entanglement content of quasi-particle excitations · dynamic bethe ansatz [39, 40] in fact suggest...

12
Entanglement Content of Quasi-Particle Excitations Olalla A. Castro-Alvaredo, 1 Cecilia De Fazio, 2 Benjamin Doyon, 3 and Istv´ an M. Sz´ ecs´ enyi 1 1 Department of Mathematics, City, University of London, 10 Northampton Square, EC1V 0HB, London, U.K. 2 Dipartimento di Fisica e Astronomia, Universit` a di Bologna, Via Irnerio 46, I-40126 Bologna, Italy 3 Department of Mathematics, King’s College London, Strand WC2R 2LS, London, U.K. (Dated: May 15, 2018) We investigate the quantum entanglement content of quasi-particle excitations in extended many- body systems. We show that such excitations give an additive contribution to the bi-partite von Neumann and R´ enyi entanglement entropies that takes a simple, universal form. It is largely in- dependent of the momenta and masses of the excitations, and of the geometry, dimension and connectedness of the entanglement region. The result has a natural quantum information theo- retic interpretation as the entanglement of a state where each quasi-particle is associated with two qubits representing their presence within and without the entanglement region, taking into account quantum (in)distinguishability. This applies to any excited state composed of finite numbers of quasi-particles with finite De Broglie wavelengths or finite intrinsic correlation length. We derive this result analytically in one-dimensional massive bosonic and fermionic free field theories and for simple setups in higher dimensions. We provide numerical evidence for the harmonic chain and the two-dimensional harmonic lattice in all regimes where excitations have quasi-particle properties. Finally, we provide supporting calculations for integrable spin chain models and other situations without particle production. Our results point to new possibilities for creating entangled states using many-body quantum systems. Introduction.— Measures of entanglement, such as the entanglement entropy (EE) [1] and entanglement negativity [2–7], have attracted much attention in re- cent years, both theoretically [8–10] and experimentally [11, 12]. Quantum entanglement encodes correlations be- tween degrees of freedom associated to independent fac- tors of the Hilbert space, and as such, it separates quan- tum correlations from the particularities of observables. As a consequence, the entanglement in extended systems encodes, in a natural fashion, universal properties of the state. For instance, at criticality, the entanglement of ground states provides an efficient measure of universal properties of quantum phase transitions, such as the (ef- fective) central change of the corresponding conformal field theory (CFT) and the primary operator content [13– 22]. Near criticality, it is universally controlled by the masses of excitations [23–25]. In states that are highly excited, with finite energy densities, the entanglement is known to give rise to local thermalisation effects: this is at the heart of the eigenstate thermalisation hypothesis [26–30], as the large entanglement between local degrees of freedom and the rest of the system effectively gener- ates a Gibbs ensemble (in the case of integrable systems, a generalized Gibbs ensemble). The entanglement effects of a finite number of excitations are less known. Some re- sults are available in critical systems: using the methods of Holzhey, Larsen and Wilczek [14], combining a geomet- ric description with Riemann uniformization techniques in CFT it was shown in [31, 32] that certain excitations, with energies tending to zero in the large volume limit, correct the ground state entanglement by power laws in the ratio of length scales. Various few-particle states have also been studied in special cases of integrable spin chains [33–37]. In this paper we propose a universal formula, with a simple quantum information theoretic interpretation, for the entanglement content of states with well-defined quasi-particle excitations. For this purpose, we study a variety of extended systems of different dimensions. We consider the von Neumann and R´ enyi EEs: these are measures of the amount of quantum entanglement, in a pure quantum state, between the degrees of freedom associated to two sets of independent observables whose union is complete on the Hilbert space. We use the setup where the Hilbert space is factorised as H A ⊗H B , accord- ing to two complementary spatial regions A and B, of typical length scales A and B , respectively (for dimen- sions higher than one, we can think of these as the di- ameters of the regions under consideration). The regions can be of generic geometry and connectedness. Quasi- particle excitations arise naturally in massive quantum field theory (QFT), where irreducible representations of the Poincar´ e group are identified with relativistic parti- cles. We first consider excited states of the massive free real boson and free Majorana fermion models, formed of finite numbers k of particles, at various momenta. We use techniques based on form factors of branch point twist fields in 1+1 dimensions as introduced in [23], and dimen- sional reduction methods [38] to access simple entangle- ment regions in higher dimensions. More generally, exci- tations in the harmonic chain and in higher-dimensional harmonic lattices can be interpreted as being composed of quasi-particles whenever the correlation length ξ is small enough, ξ min(A ,‘ B ), or the maximal De Broglie wavelength of the particles ζ = max{2π/|~ p i |}, where ~ p i are the particles’ momenta, is small enough, ζ min(A ,‘ B ). This includes the large-momenta re- gion of the chain’s or lattice’s CFT regime (beyond the applicability of the results of [31, 32]), as well as the non- universal regime, beyond QFT and CFT. We then study the harmonic chain and two-dimensional lattice numeri- cally. Quasi-particles also arise naturally in Bethe ansatz arXiv:1805.04948v1 [cond-mat.stat-mech] 13 May 2018

Upload: others

Post on 14-Oct-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Entanglement Content of Quasi-Particle Excitations · dynamic Bethe ansatz [39, 40] in fact suggest that the same calculation should also give the correct answer in massive integrable

Entanglement Content of Quasi-Particle Excitations

Olalla A. Castro-Alvaredo,1 Cecilia De Fazio,2 Benjamin Doyon,3 and Istvan M. Szecsenyi1

1Department of Mathematics, City, University of London,10 Northampton Square, EC1V 0HB, London, U.K.

2Dipartimento di Fisica e Astronomia, Universita di Bologna, Via Irnerio 46, I-40126 Bologna, Italy3Department of Mathematics, King’s College London, Strand WC2R 2LS, London, U.K.

(Dated: May 15, 2018)

We investigate the quantum entanglement content of quasi-particle excitations in extended many-body systems. We show that such excitations give an additive contribution to the bi-partite vonNeumann and Renyi entanglement entropies that takes a simple, universal form. It is largely in-dependent of the momenta and masses of the excitations, and of the geometry, dimension andconnectedness of the entanglement region. The result has a natural quantum information theo-retic interpretation as the entanglement of a state where each quasi-particle is associated with twoqubits representing their presence within and without the entanglement region, taking into accountquantum (in)distinguishability. This applies to any excited state composed of finite numbers ofquasi-particles with finite De Broglie wavelengths or finite intrinsic correlation length. We derivethis result analytically in one-dimensional massive bosonic and fermionic free field theories and forsimple setups in higher dimensions. We provide numerical evidence for the harmonic chain andthe two-dimensional harmonic lattice in all regimes where excitations have quasi-particle properties.Finally, we provide supporting calculations for integrable spin chain models and other situationswithout particle production. Our results point to new possibilities for creating entangled statesusing many-body quantum systems.

