Kekul´e valence-bond solid (KVBS) masses map onto a field ...

10
arXiv:2005.08996v1 [cond-mat.str-el] 18 May 2020 Topological terms on topological defects: a quantum Monte Carlo study Toshihiro Sato, 1 Martin Hohenadler, 1 Tarun Grover, 2 John McGreevy, 2 and Fakher F. Assaad 1, 3 1 Institut f¨ ur Theoretische Physik und Astrophysik, Universit¨ at W¨ urzburg, 97074 W¨ urzburg, Germany 2 Department of Physics, University of California at San Diego, La Jolla, CA 92093, USA 3 urzburg-Dresden Cluster of Excellence ct.qmat, Am Hubland, 97074 W¨ urzburg, Germany Dirac fermions in 2 + 1 dimensions with dynamically generated anticommuting SO(3) antiferro- magnetic (AFM) and Z2 Kekul´ e valence-bond solid (KVBS) masses map onto a field theory with a topological θ-term. This term provides a mechanism for continuous phase transitions between different symmetry-broken states: topological defects of one phase carry the charge of the other and proliferate at the transition. The θ-term implies that a domain wall of the Z2 KVBS order parameter harbors a spin-1/2 Heisenberg chain, as described by a 1 + 1 dimensional SO(3) non-linear sigma model with θ-term at θ = π. Using pinning fields to stabilize the domain wall, we show that our auxiliary-field quantum Monte Carlo simulations indeed support the emergence of a spin-1/2 chain at the Z2 topological defect. This concept can be generalized to higher dimensions where 2 + 1 dimensional SO(4) or SO(5) theories with topological terms are realized at a domain wall. Introduction.—Topological terms in field theories play an important role in our understanding of phases and critical phenomena. For instance, the differences between integer and half-integer spin-S chains are a consequence of the 2πiS pre-factor of the integer-valued θ-term that counts the winding of a unit vector over the sphere. Dirac fermions provide a very appealing route to define models that map onto field theories with topological terms [1–9]. Consider 8-flavored Dirac fermions in 2 + 1 dimensions akin to graphene. In this case, there is a maximum of five anti-commuting mass terms that could, for instance, cor- respond to an antiferromagnet (AFM) with three mass terms and a Kekul´ e valence bond solid (KVBS) with two mass terms [10]. The ten commutators of these mass terms correspond to the generators of the SO(5) group so that Dirac fermions Yukawa-coupled to these five mass terms possess an SO(5) symmetry. In the massive phase, one can integrate out the fermions to obtain a Wess- Zumino-Witten (WZW) topological term [1, 2] that is believed to be at the origin of deconfined quantum crit- icality (DQC) [11, 12]. In particular, it formalizes the Levin-Senthil picture [13] of a vortex of the Kekul´ e order harboring an emergent spin-1/2 degree of freedom. The aim of this Letter is to demonstrate numerically the consequences of topological terms in the correspond- ing field theory. We will do so by considering a model of Dirac fermions in 2+1 dimensions with reduced spatial symmetries such that the three AFM and one of the two KVBS mass terms are dynamically generated. Contrary to the generic KVBS state with a spontaneously bro- ken U(1) symmetry in the continuum, our KVBS state spontaneously breaks a Z 2 symmetry. We will refer to this state as Z 2 KVBS. Starting from the WZW topo- logical term, this symmetry reduction amounts to set- ting one component of the five-dimensional field to zero. This maps the WZW term to a θ-term at θ = π [3]. Let us assume that the phase transition observed nu- merically between the AFM and the Z 2 KVBS is con- tinuous and captured by the aforementioned field the- ory. Then, the θ-term leads to the prediction that in the Z 2 KVBS phase close to the transition, a Z 2 KVBS do- main wall harbors a spin-1/2 chain. In what follows, we will provide a model—amenable to large scale negative- sign-free auxiliary-field quantum Monte Carlo (QMC) calculations—that provides compelling results support- ing this field-theory picture. Field theory.—A theory that accounts for the phase di- agram presented in Ref. [6] (see Fig. 1(a)) contains Dirac fermions Yukawa-coupled to the AFM and Z 2 KVBS mass terms, as described by L F μ (1 2 γ 0 γ μ )+ η α χ · σ α γ 0 1 2 0 γ 5 Ψ . (1) Here, the Dirac spinors Ψ carry a sublattice index, a spin index, and a valley index. The γ -matrices act on the val- ley and sublattice spaces and satisfy the Clifford algebra {γ a b } =2δ ab . σ α with α =1, 2, 3 denote the Pauli spin-1/2 matrices. The fact that the SO(3) AFM mass terms σ α γ 0 and Z 2 KVBS mass terms 1 2 0 γ 5 anti- commute results in an SO(4) invariance of the fermionic action: a global SO(4) rotation of the four-component field φ =(φ 1 2 3 4 )=(η) is equivalent to a canonical transformation of the fermion operators. The dynamics of the field is governed by a four-component ϕ 4 action, L B . While L F has SO(4) symmetry, L B inherits the SO(3) × Z 2 symmetry of the lattice model. The Lagrangian L = L F + L B can account for many phase transitions. The Gross-Neveu transitions from semimetal to AFM or from semimetal to Z 2 KVBS in- volve a closing of the mass gap corresponding to the norm of the field φ. On the other hand, QMC simulations (see Ref. [6] and the Supplemental Material (SM)) point to a continuous transition between the AFM and Z 2 KVBS states with an emergent SO(4) symmetry. Importantly, the numerical results show that the single-particle gap remains finite across the transition. In the field theory, this implies that amplitude fluctuations of φ are frozen

Transcript of Kekul´e valence-bond solid (KVBS) masses map onto a field ...

Page 1: Kekul´e valence-bond solid (KVBS) masses map onto a field ...

arX

iv:2

005.

0899

6v1

[co

nd-m

at.s

tr-e

l] 1

8 M

ay 2

020

Topological terms on topological defects: a quantum Monte Carlo study

Toshihiro Sato,1 Martin Hohenadler,1 Tarun Grover,2 John McGreevy,2 and Fakher F. Assaad1, 3

1Institut fur Theoretische Physik und Astrophysik, Universitat Wurzburg, 97074 Wurzburg, Germany2Department of Physics, University of California at San Diego, La Jolla, CA 92093, USA

3Wurzburg-Dresden Cluster of Excellence ct.qmat, Am Hubland, 97074 Wurzburg, Germany

Dirac fermions in 2 + 1 dimensions with dynamically generated anticommuting SO(3) antiferro-magnetic (AFM) and Z2 Kekule valence-bond solid (KVBS) masses map onto a field theory witha topological θ-term. This term provides a mechanism for continuous phase transitions betweendifferent symmetry-broken states: topological defects of one phase carry the charge of the other andproliferate at the transition. The θ-term implies that a domain wall of the Z2 KVBS order parameterharbors a spin-1/2 Heisenberg chain, as described by a 1 + 1 dimensional SO(3) non-linear sigmamodel with θ-term at θ = π. Using pinning fields to stabilize the domain wall, we show that ourauxiliary-field quantum Monte Carlo simulations indeed support the emergence of a spin-1/2 chainat the Z2 topological defect. This concept can be generalized to higher dimensions where 2 + 1dimensional SO(4) or SO(5) theories with topological terms are realized at a domain wall.

