Approaching the Kosterlitz-Thouless transition for the classical XY model with tensor networks

Laurens Vanderstraeten,1, ∗ Bram Vanhecke,1 Andreas M. L¨auchli,2 and Frank Verstraete1 1Department of Physics and Astronomy, University of Ghent, Krijgslaan 281, 9000 Gent, Belgium 2Institute for Theoretical Physics, University of Innsbruck, 6020 Innsbruck, Austria We apply variational tensor-network methods for simulating the Kosterlitz-Thouless phase transi- tion in the classical two-dimensional XY model. In particular, using uniform matrix product states (MPS) with non-abelian O(2) symmetry, we compute the universal drop in the spin stiffness at the critical point. In the critical low-temperature regime, we focus on the MPS entanglement spectrum to characterize the Luttinger-liquid phase. In the high-temperature phase, we confirm the exponen- tial divergence of the correlation length and estimate the critical temperature with high precision. Our MPS approach can be used to study generic two-dimensional phase transitions with continuous symmetries.

I. INTRODUCTION transition (in the absence of a local order parameter) is the spin stiffness ρ(T ), which discontinuously drops [7] In contemporary theoretical physics the interplay be- from a finite value in the critical phase to zero in the tween symmetry and dimensionality is widely appre- gapped phase; from the renormalization analysis, the size ciated for giving rise to fascinating physical phenom- of the drop is known to be given by ena. Historically, one of the crucial results in estab- 2Tc lishing this viewpoint was the in the lim ρ(T ) = , (1) − π two-dimensional classical XY model. This model is in- T →Tc troduced by placing continuous angles {θi} on a square but the value of Tc is not known exactly. When ap- lattice and characterizing their interactions by the hamil- proaching the critical point from the gapped side, the tonian BKT transition is characterized by an exponential diver- X X gence of the correlation length [5,8], H = − cos (θi − θj) − h cos(θi),   hiji i b + ξ(T ) ∝ exp √ ,T → Tc , (2) T − Tc where hiji labels all nearest-neighbour pairs and we in- troduce an external magnetic field h for further reference. with b a non-universal parameter. The first ingredient for understanding the phase dia- The BKT phase transition in the XY model has been gram of the XY model (without magnetic field) is the a notoriously hard case for numerical simulations, be- Mermin-Wagner theorem [1,2]: No conventional long- cause of the exponentially diverging correlation length range order can exist at any finite temperature in this and the ensuing logarithmic finite-size corrections around model, because of the proliferation of spin-wave exci- the phase transition. Nonetheless, shortly after the the- tations in two dimensions. Still, one expects a phase oretical work, Monte-Carlo simulations [9, 10] provided transition in this system. At low temperatures, the sys- considerable evidence for the correctness of the theory tem can be described by a simple continuum field the- and rough estimates for the transition temperature. In ory with algebraically decaying correlation functions. A recent years, these estimates were significantly improved high-temperature expansion, however, suggests an ex- [11, 12], obtaining a value around Tc ≈ 0.8929, in agree- ponential decay of correlations in this system at suffi- ment with high-temperature expansions [13]. Still, these ciently high temperature. As Berezinskii [3] and Koster- results depend heavily on assumptions about the loga- litz and Thouless [4] (BKT) have showed, the phase tran- rithmic finite-size corrections and an improved extrapo- sition between the low- and high-temperature regime is lation [14] places the transition temperature at a higher driven by the unbinding of vortices and is, therefore, value of Tc ≈ 0.8935. Crucially, the best estimates use of a topological nature. The transition is captured by the universal value for the spin stiffness at the phase tran- a renormalization-group (RG) analysis [5,6], where in sition for pinpointing the critical point. arXiv:1907.04576v1 [cond-mat.stat-mech] 10 Jul 2019 the critical phase the long-wavelength properties are de- Tensor networks [15, 16] provide an original frame- scribed by a free bosonic field theory with continuously work for capturing the symmetry-dimensionality inter- varying critical exponents, until vortices become relevant play, both from the theoretical and the numerical side. and drive the system into a gapped phase at a criti- Although orginally devised for capturing the entangle- cal temperature Tc. The quantity that characterizes the ment in strongly-correlated quantum lattice systems, ten- sor networks are increasingly being applied to problems in . Since the framework is entirely different from traditional approaches such as Monte- ∗ [email protected] Carlo sampling, it can shed a new, entanglement-based, 2 light on statistical-mechanics problems. Indeed, tensor II. MPS FOR THE XY TRANSFER MATRIX networks encode all physical properties of a given system into local tensors, and they allow to understand the re- A. Partition function lation between physical symmetries of the local degrees of freedom and the global properties of the system in a We start by writing down the partition function for the transparent way. One particular example is the classifi- XY model as a tensor network. The partition function cation of symmetry-protected topological phases in one- at a given inverse temperature β = 1/T is given by dimensional quantum systems, which is brought back to the symmetry properties of the local tensors that make Z Y dθi Y Y Z = eβ cos(θi−θj ) eβh cos(θi). up a matrix product state (MPS) [17, 18]. This was ex- 2π tended to two dimensions, where the topological order i hiji i of a wavefunction can be related to the symmetry prop- erties of the local tensor in a projected entangled-pair In order to arrive at a tensor-network representation, state (PEPS) [19]. On the numerical side, physical sym- we introduce a duality transformation [6] that maps the metries can be explicitly incorporated in tensor-network above partition function to a representation in terms of algorithms and lead to an improved efficiency and per- bosonic degrees of freedom on the links. Such a map is formance [20–23]. obtained by introducing the following decomposition on every link in the lattice

The encoding of symmetries in tensor networks is per- N x cos(θi−θj ) X in(θi−θj ) e = lim In(x)e , formed most elegantly when working directly in the ther- N→∞ modynamic limit, because one can purely focus on the n=−N symmetry properties of the bulk tensors without bother- where I (x) are the modified Bessel functions of the first ing about what happens at the boundaries of the system. n kind. Then, by integrating over all the θ’s, the partition Yet, applying uniform tensor networks with an explicit function is transformed into encoding of the physical symmetries seem to break down when considering critical phases. For example, in a criti- N ! Y X Y cal phase a uniform MPS [24] typically favours an artifi- ns,3,ns,4 Z = lim Inl (β) Fn ,n N→∞ s,1 s,2 cial breaking of a continuous symmetry, where the asso- l∈L nl=−N s ciated order parameter decreases very slowly as the bond dimension is increased. The apparent reason for this ar- where F is a four-index tensor tificial symmetry breaking is that MPS have a built-in Z dθ limitation for the amount of entanglement in the state, F n3,n4 = eβh cos θeiθ(n1+n2−n3−n4). n1,n2 which makes it energetically favourable to break a con- 2π tinuous symmetry. This seems to imply that uniform The first product runs over all the links in the lattice, and tensor networks fail to capture the essential properties of s labels all the sites in the lattice. We can now represent critical phases with a continuous symmetry. this partition function as a network of tensors,

