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


institutetext: 1SISSA and INFN Sezione di Trieste, via Bonomea 265, 34136 Trieste, Italy.institutetext: 2International Centre for Theoretical Physics (ICTP), Strada Costiera 11, 34151 Trieste, Italy.institutetext: 3Institute for Theoretical Physics and Astrophysics and Würzburg-Dresden Cluster of Excellence ct.qmat, Julius-Maximilians-Universität Würzburg, Am Hubland, 97074 Würzburg, Germany.

Multi-charged moments of two intervals in conformal field theory

Filiberto Ares1, Pasquale Calabrese1,2, Giuseppe Di Giulio3 and Sara Murciano1 smurcian@sissa.it
Abstract

We study the multi-charged moments for two disjoint intervals in the ground state of two 1+11+1 dimensional CFTs with central charge c=1c=1 and global U(1)U(1) symmetry: the massless Dirac field theory and the compact boson (Luttinger liquid). For this purpose, we compute the partition function on the higher genus Riemann surface arising from the replica method in the presence of background magnetic fluxes between the sheets of the surface. We consider the general situation in which the fluxes generate different twisted boundary conditions at each branch point. The obtained multi-charged moments allow us to derive the symmetry resolution of the Rényi entanglement entropies and the mutual information for non complementary bipartitions. We check our findings against exact numerical results for the tight-binding model, which is a lattice realisation of the massless Dirac theory.

1 Introduction

As Schrödinger already recognised one century ago, entanglement is at the core of quantum mechanics. Nowadays it turns out to be the fundamental notion behind many quantum phenomena, from quantum algorithms nc-10 to gravity Rangamani ; nrt-09 , passing by critical phenomena and topological phases of matter intro1 ; intro2 ; eisert-2010 ; intro3 , triggering unexpected connections between apparently far branches of physics. At the center of all these ideas, we find the (Rényi) entanglement entropies which are powerful entanglement measures that provide fundamental insights about the investigated system or theory. They are defined as follows. Let us consider an extended quantum system in a pure state |Ψ|\Psi\rangle and a spatial bipartition into AA and BB. The subsystem AA is described by the reduced density matrix ρA=TrB|ΨΨ|\rho_{A}=\mathrm{Tr}_{B}|\Psi\rangle\langle\Psi| and the associated Rényi entropies are given by the moments of ρA\rho_{A} as

SnA11nlnTr[ρAn],S_{n}^{A}\equiv\frac{1}{1-n}\ln\mathrm{Tr}[\rho_{A}^{n}], (1)

where we assume that nn is an integer number. After the analytic continuation to complex values of nn, the limit n1n\to 1 of Eq. (1) yields the von Neumann entanglement entropy

S1ATr[ρAlnρA].S_{1}^{A}\equiv-\mathrm{Tr}[\rho_{A}\ln\rho_{A}]. (2)

For bipartite systems in a pure state, the von Neumann and Rényi entropies can be used as measures of the entanglement shared between the two complementary parts. One of the most interesting properties of entanglement entropies is that they are sensitive to criticality. In particular, for one-dimensional gapless systems, if AA is a single interval, then the ground state entanglement entropy breaks the area law and is proportional to the central charge of the 1+1 dimensional CFT that describes the low-energy spectrum of the system cw-94 ; hlw-94 ; cc-04 ; cc-09 .

In the case considered in this work in which AA consists of two subsystems A1A_{1} and A2A_{2}, i.e. A=A1A2A=A_{1}\cup A_{2}, the ground-state entanglement entropy depends on the full operator content of the CFT, encoding all the conformal data of the model Caraglio ; Furukawa ; twist2 ; cct-11 . It is important to remark that, in this situation, the entanglement entropies quantify the entanglement between AA and BB but not between the two parts of AA, for which one must resort to other entanglement measures such as negativity neg-qft ; neg-qft-2 ; ctt-13 ; Alba13 ; Coser3 ; Ares ; Rockwood . Nevertheless, from the entanglement entropies, it is possible to construct the following quantity, dubbed mutual information,

IA1:A2S1A1+S1A2S1A,I^{A_{1}:A_{2}}\equiv S_{1}^{A_{1}}+S_{1}^{A_{2}}-S_{1}^{A}, (3)

which is a measure of the total correlations between A1A_{1} and A2A_{2}. The computation of two-interval Rényi entanglement entropies is a difficult problem, even for minimal CFTs Dupic ; Ares , as it boils down in general to determine the partition function of the theory on a higher genus nn-sheeted Riemann surface twist2 ; cct-11 . In fact, exact analytic expressions are only available for the free theories or special limits Furukawa ; twist2 ; cct-11 ; CFH ; Alba ; Alba2 ; twist3 ; CoserTagliacozzo ; DeNobili ; Coser2 ; rtc-18 ; HLM13 ; gkt-21 ; Casini ; g-21 ; b-19 ; bds-20 . Moreover, the analytic continuation in nn to obtain Eq. (2) is still a challenging open issue.

In recent times, a question that has attracted much attention is how entanglement decomposes into the different symmetry sectors in the presence of global conserved charges lr-14 ; goldstein ; xavier . Various reasons have motivated the interest in this problem. The effect of symmetries on entanglement can be investigated experimentally fis ; Azses ; Neven ; Vitale and, moreover, understanding how entanglement arises from the symmetry sectors is crucial to better grasp some quantum features, for example in non-equilibrium dynamics fis . Also at more practical level, it can help to speed-up the numerical algorithms to simulate quantum many-body systems xavier . All that has been the breeding ground for a plethora of works that analyse the resolution of entanglement from different perspectives: spin chains lr-14 ; Vitale ; Neven ; riccarda ; SREE2dG ; goldstein2 ; MDC-19-CTM ; ccgm-20 ; wv-03 ; bhd-18 ; bcd-19 ; byd-20 ; mrc-20 ; tr-19 ; ms-21 ; amc-22 ; jones-22 ; pvcc-22 , integrable quantum field theories MDC-20 ; hcc-21 ; hcc-a-21 ; hc-20 ; chcc-a22 , CFTs goldstein ; xavier ; goldstein1 ; crc-20 ; MBC-21 ; Chen-21 ; Capizzi-Cal-21 ; Hung-Wong-21 ; cdm-21 ; boncal-21 ; eim-d-21 ; Chen-22 ; Ghasemi-22 ; mt-22 ; ms-21 , holography znm-20 ; wznm-21 ; znwm-22 ; bbcg-22 , out-of-equilibrium pbc-21-1 ; pbc-21 ; fg-21 ; pbc-22 ; sh-22 ; chen-22-2 and disordered systems trac-20 ; kusf-20 ; kusf-20b ; kufs-21-1 or topological matter clss-19 ; ms-20 ; Azses-Sela-20 ; ahn-20 ; ads-21 ; ore-21 to mention some of them. In order to analyse entanglement in each symmetry sector, quantities such as the symmetry-resolved entanglement entropy  lr-14 ; goldstein ; xavier and the symmetry-resolved mutual information pbc-21 have been proposed. As shown in Ref. goldstein , symmetry-resolved entropies are intimately related to the charged moments of the reduced density matrix ρA\rho_{A}, which were independently studied in holographic theories Belin-Myers-13-HolChargedEnt ; cms-13 ; cnn-16 ; d-16 ; d-17 ; ssr-17 ; shapourian-19 . Similarly to the moments of ρA\rho_{A}, they can be interpreted as the partition function of the field theory on a Riemann surface, which is now coupled to an external magnetic flux. Partition functions with a background gauge field have been also introduced as non-local order parameters to detect symmetry-protected topological phases in interacting fermionic systems ssr1-17 ; ssr-16 .

The symmetry resolution of entanglement in the two-interval case has not been much explored in CFT. Ref. wznm-21 studies it at large central charge, in the context of holography, while, in Ref. Chen-22 , the charged Rényi negativity is analysed for the complex free boson. Here we take a different charge for each part of AA, which leads to introduce the multi-charged moments of ρA\rho_{A}. This non-trivial generalisation of the charged moments, first considered in Ref. pbc-21 in the context of quench dynamics, is the main subject of this work. In CFT, they correspond to the partition function on the nn-sheeted Riemann surface, but with the insertion of a different magnetic flux across each subset (interval) of AA. We compute the multi-charged moments analytically for the ground state of two bidimensional CFTs with central charge c=1c=1 and global U(1)U(1) symmetry— the massless Dirac field theory and the free compact boson— generalising the expressions for the (neutral) Rényi entropies found in Refs. CFH and twist2 respectively. From the multi-charged moments, we derive the ground state symmetry-resolved entanglement entropy and mutual information of two disjoint intervals.

The paper is organised as follows: in Sec. 2, we define the symmetry-resolved entanglement and mutual information as well as the multi-charged moments, and we briefly describe the general approach to compute the latter in CFTs. We then move on to calculate the multi-charged moments for the ground state of the massless Dirac field theory in Sec. 3 and of the free compact boson in Sec. 4. In Sec. 5, we apply the previous results to obtain the symmetry-resolution of the mutual information in these theories. When possible, we benchmark the analytic expressions with exact numerical calculations for lattice models in the same universality class. We draw our conclusions in Sec. 6 and we include three appendices, with more details about the analytical and numerical computations.

2 Definitions

In this section, we first give the definition of the symmetry-resolved entanglement entropy and mutual information for a subsystem composed of two disjoint regions. We explain their relation with the multi-charged moments of the reduced density matrix, and we introduce the replica method to calculate them in CFTs.

2.1 Symmetry-resolved entanglement entropies and mutual information

As we already pointed out in Sec. 1, we take a spatial bipartition ABA\cup B of an extended quantum system in a pure state |Ψ|\Psi\rangle, with AA made of two disconnected regions, A=A1A2A=A_{1}\cup A_{2}. We assume that the system is endowed with a global U(1)U(1) symmetry generated by a local charge QQ. Given the partition of the system in different subsets, we can consider the charge operator in each of them; for example, in region AA, it can be obtained as QA=TrB(Q)Q_{A}=\mathop{\mbox{Tr}}\nolimits_{B}(Q). If |Ψ|\Psi\rangle is an eigenstate of QQ, the density matrix ρ=|ΨΨ|\rho=\ket{\Psi}\bra{\Psi} commutes with QQ, i.e. [Q,ρ]=0[Q,\rho]=0, and, by taking the trace over BB, we find that [QA,ρA]=0[Q_{A},\rho_{A}]=0. This implies that the reduced density matrix ρA\rho_{A} presents a block diagonal structure, in which each block corresponds to an eigenvalue qq\in\mathbb{Z} of QAQ_{A}. That is,

ρA=qΠqρA=q[p(q)ρA(q)],\rho_{A}=\bigoplus_{q}\Pi_{q}\rho_{A}=\bigoplus_{q}\left[p(q)\rho_{A}(q)\right], (4)

where Πq\Pi_{q} is the projector onto the eigenspace associated to the eigenvalue qq and p(q)=Tr(ΠqρA)p(q)=\mathop{\mbox{Tr}}\nolimits\left(\Pi_{q}\rho_{A}\right) is the probability of obtaining qq as the outcome of a measurement of QAQ_{A}. Notice that Eq. (4) guarantees the normalisation Tr[ρA(q)]=1\mathrm{Tr}[\rho_{A}(q)]=1 for any qq.

The amount of entanglement between AA and BB in each symmetry sector can be quantified by the symmetry-resolved Rényi entropies, defined as

SnA(q)11nlnTr[ρA(q)n].S_{n}^{A}(q)\equiv\frac{1}{1-n}\ln\mathrm{Tr}[\rho_{A}(q)^{n}]. (5)

Taking the limit n1n\to 1 in this expression, we obtain the symmetry-resolved entanglement entropy,

S1A(q)Tr[ρA(q)lnρA(q)].S_{1}^{A}(q)\equiv-\mathrm{Tr}[\rho_{A}(q)\ln\rho_{A}(q)]. (6)

According to the decomposition of Eq. (4), the total entanglement entropy in Eq. (2) can be written as nc-10

S1A=qp(q)S1A(q)qp(q)lnp(q)ScA+SnumA,S_{1}^{A}=\sum_{q}p(q)S_{1}^{A}(q)-\sum_{q}p(q)\ln p(q)\equiv S_{\textrm{c}}^{A}+S_{\textrm{num}}^{A}, (7)

where ScS_{\textrm{c}} is known as configurational entropy and quantifies the average contribution to the total entanglement of all the charge sectors fis ; wv-03 ; bhd-18 ; bcd-19 , while SnumS_{\textrm{num}} is called number entropy and takes into account the entanglement due to the fluctuations of the value of the charge within the subsystem AA fis ; kusf-20 ; kusf-20b ; ms-20 ; kufs-21 ; zshgs-20 ; kufs-21b .

Since the total charge in AA is the sum of the charge in A1A_{1} and A2A_{2}, QA=QA1+QA2Q_{A}=Q_{A_{1}}+Q_{A_{2}}, then the reduced density matrices ρA1\rho_{A_{1}}, ρA2\rho_{A_{2}} of A1A_{1} and A2A_{2} can be independently decomposed in charged sectors as we did for ρA\rho_{A} in Eq. (4). Therefore, we can define the symmetry-resolved entropies SnA1(q1)S_{n}^{A_{1}}(q_{1}), SnA2(q2)S_{n}^{A_{2}}(q_{2}) for the regions A1A_{1} and A2A_{2} analogous to Eq. (5) for AA, with q=q1+q2q=q_{1}+q_{2}. In Ref. pbc-21 , it has been proposed to define the symmetry-resolved mutual information as

IA1:A2(q)=q1=0qp(q1,qq1)[S1A1(q1)+S1A2(qq1)]S1A(q).I^{A_{1}:A_{2}}(q)=\sum_{q_{1}=0}^{q}p(q_{1},q-q_{1})\left[S_{1}^{A_{1}}(q_{1})+S_{1}^{A_{2}}(q-q_{1})\right]-S_{1}^{A}(q). (8)

The quantity p(q1,qq1)p(q_{1},q-q_{1}), normalised as

q1=0qp(q1,qq1)=1,\sum_{q_{1}=0}^{q}p(q_{1},q-q_{1})=1, (9)

is the probability that a simultaneous measurement of the charges QA1Q_{A_{1}} and QA2Q_{A_{2}} yields q1q_{1} and qq1q-q_{1}, respectively, while the charge of the whole system AA is fixed to qq. Although Eq. (8) is a natural definition, IA1:A2(q)I^{A_{1}:A_{2}}(q) is not in general a good measure of the total correlations between A1A_{1} and A2A_{2} within each charge sector since, in some cases, it can be negative pbc-21 . Nevertheless, we find interesting to investigate this quantity given that it provides a decomposition for the total mutual information (8) similar to the one reported in Eq. (7) for the entanglement entropy,

IA1:A2=qp(q)IA1:A2(q)+InumA1:A2,I^{A_{1}:A_{2}}=\sum_{q}p(q)I^{A_{1}:A_{2}}(q)+I^{A_{1}:A_{2}}_{\textrm{num}}, (10)

where InumA1:A2SnumA1+SnumA2SnumAI^{A_{1}:A_{2}}_{\textrm{num}}\equiv S^{A_{1}}_{\textrm{num}}+S^{A_{2}}_{\textrm{num}}-S^{A}_{\textrm{num}} is the number mutual information.

2.2 Charged moments and symmetry resolution

The computation of the symmetry-resolved entanglement entropies and mutual information from the definitions (5) and (8) requires the knowledge of the entanglement spectra of ρA\rho_{A}, ρA1\rho_{A_{1}} and ρA2\rho_{A_{2}} and their symmetry resolution. However, this is usually a very difficult task, in particular if one is interested in analytical expressions. Alternatively, one can employ the charged moments of the reduced density matrices. For ρA\rho_{A}, they are defined as

ZnA(α)=Tr[ρAneiαQA].Z_{n}^{A}(\alpha)=\mathop{\mbox{Tr}}\nolimits[\rho_{A}^{n}e^{i\alpha Q_{A}}]. (11)

Similar quantities can also be introduced for the two subsystems A1A_{1} and A2A_{2} that constitute AA by replacing ρA\rho_{A} and QAQ_{A} by ρAp\rho_{A_{p}} and QApQ_{A_{p}}, p=1,2p=1,2. If we take now their Fourier transform,

𝒵nA(q)=ππdα2πeiαqZnA(α),\mathcal{Z}_{n}^{A}(q)=\int_{-\pi}^{\pi}\frac{{\rm d}\alpha}{2\pi}e^{-i\alpha q}Z_{n}^{A}(\alpha), (12)

the symmetry-resolved entanglement entropies of the subsystem AA are given by goldstein ; xavier

SnA(q)=11nln[𝒵nA(q)(𝒵1A(q))n].S_{n}^{A}(q)=\frac{1}{1-n}\ln\left[\frac{\mathcal{Z}_{n}^{A}(q)}{\left(\mathcal{Z}_{1}^{A}(q)\right)^{n}}\right]. (13)

In a similar manner, replacing AA with A1A_{1} and A2A_{2}, we can obtain the symmetry-resolved entanglement entropies of the two components of AA.

Notice that, for computing the symmetry-resolved mutual information of Eq. (8), we need to determine p(q1,qq1)p(q_{1},q-q_{1}), i.e. the probability that a measurement of QA1Q_{A_{1}} and QA2Q_{A_{2}} gives q1q_{1} and qq1q-q_{1} respectively, with QAQ_{A} fixed to qq. In order to calculate it, we consider the generalisation of the charged moments in Eq. (11) introduced for the first time in Ref. pbc-21 ,

ZnA1:A2(α,β)=Tr[ρAneiαQA1+iβQA2].Z_{n}^{A_{1}:A_{2}}(\alpha,\beta)=\mathrm{Tr}\left[\rho_{A}^{n}e^{i\alpha Q_{A_{1}}+i\beta Q_{A_{2}}}\right]. (14)

We refer to them as multi-charged moments. When α=β\alpha=\beta, Eq. (14) reduces to the charged moments of A=A1A2A=A_{1}\cup A_{2} of Eq. (11). If we take the Fourier transform of Eq. (14),

𝒵nA1:A2(q1,q2)=ππdα2πdβ2πeiαq1iβq2ZnA1:A2(α,β),\mathcal{Z}_{n}^{A_{1}:A_{2}}(q_{1},q_{2})=\int_{-\pi}^{\pi}\frac{{\rm d}\alpha}{2\pi}\frac{{\rm d}\beta}{2\pi}e^{-i\alpha q_{1}-i\beta q_{2}}Z_{n}^{A_{1}:A_{2}}(\alpha,\beta), (15)

then 𝒵1A1:A2(q1,q2)\mathcal{Z}_{1}^{A_{1}:A_{2}}(q_{1},q_{2}) can be interpreted as the probability of having q1q_{1} and q2q_{2} as outcomes of a measurement of QA1Q_{A_{1}} and QA2Q_{A_{2}} respectively, independently of the value of QAQ_{A}. Therefore, it satisfies the normalisation

q1,q2𝒵1A1:A2(q1,q2)=1,\sum_{q_{1},q_{2}}\mathcal{Z}_{1}^{A_{1}:A_{2}}(q_{1},q_{2})=1, (16)

and p(q1,qq1)p(q_{1},q-q_{1}) can be calculated as the conditional probability

p(q1,qq1)=𝒵1A1:A2(q1,qq1)p(q),p(q_{1},q-q_{1})=\frac{\mathcal{Z}_{1}^{A_{1}:A_{2}}(q_{1},q-q_{1})}{p(q)}, (17)

which fulfills Eq. (9).

