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

Instability of the UU(1) spin liquid with a large spinon Fermi surface in the Heisenberg-ring exchange model on the triangular lattice

Jianhua Yang and Tao Li Department of Physics, Renmin University of China, Beijing 100872, P.R.China
Abstract

It is widely believed that the U(1)U(1) spin liquid with a large spinon Fermi surface(SFS state) can be realized in the spin-12\frac{1}{2} J1J4J_{1}-J_{4} model on the triangular lattice, when the ring exchange coupling J4J_{4} is sufficiently strong to suppress the 120 degree magnetic ordered state. This belief is supported by many variational studies on this model and seems to be consistent with the observations on the organic spin liquid materials such as κ\kappa-(BEDT-TTF)2Cu2(CN)3 and EtMe3Sb[Pd(dmit)2]2, which are systems close to their Mott transition and thus have large J4J_{4}. Here we show through systematic variational search that such a state is never favored in the J1J4J_{1}-J_{4} model on the triangular lattice. Instead, a state with broken spatial symmetry is favored in the most probable parameter regime for the SFS state and has an energy much lower than that of the SFS state and other proposed variational states. More specifically, we find that for J40.09J1J_{4}\geq 0.09J_{1}, the model favors a valence bond solid state with a 4×64\times 6 period in its local spin correlation pattern and has a variational energy that is about 5%5\% lower than that of the SFS state. This state is separated from the π\pi-flux state for J40.045J1J_{4}\leq 0.045J_{1} by an intermediate symmetry breaking phase with a zigzag pattern in its local spin correlation. We find that the variational phase diagram we got is in qualitative agreement with that obtained from exact diagonalization on a 6×66\times 6 cluster.

I Introduction

The search of quantum spin liquid in strongly frustrated quantum magnets has lasted for more than three decadesPALee1 ; Balents . A quantum spin liquid is an exotic state of matter that can host excitations with fractionalized quantum number and novel exchange statistics. Such novel excitations may be responsible for some major puzzles in strongly correlated electron systems, for example, the anomalous dynamical behaviors of some highly frustrated quantum magnets and the non-Fermi liquid behavior of the cuprate superconductorsPiazza ; Saya ; Becca ; Li1 ; PALee2 . These novel excitations may also be used to realize topologically protected quantum computation. However, even after the extensive efforts of the community in the last three decades, we are still not sure if such an exotic state of matter is indeed realized in any real material.

The difficulty in truly identifing a quantum spin liquid in a real material has multiple origins. As a typical example of state of matter that is beyond the description of the Landau paradigm, a quantum spin liquid lacks the conventional local order parameter to be detected experimentally. At the same time, a quantum spin liquid usually occurs only in a very small parameter region of model Hamiltonian in which mutually frustrating exchange couplings are delicately balanced with each other. Furthermore, the inevitable existence of impurities in real materials may obscure the distinction between genuine spin liquid behavior and some glassy behavior. These difficulties are all related to the lack of physical intuition on the nature and the mechanism of emergence of the quantum spin liquid. In fact, we seldom has the physical intuition to judge if a particular kind of quantum spin liquid can be realized in a specific model.

The U(1)U(1) spin liquid with a large spinon Fermi surface(referred to as the SFS state in the following) is a clearly an exception in this regard. This state can be roughly thought of as the descendant of a metallic state after a Mott transition, in which the charge excitation has already developed a gap but the Fermi surface remains intact. Such a situation is very likely to occur if the system is in the close vicinity of the Mott transition so that the multiple spin exchange coupling is large. Indeed, in triangular lattice spin liquid materials such as κ\kappa-(BEDT-TTF)2Cu2(CN)3 and EtMe3Sb[Pd(dmit)2]2, people do find evidence for the existence of such a quantum spin liquidKanoda1 ; Kanoda2 ; Zhou . The hypothetical charge neutral spinon Fermi surface manifests itself in the metal-like behavior of the magnetic susceptibility and the specific heat at low temperature, although the system is already a charge insulator. Thermal conductivity measurements aiming to detect itinerant Fermionic spinon lead to controversial resultsMatsuda ; Taillefer ; Lisy . Similar claims of the SFS state have also been made for other triangular magnetic systems such as 1T-TaS2 and YbMgGaO4 PALee3 ; Kanigel ; Arcon ; Zhang .