Introduction.—Topological terms in field theories playan important role in our understanding of phases andcritical phenomena. For instance, the differences betweeninteger and half-integer spin-S chains are a consequenceof the 2πiS pre-factor of the integer-valued θ-term thatcounts the winding of a unit vector over the sphere. Diracfermions provide a very appealing route to define modelsthat map onto field theories with topological terms [1–9].Consider 8-flavored Dirac fermions in 2 + 1 dimensionsakin to graphene. In this case, there is a maximum of fiveanti-commuting mass terms that could, for instance, cor-respond to an antiferromagnet (AFM) with three massterms and a Kekule valence bond solid (KVBS) with twomass terms [10]. The ten commutators of these massterms correspond to the generators of the SO(5) groupso that Dirac fermions Yukawa-coupled to these five massterms possess an SO(5) symmetry. In the massive phase,one can integrate out the fermions to obtain a Wess-Zumino-Witten (WZW) topological term [1, 2] that isbelieved to be at the origin of deconfined quantum crit-icality (DQC) [11, 12]. In particular, it formalizes theLevin-Senthil picture [13] of a vortex of the Kekule orderharboring an emergent spin-1/2 degree of freedom.

The aim of this Letter is to demonstrate numericallythe consequences of topological terms in the correspond-ing field theory. We will do so by considering a model ofDirac fermions in 2+1 dimensions with reduced spatialsymmetries such that the three AFM and one of the twoKVBS mass terms are dynamically generated. Contraryto the generic KVBS state with a spontaneously bro-ken U(1) symmetry in the continuum, our KVBS statespontaneously breaks a Z2 symmetry. We will refer tothis state as Z2 KVBS. Starting from the WZW topo-logical term, this symmetry reduction amounts to set-ting one component of the five-dimensional field to zero.This maps the WZW term to a θ-term at θ = π [3].Let us assume that the phase transition observed nu-merically between the AFM and the Z2 KVBS is con-tinuous and captured by the aforementioned field the-

ory. Then, the θ-term leads to the prediction that in theZ2 KVBS phase close to the transition, a Z2 KVBS do-main wall harbors a spin-1/2 chain. In what follows, wewill provide a model—amenable to large scale negative-sign-free auxiliary-field quantum Monte Carlo (QMC)calculations—that provides compelling results support-ing this field-theory picture.

Field theory.—A theory that accounts for the phase di-agram presented in Ref. [6] (see Fig. 1(a)) contains Diracfermions Yukawa-coupled to the AFM and Z2 KVBSmass terms, as described by

LF= Ψ†

[

∂µ(12 ⊗ γ0γµ) +

(

ηαχ

)

·

(

σα ⊗ γ012 ⊗ iγ0γ5

)]

Ψ .

(1)

Here, the Dirac spinors Ψ† carry a sublattice index, a spinindex, and a valley index. The γ-matrices act on the val-ley and sublattice spaces and satisfy the Clifford algebra{γa, γb} = 2δab. σα with α = 1, 2, 3 denote the Paulispin-1/2 matrices. The fact that the SO(3) AFM massterms σα ⊗ γ0 and Z2 KVBS mass terms 12 ⊗ iγ0γ5 anti-commute results in an SO(4) invariance of the fermionicaction: a global SO(4) rotation of the four-componentfield φ = (φ1, φ2, φ3, φ4) = (η, χ) is equivalent to acanonical transformation of the fermion operators. Thedynamics of the field is governed by a four-component ϕ4

action, LB. While LF has SO(4) symmetry, LB inheritsthe SO(3) × Z2 symmetry of the lattice model.The Lagrangian L = LF + LB can account for many

phase transitions. The Gross-Neveu transitions fromsemimetal to AFM or from semimetal to Z2 KVBS in-volve a closing of the mass gap corresponding to the normof the field φ. On the other hand, QMC simulations (seeRef. [6] and the Supplemental Material (SM)) point toa continuous transition between the AFM and Z2 KVBSstates with an emergent SO(4) symmetry. Importantly,the numerical results show that the single-particle gapremains finite across the transition. In the field theory,this implies that amplitude fluctuations of φ are frozen

Page 2: Kekul´e valence-bond solid (KVBS) masses map onto a field ...

2

and only phase fluctuations of the field need to be re-tained. Since the fermions remain massive, they can beintegrated out (in the large mass limit) to obtain

S =

dx2dτ1

g(∂uφ)

2 + iθQ , θ = π (2)

with

Q =1

12π2

dx2dτǫi,j,kǫα,β,γ,δφα∂iφβ∂j φγ∂kφδ. (3)

Here, φ(x, τ) = φ/|φ| defines a mapping from 2 + 1 di-mensional Euclidean space-time to the three-dimensionalsphere S3. For smooth field configurations with no sin-gularities at infinity Q is quantized to integer values andcorresponds to the winding of the unit four-vector φ onthe hypersphere in four dimensions, S3.We now consider a smooth domain wall of the Z2

KVBS order parameter. Such a configuration can be ob-tained by pinning the field φ at the origin and at infinity:φ(0) = (0, 0, 0, 1) and φ(∞) = (0, 0, 0,−1). Specifically,let us parameterize 2+1 dimensional Euclidean space-time with spherical coordinates, (x, τ) = rn with n aunit vector, and choose

φ(rn) = (sin(f(r))n, cos(f(r)) . (4)

Here, f is a one-to-one smooth function with boundaryconditions f(0) = 0 and f(∞) = π and describes theprofile of the domain wall. As shown in the SM, theintegration over r can now be carried out to obtain thedomain-wall action:

SDW =

dxdτ1

g(∂un)

2 + iπQDW (n) (5)

with

QDW (n) =1

dxdτn · ∂τn× ∂xn. (6)

Above we have mapped S2 (on which n is defined) toR

2. While the topological term is independent of thechoice of the profile of the domain wall, g depends on f .The action in Eq. (5) corresponds to that of the spin-1/2 Heisenberg chain [14, 15]. Thereby, the topologicalθ-term at θ = π has the important consequence that adomain wall of the Z2 KVBS order parameter harbors aspin-1/2 Heisenberg chain.