In this paper, we explore this question in more de- tail by investigating the precise sense in which uniform O O O O MPS capture critical phases with a continuous symme- try. As explained, the XY model serves as the paradig- O O O O matic example of a system where the absence of symme- Z = , try breaking leads to a critical phase, and, therefore, we O O O O take the XY model as our test case. In contrast to ear- lier tensor-network approaches for the XY model [25], we O O O O use uniform MPS methods for transfer matrices [26–28] as a means for characterizing the BKT phase transition. In the first two sections we explain the duality transfor- where every tensor O is given by mation [6] that allows us to define a row-to-row transfer matrix, approximate its fixed point as a uniform MPS n1 4 !1/2 and to compute local observables. In the next section, Y n3,n4 n4 O n2 = In (β) F we focus on the spin stiffness as the characteristic quan- i n1,n2 tity in the BKT phase transition. Afterwards, we use i=1 n3 the Luttinger-liquid formalism to characterize the criti- cal phase. Finally, in the last section, we focus on the and the virtual legs ni have infinite dimension. In prac- gapped phase and locate the critical temperature with tice, however, it will be possible to truncate these indices high precision. without loss of accuracy. We have introduced arrows on 3 the legs to indicate the signs in which the ni’s appear in establishing an optimization problem for the tensor A: the F tensor above. The fundamental object in this representation of the hΨ(A¯)| T (β, h) |Ψ(A)i max . partition function is the row-to-row transfer matrix A hΨ(A¯)|Ψ(A)i T (β, h) This optimization problem can be efficiently solved us- ing tangent-space methods for uniform MPS [26, 28]; in T (β, h) = O O O O , particular, we use the vumps algorithm [27] for finding the optimal MPS tensor. The eigenvalue, and therefore which is an operator acting on an infinite chain of bosonic the free energy, is then obtained as the contraction of degrees of freedom. The value for the partition function an infinite channel of O tensors sandwiched between the and, therefore, the free energy of the model, is deter- fixed-point MPS and its conjugate, mined by the leading eigenvalue Λ of the transfer matrix. ¯ Indeed, this leading eigenvalue is expected to scale as the Λ = hΨ(A)| T (β, h) |Ψ(A)i N number of sites per row, i.e. Λ ∼ λ x , such that the free A A A A energy per site is = O O O O ,  1 1  f(β, h) = lim − log Z ¯ ¯ ¯ ¯ NxNy →∞ β NxNy A A A A  1 1  = lim − log T (β, h)Ny  or we find λ as the leading eigenvalue (spectral radius ρ) NxNy →∞ β NxNy of the channel operator 1 = − log λ(β, h). β  A    The eigenvector corresponding to the leading eigen- λ = ρ  O  . value, refered to as the fixed point of the transfer matrix,   A¯ T (β, h) |Ψβ,hi = Λ |Ψβ,hi , Here, we have assumed that the MPS itself is normalized will be of crucial importance in all computations. In as a number of applications, it has been shown that fixed points of transfer matrices can be approximated accu- A ! rately using the variational class of matrix product states ρ = 1. (MPS) [26]. For translation-invariant transfer matrices, A¯ we can describe the fixed point as a uniform MPS de- scribed by a single tensor A, In Fig.1 we have plotted the result for the free energy per site as obtained by a variational MPS simulation of the XY model in a certain temperature range and without |Ψ(A)i = A A A A . magnetic field, where we use MPS with a bond dimension of D = 30. It is clear that the free energy shows no signs of a phase transition as it is perfectly smooth everywhere. The tensor A has virtual legs of dimension D, which we Note that, whereas the free energy cannot be directly call the MPS bond dimension, and by repeating the ten- computed using Monte-Carlo sampling, it appears as the sor on every site and contracting over the virtual legs, fundamental quantity in the variational MPS setup. we obtain a translation-invariant state. For simplicity, we take the virtual legs of the MPS to have no arrows. In terms of the fixed-point MPS, the fixed-point eigen- value equation is rephrased as B. Symmetries in the MPS representation

A A A A The case without magnetic field (h = 0) is of partic- ∝ ular importance, because the model is invariant under O O O O a global transformation θi → θi + α. This U(1) invari- ance is reflected in the tensor-network representation as a symmetry of the transfer matrix. Indeed, if h = 0 we . have Z n ,n dθ iθ(n +n −n −n ) We aim at finding a tensor for which this eigenvalue prob- F 3 4 = e 1 2 3 4 n1,n2 h=0 2π lem is obeyed in an optimal way. Since the transfer ma- = δn3+n4 . trix is hermitian, we can use the variational principle for n1+n2 4

a symmetry property of the local tensor A [29]. Specif- ically, we have that the virtual legs of the MPS tensor transform themselves according to representations of the U(1) symmetry,