Motivated by these physical expectations, a large number of theoretical efforts have been devoted to the study of the spin-12\frac{1}{2} Heisenberg-ring exchange model(the J1J4J_{1}-J_{4} model) on the triangular lattice, in which a large ring exchange coupling J4J_{4} is introduced to frustrate the conventional 120 degree order favored by the Heisenberg exchange coupling J1J_{1}. Variational studies find that when J4J_{4} is strong enough, the SFS state becomes the best variational ground state of the modelMotrunich ; SSLee1 ; Grover ; Xu ; Liu ; Lijx . This conclusion finds some support from a DMRG simulation of the modelPALee4 . However, in a more recent DMRG simulationMoore , it is found that in the most favorable parameter regime for the SFS state there is strong evidence of spatial symmetry breaking in the ground state of the model. A systematic variational investigation of the model with potential spatial symmetry breaking allowed is thus strongly called for.

In this work, we have performed a large scale variational optimization of the spin-12\frac{1}{2} J1J4J_{1}-J_{4} model on the triangular lattice without assuming any symmetry a prior. To tackle such a challenging numerical problem, we have proposed several improvements on the variational optimization algorithm. We find that while the SFS state is extremely stable locally within the subspace of the Fermionic resonating valence bond(RVB) states, it is never the true variational ground state of the spin-12\frac{1}{2} J1J4J_{1}-J_{4} model on the triangular lattice. Instead, we find that in the parameter regime which is thought to be the most favorable for the SFS state, a symmetry breaking state with a 4×64\times 6 periodicity in its local spin correlation pattern has an energy much lower than that of the SFS state. This state is separated from the π\pi-flux phase favored for small J4J_{4} by an intermediate phase with a zigzag pattern in its local spin correlation. We find that such a variational phase diagram has strong similarity with that obtained from exact diagonalization(ED) calculation on small cluster.

This paper is organized as follows. In the next section, we introduce the spin-12\frac{1}{2} J1J4J_{1}-J_{4} model studied in this work and summarize previous theoretical results about it. We then introduce the variational wave functions we adopted in our study in Sec.III. This is followed by an introduction of the optimization algorithms we used in this work in Sec.IV. In Section V, we present our numerical results from the variational optimization. Here we will present the full variational phase diagram of the J1J4J_{1}-J_{4} model and the symmetry breaking pattern of each phase in this phase diagram. A comparison with the ED phase diagram will also be presented in this section. In the sixth section, we draw conclusion from our results and discuss their implications.

II The J1J4J_{1}-J_{4} model on the triangular lattice

The model we study in this work is described by the following Hamiltonian

H=J1i,j𝐒i𝐒j+J4[i,j,k,l](Pi,j,k,l+Pi,j,k,l1)H=J_{1}\sum_{\langle i,j\rangle}\mathbf{S}_{i}\cdot\mathbf{S}_{j}+J_{4}\sum_{[i,j,k,l]}(P_{i,j,k,l}+P^{-1}_{i,j,k,l}) (1)

in which J1J_{1} denotes the Heisenberg exchange coupling between nearest-neighboring sites of the triangular lattice, J4J_{4} denotes the four-spin ring exchange coupling around every elementary rhombi of the triangular lattice. In the following, we will set J1=1J_{1}=1 as the unit of energy. This model and its various extensions have been studied by many researchersLiMing ; Motrunich ; Schmidt ; Grover ; Xu ; PALee4 ; Seki ; Moore . We note that the value of J1J_{1} in our notation differs by a factor of two from that adopted in Ref.[Xu, ; Grover, ]. Now we summarize previous results about this model.

When J4=0J_{4}=0, the model reduces to the Heisenberg model with antiferromagnetic exchange couplings between nearest-neighboring sites of the triangular lattice. It is well known that such a model possesses a 120 degree long range order in its ground state. Within the RVB framework, it was found that such a long range ordered state can be accurately described by a Bosonic RVB ansatzZhangQ . Using a general theorem proved by Seiji and Sorella relating Bosoinc and Fermionic RVB state on planar graphSeiji , this Bosonic RVB state can be connected continuously to the famous π\pi-flux phase on the triangular lattice, which is a Fermionic RVB state describing a Dirac quantum spin liquid. The short-ranged RVB state proposed by AndersonAnderson plays a key role in establishing such a marvelous connection. Although there is no true magnetic long range order in the π\pi-flux phase, the local spin correlation in this state is very close to that in the 120 degree ordered ground state(see Fig.5 below). In our study, we will concentrate on the subspace of spin rotational invariant Fermionic RVB state, within which the π\pi-flux state is the best representative of the 120 degree ordered phase. As we will see later, the π\pi-flux phase on the triangular lattice is actually the global minimum at J4=0J_{4}=0 within the subspace of Fermionic RVB state.