2.3 Charged moments in CFT

In the rest of the paper, we will analyse the previous quantities in 1+11+1-dimensional CFTs with a global U(1)U(1) symmetry. We will assume that the entire system is in the ground state and that the spatial dimension is an infinite line which we will divide into two parts AA and BB, with AA made up of two disjoint intervals, namely A=A1A2=[u1,v1][u2,v2]A=A_{1}\cup A_{2}=[u_{1},v_{1}]\cup[u_{2},v_{2}]. If we denote by 1\ell_{1} and 2\ell_{2} the lengths of the two intervals and dd their separation, we have

1=|v1u1|,2=|v2u2|,d=|u2v1|,x=12(d+1)(d+2),\ell_{1}=|v_{1}-u_{1}|,\qquad\ell_{2}=|v_{2}-u_{2}|,\qquad d=|u_{2}-v_{1}|,\qquad x=\frac{\ell_{1}\ell_{2}}{(d+\ell_{1})(d+\ell_{2})}, (18)

where we have also introduced the cross ratio xx of the four end-points, which takes values between 0 and 11.

As explained in detail in Refs. cc-04 ; cc-09 , using the path integral representation of ρA\rho_{A}, the moments ZnA(0)=Tr[ρAn]Z_{n}^{A}(0)=\mathop{\mbox{Tr}}\nolimits[\rho_{A}^{n}] are equal to the partition function of the CFT on a Riemann surface, which we call Σn\Sigma_{n}, obtained as follows. We take the complex plane where the CFT is originally defined and we perform two cuts along the intervals A1=(u1,v1)A_{1}=(u_{1},v_{1}) and A2=(u2,v2)A_{2}=(u_{2},v_{2}). Then we replicate nn times the cut plane and we glue the copies together along the cuts in a cyclical way as we illustrate in Fig. 1. We eventually obtain an nn-sheeted Riemann surface of genus n1n-1, which is symmetric under the n\mathbb{Z}_{n} cyclic permutation of the sheets.

Alternatively, instead of replicating the space-time where the CFT is initially defined, one can take nn copies of the CFT on the complex plane and quotient it by the n\mathbb{Z}_{n} symmetry under the cyclic exchange of the copies. We then get the orbifold theory CFTn/n{\rm CFT}^{\otimes n}/\mathbb{Z}_{n}. The moments ZnA(0)Z_{n}^{A}(0) are equal to the four-point correlation function on the complex plane cc-04 ; ccd-08 ,

ZnA(0)=τn(u1)τ~n(v1)τn(u2)τ~n(v2),Z_{n}^{A}(0)=\langle\tau_{n}(u_{1})\tilde{\tau}_{n}(v_{1})\tau_{n}(u_{2})\tilde{\tau}_{n}(v_{2})\rangle, (19)

where τn\tau_{n} and τ~n\tilde{\tau}_{n} are dubbed as twist and anti-twist fields k-87 ; dixon ; cc-04 ; ccd-08 . They implement in the orbifold the multivaluedness of the correlation functions on the surface Σn\Sigma_{n} when we go around its branch points. In fact, the winding around the point where τn\tau_{n} (τ~n\tilde{\tau}_{n}) is inserted maps a field 𝒪k\mathcal{O}_{k} living in the copy kk of the orbifold into the copy k+1k+1 (k1k-1), that is

τn(u)𝒪k(e2πi(zu))=𝒪k+1(zu)τn(u).\tau_{n}(u)\mathcal{O}_{k}(e^{2\pi i}(z-u))=\mathcal{O}_{k+1}(z-u)\tau_{n}(u). (20)

The twist and anti-twist fields are spinless primaries with conformal weight

hnτ=c24(n1n),h_{n}^{\tau}=\frac{c}{24}\left(n-\frac{1}{n}\right), (21)

where cc is the central charge of the initial CFT.

The charged moments of ρA\rho_{A} can also be computed employing the previous frameworks. As argued in Ref. goldstein , the operator eiαQAe^{i\alpha Q_{A}} can be interpreted as a magnetic flux between the sheets of the surface Σn\Sigma_{n}, such that a charged particle moving along a closed path that crosses all the sheets acquires a phase eiαe^{i\alpha}. For the multi-charged moments introduced in Eq. (14), we have to insert two different magnetic fluxes α\alpha and β\beta at the cuts A1A_{1} and A2A_{2} respectively, as we pictorially show in Fig. 1. They can be implemented by a local U(1)U(1) operator 𝒱α(x)\mathcal{V}_{\alpha}(x) that generates a phase shift eiαe^{i\alpha} along the real interval [x,)[x,\infty). Then the charged moments are equal to the four-point correlation function on the surface Σn\Sigma_{n}

ZnA1:A2(α,β)=ZnA(0)𝒱α(u1)𝒱α(v1)𝒱β(u2)𝒱β(v2)Σn.Z_{n}^{A_{1}:A_{2}}(\alpha,\beta)=Z_{n}^{A}(0)\langle\mathcal{V}_{\alpha}(u_{1})\mathcal{V}_{-\alpha}(v_{1})\mathcal{V}_{\beta}(u_{2})\mathcal{V}_{-\beta}(v_{2})\rangle_{\Sigma_{n}}. (22)
Refer to caption
Figure 1: Representation of the nn-sheeted Riemann surface Σn\Sigma_{n} for n=3n=3. The red edge of each cut is identified with the blue edge of the corresponding cut in the lower copy. The calculation of the multi-charged moments in Eq. (14) requires to insert different magnetic fluxes between the sheets, which we indicate by the arrows. The operator eiαQA1e^{i\alpha Q_{A_{1}}} is implemented by the flux insertions α\alpha and α-\alpha along the left interval, while eiβQA2e^{i\beta Q_{A_{2}}} corresponds to the fluxes β\beta and β-\beta at the right interval.

In the orbifold theory, the magnetic flux can be incorporated by considering the composite twist field τn,ατn𝒱α\tau_{n,\alpha}\equiv\tau_{n}\cdot\mathcal{V}_{\alpha}. Thus, if we take a field 𝒪k\mathcal{O}_{k} in the copy kk of the orbifold, then the winding (zu)e2πi(zu)(z-u)\mapsto e^{2\pi i}(z-u) around the point uu where τn,α\tau_{n,\alpha} is inserted gives rise to a phase eiα/ne^{i\alpha/n},

τn,α(u)𝒪k(e2πi(zu))=eiα/n𝒪k+1(zu)τn,α(u).\tau_{n,\alpha}(u)\mathcal{O}_{k}(e^{2\pi i}(z-u))=e^{i\alpha/n}\mathcal{O}_{k+1}(z-u)\tau_{n,\alpha}(u). (23)

The same applies to the composite anti-twist field τ~n,ατ~n𝒱α\tilde{\tau}_{n,\alpha}\equiv\tilde{\tau}_{n}\cdot\mathcal{V}_{\alpha}, which takes a field from the copy kk to k1k-1 adding a phase eiα/ne^{i\alpha/n}. Therefore, Eq. (22) can be re-expressed as the four-point function on the complex plane

ZnA1:A2(α,β)=τn,α(u1)τ~n,α(v1)τn,β(u2)τ~n,β(v2).Z_{n}^{A_{1}:A_{2}}(\alpha,\beta)=\langle\tau_{n,\alpha}(u_{1})\tilde{\tau}_{n,-\alpha}(v_{1})\tau_{n,\beta}(u_{2})\tilde{\tau}_{n,-\beta}(v_{2})\rangle. (24)

In Ref. goldstein , it is shown that, if 𝒱α\mathcal{V}_{\alpha} is a spinless primary operator with conformal weight hα𝒱h_{\alpha}^{\mathcal{V}}, then so are the composite twist and anti-twist fields, with conformal weights

hn,α=hnτ+hα𝒱n.h_{n,\alpha}=h_{n}^{\tau}+\frac{h_{\alpha}^{\mathcal{V}}}{n}. (25)

One can further consider other configurations for the magnetic fluxes between the sheets of the Riemann surface Σn\Sigma_{n}. In general, if we assume that a particle gets a different phase eiαje^{i\alpha_{j}} when it goes around each branch point, provided they satisfy the neutrality condition α1+α2+α3+α4=0\alpha_{1}+\alpha_{2}+\alpha_{3}+\alpha_{4}=0, then the partition function of this theory is given by

ZnA1:A2({αj})=ZnA(0)𝒱α1(u1)𝒱α2(v1)𝒱α3(u2)𝒱α4(v2)Σn,Z_{n}^{A_{1}:A_{2}}(\{\alpha_{j}\})=Z_{n}^{A}(0)\langle\mathcal{V}_{\alpha_{1}}(u_{1})\mathcal{V}_{\alpha_{2}}(v_{1})\mathcal{V}_{\alpha_{3}}(u_{2})\mathcal{V}_{\alpha_{4}}(v_{2})\rangle_{\Sigma_{n}}, (26)

or, in terms of the composite twist fields, by

ZnA1:A2({αj})=τn,α1(u1)τ~n,α2(v1)τn,α3(u2)τ~n,α4(v2).Z_{n}^{A_{1}:A_{2}}(\{\alpha_{j}\})=\langle\tau_{n,\alpha_{1}}(u_{1})\tilde{\tau}_{n,\alpha_{2}}(v_{1})\tau_{n,\alpha_{3}}(u_{2})\tilde{\tau}_{n,\alpha_{4}}(v_{2})\rangle. (27)

Then the multi-charged moments ZnA1:A2(α,β)Z_{n}^{A_{1}:A_{2}}(\alpha,\beta) can be treated as the particular case in which α1=α2=α\alpha_{1}=-\alpha_{2}=\alpha and, due to the neutrality condition, α3=α4=β\alpha_{3}=-\alpha_{4}=\beta.

In Sec. 3, we will compute the charged moments ZnA1:A2(α,β)Z_{n}^{A_{1}:A_{2}}(\alpha,\beta)—and more in general the partition functions ZnA1:A2({αj})Z_{n}^{A_{1}:A_{2}}(\{\alpha_{j}\})—for the massless Dirac fermion using the orbifold theory CFTn/n{\rm CFT}^{\otimes n}/\mathbb{Z}_{n}. On the other hand, in Sec. 4, we will adopt a geometric approach to obtain the multi-charged moments of the compact boson from the correlation function on the Riemann surface Σn\Sigma_{n} of Eq. (26).

3 Free massless Dirac field theory

The massless Dirac field theory is described by the action

𝒮D=dx0dx1ψ¯γμμψ,\mathcal{S}_{\rm D}=\int{\rm d}x_{0}{\rm d}x_{1}\bar{\psi}\gamma^{\mu}\partial_{\mu}\psi, (28)

where ψ¯=ψγ0\bar{\psi}=\psi^{\dagger}\gamma^{0}. The γμ\gamma^{\mu} matrices can be represented in terms of the Pauli matrices as γ0=σ1\gamma^{0}=\sigma_{1} and γ1=σ2\gamma^{1}=\sigma_{2}. The action of Eq. (28) exhibits a global U(1)U(1) symmetry: it is invariant if the fields are multiplied by a phase, i.e. ψeiαψ\psi\mapsto e^{i\alpha}\psi and ψ¯eiαψ¯\bar{\psi}\mapsto e^{-i\alpha}\bar{\psi}. By Noether’s theorem, this symmetry is related to the conservation of the charge QD=dx1ψψQ_{\rm D}=\int{\rm d}x_{1}\psi^{\dagger}\psi.

The ground state entanglement of a subsystem AA made up of multiple disjoint intervals in the ground state of this theory was first investigated in Ref. CFH . For the case of two disjoint intervals, A=A1A2A=A_{1}\cup A_{2}, it was found that the moments of ρA\rho_{A} are

ZnA(0)=cn[12(1x)]1n26n,Z_{n}^{A}(0)=c_{n}\left[\ell_{1}\ell_{2}(1-x)\right]^{\frac{1-n^{2}}{6n}}, (29)

where cnc_{n} is a non-universal constant.

In this section, we will compute the multi-charged moments of Eq. (14) in the ground state of the massless Dirac field theory. We will extend the approach introduced in Ref. CFH for the moments of Eq. (29). Similar techniques have been exploited in Ref. MDC-20 for studying the charged moments of Eq. (11), when AA is a single interval, in two dimensional free massless Dirac theories and in Ref. MBC-21 in the context of the charge imbalance resolved negativity. We will benchmark our analytical results with exact numerical calculations in a lattice model.

3.1 Charged moments

In Sec. 2, we explained that the partition function ZnA1:A2({αj})Z_{n}^{A_{1}:A_{2}}(\{\alpha_{j}\}) can be obtained either by considering the theory on a complicated Riemann surface or by replicating it nn-times and working with the orbifold on the complex plane. For the massless Dirac field theory, the latter approach is more convenient. Thus, let us take the nn-component field

Ψ=(ψ1ψ2ψn),\Psi=\begin{pmatrix}\psi_{1}\\ \psi_{2}\\ \vdots\\ \psi_{n}\end{pmatrix}, (30)

where ψj\psi_{j} is the Dirac field on the jj-th copy of the system. Eq. (23) describes the effect of the composite twist fields on the components of Ψ\Psi when going around the end-points of the subsystem A=A1A2A=A_{1}\cup A_{2}. This transformation can be encoded in the matrix

Ta=(0eia/n0eia/n(1)n1eia/n0).T_{a}=\begin{pmatrix}0&e^{ia/n}&&\\ &0&e^{ia/n}&\\ &&\ddots&\ddots\\ (-1)^{n-1}e^{ia/n}&&0\end{pmatrix}. (31)

In the general case of Eq. (27), Ψ\Psi transforms according to Tα2p1T_{\alpha_{2p-1}} when winding around the point upu_{p} and to the transpose matrix Tα2ptT_{\alpha_{2p}}^{t} when going around the point vpv_{p}, with p=1,2p=1,2. The matrix TaT_{a} in Eq. (31), sometimes called twist matrix, was introduced for the case a=0a=0 in Refs. CFH ; ccd-08 and for general aa in Ref. MDC-20 . Its eigenvalues are of the form

tk=eia/ne2πik/n,k=n12,,n12.t_{k}=e^{ia/n}e^{2\pi ik/n},\qquad k=-\frac{n-1}{2},\dots,\frac{n-1}{2}. (32)

By simultaneously diagonalising all the TαjT_{\alpha_{j}} with a unitary transformation (which is independent of αj\alpha_{j}), we can recast the replicated theory in nn decoupled fields ψk\psi_{k} on the plane, which are multi-valued,

ψk(ei2π(zup))\displaystyle\psi_{k}(e^{i2\pi}(z-u_{p})) =\displaystyle= eiα2p1/ne2πik/nψk(zup),\displaystyle e^{i\alpha_{2p-1}/n}e^{2\pi ik/n}\psi_{k}(z-u_{p}),
ψk(ei2π(zvp))\displaystyle\psi_{k}(e^{i2\pi}(z-v_{p})) =\displaystyle= eiα2p/ne2πik/nψk(zvp).\displaystyle e^{i\alpha_{2p}/n}e^{-2\pi ik/n}\psi_{k}(z-v_{p}). (33)

Notice that this technique, known as diagonalisation in the replica space, can be applied only to free theories since, otherwise, the kk-modes do not decouple. For the free massless Dirac theory, this allows us to write the Lagrangian of the replicated theory as

D,n=kk,k=ψ¯kγμμψk.\mathcal{L}_{{\rm D},n}=\sum_{k}\mathcal{L}_{k},\qquad\mathcal{L}_{k}=\bar{\psi}_{k}\gamma^{\mu}\partial_{\mu}\psi_{k}. (34)

Following this approach, the partition function of Eq. (27) factorises into

ZnA1:A2({αj})=k=n12k=n12Zk,nA1:A2({αj}),Z_{n}^{A_{1}:A_{2}}(\{\alpha_{j}\})=\prod_{k=-\frac{n-1}{2}}^{k=\frac{n-1}{2}}Z_{k,n}^{A_{1}:A_{2}}(\{\alpha_{j}\}), (35)

where Zk,nA1:A2({αj})Z_{k,n}^{A_{1}:A_{2}}(\{\alpha_{j}\}) is the partition function for a Dirac field ψk\psi_{k} with the boundary conditions of Eq. (3.1).