A † = vθ A vθ . (5) uθ

In general, these representations on the virtual level can be projective. Imposing that the MPS is invariant under group transformations implies that the tensor A has a certain block structure, where each block can be labeled by a choice of representations on each leg of the tensor. FIG. 1. Free energy per site of the XY model (h = 0) as ob- Although the characteristic physics of the XY model is tained from variational MPS simulations with bond dimension determined by the U(1) symmetry, we can also take into D = 30 and without using any symmetries on the virtual legs. account the charge-conjugation symmetry C that flips the sign of the charges in the transfer matrix. The to- tal symmetry group that we consider is the (non-abelian) group U(1) C, which corresponds to the full O(2) sym- such that the tensor O has conservation of U(1) charges. o metry of the XY model. There are two types of ir- We can introduce the operator Q reducible representations (irreps) of this O(2) symme- n1 try. First, there are the irreps corresponding to inte- ger charges n = 0, 1, 2,... ; for n = 0 there are two Q = δn2 n , (3) n1 1 one-dimensional irreps, whereas for n > 0 they are all n two-dimensional. Second, there are half-integer charges 2 1 3 n = 2 , 2 ,... , corresponding to projective representa- which counts the charge on a given leg in the tensor net- tions; these irreps are all two-dimensional. work; this operator is the generator of the U(1) symme- Since the transfer matrix only contains legs with in- try of the transfer matrix. Indeed, from the conservation teger representations, the ‘physical legs’ of the MPS all property for the tensor O the transfer matrix clearly com- transform according to integer representations as well. mutes with the symmetry operation This, however, still allows for the freedom on the virtual iθ P Q level of the MPS to impose either integer representations U(θ) = e j j , [T (β),U(θ)] = 0. or half-integer representations, giving rise to two differ- For future reference, we introduce the tensor u(θ) ent classes of O(2) invariant MPS. In gapped phases, the occurence of either set of representations is known n1 to characterize a trivial (integer) or symmetry-protected

n2 topological (half-integer) phase [17, 18]. In order to illus- uθ = δ exp (iθn1) , (4) n1 trate these different classes of MPS and to see which one

n2 is realized for the fixed point of the XY transfer matrix, we have performed simulations (i) without imposing sym- N such that the symmetry operation is U(θ) = i ui(θ). metries on the legs, (ii) with integer irreps on the legs, The Mermin-Wagner theorem now dictates that this and (iii) with half-integer irreps on the legs. In TableI U(1) symmetry cannot be broken at any finite temper- we list the corresponding values for the transfer-matrix ature. On the level of the transfer matrix this implies leading eigenvalue λ at a point in the critical phase. that the leading eigenvector (fixed point) necessarily is A first observation is that we find a higher value for λ invariant under the U(1) transformation. For the MPS when no symmetries are imposed at significantly lower approximation of the fixed point, this implies that bond dimension. This shows that a variational MPS favours the breaking of the U(1) symmetry. As we have A A A A discussed in the introduction, this artificial symmetry breaking is also observed in other critical systems with uθ uθ uθ uθ a U(1) symmetry [24]. The explanation for this effect can be sought in the fact that MPS necessarily induces a finite correlation length. Indeed, simulating a critical = A A A A system with MPS with a finite bond dimension can be thought of as slightly perturbing the system such that for any θ. The fundamental theorem of MPS now di- a gap is opened and a finite correlation length is intro- cates that we can associate this symmetry of the MPS to duced. From the perspective of an effective field theory 5

Z dθ D λ × eβh cos θeiθ(n1+n2−n3−n4)h(θ). no symmetries 19 2.5869206 2π integer charges 45 2.5869172 Using the MPS representation of the transfer-matrix half-integer charges 46 2.5869184 fixed point, we can simplify this to

TABLE I. The leading eigenvalue (per site) of the XY transfer   matrix in the critical phase (T = 0.8) as obtained by imposing A A A   an MPS approximation for the fixed point, where we have    O M O  imposed different symmetry properties on the virtual legs of   the MPS.   A¯ A¯ A¯ hh(θ)i =   . describing the critical system, the MPS adds a relevant A A A   perturbation that opens up a gap in this field theory. In    O O O  the U(1) phase, however, only symmetry-breaking terms   are relevant, such that we expect that the MPS approx-   A¯ A¯ A¯ imation induces an artificial symmetry breaking. This observation is confirmed when we compute the U(1) or- der parameter in the next section. The contractions of these infinite channels are evaluated Secondly, upon imposing the U(1) or O(2) symmetry by finding the leading eigenvectors of the channel opera- explicitly, we observe that the eigenvalue reaches a sim- tors. ilar value at a comparable bond dimension in the two Similarly, a generic nearest-neighbour two-angle ob- (normal and projective, resp.) sectors. If we again inter- servable at sites j and k is given by pret the MPS approximation as introducing a gap, this   result suggests that the critical U(1) phase can be per-   turbed into both an SPT phase and a regular phase. As  O O O O    we will see in Sec.V, in the gapped phase it can be deter- 1     mined without ambiguity what irreps should be chosen hh(θj, θk)i =  O E O  , Z   on the MPS virtual legs.    O O O O    C. Computing local observables with Local observables such as the internal energy and the n1 n2 magnetization can be represented in the tensor-network 6 !1/2 language as follows. A generic one-angle observable at Y n6 E n3 = I (β) site j ni i=1 n5 n4 1 Z dθ  Y i −βE({θi}) Z dθ Z dθ hh(θj)i = e h(θj), j k βh(cos θj +cos θk) Z 2π × h(θj, θk)e i 2π 2π  X iθj (n1+m−n5−n6) can, under the duality transform that we introduced × Im(β)e above, be represented diagrammatically as m  iθk(n2+n3−n4−m)   e .    O O O  Using the MPS fixed points that we have optimized ear-   lier for computing the free energy – i.e. without using 1     symmetries on the virtual level – we now compute the hh(θi)i =  O M O  Z   internal energy e, the entropy s and the order parameter    O O O  o  

e = −2 hcos(θi − θj)i with s = β(e − f) iθ o = he i i , n1 4 !1/2 Y as a function of temperature, and plot the results in n4 M n2 = I (β) ni Fig.2. Again, the energy and entropy show no sign of i=1 n3 the phase transition. Also, we observe that the value 6 for the energy has already converged up to an error of  = 10−6 at a bond dimension of D = 30; since the en- tropy is computed from the free energy and the internal energy, it has the same accuracy. As anticipated in the previous section, the order parameter shows a very large value in the critical region, which decreases very slowly as the bond dimension increases. This shows that the MPS breaks the continuous U(1) symmetry significantly in the critical phase, whereas in the gapped phase the symme- try is restored. The fact that the order parameter decays very slowly with increasing bond dimension shows that this is an essential property of MPS approximations for U(1) phases. On the other hand, from the convergence of the free and internal energy, we see that this does not (a) prohibit an accurate evaluation of the system’s physical properties. Note that the order parameter is identically zero if we would impose the U(1) or O(2) symmetry on the virtual level of the MPS, but, as we have seen in Tab.I, at a large variational cost in the free energy.

