Corrections to Conformal Charge at BKT Transitions

Jon Spalding1 [email protected] 1School of Physical and Mathematical Sciences, Nanyang Technological University, Singapore
(October 2, 2024)
Abstract

We present an analytical derivation of the lowest, 1stsuperscript1𝑠𝑡1^{st}1 start_POSTSUPERSCRIPT italic_s italic_t end_POSTSUPERSCRIPT-order corrections to the conformal charge for finite-sized systems at Berezinskii-Kosterlitz-Thouless transitions and compare the result with careful observations from the quantum Heisenberg model with periodic boundary conditions in one spatial dimension.

pacs:
Valid PACS appear here

I Introduction

One-dimensional quantum and two-dimensional classical critical theories exhibit conformal invariance, with many properties involving a universal quantity “c” known as the central charge or conformal anomaly (from the stress-energy tensor Virasoro algebra and stress-energy tensor correlation function, respectively). It appears as the response of a theory to changes in metric, provides the first universal finite-size correction term to the free energy for classical systems, and the specific heat capacity for quantum systems. It can also be interpreted as the number of bosonic fields in a theory [1, 2, 3].

Central charge gained attention as a coefficient on the universal logarithmic contribution to the bipartite von-Neumann entanglement entropy of ground-state (pure) wavefunctions in one spatial dimension [4, 5, 6] and renyi entropies [7] and at finite temperatures [8]. This paralleled the development of DMRG and MPS methods [9, 10, 11] which implicitly rely on entanglement entropy to minimize a variational wavefunction. These methods provide convenient access to high-precision measurements of entanglement with far greater ease than any other observable in an MPS or other numerical method such as QMC [12]. This feature motivates the numerical component of the current study.

BKT transitions were the first topological phase transitions discovered for interacting many-body systems, rewarded by a Nobel prize in 2016 [13, 14, 15].

More recently, analytical calculations based on conformal field theory provided the precise form of the entanglement entropy with open and periodic boundaries in infinite-size chains, before finite-size effects were observed numerically [16] and then treated analytically [17, 18]. For a finite-size chain with periodic boundaries,

SvN=S0+c3.subscript𝑆𝑣𝑁subscript𝑆0𝑐3.S_{vN}=S_{0}+\frac{c}{3}\ell\text{.}italic_S start_POSTSUBSCRIPT italic_v italic_N end_POSTSUBSCRIPT = italic_S start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + divide start_ARG italic_c end_ARG start_ARG 3 end_ARG roman_ℓ . (1)

Here we define log(L/πsin(πx/L))𝑙𝑜𝑔𝐿𝜋𝑠𝑖𝑛𝜋𝑥𝐿\ell\equiv log(L/\pi sin(\pi x/L))roman_ℓ ≡ italic_l italic_o italic_g ( italic_L / italic_π italic_s italic_i italic_n ( italic_π italic_x / italic_L ) ) as the length scale for periodic boundaries. We point out that the coefficient c simply indicates the relative concavity, or amplitude of the curvature, of a plot of the entanglement. In [17], it was determined that in the case of a marginally irrelevant operator (the case at a BKT transition), the central charge obtains a correction proportional to log()3𝑙𝑜𝑔superscript3log(\ell)^{-3}italic_l italic_o italic_g ( roman_ℓ ) start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT. As we show below, this term can be re-interpreted as a finite-size correction to c itself in addition to a contribution to the entanglement entropy at BKT transitions [19, 20, 16, 21].

If the exact form of this correction were known, including any coefficients, it would provide useful information at classical BKT transitions in two dimensions as well as quantum BKT transitions in one dimension. In the quantum case, it would enable precise and convenient identification of BKT transitions from numerical data [22, 23] although this is already possible knowing the general, though imprecise, nature of the corrections. That is, there is a well-defined peak in the central charge at BKT transitions as described and then implemented in [17, 22].