The main difference between the partition functions Zk,nA1:A2({αj})Z_{k,n}^{A_{1}:A_{2}}(\{\alpha_{j}\}) and the standard computation of Ref. CFH for Rényi entropies is that the boundary conditions of the multi-valued fields around the branch points now depend on the phases αj\alpha_{j}. This multivaluedness can be removed, as done in CFH for αj=0\alpha_{j}=0, by introducing an external U(1)U(1) gauge field AμkA_{\mu}^{k} coupled to single-valued fields ψ~k\tilde{\psi}_{k}. In fact, if we apply the singular gauge transformation

ψk(x)=eix0xdyμAμkψ~k(x),\psi_{k}(x)=e^{i\int_{x_{0}}^{x}\mathrm{d}y^{\mu}A_{\mu}^{k}}\tilde{\psi}_{k}(x), (36)

then the Lagrangian for the kk-th mode can be rewritten as

k=ψ~¯kγμ(μ+iAμk)ψ~k,\mathcal{L}_{k}=\bar{\tilde{\psi}}_{k}\gamma^{\mu}\left(\partial_{\mu}+iA_{\mu}^{k}\right)\tilde{\psi}_{k}, (37)

with the advantage of absorbing the phase around the end-points of A1A2A_{1}\cup A_{2} into the gauge field. The only requirement that AμkA_{\mu}^{k} in Eq. (36) must satisfy is that, integrated along any closed curve 𝒞\mathcal{C} that encircles the end-points of AA, the boundary conditions of Eq. (3.1) for ψk\psi_{k} must be reproduced. For this purpose, we require

𝒞updyμAμk=2πknα2p1n,𝒞vpdyμAμk=2πknα2pn,\oint_{\mathcal{C}_{u_{p}}}{\rm d}y^{\mu}A_{\mu}^{k}=-\frac{2\pi k}{n}-\frac{\alpha_{2p-1}}{n},\qquad\oint_{\mathcal{C}_{v_{p}}}{\rm d}y^{\mu}A_{\mu}^{k}=\frac{2\pi k}{n}-\frac{\alpha_{2p}}{n}, (38)

where 𝒞up\mathcal{C}_{u_{p}} and 𝒞vp\mathcal{C}_{v_{p}} are closed contours around the end-points of the pp-th interval. Moreover, we have to impose that, if 𝒞\mathcal{C} does not enclose any end-point, then 𝒞dyμAμk=0\oint_{\mathcal{C}}{\rm d}y^{\mu}A_{\mu}^{k}=0. Applying the Stoke’s theorem, the conditions of Eqs. (38) can be expressed in differential form,

12πϵμννAμk(x)=p=12[(α2p12πn+kn)δ(xup)+(α2p2πnkn)δ(xvp)].\frac{1}{2\pi}\epsilon^{\mu\nu}\partial_{\nu}A_{\mu}^{k}(x)=\sum_{p=1}^{2}\left[\left(\frac{\alpha_{2p-1}}{2\pi n}+\frac{k}{n}\right)\delta(x-u_{p})+\left(\frac{\alpha_{2p}}{2\pi n}-\frac{k}{n}\right)\delta(x-v_{p})\right]. (39)

Once the transformation of Eq. (36) is performed, the partition function Zk,nA1:A2({αj})Z_{k,n}^{A_{1}:A_{2}}(\{\alpha_{j}\}) of the kk-th mode is equal to the vacuum expectation value

Zk,nA1:A2({αj})=eid2xjkμAμk,Z_{k,n}^{A_{1}:A_{2}}(\{\alpha_{j}\})=\left\langle e^{i\int d^{2}xj^{\mu}_{k}A_{\mu}^{k}}\right\rangle, (40)

where jkμψ~¯kγμψ~kj^{\mu}_{k}\equiv\bar{\tilde{\psi}}_{k}\gamma^{\mu}\tilde{\psi}_{k} is the conserved Dirac current for each mode. Eq. (40) can be easily computed via bosonisation CFH , which allows us to write the current in terms of the dual scalar field ϕk\phi_{k} such that jkμ=ϵμννϕk/πj^{\mu}_{k}=\epsilon^{\mu\nu}\partial_{\nu}\phi_{k}/\sqrt{\pi}. If we use this result in Eq. (40), and we apply Eq. (39), then Zk,nA1:A2({αj})Z_{k,n}^{A_{1}:A_{2}}(\{\alpha_{j}\}) is equal to the following correlation function of vertex operators Va(y)=eiaϕk(y)V_{a}(y)=e^{-ia\phi_{k}(y)}

Zk,nA1:A2({αj})=Vkn+α12πn(u1)Vkn+α22πn(v1)Vkn+α32πn(u2)Vkn+α42πn(v2).Z_{k,n}^{A_{1}:A_{2}}(\{\alpha_{j}\})=\langle V_{\frac{k}{n}+\frac{\alpha_{1}}{2\pi n}}(u_{1})V_{-\frac{k}{n}+\frac{\alpha_{2}}{2\pi n}}(v_{1})V_{\frac{k}{n}+\frac{\alpha_{3}}{2\pi n}}(u_{2})V_{-\frac{k}{n}+\frac{\alpha_{4}}{2\pi n}}(v_{2})\rangle. (41)

Notice that the neutrality condition α1+α2+α3+α4=0\alpha_{1}+\alpha_{2}+\alpha_{3}+\alpha_{4}=0 ensures that the latter correlator does not vanish. The correlation function of vertex operators in the complex plane is well-known (see, for instance, Ref. difrancesco ) and, therefore, Eq. (41) can be easily calculated. Plugging the result into Eq. (35) and performing the product over kk, we obtain

ZnA1:A2({αj})[12(1x)]1n26n[dα2α31α1α22α3α4(d+1)α1α3(d+2)α2α4(d+1+2)α1α4]12π2n,Z_{n}^{A_{1}:A_{2}}(\{\alpha_{j}\})\propto\\ \left[\ell_{1}\ell_{2}(1-x)\right]^{\frac{1-n^{2}}{6n}}\left[d^{\alpha_{2}\alpha_{3}}\ell_{1}^{\alpha_{1}\alpha_{2}}\ell_{2}^{\alpha_{3}\alpha_{4}}(d+\ell_{1})^{\alpha_{1}\alpha_{3}}(d+\ell_{2})^{\alpha_{2}\alpha_{4}}(d+\ell_{1}+\ell_{2})^{\alpha_{1}\alpha_{4}}\right]^{\frac{1}{2\pi^{2}n}}, (42)

where xx is the cross-ratio defined in Eq. (18). When we take α1=α2=α\alpha_{1}=-\alpha_{2}=\alpha and α3=α4=β\alpha_{3}=-\alpha_{4}=\beta in this expression, we get the multi-charged moments in Eq. (14) as

ZnA1:A2(α,β)\displaystyle Z_{n}^{A_{1}:A_{2}}(\alpha,\beta) =\displaystyle= cn;α,β[12(1x)]1n26n[(1x)αβ1α22β2]12π2n.\displaystyle c_{n;\alpha,\beta}\left[\ell_{1}\ell_{2}(1-x)\right]^{\frac{1-n^{2}}{6n}}\left[\left(1-x\right)^{-\alpha\beta}\ell_{1}^{-\alpha^{2}}\ell_{2}^{-\beta^{2}}\right]^{\frac{1}{2\pi^{2}n}}. (43)

We assume that all the length scales in this formula have been regularised through a UV cutoff which is included in the multiplicative constant cn;α,βc_{n;\alpha,\beta}.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 2: Analysis of Z1A1:A2(α,β)Z_{1}^{A_{1}:A_{2}}(\alpha,\beta) for the tight-binding model. In the top and middle panels, we show Z1A1:A2(α,β)Z_{1}^{A_{1}:A_{2}}(\alpha,\beta) as a function of α\alpha at fixed β\beta for different intervals of lengths 1,2\ell_{1},\ell_{2}, separated by a distance d>0d>0. In this case, the solid lines are the theoretical predictions of Eq. (43), taking Eq. (47) for cn;α,βc_{n;\alpha,\beta}. In the bottom left panel, we repeat the analysis but considering adjacent intervals (d=0d=0) and the solid curves correspond to Eq. (45), with c~n;α,β\tilde{c}_{n;\alpha,\beta} conjectured in Eq. (51). In the bottom right panel, we plot Z1A1:A2(α,β)Z_{1}^{A_{1}:A_{2}}(\alpha,\beta) as a function of the cross-ratio xx. Here the solid lines correspond to Eq. (43) using the exact expression for cn;α,βc_{n;\alpha,\beta} of Eq. (47), while for the dashed curves we have considered instead the quadratic approximation for this constant of Eq. (49). In all the cases, the points are the exact numerical values for Z1A1:A2(α,β)Z_{1}^{A_{1}:A_{2}}(\alpha,\beta) calculated as described in Appendix A.

An interesting case to analyse is when the two intervals A1A_{1} and A2A_{2} become adjacent; that is, when d0d\to 0. In that limit, the cross-ratio xx tends to one such that

1x=d1+212+O(d2),1-x=d\frac{\ell_{1}+\ell_{2}}{\ell_{1}\ell_{2}}+O(d^{2}), (44)

and Eq. (43) vanishes. Nevertheless, in this regime, the distance dd must be regarded as another UV cutoff, which can be absorbed in the multiplicative constant cn;α,βc_{n;\alpha,\beta}. Therefore, from Eqs. (44) and (43), one expects

ZnA1:A2(α,β)=c~n;α,β(1+2)1n26n[1αβα22αββ2(1+2)αβ]12π2n,Z_{n}^{A_{1}:A_{2}}(\alpha,\beta)=\tilde{c}_{n;\alpha,\beta}(\ell_{1}+\ell_{2})^{\frac{1-n^{2}}{6n}}\left[\frac{\ell_{1}^{\alpha\beta-\alpha^{2}}\ell_{2}^{\alpha\beta-\beta^{2}}}{(\ell_{1}+\ell_{2})^{\alpha\beta}}\right]^{\frac{1}{2\pi^{2}n}}, (45)

which agrees with the fact that, according to Eq. (24), the multi-charged moments ZnA1:A2(α,β)Z_{n}^{A_{1}:A_{2}}(\alpha,\beta) must tend to the three-point function of primaries τn,α(u1)𝒱α+β(v1)τ~n,β(v2)\langle\tau_{n,\alpha}(u_{1})\mathcal{V}_{-\alpha+\beta}(v_{1})\tilde{\tau}_{n,-\beta}(v_{2})\rangle, in the limit d0d\to 0.

In Fig. 2, we check the expressions obtained in Eqs. (43) and  (45) for ZnA1:A2(α,β)Z_{n}^{A_{1}:A_{2}}(\alpha,\beta) with exact numerical calculations performed in the tight-binding model, which is a chain of non-relativistic free fermions whose scaling limit is described by the massless Dirac field theory. The details of the numerical techniques employed are given in Appendix A. In order to compare Eqs. (43) and (45) with the numerical data, we need the concrete expression of the non-universal factors cn;α,βc_{n;\alpha,\beta} and c~n;α,β\tilde{c}_{n;\alpha,\beta} for this particular model. When the two intervals A1A_{1} and A2A_{2} are far away, that is in the limit dd\to\infty, the multi-charged moments of A1A2A_{1}\cup A_{2} factorise into those of A1A_{1} and A2A_{2},

limdZnA1:A2(α,β)=ZnA1(α)ZnA2(β).\lim_{d\to\infty}Z_{n}^{A_{1}:A_{2}}(\alpha,\beta)=Z_{n}^{A_{1}}(\alpha)Z_{n}^{A_{2}}(\beta). (46)

Therefore, one expects cn;α,βc_{n;\alpha,\beta} to be the product of the two non-universal constants associated to A1A_{1} and A2A_{2} as single intervals. The latter were obtained for the tight-binding model in Ref. riccarda by exploiting the asymptotic properties of Toeplitz determinants. We can also apply here those results, taking into account that each interval is associated to a different flux, either α\alpha or β\beta. Then we have

cn;α,β=e[13(n1n)α22π2nβ22π2n]log2+Υ(n,α)+Υ(n,β),c_{n;\alpha,\beta}=e^{\left[-\frac{1}{3}\left(n-\frac{1}{n}\right)-\frac{\alpha^{2}}{2\pi^{2}n}-\frac{\beta^{2}}{2\pi^{2}n}\right]\log 2+\Upsilon(n,\alpha)+\Upsilon(n,\beta)}, (47)

where riccarda

Υ(n,α)=nidw[tanh(πw)tanh(πnw+iα/2)]lnΓ(12+iw)Γ(12iw).\Upsilon{(n,\alpha)}={ni}\int_{-\infty}^{\infty}{\rm d}w[\tanh(\pi w)-\tanh(\pi nw+i\alpha/2)]\ln\frac{\Gamma(\frac{1}{2}+iw)}{\Gamma(\frac{1}{2}-iw)}. (48)

Expanding Υ(n,α)\Upsilon(n,\alpha) up to quadratic order in α\alpha, then Eq. (47) can be approximated as

cn;α,βe13(n1n)ln2+2Υ(n,0)ζn2π2n(α2+β2),c_{n;\alpha,\beta}\approx e^{-\frac{1}{3}\left(n-\frac{1}{n}\right)\ln 2+2\Upsilon(n,0)-\frac{\zeta_{n}}{2\pi^{2}n}(\alpha^{2}+\beta^{2})}, (49)

where ζn=ln22π2nγ2(n)\zeta_{n}=\ln 2-2\pi^{2}n\gamma_{2}(n) and

γ2(n)=122Υ(n,α)α2|α=0=ni40dw[tanh3(πnw)tanh(πnw)]logΓ(12+iw)Γ(12iw).\gamma_{2}(n)=\frac{1}{2}\left.\frac{\partial^{2}\Upsilon(n,\alpha)}{\partial\alpha^{2}}\right|_{\alpha=0}=\frac{ni}{4}\int_{0}^{\infty}{\rm d}w[\tanh^{3}(\pi nw)-\tanh(\pi nw)]\log\frac{\Gamma(\frac{1}{2}+iw)}{\Gamma(\frac{1}{2}-iw)}. (50)

In the limit of adjacent intervals, given by Eq. (45), the multiplicative constant c~n;α,β\tilde{c}_{n;\alpha,\beta} cannot be determined using the known results for Toeplitz determinants. However, we can conjecture an analytical approximation for it at quadratic order in α\alpha and β\beta. When d0d\to 0, we can associate to each end-point of the intervals u1u_{1}, v1=u2v_{1}=u_{2} and v2v_{2} the fluxes α\alpha, βα\beta-\alpha and β-\beta respectively. From the results for one interval of Ref. riccarda , one can conjecture that each end-point with flux α\alpha contributes with a factor eζn4π2nα2e^{-\frac{\zeta_{n}}{4\pi^{2}n}\alpha^{2}} to the constant c~n;α,β\tilde{c}_{n;\alpha,\beta}, if we restrict to terms up to order α2\alpha^{2}. Therefore, the combination of all the fluxes in our case should contribute with a total factor eζn4π2n(α2+(βα)2+β2)e^{-\frac{\zeta_{n}}{4\pi^{2}n}(\alpha^{2}+(\beta-\alpha)^{2}+\beta^{2})}. We then expect that c~n;α,β\tilde{c}_{n;\alpha,\beta} should be well be approximated by

c~n;α,β=e16(n1n)ln2+Υ(n,0)ζn2π2n(α2+β2αβ).\tilde{c}_{n;\alpha,\beta}=e^{-\frac{1}{6}\left(n-\frac{1}{n}\right)\ln 2+\Upsilon(n,0)-\frac{\zeta_{n}}{2\pi^{2}n}(\alpha^{2}+\beta^{2}-\alpha\beta)}. (51)

When α=β=0\alpha=\beta=0, the expression above simplifies to the multiplicative constant for the moments of a single interval obtained in Ref. JK04 . In spite of the heuristic reasoning of this result, in Fig. 2, we check its validity by comparing it against exact numerical data.

In the top and middle panels of Fig. 2, we study Z1A1:A2(α,β)Z_{1}^{A_{1}:A_{2}}(\alpha,\beta) as function of α\alpha, for various values of β\beta, 1\ell_{1}, 2\ell_{2} and d>0d>0. The points correspond to the exact numerical values obtained as we described in Appendix A, while the solid curves are the analytic prediction of Eq. (43), taking as multiplicative constant cn,α,βc_{n,\alpha,\beta} that in Eq. (47). We find an excellent agreement. In the bottom left panel, we analyse the case of adjacent intervals (d=0d=0). The numerical data for Z1A1:A2(α,β)Z_{1}^{A_{1}:A_{2}}(\alpha,\beta) are in very good agreement with the analytical expression (solid curves) of Eq. (45), using Eq. (51) as multiplicative constant. Finally, in the bottom right panel, we plot Z1A1:A2(α,β)Z_{1}^{A_{1}:A_{2}}(\alpha,\beta) as function of the cross-ratio xx, for various values of α\alpha, β\beta, 1\ell_{1} and 2\ell_{2}. The curves represent the prediction of Eq. (43). The continuous ones correspond to take as non-universal constant cn;α,βc_{n;\alpha,\beta} the full expression of Eq. (47), while for the dashed ones we have used the quadratic approximation of Eq. (49). The agreement between the analytic prediction and the numerical data is extremely good, even considering the quadratic approximation for the non-universal constant. As expected, this agreement is better for small values of α\alpha and β\beta, while, around ±π\pm\pi, we need to take larger subsystem sizes 1\ell_{1}, 2\ell_{2} in order to suppress the finite-size corrections which are well-known and characterised for the charged moments with a single flux insertion riccarda .

4 Free compact boson

The second theory we focus on in this manuscript is the free compact boson, which is the CFT of the Luttinger liquid, and whose action reads

𝒮b=18πdx0dx1μφμφ.\mathcal{S}_{\rm b}=\frac{1}{8\pi}\int{\rm d}x_{0}{\rm d}x_{1}\partial_{\mu}\varphi\partial^{\mu}\varphi. (52)

The target space of the real field φ\varphi is compactified on a circle of radius RR, i.e. φφ+2πmR\varphi\sim\varphi+2\pi mR with mm\in\mathbb{Z}. The compactification radius RR is related to the Luttinger parameter KK as R=2/KR=\sqrt{2/K}. The action of Eq. (52) is invariant under the transformation φφ+α\varphi\mapsto\varphi+\alpha which, due to the compact nature of φ\varphi, realises a U(1)U(1) global symmetry. The associated conserved charge is Qb=12πdx1x1φQ_{\rm b}=\frac{1}{2\pi}\int{\rm d}x_{1}\partial_{x_{1}}\varphi.

