Fidelity spectrum: A tool to probe the property of a quantum phase
Chi Yu Wing†, , Gu Shi-Jian
Department of Physics and ITP, The Chinese University of Hong Kong, Hong Kong, China

 

† Corresponding author. E-mail: wcyu@phy.cuhk.edu.hk

Project supported by the Earmarked Research Grant from the Research Grants Council of HKSAR, China (Grant No. CUHK 401212).

Abstract
Abstract

Fidelity measures the similarity between two states and is widely adapted by the condensed matter community as a probe of quantum phase transitions in many-body systems. Despite its success in witnessing quantum critical points, information about the fine structure of a quantum phase one can get from this approach is still limited. Here, we proposed a scheme called fidelity spectrum. By studying the fidelity spectrum, one can obtain information about the characteristics of a phase. In particular, we investigated the spectra in the one-dimensional transverse-field Ising model and the two-dimensional Kitaev model on a honeycomb lattice. It was found that the spectra have qualitative differences in the critical and non-critical regions of the two models. From the distributions of them, the dominating k modes in a particular phase could also be captured.

1. Introduction

A quantum phase transition (QPT)[1,2] is driven by quantum fluctuations in the ground state of a many-body system. As an external parameter such as a magnetic field varies, the qualitative structure of the system's ground-state wavefunction exhibits an abrupt change across the quantum critical point. With this observation, researchers were motivated to use the fidelity, which is a concept borrowed from the quantum information theory, to study QPTs.[3,4]

Fidelity measures the similarity between two states. Mathematically, the static ground-state fidelity is defined as the magnitude of the overlap between two ground states differing from each other by a small value in the parameter space. Within one phase of the system, the two states have similar structures and thus the fidelity is close to unity. On the other hand, the ground states in distinct phases of the system are significantly different from each other. The fidelity is thus expected to show a sudden drop across the quantum critical point. The significance to use fidelity and its variations, such as fidelity susceptibility,[5,6] generalized fidelity susceptibility,[7] fidelity per site,[8,9] reduced fidelity,[10,11] operator fidelity,[12,13] and density-functional fidelity,[14] as a probe of QPTs have been testified in a number of condensed matter models (for a comprehensive review, please refer to Ref. [15]). Recently, an equality relating the fidelity susceptibility and the spectral function has been derived.[16] This makes the fidelity susceptibility to be directly measurable in experiments via neutron scattering or the angle-resolved photoemission spectroscopy (ARPES) techniques.

Although the fidelity can well witness a quantum critical point and provide us the critical exponents in continuous phase transitions,[1720] the information about the nature and the symmetry of a phase that one could obtain is limited. With the motivation to gain more insight about a quantum phase from the fidelity approach, we proposed the fidelity spectrum in this work. Recently, Gu et al.[21] and Sacramento et al.[22] also introduced the concepts of mode fidelity and fidelity spectrum, respectively, with similar motivations.

In Sacramento and his collaborators' proposal, the fidelity spectrum was defined through the fidelity operator where ρ1 and ρ2 are the reduced density matrices. The focus was on the partial state of the system. However, the global ground state of the system usually plays a role in a QPT. Especially in a topological QPT in which the global topological order of the system changes,[23] it may be insufficient to just consider the partial state of the system to characterize the quantum phase. At this point, we were motivated to define the fidelity spectrum using the global ground-state wavefunction of the system. We investigated the distributions of the spectra in the one-dimensional (1D) Ising model in a transverse field and the two-dimensional (2D) Kitaev model on a honeycomb lattice. The spectra were found to exhibit qualitative different behaviors in different regimes of the models. From the distributions of the spectra, one can determine the dominating k modes in a phase.

The paper is organized as follows. In Section 2, we established a general mathematical framework of the fidelity spectrum. The fidelity spectrum was studied in the 1D transverse-field Ising model, which exhibits a continuous QPT, and the honeycomb lattice Kitaev model, which exhibits a topological QPT. The results are discussed in Sections 3 and 4, respectively. Finally, a summary is given in Section 5.