Model.—The QMC simulations presented in Ref. [6]for the honeycomb lattice support a direct and continu-ous transition between the AFM and Z2 KVBS with anemergent SO(4) symmetry and, in principle, provide acase to test the above predictions. However, irrespec-tive of how one places the pinning fields on the hon-eycomb lattice, translation symmetry along the domainwall will be broken. Since gaplessness of the spin-1/2chain is protected by a mixed anomaly between trans-lations and time-reversal or spin rotations, dimerization

(a) (b)

(c)

4

6

-8 -6 -4 -2 0 2 4 6 8

i 1

i2

-0.4

0 0.4

1/h01/hc

U

Uc

π

π

π

π

eiπ/2eiπ/2

1

1

a1

a2

ξ −ξ

c

Semimetal

AFM

Z2KVBS

FIG. 1. (a) Schematic ground-state phase diagram withsemimetallic, AFM, and Z2 KVBS phases [6]. Results cor-respond to scans along the dashed line. (b) In our model,fermions acquire a π-flux when circulating around a plaque-tte. Bonds labeled by a solid circle accommodate an Isingspin that couples to the fermions with magnitude ±ξ. Weconsider periodic (open) boundary conditions in the a1 (a2)directions and freeze the Ising spins on the open boundaryto impose a domain wall. (c) Real-space bond energy change

∆Bi,al(see text). Here, L1 = 30 and L2 = 17.

along the domain wall will occur. Hence, even if a spinchain emerges at the domain wall, it will gap out dueto the choice of lattice discretization. To avoid this, wehave reformulated the model of Ref. [6] on the π-fluxsquare lattice. This provides a lattice discretization ofDirac fermions with a C4 symmetry (as opposed to C3 forthe honeycomb lattice). The model Hamiltonian readsH = Hf + Hs + Hfs (see Fig. 1(b)) with

Hf=∑

〈ij〉,σ

tij c†iσ cjσ + U

i

(ni↑ −12 )(ni↓ −

12 ), (7)

Hs= J∑

〈ij,kl〉

szij szkl − h

〈ij〉

sxij , Hfs =∑

〈ij〉,σ

tijξij szij c

†iσ cjσ.

While Hf corresponds to the half-filled Hubbard modelon the π-flux square lattice, Hs is a ferromagnetic,transverse-field Ising model defined on the bonds 〈ij〉 ofthe square lattice. Hfs accounts for the coupling betweenDirac fermions and Ising spins. The Hubbard interactionand the fermion-spin coupling can dynamically generateSO(3) AFM order and Z2 KVBS order (ferromagneticorder of the Ising spins), respectively. For the numeri-cal simulations we used the ALF (Algorithms for LatticeFermions) implementation [16] of the well-establishedfinite-temperature auxiliary-field QMC method [17, 18].Our model can be simulated without encountering thenegative sign problem. Henceforth, we use t = 1 as theenergy unit, set J = −1, ξ = 0.5, and U = 7. An in-verse temperature β = 30 (with Trotter discretization

Page 3: Kekul´e valence-bond solid (KVBS) masses map onto a field ...

3

∆τ = 0.1) yields results representative of the groundstate. QMC results on torus geometries detailed in theSM suggest a continuous AFM–Z2 KVBS transition withemergent SO(4) symmetry at 1/hc ≈ 0.270.

To pin a domain wall configuration, we consider acylindrical geometry and freeze the Ising spins at theedges to sz(i1,−n),(i1,−n+1) = 1 and sz(i1,n−1),(i1,n)

= −1where L2 = 2n + 1. Importantly, and taking into ac-count the gauge freedom to define the π-flux model,translation symmetry by a1 is present. The model withpinning fields has a mirror symmetry corresponding tothe combined transformations c†(i1,i2),σ → c†(i1,−i2),σ

and

sz(i1,i2),(i1,i2+1) → −sz(i1,i2),(i1,−i2−1).

Numerical results.—To detect the profile of the do-main wall, we measure the bond kinetic energy ∆Bi,al

=

〈Bi,al〉 − 〈B〉. Here, Bi,al

=∑

σ ti,i+al(c†iσ ci+alσ

+

c†i+alσciσ) and B = (2L1L2)

−1∑

i,l Bi,alwhere l = 1, 2.

Figure 1(c) shows this quantity. The aforementionedtranslation and mirror symmetries are readily seen.

As discussed above, the field theory of the domain wallis described by an SO(3) non-linear sigma model withθ-term at θ = π in 1+1 dimensions. We expect thistheory to have an emergent SO(4) [19, 20] symmetry re-flecting the fact that spin-spin and dimer-dimer correla-tions decay with the same power law but with differentlogarithmic corrections: (−1)r(ln r)1/2r−1 for the spin[21–23] and (−1)r(ln r)−3/2r−1 for the dimer [23]. In

Figs. 2(a-c) we plot the spin [CS(i) = 〈Si · S0〉], dimer

[CD(i) = 〈(Di−〈Di〉)·(D0−〈D0〉)〉] and bond [CB(i) =〈(Bi,a1

−〈Bi,a1〉)·(B0,a1

−〈B0,a1〉)〉] correlators as a func-

tion of the conformal distance x = L1sin(πi1/L1) [24].

Here, Si =∑

σσ′ c†iσσσσ′ ciσ′ , and Di = Si · Si+a1

.The bond and dimer correlations share the same sym-metries so that we expect them to decay with the samepower law. We compare our results with those for thehalf-filled Hubbard chain at U/t = 4. While there is re-markable agreement between the spin correlations (seeFig. 2(a)) it appears that we have to reach longer lengthscales in the domain wall calculation to observe the 1/rpower law decay for the dimer and bond correlations (seeFig. 2(b),(c)). A possible interpretation of these numer-ical results is that the Hubbard model is closer to theemergent SO(4) conformal field theory than the domainwall SO(3) theory of Eq. (5).

The field theory interpretation of the domain wall hasconsequences. It should be independent of the choice ofthe lattice discretization—provided that it does not breakrelevant symmetries such as translation along the domainwall—and the lattice constant should correspond to ahigh-energy scale. In Fig 2(d), we check that the domainwall extends over many lattice sites in the perpendiculardirection. In particular, for the value of transverse field hconsidered, the domain wall extends over several latticespacings and the data are consistent with CS(x, i2) ∼e−|i2|/ξx−1 where ξ ∼ 4. Varying 1/h in the Z2 KVBS

10-5

10-4

10-3

10-2

10-1

100

L1=14L1=18L1=22L1=26L1=30

(a) (-1)xCS(x,0)

L1=14L1=18L1=22L1=26L1=30

(b) (-1)xCD(x,0)

L1=14L1=18L1=22L1=26L1=30

10-5

10-4

10-3

10-2

10-1

100

4 6 8 10 30x

(c) (-1)xCB(x,0)

x

(d) (-1)xC

S(x,i2)

4 6 8 10 30

i2 =0i2 =1i2 =2i2 =3i2 =4i2 =5

i2 =6i2 =7i2 =8

FIG. 2. Real-space correlation functions of the model (7) for(a) spin, (b) dimer, and (c) bond at the domain wall of Z2

KVBS (solid symbols). Here L2 = 17 and x = L1sin(πi1/L1)is the conformal distance [24]. Open symbols indicates QMCresults for the one-dimensional Hubbard model with U/t = 4and at half filling. (d) Real-space correlation functions of themodel (7) for spin along the domain wall. Here, L1 = 30 andL2 = 17. Solid line denotes (−1)xx−1.

will merely change the profile of the domain wall, therebychanging the length scale ξ but not the properties of thespin-1/2 chain along the domain wall (see SM). If 1/hdrops below 1/hc into the AFM phase, spinons will bindand we expect long-range order to develop within thedomain wall. Calculations confirming this point of viewcan be found in the SM.