The moments of the ground state reduced density matrix of this theory are well-known. An exact analytic expression for the two-interval case was obtained in Ref. twist2 , which was generalised to an arbitrary number of disjoint intervals in Ref. CoserTagliacozzo . In particular, for two intervals, it reads twist2

ZnA(0)=cn[12(1x)]1n26nn(x),Z_{n}^{A}(0)=c_{n}\left[\ell_{1}\ell_{2}(1-x)\right]^{\frac{1-n^{2}}{6n}}\mathcal{F}_{n}(x), (53)

where cnc_{n} is a non-universal constant and

n(x)=Θ(𝟎|Γ(x)/K)Θ(𝟎|Γ(x)K)[Θ(𝟎|Γ(x))]2.\mathcal{F}_{n}(x)=\frac{\Theta(\boldsymbol{0}|\Gamma(x)/K)\Theta(\boldsymbol{0}|\Gamma(x)K)}{\left[\Theta(\boldsymbol{0}|\Gamma(x))\right]^{2}}. (54)

We denote by Θ\Theta the Riemann-Siegel Theta function

Θ[𝜺𝜹](𝒖|Ω)𝒎n1eiπ(𝒎+𝜺)tΩ(𝒎+𝜺)+2πi(𝒎+𝜺)t(𝒖+𝜹),\Theta\!\left[\begin{smallmatrix}\boldsymbol{\varepsilon}\\ \boldsymbol{\delta}\end{smallmatrix}\right]\!(\boldsymbol{u}|\Omega)\equiv\sum_{\boldsymbol{m}\in\mathbb{Z}^{n-1}}e^{i\pi(\boldsymbol{m}+\boldsymbol{\varepsilon})^{t}\cdot\Omega(\boldsymbol{m}+\boldsymbol{\varepsilon})+2\pi i(\boldsymbol{m}+\boldsymbol{\varepsilon})^{t}\cdot(\boldsymbol{u}+\boldsymbol{\delta})}, (55)

with characteristics 𝜺,𝜹(/2)n1\boldsymbol{\varepsilon},\boldsymbol{\delta}\in(\mathbb{Z}/2)^{n-1}, 𝒖n1\boldsymbol{u}\in\mathbb{C}^{n-1} and Ω\Omega a complex (n1)×(n1)(n-1)\times(n-1) matrix. In Eq. (54), the characteristics are zero 𝟎=(0,,0)\boldsymbol{0}=(0,\dots,0) and, therefore, we have used the standard shorthand notation Θ(𝒖|Ω)Θ[𝟎𝟎](𝒖|Ω)\Theta(\boldsymbol{u}|\Omega)\equiv\Theta\!\left[\begin{smallmatrix}\boldsymbol{0}\\ \boldsymbol{0}\end{smallmatrix}\right]\!(\boldsymbol{u}|\Omega). The matrix Γ(x)\Gamma(x) in Eq. (54) has entries given by

Γrs(x)=2inl=1n1cos[2πl(rs)n]sin(πln)βl/n(x),r,s=1,,n1,\Gamma_{rs}(x)=\frac{2i}{n}\sum_{l=1}^{n-1}\cos\left[\frac{2\pi l(r-s)}{n}\right]\sin\left(\frac{\pi l}{n}\right)\beta_{l/n}(x),\quad r,s=1,\dots,n-1, (56)

and

βp(x)=Ip(1x)Ip(x),\beta_{p}(x)=\frac{I_{p}(1-x)}{I_{p}(x)}, (57)

with Ip(x)2F1(p,1p,1,1x)I_{p}(x)\equiv\,_{2}F_{1}(p,1-p,1,1-x). The function n(x)\mathcal{F}_{n}(x) is invariant under x1xx\mapsto 1-x and it is normalised such that n(0)=n(1)=1\mathcal{F}_{n}(0)=\mathcal{F}_{n}(1)=1. Although the moments of ρA\rho_{A} are known for all the integer nn, its analytic continuation to complex nn and, consequently, the von Neumann entropy of Eq. (2) is still not available for all the values of the Luttinger parameter.

A remarkable contact point between the theories described by the actions of Eqs. (28) and (52) is the case K=1K=1. Notice that, when the Luttinger parameter takes this value, the function n(x)\mathcal{F}_{n}(x) in Eq. (54) simplifies to n(x)\mathcal{F}_{n}(x)=1 and the moments of Eq. (53) for the massless compact boson present the same universal dependence on 1\ell_{1}, 2\ell_{2} and xx as the ones in Eq. (29) for the massless Dirac fermion. A detailed discussion on this identity can be found in Ref. HLM13 , where it is explained the reason why, although these two theories are not related by a duality, their partition functions on the Riemann surfaces Σn\Sigma_{n} arising in the two-interval replica method are actually equal. Here we find that this identity extends to the partition functions ZnA1:A2({αj})Z_{n}^{A_{1}:A_{2}}(\{\alpha_{j}\}) on the surface Σn\Sigma_{n} with different twisted boundary conditions at each branch point. In general, when AA is made up of more than two intervals and the Rényi index nn is larger than two, the moments of the reduced density matrix in these CFTs (and the corresponding Rényi entropies) are different HLM13 .

4.1 Charged moments

We now generalise the result (53) to the multi-charged moments in Eq. (14). Starting from Eq. (22), we will compute them as the four-point function of the field 𝒱α\mathcal{V}_{\alpha} on the Riemann surface Σn\Sigma_{n}. Since the U(1)U(1) conserved current is proportional to x1φ\partial_{x_{1}}\varphi, 𝒱α\mathcal{V}_{\alpha} is identified in this case with the vertex operator goldstein

𝒱α(z)=eiα2πφ(z),\mathcal{V}_{\alpha}(z)=e^{i\frac{\alpha}{2\pi}\varphi(z)}, (58)

which has conformal dimensions

hα𝒱=h¯α𝒱=(α2π)2K2.h_{\alpha}^{\mathcal{V}}=\bar{h}_{\alpha}^{\mathcal{V}}=\left(\frac{\alpha}{2\pi}\right)^{2}\frac{K}{2}. (59)

In the following, it will be useful to introduce the rescaled Luttinger paratemeter η=K/(2π2)\eta=K/(2\pi^{2}) in order to lighten the expressions.

Without loss of generality, let us consider that the end-points of subsystem AA are u1=0u_{1}=0, v1=xv_{1}=x, u2=1u_{2}=1 and v2=v_{2}=\infty. Using Eq. (27), and given that the composite twist fields are primaries, we can eventually obtain the expression for an arbitrary set of end-points through a global conformal transformation. Therefore, according to Eq. (22), the multi-charged moments can be derived from the four-point correlation function of the vertex operators of Eq (58)

ZnA1:A2({αj})=ZnA(0)𝒱α1(0)𝒱α2(x)𝒱α3(1)𝒱α4()Σn(x)Z_{n}^{A_{1}:A_{2}}(\{\alpha_{j}\})=Z_{n}^{A}(0)\langle\mathcal{V}_{\alpha_{1}}(0)\mathcal{V}_{\alpha_{2}}(x)\mathcal{V}_{\alpha_{3}}(1)\mathcal{V}_{\alpha_{4}}(\infty)\rangle_{\Sigma_{n}(x)} (60)

on the nn-sheeted Riemann surface Σn(x)\Sigma_{n}(x) with branch points at 0, xx, 11 and \infty. This surface of genus n1n-1 can be described by the complex curve

yn=z(z1)zx.y^{n}=\frac{z(z-1)}{z-x}. (61)

The correlator of vertex operators on a general Riemann surface of arbitrary genus was obtained in Ref. Verlinde . In order to give the explicit expression in our case, we need to introduce some notions about Riemann surfaces Fay .

There are different parameterisations of the moduli space of genus n1n-1 Riemann surfaces. One possibility is through the matrix of periods, which we denote by Γ\Gamma. This is a (n1)×(n1)(n-1)\times(n-1) symmetric matrix with positive definite imaginary part. Notice that, according to Eq. (61), the Riemann surface Σn(x)\Sigma_{n}(x) is parametrised by the cross-ratio xx. Therefore, the corresponding matrix of periods only depends on xx, i.e. Γ=Γ(x)\Gamma=\Gamma(x). In order to define it, we need first to specify a particular homology basis for Σn(x)\Sigma_{n}(x), i.e. a basis of 2(n1)2(n-1) oriented non-contractible curves on the surface, which we denote by ara_{r} and brb_{r}, with r=1,,n1r=1,\dots,n-1. The detailed description of the specific basis that we consider is given in Appendix B. We also have to choose a basis of holomorphic differentials νr\nu_{r}, r=1,,n1r=1,\dots,n-1, normalised with respect to the ara_{r} cycles. That is,

ardzνs(z)=δr,s,r,s=1,,n1.\oint_{a_{r}}{\rm d}z\nu_{s}(z)=\delta_{r,s},\quad r,s=1,\dots,n-1. (62)

Then the matrix of periods is defined as

Γrs=brdzνs(z).\Gamma_{rs}=\oint_{b_{r}}{\rm d}z\nu_{s}(z). (63)

For the surface Σn(x)\Sigma_{n}(x), the normalised holomorphic differentials read

νr(z)=1πnl=1n1ei2π(r1)lnsin(πl/n)Il/n(x)(z(z1))l/n(zx)1+l/n.\nu_{r}(z)=\frac{1}{\pi n}\sum_{l=1}^{n-1}\frac{e^{-\frac{i2\pi(r-1)l}{n}}\sin(\pi l/n)}{I_{l/n}(x)}(z(z-1))^{-l/n}(z-x)^{-1+l/n}. (64)

In Appendix B, we thoroughly explain the derivation of the expression for νr\nu_{r}. Inserting it in Eq. (63), it is then easy to show that the entries of the matrix of periods Γ(x)\Gamma(x) are precisely those of Eq. (56).

If we now consider four vertex operators inserted at generic points in the surface Σn(x)\Sigma_{n}(x) and with arbitrary dimensions satisfying the neutrality condition α1+α2+α3+α4=0\alpha_{1}+\alpha_{2}+\alpha_{3}+\alpha_{4}=0, then its correlation function is of the form Verlinde

𝒱α1(z1)𝒱α2(z2)𝒱α3(z3)𝒱α4(z4)Σn(x)=1j<j4|E(zj,zj)eπIm[𝒘(zj)𝒘(zj)]tIm[Γ(x)1]Im[𝒘(zj)𝒘(zj)]|αjαjη.\langle\mathcal{V}_{\alpha_{1}}(z_{1})\mathcal{V}_{\alpha_{2}}(z_{2})\mathcal{V}_{\alpha_{3}}(z_{3})\mathcal{V}_{\alpha_{4}}(z_{4})\rangle_{\Sigma_{n}(x)}=\\ \prod_{1\leq j<j^{\prime}\leq 4}\left|E(z_{j},z_{j^{\prime}})e^{-\pi{\rm Im}[\boldsymbol{w}(z_{j})-\boldsymbol{w}(z_{j^{\prime}})]^{t}\cdot{\rm Im}[\Gamma(x)^{-1}]{\rm Im}[\boldsymbol{w}(z_{j})-\boldsymbol{w}(z_{j^{\prime}})]}\right|^{\alpha_{j}\alpha_{j^{\prime}}\eta}. (65)

In this expression, we denote by E(z,z)E(z,z^{\prime}) the prime form of the surface Σn(x)\Sigma_{n}(x), which we will define precisely later, and 𝒘(z)=(w1(z),,wn1(z))\boldsymbol{w}(z)=(w_{1}(z),\dots,w_{n-1}(z)) is the Abel-Jacobi map, which relates a point zz in the surface Σn(x)\Sigma_{n}(x) to a point 𝒘(z)\boldsymbol{w}(z) in the genus n1n-1 complex torus n1/Λ\mathbb{C}^{n-1}/\Lambda, where Λ=n1+Γn1\Lambda=\mathbb{Z}^{n-1}+\Gamma\mathbb{Z}^{n-1}. This map can be written in terms of the normalised holomorphic differentials of Eq. (64) as

wr(z)=0zdzνr(z)(modΛ),w_{r}(z)=\int_{0}^{z}{\rm d}z^{\prime}\nu_{r}(z^{\prime})\quad(\rm{mod}\,\Lambda), (66)

where we have taken as origin the branch point z=0z=0. The images under the Abel-Jacobi map of the points z=0z=0, xx, 11, and \infty, where the vertex operators in Eq. (60) are inserted, can be easily computed using Eq. (64) and applying the identities of Eq. (127). Then we find

𝒘(0)\displaystyle\boldsymbol{w}(0) =\displaystyle= 𝟎,\displaystyle\boldsymbol{0}, (67)
𝒘(x)\displaystyle\boldsymbol{w}(x) =\displaystyle= 𝒒,\displaystyle\boldsymbol{q}, (68)
𝒘(1)\displaystyle\boldsymbol{w}(1) =\displaystyle= 𝒒+i𝒑(x),\displaystyle\boldsymbol{q}+i\boldsymbol{p}(x), (69)
𝒘()\displaystyle\boldsymbol{w}(\infty) =\displaystyle= i𝒑(x),\displaystyle i\boldsymbol{p}(x), (70)

where 𝒒=(1/n,,1/n)\boldsymbol{q}=(1/n,\dots,1/n) and 𝒑(x)=(p1(x),,pn1(x))\boldsymbol{p}(x)=(p_{1}(x),\dots,p_{n-1}(x)) with

pr(x)=1nl=1n1[cos[2πl(r1)n]sin(πln)+sin[2πl(r1)n]cos(πln)]βl/n(x).p_{r}(x)=-\frac{1}{n}\sum_{l=1}^{n-1}\left[\cos\left[\frac{2\pi l(r-1)}{n}\right]\sin\left(\frac{\pi l}{n}\right)+\sin\left[\frac{2\pi l(r-1)}{n}\right]\cos\left(\frac{\pi l}{n}\right)\right]\beta_{l/n}(x). (71)

Therefore, for the case z1=0z_{1}=0, z2=xz_{2}=x, z3=1z_{3}=1, z4=z_{4}=\infty, Eq. (65) simplifies to

𝒱α1(0)𝒱α2(x)𝒱α3(1)𝒱α4()Σn(x)=Mn(x)(α1+α2)(α3+α4)η1j<j4|E(zj,zj)|αjαjη,\langle\mathcal{V}_{\alpha_{1}}(0)\mathcal{V}_{\alpha_{2}}(x)\mathcal{V}_{\alpha_{3}}(1)\mathcal{V}_{\alpha_{4}}(\infty)\rangle_{\Sigma_{n}(x)}=M_{n}(x)^{(\alpha_{1}+\alpha_{2})(\alpha_{3}+\alpha_{4})\eta}\prod_{1\leq j<j^{\prime}\leq 4}\left|E(z_{j},z_{j^{\prime}})\right|^{\alpha_{j}\alpha_{j^{\prime}}\eta}, (72)

where

Mn(x)=eπ𝒑(x)t[ImΓ(x)]1𝒑(x).M_{n}(x)=e^{-\pi\boldsymbol{p}(x)^{t}\cdot[{\rm Im}\Gamma(x)]^{-1}\boldsymbol{p}(x)}. (73)

Let us now focus on the prime form E(z,z)E(z,z^{\prime}). It can be defined as Fay

E(z,z)=Θ𝟏𝟐(𝒘(z)𝒘(z)|Γ(x))g(z)g(z),E(z,z^{\prime})=\frac{\Theta_{\boldsymbol{\frac{1}{2}}}(\boldsymbol{w}(z)-\boldsymbol{w}(z^{\prime})|\Gamma(x))}{\sqrt{g(z)}\sqrt{g(z^{\prime})}}, (74)

where Θ𝟏𝟐\Theta_{\boldsymbol{\frac{1}{2}}} is a shorthand notation for the Theta function of Eq. (55) with both characteristics equal to (1/2,0,,0)(/2)n1(1/2,0,\dots,0)\in(\mathbb{Z}/2)^{n-1} and g(z)g(z) is

g(z)=r=1n1νr(z)urΘ𝟏𝟐(𝒖|Γ(x))|𝒖=𝟎.g(z)=\sum_{r=1}^{n-1}\nu_{r}(z)\partial_{u_{r}}\left.\Theta_{\boldsymbol{\frac{1}{2}}}(\boldsymbol{u}|\Gamma(x))\right|_{\boldsymbol{u}=\boldsymbol{0}}. (75)

Notice that the holomorphic differentials νr(z)\nu_{r}(z) in Eq. (64) and, therefore, g(z)g(z) are singular at the branch points of the curve that defines the surface Σn(x)\Sigma_{n}(x). This means that the correlation function (65) is in principle not well-defined. In order to solve this issue, the vertex operators inserted at the branch points have to be regularised by redefining them as a proper limit from a non-singular point. We will first extract and remove from g(z)g(z) the divergent terms at the branch points. Then, by considering the limit in which the distance between A1A_{1} and A2A_{2} tends to infinity, we will fix the correct definition of the regularised vertex operators at the branch points.