Introduction.— Measures of entanglement, such asthe entanglement entropy (EE) [1] and entanglementnegativity [2–7], have attracted much attention in re-cent years, both theoretically [8–10] and experimentally[11, 12]. Quantum entanglement encodes correlations be-tween degrees of freedom associated to independent fac-tors of the Hilbert space, and as such, it separates quan-tum correlations from the particularities of observables.As a consequence, the entanglement in extended systemsencodes, in a natural fashion, universal properties of thestate. For instance, at criticality, the entanglement ofground states provides an efficient measure of universalproperties of quantum phase transitions, such as the (ef-fective) central change of the corresponding conformalfield theory (CFT) and the primary operator content [13–22]. Near criticality, it is universally controlled by themasses of excitations [23–25]. In states that are highlyexcited, with finite energy densities, the entanglement isknown to give rise to local thermalisation effects: this isat the heart of the eigenstate thermalisation hypothesis[26–30], as the large entanglement between local degreesof freedom and the rest of the system effectively gener-ates a Gibbs ensemble (in the case of integrable systems,a generalized Gibbs ensemble). The entanglement effectsof a finite number of excitations are less known. Some re-sults are available in critical systems: using the methodsof Holzhey, Larsen and Wilczek [14], combining a geomet-ric description with Riemann uniformization techniquesin CFT it was shown in [31, 32] that certain excitations,with energies tending to zero in the large volume limit,correct the ground state entanglement by power laws inthe ratio of length scales. Various few-particle states havealso been studied in special cases of integrable spin chains[33–37].

In this paper we propose a universal formula, witha simple quantum information theoretic interpretation,

for the entanglement content of states with well-definedquasi-particle excitations. For this purpose, we study avariety of extended systems of different dimensions. Weconsider the von Neumann and Renyi EEs: these aremeasures of the amount of quantum entanglement, ina pure quantum state, between the degrees of freedomassociated to two sets of independent observables whoseunion is complete on the Hilbert space. We use the setupwhere the Hilbert space is factorised as HA⊗HB , accord-ing to two complementary spatial regions A and B, oftypical length scales `A and `B , respectively (for dimen-sions higher than one, we can think of these as the di-ameters of the regions under consideration). The regionscan be of generic geometry and connectedness. Quasi-particle excitations arise naturally in massive quantumfield theory (QFT), where irreducible representations ofthe Poincare group are identified with relativistic parti-cles. We first consider excited states of the massive freereal boson and free Majorana fermion models, formed offinite numbers k of particles, at various momenta. We usetechniques based on form factors of branch point twistfields in 1+1 dimensions as introduced in [23], and dimen-sional reduction methods [38] to access simple entangle-ment regions in higher dimensions. More generally, exci-tations in the harmonic chain and in higher-dimensionalharmonic lattices can be interpreted as being composedof quasi-particles whenever the correlation length ξ issmall enough, ξ � min(`A, `B), or the maximal DeBroglie wavelength of the particles ζ = max{2π/|~pi|},where ~pi are the particles’ momenta, is small enough,ζ � min(`A, `B). This includes the large-momenta re-gion of the chain’s or lattice’s CFT regime (beyond theapplicability of the results of [31, 32]), as well as the non-universal regime, beyond QFT and CFT. We then studythe harmonic chain and two-dimensional lattice numeri-cally. Quasi-particles also arise naturally in Bethe ansatz

arX

iv:1

805.

0494

8v1

[co

nd-m

at.s

tat-

mec

h] 1

3 M

ay 2

018

Page 2: Entanglement Content of Quasi-Particle Excitations · dynamic Bethe ansatz [39, 40] in fact suggest that the same calculation should also give the correct answer in massive integrable

2

integrable models, where the conditions ξ � min(`A, `B)and ζ � min(`A, `B) can be given precise meanings.We study few-particle excitations in generic states ofthe Bethe form. In this case, the calculation is elemen-tary, and complements some of the calculations done in[33]. Zamolodchikov’s ideas underpinning the thermo-dynamic Bethe ansatz [39, 40] in fact suggest that thesame calculation should also give the correct answer inmassive integrable QFT, thus including the effects of in-teractions. In all cases, we find a universal result in thelimit min(ξ, ζ)� min(`A, `B), independent of the modelstudied, of the connectedness or shape of the entangle-ment region, and of the dimension. The result extendsthe “semiclassical” form discussed in the context of spinchains in [33]. It has a very natural qubit interpretationwhere qubits representing the particles are entangled ac-cording to the particles’ distribution in space, taking intoaccount quantum indistinguishability in the bosonic case.The qubit interpretation indicates that quasi-particle ex-citations in many-body models may provide a simple wayof generating entangled states with easily adjustable pa-rameters.

0.0 0.2 0.4 0.6 0.8 1.00.00.10.20.30.40.50.60.7

r

DSn1

0.0 0.2 0.4 0.6 0.8 1.00.0

0.5

1.0

1.5

r

DSn6

FIG. 1. The functions (3) and (8) for k = 1, 6 and n =2, 3, 5, 8, 11, 17 and the limits n→ 1 (von Neumann) and n→∞ (single-copy). The outer-most curve is the von Neumannentropy and the inner-most curve is the single-copy entropy.

Results.— Consider a bi-partition of a system C =A ∪ B in state |Ψ〉 composed of a number k of quasi-particles. In infinite volume, the notion of quasi-particlesis very natural via the theory of scattering states [41, 42],and one expects that in finite but large volumes there arecorresponding excited states, unambiguously defined upto exponentially decaying corrections in the volume. Wedefine the reduced density matrix associated to subsys-tem A as ρA = TrB(|Ψ〉〈Ψ|). The Renyi EE is the Renyientropy of this reduced density matrix,

SΨn (A,B) =

log TrρnA1− n

. (1)

From (1) we may compute the von Neumann EE asSΨ

1 (A,B) := limn→1 SΨn (A,B) and the so-called single-

copy entropy [43–45] as SΨ∞(A,B) := limn→∞ SΨ

n (A,B).For large system size and fixed entanglement region, oneexpects the entanglement entropies to tend to those ofthe ground state. We therefore concentrate on the non-trivial limit where both the full system C and the en-tanglement region A are large, scaled simultaneously,A 7→ λA and B 7→ λB. Let r = Vold (A)/Vold (C) be

the ratio of the d-dimensional hypervolume of the re-gion to that of the system. We compute the difference∆SΨ

n (A,B) = SΨn (A,B) − S0

n(A,B) between the Renyientropy in the excited state |Ψ〉 and in the ground (vac-uum) state |0〉, in this limit,

∆SΨn (r) := lim

λ→∞∆SΨ

n (λA, λB). (2)

This is the contribution of the excitations to the entan-glement, or “excess entanglement” as named in [31, 32].

We find that for a wide variety of quantum systems,the results depend only on the proportion r of the sys-tem’s volume occupied by the entanglement region, andare largely independent of the momenta of the quasi-particles. Suppose the state is formed of k particles ofequal momenta. Denoting ∆SΨ

n (r) = ∆Skn(r), we find

∆Skn(r) =

logk∑q=0

fkq (r)n

1− n, ∆Sk1 (r) = −

k∑q=0

fkq (r) log fkq (r)

(3)

with fkq (r) =

(kq

)rq(1 − r)k−q. For a state composed

of k particles divided into groups of ki particles of equalmomenta ~pi, with i = 1, 2, . . . and

∑i ki = k, we denote

∆SΨn (r) = ∆Sk1,k2,...

n (r) and have

∆Sk1,k2,...n (r) =

∑i

∆Skin (r). (4)

In particular, for k particles of distinct momenta the re-sult is k times that for a single particle, which is

∆S1n(r) =

log(rn + (1− r)n)

1− n,

∆S11(r) = −r log r − (1− r) log(1− r). (5)

We observe that in all cases, the entanglement is max-imal at r = 1/2. For k distinct-momenta particles, themaximum is k log 2, while when some particles have coin-ciding momenta, the maximal value is smaller. Interest-ingly, single-copy entropies present non-analytic features.For distinct momenta, we have

∆S1∞(r) =

{− log(1− r) for 0 ≤ r < 1

2− log r for 1

2 ≤ r ≤ 1.(6)

Again, the result is just multiplied by k for a state con-sisting of k distinct-momentum particles. For equal mo-menta it is a function which is non-differentiable at kpoints in the interval r ∈ (0, 1) (generalizing (6)). Thepositions of these cusps are given by the values

r =1 + q

1 + kfor q = 0, . . . , k − 1, (7)

and the single copy entropy is given by

∆Sk∞(r) = − log fkq (r) forq

1 + k≤ r < 1 + q

1 + k(8)

and q = 0, . . . , k.