2. Fidelity spectrum

Consider a general form of Hamiltonian,

where HI is the driving Hamiltonian and λ is an external driving parameter. Let {|Ψn(λ)〉} be a set of complete orthogonal eigen-basis of the Hamiltonian at λ with corresponding eigenenergies En(λ), i.e., H(λ)|Ψn(λ)〉 = En(λ)|Ψn(λ)〉. Focusing on QPTs that are not induced by the ground-state level crossing, the ground state of the Hamiltonian in a finite system is non-degenerate. For a small change δλ in the external driving parameter, the non-degenerate perturbation theory could be applied. To the first-order correction, the perturbed ground state is

where and

is the fidelity susceptibility.[5,6]

Consider the overlap

If the driving parameter λ is time-dependent, the overlap would be related to the quantum work distribution in quench dynamics.[2426] For n = 0, the overlap is just the fidelity[3,4] as given by

For n ≠ 0, we have

where

The fidelity spectrum we discussed in this work is defined by the set {Mn(λ)}, which is independent of δλ.

Physically, the fidelity spectrum tells us how much the ground state of the final Hamiltonian comes from different eigenstates of the initial Hamiltonian. When the system is under an external perturbation δλHI, the perturbed ground state is in general a superposition of the eigenstates of the unperturbed Hamiltonian. The contribution from the ground state is given by the fidelity while the contribution from the n-th excited state of the original Hamiltonian is governed by Mn. As one can see from Eq. (7), the fidelity spectrum depends on the scattering amplitude of the driving Hamiltonian and also the energy gap between the ground state and the excited state. By investigating the distribution of {Mn}, we expect to gain information about the modes that are dominating in a quantum phase.

To have a quantitative description of the distribution of {Mn}, we introduce a quantity

Since Mn has a probability interpretation of the final ground state being contributed from an excited state of the original Hamiltonian and it is normalized, i.e., Σn≠0 |Mn(λ)|2 = 1, SF can be regarded as an analogy to the Shannon entropy characterizing the distribution. The more spread the distribution of {Mn} is, the larger SF is expected. While for a less spread distribution, SF would be smaller.

3. One-dimensional transverse-field Ising model

The 1D transverse-field Ising model is one of the simplest many-body systems, which is exactly solvable and exhibits a QPT.[2731] The Hamiltonian of the model is given by

where , with κ = x,y,z, is the Pauli matrix, N is the number of spins, and h characterizes the strength of the external field in the unit of the coupling constant. Periodic boundary condition is adapted here.

The model’s Hamiltonian can be diagonalized through standard procedure using the Jordan–Wigner transformation, Fourier transformation, and Bogoliubov transformation.[1] After these transformations, equation (9) becomes

where

and k is the momentum wavevector with values of (2m + 1)π/N for m = −N/2,−N/2+1,…, N/2−1. In the thermodynamic limit, the system is gapless only at h = 1 and gapped for other values of h > 0. A continuous QPT takes place at hc = 1. For 0 ≤ h < 1, the system is a ferromagnet, while it is paramagnetic for h > 1. Recently, the system has been experimentally realized by tuning a strong transverse magnetic field in CoNb2O6 (cobalt niobate). By detecting the low-energy excitation in neutron scattering experiment, a continuous QPT was observed in the material.[32]

To investigate the fidelity spectrum in Eq. (7), we take the driving Hamiltonian as

The fidelity spectrum was calculated to be

where k is used as an index to characterize an excited state in which a pair of quasiparticles with momentum k and −k is created. The fidelity susceptibility is given by[33]

Our main target is to investigate the distribution of the fidelity spectrum, which is closely related to the probability distribution of the k mode excitations discussed in the context of quench dynamics.[34,35]

Figure 1 shows the distribution of Mk over the k-space for various values of h. From the figure, we see that for h away from the critical point, the distribution of the fidelity spectrum is more widely spread. However, at the critical field h = 1, the distribution is dominated by the zero-th mode and shows an exponential decaying trend away from this mode. One may understand this behavior by referring to Eq. (7) and noting that the energy gap between the ground state and the first excited state vanishes at the critical point.

Fig. 1. A plot of Mk in Eq. (13) distributed over the k-space for the 1D transverse-field Ising model. Here N = 100 and k = (2m + 1)π/N. Only positive k’s are shown.