The ground state of the model with J40J_{4}\neq 0 is much more complicated and is still under strong debate. ED study on small clustersLiMing find that some kind of spin liquid of unknown character may be realized for large J4J_{4}. Driven by the experimental claim of possible spin liquid behavior in triangular lattice organic salt material κ\kappa-(BEDT-TTF)2Cu2(CN)3Kanoda1 , the model is revisited by Moturnich in 2005 with variational Monte Carlo methodMotrunich . It is found that a U(1)U(1) spin liquid state with a large spinion Fermi surface is favored for large ring exchange coupling. Similar conclusions are also reached from other variational studies in the large J4J_{4} regimeGrover ; Xu ; Liu ; Lijx . The intermediate phase between the 120 degree ordered phase and the SFS state is proposed to take the form of a Z2Z_{2} spin liquid state with either an extended ss-wave, a dx2y2d_{x^{2}-y^{2}}-wave or a dx2y2+idxyd_{x^{2}-y^{2}}+id_{xy}-wave spinon pairing.

The SFS state is believed to host Fermionic spinon excitation around the hypothetical spinon Fermi surface. This seems to be consistent with the experimental observation of a linear-in-T specific heat and a constant magnetic susceptibility on the triangular lattice organic salt material κ\kappa-(BEDT-TTF)2Cu2(CN)3, which is thought to be described by the J1J4J_{1}-J_{4} model. However, gauge fluctuation beyond the mean field description is argued to generate singular correction to the specific heatMotrunich ; SSLee1 of the form CvT2/3C_{v}\propto T^{2/3}, which is never observed. Driven by the tension between theories and experiments, several novel spin liquids other than the SFS phase have been proposed over the years. For example, the author of Ref.[Xu, ] proposed a Z2Z_{2} spin liquid with a fully gapped gauge fluctuation spectrum and a spinon dispersion with quadratic band touching(QBT) at the Γ\Gamma point. The spinon excitation above the QBT point enjoys a finite density of state but is free from singular gauge field fluctuation corrections. Such a novel state is found to have a slightly lower variational energy in the intermediate region of the ring exchange coupling than both the SFS state and the nematic spin liquid mentioned above. Another proposal is to assume spinon pairing at nonzero total momentum so that the spinon Fermi surface is only partially gappedYao . This proposal has no support from the calculation of the J1J4J_{1}-J_{4} model.

In Ref.[Li2, ], we show that the singular gauge fluctuation correction around the SFS state argued before is actually a theoretical artifact of the Gaussian approximation. When we go beyond the Gaussian level, the gauge fluctuation around the SFS state can only contribute a subleading correction to the specific heat of the form CvT2C_{v}\propto T^{2}, even if the SFS state is indeed the true ground state of the J1J4J_{1}-J_{4} model. The new theory also provides a unified mechanism for spin fractionalization in both 1D and 2D quantum magnets. Such a new mechanism is built on the nontrivial topological character of the Gutzwiller projected mean field state, rather than the deconfinement of slave particles.

The J1J4J_{1}-J_{4} model has also been studied by DMRG simulationsPALee4 ; Moore . In an attempt to account for the possible spin liquid behavior found in 1T-TaS2PALee3 ; Kanigel ; Arcon , a rather complicate triangular material argued to be described by an approximate J1J4J_{1}-J_{4} model at low energy, the authors of Ref.[PALee4, ] revisited the J1J4J_{1}-J_{4} model with DMRG. They found that a paramagnetic state without any detectable symmetry breaking pattern is realized at large value of J4J_{4}. This state possesses a spin structure factor with an approximate 2kFk_{F} peak expected for a spin liquid with a large spinon Fermi surface. However, such a claim is challenged by a more recent DMRG simulation on the same modelMoore , in which the authors report a zigzag type symmetry breaking phase in the parameter regime thought to be the most favorable for the SFS state. A more thorough investigation is thus clearly called for to determine if the SFS state can indeed be realized in this model.

III The variational wave functions

In this work, we describe the ground state of the J1J4J_{1}-J_{4} model with the Fermionic RVB state of the form

|fRVB=PG{ik,jk}k=1N/2a(ik,jk)Pik,jk|0|f-RVB\rangle=P_{G}\sum_{\{i_{k},j_{k}\}}\prod_{k=1}^{N/2}a(i_{k},j_{k})\ P_{i_{k},j_{k}}|0\rangle (2)