Page 3: Entanglement Content of Quasi-Particle Excitations · dynamic Bethe ansatz [39, 40] in fact suggest that the same calculation should also give the correct answer in massive integrable

3

The results take their full meaning under a quan-tum information theoretic interpretation that combinesa “semiclassical” picture of particles with quantum in-distinguishability. Consider a bi-partite Hilbert spaceH = Hint ⊗ Hext. Each factor Hint ' Hext is a ten-sor product ⊗iHki of Hilbert spaces Hki ' Cki+1 for kiindistinguishable qubits, with, as above,

∑i ki = k. We

associate Hint with the interior of the region A and Hext

with its exterior, and we identify the qubit state 1 withthe presence of a particle and 0 with its absence. We con-struct the state |Ψqb〉 ∈ H under the picture accordingto which equal-momenta particles are indistinguishable,and a particle can lie anywhere in the full volume of thesystem with flat probability. That is, any given particlehas probability r of lying within A, and 1−r of lying out-side of it. We make a linear combination of qubit statesfollowing this picture, with coefficients that are squareroots of the total probability of a given qubit configu-ration, taking proper care of (in)distinguishability. Forinstance, for a single particle,

|Ψqb〉 =√r |1〉 ⊗ |0〉+

√1− r |0〉 ⊗ |1〉 (9)

as either the particle is in the region, with probability r,or outside of it, with probability 1 − r. If two particlesof coinciding momenta are present, then we have

|Ψqb〉 =√r2 |2〉 ⊗ |0〉+

√2r(1− r) |1〉 ⊗ |1〉

+√

(1− r)2 |0〉 ⊗ |2〉 (10)

as either the two particles are in the region, with proba-bility r2, or one is in the region and one outside of it (nomatter which one), with probability 2r(1 − r), or bothare outside the region, with probability (1− r)2. For twoparticles of different momenta,

|Ψqb〉 =√r2 |11〉 ⊗ |00〉+

√(1− r)2 |00〉 ⊗ |11〉

+√r(1− r) (|10〉 ⊗ |01〉+ |01〉 ⊗ |10〉) (11)

counting the various ways two distinct particles can bedistributed inside or outside the region. Higher-particlestates can be constructed similarly. The results statedabove are then equivalent to the identification ∆SΨ

n (r) =

∆SΨqbn (r), where

∆SΨqbn (r) =

log Trρnint

1− n, ρint = Trext|Ψqb〉〈Ψqb|. (12)

Methods.— In general, the quantity ∆SΨn can be com-

puted using the replica method [13, 14]. In this context,one evaluates traces of powers of the reduced density ma-trix ρA. After reinterpretation of such traces, this boilsdown to ratios of expectation values of a twist operator,acting on a replica model composed of n independentcopies of the original theory. The operator T(A,B) actsas a cyclic permutation of the copies i 7→ i+ 1 mod n onHA, and as the identity on HB , and (2) is expressed as

∆SΨn (r) = lim

λ→∞

1

1− nlog

[n〈Ψ|T(λA, λB)|Ψ〉nn〈0|T(λA, λB)|0〉n

], (13)

where |0〉n is the vacuum state. Both |0〉n and the state|Ψ〉n have the structure

|Ψ〉n = |Ψ〉1 ⊗ |Ψ〉2 ⊗ · · · ⊗ |Ψ〉n. (14)

Here |Ψ〉i ' |Ψ〉 is the k-particle excited state of interest,implemented in the ith copy.

In one dimension, A is in general a union of segments.Then, T(A,B) is expressed as a product of branch-pointtwist fields [23], supported on the boundary points ofthese segments. Branch point twist fields are twist fieldsassociated to the cyclic permutation, a symmetry of thereplica model. Let us consider the case A = [0, `] in asystem of length L. Then T(A,B) = T (0)T (`), whereT is the branch point twist field and T is its hermitianconjugate. Expression (13) may be used by expandingtwo-point functions of branch-point twist fields in thebasis |Φ〉 of quasi-particles,

n〈Ψ|T (0)T (`)|Ψ〉n =∑Φ

e−iPΦ`|n〈Ψ|T (0)|Φ〉|2 (15)

where PΦ are the momentum eigenvalues (in finite vol-ume, they are quantised, and the set of states is discrete).

Using (13) with (15) in integrable 1+1-dimensionalQFT presents however a number of challenges. Via anextension of the form factor program [46, 47], matrix el-ements of branch-point twist fields in infinite volume areknown exactly [23], and have been used successfully inthe vacuum. But they cannot be used in order to evalu-ate the limit L → ∞ in (13), as in excited states, diver-gencies occur in the expansion (15) whenever momentaof intermediate particles in |Φ〉 coincide with momenta ofparticles in the state |Ψ〉n. One must first evaluate finite-volume matrix elements, re-sum the series (15), and thentake the limit. Finite-volume matrix elements of ordinarylocal fields in integrable QFT have been studied recently[48, 49]. They are simply related to infinite-volume ma-trix elements up to exponentially decaying terms in L.They are evaluated at momenta that are quantised ac-cording to the Bethe-Yang equations based on the two-particle scattering matrix of the integrable model. Butfor twist fields, the theory has not been developed yet.In particular, the twist properties affect the quantisa-tion condition of the individual momenta of the quasi-particles. We have solved these problems for the massivefree real boson and the massive free Majorana fermionin periodic space. By performing the summation overintermediate states at large L, noting that the so-called“kinematic singularities” of infinite-volume matrix ele-ments provide the leading contribution, we have derivedthe full results (3) and (4). The details are technical, andpresented in a separate paper [50].

The qubit interpretation presented earlier suggeststhat our results need not be restricted to one-dimensionalQFT. To test this claim, we performed a numerical eval-uation of the quantity ∆SΨ

n (r) in the harmonic chain andthe two-dimensional harmonic lattice. We present someof our results in FIG. 2 and in the supplementary material(SM). We used wave-functional methods in order to rep-resent the state |Ψ〉. In the finite-volume Klein-Gordon

Page 4: Entanglement Content of Quasi-Particle Excitations · dynamic Bethe ansatz [39, 40] in fact suggest that the same calculation should also give the correct answer in massive integrable

4

´

´

´

´´ ´ ´

´

´

´

´á

á

á

á

áá á

á

á

á

á´

´

´

´

´´ ´

´

´

´

´á

á

á

á

á

áá

á

á

á

á

0.0 0.2 0.4 0.6 0.8 1.00.00.20.40.60.81.01.21.4

r

S2YHrL

ææ

ææ

ææ

ææ

ææ

ææ

ææ

ææææææ

-0.5 0.0 0.5 1.0 1.5 2.0

-1

0

1

2

3

log@mLD

-log@log 2-DS21H1�2LD

FIG. 2. Numerical results for the case n = 2 on the toriclattice [0, L]2 with L = 50 and lattice spacing ∆x = 1.Left. Two-particle states |Ψ〉 = |~p1, ~p2〉. Squares are formass m = 1 and small momenta, crosses are for massm = 0.001 and large momenta. The upper curve is formula(4) for ∆S1,1