Quantitatively, we can also obtain the dominating mode by considering dMk/dk = 0 which gives cosk = 2h/(1 + h2). For h = 1, k = 0 is a solution to the equation. However, note that for N is even, k = (2m + 1)π/N and the k = 0 mode is not allowed in a finite system. Instead, Mk peaks at k = π/N, which in turn tends to zero as N tends to infinity.

Figure 2 shows a plot of SF as a function of h for various system sizes. At h = 1, the fidelity spectrum is dominated by the zero-th mode. SF is thus a minimum as expected. On the other hand, the fidelity spectra in the non-critical region are more spread and SF becomes larger.

Fig. 2. SF as a function of h for various system size in the 1D transverse-field Ising model.

In addition, a point worth noting from the figure is that SF is insensitive to the system size at the critical point while it shows a non-trivial scaling behavior with N for h ≠ 1. For h = 1, the size dependence of SF can be obtained analytically. Putting Eq. (13) into Eq. (8), we have

From Eq. (14), χF(h = 1) = N(N − 1)/32 and we obtain

where A and B are some constants independent of N. In the limit where N → ∞, SF tends to a constant.

Figure 3 shows a plot of the scaling dependence of SF with the system size. From linear fitting, we obtain

In the thermodynamic limit, SF tends to a constant at the critical point as expected from the analytical analysis above. On the other hand, SF shows a logarithmic divergence away from the critical point. This implies that fidelity spectrum becomes more and more dispersive as the system size increases.

Fig. 3. (Left) The scaling behavior of SF at the critical point of the 1D transverse-field Ising model. The straight line shows the linear fit of −1.89lnN/N + 0.89. (Right) The scaling behavior of SF in the non-critical region. The slope of the straight line is 1 (maximum error = ±0.001).
4. The Kitaev model on a honeycomb lattice

The Kitaev model was first introduced by Kitaev to study anyonic statistics.[36] The ground state of the model can be exactly obtained and it is found to possess a topological QPT.[36,37] Its ground-state phase diagram consists of gapless phases with non-Abelian anyon excitations and gapped phases with Abelian anyon excitations. As an exactly solvable 2D model exhibiting such an interesting phase diagram, the Kitaev model and its extensions have been intensively studied in the field of condensed matter physics (for examples, please see Refs. [38]–[42]). Besides, the model also captured much attention in the field of quantum information processing for its potential to realize the fault-tolerant quantum computation.[43]

The model consists of 1/2 spins located at the vertices of a honeycomb lattice. They are interacting with the nearest neighbors through three types of bonds, namely, x, y, and z bond, depending on their directions (see Fig. 4). The Hamiltonian reads [36]

where j, k denote the vertex of the corresponding bonds. Jx, Jy, and Jz are dimensionless coupling constants in x, y, and z directions, respectively.

Fig. 4. (a) A sketch of the Kitaev model on a honeycomb lattice. The lattice can be divided into two equivalent triangular sublattices denoted by the black and white dots. (b) A deformation of the honeycomb lattice on the left into a topologically equivalent brick-wall lattice. ex and ey are the unit vectors connecting the mid-point of the z bonds.

Consider the plaquette labeled as p in Fig. 4. The operator commutes with the Hamiltonian for all p. Obviously, Wp of different plaquettes also commute. Therefore, the Hilbert space can be divided into subspaces with wp = ±1 for each plaquette. The ground state lies in the subspace in which all wp are equal to one. This is the key feature that makes an exact solution to the model possible.[36,37]

The Hamiltonian in Eq. (18) can be diagonalized by two different methods.[36,37] Following the approach introduced by Chen et al. in Ref. [37], the honeycomb lattice is first deformed into a topologically equivalent brick-wall lattice by stretching the lattice horizontally (see Fig. 4). Defining the Jordan–Wigner transformation as

where n and m are the column and row indices, respectively. Ñ is the number of sites in each row, and cn,m and are fermion annihilation and creation operators, respectively. The superscript s = b, w denotes the black and white sites, respectively.

Introducing the Majorana fermion operators

the Hamiltonian is then transformed into

where r is the coordinate to the mid-point of a z bond connecting the black and white sites at (n,m) and (n,m + 1). and is related to Wp by . By taking all αr = 1, the Hamiltonian becomes quadratic and can be diagonalized through Bogoliubov transformation into

where

bk and is another set of fermionic operators satisfying the usual anti-commutation relations. kx(y) is the momentum wavevector with values of (2mx(y) + 1)π/L for mx(y) = −L/2,…,L/2 − 1.