In this article, we present an analytical derivation of the finite-size correction to c at BKT transitions:

ceff()=1+R(y01+y0)3subscript𝑐eff1𝑅superscriptsubscript𝑦01subscript𝑦03c_{\rm eff}(\ell)=1+R\left(\frac{y_{0}}{1+y_{0}\ell}\right)^{3}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ( roman_ℓ ) = 1 + italic_R ( divide start_ARG italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 1 + italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_ℓ end_ARG ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT (2)

in which R=3/16𝑅316R=3/16italic_R = 3 / 16 is a universal constant, and y0=1subscript𝑦01y_{0}=1italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1 is the topological defect fugacity. For the remainder of the paper we use =log(L/π)𝑙𝑜𝑔𝐿𝜋\ell=log(L/\pi)roman_ℓ = italic_l italic_o italic_g ( italic_L / italic_π ) at the midpoint of the spin chain. This universally modifies equation 1 for BKT transitions by increasing the curvature of the entanglement. We follow this derivation with a careful numerical study of ground state entropy via exact diagonalization and DMRG for the XXZ model in one dimension with periodic boundary conditions.

II Analytical Study

We assume that, because they have to be consistent with Zamolodchikov’s c𝑐citalic_c-theorem, the leading log corrections to ceffsubscript𝑐effc_{\rm eff}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT are the same independent of whether they measure the finite-size scaling on a cylinder, the entanglement of an interval, or any other physical quantity [24, 25].

If we have a set of (almost) marginal couplings {gk}subscript𝑔𝑘\{g_{k}\}{ italic_g start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT } coupling to operators ΦksubscriptΦ𝑘\Phi_{k}roman_Φ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT, the RG equations in general take the form

g˙k=ykgk+πijcijkgigj+subscript˙𝑔𝑘subscript𝑦𝑘subscript𝑔𝑘𝜋subscript𝑖𝑗subscript𝑐𝑖𝑗𝑘subscript𝑔𝑖subscript𝑔𝑗\dot{g}_{k}=y_{k}g_{k}+\pi\sum_{ij}c_{ijk}g_{i}g_{j}+\cdotsover˙ start_ARG italic_g end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_g start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + italic_π ∑ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i italic_j italic_k end_POSTSUBSCRIPT italic_g start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_g start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + ⋯

where yksubscript𝑦𝑘y_{k}italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT is the RG eigenvalue at the fixed point and cijksubscript𝑐𝑖𝑗𝑘c_{ijk}italic_c start_POSTSUBSCRIPT italic_i italic_j italic_k end_POSTSUBSCRIPT is the coefficient in the OPE

ΦiΦj=kcijkΦk.subscriptΦ𝑖subscriptΦ𝑗subscript𝑘subscript𝑐𝑖𝑗𝑘subscriptΦ𝑘.\Phi_{i}\cdot\Phi_{j}=\sum_{k}c_{ijk}\Phi_{k}\text{.}roman_Φ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⋅ roman_Φ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i italic_j italic_k end_POSTSUBSCRIPT roman_Φ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT .

In the CFT normalization where the 2-point functions Φi(r)Φi(0)delimited-⟨⟩subscriptΦ𝑖𝑟subscriptΦ𝑖0\langle\Phi_{i}(r)\Phi_{i}(0)\rangle⟨ roman_Φ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_r ) roman_Φ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( 0 ) ⟩ are normalized to 1 at separation r=1𝑟1r=1italic_r = 1, the coefficients cijksubscript𝑐𝑖𝑗𝑘c_{ijk}italic_c start_POSTSUBSCRIPT italic_i italic_j italic_k end_POSTSUBSCRIPT are universal and symmetric in the indices.

To this order the RG equations are then gradient flows

g˙k=(/gk)C~({gi})subscript˙𝑔𝑘subscript𝑔𝑘~𝐶subscript𝑔𝑖\dot{g}_{k}=(\partial/\partial g_{k})\widetilde{C}(\{g_{i}\})over˙ start_ARG italic_g end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = ( ∂ / ∂ italic_g start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) over~ start_ARG italic_C end_ARG ( { italic_g start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } )

where

C~({gi})=12kykgk2+13πijkcijkgigjgk+~𝐶subscript𝑔𝑖12subscript𝑘subscript𝑦𝑘superscriptsubscript𝑔𝑘213𝜋subscript𝑖𝑗𝑘subscript𝑐𝑖𝑗𝑘subscript𝑔𝑖subscript𝑔𝑗subscript𝑔𝑘\widetilde{C}(\{g_{i}\})=\frac{1}{2}\sum_{k}y_{k}g_{k}^{2}+\frac{1}{3}\pi\sum_% {ijk}c_{ijk}g_{i}g_{j}g_{k}+\cdotsover~ start_ARG italic_C end_ARG ( { italic_g start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_y start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_g start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG 1 end_ARG start_ARG 3 end_ARG italic_π ∑ start_POSTSUBSCRIPT italic_i italic_j italic_k end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i italic_j italic_k end_POSTSUBSCRIPT italic_g start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_g start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_g start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT + ⋯