2 , with numerical results for distinct momenta~p1 = (0, 0), ~p2 = (0.26, 0) = (4π/L, 0) (squares) and ~p1 =(2.51, 1.26) = (40π/L, 20π/L), ~p2 = (3.14, 0) = (50π/L, 0)(crosses). The lower curve is formula (3) for ∆S2

2 , with numer-ical results for equal momenta ~p1 = ~p2 = (0.13, 0) = (2π/L, 0)(squares) and ~p1 = ~p2 = (2.51, 1.26) (crosses). Right. Ap-proach of ∆S1

2(1/2) to the analytical value log 2 for the one-particle state |Ψ〉 = |~p〉 with ~p = (0, 0) as a function of L.This shows a linear approach for large mL. The solid lineis the fit 0.527 − 1.783 log(mL) on the last 8 data points(mL ∈ [6.5, 10]).

theory, the vacuum wave functional takes the form

〈ϕ|0〉 ∝ exp

[−1

2

∫C×C

ddxddy K(~x− ~y)ϕ(~x)ϕ(~y)

](16)

where K(~x) =∑

~p Vold (C)−1E~p ei~p·~x. Excited state

wave functionals are obtained by applying the operator

A†(~p) =

∫C

ddx ei~p·~x (E~pϕ(~x)− i$(~x))√2E~p Vold (C)

, [A~p, A†~q] = δ~p,~q,

(17)

(where E~p =√~p2 +m2) with the representation of

the canonical momentum $(~x) = −iδ/δϕ(~x) satisfying[ϕ(~x), $(~y)] = iδ(~x−~y). This generates factors which arepolynomial functionals of ϕ(~x). The operator T(A,B) iseasily implemented on the space of field configurations.The ratio (13) then becomes a Gaussian average of poly-nomial functionals of the fields. Discretising space to afinite lattice spacing ∆x modifies the dispersion relationto E2

~p = m2 + 4∑di=1 sin2(pi∆x/2)/(∆x)2. Numerical

results in the one-dimensional case are discussed in moredetail in [50], where both QFT and non-universal pa-rameter regimes are seen to agree with our predictions,for connected and disconnected regions. We concentratehere on the two-dimensional periodic square lattice onC = [0, L]2. We choose a set of subregions A = [0, `]2

for values of r = `2/L2 ranging between 0 and 1. In or-der to establish the validity of the requirements on thecorrelation length ξ and the De Broglie wavelength ζ, weexplore two distinct regimes: that of small ξ but large ζ,and that of small ζ but large ξ, in both cases looking attwo-particle states with equal and with distinct rapidi-ties. We find excellent agreement with formulae (4) for∆S1,1

n and (3) for ∆S2n, respectively, see FIG. 2. Note

that the configuration we have chosen is not symmet-ric: regions A and B have different shapes. Nevertheless,

the symmetry r 7→ 1 − r is correctly recovered in theregime of validity of formulae (3) and (4). We have ex-plored other shapes of the region A, obtaining similaraccuracy, and have analysed regimes where both ξ andζ are small, finding even greater accuracy. We have alsoanalysed the breaking of formulae (3) and (4) away fromtheir regime of validity. The approach to the maximumlog 2 in the case of a single particle with r = 1/2 (thismaximal value is supported by general arguments [34]) isshown in FIG. 2, where the correlation length is varied;we observe an algebraic approach at large mL.

For particular choices of the region A, it is possible toshow analytically the results (3–8) in the massive Klein-Gordon and Majorana models of any higher dimension,by dimensional reduction [38]. Consider the slab-like re-gions A = [0, `] × C⊥ in C = [0, L] × C⊥ where C⊥ issome d − 1-dimensional space. From the d-dimensionalKlein-Gordon fields ϕ(~x, t) and $(~x, t) construct

ϕ(x1, t) =

∫C⊥

dd−1x⊥ϕ(x1, ~x⊥, t)√Vold−1 (C⊥)

(18)

and similarly for $(x1, t) in terms of $(~x, t). One ob-serves that ϕ(x, t) and $(x, t) are canonically normalisedone-dimensional Klein-Gordon fields, and that the di-mensional reduction map preserves the vacuum [38]. Italso preserves the many-particle states when all momentapoint in the x1 direction: with ~p = (p, 0, . . . , 0) the ex-pression (17) gives A†(~p) = A†(p). Therefore, the quan-tity n〈Ψ|T(λA, λB)|Ψ〉n in d dimensions, is proportionalto n〈Ψ|T (0)T (`)|Ψ〉n in 1 dimension. The singularityas ` → 0 is dimension dependent, but in the ratio (13),this cancels out, and there is exact equality. See [50] formore details. This analysis extends to other quasi-one-dimensional configurations.

Finally, we establish that our results hold beyond freetheories. We analyse the quantity ∆SΨ

n (r) in interact-ing states of the Bethe ansatz form. Previous analysesexist [33, 37], which however concentrated on less univer-sal regimes. In the ferromagnetic Heisenberg chain, two-particle states with respect to the ferromagnetic vacuumhave the simple form

∑x,y∈Z e

ipx+iqySsgn(x−y)(p, q)| ↑· · · ↓x · · · ↑ · · · ↓y · · · ↑〉, where Sε(p, q) is the Betheansatz scattering matrix. As in the context of the ther-modynamic Bethe ansatz formalism of integrable QFT[39, 40], for the purpose of evaluating large-distancequantities these are abstract states representing two-particle asymptotic states, with Sε(p, q) the two-bodyscattering matrix of the field theory. States of the Betheansatz form thus are expected to provide large-distanceresults of great generality in integrable models. We haveanalysed such one- and two-particle states, and foundthat formulae (3) and (4) hold, see the SM. There is noneed to fix the momenta via the Bethe ansatz; with equalmomenta S2

n(r) is indeed reproduced, extending previousresults. Bound states of the Heisenberg chains (Bethestrings) have been studied [33]; these have an intrinsiclength scale ξ (inversely proportional to the boundingenergy), and one can see that in the regimes discussedabove, S1

n(r) is indeed reproduced.

Page 5: Entanglement Content of Quasi-Particle Excitations · dynamic Bethe ansatz [39, 40] in fact suggest that the same calculation should also give the correct answer in massive integrable

5

Discussion.— It is remarkable that the entanglementof a wide variety of many-body quantum systems admitssuch a simple and universal “qubit” interpretation. Thiscombines a semiclassical picture of localised particlescontrolled by correlation lengths and De Broglie wave-lengths, with the quantum effect of (in)distinguishability.The applicability of equations (3–8) to higher dimensionsis particularly significant, showing that a large amount ofgeometric information is irrelevant. Their application toQFT is also interesting: QFT locality is formally basedon the vanishing of space-like commutation relations, noton particles, yet our results show how quantum entangle-ment clearly “sees” localised particles. The results holdbeyond the QFT regime, as we checked in quantum har-monic lattices and in Bethe ansatz excitations of quan-tum spin chains. Going beyond integrability, the resultsare fully expected to hold when no particle productionoccurs, for instance in QFT one-particle states, and two-particle states below the particle production threshold.In fact, any one- and two-particle excitations of Bethe-ansatz form will have EE described by (3)-(8), such as inspin-preserving quantum chains, integrable or not. Therelation (12) also suggests that quasi-particle excitationsin extended systems of any dimension can be used tocreate simple entangled states with controllable entan-glement, where the control parameter is the region-to-system volume ratio r. It would be interesting to investi-gate the possible applications of such a result in the areaof quantum information.