For

the system is gapless in the thermodynamic limit. In this gapless phase, the excitation is described by non-Abelian anyons. For the coupling constants outside the range given by Eq. (30), the system is in a gapped phase and obeys Abelian anyonic statistics. In the intermediate regime, a topological quantum phase transition takes place. A schematic phase diagram on the Jx + Jy + Jz = 1 plane is shown in the inset of Fig. 6.

To calculate the ground-state fidelity spectrum, let us consider varying Jz along the Jx = Jy line on the Jx + Jy + Jz = 1 plane (see the red dashed line in the inset of Fig. 6). The interaction Hamiltonian is given by

The fidelity spectrum was found to be

and the fidelity susceptibility can be calculated by[44]

The distribution of Mk over the k space for Jz = 0.1, 0.3, 0.5, 0.7, 0.9 are shown in the 3D plots in Fig. 5. In the gapless phase where Jz < 0.5, the fidelity spectrum consists of four peaks. These peaks become more concentrated to the zone center as the value of Jz approaches the critical point at Jz = 0.5. While in the gapped phase (Jz > 0.5), the spectrum becomes more dispersive and there is no singular peak.

Fig. 5. 3D plots of the distributions of Mk over the k space for (a) Jz = 0.1, (b) 0.3, (c) 0.5, (d) 0.7, (e) 0.9 in the Kitaev model. Here kx = (2mx + 1)π/L and ky = (2my + 1)π/L, and L = 100.

The behavior of the fidelity spectrum in the gapless phase of the Kitaev model is similar to that at the critical point of the Ising model in which the energy gap vanishes. When the system is gapless in the energy spectrum, the ground state is readily excited to other states. From the peaks in the fidelity spectrum, one can immediately realize which modes’ contributions are the largest for various values of Jz.

Figure 6 shows a plot of SF in Eq. (8) as a function of Jz for various system sizes. From the figure, we see that SF for Jz < 0.5 is smaller than that for Jz > 0.5. This echoes the fact that the fidelity spectrum in the gapless phase is less dispersive than that in the gapped phase.

Fig. 6. SF as a function of Jz for various system size. The lines represent the cases for L = 100, 200, 400, 800, and 1600, respectively, when going upward. The inset shows a schematic phase diagram on the Jx + Jy + Jz = 1 plane of the Kitaev model. The A phase is gapped, while the B phase is gapless. A topological QPT takes place at Jz = 0.5 along the Jx = Jy line (as indicated by the dashed line).

Moreover, SF in the gapless phase is an oscillating function of Jz. This feature comes from the periodic level-crossing occurred in the excited states of the system. The oscillation patterns are magnified in Fig. 7. As the system size increases, the number of oscillation cycles increases. From the linear fitting on the right panel in Fig. 7, we find that

where ω0 is the frequency of oscillation. From Fig. 6, we also recognize that SF exhibits some non-trivial scaling behaviors. It increases with the system size. Its scaling behavior is analyzed in Fig. 8. For Jz < 0.5, SF is oscillating and we take the values enclosed by the upper envelopes to perform the scaling analysis. By linear fitting, we find that

Fig. 7. (a) Oscillating pattern of SF in the gapless phase. (b) Size dependence of the number of peaks (np) in SF in the gapless phase. The slope of the straight line is 0.991 ± 0.004.
Fig. 8. Scaling behavior of SF for various values of Jz. The straight lines are the linear fitting of the data points.

Interestingly, like the Ising model, the pre-factor of the size dependence of SF in the gapped region is also equal to unity. If we express the scaling behaviors in the two models both in terms of N, we have

in a gapped region. Whether this relation is general for other gapped phases in many-body systems or not still needs further exploration.

5. Summary

In summary, we investigated the fidelity spectra in the 1D transverse-field Ising model which exhibits a continuous QPT, and in the 2D honeycomb lattice Kitaev model which exhibits a topological QPT. We found that the fidelity spectrum shows qualitatively different behaviors in different phases of the model. A common feature about the spectrum is that the distribution is dominated by certain k modes in the gapless region while it is more dispersive in the gapped region.