Further evidence for an emergent spin-1/2 Heisenbergchain can be obtained from the dynamical spin struc-ture factor at the domain wall of the Z2 KVBS shown inthe left panel of Fig. 3. The key feature of the spin-1/2chain is that the low-lying excitations are well describedby the two-spinon continuum revealed by the dynami-cal spin structure factor. In the thermodynamic limit,such excitations have a support in the wave vector, q,versus frequency, ω, plane with lower and upper boundsgiven by ωl(q) ∼ π/2| sin(q)| and ωu(q) ∼ π| sin(q/2)|,respectively. Experimental as well as theoretical cal-culations of the dynamical spin structure factor canbe found in Ref. [25]. Figure 3 shows the results forthe dynamical spin structure factor CS(q, ω) at the do-main wall of the Z2 KVBS. In our QMC simulations,CS(q, ω) was obtained via the analytic continuation of

Page 4: Kekul´e valence-bond solid (KVBS) masses map onto a field ...

4

0

1

2

3

0 π/2 π

ω

2π3π/2q

0.01

0.1

1

10

FIG. 3. Dynamical spin structure factor CS(q, ω) at the do-main wall of the Z2 KVBS (left panel). Here, L1 = 26 andL2 = 17. From the calculation of the single-particle gap ontorus geometries (see the SM) we estimate the particle-holecontinuum to lie above ω ≃ 1.5. Right panel: dynamical spinstructure factor for the Hubbard model on a 46-site chain atU/t = 4 and at half filling.

the imaginary-time-displaced spin correlation functionsCS(q, τ) = 〈Sq(τ) · S−q(0)〉 = L1

−1 ∑

i1eiqi1 〈S(i1,0)(τ) ·

S0(0)〉 at i2 = 0. Below the particle-hole continuum, es-timated to start at ω > 1.5t, our QMC results reproducethe well-known features of the two-spinon continuum [seethe right panel of Fig. 3].Summary and discussion.—We have shown how to

probe topological terms in lattice realizations of field the-ories by pinning defects. The explicit example providedin this work is based on a model where the effective fieldtheory has a θ-term at θ = π in 2+1 dimensions withemergent SO(4) symmetry. In the lattice realization ofthis model, the SO(4) symmetry reduces to SO(3)× Z2

and we consider a domain wall of the Z2 field. The emer-gent SO(4) symmetry then suggests that the domain wallharbors a spin-1/2 chain. Our numerical results confirmthis point of view.The above argument holds for continuous transitions

with emergent symmetries. There is an ongoing de-bate on the nature of the generic DQCP between AFMand VBS [26–28] (or quantum spin Hall (QSH) and s-wave superconducting (SSC) [29]) orders with emergentSO(5) symmetry [30]. Compelling evidence for a con-tinuous transition as well as emergent SO(5) symmetryfrom finite-size calculations has been put forward. How-ever the critical exponents stand at odds with the boot-strap bounds [31]. To resolve this apparent contradic-tion, one can conjecture [32–34] that an SO(5) conformalfield theory indeed exists in spatial dimensions slightlygreater than two that, however, collides with anotherfixed point and becomes complex upon tuning the di-mension down to two. Proximity to fixed-point collisionis at the origin of a very slow renormalization group flowand associated very long correlation lengths [32, 35]. Infact, recent simulations of the SO(5) non-linear sigmamodel with a WZW topological term support this point

of view [36]. Very similar arguments can be applied tothe present case where weakly first-order transitions werereported for similar symmetry classes [37–39]. Hence, aweakly first-order transition does not impair the notionthat topological terms can play a dominant role at inter-mediate length scales.

Our approach can also be applied to other models. Forthe AFM-VBS transition, pinning a C4 vortex on a sys-tem with open boundary conditions should result in aspinon that can be probed via the spin susceptibility. Inthe context of the QSH-SSC transition of Ref. [29] it ispossible to pin a skyrmion of the O(3) QSH order pa-rameter. Since topology states that the skyrmion carriescharge 2e [5] this should result in a doping of the system.

Our observation may have some utility in three dimen-sions, where there are also pairs of ordered phases forwhich the disorder operators for one phase are chargedunder the symmetry broken by the other. A simple ex-ample involves a cubic-lattice AFM and a cubic-latticeVBS [40]. As in two dimensions, the skyrmions carrylattice-symmetry quantum numbers, and the defects ofthe VBS pattern (which are hedgehogs) carry spin. Thiscan be encoded in a sigma model with a WZW term,now with softly-broken SO(6) ⊃ SO(3)AFM × SO(3)VBS

symmetry. However, in addition to the usual possibilityof a first-order transition [41], a direct transition betweenthese two phases can also be preempted by an intermedi-ate disordered phase for the following reason: in contrastto two dimensions, compact abelian gauge theory withsmall amounts of charged matter (QED) has a (famil-iar) deconfined phase in three dimensions. But, as in theabove discussion, the WZW term still has consequenceswithin the ordered phases. For definiteness and similar-ity with our example above, consider breaking the cubiclattice symmetry down to Z2 × Z2, where the second Z2

represents reflections in z, say. The associated sigmamodel then has SO(5) ⊃ SO(3)AFM × (Z2 ×Z2)VBS sym-metry with a θ-term at θ = π. In analogy to the caseconsidered here, the domain wall of the Z2 part of theVBS order parameter transverse to the z direction willhost an SO(4) non-linear sigma model at θ = π, now in2+ 1 dimensions. This is a description of the deconfinedcritical point between AFM and KVBS orders in two di-mensions. Along a domain wall of the VBS pattern, spinand VBS correlations are predicted to be long-ranged,with the same exponents.

Another possibility is to break the cubic lattice sym-metry down to C4 × Z2, where again the Z2 representsreflections in z. The associated sigma model then hasSO(6) ⊃ SO(3)AFM × (SO(2)× Z2)VBS symmetry with aWZW term. Now, the domain wall of the Z2 part of theVBS order parameter, transverse to the z direction, willhost an SO(5) non-linear sigma model with a WZW term,now in 2+1 dimensions. This construction could providean alternative for the Landau-level projection formula-tion of this theory [36, 42]. This is a description of the

Page 5: Kekul´e valence-bond solid (KVBS) masses map onto a field ...

5

DQCP between AFM and VBS.

We thank J. S. Hofmann, M. Oshikawa, Z. Wang, C.Xu and Yi-Zhuang You for insightful discussions. Wethank the Gauss Centre for Supercomputing (SuperMUCat the Leibniz Supercomputing Centre) for generous al-location of supercomputing resources. The research hasbeen supported by the Deutsche Forschungsgemeinschaftthrough grant numbers AS 120/15-1 (TS), AS120/14-1(FFA), the Wurzburg-Dresden Cluster of excellence onComplexity and Topology in Quantum Matter - ct.qmat(EXC 2147, project-id 39085490) (FFA) and SFB 170 To-CoTronics (MH). TG is supported by the National Sci-ence Foundation under Grant No. DMR-1752417, and asan Alfred P. Sloan Research Fellow. FFA and TG thankthe BaCaTeC for partial financial support.