In terms of this Zamolodchikov’s c𝑐citalic_c-function is

C({gi})=c6π2C~({gi})+.𝐶subscript𝑔𝑖𝑐6superscript𝜋2~𝐶subscript𝑔𝑖.C(\{g_{i}\})=c-6\pi^{2}\widetilde{C}(\{g_{i}\})+\cdots\text{.}italic_C ( { italic_g start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } ) = italic_c - 6 italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over~ start_ARG italic_C end_ARG ( { italic_g start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } ) + ⋯ .

Now specialise to the BKT transition (or any model which maps onto it in the long-distance limit). There are two marginal operators. The CFT is defined by the gaussian model action

S=g4π(ϕ)2d2r𝑆𝑔4𝜋superscriptitalic-ϕ2superscript𝑑2𝑟S=\frac{g}{4\pi}\int(\partial\phi)^{2}d^{2}ritalic_S = divide start_ARG italic_g end_ARG start_ARG 4 italic_π end_ARG ∫ ( ∂ italic_ϕ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_r

One marginal operator is (ϕ)2proportional-toabsentsuperscriptitalic-ϕ2\propto(\partial\phi)^{2}∝ ( ∂ italic_ϕ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The others are the vortex (anti-vortex) operators V𝑉Vitalic_V and V¯¯𝑉\bar{V}over¯ start_ARG italic_V end_ARG with corresponding fugacity y𝑦yitalic_y.

The scaling dimension is g/2𝑔2g/2italic_g / 2, so the RG equation for y𝑦yitalic_y is

y˙=(2g/2)y.˙𝑦2𝑔2𝑦.\dot{y}=(2-g/2)y\text{.}over˙ start_ARG italic_y end_ARG = ( 2 - italic_g / 2 ) italic_y .

The BKT point is g=4𝑔4g=4italic_g = 4.

We need to get the right normalization for (ϕ)2superscriptitalic-ϕ2(\partial\phi)^{2}( ∂ italic_ϕ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Using

ϕ(r1)ϕ(r2)=(1/g)log|r1r2|+const.delimited-⟨⟩italic-ϕsubscript𝑟1italic-ϕsubscript𝑟21𝑔subscript𝑟1subscript𝑟2const\langle\phi(r_{1})\phi(r_{2})\rangle=-(1/g)\log|r_{1}-r_{2}|+{\rm const.}⟨ italic_ϕ ( italic_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) italic_ϕ ( italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ⟩ = - ( 1 / italic_g ) roman_log | italic_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | + roman_const .