Here

Pik,jk=[fik,fjk,fik,fjk,]P_{i_{k},j_{k}}=[f^{\dagger}_{i_{k},\uparrow}f^{\dagger}_{j_{k},\downarrow}-f^{\dagger}_{i_{k},\downarrow}f^{\dagger}_{j_{k},\uparrow}] (3)

creates a Fermionic spin singlet pair(valence bond) between site iki_{k} and jkj_{k}. fi,σf^{\dagger}_{i,\sigma} is the Fermion creation operator on site ii and with spin σ\sigma. |0|0\rangle is the vacuum of the ff-Fermion. a(ik,jk)a(i_{k},j_{k}) is the RVB amplitude of the kk-th valence bond and satisfies a(ik,jk)=a(jk,ik)a(i_{k},j_{k})=a(j_{k},i_{k}). {ik,jk}\sum_{\{i_{k},j_{k}\}} denotes the sum over all valence bond configurations on the lattice. PGP_{G} denotes the Gutzwiller projection introduced to enforce the single occupancy constraint on the ff-Fermions.

III.1 General Fermionic RVB states

The Fermionic RVB state can be expanded in the Ising basis as

|fRVB={σ1,..σN}Ψ(σ1,.,σN)|σ1,.,σN|f-RVB\rangle=\sum_{\{\sigma_{1},.....\sigma_{N}\}}\Psi(\sigma_{1},....,\sigma_{N})|\sigma_{1},....,\sigma_{N}\rangle (4)

in which

|σ1,.,σN=k=1N/2fik,fjk,|0|\sigma_{1},....,\sigma_{N}\rangle=\prod_{k=1}^{N/2}f^{\dagger}_{i_{k},\uparrow}f^{\dagger}_{j_{k},\downarrow}|0\rangle (5)

denotes an Ising basis written in terms of the Fermion Fock state

Ψ(σ1,.,σN)=Det[𝐀]\Psi(\sigma_{1},....,\sigma_{N})=Det[\mathbf{A}] (6)

is the corresponding wave function amplitude. Here and in the following, we will use iki_{k} and jkj_{k} to denote the locations of the kk-th up and the kk-th down spin in the Ising basis |σ1,.,σN|\sigma_{1},....,\sigma_{N}\rangle, rather than the two end points of the kk-th valence bond. 𝐀\mathbf{A} is a N2×N2\frac{N}{2}\times\frac{N}{2} matrix with its matrix element given by a(ik,jk)a(i_{k},j_{k^{\prime}}) at its kk-th row and kk^{\prime}-th column.

In practice, we can treat the RVB amplitude a(ik,jk)a(i_{k},j_{k}) as variational parameter directly. The variational state constructed in this way will be referred to as the general RVB state(gRVB) in the following discussion. The number of variational parameters in the gRVB state increases very rapidly with the system size. As we will see in the next section, such an unfavorable feature of the gRVB state is compensated partly by the fact that the calculation of the energy gradient in the gRVB state is rather cheap.

III.2 Fermionic RVB states generated from mean field ansatzs

The Fermionic RVB state can also be generated by Gutzwiller projection of mean field ground state of the following Bogoliubov-de Gennes Hamiltonian

HMF={i,j},σ(χi,jfi,σfj,σ+h.c.)+{i,j}(Δi,jfi,fj,+h.c.)H_{MF}=-\sum_{\{i,j\},\sigma}(\chi_{i,j}f^{\dagger}_{i,\sigma}f_{j,\sigma}+h.c.)+\sum_{\{i,j\}}(\Delta_{i,j}f^{\dagger}_{i,\uparrow}f^{\dagger}_{j,\downarrow}+h.c.) (7)

Here the condition Δi,j=Δj,i\Delta_{i,j}=\Delta_{j,i} is imposed to enforce spin rotational symmetry of the variational ground state. In general, both χi,j\chi_{i,j} and Δi,j\Delta_{i,j} can be chosen complex. In our work, χi,j\chi_{i,j} and Δi,j\Delta_{i,j} will be chosen real and be restricted to the nearest neighboring bonds.They are otherwise free of any assumption.

HMFH_{MF} is usually referred to as a mean field ansatz or a variational ansatz of the Fermionic RVB state. To generate |fRVB|f-RVB\rangle, we rewrite HMFH_{MF} in the following form

HMF=ψ(𝝌𝚫𝚫𝝌)ψ=ψ𝐌ψH_{MF}=\psi^{\dagger}\left(\begin{array}[]{cc}-\bm{\chi}&\bm{\Delta}\\ \bm{\Delta}^{\dagger}&\bm{\chi}^{*}\end{array}\right)\psi=\psi^{\dagger}\mathbf{M}\psi (8)

in which

ψ=(f1,,,fN,,f1,,.,fN,)\psi^{\dagger}=(f^{\dagger}_{1,\uparrow},...,f^{\dagger}_{N,\uparrow},f_{1,\downarrow},....,f_{N,\downarrow}) (9)