[1] A. Abanov and P. Wiegmann,Nuclear Physics B 570, 685 (2000).

[2] A. Tanaka and X. Hu,Phys. Rev. Lett. 95, 036402 (2005).

[3] T. Senthil and M. P. A. Fisher,Phys. Rev. B 74, 064405 (2006).

[4] J. Lee and S. Sachdev,Phys. Rev. Lett. 114, 226801 (2015).

[5] T. Grover and T. Senthil,Phys. Rev. Lett. 100, 156804 (2008).

[6] T. Sato, M. Hohenadler, and F. F. Assaad,Phys. Rev. Lett. 119, 197203 (2017).

[7] Z.-X. Li, Y.-F. Jiang, S.-K. Jian, and H. Yao,Nature Communications 8, 314 (2017).

[8] M. Ippoliti, R. S. K. Mong, F. F. Assaad, and M. P.Zaletel, Phys. Rev. B 98, 235108 (2018).

[9] Y. Liu, Z. Wang, T. Sato, M. Hohenadler,C. Wang, W. Guo, and F. F. Assaad,Nature Communications 10, 2658 (2019).

[10] S. Ryu, C. Mudry, C.-Y. Hou, and C. Chamon,Phys. Rev. B 80, 205319 (2009).

[11] T. Senthil, L. Balents, S. Sachdev, A. Vishwanath, andM. P. A. Fisher, Phys. Rev. B 70, 144407 (2004).

[12] T. Senthil, A. Vishwanath, L. Balents, S. Sachdev, andM. P. A. Fisher, Science 303, 1490 (2004).

[13] M. Levin and T. Senthil,Phys. Rev. B 70, 220403 (2004).

[14] F. D. M. Haldane, Phys. Rev. Lett. 50, 1153 (1983).[15] C. Mudry, Lecture Notes on Field Theory in Condensed

Matter Physics (World Scientific Publishing Company,2014).

[16] M. Bercx, F. Goth, J. S. Hofmann, and F. F. Assaad,SciPost Phys. 3, 013 (2017).

[17] R. Blankenbecler, D. J. Scalapino, and R. L. Sugar,Phys. Rev. D 24, 2278 (1981).

[18] F. Assaad and H. Evertz, in Computational Many-

Particle Physics, Lecture Notes in Physics, Vol. 739,edited by H. Fehske, R. Schneider, and A. Weiße(Springer, Berlin Heidelberg, 2008) pp. 277–356.

[19] I. Affleck, Phys. Rev. Lett. 55, 1355 (1985).[20] I. Affleck and F. D. M. Haldane,

Phys. Rev. B 36, 5291 (1987).

[21] I. Affleck, D. Gepner, H. J. Schulz, and T. Ziman,J. Phys. A: Math. Gen. 22, 511 (1989).

[22] R. R. P. Singh, M. E. Fisher, and R. Shankar,Phys. Rev. B 39, 2562 (1989).

[23] T. Giamarchi and H. J. Schulz,Phys. Rev. B 39, 4620 (1989).

[24] J. Cardy, Scaling and Renormalization in Statistical

Physics (Cambridge University Press, 1996).[25] B. Lake, D. A. Tennant, J.-S. Caux, T. Barthel,

U. Schollwock, S. E. Nagler, and C. D. Frost,Phys. Rev. Lett. 111, 137205 (2013).

[26] H. Shao, W. Guo, and A. W. Sandvik,Science 352, 213 (2016).

[27] A. Nahum, J. T. Chalker, P. Serna, M. Ortuno, andA. M. Somoza, Phys. Rev. X 5, 041048 (2015).

[28] A. W. Sandvik and B. Zhao, arXiv:2003.14305 (2020),arXiv:2003.14305 [cond-mat.str-el].

[29] Y. Liu, Z. Wang, T. Sato, M. Hohenadler,C. Wang, W. Guo, and F. F. Assaad,Nature Communications 10, 2658 (2019).

[30] A. Nahum, P. Serna, J. T. Chalker, M. Ortuno, andA. M. Somoza, Phys. Rev. Lett. 115, 267203 (2015).

[31] D. Poland, S. Rychkov, and A. Vichi,Rev. Mod. Phys. 91, 015002 (2019).

[32] C. Wang, A. Nahum, M. A. Metlitski, C. Xu, andT. Senthil, Phys. Rev. X 7, 031051 (2017).

[33] A. Nahum, arXiv:1912.13468 (2019),arXiv:1912.13468 [cond-mat.str-el].

[34] R. Ma and C. Wang, arXiv:1912.12315 (2019),arXiv:1912.12315 [cond-mat.str-el].

[35] I. F. Herbut and L. Janssen,Phys. Rev. Lett. 113, 106401 (2014).

[36] Z. Wang, Michael P. Zaletel, R. S. K. Mong,and F. F. Assaad, arXiv:2003.08368 (2020),arXiv:2003.08368 [cond-mat.str-el].

[37] E. Torres, L. Weber, L. Janssen, S. Wessel,and M. M. Scherer, arXiv:1911.01244 (2019),arXiv:1911.01244 [cond-mat.str-el].

[38] P. Serna and A. Nahum, arXiv:1805.03759 (2018),arXiv:1805.03759 [cond-mat.str-el].

[39] B. Zhao, P. Weinberg, and A. W. Sandvik,Nature Physics 15, 678 (2019).

[40] P. Hosur, S. Ryu, and A. Vishwanath,Phys. Rev. B 81, 045120 (2010).

[41] M. S. Block and R. K. Kaul,Phys. Rev. B 86, 134408 (2012).

[42] M. Ippoliti, R. S. K. Mong, F. F. Assaad, and M. P.Zaletel, Phys. Rev. B 98, 235108 (2018).

[43] C. Wu and S.-C. Zhang,Phys. Rev. B 71, 155115 (2005).

[44] K. Binder, Z. Phys. B Con. Mat. 43, 119 (1981).[45] S. Pujari, T. C. Lang, G. Murthy, and R. K. Kaul,

Phys. Rev. Lett. 117, 086404 (2016).[46] I. F. Herbut, V. Juricic, and B. Roy,

Phys. Rev. B 79, 085116 (2009).[47] F. F. Assaad and D. Wurtz,

Phys. Rev. B 44, 2681 (1991).

Page 6: Kekul´e valence-bond solid (KVBS) masses map onto a field ...

6

SUPPLEMENTAL MATERIAL

In this supplemental material we will first derive thedomain wall action: a (1+1) dimensional SO(3) non-linear sigma model with θ-term at θ = π. Next we will re-view our auxiliary-field quantum Monte Carlo (AFQMC)simulations on the torus. They support the point of viewthat—on the accessible lattice sizes–our model indeedshows a continuous transition between an SO(3) anti-ferromagnet (AFM) and Z2 KVBS with emergent SO(4)symmetry. We then provide more data on the profile ofthe domain wall and spin dynamics along the domainwall when tuning through the transition. Finally, we willshow that AFQMC simulations of the one-dimensionalrepulsive Hubbard model at half filing reproduce the ex-pected power-law decay of real-space spin, dimer, andbond correlation functions.