III. THE SPIN STIFFNESS

The phase transition in the XY model can, in the ab- sence of a local order parameter, be characterized by the (b) so-called spin stiffness. This quantity is defined as the response to a twist field ~v, which rotates the angles as

θi → θi + ~v · ~ni.

If we take the twist field along the y axis, this modifies the classical hamiltonian (without magnetic field) to

X X Hv = − cos(θi − θj) − cos(θi − θj + v).

hijix hijiy (c)

On the level of the partition function, this introduces an FIG. 2. Observables from MPS simulations with bond dimen- sion D = 30 (no virtual symmetries). In (a) we plot the or- extra phase factor on the vertical links, der parameter, showing significant symmetry breaking in the critical region; the inset shows that this value (for T = 0.7) of the order parameter decreases very slowly with the bond dimension. In (b) we plot the internal energy, showing no N ! Y X sign of a phase transition; here, the inset shows good conver- Z = I (β)einlv v nl gence even for small bond dimensions (T = 0.7). In (c) we l∈Ly nl=−N plot the entropy per site, which is easily evaluated from the N ! free and internal energy; we have renormalized the entropy Y X Y 1 × I (β) F ns,3,ns,4 . as s(β) → s(β) − s(∞), with s(∞) = log( ), such that the nl ns,1,ns,2 2π zero-temperature limit yields a zero entropy and the entropy l∈Lx nl=−N s is positive everywhere. 7

Incorporating these extra phase factors, we can represent We can work around this by computing the spin stiff- the partition function as ness as a two-point function. First we note that we can rewrite the expression for the spin stiffness as

2 1 ∂ Zv O O O ρ = − lim , N→∞ 2 NβZ(0) ∂v v=0 uv uv uv where the twisted partition function can be written as O O O Zv = , uv uv uv Ov Ov Ov Ov O O O Ov Ov Ov Ov uv uv uv Zv = , Ov Ov Ov Ov

Ov Ov Ov Ov where the tensor uv was defined in Eq. (4). We find for the transfer matrix after the twist

with the twisted tensor O O O O Tv(β) = n1 1/2 uv uv uv uv 4 ! Y in v n ,n n4 O n2 = I (β) e 3 F 3 4 . v ni n1,n2 = T (β)U(v) = U(v)T (β), i=1 n3 because U(v) commutes with the transfer matrix. This In the tensor-network representation of Z we can easily implies that all eigenvectors of the transfer matrix remain v differentiate with respect to v. Indeed, the first derivative unchanged after a twist and the eigenvalues are multi- is given by plied by a phase. In particular, the fixed point remains unchanged and, as the fixed point lives in the charge- zero sector, so does the leading eigenvalue. This, in turn, implies that the free energy density is a constant when O O O applying a twist 1 ∂Zv = O R O , f(v) = f(0), N ∂v v=0 O O O and that the spin stiffness, defined as

2 ∂ f ρ = 2 , ∂v v=0 ∂Ov with the tensor R = ∂v v=0, or is identically zero for all temperatures. n1 This is a surprising result, because ρ is supposed to 4 !1/2 jump discontinuously at the transition. The reason for Y n3,n4 n4 R n2 = in3 In (β) F . this discrepancy lies in the fact that we work directly in i n1,n2 i=1 the thermodynamic limit. Indeed, the spin stiffness is n3 typically defined in a system that is finite and has peri- odic boundary conditions in the direction of the twist. For the second derivative we have to differentiate two In that setting, applying the twist field is equivalent different tensors, and twice the same tensor. The result to imposing twisted boundary conditions – this can be is given by (the site i is arbitrary) observed from the above representation of the partition function, where the operator U(v) can be pulled through the lattice – and is, therefore, a finite-size property. It is only after it has been defined on this periodic sys- O O O 2 tem, that the infinite-size limit can be taken. The above 1 ∂ Zv i = O S O definition of the spin stiffness in terms of the infinite-size 2 N ∂v v=0 transfer matrix, on the other hand, assumes open bound- O O O ary conditions. If the transfer matrix has no gap in the thermodynamic limit, both definitions are not equivalent. 8

FIG. 3. The spin stiffness as a function of temperature for a set of different values of the magnetic field: h = 10−3 (green), h = 10−4 (purple), h = 10−5 (yellow), h = 10−6 (red), h = 10−7 (blue). The full lines were computed with a bond dimension 2T D = 150, the dashed lines are D = 90. We have also plotted the straight line π , which is known to intersect the curve for the critical temperature [Eq. (1)]; for h = 10−7 and D = 150 we find an intersection at T = 0.899.

with X O O O Sy = sin(θi − θj). j X hijiy + O O R , (6) j6=i i In this form, it can be evaluated in Monte-Carlo simu- R O O lations on a system with periodic boundary conditions without explicitly applying a twist [14, 30]. Again, the evaluation of this two-point function is iden-

2 tically zero on an infinite system with open boundary ∂ Ov where we have introduced the tensor S = ∂v2 , or v=0 conditions in the presence of an unbroken U(1) symme- try. Indeed, if we represent the infinite upper plane of n1 the above expression by the fixed point of the transfer 4 !1/2 2 Y n ,n matrix, we observe that n4 S n2 = (in ) × I (β) F 3 4 . 3 ni n1,n2 i=1 A A A A n3 X j This reduces the spin stiffness to a summation of two- O O R O j point functions, and therefore has the form of a structure factor. We should note that bringing the factor in3 down A A A A in the tensor is equivalent to introducing a factor sin(θi − X O O O O θj) in the partition function because of the identities = i , j j ∞ Q X inθ x cos θ (in)In(x)e = e sin θ n=−∞ which vanishes because the fixed point is U(1) symmetric. ∞ X This is, of course, consistent with the vanishing of the (in)2I (x)einθ = ex cos θ(sin2 θ − cos θ). n spin stiffness as defined above. n=−∞ For evaluating the spin stiffness, we can, however, in- The spin stiffness can therefore be brought into the form troduce a magnetic field in the hamiltonian, which breaks the U(1) symmetry of the model and induces a gap. In 1 the presence of this extra field, the spin stiffness as de- ρ = − hS2i − hcos(θ − θ )i Nβ y i j fined by the above structure factor [Eq. (6)] can be eval- 9 uated in an infinite system with open boundary condi- tions. Similarly to taking the infinite-size limit for a pe- riodic system, we can then take the limit h → 0 to obtain the result in the zero-field case. In Fig.3 the results for the spin stiffness (as defined from the structure factor in Eq. (6)) as a function of temperature are given for a few values of the magnetic field. These results were obtained by using the channel-environment construction of Ref. 31 for evaluating structure factors in two-dimensional tensor networks. The figure shows that the drop in the spin stiff- ness becomes sharper as the magnetic field is decreased, but for very small values of h the effects of a finite bond dimension become more pronounced. (a)