Acknowledgements: We are grateful to EPSRC forfunding through the standard proposal “Entangle-ment Measures, Twist Fields, and Partition Functionsin Quantum Field Theory” under reference numbersEP/P006108/1 and EP/P006132/1. We would also liketo thank Vincenzo Alba for discussions and for bringingreference [33] to our attention.

[1] C. H. Bennett, H. J. Bernstein, S. Popescu, and B. Schu-macher, Concentrating partial entanglement by local op-erations, Phys. Rev. A53, 2046–2052 (1996).

[2] G. Vidal and R. F. Werner, Computable measure ofentanglement, Phys. Rev. A65, 032314 (2002).

[3] K. Zyczkowski, P. Horodecki, A. Sanpera, and M. Lewen-stein, Volume of the set of separable states, Phys. Rev.A58, 883–892 (1998).

[4] M. B. Plenio, Logarithmic Negativity: A full entangle-ment monotone that is not convex, Phys. Rev. Lett. 95,090503 (2005).

[5] M. B. Plenio, Erratum: logarithmic negativity: a fullentanglement monotone that is not convex, Phys. Rev.Lett. 95, 119902 (2005).

[6] K. Audenaert, J. Eisert, M. B. Plenio, and R. F. Werner,Entanglement properties of the harmonic chain, Phys.Rev. A66, 042327 (2002).

[7] J. Eisert, Entanglement in quantum information theory(PhD Thesis), quant-ph/0610253 (2006).

[8] L. Amico, R. Fazio, A. Osterloh, and V. Vedral, En-tanglement in many-body systems, Rev. Mod. Phys. 80,517–576 (2008).

[9] P. Calabrese, J. Cardy, and B. Doyon (ed), Entanglemententropy in extended quantum systems, J. Phys. A42,500301 (2009).

[10] J. Eisert, M. Cramer, and M. B. Plenio, Colloquium:Area laws for the entanglement entropy, Rev. Mod. Phys.82, 277–306 (2010).

[11] R. Islam, R. Ma, P. M. Preiss, M. E. Tai, A. Lukin,M. Rispoli, and M. Greiner, Measuring entanglement en-tropy in a quantum many-body system, Nature 528, 77–83 (2015).

[12] A. M. Kaufman, M. E. Tai, A. Lukin, M. Rispoli, R.Schittko, P. M. Preiss, and M. Greiner, Quantum ther-malization through entanglement in an isolated many-body system, Science 353, 794–800 (2016).

[13] C. J. Callan and F. Wilczek, On geometric entropy, Phys.Lett. B333, 55–61 (1994).

[14] C. Holzhey, F. Larsen, and F. Wilczek, Geometric andrenormalized entropy in conformal field theory, Nucl.Phys. B424, 443–467 (1994).

[15] G. Vidal, J. I. Latorre, E. Rico, and A. Kitaev, Entangle-ment in quantum critical phenomena, Phys. Rev. Lett.90, 227902 (2003).

[16] J. I. Latorre, E. Rico, and G. Vidal, Ground state entan-glement in quantum spin chains, Quant. Inf. Comput. 4,48–92 (2004).

[17] P. Calabrese and J. L. Cardy, Entanglement entropy andquantum field theory, J. Stat. Mech. 0406, P002 (2004).

[18] P. Calabrese and J. L. Cardy, Evolution of entanglemententropy in one-dimensional systems, J. Stat. Mech. 0504,P010 (2005).

[19] P. Calabrese, J. Cardy, and E. Tonni, Entanglement en-tropy of two disjoint intervals in conformal field theory,J. Stat. Mech. 0911, P11001 (2009).

[20] P. Calabrese, J. Cardy, and E. Tonni, Entanglement neg-ativity in quantum field theory, Phys. Rev. Lett. 109,130502 (2012).

[21] P. Calabrese, J. Cardy, and E. Tonni, Entanglement neg-ativity in extended systems: a field theoretical approach,J. Stat. Mech. 1302, P02008 (2013).

[22] D. Bianchini, O. Castro-Alvaredo, B. Doyon, E. Levi,and F. Ravanini, Entanglement entropy of non-unitaryconformal field theory, J.Phys. A48, 04FT01 (2015).

[23] J. L. Cardy, O. A. Castro-Alvaredo, and B. Doyon, Formfactors of branch-point twist fields in quantum integrablemodels and entanglement entropy, J. Stat. Phys. 130,129–168 (2008).

[24] B. Doyon, Bi-partite entanglement entropy in massivetwo-dimensional quantum field theory, Phys. Rev. Lett.102, 031602 (2009).

[25] O. Blondeau-Fournier, O.A. Castro-Alvaredo, andB. Doyon, Universal scaling of the logarithmic negativ-ity in massive quantum field theory, J. Phys. A49(12),125401 (2016).

[26] R. V. Jensen and R. Shankar, Statistical behavior indeterministic quantum systems with few degrees of free-dom, Phys. Rev. Lett. 54, 1879–1882 (1985).

[27] J. M. Deutsch, Quantum statistical mechanics in a closedsystem, Phys. Rev. A43, 2046–2049 (1991).

[28] M. Srednicki, Chaos and quantum thermalization, Phys.Rev. E 50, 888–901 (1994).

[29] H. Tasaki, From quantum dynamics to the canonical dis-tribution: general picture and a rigorous example, Phys.Rev. Lett. 80, 1373–1376 (1998).

[30] M. Rigol, V. Dunjko, and M. Olshanii, Thermalizationand its mechanism for generic isolated quantum systems,Nature 452, 854–858 (2008).

[31] F. C. Alcaraz, M. I. Berganza, and G. Sierra, Entangle-ment of low-energy excitations in conformal field theory,

Page 6: Entanglement Content of Quasi-Particle Excitations · dynamic Bethe ansatz [39, 40] in fact suggest that the same calculation should also give the correct answer in massive integrable

6

Phys. Rev. Lett. 106, 201601 (2011).[32] M. I. Berganza, F. C. Alcaraz, and G. Sierra, Entan-

glement of excited states in critical spin chains, J. Stat.Mech. 1201, P01016 (2012).

[33] J. Molter, T. Barthel, U. Schollwock, and V. Alba, Boundstates and entanglement in the excited states of quantumspin chains, J. Stat. Mech. 2014(10), P10029 (2014).

[34] I. Pizorn, Universality in entanglement of quasiparticleexcitations, arXiv:1202.3336 (2012).

[35] M. Storms and R. R. P. Singh, Entanglement in groundand excited states of gapped free-fermion systems andtheir relationship with Fermi surface and thermodynamicequilibrium properties, Phys. Rev. E89, 012125 (2014).

[36] R. Berkovits, Two-particle excited states entanglemententropy in a one-dimensional ring, Phys. Rev. B87,075141 (2013).

[37] V. Alba, M. Fagotti, and P. Calabrese, Entanglement en-tropy of excited states, J. Stat. Mech 2009(10), P10020(2009).

[38] B. Doyon, A. Lucas, K. Schalm, and M. J. Bhaseen, Non-equilibrium steady states in the Klein-Gordon theory, J.Phys. A48(9), 095002 (2015).

[39] A. Zamolodchikov, Thermodynamic Bethe ansatz in rel-ativistic models. Scaling three state Potts and Lee-Yangmodels, Nucl. Phys. B342, 695–720 (1990).

[40] T. R. Klassen and E. Melzer, The Thermodynamics ofpurely elastic scattering theories and conformal pertur-

bation theory, Nucl. Phys. B350, 635–689 (1991).[41] R. Eden, P. Landshoff, D. Olive, and J. C. Polkinghorne,

The analytic S-matrix, Cambridge University Press(1966).