SPIN-1/2 CHAIN PINNED AT THE Z2 DOMAINWALL: A FIELD THEORETIC APPROACH

In this section, we will detail the calculation that showsthat starting from the θ-term at θ = π, a domain wall ofZ2 KVBS nucleates a spin-1/2 chain. Our starting pointis the field theory discussed in the main text:

S =

dx2dτ1

g(∂uφ)

2 + iθQ , θ = π (8)

with

Q =1

12π2

dx2dτǫi,j,kǫα,β,γ,δφα∂iφβ∂j φγ∂kφδ. (9)

Here, φ = φ(x, τ) defines a mapping between the 2 + 1dimensional Euclidean space-time to the unit sphere infour dimensions: S3. For any smooth field configuration,Q counts the winding of the unit four-vector φ = φ/|φ|on the hypersphere S3 and takes an integer value. Thevalue of θ is thereby of great importance. One can startfrom the Wess-Zumino-Witten term that appears in asetting where the Dirac fermions couple symmetricallyto a quintuplet of anti-commuting mass terms and setone mass to zero. A calculation will then lead to theaforementioned θ-term at θ = π [3]. Here, our aim is toconsider a domain-wall configuration. For the calculationto be well defined, we have to make sure that the fieldconfiguration φ(x, τ) has no singularities. To achievethis goal, it is useful to include the point at infinity inEuclidean space, so that the base space is topologically athree-sphere. We can specify coordinates by embeddingthe three-sphere into R

4 Euclidean space by

x1 = cos(ϕ1)

x2 = sin(ϕ1) cos(ϕ2)

x3 = sin(ϕ1) sin(ϕ2) cos(ϕ3)

x4 = sin(ϕ1) sin(ϕ2) sin(ϕ3) (10)

where ϕ1 ∈ [0, π], ϕ2 ∈ [0, π], and ϕ3 ∈ [0, 2π]. Sincethe topological term is independent of the choice of themetric we obtain with this parameterization:

Q =1

12π2

∫ π

0

dϕ1

∫ π

0

dϕ2

∫ 2π

0

dϕ3

ǫi,j,kǫα,β,γ,δφα∂iφβ∂j φγ∂kφδ (11)

with ∂i =∂

∂ϕi

. With this compactification of R3 we candefine a smooth domain-wall configuration as:

φ4 = cos(f(ϕ1))

φ3 = sin(f(ϕ1))n3

φ2 = sin(f(ϕ1))n2

φ1 = sin(f(ϕ1))n1 (12)

Here n(ϕ2, ϕ3) = (n1, n2, n3). From |φ| = 1 follows that|n| = 1. ϕ2 and ϕ3 define a point on the unit sphere inthree dimensions, S2, that is isomorphic to R

2. Hence,n defines a mapping from R

2 to the unit sphere in R3.

We can now return to φ. The function f(ϕ1) definesa generic domain wall and is required to satisfy the fol-lowing properties. f(0) = 0, f(π) = π and f(ϕ1) is aone-to-one mapping from [0, π] to [0, π]. At infinity in

R3, corresponding to ϕ1 = 0, φ = (0, 0, 0, 1) and at the

origin, ϕ1 = π, φ = (0, 0, 0,−1). The pinning of φ atthe origin and at infinity defines a Z2 domain wall in thefourth component of the field. This choice of the field φ

explicitly breaks SO(4) symmetry down to SO(3) corre-sponding to rotations of the n vector.

For the domain-wall field configuration of Eq. 12, wecan evaluate the θ-term by explicitly carrying out theintegration over ϕ1. Since f(ϕ1) is a one-to-one mappingand Q is a topological quantity, the value of Q for thedomain wall field of Eq. 12 is identical to the deformeddomain wall:

φ4 = cos(ϕ1)

φ3 = sin(ϕ1)n3

φ2 = sin(ϕ1)n2

φ1 = sin(ϕ1)n1. (13)

Inserting this result in Q and evaluating the integral overϕ1 gives

QDW =1

∫ π

0

dϕ2

∫ 2π

0

dϕ3 n · ∂ϕ2n× ∂ϕ3

n. (14)

In summary, the domain-wall action reads

SDW =

dxdτ1

g(∂un)

2 + iπQDW, (15)

which corresponds to the coherent spin state path inte-gral of the spin-1/2 chain [14, 15].

Page 7: Kekul´e valence-bond solid (KVBS) masses map onto a field ...

7

CONTINUOUS AFM-Z2 KVBS TRANSITIONWITH EMERGENT SO(4) SYMMETRY

As mentioned in the main text, our calculations in [6]were carried out on the honeycomb lattice. Here, we mapout the bulk phase diagram on a torus geometry for theπ-flux lattice and show that the model has a similar phasediagram.We consider a model of Dirac fermions in 2+1 dimen-

sions with Hamiltonian H = Hf + Hs + Hfs (see Fig.1(b)of the main text):

Hf=∑

〈ij〉,σ

tij c†iσ cjσ + U

i

(ni↑ −12 )(ni↓ −

12 ),

Hs= J∑

〈ij,kl〉

szij szkl − h

〈ij〉

sxij ,

Hfs=∑

〈ij〉,σ

tijξij szij c

†iσ cjσ. (16)

Here Hf corresponds to the half-filled Hubbard model onthe π-flux square lattice. ciσ(c

†iσ) is the fermionic anni-

hilation (creation) operator at site i with spin σ =↑, ↓,

and niσ ≡ c†iσ ciσ. Hs is a ferromagnetic, transverse-fieldIsing model where the Ising spins—described by the Paulispin operators sαi,j(α = x, y, z)—live on the bonds con-necting fermionic sites i and j. J is the ferromagneticnearest-neighbor interaction and h the transverse field.Hfs accounts for the coupling (ξij = ±ξ) between Dirac

fermions and Ising spins. Our model Hamiltonian H hasan SU(2) spin symmetry as well as a Z2 symmetry corre-sponding to invariance under the combined operation ofinversion and szij → −szij . Under inversion Hfs → −Hfs,so that the energy does not depend on the sign of ξ andthe two possible Kekule patterns related by ξ → −ξ aredegenerate. The Hubbard interaction and the fermion-spin coupling have the potential to dynamically generateSO(3) AFM order and Z2 KVBS order (ferromagneticorder of the Ising spins), respectively.We used the ALF (Algorithms for Lattice Fermions)

implementation [16] of the well-established finite-temperature auxiliary-field QMC method [17, 18]. Simu-lations of our model are free of the negative sign problem.To see this, one can first carry out a partial particle-holecanonical transformation: ci,↑ → eiQ·ic†i,↑ and ci,↓ → ci,↓with Q = (π, π)/a. This changes the sign of the Hub-bard interaction from positive to negative. One can theuse time reversal symmetry to show that the eigenval-ues of the Fermion determinant come in complex conju-gate pairs [43]. We simulated lattices with L × L unitcells (each containing four Dirac fermions and four Isingspins) and periodic boundary conditions. Henceforth, weuse t = 1 as the energy unit, set J = −1, ξ = 0.5, andU = 7. All the data are for the Trotter discretization∆τ = 0.1. In the considered parameter range, an in-verse temperature β = 30 was sufficient to obtain results