IV. LUTTINGER LIQUID MAPPING

In the previous section we have computed the spin stiff- ness by explicitly breaking the U(1) symmetry via an ex- ternal magnetic field. In this section, we show that we can use the Luttinger-liquid formalism [32] to compute the same quantity without an explicit breaking of the symmetry, directly in the thermodynamic limit. Early on, it was realized [6] that the proper effective field theory for the partition function is given by the sine-Gordon model [33], which is described by the one- dimensional quantum hamiltonian Z 1  2 u 2 H = dx uK (∇θ(x)) + (∇φ(x)) (b) SG 2π K Z + g dx cos(2φ(x)).

The first line is the hamiltonian of the Luttinger-liquid field theory [32] describing the critical spin-wave excita- tions, and the second line adds vortices to the picture. The microscopic U(1) symmetry of the model is reflected in the sine-Gordon field theory by the generator 1 Z − dx∇φ(x). π Since vortices are irrelevant in the critical phase, we ex- pect that under an RG transformation the vortices will drop from the sine-Gordon hamiltonian and the parame- (c) ters u and K will be renormalized. In other words, we ex- pect that the low-energy properties of the XY model will FIG. 4. The XY model at T = 0.7 as a Luttinger liquid. be described by an effective Luttinger-liquid field theory In (a) we have plotted the dispersion relation of the transfer Z   matrix as defined by Eq. (7); the dashed line is a straight line 1 2 u˜ 2 with a slope equal to one, showing that we find a Luttinger HLL = dx u˜K˜ (∇θ(x)) + (∇φ(x)) , 2π K˜ velocity exactly equal to one. In (b) we have plotted the correlation length versus the entanglement entropy obtained with the effective parametersu ˜ and K˜ determined by the in MPS approximations for the fixed point with increasing inverse temperature β. As soon as the effective Luttinger bond dimension (D = 50 : 25 : 175); the fit is made from parameter reaches K˜ = 2, however, vortices become rel- the three last points according to the form Eq. (8), yielding a evant, leading to a gapped phase where the φ field is value for the central charge c = 1.003. In (c) we have plotted locked in the minima of the cosine term; the elementary the response in the free energy (left) and the expectation value for Q with a quadratic and linear fit, resp. The same value excitations are kinks and anti-kinks between the different i κ ≈ 1.10953 is found from both fits. ground states. 10

We can lift this notion of an effective field theory to the This shift in the transfer matrix does not affect its eigen- level of a transfer matrix, where we formally introduce a vectors, but reshuffles the eigenvalues. Therefore the Luttinger-liquid transfer matrix fixed point |ψµi will change as a function of µ, and will give rise to a finite expectation value of the generator. TLL = exp (−HLL) . The associated compressbility The idea is that this captures the low-energy behavior d hψµ| Qi |ψµi κ = of the XY transfer matrix, where the parametersu ˜ and dµ K˜ depend on the inverse temperature β. We confirm µ=0 the low-energy correspondence of the XY transfer matrix will therefore give a finite value. This, in turn, yields a T (β) with this effective Luttinger-liquid form in three direct estimate of the effective Luttinger parameter K˜ . different ways. Since Q commutes with the transfer matrix, the parti- The first correspondence can be found by computing tion function will only be affected in second order, and the low-lying spectrum of the XY transfer matrix, us- the second derivative yields the same value for the com- ing the MPS quasiparticle ansatz for the low-lying ex- pressibility, cited states [26, 28]. We can label the excitations with a momentum p, yielding a dispersion relation ω(p) as the d2 log λ (β) κ = − µ , logarithm of its eigenvalues, 2 dµ µ=0 hφ(p)| T (β) |φ(p)i ω(p) = − log , (7) where λµ(β) is the usual eigenvalue per site of the transfer hφ(p)|φ(p)i matrix

1/Nx and observe that we find a gapless spectrum with a lin- λµ(β) = lim hψµ| Tµ(β) |ψµi . ear dispersion relation ω(p) =u ˜|p|. Even stronger, we Nx→∞ find the effective velocityu ˜ = 1 for all values of the We can compute this compressibility straightforwardly temperature within the critical phase (see Fig.4(a)). using the uniform MPS framework. Indeed, as explained It is, of course, expected that the effective field theory above, the eigenvalue λµ(β) is the quantity that is vari- is isotropic, but this serves as an excellent test of our ationally optimized in an MPS fixed-point simulation, transfer-matrix approach. whereas the charge density hQii can be easily computed A second test consists of measuring the central charge, as an expectation value. In Fig.4 it is shown that both which can be determined in an MPS simulation by com- quantities yield a consistent numerical value for the com- paring the scaling of the entanglement entropy and the pressibility. correlation length as a function of the bond dimension. It We should note, however, that the transfer matrix is known that, for a system that is described by a confor- Tµ(β) is equivalent to the one of the twisted XY model mal field theory the scaling is determined by the central but with an imaginary value for the twist field charge c as [34, 35] c Tµ(β) = T (β)U(iµ). SD ∝ log(ξD). (8) 6 Therefore, the effective Luttinger parameter that we have In Fig.4(b) we show clear evidence for a central charge defined here is related to the spin stiffness from before, c = 1, which is precisely the value for a Luttinger liquid. ˜ An estimate for the effective Luttinger parameter K˜ K = πβρ. is obtained by computing the response to a chemical po- This correspondence, which can be readily seen from the tential. It is easily seen [32] that adding the generator of mapping of the XY model to the sine-Gordon field theory the U(1) symmetry, [6, 33], yields the famous [7] value for the spin stiffness µ Z  at the critical temperature ρ = 2Tc/π. TLL → exp dx∇φ(x) TLL In the previous section, we had anticipated that com- π puting the spin stiffness without introducing a symmetry- yields a compressilbity that is equal to breaking term would not be possible in the thermody- namic limit directly. The reason that we are here able 1 d K˜ to compute the spin stiffness without breaking the U(1) κ = − h∇φi = . π dµ µ uπ˜ symmetry consists in the fact that we have expressed it as a thermodynamic quantity (the compressibility) for We have an explicit microscopic form for the generator which the extensivity properties of the uniform MPS sim- of the U(1), so we can explicity implement a chemical ulations are ideally suited. This thermodyamic quantity, potential on the level of the microscopic transfer matrix however, is necessarily formulated on the level of the as transfer matrix after the duality transformation, since  µ   µ  an imaginary twist does not translate to a realistic mod- T (β) = exp − Q T (β) exp − Q . µ 2 2 ification of the classical XY hamiltonian. 11