and Wick’s theorem,

(ϕ(r1))2(ϕ(r2))2=delimited-⟨⟩superscriptitalic-ϕsubscript𝑟12superscriptitalic-ϕsubscript𝑟22absent\langle(\partial\phi(r_{1}))^{2}(\partial\phi(r_{2}))^{2}\rangle=⟨ ( ∂ italic_ϕ ( italic_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( ∂ italic_ϕ ( italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ =
(2/g2)(1μ2νlog|r1r2|)(1μ2νlog|r1r2|)2superscript𝑔2subscriptsuperscript𝜇1subscriptsuperscript𝜈2subscript𝑟1subscript𝑟2subscriptsuperscript𝜇1subscriptsuperscript𝜈2subscript𝑟1subscript𝑟2(2/g^{2})(\partial^{\mu}_{1}\partial^{\nu}_{2}\log|r_{1}-r_{2}|)(\partial^{\mu% }_{1}\partial^{\nu}_{2}\log|r_{1}-r_{2}|)( 2 / italic_g start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( ∂ start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∂ start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_log | italic_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | ) ( ∂ start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∂ start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_log | italic_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | )
=(2/g2)(δμνr22rμrνr3)(δμνr22rμrνr3)=4/g2r4absent2superscript𝑔2superscript𝛿𝜇𝜈superscript𝑟22superscript𝑟𝜇superscript𝑟𝜈superscript𝑟3superscript𝛿𝜇𝜈superscript𝑟22superscript𝑟𝜇superscript𝑟𝜈superscript𝑟34superscript𝑔2superscript𝑟4=(2/g^{2})\left(\frac{\delta^{\mu\nu}}{r^{2}}-\frac{2r^{\mu}r^{\nu}}{r^{3}}% \right)\left(\frac{\delta^{\mu\nu}}{r^{2}}-\frac{2r^{\mu}r^{\nu}}{r^{3}}\right% )=\frac{4/g^{2}}{r^{4}}= ( 2 / italic_g start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( divide start_ARG italic_δ start_POSTSUPERSCRIPT italic_μ italic_ν end_POSTSUPERSCRIPT end_ARG start_ARG italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - divide start_ARG 2 italic_r start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT italic_r start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT end_ARG start_ARG italic_r start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG ) ( divide start_ARG italic_δ start_POSTSUPERSCRIPT italic_μ italic_ν end_POSTSUPERSCRIPT end_ARG start_ARG italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - divide start_ARG 2 italic_r start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT italic_r start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT end_ARG start_ARG italic_r start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG ) = divide start_ARG 4 / italic_g start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_r start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT end_ARG

(The extra 2 comes from the two different Wick contractions.) Thus at the BKT point the correctly normalized operator is Φ=2(ϕ)2Φ2superscriptitalic-ϕ2\Phi=2(\partial\phi)^{2}roman_Φ = 2 ( ∂ italic_ϕ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

The perturbation of the action is

δg4π(ϕ)2=xΦ𝛿𝑔4𝜋superscriptitalic-ϕ2𝑥Φ\frac{\delta g}{4\pi}(\partial\phi)^{2}=x\Phidivide start_ARG italic_δ italic_g end_ARG start_ARG 4 italic_π end_ARG ( ∂ italic_ϕ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_x roman_Φ

where x=δg/(8π)𝑥𝛿𝑔8𝜋x=\delta g/(8\pi)italic_x = italic_δ italic_g / ( 8 italic_π ). In terms of this quantity the RG equation for y𝑦yitalic_y becomes

y˙=4πxy˙𝑦4𝜋𝑥𝑦\dot{y}=-4\pi xyover˙ start_ARG italic_y end_ARG = - 4 italic_π italic_x italic_y

Thus

C~=2πxy2+~𝐶2𝜋𝑥superscript𝑦2\widetilde{C}=-2\pi xy^{2}+\cdotsover~ start_ARG italic_C end_ARG = - 2 italic_π italic_x italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ⋯

and the other RG equation is therefore

x˙=2πy2˙𝑥2𝜋superscript𝑦2\dot{x}=-2\pi y^{2}over˙ start_ARG italic_x end_ARG = - 2 italic_π italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT

Note that we have avoided working out the correct normalization for V𝑉Vitalic_V.

The RG trajectory corresponding to the BKT transition line is found by setting y=αx𝑦𝛼𝑥y=\alpha xitalic_y = italic_α italic_x, so, dividing the two equations we have α=2/α𝛼2𝛼\alpha=2/\alphaitalic_α = 2 / italic_α, that is α=2𝛼2\alpha=\sqrt{2}italic_α = square-root start_ARG 2 end_ARG.

Refer to caption
Figure 1: The RG flow is towards the BKT fixed point.

Along this trajectory,

y˙=22πy2˙𝑦22𝜋superscript𝑦2\dot{y}=-2\sqrt{2}\pi y^{2}over˙ start_ARG italic_y end_ARG = - 2 square-root start_ARG 2 end_ARG italic_π italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT

so that

y()=y01+22πy0𝑦subscript𝑦0122𝜋subscript𝑦0y(\ell)=\frac{y_{0}}{1+2\sqrt{2}\pi y_{0}\ell}italic_y ( roman_ℓ ) = divide start_ARG italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 1 + 2 square-root start_ARG 2 end_ARG italic_π italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_ℓ end_ARG

where \ellroman_ℓ is the RG scale, logLsimilar-toabsent𝐿\sim\log L∼ roman_log italic_L for the entanglement problem.