To characterize the distribution of the fidelity spectrum quantitatively, we introduced the quantity SF in analogy to the Shannon entropy. It was found that at the critical point of the Ising model, SF ∼ ln N/N while it scales with lnN in the non-critical region. The latter scaling behavior also holds for the Kitaev model. The universality of the relation in Eq. (36) requires further exploration.

Reference
1Sachdev S1999Quantum Phase TransitionsCambridgeCambridge University Press
2Carr L2011Understanding Quantum Phase TransitionsBoca RatonCRC Press
3Quan H TSong ZLiu X FZanardi PSun C P 2006 Phys. Rev. Lett. 96 140604
4Zanardi PPaunković N 2006 Phys. Rev. E 74 031123
5You W LLi Y WGu S J2007Phys. Rev. E76022101
6Zanardi PGiorda PCozzini M 2007 Phys. Rev. Lett. 99 100603
7You W LHe L 2015 J. Phys.: Condens. Matter 27 205601
8Zhou H QBarjaktarevič J P 2008 J. Phys. A 41 412001
9Zhou H QOrǔs RVidal G 2008 Phys. Rev. Lett. 100 080601
10Gorin TProsen TSeligman T HŽnidarič M 2006 Phys. Rep. 435 33
11Paunković NSacramento PNogueira PVieira VDugaev V 2008 Phys. Rev. A 77 052302
12Lu X MSun ZWang XZanardi P 2008 Phys. Rev. A 78 032309
13Wang XSun ZWang Z D 2009 Phys. Rev. A 79 012105
14Gu S J 2009 Chin. Phys. Lett. 26 026401
15Gu S J 2010 Int. J. Mod. Phys. B 24 4371
16Gu S JYu W C 2014 Europhys. Lett. 108 20002
17Venuti L CZanardi P 2007 Phys. Rev. Lett. 99 095701
18Gu S JKwok H MNing W QLin H Q 2008 Phys. Rev. B 77 245109
19Albuquerque A FAlet FSire CCapponi S 2010 Phys. Rev. B 81 064418
20Rams MDamski B 2011 Phys. Rev. Lett. 106 055701
21Gu S JYu W CLin H Q 2013 Int. J. Mod. Phys. B 27 1350106
22Sacramento P DPaunković NVieira V R 2011 Phy. Rev. A 84 062318
23Wen X G2004Qunatum Field Theory of Many-body SystemsNew YorkOxford University
24Tasaki H2000arXiv:0009244[cond-mat]
25Kurchan J2007J. Stat. Mech. Theor. Exp.2007P07005
26Mukamel S 2003 Phys. Rev. Lett. 90 170604
27Pfeuty P 1970 Ann. Phys. 57 79
28Elliott R JPfeuty PWood C 1970 Phys. Rev. Lett. 25 443
29Jullien RPfeuty PFields J NDoniach S 1978 Phys. Rev. B 18 3568
30Barouch EMcCoy B M 1970 Phys. Rev. A 2 1075
31Barouch EMcCoy B M 1971 Phys. Rev. A 3 786
32Coldea RTennant DWheeler EWawrzynska EPrabhakaran DTelling MHabicht KSmeibidl PKiefer K 2010 Science 327 177
33Damski B 2013 Phys. Rev. E 87 052131
34Deng SOrtiz GViola L 2011 Phys. Rev. B 83 094304
35Rams MDamski B 2011 Phys. Rev. A 84 032324
36Kitaev A2006Ann. Phys.(N.Y.)443312
37Chen H DNussinov Z 2008 J. Phys. A: Math. Theor. 41 075001
38Feng X YZhang G MXiang T 2007 Phys. Rev. Lett. 98 087204
39Lee D HZhang G MXiang T 2007 Phys. Rev. Lett. 99 196805
40Baskaran GMandal SShankar R 2007 Phys. Rev. Lett. 98 247201
41Mondal SSen DSengupta K 2008 Phys. Rev. B 78 045101
42Shi X FYu YYou J QNori F 2009 Phys. Rev. B 79 134431
43Kitaev A 2003 Ann. Phys. 303 2
44Yang SGu S JSun C PLin H Q 2008 Phys. Rev. A 78 012304