[42] M. Peskin and D. Schroder, An introduction to quantumfield theory, CRC Press (1995).

[43] J. Eisert and M. Cramer, Single-copy entanglement incritical quantum spin chains, Phys. Rev. A 72, 042112(2005).

[44] I. Peschel and J. Zhao, On single-copy entanglement, J.Stat. Mech. 2005(11), P11002 (2005).

[45] A. Dimic and B. Dakic, Single-copy entanglement detec-tion, Nature Quantum Information 1(4), 11 (2018).

[46] M. Karowski and P. Weisz, Exact S matrices and form-factors in (1+1)-dimensional field theoretic models withsoliton behavior, Nucl. Phys. B139, 455–476 (1978).

[47] F. Smirnov, Form factors in completely integrable modelsof quantum field theory, Adv. Series in Math. Phys. 14,World Scientific, Singapore (1992).

[48] B. Pozsgay and G. Takacs, Form-factors in finite volumeI: form-factor bootstrap and truncated conformal space,Nucl. Phys. B788, 167–208 (2008).

[49] B. Pozsgay and G. Takacs, Form factors in finite vol-ume. II. Disconnected terms and finite temperature cor-relators, Nucl. Phys. B788, 209–251 (2008).

[50] O.A. Castro-Alvaredo, C. De Fazio, B. Doyon, andI.M. Szecsenyi, Entanglement content of quantum parti-cle excitations I. Free field theory, work in progress.

Supplementary Material

Renyi Entropy of one- and two-Magnon States in Gapped Quantum Spin Chains

In this section we present a derivation of the nth Renyi entropy for a one-magnon state and the 2nd Renyi entropyof a two-magnon state in a generic gapped quantum spin- 1

2 chain of length L. There is some overlap with calculationspresented in the appendices of [33] but the focus is slightly different here. An obvious example would be the XXZmodel in the gapped regime but other models can also be included, as long as their excited states can be representedas

|Ψ〉1 =1√L

L∑j=1

eipj |j〉, (19)

and

|Ψ〉2 =1√N

L∑j1,j2=1

Sj1j2eip1j1+ip2j2 |j1j2〉, (20)

where

Sj1j2 = (S∗j1j2)−1 =

eiϕ for j1 > j21 for j1 < j20 for j1 = j2

(21)

for a two-particle excited state. The states |j〉 and |j1j2〉 represent tensor product states where all spins are up, exceptthe spin at position j or the spins at positions j1, j2, respectively. The normalization N of the two-particle excitedstate will be discussed below. In our computations we will consider a bi-partition such that

spins {1, . . . , `} ∈ A and spins {`+ 1, . . . , L} ∈ B (22)

The choice of the spins is relevant in the computations below as it determines the value of the scattering phases (21).However, it is easy to show that this choice is not essential. That is, the results would be identical in the regions Aand B are not simply-connected. Finally we note that we will use the notation

r :=`

L(23)

to denote the dimensionless ratio of length scales in the problem.

Page 7: Entanglement Content of Quasi-Particle Excitations · dynamic Bethe ansatz [39, 40] in fact suggest that the same calculation should also give the correct answer in massive integrable

7

One-Magnon State

Consider the state (19). The basic states are normalized as 〈j1|j2〉 = δj1j2 and therefore the state above has norm1. We would like to compute the Renyi entropy of region A for the state |Ψ〉. This computation is actually verysimple and the results have been presented for instance in [36]. Let us first compute the reduced density matrix

ρA = TrB(|Ψ〉11〈Ψ|). (24)

This is

ρA =1

LTrB

L∑j,k=1

eip(j−k)|j〉〈k|

=

1

L

∑j,k∈A

eip(j−k)|j〉〈k|TrB (|0〉〈0|) +1

L|0〉〈0|

∑j,k∈B

eip(j−k)TrB (|j〉〈k|)

+1

L

∑j∈A

eipj |j〉〈0|∑k∈B

e−ipkTrB (|0〉〈k|) +1

L

∑k∈A

e−ipk|0〉〈k|∑j∈B

eipjTrB (|j〉〈0|)

=1

L

∑j,k∈A

eip(j−k)|j〉〈k|+ L− `L|0〉〈0|, (25)

where |0〉 represents the vacuum state. To compute the Renyi entropy we need to take the n-th power of the matrixabove and then the trace thereof. When doing so the two contributions do not mix so we can write:

TrAρnA = TrA(

1

L

∑j,k∈A

eip(j−k)|j〉〈k|)n + TrA(L− `L|0〉〈0|)n (26)

The first term is

TrA(1

L

∑j,k∈A

eip(j−k)|j〉〈k|)n = TrA

1

Ln

∑k1,...,kn∈Aj1,...,jn∈A

eipj1 |j1〉

[n−1∏i=1

eip(ji+1−ki)〈ki|ji+1〉

]e−ipkn〈kn|

= TrA

1

Ln

∑k1,...,kn∈Aj1,...,jn∈A

eipj1 |j1〉

[n−1∏i=1

n−1∏i=1

eip(ji+1−ki)δkiji+1

]e−ipkn〈kn|

=

1

Ln

∑k1,...,kn∈Aj1,...,jn∈A

δj1kn

[n−1∏i=1

δkiji+1

]=

`n

Ln= rn. (27)

The second term is simply

TrA(L− `L|0〉〈0|)n =

(L− `)n

Ln= (1− r)n. (28)

Which gives the known formula for the entanglement of a single excitation:

S1n(r) =

log(rn + (1− r)n)

1− n. (29)

Note that if the ground state is factorizable (as in the XXZ gapped chain) and therefore has zero entropy the resultgives us the entanglement of the excited state directly. Note that all functions (29) have a maximum for r = 1

2

S1n(1/2) = log 2, (30)

including the limits n → 1 and n → ∞ (see FIG. 1). The fact that one-particle excitations contribute log 2 to theentanglement entropy was studied in detail in [34] where it was shown to be case even for a non-integrable spin chainmodel.

Two-Magnon State

The effect of the presence of non-trivial interaction can be explored by considering a two-magnon state such as (20).

Page 8: Entanglement Content of Quasi-Particle Excitations · dynamic Bethe ansatz [39, 40] in fact suggest that the same calculation should also give the correct answer in massive integrable

8

State Normalization

We will start by fixing the normalization of the state, N . The norm is

2〈Ψ|Ψ〉2 =1

N

L∑j1,j2,j′1j

′2=1

Sj1j2S∗j′1j

′2eip1(j1−j′1)+ip2(j2−j′2)〈j′1j′2|j1j2〉

=1

N

L∑j1,j2,j′1j

′2=1

Sj1j2S∗j′1j

′2eip1(j1−j′1)+ip2(j2−j′2)[δj1j′1δj2j′2 + δj1j′2δj2j′1 ]

=1

N

L∑j1,j2=1

[1− δj1j2 + Sj1j2S

∗j2j1e

ip12j12], (31)

where p12 := p1 − p2 and j12 := j1 − j2. Using the definition of the S-matrix we have that we can rewrite the sum as

2〈Ψ|Ψ〉2 =L(L− 1)

N+

1

N

L∑j1=2

j1−1∑j2=1

[Sj1j2S

∗j2j1e

ip12j12 + Sj2j1S∗j1j2e

ip12j21]

=L(L− 1)

N+

1

N

L∑j1=2

j1−1∑j2=1

[eiϕeip12j12 + e−iϕeip12j21

]=L(L− 1)

N+

1

N

L∑j1=2

j1−1∑j2=1

[2 cos(ϕ+ p12j12)]