Along the transition line

C~=2πy3~𝐶2𝜋superscript𝑦3\widetilde{C}=-\sqrt{2}\pi y^{3}over~ start_ARG italic_C end_ARG = - square-root start_ARG 2 end_ARG italic_π italic_y start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT

so the c-function becomes

ceff()=1+(6π2)(2π)(y01+22πy0)3subscript𝑐eff16superscript𝜋22𝜋superscriptsubscript𝑦0122𝜋subscript𝑦03c_{\rm eff}(\ell)=1+(6\pi^{2})(\sqrt{2}\pi)\left(\frac{y_{0}}{1+2\sqrt{2}\pi y% _{0}\ell}\right)^{3}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ( roman_ℓ ) = 1 + ( 6 italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( square-root start_ARG 2 end_ARG italic_π ) ( divide start_ARG italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 1 + 2 square-root start_ARG 2 end_ARG italic_π italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_ℓ end_ARG ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT

Since y0subscript𝑦0y_{0}italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (which is like the bare vortex fugacity) is arbitrary, we can rescale 22πy0y022𝜋subscript𝑦0subscript𝑦02\sqrt{2}\pi y_{0}\to y_{0}2 square-root start_ARG 2 end_ARG italic_π italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT → italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT in which case this becomes

ceff()=1+38(y01+y0)3subscript𝑐eff138superscriptsubscript𝑦01subscript𝑦03c_{\rm eff}(\ell)=1+\frac{3}{8}\left(\frac{y_{0}}{1+y_{0}\ell}\right)^{3}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ( roman_ℓ ) = 1 + divide start_ARG 3 end_ARG start_ARG 8 end_ARG ( divide start_ARG italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 1 + italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_ℓ end_ARG ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT

To apply this to the entanglement problem for a single interval, we should recall that in the pure CFT

S(L)(c/3)similar-to𝑆𝐿𝑐3S(L)\sim(c/3)\ellitalic_S ( italic_L ) ∼ ( italic_c / 3 ) roman_ℓ

Here, we use the results from Casini and Huerta [20] that implies we may interpret that S𝑆\frac{\partial S}{\partial\ell}divide start_ARG ∂ italic_S end_ARG start_ARG ∂ roman_ℓ end_ARG obeys Zamolodchikov’s c-theorem, that is,

ceff(l)=3S.subscript𝑐effl3𝑆.c_{\rm eff}(\textit{l})=3\frac{\partial S}{\partial\ell}\text{.}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ( l ) = 3 divide start_ARG ∂ italic_S end_ARG start_ARG ∂ roman_ℓ end_ARG .

Then by integrating we get the next higher correction to the entanglement entropy,

S(L)(1/3)logL116y02(1+y0logL)2+.similar-to𝑆𝐿13𝐿116superscriptsubscript𝑦02superscript1subscript𝑦0𝐿2.S(L)\sim(1/3)\log L-\frac{1}{16}\frac{y_{0}^{2}}{(1+y_{0}\log L)^{2}}+\cdots% \text{.}italic_S ( italic_L ) ∼ ( 1 / 3 ) roman_log italic_L - divide start_ARG 1 end_ARG start_ARG 16 end_ARG divide start_ARG italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ( 1 + italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_log italic_L ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + ⋯ .

Note that the value R=3/8𝑅38R=3/8italic_R = 3 / 8 contradicts our curve fit result, 3/163163/163 / 16 below.

III Numerical Study

In this section, we present numerical investigations that confirm the universal, analytical derivation in the prior section. Because the analytical corrections to c are universal at all finite-size BKT phase transitions, we choose a convenient model and method. That is, we extract c from entanglement entropy in the antiferromagnetic Heisenberg model in one spatial dimension with the DMRG method applied to periodic boundary conditions.