As a final signature of the Luttinger-liquid phase, we investigate the entanglement spectrum of the MPS fixed point. As was observed in Ref. 36, the low-lying part of the entanglement spectrum for a bipartition of the MPS, should resemble the energy spectrum of a boundary con- formal field theory (CFT). In Fig.5 we plot the entangle- ment spectrum of the MPS fixed point, where we have imposed O(2) symmetry on the MPS tensor such that we can label the spectrum with the appropriate quantum numbers. We observe that the spectrum has a quadratic envelope, and that we obtain an equidistant spectrum after rescaling the different sectors, in perfect correspon- dence with the spectrum of a free-boson boundary CFT. Moreover, from the rescaling parameter we can deduce an estimate of the Luttinger parameter K˜ , which nicely (a) converges to the same value as the one we find using the compressibility.

V. THE GAPPED PHASE

The characterization of the gapped phase of the XY transfer matrix using uniform MPS is a lot more straight- forward. Indeed, we expect that the fixed point of a gapped transfer matrix can be approximated by an MPS with arbitrary precision. Therefore, we no longer expect that the MPS will spontaneously break the U(1) or O(2) symmetry, and we can safely use the fundamental the- orem to realize that the virtual legs of the MPS should transform under (projective) representations as well (ac- cording to Eq. (5)). (b) As explained in Sec.IIB, we are ignorant on which representations should be chosen on the virtual legs of FIG. 5. Entanglement spectrum of the fixed-point MPS of the MPS. For that reason, we first plot the entangle- the XY transfer matrix at T = 0.8. We have imposed the full ment spectrum of a fixed-point MPS around temperature O(2) symmetry on the MPS tensor, which implies that the T = 1.2 without explicit symmetries on the MPS repre- entanglement spectrum is labeled by the irreps on the virtual sentation (see Fig.6). We find both isolated and twofold bonds: two one-dimensional irreps with charge q = 0 (blue, degenerate Schmidt values, which point to integer repre- red), and two-dimensional irreps with charges q = 1, 2,... . In sentations on the virtual legs. Indeed, in Fig.6 we plot (a) we plot the bare entanglement spectrum, and we find that the envelope of the entanglement spectrum follows a quadratic the entanglement spectrum with integer representations form (striped line). In (b) we have shifted the different sectors on the virtual legs imposed, showing that the isolated such that the lowest value is zero; this produces a nice free- values correspond to either of the two one-dimensional boson boundary CFT spectrum. irreps with n = 0 charge sector; the twofold degenerate ones correspond to n = 1, 2,... . One of the hallmarks of the BKT transition is the this to exponential divergence of the correlation length when b p approaching the critical point from the gapped side log ξ = √ + c + d T − Tc, (9) T − T [Eq. (2)]. Using MPS, we can confirm this behaviour and c use this form to obtain an estimate for the critical point. where the extra terms are added to account for deviations The correlation length is a notoriously hard quantity to away from the critical point. This fit yields an estimate converge in MPS simulations, but using the second gap in for the correlation length of Tc = 0.8930(1), which agrees the transfer matrix (typically denoted as δ) it is possible well with other numerical results (see Tab.II). We should to extrapolate its value in a reliable way [37]. In Fig.7 we note, however, that estimates for the critical point can plot this extrapolation procedure for T = 0.93, yielding depend strongly on higher-order corrections to the scaling an accurate value for a correlation length of more than a behavior – for the finite-size extrapolation of the spin thousand sites. In Fig.7 we then plot the extrapolated stiffness this is clearly the case [14]. We leave the careful correlation lengths as a function of temperature, and fit incorporation of higher-order contributions to the scaling 12

FIG. 6. The entanglement spectrum of the fixed-point MPS at T = 1.216, where we impose O(2) invariance on the vir- tual legs (left) and without explicit symmetries (right). On (a) the left, we label the different O(2) representations as follows: n = 0 (blue crosses), q = ±1 (red), q ± 2 (orange) and q = ±3 (purple), whereas on the right we don’t have no labeling; we only show the Schmidt values above 10−4. The perfect cor- respondence of the entanglement spectra with and without explicit symmetries on the virtual legs shows that the MPS representation for the fixed-point only contains integer repre- sentations of O(2). of the correlation length, and a more accurate estimation of the critical point, for further study. The gapped phase is further characterized by the low- lying spectrum of the transfer matrix, which we define as before in Eq.7. In Fig.8 we have plotted the spectrum at a temperature T = 1.3, showing an isolated two-fold (b) degenerate quasiparticle line; the excited states on this line carry U(1) quantum numbers q = ±1. Above this el- FIG. 7. In (a) we plot the extrapolation procedure for T = ementary one-particle excitation, we find the two-particle 0.93 yielding a value for the correlation length ξ = 1463(4); continuum and, interestingly, around momentum p = π the maximal dimension in each block was set at Dmax = 512, we find a slightly bound state below the continuum with yielding a total MPS bond dimension of D = 3754. In (b) charge q = 0. we plot the extrapolated correlation lengths as a function of temperature. We fit this to the KT form [Eq. (9)] (red line), yielding a value for the critical temperature Tc = 0.8930(1).