Close to the branch points z=0z=0, xx, 11, and \infty, the holomorphic normalised differentials νr(z)\nu_{r}(z) of Eq. (64) behave as

νr(z+ϵ)=ϵ1nn[νr()(z)+O(ϵ1/n)],\nu_{r}(z+\epsilon)=\epsilon^{\frac{1-n}{n}}\left[\nu_{r}^{(*)}(z)+O(\epsilon^{1/n})\right], (76)

with |ϵ|1|\epsilon|\ll 1 ,

νr()(z)={x1/nQr,n(x),z=0,eiπ(4r3)n(x(1x))1/nQr,n(x),z=x,(1x)1/nQr,n(x),z=1,e4πi(r1)nQr,n(x),z=,\nu_{r}^{(*)}(z)=\left\{\begin{array}[]{ll}x^{-1/n}Q_{r,n}(x),&z=0,\\ e^{-\frac{i\pi(4r-3)}{n}}(x(1-x))^{-1/n}Q_{r,n}(x),&z=x,\\ (1-x)^{-1/n}Q_{r,n}(x),&z=1,\\ e^{-\frac{4\pi i(r-1)}{n}}Q_{r,n}(x),&z=\infty,\end{array}\right. (77)

and

Qr,n(x)=e2πi(r1)nsin(π/n)πnI1/n(x).Q_{r,n}(x)=e^{\frac{2\pi i(r-1)}{n}}\frac{\sin(\pi/n)}{\pi nI_{1/n}(x)}. (78)

Observe that, in the four singularities, the divergent term when ϵ0\epsilon\to 0 is a global factor ϵ1nn\epsilon^{\frac{1-n}{n}} and, once we take it out, the subleading corrections in ϵ\epsilon vanish. Therefore, these singularities can be removed in the correlation function of Eq (65) by defining the vertex operators at the branch points as the limit

𝒱α()(z)=limϵ0(κnϵn1n)2hα𝒱𝒱α(z+ϵ),z=0,x,1,.\mathcal{V}_{\alpha}^{(*)}(z)=\lim_{\epsilon\to 0}\left(\kappa_{n}\epsilon^{\frac{n-1}{n}}\right)^{2h_{\alpha}^{\mathcal{V}}}\mathcal{V}_{\alpha}(z+\epsilon),\quad z=0,x,1,\infty. (79)

In this definition, we have included a possible global rescaling factor κn\kappa_{n}, which may depend on the genus of the surface, and we will adjust by studying the limit of large separation between the two intervals. If we replace in Eq. (72) the vertex operators by the regularised ones introduced in Eq. (79), then the resulting correlation function can be written in the form

𝒱α1()(0)𝒱α2()(x)𝒱α3()(1)𝒱α4()()Σn(x)==κn2hTMn(x)(α1+α2)(α3+α4)η1j<j4|E()(zj,zj)|αjαjη,\langle\mathcal{V}_{\alpha_{1}}^{(*)}(0)\mathcal{V}_{\alpha_{2}}^{(*)}(x)\mathcal{V}_{\alpha_{3}}^{(*)}(1)\mathcal{V}_{\alpha_{4}}^{(*)}(\infty)\rangle_{\Sigma_{n}(x)}=\\ =\kappa_{n}^{2h_{T}}M_{n}(x)^{(\alpha_{1}+\alpha_{2})(\alpha_{3}+\alpha_{4})\eta}\prod_{1\leq j<j^{\prime}\leq 4}|E^{(*)}(z_{j},z_{j^{\prime}})|^{\alpha_{j}\alpha_{j^{\prime}}\eta}, (80)

where hT=hα1+hα2+hα3+hα4h_{T}=h_{\alpha_{1}}+h_{\alpha_{2}}+h_{\alpha_{3}}+h_{\alpha_{4}} and E()(zj,zj)E^{(*)}(z_{j},z_{j}^{\prime}) stands for the regularised prime form

E()(zj,zj)=Θ𝟏𝟐(𝒘(zj)𝒘(zj)|Γ(x))g()(zj)g()(zj),E^{(*)}(z_{j},z_{j^{\prime}})=\frac{\Theta_{\boldsymbol{\frac{1}{2}}}(\boldsymbol{w}(z_{j})-\boldsymbol{w}(z_{j^{\prime}})|\Gamma(x))}{\sqrt{g^{(*)}(z_{j})}\sqrt{g^{(*)}(z_{j^{\prime}})}}, (81)

with

g()(zj)=r=1n1νr()(zj)urΘ𝟏𝟐(𝒖|Γ(x))|𝒖=𝟎,g^{(*)}(z_{j})=\sum_{r=1}^{n-1}\nu_{r}^{(*)}(z_{j})\partial_{u_{r}}\left.\Theta_{\boldsymbol{\frac{1}{2}}}(\boldsymbol{u}|\Gamma(x))\right|_{\boldsymbol{u}=\boldsymbol{0}}, (82)

and the expressions of ν()(z)\nu^{(*)}(z) at the branch points are those given in Eq. (77).

In Appendix C, we conjecture and numerically check the following identities for the regularised prime forms that appear in Eq. (80),

|E()(0,x)|\displaystyle|E^{(*)}(0,x)| =\displaystyle= nx1/n,\displaystyle nx^{1/n}, (83)
|E()(x,1)|\displaystyle|E^{(*)}(x,1)| =\displaystyle= n(1x)1/nMn(x),\displaystyle\frac{n(1-x)^{1/n}}{M_{n}(x)}, (84)
|E()(1,)|\displaystyle|E^{(*)}(1,\infty)| =\displaystyle= n,\displaystyle n, (85)

and

|E()(0,1)|=|E()(0,)|=|E()(x,)|=nMn(x).|E^{(*)}(0,1)|=|E^{(*)}(0,\infty)|=|E^{(*)}(x,\infty)|=\frac{n}{M_{n}(x)}. (86)

Plugging them into Eq. (80), we finally find

𝒱α1()(0)𝒱α2()(x)𝒱α3()(1)𝒱α4()()Σn(x)=(κnn)2hTxα1α2ηn(1x)α2α3ηn.\langle\mathcal{V}_{\alpha_{1}}^{(*)}(0)\mathcal{V}_{\alpha_{2}}^{(*)}(x)\mathcal{V}_{\alpha_{3}}^{(*)}(1)\mathcal{V}_{\alpha_{4}}^{(*)}(\infty)\rangle_{\Sigma_{n}(x)}=\left(\frac{\kappa_{n}}{n}\right)^{2h_{T}}x^{\frac{\alpha_{1}\alpha_{2}\eta}{n}}(1-x)^{\frac{\alpha_{2}\alpha_{3}\eta}{n}}. (87)

In Eq. (60), once the vertex operators 𝒱α\mathcal{V}_{\alpha} are replaced by the regularised ones 𝒱α()\mathcal{V}_{\alpha}^{(*)}, we can exploit Eq. (87) to get

ZnA1:A2({αj})(κnn)2hT(x(1x))1n26n(xα1α2(1x)α2α3)ηnn(x).Z_{n}^{A_{1}:A_{2}}(\{\alpha_{j}\})\propto\left(\frac{\kappa_{n}}{n}\right)^{2h_{T}}\left(x(1-x)\right)^{\frac{1-n^{2}}{6n}}\left(x^{\alpha_{1}\alpha_{2}}(1-x)^{\alpha_{2}\alpha_{3}}\right)^{\frac{\eta}{n}}\mathcal{F}_{n}(x). (88)

In particular, when α1=α2=α\alpha_{1}=-\alpha_{2}=\alpha and α3=α4=β\alpha_{3}=-\alpha_{4}=\beta, we get the multi-charged moments ZnA1:A2(α,β)Z_{n}^{A_{1}:A_{2}}(\alpha,\beta). After a global conformal transformation to a subsystem AA with arbitrary end-points (u1,v1,u2,v2)(u_{1},v_{1},u_{2},v_{2}), we obtain the following result

ZnA1:A2(α,β)=cn;α,β(κnn)2hT(12(1x))1n26n(1α22β2(1x)αβ)ηnn(x).Z_{n}^{A_{1}:A_{2}}(\alpha,\beta)=c_{n;\alpha,\beta}\left(\frac{\kappa_{n}}{n}\right)^{2h_{T}}\left(\ell_{1}\ell_{2}(1-x)\right)^{\frac{1-n^{2}}{6n}}\left(\ell_{1}^{\alpha^{2}}\ell_{2}^{\beta^{2}}(1-x)^{\alpha\beta}\right)^{-\frac{\eta}{n}}\mathcal{F}_{n}(x). (89)

Note that the rescaling factor κn\kappa_{n}, which was introduced in the definition of the regularised vertex operators at the branch points, is still undetermined. We can fix it by analysing the limit in which the two intervals A1A_{1} and A2A_{2} are far, i.e. dd\to\infty, as done for the Dirac theory. In that case, the charged moments ZnA1:A2(α,β)Z_{n}^{A_{1}:A_{2}}(\alpha,\beta) must verify Eq. (46). Since n(0)=1\mathcal{F}_{n}(0)=1 and the constant cn;α,βc_{n;\alpha,\beta} factorises into those for the intervals A1A_{1} and A2A_{2}, then Eq. (89) satisfies the limit dd\to\infty of Eq. (46) if κn=n\kappa_{n}=n.

In conclusion, for the massless compact boson, the partition function on the surface Σn\Sigma_{n} with general twisted boundary conditions is of the form

ZnA1:A2({αj})ZnA(0)[dα2α31α1α22α3α4(d+1)α1α3(d+2)α2α4(d+1+2)α1α4]K2π2n,\frac{Z_{n}^{A_{1}:A_{2}}(\{\alpha_{j}\})}{Z_{n}^{A}(0)}\propto\left[d^{\alpha_{2}\alpha_{3}}\ell_{1}^{\alpha_{1}\alpha_{2}}\ell_{2}^{\alpha_{3}\alpha_{4}}(d+\ell_{1})^{\alpha_{1}\alpha_{3}}(d+\ell_{2})^{\alpha_{2}\alpha_{4}}(d+\ell_{1}+\ell_{2})^{\alpha_{1}\alpha_{4}}\right]^{\frac{K}{2\pi^{2}n}}, (90)

and the multi-charged moments are

ZnA1:A2(α,β)=cn;α,β(12(1x))1n26n(1α22β2(1x)αβ)K2π2nn(x).Z_{n}^{A_{1}:A_{2}}(\alpha,\beta)=c_{n;\alpha,\beta}\left(\ell_{1}\ell_{2}(1-x)\right)^{\frac{1-n^{2}}{6n}}\left(\ell_{1}^{\alpha^{2}}\ell_{2}^{\beta^{2}}(1-x)^{\alpha\beta}\right)^{-\frac{K}{2\pi^{2}n}}\mathcal{F}_{n}(x). (91)

When the Luttinger parameter is K=1K=1, then n(x)=1\mathcal{F}_{n}(x)=1, and the partition function of Eq. (91) is equal to the one obtained in Eq. (42) for the massless Dirac field, as we anticipated at the beginning of this section. Interestingly, the factor in Eq. (91) due to the magnetic fluxes is the same, when α=β\alpha=\beta, as the one derived in Ref. wznm-21 for a large central charge CFT with the Luttinger parameter KK replaced by the level of the Kac-Moody algebra of that theory.

5 Symmetry resolution

In this section, we apply the approach described in Sec. 2.1 in order to evaluate the symmetry resolution of the mutual information in the two CFTs analysed in Secs. 3 and 4 from the expressions obtained there for their multi-charged moments ZnA1:A2(α,β)Z_{n}^{A_{1}:A_{2}}(\alpha,\beta).

5.1 Fourier transforms

The first step is to determine the Fourier transform (15) of the multi-charged moments. We need to know how the non-universal constant cn;α,βc_{n;\alpha,\beta} does depend on α\alpha and β\beta. In Sec. 3, we have concluded that, for the tight-binding model, it can be well approximated if we only take into account the quadratic terms in α\alpha and β\beta. In the following, we will assume that this is in general a good approximation  xavier . Therefore, we will take

cn;α,β=cn;0,0λn(α2+β2)K2π2n.c_{n;\alpha,\beta}=c_{n;0,0}\lambda_{n}^{-\frac{(\alpha^{2}+\beta^{2})K}{2\pi^{2}n}}. (92)

In the case of the tight-binding model (K=1K=1), we obtained in Eq. (49) that λn=eζn\lambda_{n}=e^{\zeta_{n}}.

Therefore, applying Eq. (92) in the result of Eq. (91), the multi-charged moments can be rewritten as

ZnA1:A2(α,β)=ZnA(0)[(1x)αβ~1α2~2β2]K2π2n,Z_{n}^{A_{1}:A_{2}}(\alpha,\beta)=Z_{n}^{A}(0)\left[\left(1-x\right)^{-\alpha\beta}\tilde{\ell}_{1}^{-\alpha^{2}}\tilde{\ell}_{2}^{-\beta^{2}}\right]^{\frac{K}{2\pi^{2}n}}, (93)

where ~p=λnp\tilde{\ell}_{p}=\lambda_{n}\ell_{p}. The evaluation of Eq. (15) using the expression above yields the following multivariate Gaussian function for the Fourier modes of the multi-charged moments

𝒵nA1:A2(q1,q2)=ZnA(0)nπe2π2nq12ln~2+q22ln~1+q1q2ln(1x)K[4ln(~1)ln(~2)ln2(1x)]K4ln(~1)ln(~2)ln2(1x).\mathcal{Z}_{n}^{A_{1}:A_{2}}(q_{1},q_{2})=\frac{Z_{n}^{A}(0)n\pi e^{-2\pi^{2}n\frac{q_{1}^{2}\ln\tilde{\ell}_{2}+q_{2}^{2}\ln\tilde{\ell}_{1}+q_{1}q_{2}\ln(1-x)}{K[4\ln(\tilde{\ell}_{1})\ln(\tilde{\ell}_{2})-\ln^{2}(1-x)]}}}{K\sqrt{4\ln(\tilde{\ell}_{1})\ln(\tilde{\ell}_{2})-\ln^{2}(1-x)}}. (94)

Notice that the Luttinger parameter KK enters in the Gaussian factor as an overall rescaling of its variance.

In the limit of large separation between the intervals, i.e. dd\to\infty (x0x\to 0), Eq. (94) tends to

limd𝒵nA1:A2(q1,q2)ZnA(0)=nπ2Kenπ2q122Kln~1ln~1enπ2q222Kln~2ln~2,\lim_{d\to\infty}\frac{\mathcal{Z}_{n}^{A_{1}:A_{2}}(q_{1},q_{2})}{Z_{n}^{A}(0)}=\frac{n\pi}{2K}\frac{e^{-\frac{n\pi^{2}q_{1}^{2}}{2K\ln\tilde{\ell}_{1}}}}{\sqrt{\ln\tilde{\ell}_{1}}}\frac{e^{-\frac{n\pi^{2}q_{2}^{2}}{2K\ln\tilde{\ell}_{2}}}}{\sqrt{\ln\tilde{\ell}_{2}}}, (95)

namely 𝒵nA1:A2(q1,q2)\mathcal{Z}_{n}^{A_{1}:A_{2}}(q_{1},q_{2}) factorises into the contributions of A1A_{1} and A2A_{2}. This is consistent with the probabilistic interpretation for the case n=1n=1: the outcomes of the charge measurements in the two intervals are independently distributed when the separation between A1A_{1} and A2A_{2} is large enough. On the other hand, in the limit of two adjacent intervals, i.e. d0d\to 0 (x1x\to 1), the multi-charged moments have the form (see also Eq. (45))

limd0ZnA1:A2(α,β)ZnA(0)=[~1αβα2~2αββ2(~1+~2)αβ]K2π2n,\lim_{d\to 0}\frac{Z_{n}^{A_{1}:A_{2}}(\alpha,\beta)}{Z_{n}^{A}(0)}=\left[\frac{\tilde{\ell}_{1}^{\,\alpha\beta-\alpha^{2}}\tilde{\ell}_{2}^{\,\alpha\beta-\beta^{2}}}{(\tilde{\ell}_{1}+\tilde{\ell}_{2})^{\alpha\beta}}\right]^{\frac{K}{2\pi^{2}n}}, (96)

whose Fourier transform is

limd0𝒵nA1:A2(q1,q2)ZnA1A2(0)=nπe2π2nq12ln~2+q22ln~1+q1q2[ln(~1~2)ln(~1+~2))]4Kln(~1)ln(~2)K[ln(~1~2)ln(~1+~2))]2K4ln(~1)ln(~2)[ln(~1~2)ln(~1+~2)]2.\lim_{d\to 0}\frac{\mathcal{Z}_{n}^{A_{1}:A_{2}}(q_{1},q_{2})}{Z_{n}^{A_{1}\cup A_{2}}(0)}=\frac{n\pi e^{-2\pi^{2}n\frac{q_{1}^{2}\ln\tilde{\ell}_{2}+q_{2}^{2}\ln\tilde{\ell}_{1}+q_{1}q_{2}\left[\ln(\tilde{\ell}_{1}\tilde{\ell}_{2})-\ln\left(\tilde{\ell}_{1}+\tilde{\ell}_{2})\right)\right]}{4K\ln(\tilde{\ell}_{1})\ln(\tilde{\ell}_{2})-K\left[\ln(\tilde{\ell}_{1}\tilde{\ell}_{2})-\ln\right(\tilde{\ell}_{1}+\tilde{\ell}_{2})\left)\right]^{2}}}}{K\sqrt{4\ln(\tilde{\ell}_{1})\ln(\tilde{\ell}_{2})-\left[\ln(\tilde{\ell}_{1}\tilde{\ell}_{2})-\ln(\tilde{\ell}_{1}+\tilde{\ell}_{2})\right]^{2}}}. (97)

Setting α=β\alpha=\beta in Eq. (93), we obtain the charged moments (11) with a single flux

ZnA(α)=ZnA(0)[(1x)~1~2]α2K2π2n.Z_{n}^{A}(\alpha)=Z_{n}^{A}(0)\left[(1-x)\tilde{\ell}_{1}\tilde{\ell}_{2}\right]^{-\frac{\alpha^{2}K}{2\pi^{2}n}}. (98)

In this case, performing the Fourier transform of Eq. (12), we end up with

𝒵nA(q)=ZnA(0)πn2Kln[(1x)~1~2]eπ2nq22Kln[(1x)~1~2].\mathcal{Z}_{n}^{A}(q)=\frac{Z_{n}^{A}(0)\sqrt{\pi n}}{\sqrt{2K\ln\left[(1-x)\tilde{\ell}_{1}\tilde{\ell}_{2}\right]}}e^{-\frac{\pi^{2}nq^{2}}{2K\ln\left[(1-x)\tilde{\ell}_{1}\tilde{\ell}_{2}\right]}}. (99)

Taking n=1n=1 in this expression, we obtain the probability p(q)p(q) of having charge qq in the subsystem AA, namely p(q)=𝒵1(q)p(q)=\mathcal{Z}_{1}(q). Now we can plug it together with the result for 𝒵nA1:A2(q1,q2)\mathcal{Z}^{A_{1}:A_{2}}_{n}(q_{1},q_{2}) found in Eq. (94) into Eq. (17) to obtain the conditional probability p(q1,q2)p(q_{1},q_{2}) of having charge q1q_{1} and q2=qq1q_{2}=q-q_{1} in the intervals A1A_{1} and A2A_{2} if the total charge in AA is qq,

p(q1,q2)=2πln[(1x)~1~2]K[4ln(~1)ln(~2)ln2(1x)]e2π2K[q12ln~2+q22ln~1+q1q2ln(1x)4ln(~1)ln(~2)ln2(1x)+(q1+q2)24ln[(1x)~1~2]].p(q_{1},q_{2})=\sqrt{\frac{2\pi\ln\left[(1-x)\tilde{\ell}_{1}\tilde{\ell}_{2}\right]}{K[4\ln(\tilde{\ell}_{1})\ln(\tilde{\ell}_{2})-\ln^{2}(1-x)]}}e^{-\frac{2\pi^{2}}{K}\left[\frac{q_{1}^{2}\ln\tilde{\ell}_{2}+q_{2}^{2}\ln\tilde{\ell}_{1}+q_{1}q_{2}\ln(1-x)}{4\ln(\tilde{\ell}_{1})\ln(\tilde{\ell}_{2})-\ln^{2}(1-x)}+\frac{(q_{1}+q_{2})^{2}}{4\ln\left[(1-x)\tilde{\ell}_{1}\tilde{\ell}_{2}\right]}\right]}. (100)

In this expression, ~p=λ1p\tilde{\ell}_{p}=\lambda_{1}\ell_{p}; in particular, for the tight-binding model λ1=eln2+1+γE\lambda_{1}=e^{\ln 2+1+\gamma_{\rm E}}, with γE\gamma_{\rm E} the Euler-Mascheroni constant. As a non-trivial consistency check, we have verified that the probability functions we have obtained satisfy the normalisation conditions

𝒵1A1:A2(q1,q2)dq1dq2=1,qp(q1,qq1)dq1=1,\int_{-\infty}^{\infty}\mathcal{Z}_{1}^{A_{1}:A_{2}}(q_{1},q_{2})\mathrm{d}q_{1}\mathrm{d}q_{2}=1,\quad\int_{-\infty}^{q}p(q_{1},q-q_{1})\mathrm{d}q_{1}=1, (101)

in agreement with Eqs. (17) and (16).