The antiferromagnetic Heisenberg model can be thought of as the BKT transition separating a disordered critical phase from a gapped ordered phase at the isotropic point, Δ=1Δ1\Delta=1roman_Δ = 1, in the spin-1/2 XXZ model:

H=isixsi+1x+siysi+1y+Δsizsi+1z.𝐻subscript𝑖subscriptsuperscript𝑠𝑥𝑖subscriptsuperscript𝑠𝑥𝑖1subscriptsuperscript𝑠𝑦𝑖subscriptsuperscript𝑠𝑦𝑖1Δsubscriptsuperscript𝑠𝑧𝑖subscriptsuperscript𝑠𝑧𝑖1.H=\sum_{i}s^{x}_{i}s^{x}_{i+1}+s^{y}_{i}s^{y}_{i+1}+\Delta s^{z}_{i}s^{z}_{i+1% }\text{.}italic_H = ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT italic_x end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT + italic_s start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT + roman_Δ italic_s start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT italic_z end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT . (3)

The entanglement entropy is computed using the Schmidt decomposition between each pair of of sites:

Sj=i,jλi,j2log(λi,j2).subscript𝑆𝑗subscript𝑖𝑗superscriptsubscript𝜆𝑖𝑗2superscriptsubscript𝜆𝑖𝑗2.S_{j}=-\sum_{i,j}\lambda_{i,j}^{2}\log{(\lambda_{i,j}^{2})}\text{.}italic_S start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = - ∑ start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_log ( italic_λ start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) . (4)

This is very convenient in DMRG because the algorithm uses the Schmidt values for performing the truncation scheme, and so the Schmidt values are readily available. Note that the raw data for this project consisted of the spatially-dependent entropy, Sjsubscript𝑆𝑗S_{j}italic_S start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, at each pair of lattice sites, (j,j+1). From this data, we evaluate the derivative with respect to j at the midpoint j = L/2.

Once we have central charge for each system size we performed both nonlinear and linear regressions to fit to the analysis derived above. The curve fitting parameters include the bare topological defect fugacity, y0subscript𝑦0y_{0}italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, and a universal coefficient that we arbitrarily call “R”:

ceff()=1+R(y01+y0)3.subscript𝑐eff1𝑅superscriptsubscript𝑦01subscript𝑦03.c_{\rm eff}(\ell)=1+R\left(\frac{y_{0}}{1+y_{0}\ell}\right)^{3}\text{.}italic_c start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ( roman_ℓ ) = 1 + italic_R ( divide start_ARG italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 1 + italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_ℓ end_ARG ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT . (5)

As mentioned, we are restricted to periodic boundary conditions which are known to drastically reduce the accessible system sizes for the DMRG algorithm. As a result, we have reliable data for systems up to length 120 sites. For sizes under 32 sites, we used exact diagonalization to provide a check against the convergence accuracy of the DMRG results. We include those results in our final results.

The first data we present is a plot of the central charge vs. system size, along with a nonlinear curve fit to extract the two fitting parameters of interest (fig. 2).

Refer to caption
Figure 2: Curve fit of ceffsubscript𝑐𝑒𝑓𝑓c_{eff}italic_c start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT in equation (5) at chain midpoint for all sizes available.

One of the conclusions drawn from that plot is that plotting and fitting functions with logarithmic dependencies can be tricky, so we also found a linearized version as follows:

C=M(Y+)𝐶𝑀𝑌C=M(Y+\ell)italic_C = italic_M ( italic_Y + roman_ℓ ) (6)

where \ellroman_ℓ is the independent variable,

log(Lπsin(πx/L))𝐿𝜋𝜋𝑥𝐿\ell\equiv\log\left(\frac{L}{\pi}\sin{(\pi x/L)}\right)roman_ℓ ≡ roman_log ( divide start_ARG italic_L end_ARG start_ARG italic_π end_ARG roman_sin ( italic_π italic_x / italic_L ) )

which is simply log(L/π)𝐿𝜋\log{(L/\pi)}roman_log ( italic_L / italic_π ) at the chain midpoint. Next, we define