VI. OUTLOOK transition. First of all, despite the fact that an MPS In this paper we have investigated the classical two- breaks U(1) symmetry, we can evaluate local observables dimensional XY model using uniform MPS methods. We with very high precision in the critical phase. Secondly, have shown that an MPS approximation for the fixed we have shown that the spin stiffness can be evaluated by point of the XY transfer matrix breaks the U(1) heavily introducing a small magnetic field h, and taking the limit in the critical phase, which is expected because the MPS h → 0. The mapping to an effective Luttinger-liquid field always induces a finite correlation length in the system. theory can be made explicit using the MPS framework by In a similar vein, a uniform MPS calculation of the spin computing the central charge, the dispersion relation and stiffness in the critical U(1) phase is always zero. The the compressibility; the latter, which is the response to a reason for the latter was sought in the fact that uniform twist with imaginary magnitude, is used to find very ac- MPS work in the thermodynamic limit directly, whereas curate values for the effective Luttinger parameter K˜ . In the spin stiffness is a quantity that is necessarily defined addition, we find that the entantglement spectrum of the in a finite periodic system; only for systems with a gap MPS is in agreement with a boundary CFT spectrum. In the two definitions intersect. the gapped phase, the MPS leaves the (non-abelian) O(2) Nonetheless, we showed that uniform MPS are an ideal symmetry unbroken, which we can use to find accurate framework for characterizing the XY model and its phase values for the correlation length upon approaching the 13

Monte Carlo (1979) [10] 0.89 Monte Carlo (2005) [11] 0.8929(1) series expansion (2009) [13] 0.89286(8) Monte Carlo (2012) [12] 0.89289(6) Monte Carlo (2013) [14] 0.8935(1) tensor- (2014) [25] 0.8921(19) uniform MPS (current work) 0.8930(1)

TABLE II. Numerical estimates for the critical temperature

in this paper might prove very interesting in this con- text. The results in this paper will prove instrumental in the program of simulating systems with unbroken continu- ous symmetries with uniform tensor networks and should, in particular, be useful in the study of two-dimensional quantum spin liquids with PEPS. Indeed, the norm of a PEPS can be naturally interpreted as a two-dimensional partition function and the question often poses itself in what phase the corresponding PEPS transfer matrix is. The paradigmatic example here is the resonating valence- bond (RVB) wavefunction on the square lattice, for which the transfer matrix is known to be in a U(1) phase [43, 44], and also other symmetric PEPS parametriza- tion for (chiral) spin liquids seem to give rise to critical FIG. 8. The spectrum ω(p) of the transfer matrix at temper- transfer matrices [45]. The relation between the critical ature T = 1.3. The blue line is elementary excitation branch with charge q = ±1, the red line is the edge of the two-particle properties of the transfer matrix, which are esssentially continuum. The yellow dots are excitation energies that fall the properties of a two-dimensional classical system, the below the continuum edge, signalling a bound state in the symmetries of the PEPS tensors, and the quantum prop- q = 0 sector. erties of the PEPS wavefunction (physical correlation functions, entanglement spectra, etc.) remains, however, largely unexplored. critical point; from fitting the exponential divergence of In a different direction, the two-dimensional partition the correlation length we find T ≈ 0.8930, in agreement c functions that we have considered here, can be naturally with other numerical studies. lifted to the quantum level by promoting the charges on We expect that the current setup can be applied to the bonds to quantum-mechanical degrees of freedom. other two-dimensional classical systems with continu- Upon doing that, we find quantum-mechanical wavefunc- ous symmetries. Whereas the standard ferromagnetic tions for U(1) gauge theories on the lattice. This PEPS Heisenberg model has no phase transition [38], the so- construction can be generalized to a whole variational called RP2 models with a classical hamiltonian H = 2 class of states that are ideally suited to study the phase − P (~s · ~s ) (with ~s a three-dimensional unit vec- hiji i j i diagram of two-dimensional lattice gauge theories, and, tor) potentially hosts Z2 vortices that drive a phase in order to understand the phase transitions, we will need transition [39]. A similar phase transition might be the tools that were explored in this paper. present in the frustrated antiferromagnetic Heisenberg model on the triangular lattice [40, 41]. Also, our meth- We acknowledge inspiring discussions with Nick Bult- ods can be readily applied for simulating KT transi- inck, Thierry Giamarchi, Jutho Haegeman, and Masaki tions in one-dimensional quantum systems. The relation Oshikawa. This work was supported by the Flemish Re- between these classical topological transitions and one- search Foundation, the Austrian Science Fund (ViCoM, dimensional SPT phases [42] that we have investigated FoQuS), and the European Commission (QUTE 647905).

[1] N. D. Mermin and H. Wagner, Absence of ferromag- 1133 (1966). netism or antiferromagnetism in one- or two-dimensional [2] P. C. Hohenberg, Existence of Long-Range Order in One isotropic heisenberg models, Physical Review Letters 17, and Two Dimensions, Physical Review 158, 383 (1967). 14