0

0.5

1

L=4

L=6

L=8

L=10

L=12

0.25 0.3

RA

FM

(a)

1/h0.25 0.3

L=4L=6L=8L=10L=12

0

0.5

1

1/h

(b)

RK

VB

S

1/h

c

0.25

0.3

0 0.1 0.2

AFMZ

2 KVBS

1/L

(c)

FIG. 4. Correlation ratios for (a) AFM and (b) Z2 KVBSstates at U = 7. (c) Extrapolation of crossing points ofcorrelation ratios for L and L + 2 gives the critical value(

hAFMc

)

−1

≈(

hKVBSc

)

−1

≈ 0.270.

representative of the ground state.The dashed line in Fig. 1 (a) of the main text indicates

the scan of the phase diagram obtained by tuning h ata fixed U = 7. Along this line we observe two insulat-ing phases corresponding to the SO(3) AFM and the Z2

KVBS.To map out the phase diagram we compute equal-

time correlation functions of fermion spin Si =∑

σσ′ c†iσσσσ′ ciσ′ , fermion bond Bij =

σ tij(c†iσ cjσ +

c†jσ ciσ), and Ising spin szij . Due to the larger unitcell, these correlation functions are 4 × 4 matrices ofthe form CO

Rγ,R′δ = 〈(ORγ − 〈ORγ〉) · (OR′δ − 〈OR′δ〉)〉

(O = S, B, s) whereR,R′ labels the unit cell and γ, δ theorbitals. After diagonalizing the corresponding structurefactors

COγδ(q) =

1

L2

RR′

CORγ,R′δe

iq·(R−R′), (17)

we calculated the renormalisation-group invariant corre-lation ratio [44, 45]

RO = 1−λ1(q0 + δq)

λ1(q0)(18)

using the largest eigenvalue λ1(q); q0 is the ordering wavevector, q0+ δq a neighboring wave vector. By definition,RO → 1 for L → ∞ in the corresponding ordered state,whereas RO → 0 in the disordered state. At the criti-cal point, RO is scale-invariant for sufficiently large L sothat results for different system sizes cross. Figures 4 (a)

Page 8: Kekul´e valence-bond solid (KVBS) masses map onto a field ...

8

0.8

0.9

1

0.25 0.3

L=4L=6L=8L=10L=12

∂F/∂h

1/h

(a)

0

0.5

1

0 0.1 0.2 0.3

1/h=0.2781/h=0.2751/h=0.2721/h=0.2701/h=0.267

∆sp

(K)

1/L

(b)

0ReM

0ImM

0 max.

1

10

0.25 0.3

L=6L=8L=10L=12

(c) (d)

σK

VB

S/σ

AF

M

1/h

FIG. 5. (a) Free-energy derivative and (b) single-particle gapat the Dirac point q = K. (c) Ratio of the standard devia-tions of the AFM and Z2 KVBS order parameters. (d) Jointprobability distribution of the two order parameters at thecritical point for L = 8. Here, U = 7 and β = 30.

and (b) show the results. Based on the assumption of adynamical critical exponent z = 1 [46] we set L = β tocarry out the finite-size scaling. The onset of the AFMis detected from the crossing of RAFM ≡ RS [Fig. 4(a)],whereas the onset of Kekule order can be detected eitherfrom RKVBS ≡ Rs [Fig. 4(b)] or from RB. The finite-sizescaling of the crossing points yields a single critical pointof h−1

c ≈ 0.270 shown in Fig. 4 (c), pointing to a directphase transition between AFM and Z2 KVBS.Figure 5 (a) plots the free-energy derivative

∂F

∂h=

1

4L2

〈ij〉

sxij

. (19)

The absence of a discontinuity in this quantity favors acontinuous transition.For the derivation of a low-energy effective field the-

ory is important to confirm that the single particle gapremains finite across the transition. We measured theimaginary-time displaced Green’s function G(q, τ) andobtained the single-particle gap ∆sp(q) from

G(q, τ) ∝ exp(−τ∆sp(q)) (20)

at large imaginary time τ . We found that the single-particle gap at the Dirac point, shown in Fig. 5(b), re-mains clearly nonzero.To verify whether the critical point has an emergent

SO(4) symmetry, we measured the standard deviations

σO =

〈O2〉 − 〈O〉2 of the AFM (σAFM ≡ σS) and Z2

KVBS (σKVBS ≡ σs) order parameters. These quantitiesare in general independent but become locked togetherand can be combined into a four-component order pa-rameter if an SO(4) symmetry—unifying the two orderparameters—emerges at the critical point. In this case,the ratio σKVBS/σAFM will become universal at the crit-ical point. This is confirmed by the result in Fig. 5(c).Moreover, the emergent SO(4) symmetry can also be con-firmed in the joint probability distribution of the two or-der parameters determined from QMC snapshots,

M = σAFMei0 + σKVBSe

2 . (21)

If the SO(4) symmetry emerges at the critical point, thequantity M (after normalization of each order parame-ter) should reveal a circular distribution, as confirmed atthe critical point by Fig. 5(d).

SPIN DYNAMICS AROUND THE Z2 DOMAINWALL WITH VARYING 1/h

In the main text we set 1/h = 0.275 that places us inthe Z2 KVBS phase close to the critical point. Here wevary 1/h across the transition and investigate the fate ofthe domain wall. In Fig. 6 we map out the profile of thedomain wall by considering ∆Bi,a2

across the domainwall (left panels) the spin correlations along the domainwall (right panels). These quantities are defined in themain text.

The profile of the domain wall in the KVBS phase isexpected to be inversely proportional to the stiffness. In-deed the QMC results shown in Figs. 6 (a)-(f) supportthat upon moving away from the critical point deep intothe Z2 KVBS phase the domain wall becomes more pro-nounced. Accordingly the width around the domain wallwhere one observes 1/r decay of the spin-spin correlationsreduces.

In the AFM phase there is no scale that confines thewidth of the domain wall other than the width of thelattice L2. We equally expect the spin-spin correlationsto show long range order. In Fig. 6 (g)-(h) we considera very large value of h = 30. As apparent from Fig. 6(g) the profile of the domain wall is very flat around thecenter of the cylinder. The spin correlations show a slightupturn but remarkably do not provide clear evidence oflong range order. We can understand this in the follow-ing way. The AFM state originates from the binding ofspinons and this will occur on a given length scale. IfL2 is comparable to this length scale, then long rangeorder will be hard to detect. We hence conjecture thatas L2 grows spin-spin correlations will develop clear signsof ordering.

Page 9: Kekul´e valence-bond solid (KVBS) masses map onto a field ...