In Fig. 3, we compare the expression for 𝒵1(q1,q2)\mathcal{Z}_{1}(q_{1},q_{2}) found in Eq. (94) for the case of disjoint intervals with the exact numerical results obtained for the tight-binding model using the methods of Appendix A. The agreement is excellent. We remark that in Fig. 3 there is no free parameter when matching the analytical prediction with the numerical data since we know the expression of the non-universal constants for this particular system. In Fig. 4, we have repeated the same analysis in the case of adjacent intervals (d=0d=0), checking the validity of Eq. (97).

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 3: Probability 𝒵1(q1,q2)\mathcal{Z}_{1}(q_{1},q_{2}) for the tight-binding model as a function of q1q_{1} at fixed q2q_{2} and for two disjoint intervals of lengths 1\ell_{1}, 2\ell_{2}, and separated by a distance dd. The points are the exact numerical values calculated using the methods of Appendix A. The solid line is the theoretical prediction in Eq. (94) taking for the non-universal constants the corresponding values for the tight-binding model indicated in the main text.
Refer to caption
Refer to caption
Refer to caption
Figure 4: Probability 𝒵1(q1,q2)\mathcal{Z}_{1}(q_{1},q_{2}) as a function of q1q_{1} at fixed q2q_{2} for two adjacent intervals of sizes 1\ell_{1} and 2\ell_{2}. The solid blue line is the theoretical prediction in Eq. (97).

5.2 Symmetry-resolved mutual information

We compute now the symmetry-resolved mutual information defined in Eq. (8). We need the probability p(q1,qq1)p(q_{1},q-q_{1}) derived in Eq. (100) as well as the symmetry-resolved entropies for AA and its parts A1A_{1} and A2A_{2} separately. For the entropies of A1A_{1} and A2A_{2}, we can use the results for a single interval obtained in Ref. riccarda while, for the full subsystem AA, it can be derived from the Fourier transform of the charged moments determined in Eq. (99) by applying Eq. (13). The three symmetry-resolved entropies can eventually be written in the form

S1X(q)=S1X12ln(2KπlnδσΛ)12σπ4ξ2(Kln(λ1σΛ))2+σq2π4ξ(Kln(λ1σΛ))2.S_{1}^{X}(q)=S_{1}^{X}-\frac{1}{2}\ln\left(\frac{2K}{\pi}\ln\delta^{\sigma}\Lambda\right)-\frac{1}{2}-\sigma\pi^{4}\frac{\xi^{2}}{(K\ln(\lambda_{1}^{\sigma}\Lambda))^{2}}+\sigma q^{2}\pi^{4}\frac{\xi}{(K\ln(\lambda_{1}^{\sigma}\Lambda))^{2}}. (102)

where S1XS_{1}^{X} is the total entanglement entropy of subsystem XX. In this expression, when X=ApX=A_{p}, we have to take Λ=p\Lambda=\ell_{p} and σ=1\sigma=1 while, if X=A1A2X=A_{1}\cup A_{2}, then Λ=12(1x)\Lambda=\ell_{1}\ell_{2}(1-x) and σ=2\sigma=2. The auxiliary quantities δ\delta and ξ\xi in Eq. (102) are defined in terms of λn\lambda_{n} as

lnδ=lnλ1+2π2ξ,ξ=12π2n(lnλn)|n=1\ln\delta=\ln\lambda_{1}+2\pi^{2}\xi,\qquad\xi=-\frac{1}{2\pi^{2}}\partial_{n}(\ln\lambda_{n})|_{n=1} (103)

For the tight-binding model, we know the explicit value of these non-universal constants,

lnδ=2π2γ2(1)+ln2,ξ=γ2(1)+γ2(1).\ln\delta=2\pi^{2}\gamma^{\prime}_{2}(1)+\ln 2,\qquad\xi=\gamma_{2}(1)+\gamma^{\prime}_{2}(1). (104)

From Eqs. (100) and (102), we can now obtain an explicit expression for the symmetry-resolved mutual information. Since the conditional probability p(q1,qq1)p(q_{1},q-q_{1}) satisfies Eq. (101), we have

IA1:A2(q)\displaystyle I^{A_{1}:A_{2}}(q) =\displaystyle= IA1:A212ln[2Kπln(~1δ)ln(~2δ)ln(~1δ~2δ(1x))]122q2π4ξK2ln2(~1~2(1x))\displaystyle I^{A_{1}:A_{2}}-\frac{1}{2}\ln\left[\frac{2K}{\pi}\frac{\ln(\tilde{\ell}_{1}^{\delta})\ln(\tilde{\ell}_{2}^{\delta})}{\ln\left(\tilde{\ell}_{1}^{\delta}\tilde{\ell}_{2}^{\delta}(1-x)\right)}\right]-\frac{1}{2}{\color[rgb]{0,0,1}-}2q^{2}\pi^{4}\frac{\xi}{K^{2}\ln^{2}\left(\tilde{\ell}_{1}\tilde{\ell}_{2}(1-x)\right)} (105)
π4ξ2K2(1ln~1+1ln~21ln(~1~2(1x)))\displaystyle-\pi^{4}\frac{\xi^{2}}{K^{2}}\left(\frac{1}{\ln\tilde{\ell}_{1}}+\frac{1}{\ln\tilde{\ell}_{2}}-\frac{1}{\ln(\tilde{\ell}_{1}\tilde{\ell}_{2}(1-x))}\right)
+π4K2ξp(q1,qq1)[q12ln2~1+(qq1)2ln2~2]dq1,\displaystyle+\frac{\pi^{4}}{K^{2}}\xi\int_{-\infty}^{\infty}p(q_{1},q-q_{1})\left[\frac{q_{1}^{2}}{\ln^{2}\tilde{\ell}_{1}}+\frac{(q-q_{1})^{2}}{\ln^{2}\tilde{\ell}_{2}}\right]{\rm d}q_{1},

where IA1:A2I^{A_{1}:A_{2}} is the total mutual information of Eq. (3) and we have introduced the rescaled subsystem length ~pδ=δp\tilde{\ell}_{p}^{\delta}=\delta\ell_{p}. We plot this function in Fig. 5. As we anticipated, the symmetry-resolved mutual information is not a good measure of the total correlations between A1A_{1} and A2A_{2} in each symmetry sector since it can assume negative values.

Refer to captionRefer to caption
Figure 5: Symmetry-resolved mutual information of Eq. (105) in the tight-binding model. We plot our analytical prediction for different combinations of 1,2\ell_{1},\ell_{2} as a function of the cross-ratio xx (left panel) and qq (right panel).

Finally, we can also derive the number mutual information defined in Eq. (10). Applying in that formula the result for IA1:A2(q)I^{A_{1}:A_{2}}(q) obtained in Eq. (105), we have

InumA1:A2\displaystyle I^{A_{1}:A_{2}}_{\textrm{num}} =\displaystyle= 12ln[2Kπln(~1δ)ln(~2δ)ln(~1δ~2δ(1x))]+12+2π2ξKln(~1~2(1x))\displaystyle\frac{1}{2}\ln\left[\frac{2K}{\pi}\frac{\ln(\tilde{\ell}_{1}^{\delta})\ln(\tilde{\ell}_{2}^{\delta})}{\ln\left(\tilde{\ell}_{1}^{\delta}\tilde{\ell}_{2}^{\delta}(1-x)\right)}\right]+\frac{1}{2}+\frac{2\pi^{2}\xi}{K\ln\left(\tilde{\ell}_{1}\tilde{\ell}_{2}(1-x)\right)}
π4ξ2K2(1ln~1+1ln~21ln(~1~2(1x)))\displaystyle-\pi^{4}\frac{\xi^{2}}{K^{2}}\left(\frac{1}{\ln\tilde{\ell}_{1}}+\frac{1}{\ln\tilde{\ell}_{2}}-\frac{1}{\ln(\tilde{\ell}_{1}\tilde{\ell}_{2}(1-x))}\right)
π4ξK2𝒵1A1:A2(q1,qq1)[q12ln2~1+(qq1)2ln2~2]dq1dq.\displaystyle-\pi^{4}\frac{\xi}{K^{2}}\int_{-\infty}^{\infty}\mathcal{Z}_{1}^{A_{1}:A_{2}}(q_{1},q-q_{1})\left[\frac{q_{1}^{2}}{\ln^{2}\tilde{\ell}_{1}}+\frac{(q-q_{1})^{2}}{\ln^{2}\tilde{\ell}_{2}}\right]{\rm d}q_{1}{\rm d}q.
(106)

Since q=q1+q2q=q_{1}+q_{2} and

𝒵1A1:A2(q1,q2)qp2dq1dq2=Klnp~π2,p=1,2,\int_{-\infty}^{\infty}\mathcal{Z}_{1}^{A_{1}:A_{2}}(q_{1},q_{2})q_{p}^{2}{\rm d}q_{1}{\rm d}q_{2}=\frac{K\ln\tilde{\ell_{p}}}{\pi^{2}},\quad p=1,2, (107)

then Eq. (5.2) becomes

InumA1:A2=12ln[2Kπln(~1δ)ln(~2δ)ln(~1δ~2δ(1x))]+12π4ξ2K2(1ln~1+1ln~21ln(~1~2(1x)))π2ξK(1ln~1+1ln~22ln(~1~2(1x))).I^{A_{1}:A_{2}}_{\textrm{num}}=\frac{1}{2}\ln\left[\frac{2K}{\pi}\frac{\ln(\tilde{\ell}_{1}^{\delta})\ln(\tilde{\ell}_{2}^{\delta})}{\ln(\tilde{\ell}_{1}^{\delta}\tilde{\ell}_{2}^{\delta}(1-x))}\right]+\frac{1}{2}-\pi^{4}\frac{\xi^{2}}{K^{2}}\left(\frac{1}{\ln\tilde{\ell}_{1}}+\frac{1}{\ln\tilde{\ell}_{2}}-\frac{1}{\ln(\tilde{\ell}_{1}\tilde{\ell}_{2}(1-x))}\right)\\ -\pi^{2}\frac{\xi}{K}\left(\frac{1}{\ln\tilde{\ell}_{1}}+\frac{1}{\ln\tilde{\ell}_{2}}-\frac{2}{\ln(\tilde{\ell}_{1}\tilde{\ell}_{2}(1-x))}\right). (108)

In the limit 1,2,d\ell_{1},\ell_{2},d\to\infty, this expression behaves as

InumA1:A212ln[2Kπln1ln2ln(12(1x))]+12.I^{A_{1}:A_{2}}_{{\rm num}}\sim\frac{1}{2}\ln\left[\frac{2K}{\pi}\frac{\ln\ell_{1}\ln\ell_{2}}{\ln\left(\ell_{1}\ell_{2}(1-x)\right)}\right]+\frac{1}{2}. (109)

This result resembles the one for the number entropy of a single interval (see e.g. riccarda ), where a double logarithmic correction in the subsystem length also appears, even though, in our case, the dependence on the parameters 1,2,d\ell_{1},\ell_{2},d is more involved. On the other hand, it is a simple function of the Luttinger parameter KK since, as we already pointed out, the only effect of KK in the Gaussian factor of the Fourier transform of the multi-charged moments is renormalising its variance.

6 Conclusions

In this manuscript, we have computed the multi-charged moments of two intervals in the ground state of the free massless Dirac field and the massless compact boson, with arbitrary compactification radius. Using the replica approach, the multi-charged moments are given by the partition function of the theory on a higher genus Riemann surface with a different magnetic flux inserted in each interval. We have carried out the analysis of such partition function for the two CFTs under investigation in full generality, allowing the background magnetic flux to generate a different twisted boundary condition at each end-point of the intervals. In the case of the Dirac field, we have adapted the diagonalisation in the replica space method of Ref. CFH , to account for the different monodromy of the fields at each end-point. In the compact boson theory, we have chosen a geometric approach, and we have directly considered the four-point correlator on the Riemann surface of the vertex operators that implement the flux. It turns out that the known formulae for such correlator diverge in our case Verlinde . Once we properly regularised them, we have obtained a cumbersome expression for the multi-charged moments in terms of Riemann-Siegel Theta functions. Nevertheless, we have found several remarkable identities concerning the prime forms of the Riemann surface that allow to dramatically simplify the final result for the multi-charged moments. The factor due to the magnetic fluxes is eventually an algebraic function of the lengths of the intervals and their separation—identical to the one obtained in the Dirac theory. In fact, for a certain value of the compactification radius, the multi-charged moments of both theories are equal, generalising the known identity between their two-interval Rényi entropies HLM13 .

Given the simple expression obtained for the multi-charged moments, we can easily calculate their Fourier transform, which has a Gaussian form. From it, we have finally derived formulae for several interesting quantities such as the joint probability distribution of the charge for simultaneous measures in the two intervals, the symmetry-resolved mutual information pbc-21 and the number mutual information.

Let us conclude this manuscript discussing few outlooks. The multi-charged moments analysed here can be used to study the symmetry decomposition of the negativity in imbalance sectors goldstein1 . This is a measure of entanglement in mixed states which involves the partial transposition of the reduced density matrix. In the replica approach, this operation can be performed by properly fixing the different fluxes and exchanging the end-points of the transposed interval. A further generalisation is to identify the holographic dual of the multi-charged moments, which would be the starting point to compute the symmetry-resolved mutual information in the AdS/CFT correspondence, as done for the entanglement entropy in Ref. znm-20 . Partition functions with twisted boundary conditions, as the ones considered here, have been also proposed as non-local order parameters to distinguish various topological phases of spin chains ssr1-17 ; ssr-16 . We think that our analysis for the multi-charged moments can be useful to make progresses also in that direction. Finally, even though far beyond the scope of this manuscript, it would be interesting to obtain a rigorous proof of the identities for the prime forms that we have found and numerically checked here.

Acknowledgments

We thank Benoit Estienne, Yacine Ikhlef and Tamara Grava for useful discussions. PC, FA and SM acknowledge support from ERC under Consolidator grant number 771536 (NEMO).

Appendices

Appendix A Numerical Methods

For the numerical test of our field theory results, we consider the following lattice discretisation of the Dirac fermion, known as tight-binding model,

H=12j=(cj+1cj+h.c.),H=-\frac{1}{2}\sum_{j=-\infty}^{\infty}(c^{\dagger}_{j+1}c_{j}+\mathrm{h.c.}), (110)

where cjc_{j}^{\dagger} and cjc_{j} are fermionic creation and annihilation operators that satisfy the anti-commutation relations {cj,ck}=δjk\{c_{j},c^{\dagger}_{k}\}=\delta_{jk}. In terms of them, the charge operator reads

Q=j=(cjcj12).Q=\sum_{j=-\infty}^{\infty}\left(c^{\dagger}_{j}c_{j}-\frac{1}{2}\right). (111)

The two-point correlation functions in the ground state of Eq. (110) are of the form

cjck=sin(π(jk))2π(jk),\braket{c^{\dagger}_{j}c_{k}}=\frac{\sin(\pi(j-k))}{2\pi(j-k)}, (112)

and, due to the particle number conservation, cjck=0\langle c_{j}c_{k}\rangle=0. As well-known p-03 ; pe-09 , the moments Tr[ρAn]{\rm Tr}[\rho_{A}^{n}] can be calculated from the restriction of the two-point correlation matrix to the subsystem AA, that is (CA)j,k=cjck(C_{A})_{j,k}=\langle c_{j}^{\dagger}c_{k}\rangle, with j,kAj,k\in A. The charged moments ZnA=A1A2(α)Z^{A=A_{1}\cup A_{2}}_{n}(\alpha) can also be easily computed numerically in terms of the matrix CAC_{A} using the formula goldstein ; riccarda

ZnA=A1A2(α)=j=11+2[(εj)neiα/2+(1εj)neiα/2],Z_{n}^{A=A_{1}\cup A_{2}}(\alpha)=\displaystyle\prod_{j=1}^{\ell_{1}+\ell_{2}}[(\varepsilon_{j})^{n}e^{i\alpha/2}+(1-\varepsilon_{j})^{n}e^{-i\alpha/2}], (113)

where εj\varepsilon_{j} are the eigenvalues of CAC_{A} and p\ell_{p} is the number of sites in the interval ApA_{p}. In the case of the multi-charged moments ZA1:A2(α,β)Z^{A_{1}:A_{2}}(\alpha,\beta) defined in Eq. (14), the method used to compute ZnA(α)Z_{n}^{A}(\alpha) can not be applied since ρA\rho_{A} does not commute with the charges QA1Q_{A_{1}} and QA2Q_{A_{2}} of the two parts of AA. Following Ref. pbc-21 (which was based on gec-18 ), we rewrite Eq. (14) as

Z1A1:A2(α,β)=Z~ATrA(ρAρ~A),Z^{A_{1}:A_{2}}_{1}(\alpha,\beta)=\tilde{Z}_{A}\mathrm{Tr}_{A}(\rho_{A}\tilde{\rho}_{A}), (114)

where

ρ~A=1Z~AeiαQA1+iβQA2,Z~A=TrA(eiαQA1+iβQA2).\tilde{\rho}_{A}=\frac{1}{\tilde{Z}_{A}}e^{i\alpha Q_{A_{1}}+i\beta Q_{A_{2}}},\,\quad\tilde{Z}_{A}=\mathrm{Tr}_{A}(e^{i\alpha Q_{A_{1}}+i\beta Q_{A_{2}}}). (115)

Although ρ~A\tilde{\rho}_{A} is not a density matrix, it is a Gaussian operator with an associated two-point correlation matrix, C~A\tilde{C}_{A}, given by

(C~A)kj=δkj{eiα1+eiαjA1,eiβ1+eiβjA2.(\tilde{C}_{A})_{kj}=\delta_{kj}\begin{cases}\frac{e^{i\alpha}}{1+e^{i\alpha}}\quad j\in A_{1},\\ \frac{e^{i\beta}}{1+e^{i\beta}}\quad j\in A_{2}.\end{cases} (116)

Applying the rules for the composition of Gaussian operators fc-10 and introducing W=2CA𝕀W=2C_{A}-\mathbb{I}, we get pbc-21

Z1A1:A2(α,β)=(eiα/2+eiα/2)1(eiβ/2+eiβ/2)2det(𝟙1+2+Wαβ2),Z^{A_{1}:A_{2}}_{1}(\alpha,\beta)=(e^{-i\alpha/2}+e^{i\alpha/2})^{\ell_{1}}(e^{-i\beta/2}+e^{i\beta/2})^{\ell_{2}}\mathrm{det}\left(\frac{\mathbbm{1}_{\ell_{1}+\ell_{2}}+W_{\alpha\beta}}{2}\right), (117)

where

Wαβ=(W11W12W21W22)(eiα1eiα+1𝟙100eiβ1eiβ+1𝟙2),W_{\alpha\beta}=\begin{pmatrix}W_{11}&W_{12}\\ W_{21}&W_{22}\\ \end{pmatrix}\begin{pmatrix}\frac{e^{i\alpha}-1}{e^{i\alpha}+1}\mathbbm{1}_{\ell_{1}}&0\\ 0&\frac{e^{i\beta}-1}{e^{i\beta}+1}\mathbbm{1}_{\ell_{2}}\\ \end{pmatrix}, (118)

and the notation WppW_{pp^{\prime}}, p,p=1,2p,p^{\prime}=1,2, refers to correlations between sites in ApA_{p} and ApA_{p^{\prime}}. This result allows to exactly compute the multi-charged moments in the tight-binding model for different values of α\alpha and β\beta, as showed in Fig. 2. The Fourier transform of Z1A1:A2(α,β)Z_{1}^{A_{1}:A_{2}}(\alpha,\beta) gives the quantities 𝒵1A1:A2(q1,q2)\mathcal{Z}_{1}^{A_{1}:A_{2}}(q_{1},q_{2}) analysed in Fig. 3 and Fig. 4.