=L(L− 1)

N+

1

N

(L− 1) cosϕ− L cos(ϕ+ p12) + cos(ϕ+ Lp12)

cos p12 − 1. (32)

Thus,

N = L(L− 1) +(L− 1) cosϕ− L cos(ϕ+ p12) + cos(ϕ+ Lp12)

cos p12 − 1. (33)

This same formula was given also in one of the appendices in [33]. Clearly, for L large and p12 6= 0, the secondcontribution is sub-leading so that

N ≈ L2 for L large. (34)

Computation of the Reduced Density Matrix

We now need to proceed as for the first example. First we compute the reduce density matrix

ρA =1

NTrB

L∑j1,j2,k1,k2=1

Sj1j2S∗k1k2

eip1(j1−k1)+ip2(j2−k2)|j1j2〉〈k1k2|

. (35)

The only non-vanishing contributions to this trace come from those terms were either all indices are in A, all are inB or two indices are in A and two indices are in B. This gives the following six contributions:

ρA =1

N

∑j1,j2,k1,k2∈A

Sj1j2S∗k1k2

eip1(j1−k1)+ip2(j2−k2)|j1j2〉〈k1k2|TrB(|0〉〈0|)

+1

N

∑j1,j2,k1,k2∈B

Sj1j2S∗k1k2

eip1(j1−k1)+ip2(j2−k2)|0〉〈0|TrB (|j1j2〉〈k1k2|)

+1

N

∑j1,k1∈A,j2,k2∈B

Sj1j2S∗k1k2

eip1(j1−k1)+ip2(j2−k2)|j1〉〈k1|TrB (|j2〉〈k2|)

+1

N

∑j1,k2∈A,j2,k1∈B

Sj1j2S∗k1k2

eip1(j1−k1)+ip2(j2−k2)|j1〉〈k2|TrB (|j2〉〈k1|)

+1

N

∑j2,k1∈A,j1,k2∈B

Sj1j2S∗k1k2

eip1(j1−k1)+ip2(j2−k2)|j2〉〈k1|TrB (|j1〉〈k2|)

+1

N

∑j2,k2∈A,j1,k1∈B

Sj1j2S∗k1k2

eip1(j1−k1)+ip2(j2−k2)|j2〉〈k2|TrB (|j1〉〈k1|) (36)

Page 9: Entanglement Content of Quasi-Particle Excitations · dynamic Bethe ansatz [39, 40] in fact suggest that the same calculation should also give the correct answer in massive integrable

9

Noting that

TrB(|j1〉〈k1|) = δj1k1, TrB(|j1j2〉〈k1k2|) = δj1k1

δj2k2+ δj1k2

δj2k1, (37)

and grouping all four last terms together by relabelling the summation indices, the reduced density matrix simplifiesto:

ρA =1

N

∑j1,j2,k1,k2∈A

Sj1j2S∗k1k2

eip1(j1−k1)+ip2(j2−k2)|j1j2〉〈k1k2|

+1

N

∑j1,j2,k1,k2∈B

Sj1j2S∗k1k2

eip1(j1−k1)+ip2(j2−k2)|0〉〈0| (δj1k1δj2k2

+ δj1k2δj2k1

)

+1

N

∑j1,k1∈A,j2,k2∈B

(Sj1j2S∗k1k2

eip1(j1−k1)+ip2(j2−k2) + Sj1j2S∗k2k1

eip1(j1−k2)+ip2(j2−k1)

+Sj2j1S∗k1k2

eip1(j2−k1)+ip2(j1−k2) + Sj2j1S∗k2k1

eip1(j2−k2)+ip2(j1−k1))|j1〉〈k1| δj2k2. (38)

In the last sum all S-matrices have one index in region A and the other in region B so we can use the S-matrixdefinition to simplify the expression. Applying the δ-functions we can write,

ρA =1

N(ρ

(1)A + ρ

(2)A + ρ

(3)A ), (39)

where we introduced the matrices

ρ(1)A =

∑j1,j2,k1,k2∈A

Sj1j2S∗k1k2

eip1(j1−k1)+ip2(j2−k2)|j1j2〉〈k1k2| (40)

ρ(2)A =

∑j1,j2∈B

(Sj1j2S∗j1j2 + Sj1j2S

∗j2j1e

ip12j12)|0〉〈0| (41)

ρ(3)A =

∑j2∈B

j1,k1∈A

[eip1(j1−k1) + eip2(j1−k1) + eip1j12+ip2(j2−k1)−iϕ + eip2j12+ip1(j2−k1)+iϕ]|j1〉〈k1|. (42)

2nd Renyi Entropy

To compute the second Renyi entropy we must compute the square of the expression above and then take the traceover subspace A. It is easy to show that the only contributions to the trace will come from the squaring each of thethree terms above separately, so we can ignore cross-terms that will vanish under the trace.

For the first term squaring and taking the trace gives

TrA((ρ(1)A )2) =

∑j′1,j′2,k′

1,k′2∈A

j1,j2,k1,k2∈A

Sj1j2S∗k1k2

Sj′1j′2S∗k′1k

′2eip1(j1+j′1−k1−k′1)+ip2(j2+j′2−k2−k′2)

×(δj1k′1δj2k′2 + δj1k′2δj2k′1)(δk1j′1δk2j′2

+ δk1j′2δk2j′1

)

=∑

j1,j2,k1,k2∈A

Sj1j2S∗k1k2

eip1(j1−k1)+ip2(j2−k2)

×(Sk1k2

S∗j1j2eip1(k1−j1)+ip2(k2−j2) + Sk2k1

S∗j1j2eip1(k2−j1)+ip2(k1−j2)

+Sk1k2S∗j2j1e

ip1(k1−j2)+ip2(k2−j1) + Sk2k1S∗j2j1e

ip1(k2−j2)+ip2(k1−j1)), (43)

where, in the first equality, we used the fact that

TrA(|j1j2〉〈k1k2|j′1j′2〉〈k′1k′2|) = (δj1k′1δj2k′2 + δj1k′2δj2k′1)(δk1j′1δk2j′2

+ δk1j′2δk2j′1

). (44)

Employing the explicit form of the S-matrices, the expression can be greatly simplified and factorizes as

TrA((ρ(1)A )2) =

[ ∑k1k2∈A

(1− δk1k2

+ S∗k1k2Sk2k1

eip12k21)]2

. (45)

Page 10: Entanglement Content of Quasi-Particle Excitations · dynamic Bethe ansatz [39, 40] in fact suggest that the same calculation should also give the correct answer in massive integrable

10

The sum over the S-matrices can be computed by for instance splitting it into the contribution with k1 > k2 and thecontribution with k1 < k2. For example∑

k1k2∈A

S∗k1k2Sk2k1

eip12k21 =∑

k1k2∈A

S∗k1k2Sk2k1

eip12k21

=∑k1=1

k1−1∑k2=1

(S∗k1k2Sk2k1e

ip12k21 + S∗k2k1Sk1k2e

ip12k12)

=∑k1=1

k1−1∑k2=1

(e−iϕeip12k21 + eiϕeip12k12)

= 2∑k1=1

k1−1∑k2=1

cos(ϕ+ p12k12)

=(`− 1) cosϕ− ` cos(ϕ+ p12) + cos(ϕ+ `p12)

cos p12 − 1. (46)

Thus, the full contribution to the 2nd Renyi entropy (up to normalization of the state) is given by

TrA((ρ(1)A )2) =

[`(`− 1) +

(`− 1) cosϕ− ` cos(ϕ+ p12) + cos(ϕ+ `p12)

cos p12 − 1

]2

(47)