Y1y0𝑌1subscript𝑦0Y\equiv\frac{1}{y_{0}}italic_Y ≡ divide start_ARG 1 end_ARG start_ARG italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG

and

M1R1/3𝑀1superscript𝑅13M\equiv\frac{1}{R^{1/3}}italic_M ≡ divide start_ARG 1 end_ARG start_ARG italic_R start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT end_ARG

so that the dependent variable is

C1(c(x,L)1)1/3.𝐶1superscript𝑐𝑥𝐿113.C\equiv\frac{1}{(c(x,L)-1)^{1/3}}\text{.}italic_C ≡ divide start_ARG 1 end_ARG start_ARG ( italic_c ( italic_x , italic_L ) - 1 ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT end_ARG . (7)

With these new variables, figure 2 becomes 3. In this form it is clear that we are using the correct fitting function because the data closely matches the predictions, including a nonzero intercept at =00\ell=0roman_ℓ = 0.

Refer to caption
Figure 3: Linear fit of ceffsubscript𝑐𝑒𝑓𝑓c_{eff}italic_c start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT transformed to the linear form, equation (6).

In the appendix, we provide some additional investigations regarding these logarithmic, nonlinear regressions. We also provide data on the relative error in the parameter R for both large and small sizes L when the DMRG truncation error is taken into account.

IV Conclusions

In this paper we have provided an elegant derivation that redefines the central charge as the first derivative of entropy with respect to logarithm of length scale. As such, this “effective” finite-size corrected value of c will apply equally to classical and quantum critical systems with a conformal description. Also, we have only considered corrections to c due to finite size effects, but other parameters (fields, temperature, truncation error in finite and infinite MPS) should also follow.

The case of spin chains with open boundaries provides another opportunity to study the finite-size corrections to central charge by analytical arguments as we have done here for periodic boundaries. Our prior work [23] provides a starting point to these studies, including methods to deal with bond-alternation in the entropy as a function of position and strategies to extrapolate curve fits to the middle of the lattice, named “scaling to the middle.” Note that equation 1 is modified for open boundaries by replacing the 3 with a 6 and multiplying L by 2 within the logarithm [23].

Although our results presented here offer reliable measurements for central charge for the system sizes studied, it is possible that the asymptotic value of ceffsubscript𝑐𝑒𝑓𝑓c_{eff}italic_c start_POSTSUBSCRIPT italic_e italic_f italic_f end_POSTSUBSCRIPT for larger sizes may vary as indicated by figures 5 and 6 when both R and y0subscript𝑦0y_{0}italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT are allowed to vary. If these trends hold for larger size spin chains, additional terms in the operator product expansion may be needed to account for higher-order finite-size effects.

In addition, allowing y0subscript𝑦0y_{0}italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT to vary may offer a generic, indirect method of observing topological defect properties in quantum spin chains from entanglement entropy. We have effectively demonstrated that topological defects contribute to the entanglement entropy via the central charge for finite-size systems.

Extending these approaches to other defect-driven phase transitions is one possible application of this work.

V Acknowledgements

We thank John Cardy for the BKT analysis presented here. This research was supported in part by the NSF under grant DMR-1411345 and by University of California, Riverside’s GRMP fellowship. This work used the Extreme Science and Engineering Discovery Environment (XSEDE) COMET at the San Diego Supercomputer Center through allocation TG-DMR170082 [26] as well as University of California Riverside’s High Perfomance Computing Center.

VI Appendix

VI.1 Study of size-dependence of curve fit parameters

Refer to caption
Figure 4: Plot of the curve fit parameter, R, for each possible set of 4 neighboring data points as a function of system size, including very small and very large. Fugacity y0subscript𝑦0y_{0}italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is fixed to 1.
Refer to caption
Figure 5: Plot of the curve fit parameter, R, for each possible set of 4 neighboring data points as a function of system size. R is clearly increasing as system size increases. Fugacity y0subscript𝑦0y_{0}italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is allowed to vary. From this perspective it appears R might asymptotically increase to the value 3/8.
Refer to caption
Figure 6: y0subscript𝑦0y_{0}italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT plotted as a function of system size. Like for coefficient R plotted above, y0subscript𝑦0y_{0}italic_y start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT has poor results for sizes less than 20 and greater than 120, so in this plot we restrict to the smooth portion of the data.