Appendix B Derivation of the normalised holomorphic differentials

In this Appendix, we present the detailed derivation of the expression given in Eq. (64) for the normalised holomorphic differentials νr(z)\nu_{r}(z) of the Riemann surface Σn(x)\Sigma_{n}(x). Recall that, according to Eq. (62), they are normalised with respect to the homology basis {ar,br}\{a_{r},b_{r}\}. In order to define this basis, it is convenient to introduce an auxiliary basis of non-contractible cycles, which we call a~r\tilde{a}_{r} and b~r\tilde{b}_{r}. The cycle a~r\tilde{a}_{r} encloses anticlockwise the branch cut (u1,v1)(u_{1},v_{1}) in the rr-sheet of the surface. The dual cycle b~r\tilde{b}_{r} connects anticlockwise the rr and nn sheets through the branch cut (u1,v1)(u_{1},v_{1}) and then it goes back to the rr sheet through the branch cut (u2,v2)(u_{2},v_{2}). In Fig. 6, we draw the auxiliary homology basis in the case n=3n=3. The cycles ara_{r} and brb_{r} are defined in terms of the auxiliary ones by

ar=k=1ra~k,br=b~rb~r+1,a_{r}=\sum_{k=1}^{r}\tilde{a}_{k},\quad b_{r}=\tilde{b}_{r}-\tilde{b}_{r+1}, (119)

with r=1,,n1r=1,\dots,n-1 and b~n=0\tilde{b}_{n}=0.

We can now obtain νr(z)\nu_{r}(z) by following the usual procedure considered in the literature, see e.g. Enolskii ; Enolskii2 . We first take the basis of canonical holomorphic differentials of the surface Σn(x)\Sigma_{n}(x),

μr(z)=(z(z1))r/n(zx)r/n1,r=1,,n1.\mu_{r}(z)=(z(z-1))^{-r/n}(z-x)^{r/n-1},\quad r=1,\dots,n-1. (120)

If we denote by 𝒜\mathcal{A} and \mathcal{B} the (n1)×(n1)(n-1)\times(n-1) matrices with entries

𝒜r,s=ardzμs(z),r,s=brdzμs(z),\mathcal{A}_{r,s}=\oint_{a_{r}}{\rm d}z\mu_{s}(z),\quad\mathcal{B}_{r,s}=\oint_{b_{r}}{\rm d}z\mu_{s}(z), (121)

then the normalised holomorphic differentials νr(z)\nu_{r}(z) can be calculated from μr(z)\mu_{r}(z) such that

νr(z)=l=1n1μl(z)𝒜l,r1,\nu_{r}(z)=\sum_{l=1}^{n-1}\mu_{l}(z)\mathcal{A}^{-1}_{l,r}, (122)

where 𝒜l,r1\mathcal{A}^{-1}_{l,r} are the entries of the inverse of the matrix 𝒜\mathcal{A}. Furthermore, by combining Eq. (63) with Eqs. (121) and (122), the matrix of periods Γ(x)\Gamma(x) in Eq. (56) can be obtained from

Γ(x)=𝒜1.\Gamma(x)=\mathcal{B}\mathcal{A}^{-1}. (123)

To compute the matrices 𝒜\mathcal{A} and \mathcal{B}, it is useful to consider the auxiliary homology basis {a~r,b~r}\{\tilde{a}_{r},\tilde{b}_{r}\}.

Refer to caption
Figure 6: Auxiliary homology basis of the surface Σn(x)\Sigma_{n}(x) for n=3n=3.

The advantage of taking this basis is that it is easier to calculate the contour integrals of μr(z)\mu_{r}(z) along the cycles a~r\tilde{a}_{r} and b~r\tilde{b}_{r} than along ara_{r} and brb_{r}. In fact, let us denote by 𝒜~\tilde{\mathcal{A}} and ~\tilde{\mathcal{B}} the (n1)×(n1)(n-1)\times(n-1) matrices analogous to 𝒜\mathcal{A} and \mathcal{B} but integrating now along the auxiliary cycles a~r\tilde{a}_{r} and b~r\tilde{b}_{r} respectively,

𝒜~r,s=a~rdzμs(z),~r,s=b~rdzμs(z).\tilde{\mathcal{A}}_{r,s}=\oint_{\tilde{a}_{r}}{\rm d}z\mu_{s}(z),\quad\tilde{\mathcal{B}}_{r,s}=\oint_{\tilde{b}_{r}}{\rm d}z\mu_{s}(z). (124)

Then, by taking the correct branches for μr(z)\mu_{r}(z), we find that

𝒜~r,s\displaystyle\tilde{\mathcal{A}}_{r,s} =\displaystyle= e2πi(r1)sn(e2πisn1)0xdzμs(z)\displaystyle e^{\frac{2\pi i(r-1)s}{n}}\left(e^{-\frac{2\pi is}{n}}-1\right)\int_{0}^{x}{\rm d}z\mu_{s}(z) (125)
=\displaystyle= 2πieπi(2r3)snIs/n(x),\displaystyle 2\pi ie^{\frac{\pi i(2r-3)s}{n}}I_{s/n}(x),

and

~r,s\displaystyle\tilde{\mathcal{B}}_{r,s} =\displaystyle= e2πi(r1)sn(e2πirsn1)x1dzμs(z)\displaystyle e^{\frac{2\pi i(r-1)s}{n}}\left(e^{-\frac{2\pi irs}{n}}-1\right)\int_{x}^{1}{\rm d}z\mu_{s}(z) (126)
=\displaystyle= 2πieiπ(r3)snsin(πrsn)sin(πsn)Is/n(1x),\displaystyle-2\pi ie^{\frac{i\pi(r-3)s}{n}}\frac{\sin\left(\frac{\pi rs}{n}\right)}{\sin\left(\frac{\pi s}{n}\right)}I_{s/n}(1-x),

where we have employed the identities

0xdzμs(z)=πsin(πsn)Is/n(x),x1dzμs(z)=πeiπsnsin(πsn)Is/n(1x).\int_{0}^{x}{\rm d}z\mu_{s}(z)=-\frac{\pi}{\sin\left(\frac{\pi s}{n}\right)}I_{s/n}(x),\quad\int_{x}^{1}{\rm d}z\mu_{s}(z)=\frac{\pi e^{-\frac{i\pi s}{n}}}{\sin\left(\frac{\pi s}{n}\right)}I_{s/n}(1-x). (127)

If we apply now the relation of Eq. (119) between both homology basis, then we directly obtain the matrices 𝒜\mathcal{A},

𝒜r,s=k=1r𝒜~k,s=2πieiπ(r2)snsin(πrsn)sin(πsn)Is/n(x),\displaystyle\mathcal{A}_{r,s}=\sum_{k=1}^{r}\tilde{\mathcal{A}}_{k,s}=2\pi ie^{\frac{i\pi(r-2)s}{n}}\frac{\sin\left(\frac{\pi rs}{n}\right)}{\sin\left(\frac{\pi s}{n}\right)}I_{s/n}(x), (128)

and \mathcal{B}

r,s\displaystyle\mathcal{B}_{r,s} =\displaystyle= ~r,s~r+1,s\displaystyle\tilde{\mathcal{B}}_{r,s}-\tilde{\mathcal{B}}_{r+1,s} (129)
=\displaystyle= 2πieiπ(r2)sn[sin(π(r+1)sn)eiπsnsin(πrsn)]Is/n(1x)sin(πsn).\displaystyle 2\pi ie^{\frac{i\pi(r-2)s}{n}}\left[\sin\left(\frac{\pi(r+1)s}{n}\right)-e^{-\frac{i\pi s}{n}}\sin\left(\frac{\pi rs}{n}\right)\right]\frac{I_{s/n}(1-x)}{\sin\left(\frac{\pi s}{n}\right)}.

The entries of the inverse of 𝒜\mathcal{A} are

𝒜r,s1=e2πi(s1)rnsin(πrn)πnIr/n(x).\mathcal{A}_{r,s}^{-1}=\frac{e^{-\frac{2\pi i(s-1)r}{n}}\sin\left(\frac{\pi r}{n}\right)}{\pi nI_{r/n}(x)}. (130)

Using Eqs. (122) and (130), we finally arrive at the expression written in Eq. (64) for the normalised holomorphic differentials νr(z)\nu_{r}(z). The matrix of periods in Eq. (56) can be directly obtained by plugging Eqs. (129) and (130) into Eq. (123).

Appendix C Prime form identities

The identities for the regularised prime forms E()(z,z)E^{(*)}(z,z^{\prime}) of Eqs.(83)-(86) can be easily proved when n=2n=2 (genus one). In that case, the matrix of periods of Eq. (56) is a scalar Γ(x)=iβ1/2(x)\Gamma(x)={\rm i}\beta_{1/2}(x) and the Theta function Θ12\Theta_{\frac{1}{2}} that appears in Eq. (81) reduces to a Jacobi theta function, Θ12(𝒖|Γ(x))=ϑ1(𝒖|Γ(x))\Theta_{\frac{1}{2}}(\boldsymbol{u}|\Gamma(x))=-\vartheta_{1}(\boldsymbol{u}|\Gamma(x)). The images under the Abel-Jacobi map in Eqs. (68)-(70) are in this case

𝒘(x)=12,𝒘(1)=12+Γ(x)2,𝒘()=Γ(x)2.\boldsymbol{w}(x)=\frac{1}{2},\quad\boldsymbol{w}(1)=\frac{1}{2}+\frac{\Gamma(x)}{2},\quad\boldsymbol{w}(\infty)=\frac{\Gamma(x)}{2}. (131)

Here we illustrate the proof of Eq. (84). The rest of identities can be checked in a similar way by applying the different relations in Sec. 20.2 of Ref. nist . If n=2n=2, Eq. (81) takes the following form for z=xz=x, z=1z^{\prime}=1,

|E()(x,1)|=2πx1/4(1x)1/2I1/2(x)|ϑ1(Γ(x)/2|Γ(x))uϑ1(0|Γ(x))|,|E^{(*)}(x,1)|=2\pi x^{1/4}(1-x)^{1/2}I_{1/2}(x)\left|\frac{\vartheta_{1}(\Gamma(x)/2|\Gamma(x))}{\partial_{u}\vartheta_{1}(0|\Gamma(x))}\right|, (132)

where Ip(x)I_{p}(x) is defined below Eq. (57). If we apply the half-period translation ϑ1(Γ/2|Γ)=iM2(x)1ϑ4(0|Γ)\vartheta_{1}(\Gamma/2|\Gamma)=iM_{2}(x)^{-1}\vartheta_{4}(0|\Gamma), the equality

uϑ1(0|Γ)=πϑ2(0|Γ)ϑ3(0|Γ)ϑ4(0|Γ),\partial_{u}\vartheta_{1}(0|\Gamma)=\pi\vartheta_{2}(0|\Gamma)\vartheta_{3}(0|\Gamma)\vartheta_{4}(0|\Gamma), (133)

and the relation between the hypergeometric function and the Jacobi theta function I1/2(x)=ϑ3(0|Γ)2I_{1/2}(x)=\vartheta_{3}(0|\Gamma)^{2}, we find

|E()(x,1)|=2x1/4(1x)1/2M2(x)ϑ3(0|Γ)ϑ2(0|Γ).|E^{(*)}(x,1)|=\frac{2x^{1/4}(1-x)^{1/2}}{M_{2}(x)}\frac{\vartheta_{3}(0|\Gamma)}{\vartheta_{2}(0|\Gamma)}. (134)

Finally, employing the well-known equality

x1/4=ϑ2(0|Γ)ϑ3(0|Γ),x^{1/4}=\frac{\vartheta_{2}(0|\Gamma)}{\vartheta_{3}(0|\Gamma)}, (135)

we obtain Eq. (84).

The results for n=2n=2 led us to conjecture the generalisation of Eqs. (83)-(86) for higher genus. Unfortunately, we have not been able to find a proof for them, although they can be tested numerically with great accuracy, as we show in Fig. 7. The dots are the result of evaluating numerically the definition in Eq. (81) of E()(z,z)E^{(*)}(z,z^{\prime}), while the solid lines are the functions on the right hand side of the identities in Eqs. (83)-(86).

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 7: Numerical check of Eqs. (83)-(86) involving the regularised prime form E()(z,z)E^{(*)}(z,z^{\prime}). We test them for different values of the genus n1n-1. The points correspond to the direct numerical evaluation of the definition in Eq. (81) of E()(z,z)E^{(*)}(z,z^{\prime}). The continuous lines are the plot of the functions on the right hand side of Eqs. (83)-(86).

