Phys. Rev. E, vol. 79, p. 061103,Jun 2009.

88

Page 90: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

[12] Y. Aharonov and D. Bohm, “Significance of electromagnetic potentials inthe quantum theory,” Phys. Rev., vol. 115, pp. 485–491, Aug 1959.

[13] M. Nakahara, Geometry, Topology and Physics. CRC Press, 2003.

[14] I. Bengtsson and K. Zyczkowski, Geometry of Quantum States. CambridgeUniversity Press, 2006.

[15] J. P. Provost and G. Vallee, “Riemannian structure on manifolds of quan-tum states,” Comm. Math. Phys., vol. 3, pp. 289–301, 1980.

[16] M. V. Berry, “Quantal phase factors accompanying adiabatic changes,”Proceedings of the Royal Society of London A: Mathematical, Physical andEngineering Sciences, vol. 392, no. 1802, pp. 45–57, 1984.

[17] F. Wilczek and A. Zee, “Appearance of gauge structure in simple dynamicalsystems,” Phys. Rev. Lett., vol. 52, 1984.

[18] A. Uhlmann, “Parallel transport and “quantum holonomy” along densityoperators,” Reports on Mathematical Physics, vol. 24, pp. 229–240, Oct.1986.

[19] H.-J. Sommers and K. Zyczkowski, “Bures volume of the set of mixedquantum states,” J. Phys. A, vol. 36, 2003.

[20] D. Bures, “An extension of kakutani’s theorem on infinite product measuresto the tensor product of semifinite w∗-algebras,” Trans. Amer. Math. Soc.,vol. 135, pp. 199–212, 1969.

[21] S. L. Braunstein and C. M. Caves, “Statistical distance and the geometryof quantum states,” Phys. Rev. Lett., vol. 72, pp. 3439–3443, May 1994.

[22] R. Jozsa, “Fidelity for mixed quantum states,” Journal of Modern Optics,vol. 41, no. 12, pp. 2315–2323, 1994.

[23] A. T. Rezakhani, D. F. Abasto, D. A. Lidar, and P. Zanardi, “Intrinsicgeometry of quantum adiabatic evolution and quantum phase transitions,”Phys. Rev. A, vol. 82, p. 012321, Jul 2010.

[24] P. Zanardi, L. Campos Venuti, and P. Giorda, “Bures metric over thermalstate manifolds and quantum criticality,” Phys. Rev. A, vol. 76, p. 062318,Dec 2007.

[25] M. Kolodrubetz, P. Mehta, and A. Polkovnikov, “Geometry and non-adiabatic response in quantum and classical systems,” arXiv preprintarXiv:1602.01062, 2016.

[26] G. Clos, D. Porras, U. Warring, and T. Schaetz, “Time-resolved observa-tion of thermalization in an isolated quantum system,” Phys. Rev. Lett.,vol. 117, p. 170401, Oct 2016.

89

Page 91: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

[27] H.-P. Breuer and F. Petruccione, The Theory of Open Quantum Systems.Oxford University Press, 2007.

[28] A. M. Kaufman, M. E. Tai, A. Lukin, M. Rispoli, R. Schittko, P. M. Preiss,and M. Greiner, “Quantum thermalization through entanglement in anisolated many-body system,” Science, vol. 353, no. 6301, pp. 794–800, 2016.

[29] J. Von Neumann, “Quote,” Scientific American, vol. 225, no. 3, p. 180,1971.

[30] M. A. Nielsen and I. L. Chuang, Quantum Computation and QuantumInformation. Cambridge University Press, 2000.

[31] C. A. Fuchs and C. M. Caves, “Ensemble-dependent bounds for accessibleinformation in quantum mechanics,” Phys. Rev. Lett., vol. 73, pp. 3047–3050, Dec 1994.

[32] S. R. B. Eynard, T. Kimura, “Lecture notes on random matrices,” arXivpreprint arXiv:1510.04430, 2015.

[33] A. Haar, “Der massbegriff in der theorie der kontinuierlichen gruppen,”Annals of Mathematics, vol. 34, no. 1, pp. 147–169, 1933.

[34] M. Ozols, “How to generate a random unitary matrix.” unpublished essayon http://home.lu.lv/ sd20008/, 2009.

[35] F. Mezzadri, “How to generate random matrices from the classical compactgroups,” Notices of the American Mathematical Society, vol. 54, pp. 592 –604, 5 2007.

[36] J. J. Rotman, An Introduction to the Theory of Groups. Springer, 4th ed.,1999.

[37] A. O. Caldeira and A. J. Leggett, “Influence of dissipation on quantumtunneling in macroscopic systems,” Phys. Rev. Lett., vol. 46, pp. 211–214,Jan 1981.

[38] S. Sen, “Average entropy of a quantum subsystem,” Physical Review Let-ters, vol. 77, pp. 1–3, July 1996.

[39] C. Itzykson and J. Zuber, “The planar approximation. ii,” Journal of Math-ematical Physics, vol. 21, no. 3, pp. 411–421, 1980.

[40] B. Collins and P. Sniady, “Integration with respect to the haar measureon unitary, orthogonal and symplectic group,” Communications in Mathe-matical Physics, vol. 264, no. 3, pp. 773–795, 2006.

[41] P. W. Brouwer and C. W. J. Beenakker, “Diagrammatic method of inte-gration over the unitary group, with applications to quantum transport inmesoscopic systems,” Journal of Mathematical Physics, vol. 37, pp. 4904–4934, Oct. 1996.

90

Page 92: Thermalization in Quantum Information Geometry · 1.2 Quantum Geometry There is a new emerging school of thought in the eld of quantum condensed matter and quantum statistical physics.

[42] E. Lubkin, “Entropy of an n-system from its correlation with a k-resevoir,”Journal of Mathematical Physics, vol. 19, no. 5, pp. 1028–1031, 1978.

[43] R. P. Feynman, “Forces in molecules,” Phys. Rev., vol. 56, pp. 340–343,Aug 1939.

[44] Y. Ma, S. Chen, H. Fan, and W. Liu, “Abelian and non-abelian quantumgeometric tensor,” Phys. Rev. B, vol. 81, 2010.

[45] V. Gritsev and A. Polkovnikov, “Dynamical quantum hall effect in theparameter space,” PNAS, vol. 109, 2012.

[46] B. De Wit, “Lecture notes: Introduction to gauge theories and the standardmodel,” October 1995.

91