VI.2 Study of Truncation Effects in Curve Fit Results

Refer to caption
Figure 7: Plot of coefficient R vs the DMRG truncation parameter for system sizes 24 to 30 plotted with the exact diagonalization result. The error in R is at the 5thsuperscript5𝑡5^{th}5 start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT decimal place.
Refer to caption
Figure 8: Similar to figure 7 but for system sizes 112 to 118. The error in R for our data is approximately the difference between the first two data points in this figure, which is less than 0.001, but still discernible in this plot. It is noteworthy that the same truncation error in DMRG results in a larger error in R when the system size is increased; error increases by factor of 100.

References

  • Francesco et al. [1997] P. D. Francesco, P. Mathieu, and D. Sénéchal, Conformal Field Theory (Springer, 1997).
  • Blöte et al. [1986] H. W. J. Blöte, J. L. Cardy, and M. P. Nightingale, Phys. Rev. Lett. 56, 742 (1986).
  • Affleck [1986] I. Affleck, Phys. Rev. Lett. 56, 746 (1986).
  • Holzhey et al. [1994] C. Holzhey, F. Larsen, and F. Wilczek, Nuclear Physics B 424, 443 (1994).
  • Vidal et al. [2003] G. Vidal, J. I. Latorre, E. Rico, and A. Kitaev, Phys. Rev. Lett. 90, 227902 (2003).
  • Korepin [2004a] V. E. Korepin, Phys. Rev. Lett. 92, 096402 (2004a).
  • Jin and Korepin [2004] B.-Q. Jin and V. E. Korepin, Journal of statistical physics 116, 79 (2004).
  • Korepin [2004b] V. E. Korepin, Physical review letters 92, 096402 (2004b).
  • White [1992] S. R. White, Physical review letters 69, 2863 (1992).
  • White [1993] S. R. White, Physical Review B 48, 10345 (1993).
  • Schollwöck [2011] U. Schollwöck, Annals of Physics 326, 96 (2011), january 2011 Special Issue.
  • Hastings et al. [2010] M. B. Hastings, I. González, A. B. Kallin, and R. G. Melko, Physical review letters 104, 157201 (2010).
  • Kosterlitz and Thouless [2018] J. M. Kosterlitz and D. J. Thouless, in Basic Notions Of Condensed Matter Physics (CRC Press, 2018) pp. 493–515.
  • Berezinskii [1971] V. Berezinskii, Sov. Phys. JETP 32, 493 (1971).
  • Berezinskii [1972] V. Berezinskii, Sov. Phys. JETP 34, 610 (1972).
  • Laflorencie et al. [2006] N. Laflorencie, E. S. Sørensen, M.-S. Chang, and I. Affleck, Phys. Rev. Lett. 96, 100603 (2006).
  • Cardy and Calabrese [2010] J. Cardy and P. Calabrese, Journal of Statistical Mechanics: Theory and Experiment 2010, P04023 (2010).
  • Fagotti and Calabrese [2011] M. Fagotti and P. Calabrese, Journal of Statistical Mechanics: Theory and Experiment 2011, P01017 (2011).
  • Calabrese and Cardy [2009] P. Calabrese and J. Cardy, Journal of Physics A: Mathematical and Theoretical 42, 504005 (2009).
  • Casini and Huerta [2004] H. Casini and M. Huerta, Physics Letters B 600, 142 (2004).
  • Calabrese et al. [2010] P. Calabrese, M. Campostrini, F. Essler, and B. Nienhuis, Phys. Rev. Lett. 104, 095701 (2010).
  • Nishimoto [2011] S. Nishimoto, Phys. Rev. B 84, 195108 (2011).
  • Spalding et al. [2019] J. Spalding, S.-W. Tsai, and D. K. Campbell, Phys. Rev. B 99, 195445 (2019).
  • Zamolodchikov [1986] A. B. Zamolodchikov, JETP Lett. 43, 731 (1986).
  • Cardy [1988] J. L. Cardy, Les Houches  (1988).
  • Towns et al. [2014] J. Towns, T. Cockerill, M. Dahan, I. Foster, K. Gaither, A. Grimshaw, V. Hazlewood, S. Lathrop, D. Lifka, G. D. Peterson, R. Roskies, J. R. Scott, and N. Wilkins-Diehr, Computing in Science & Engineering 16, 62 (2014).