References

  • (1) M. A. Nielsen and I. L. Chuang, Quantum computation and quantum information, Cambridge University Press, Cambridge, UK, 10th anniversary ed. (2010).
  • (2) T. Nishioka, S. Ryu, and T. Takayanagi, Holographic entanglement entropy: an overview, J. Phys. A 42, 504008 (2009).
  • (3) M. Rangamani and T. Takanayagi, Holographic Entanglement Entropy, Lect. Notes Phys. 931 (2017).
  • (4) L. Amico, R. Fazio, A. Osterloh, and V. Vedral, Entanglement in many-body systems, Rev. Mod. Phys. 80, 517 (2008).
  • (5) P. Calabrese, J. Cardy, and B. Doyon, Entanglement entropy in extended quantum systems, J. Phys. A 42, 500301 (2009).
  • (6) J. Eisert, M. Cramer, and M. B. Plenio, Area laws for the entanglement entropy, Rev. Mod. Phys. 82, 277 (2010).
  • (7) N. Laflorencie, Quantum entanglement in condensed matter systems, Phys. Rep. 643, 1 (2016).
  • (8) C. G. Callan and F. Wilczek, On Geometric Entropy, Phys. Lett. B 333, 55 (1994).
  • (9) C. Holzhey, F. Larsen, and F. Wilczek, Geometric and renormalized entropy in conformal field theory, Nucl. Phys. B 424, 443 (1994).
  • (10) P. Calabrese and J. Cardy, Entanglement entropy and quantum field theory, J. Stat. Mech. (2004) P06002.
  • (11) P. Calabrese and J. Cardy, Entanglement entropy and conformal field theory, J. Phys. A 42, 504005 (2009).
  • (12) M. Caraglio and F. Gliozzi, Entanglement entropy and twist fields, JHEP 11 (2008) 076.
  • (13) S. Furukawa, V. Pasquier, and J. Shiraishi, Mutual Information and Boson Radius in c=1c=1 Critical Systems in One Dimension, Phys. Rev. Lett. 102, 170602 (2009).
  • (14) P. Calabrese, J. Cardy, and E. Tonni, Entanglement entropy of two disjoint intervals in conformal field theory, J. Stat. Mech. (2009) P11001.
  • (15) P. Calabrese, J. Cardy, and E. Tonni, Entanglement entropy of two disjoint intervals in conformal field theory II, J. Stat. Mech. (2011) P01021.
  • (16) P. Calabrese, J. Cardy, and E. Tonni, Entanglement negativity in quantum field theory, Phys. Rev. Lett. 109, 130502 (2012).
  • (17) P. Calabrese, J. Cardy, and E. Tonni, Entanglement negativity in extended systems: a quantum field theory approach, J. Stat. Mech. (2013) P02008 .
  • (18) P. Calabrese, L. Tagliacozzo, and E. Tonni, Entanglement negativity in the critical Ising chain, J. Stat. Mech. (2013) P05002.
  • (19) V. Alba, Entanglement negativity and conformal field theory: a Monte Carlo study, J. Stat. Mech. (2013) P05013.
  • (20) A. Coser, E. Tonni, and P. Calabrese, Towards the entanglement negativity of two disjoint intervals for a one dimensional free fermion, J. Stat. Mech. (2016) 033116.
  • (21) F. Ares, R. Santachiara, and J. Viti, Crossing-symmetric Twist Field Correlators and Entanglement Negativity in Minimal CFTs, JHEP 10 (2021) 175.
  • (22) G. Rockwood, Replicated Entanglement Negativity for Disjoint Intervals in the Ising and Free Compact Boson Conformal Field Theories, arXiv:2203.04339.
  • (23) T. Dupic, B. Estienne, and Y. Ikhlef, Entanglement entropies of minimal models from null-vectors, SciPost Phys. 4, 031 (2018).
  • (24) H. Casini, C. D. Fosco, and M. Huerta, Entanglement and alpha entropies for a massive Dirac field in two dimensions, J. Stat. Mech. (2005) P07007.
  • (25) H. Casini and M. Huerta, Reduced density matrix and internal dynamics for multicomponent regions, Class. Quant. Grav. 26, 185005 (2009)
  • (26) V. Alba, L. Tagliacozzo, and P. Calabrese, Entanglement entropy of two disjoint blocks in critical Ising models, Phys. Rev. B 81, 060411(R) (2010).
  • (27) V. Alba, L. Tagliacozzo, and P. Calabrese, Entanglement entropy of two disjoint intervals in c=1c=1 theories, J. Stat. Mech. (2011) P06012.
  • (28) M. A. Rajabpour and F. Gliozzi, Entanglement entropy of two disjoint intervals from fusion algebra of twist fields, J. Stat. Mech. (2012) P02016.
  • (29) M. Headrick, A. Lawrence, and M. Roberts, Bose-Fermi duality and entanglement entropies, J. Stat. Mech. (2013) P02022.
  • (30) A. Coser, L. Tagliacozzo, and E. Tonni, On Rényi entropies of disjoint intervals in conformal field theory, J. Stat. Mech. (2014) P01008.
  • (31) C. De Nobili, A. Coser, and E. Tonni, Entanglement entropy and negativity of disjoint intervals in CFT: some numerical extrapolations, J. Stat. Mech. (2015) P06021.
  • (32) A. Coser, E. Tonni, and P. Calabrese, Spin structures and entanglement of two disjoint intervals in conformal field theories, J. Stat. Mech. (2016) 053109.
  • (33) P. Ruggiero, E. Tonni, and P. Calabrese, Entanglement entropy of two disjoint intervals and the recursion formula for conformal blocks, J. Stat. Mech. (2018) 113101.
  • (34) A. Bastianello, Rényi entanglement entropies for the compactified massless boson with open boundary conditions, JHEP 10 (2019) 141.
  • (35) A. Bastianello, J. Dubail, and J.M. Stéphan, Entanglement entropies of inhomogeneous Luttinger liquids, J. Phys. A: Math. Theor. 53, 155001 (2020).
  • (36) T. Grava, A. P. Kels, and E. Tonni, Entanglement of Two Disjoint Intervals in Conformal Field Theory and the 2D Coulomb Gas on a Lattice, Phys. Rev. Lett. 127, 141605 (2021).
  • (37) M. Gerbershagen, Monodromy methods for torus conformal blocks and entanglement entropy at large central charge, JHEP 08 (2021) 143.
  • (38) N. Laflorencie and S. Rachel, Spin-resolved entanglement spectroscopy of critical spin chains and Luttinger liquids, J. Stat. Mech. (2014) P11013.
  • (39) M. Goldstein and E. Sela, Symmetry-Resolved Entanglement in Many-Body Systems, Phys. Rev. Lett. 120, 200602 (2018).
  • (40) J. C. Xavier, F. C. Alcaraz, and G. Sierra, Equipartition of the entanglement entropy, Phys. Rev. B 98, 041106 (2018).
  • (41) A. Lukin, M. Rispoli, R. Schittko, M. E. Tai, A. M. Kaufman, S. Choi, V. Khemani, J. Leonard, and M. Greiner, Probing entanglement in a many-body localized system, Science 364, 256 (2019).
  • (42) D. Azses, R. Haenel, Y. Naveh, R. Raussendorf, E. Sela, and E. G. Dalla Torre, Identification of Symmetry-Protected Topological States on Noisy Quantum Computers, Phys. Rev. Lett. 125, 120502 (2020).
  • (43) A. Neven, J. Carrasco, V. Vitale, C. Kokail, A. Elben, M. Dalmonte, P. Calabrese, P. Zoller, B. Vermersch, R. Kueng, and B. Kraus, Symmetry-resolved entanglement detection using partial transpose moments, Npj Quantum Inf. 7, 152 (2021).
  • (44) V. Vitale, A. Elben, R. Kueng, A. Neven, J. Carrasco, B. Kraus, P. Zoller, P. Calabrese, B. Vermersch, and M. Dalmonte, Symmetry-resolved dynamical purification in synthetic quantum matter, SciPost Phys. 12, 106 (2022).
  • (45) R. Bonsignori, P. Ruggiero, and P. Calabrese, Symmetry resolved entanglement in free fermionic systems, J. Phys. A 52, 475302 (2019).
  • (46) S. Fraenkel and M. Goldstein, Symmetry resolved entanglement: Exact results in 1d and beyond, J. Stat. Mech. (2020) 033106.
  • (47) N. Feldman and M. Goldstein, Dynamics of Charge-Resolved Entanglement after a Local Quench, Phys. Rev. B 100, 235146 (2019).
  • (48) S. Murciano, G. Di Giulio, and P. Calabrese, Symmetry resolved entanglement in gapped integrable systems: a corner transfer matrix approach, SciPost Phys. 8, 046 (2020).
  • (49) P. Calabrese, M. Collura, G. Di Giulio, and S. Murciano, Full counting statistics in the gapped XXZ spin chain, EPL 129, 60007 (2020).
  • (50) H. M. Wiseman and J. A. Vaccaro, Entanglement of Indistinguishable Particles Shared between Two Parties, Phys. Rev. Lett. 91, 097902 (2003).
  • (51) H. Barghathi, C. M. Herdman, and A. Del Maestro, Rényi Generalization of the Accessible Entanglement Entropy, Phys. Rev. Lett. 121, 150501 (2018).
  • (52) H. Barghathi, E. Casiano-Diaz, and A. Del Maestro, Operationally accessible entanglement of one dimensional spinless fermions, Phys. Rev. A 100, 022324 (2019).
  • (53) H. Barghathi, J. Yu, and A. Del Maestro Theory of noninteracting fermions and bosons in the canonical ensemble, Phys. Rev. Res. 2, 043206 (2020).
  • (54) S. Murciano, P. Ruggiero, and P. Calabrese, Symmetry resolved entanglement in two-dimensional systems via dimensional reduction, J. Stat. Mech. (2020) 083102.
  • (55) M. T. Tan and S. Ryu, Particle Number Fluctuations, Rényi and Symmetry-resolved Entanglement Entropy in Two-dimensional Fermi Gas from Multi-dimensional bosonisation, Phys. Rev. B 101, 235169 (2020).
  • (56) Z. Ma, C. Han, Y. Meir, and E. Sela, Symmetric inseparability and number entanglement in charge conserving mixed states, Phys. Rev. A 105, 042416 (2022).
  • (57) F. Ares, S. Murciano, and P. Calabrese, Symmetry-resolved entanglement in a long-range free-fermion chain, J. Stat. Mech. (2022) 063104.
  • (58) N. G. Jones, Symmetry-resolved entanglement entropy in critical free-fermion chains, J. Statist. Phys. 188 (2022) 28.
  • (59) L. Piroli, E. Vernier, M. Collura, and P. Calabrese, Thermodynamic symmetry resolved entanglement entropies in integrable systems, J. Stat. Mech. (2022) 073102.
  • (60) S. Murciano, G. Di Giulio, and P. Calabrese, Entanglement and symmetry resolution in two dimensional free quantum field theories, JHEP 08 (2020) 073.
  • (61) D. X. Horvath, L. Capizzi, and P. Calabrese, U(1) symmetry resolved entanglement in free 1+1 dimensional field theories via form factor bootstrap, JHEP 05 (2021) 197.
  • (62) D. X. Horvath, P. Calabrese, and O. A. Castro-Alvaredo, Branch Point Twist Field Form Factors in the sine-Gordon Model II: Composite Twist Fields and Symmetry Resolved Entanglement, SciPost Phys. 12, 088 (2022).
  • (63) D. X. Horvath and P. Calabrese, Symmetry resolved entanglement in integrable field theories via form factor bootstrap, JHEP 11 (2020) 131.
  • (64) L. Capizzi, D. X. Horvath, P. Calabrese, and O. A. Castro-Alvaredo, Entanglement of the 3-State Potts Model via Form Factor Bootstrap: Total and Symmetry Resolved Entropies, JHEP 05 (2022) 113.
  • (65) E. Cornfeld, M. Goldstein, and E. Sela, Imbalance Entanglement: Symmetry Decomposition of Negativity, Phys. Rev. A 98, 032302 (2018).
  • (66) L. Capizzi, P. Ruggiero, and P. Calabrese, Symmetry resolved entanglement entropy of excited states in a CFT, J. Stat. Mech. (2020) 073101.
  • (67) S. Murciano, R. Bonsignori, and P. Calabrese, Symmetry decomposition of negativity of massless free fermions, SciPost Phys. 10, 111 (2021).
  • (68) H.-H. Chen, Symmetry decomposition of relative entropies in conformal field theory, JHEP 07 (2021) 084.
  • (69) L. Capizzi and P. Calabrese, Symmetry resolved relative entropies and distances in conformal field theory, JHEP 10 (2021) 195.
  • (70) L. Hung and G. Wong, Entanglement branes and factorization in conformal field theory, Phys. Rev. D 104, 026012 (2021).
  • (71) P. Calabrese, J. Dubail, and S. Murciano, Symmetry-resolved entanglement entropy in Wess-Zumino-Witten models, JHEP 10 (2021) 067.
  • (72) R. Bonsignori and P. Calabrese, Boundary effects on symmetry resolved entanglement, J. Phys. A 54, 015005 (2021).
  • (73) B. Estienne, Y. Ikhlef, and A. Morin-Duchesne, Finite-size corrections in critical symmetry-resolved entanglement, SciPost Phys. 10, 054 (2021).
  • (74) H.-H. Chen, Charged Rényi negativity of massless free bosons, JHEP 02 (2022) 117.
  • (75) A. Milekhin and A. Tajdini, Charge fluctuation entropy of Hawking radiation: a replica-free way to find large entropy, arXiv:2109.03841.
  • (76) M. Ghasemi, Universal Thermal Corrections to Symmetry-Resolved Entanglement Entropy and Full Counting Statistics, arXiv:2203.06708.
  • (77) S. Zhao, C. Northe, and R. Meyer, Symmetry-Resolved Entanglement in AdS3/CFT2 coupled to U(1)U(1) Chern-Simons Theory, JHEP 07 (2021) 030.
  • (78) K. Weisenberger, S. Zhao, C. Northe, and R. Meyer, Symmetry-resolved entanglement for excited states and two entangling intervals in AdS3/CFT2, JHEP 12 (2021) 104.
  • (79) S. Zhao, C. Northe, K. Weisenberger, and R. Meyer, Charged Moments in W3W_{3} Higher Spin Holography, JHEP 05 (2022) 166.
  • (80) S. Baiguera, L. Bianchi, S. Chapman, and D. A. Galante, Shape Deformations of Charged Rényi Entropies from Holography, JHEP 06 (2022) 068.
  • (81) G. Parez, R. Bonsignori, and P. Calabrese, Quasiparticle dynamics of symmetry resolved entanglement after a quench: the examples of conformal field theories and free fermions, Phys. Rev. B 103, L041104 (2021).
  • (82) G. Parez, R. Bonsignori, and P. Calabrese, Exact quench dynamics of symmetry resolved entanglement in a free fermion chain, J. Stat. Mech. (2021) 093102.
  • (83) S. Fraenkel and M. Goldstein, Entanglement Measures in a Nonequilibrium Steady State: Exact Results in One Dimension, SciPost Phys. 11, 085 (2021).
  • (84) G. Parez, R. Bonsignori, and P. Calabrese, Dynamics of charge-imbalance-resolved entanglement negativity after a quench in a free-fermion mode, J. Stat. Mech. (2022) 053103.
  • (85) S. Scopa and D. X. Horvath, Exact hydrodynamic description of symmetry-resolved Rényi entropies after a quantum quench, J. Stat. Mech. (2022) 083104.
  • (86) H.-H. Chen, Dynamics of charge imbalance resolved negativity after a global quench in free scalar field theory, JHEP 08 (2022) 146.
  • (87) X. Turkeshi, P. Ruggiero, V. Alba, and P. Calabrese, Entanglement equipartition in critical random spin chains, Phys. Rev. B 102, 014455 (2020).
  • (88) M. Kiefer-Emmanouilidis, R. Unanyan, J. Sirker, and M. Fleischhauer, Bounds on the entanglement entropy by the number entropy in non-interacting fermionic systems, SciPost Phys 8, 083 (2020).
  • (89) M. Kiefer-Emmanouilidis, R. Unanyan, J. Sirker, and M. Fleischhauer, Evidence for unbounded growth of the number entropy in many-body localized phases, Phys. Rev. Lett. 124, 243601 (2020).
  • (90) M. Kiefer-Emmanouilidis, R. Unanyan, M. Fleischhauer, and J. Sirker, Absence of true localization in many-body localized phases, Phys. Rev. B 103, 024203 (2021).
  • (91) E. Cornfeld, L. A. Landau, K. Shtengel, and E. Sela, Entanglement spectroscopy of non-Abelian anyons: Reading off quantum dimensions of individual anyons, Phys. Rev. B 99, 115429 (2019).
  • (92) K. Monkman and J. Sirker, Operational Entanglement of Symmetry-Protected Topological Edge States, Phys. Rev. Res. 2, 043191 (2020).
  • (93) D. Azses and E. Sela, Symmetry resolved entanglement in symmetry protected topological phases, Phys. Rev. B 102, 235157 (2020).
  • (94) D. Azses, R. Haenel, Y. Naveh, R. Raussendorf, E. Sela, and E. G. Dalla Torre, Identification of Symmetry-Protected Topological States on Noisy Quantum Computers, Phys. Rev. Lett. 125, 120502 (2020).
  • (95) D. Azses, E. G. Dalla Torre, and E. Sela, Observing Floquet topological order by symmetry resolution, Phys. Rev. B 104, L220301 (2021).
  • (96) B. Oblak, N. Regnault, and B. Estienne, Equipartition of Entanglement in Quantum Hall States, Phys. Rev. B 105, 115131 (2022).
  • (97) A. Belin, L.-Y. Hung, A. Maloney, S. Matsuura, R. C. Myers, and T. Sierens, Holographic charged Rényi entropies, JHEP 12 (2013) 059.
  • (98) P. Caputa, G. Mandal, and R. Sinha, Dynamical entanglement entropy with angular momentum and U(1)U(1) charge, JHEP 11 (2013) 052.
  • (99) P. Caputa, M. Nozaki, and T. Numasawa, Charged Entanglement Entropy of Local Operators, Phys. Rev. D 93, 105032 (2016).
  • (100) J. S. Dowker, Conformal weights of charged Rényi entropy twist operators for free scalar fields in arbitrary dimensions, J. Phys. A 49, 145401 (2016).
  • (101) J. S. Dowker, Charged Rényi entropies for free scalar fields, J. Phys. A 50, 165401 (2017).
  • (102) H. Shapourian, K. Shiozaki, and S. Ryu, Partial time-reversal transformation and entanglement negativity in fermionic systems, Phys. Rev. B 95, 165101 (2017).
  • (103) H. Shapourian, P. Ruggiero, S. Ryu, and P. Calabrese, Twisted and untwisted negativity spectrum of free fermions, SciPost Phys. 7, 037 (2019).
  • (104) H. Shapourian, K. Shiozaki, and S. Ryu, Many-Body Topological Invariants for Fermionic Symmetry-Protected Topological Phases, Phys. Rev. Lett. 118, 216402 (2017).
  • (105) K. Shiozaki, H. Shapourian, and S. Ryu, Many-body topological invariants in fermionic symmetry protected topological phases: Cases of point group symmetries, Phys. Rev. B 95, 205139 (2017).
  • (106) M. Kiefer-Emmanouilidis, R. Unanyan, M. Fleischhauer, and J. Sirker, Slow delocalization of particles in many-body localized phases, Phys. Rev. B 103, 024203 (2021).
  • (107) Y. Zhao, D. Feng, Y. Hu, S. Guo, and J. Sirker, Entanglement dynamics in the three-dimensional Anderson model, Phys. Rev. B 102, 195132 (2020).
  • (108) M. Kiefer-Emmanouilidis, R. Unanyan, M. Fleischhauer, and J. Sirker, Unlimited growth of particle fluctuations in many-body localized phases, Ann. Phys. 168481 (2021).
  • (109) J. L. Cardy, O. A. Castro-Alvaredo, and B. Doyon, Form factors of branch-point twist fields in quantum integrable models and entanglement entropy, J. Stat. Phys. 130, 129 (2008).
  • (110) V. Knizhnik, Analytic fields on Riemann surfaces. II, Comm. Math. Phys. 112, 567 (1987).
  • (111) L. J. Dixon, D. Friedan, E. J. Martinec, and S. H. Shenker, The Conformal Field Theory Of Orbifolds, Nucl. Phys. B 282, 13 (1987).
  • (112) B.-Q. Jin and V. E. Korepin, Quantum spin chain, Toeplitz determinants and Fisher-Hartwig conjecture, J. Stat. Phys. 116, 79 (2004).
  • (113) P. Di Francesco, P. Mathieu, and D. Senechal, Conformal Field Theory, Springer-Verlag, New York, (1997).
  • (114) J. D. Fay, Theta functions on Riemann surfaces, Springer Notes in Mathematics 352 (Springer, 1973).
  • (115) E. Verlinde and H. Verlinde, Chiral Bosonization, Determinants and the String Partition Function, Nucl. Phys. B, 288, 357 (1987).
  • (116) S. Groha, F. H. L. Essler, and P. Calabrese, Full Counting Statistics in the Transverse Field Ising Chain, SciPost Phys. 4, 043 (2018).
  • (117) V. Enolskii and T. Grava, Singular N\mathbb{Z}_{N} curves, Riemann-Hilbert problem and modular solutions of the Schlesinger equation, Int. Math. Res. Not. 32, 1619 (2004).
  • (118) V. Enolskii and T. Grava, Thomae type formulae for singular N\mathbb{Z}_{N} curves, Lett. Math. Phys. 76, 187 (2006).
  • (119) I. Peschel, Calculation of reduced density matrices from correlation functions, J. Phys. A 36, L205 (2003).
  • (120) I. Peschel and V. Eisler, Reduced density matrices and entanglement entropy in free lattice models, J. Phys. A 42, 504003 (2009).
  • (121) M. Fagotti and P. Calabrese, Entanglement entropy of two disjoint blocks in XY chains, J. Stat. Mech. (2010) P04016.
  • (122) F. W. J. Olver, A. B. Olde Daalhuis, D. W. Lozier, B. I. Schneider, R. F. Boisvert, C. W. Clark, B. R. Miller, and B. V. Saunders (ed), NIST Digital Library of Mathematical Functions, (Gaithersburg: National Institute of Standards and Technology) (http://dlmf.nist.gov/).