[3] V. L. Berezinskii, Destruction of long-range order in one- try, Physical Review A 82, 050301 (2010). dimensional and two-dimensional systems possessing a [22] A. Weichselbaum, Non-abelian symmetries in tensor net- continuous symmetry group. ii. quantum systems, Soviet works: A quantum symmetry space approach, Annals of Journal of Experimental and Theoretical Physics 34, 610 Physics 327, 2972 (2012). (1972). [23] V. Zauner-Stauber, L. Vanderstraeten, J. Haegeman, [4] J. M. Kosterlitz and D. J. Thouless, Ordering, metasta- I. P. McCulloch, and F. Verstraete, Topological nature bility and phase transitions in two-dimensional systems, of spinons and holons: Elementary excitations from ma- Journal of Physics C: Solid State Physics 6, 1181 (1973). trix product states with conserved symmetries, Physical [5] J. M. Kosterlitz, The critical properties of the two- Review B 97, 235155 (2018). dimensional xy model, Journal of Physics C: Solid State [24] D. Draxler, J. Haegeman, T. J. Osborne, V. Stojevic, Physics 7, 1046 (1974). L. Vanderstraeten, and F. Verstraete, Particles, holes, [6] J. V. Jos´e,L. P. Kadanoff, S. Kirkpatrick, and D. R. Nel- and solitons: A matrix product state approach, Physical son, Renormalization, vortices, and symmetry-breaking Review Letters 111, 020402 (2013). perturbations in the two-dimensional planar model, Phy- [25] J. F. Yu, Z. Y. Xie, Y. Meurice, Y. Liu, A. Denbleyker, ical Review B 16, 1217 (1977). H. Zou, M. P. Qin, J. Chen, and T. Xiang, Tensor [7] D. R. Nelson and J. M. Kosterlitz, Universal jump in the renormalization group study of classical xy model on the superfluid density of two-dimensional superfluids, Physi- square lattice, Physical Review E 89, 013308 (2014). cal Review Letters 39, 1201 (1977). [26] J. Haegeman and F. Verstraete, Diagonalizing Transfer [8] P. Minnhagen, The two-dimensional coulomb gas, vor- Matrices and Matrix Product Operators: A Medley of tex unbinding, and superfluid-superconducting films, Re- Exact and Computational Methods, Annual Review of views of Modern Physics 59, 1001 (1987). Condensed Matter Physics 8, 355 (2017). [9] S. Miyashita, H. Nishimori, A. Kuroda, and M. Suzuki, [27] M. T. Fishman, L. Vanderstraeten, V. Zauner-Stauber, Monte carlo simulation and static and dynamic critical J. Haegeman, and F. Verstraete, Faster methods for con- behavior of the plane rotator model, Progress of Theo- tracting infinite two-dimensional tensor networks, Phys- retical Physics 60, 1669 (1978). ical Review B 98, 235148 (2018). [10] J. Tobochnik and G. V. Chester, Monte carlo study of the [28] L. Vanderstraeten, J. Haegeman, and F. Verstraete, planar , Physical Review B 20, 3761 (1979). Tangent-space methods for uniform matrix product [11] M. Hasenbusch, The two-dimensional xy model at the states, SciPost Phys. Lect. Notes , 7 (2019). transition temperature: a high-precision monte carlo [29] D. Perez-Garcia, F. Verstraete, M. Wolf, and J. Cirac, study, Journal of Physics A: Mathematical and General Matrix product state representations, arXiv (2006), 38, 5869 (2005). quant-ph/0608197. [12] Y. Komura and Y. Okabe, Large-scale monte carlo sim- [30] A. W. Sandvik, Computational studies of quantum spin ulation of two-dimensional classical xy model using mul- systems, AIP Conference Proceedings 1297, 135 (2010). tiple gpus, Journal of the Physical Society of Japan 81, [31] L. Vanderstraeten, M. Mari¨en, F. Verstraete, and 113001 (2012). J. Haegeman, Excitations and the tangent space of [13] H. Arisue, High-temperature expansion of the magnetic projected entangled-pair states, Physical Review B 92, susceptibility and higher moments of the correlation 201111 (2015). function for the two-dimensional xy model, Physical Re- [32] T. Giamarchi, Quantum Physics in One Dimension (Ox- view E 79, 011107 (2009). ford University Press, 2004). [14] Y.-D. Hsieh, Y.-J. Kao, and A. W. Sandvik, Finite- [33] L. Benfatto, C. Castellani, and T. Giamarchi, size scaling method for the berezinskii-kosterlitz-thouless Berezinskii-kosterlitz-thouless transition within the transition, Journal of Statistical Mechanics: Theory and sine-gordon approach: The role of the vortex-core Experiment 2013, P09001 (2013). energy, in 40 Years of Berezinskii-Kosterlitz-Thouless [15] F. Verstraete, V. Murg, and J. I. Cirac, Matrix prod- Theory, pp. 161–199. uct states, projected entangled pair states, and varia- [34] F. Pollmann, S. Mukerjee, A. M. Turner, and J. E. tional renormalization group methods for quantum spin Moore, Theory of finite-entanglement scaling at one- systems, Advances in Physics 57, 143 (2008). dimensional quantum critical points, Physical Review [16] R. Or´us,A practical introduction to tensor networks: Letters 102, 255701 (2009). Matrix product states and projected entangled pair [35] B. Pirvu, G. Vidal, F. Verstraete, and L. Tagliacozzo, states, Annals of Physics 349, 117 (2014). Matrix product states for critical spin chains: Finite-size [17] F. Pollmann, A. M. Turner, E. Berg, and M. Oshikawa, versus finite-entanglement scaling, Physical Review B 86, Entanglement spectrum of a topological phase in one di- 075117 (2012). mension, Physical Review B 81, 064439 (2010). [36] A. M. Laeuchli, Operator content of real-space entangle- [18] X. Chen, Z.-C. Gu, and X.-G. Wen, Classification of ment spectra at conformal critical points, arXiv (2013), gapped symmetric phases in one-dimensional spin sys- 1303.0741. tems, Physical Review B 83, 035107 (2011). [37] M. M. Rams, P. Czarnik, and L. Cincio, Precise extrap- [19] N. Schuch, J. I. Cirac, and D. P´erez-Garc´ıa, Peps as olation of the correlation function asymptotics in uni- ground states: Degeneracy and topology, Annals of form tensor network states with application to the bose- Physics 325, 2153 (2010). hubbard and xxz models, Physical Review X 8, 041033 [20] I. P. McCulloch and M. Gul´acsi,The non-abelian density (2018). matrix renormalization group algorithm, Europhysics [38] S. H. Shenker and J. Tobochnik, Monte carlo Letters 57, 852 (2002). renormalization-group analysis of the classical heisenberg [21] S. Singh, R. N. C. Pfeifer, and G. Vidal, Tensor net- model in two dimensions, Physical Review B 22, 4462 work decompositions in the presence of a global symme- (1980). 15

[39] S. Solomon, Vortices and phase transitions in two- in crystal surfaces and valence-bond phases in quantum dimensional non-abelian spin models, Physics Letters B spin chains, Physical Review B 40, 4709 (1989). 100, 492 (1981). [43] N. Schuch, D. Poilblanc, J. I. Cirac, and D. P´erez-Garc´ıa, [40] H. Kawamura and S. Miyashita, Phase transition of the Resonating valence bond states in the peps formalism, two-dimensional heisenberg antiferromagnet on the tri- Physical Review B 86, 115108 (2012). angular lattice, Journal of the Physical Society of Japan [44] J.-Y. Chen and D. Poilblanc, Topological Z2 resonating- 53, 4138 (1984). valence-bond spin liquid on the square lattice, Physical [41] H. Kawamura, A. Yamamoto, and T. Okubo, Z2-vortex Review B 97, 161107 (2018). ordering of the triangular-lattice heisenberg antiferro- [45] J.-Y. Chen, L. Vanderstraeten, S. Capponi, and D. Poil- magnet, Journal of the Physical Society of Japan 79, blanc, Non-abelian chiral spin liquid in a quantum anti- 023701 (2010). ferromagnet revealed by an ipeps study, Physical Review [42] M. den Nijs and K. Rommelse, Preroughening transitions B 98, 184409 (2018).