9

(b) (-1)xC

S(x,i2)

(d) (-1)xC

S(x,i2)

(f) (-1)xC

S(x,i2)

-0.5

0

0.5

-5 0 5i2

(a) Bi,a2

-0.5

0

0.5

-5 0 5i2

(c) Bi,a2

-0.5

0

0.5

-5 0 5i2

(e) Bi,a2

-0.5

0

0.5

-5 0 5i2

(g) Bi,a2 (h) (-1)xC

S(x,i2)

4 6 8 10 3010

-5

10-4

10-3

10-2

10-1

x

4 6 8 10 3010

-5

10-4

10-3

10-2

10-1

x

i2 =0i2 =1i2 =2i2 =3i2 =4

i2 =5i2 =6i2 =7i2 =8

4 6 8 10 3010

-5

10-4

10-3

10-2

10-1

x

4 6 8 10 3010

-5

10-4

10-3

10-2

10-1

x

i2 =0i2 =1i2 =2i2 =3i2 =4

i2 =5i2 =6i2 =7i2 =8

i2 =0i2 =1i2 =2i2 =3i2 =4

i2 =5i2 =6i2 =7i2 =8

i2 =0i2 =1i2 =2i2 =3i2 =4

i2 =5i2 =6i2 =7i2 =8

FIG. 6. Real-space bond energy change ∆Bi,a2at i = (i1 =

0, i2) (left panel) and real-space spin-spin correlation func-tions along the domain wall of Z2 KVBS CS(x, i2) (rightpanel) for L1 = 30 and L2 = 17. Here x = L1sin(πi1/L1)is the conformal distance. Parameters used are 1/h = 0.333[(a) and (b) in the Z2 KVBS phase], 1/h = 0.299 [(c) and (d)in the Z2 KVBS phase], 1/h = 0.275 [(e) and (f) in the Z2

KVBS phase close to the critical point], and 1/h = 0.033 [(g)and (h) in the AFM phase]. Here, U = 7 and β = 30.

ONE DIMENSIONAL HUBBARD MODEL ATHALF FILLING

In this section, we present QMC results of the one-dimensional repulsive Hubbard model at half filling. TheHamiltonian is

H = t∑

i,σ

c†iσ ci+1σ +H.c.+ U∑

i

(ni↑ −12 )(ni↓ −

12 ),

(22)

where t is the nearest-neighbor hopping amplitude andU is the Hubbard repulsion. ciσ(c

†iσ) is the fermionic an-

10-4

10-3

10-2

10-1

100

10 100

L=30L=110L=190L=290L=410

(a) (-1)xCS(x,0)

10 100

L=30L=110L=190L=290L=410

10-4

10-3

10-2

10-1

100

(b) (-1)xCD(x,0)

10-4

10-3

10-2

10-1

100

10 100

L=30L=110L=190L=290L=410

(c) (-1)xCB(x,0)

0.5

1

1.5

0 100 200

O=SO=DO=B

L/2

(d) bO

( ,L/2)

xx

x

FIG. 7. Real-space correlation functions of the one-dimensional repulsive Hubbard model at half filling for (a)spin, (b) dimer, and (c) bond. Here x = Lsin(πr/L) is theconformal distance. Solid line correspond to (−1)xx−1. (d)bS(q, L/2), bD(q, L/2), and bB(q, L/2) at q = π (see text).

nihilation (creation) operator at site i with spin σ =↑, ↓,

and niσ ≡ c†iσ ciσ. For the numerical simulations we usedthe ALF (Algorithms for Lattice Fermions) implementa-tion [16] of the well-established auxiliary-field quantumMonte Carlo (AFQMC) method [17, 18]. We carried outground-state simulations with the projective AFQMC al-gorithm, which is based on the equation

〈O〉H = limΘ→∞

〈ψT |e−Θ

2HOe−

Θ

2H |ψT 〉

〈ψT |e−ΘH |ψT 〉, (23)

where Θ is a projection parameter. The trial wave func-tion |ψT 〉 is chosen to correspond to the ground stateof the noninteracting Hamiltonian. We simulated latticesizes ranging from L = 14 to 510 with periodic boundaryconditions. Henceforth, we use t = 1 as the energy unitand set U = 4. All the data were obtained for Θ = 30(and Trotter discretization ∆τ = 0.1), sufficient to ob-tain results representative of the ground state for theparameters considered.We measured real-space correlation functions of spin

Si =∑

σσ′ c†iσσσσ′ ciσ′ , dimer Di = Si · Si+1, and bond

Bi =∑

σ(c†iσ ci+1σ + c†i+1σ ciσ),

CS(r)= 〈Sr · S0〉, (24)

CD(r)= 〈(Dr − 〈Dr〉) · (D0 − 〈D0〉)〉, (25)

CB(r)= 〈(Br − 〈Br〉) · (B0 − 〈B0〉)〉. (26)

Page 10: Kekul´e valence-bond solid (KVBS) masses map onto a field ...

10

The ground state of the SU(2) symmetric isotropic spin-1/2 Heisenberg chain is conformally invariant. Spincorrelation functions scale as ∼ (−1)r(ln r)1/2r−1, withmultiplicative logarithmic corrections [21–23]. Ow-ing to emergent SO(4) symmetry [19, 20], dimercorrelation functions exhibit the same exponent butdifferent logarithmic corrections; they decay as ∼(−1)r(ln r)−3/2r−1 [23]. Here we consider the half-filledHubbard Hamiltonian that maps onto the Heisenbergmodel.

Figures 7(a) and (b) show QMC results for spin anddimer correlation functions, which indicate that the cor-relation functions both yield consistent power-law decayat large conformal distances, x = Lsin(πr/L). The bondcorrelations have the same symmetry properties as thedimer correlations so that we expect the same power-lawdecay. This can be confirmed by the QMC results shownin Fig. 7 (c). The exponent of the power-law decay canbe detected using the quantity [47]

bO(q, L/2)≡ CO(q)

−CO(q + 2π/L)− CO(q − 2π/L), (27)

where CO(q) corresponds to the static structure fac-tor of the local observable O. One can show thatthe real space correlations at distance CO(L/2) =14L

q eiqL/2bO(q, L/2). This formula singles out the

wave numbers where the static structure factor showsnon-analytical behavior. In our case, this is the case atq = π so that

CO(L/2) ∝1

LeiπL/2bO(q = π, L/2). (28)

Assuming that CO(L/2) ∝ αeiπL/2(L/2)−K(ln(L/2))γ

then

ln(bO(π, L/2)) ∝ α′ + (−K + 1) ln(L/2) + γ ln(ln(L/2)).(29)

Figure 7(d) plots this quantity for spin, dimer, and bondcorrelations as a function of L/2. All three quantitiesare nearly constant for large values of L/2, thus confirm-ingK = 1. Differences in bD(π, L/2) and bS(π, L/2) arisefrom the third term on the right-hand side of Eq. (29), re-flecting the different logarithmic corrections. Since bondcorrelations have the same symmetry properties as thedimer correlations, the logarithmic corrections are ex-pected to be consistent.