Clearly, for L, ` large and p12 6= 0 and considering (34) the leading contribution to the entanglement entropy is

1

N2TrA((ρ

(1)A )2) ≈ r4. (48)

It is easy to show that the contribution to the 2nd Renyi entropy of the the second term in (39) is identical to theabove, with the replacement `→ L− `. That is,

TrA((ρ(2)A )2) =[

(L− `)(L− `− 1) +(L− `− 1) cosϕ− (L− `) cos(ϕ+ p12) + cos(ϕ+ (L− `)p12)

cos p12 − 1

]2

. (49)

for L, ` large and p12 6= 0 the leading contribution is

1

N2TrA((ρ

(2)A )2) ≈ (1− r)4. (50)

Let us now consider then the third term ρ(3)A . After computing the square and then the trace over subspace A and

noting that

TrA(|j1〉〈k1|j′1〉〈k′1|) = δj1k′1δj′1k1, (51)

we find

TrA((ρ(3)A )2) =

∑j2,j′2∈B

j1,k1∈A

(eip1(j1−k1) + eip2(j1−k1) + e−iϕeip1j12+ip2(j2−k1) + eiϕeip2j12+ip1(j2−k1))

×(eip1(k1−j1) + eip2(k1−j1) + e−iϕeip1(k1−j′2)+ip2(j′2−j1) + eiϕeip2(k1−j′2)+ip1(j′2−j1))) (52)

Expanding the product and simplifying we end up with the sum

TrA((ρ(3)A )2) =

∑j2,j′2∈B

j1,k1∈A

[2 + 2 cos(p12(j1 − k1)) + 2 cos(ϕ+ p12(j′2 − j1))

+2 cos(ϕ+ p12(j′2 − k1)) + 2 cos(ϕ+ p12j21) + 2 cos(ϕ+ p12(j2 − k1))

+2 cos(2ϕ+ p12(j′2 + j2 − j1 − k1)) + 2 cos(p12(j2 − j′2))] (53)

This is equal to

TrA((ρ(3)A )2) = 2`2(L− `)2 + 2(L− `)2 cos(`p12)− 1

cos p12 − 1+ 2`2

cos(L− `)p12 − 1

cos p12 − 1.

+2cos(2ϕ+ Lp12) sin2 `p12

2 sin2 (L−`)p12

2

sin4 p12

2

+8`(L− `)cos(ϕ+ Lp12

2 ) sin `p12

2 sin p12(L−`)2

sin2 p12

2

(54)

Page 11: Entanglement Content of Quasi-Particle Excitations · dynamic Bethe ansatz [39, 40] in fact suggest that the same calculation should also give the correct answer in massive integrable

11

For `, L large and p12 6= 0 the first term in the sum is leading so the leading contribution to the Renyi entropy is

1

N2TrA((ρ

(3)A )2) ≈ 2r2(1− r)2. (55)

Putting the contributions (48), (50) and (55) together we find that the 2nd Renyi entropy of a two-magnon state forlarge volume and region size is, as expected.

S1,12 (r) = −2 log(r2 + (1− r)2) = 2S1

2(r), (56)

that is, as expected, twice the entropy of a single excitation. Crucially, for large volume, the result is independent ofthe scattering matrix.

Equal Momenta

Although in many cases (such as the XXZ chain) the momenta of the magnons is required to be distinct, it is possibleto consider spin chains with bosonic statistics, so that p12 = 0 is allowed. In such cases the formulae of the previoussection can still be used, but the terms involving trigonometric functions, which were negligible at large volume, areno-longer so. Indeed, they now become of the same order as the leading terms and it is a simple calculation to showthat:

N = L(L− 1)(1 + cosϕ), (57)

and

limp12→0

TrA((ρ(1)A )2) = `2(`− 1)2(1 + cosϕ)2, (58)

limp12→0

TrA((ρ(2)A )2) = (L− `)2(L− `− 1)2(1 + cosϕ)2, (59)

limp12→0

TrA((ρ(3)A )2) = 4`2(L− `)2(1 + cosϕ)2. (60)

Putting these results together, we have that the entanglement entropy is independent of the interaction and this isirrespective of whether or not volume is large. For large volume we find that

S22(r) = − log(r4 + 4r2(1− r)2 + (1− r)4), (61)

which is different from the expression for distinct momenta of the previous section. The expression agrees withour findings for the free massive boson QFT and the harmonic chain, showing that also for discrete systems theentanglement entropy encodes information about the fermionic/bosonic nature of the quasi-particles. From (58)-(60)it is possible to show the the next to leading order correction for equal momenta is:

S22(r) = − log(r4 + 4r2(1− r)2 + (1− r)4) +

1

L

[1− 2

r3 + (1− r)3

r4 + 4r2(1− r)2 + (1− r)4

]+O(L−2) (62)

In particular at r = 12 were entanglement is maximal [34] this gives the next-to-leading order correction

S22(1/2) = log

8

3+

1

3L+O(L−2) (63)

It would be interesting to derive such higher order corrections from a QFT computation. A discussion of finize-sizecorrections to the von Neumann entropy of different kinds of excitations in the XXZ spin- 1

2 chain was also providedin several of the appendices of [33].

Large Volume Corrections to the 2nd Renyi Entropy in the Harmonic Chain

In this paper we have not presented numerical results for the one-dimensional harmonic chain, whose scaling limitis described by a real massive free boson. Our focus has been on higher dimensions as we intent to discuss the one-dimensional case in much detail in a future work [50]. However, it is interesting to present here some data regarding

Page 12: Entanglement Content of Quasi-Particle Excitations · dynamic Bethe ansatz [39, 40] in fact suggest that the same calculation should also give the correct answer in massive integrable

12

æ

æ

æ

æ

æ

ææ

ææ

æææ

ææ

ææ

æææææææææææææ

ææææ

æ

-0.50.0 0.5 1.0 1.5 2.0 2.5

1.52.02.53.03.54.04.55.0

log@mLD

-lo

g@lo

g2-

DS 21 H1

�2LD

Harmonic Chain

æ

æ

æ

æ

æ

æ

æ

æ

æ

æ

æ

æ

æ

æ

æ

2.3 2.4 2.5 2.6 2.7 2.8

4.04.24.44.64.8

log@mLD

-lo

g@lo

g2-

DS 21 H1

�2LD

Harmonic Chain

FIG. 3. Left. The figure shows the approach to the value log 2 of the 2nd Renyi entropy of a single excitation in a 1D harmonicchain. For large enough volume mL we see linear behaviour with positive slope, meaning that the leading correction is of theform β/Lα for some β and α. Right. The figure shows a selection of the points on the first figure which are well fitted by thelinear fit −0.613 + 1.974 log(mL) also shown. The fit changes slightly depending on which points are selected. The behaviourstrongly suggests (negative) finite-size corrections of order 1/(mL)2.

the next-to-leading order corrections to the maximal entanglement of an excitation, as they display similar featuresas in two dimensions (see FIG. 2).

An interesting feature of the results above is that the leading corrections to saturation of the entanglement is oforder 1/(mL)2. From form factor calculations in QFT one would expect a leading correction of order 1/(mL) [50] andit would be interesting to investigate whether or not this correction is also vanishing in QFT. The behaviour aboveis very similar to what we have found for the harmonic lattice. In that case our data are a bit more limited as we donot have access to extremely large surfaces but, as seen in FIG. 2 they also point towards a 1/Lα with α > 1 negativecorrection to the maximal entanglement of a one-particle state.