Here 𝝌\bm{\chi} and 𝚫\bm{\Delta} are N×NN\times N matrices with χi,j\chi_{i,j} and Δi,j\Delta_{i,j} as their matrix elements. HMFH_{MF} can be diagonalized by the following unitary transformation

ψ=(𝒖𝒗𝒗𝒖)γ\psi=\left(\begin{array}[]{cc}\bm{u}&\bm{v}\\ -\bm{v}&\bm{u}\end{array}\right)\gamma (10)

in which 𝒖\bm{u} and 𝒗\bm{v} are N×NN\times N matrices. The diagonalized Hamiltonian takes the form of

HMF=γ(𝑬00𝑬)γH_{MF}=\gamma^{\dagger}\left(\begin{array}[]{cc}\bm{E}&0\\ 0&-\bm{E}\end{array}\right)\gamma (11)

in which 𝑬\bm{E} is a N×NN\times N diagonal matrix with positive definite diagonal matrix elements. When 𝚫0\bm{\Delta}\neq 0, the RVB amplitude a(ik,jk)a(i_{k},j_{k}) of the corresponding Fermionic RVB state is given by

a(ik,jk)=(𝒖1𝒗)ik,jk.a(i_{k},j_{k})=(\bm{u}^{-1}\bm{v})_{i_{k},j_{k}}. (12)

When 𝚫=0\bm{\Delta}=0, the RVB amplitude is given by

a(ik,jk)=Em<μφm(ik)φm(jk)a(i_{k},j_{k})=\sum_{E_{m}<\mu}\varphi^{*}_{m}(i_{k})\varphi_{m}(j_{k}) (13)

in which φm(jk)\varphi_{m}(j_{k}) is the eigenvector of the matrix 𝝌\bm{\chi} with eigenvalue EmE_{m}. Here μ\mu denotes the chemical potential of the Fermionic spinon at half filling.

III.3 Particle-hole transformation

In practical variational calculation, it is often convenient to adopt a particle-hole transformation on the down-spin Fermions, which is given by

fi,f~i,f^{\dagger}_{i,\downarrow}\rightarrow\tilde{f}_{i,\downarrow} (14)

The mean field ground state of HMFH_{MF} is then constructed by filling up the lowest NN eigenvectors of 𝐌\mathbf{M}. More specifically, we can expand the RVB state as

|fRVB={i1,iN/2}Ψ~(i1,,iN/2)k=1N/2fik,f~ik,|0~|f-RVB\rangle=\sum_{\{i_{1},...i_{N/2}\}}\tilde{\Psi}(i_{1},...,i_{N/2})\prod_{k=1}^{N/2}f^{\dagger}_{i_{k},\uparrow}\tilde{f}^{\dagger}_{i_{k},\downarrow}|\tilde{0}\rangle (15)

in which

|0~=i=1Nfik,|0|\tilde{0}\rangle=\prod_{i=1}^{N}f^{\dagger}_{i_{k},\downarrow}|0\rangle (16)

is the reference state with all sites occupied by down spin Fermions(or the vacuum of f~\tilde{f} operators). Here i1,,iN/2i_{1},...,i_{N/2} denote the locations of the N/2N/2 up spin Fermions. Note that they are also the locations of the N/2N/2 holes of the down spin Fermion. The wave function in the particle-hole transformed picture then takes the form of

Ψ~(i1,,iN/2)=Det[𝚽]\tilde{\Psi}(i_{1},...,i_{N/2})=Det[\bm{\Phi}] (17)

in which 𝚽\bm{\Phi} is a N×NN\times N matrix of the form

𝚽=(ϕ1(i1)..ϕN(i1)....ϕ1(iN/2)..ϕN(iN/2)ϕ1(i1+N)..ϕN(i1+N)....ϕ1(iN/2+N)..ϕN(iN/2+N))\displaystyle\bm{\Phi}=\left(\begin{array}[]{cccc}\phi_{1}(i_{1})&.&.&\phi_{N}(i_{1})\\ .&.&.&.\\ \phi_{1}(i_{N/2})&.&.&\phi_{N}(i_{N/2})\\ \phi_{1}(i_{1}+N)&.&.&\phi_{N}(i_{1}+N)\\ .&.&.&.\\ \phi_{1}(i_{N/2}+N)&.&.&\phi_{N}(i_{N/2}+N)\end{array}\right)

Here ϕn(i)\phi_{n}(i) denotes the ii-th component of the nn-th eigenvector of the matrix 𝐌\mathbf{M} with negative eigenvalue.

III.4 Number of variational parameters

In this study, we use either the general Fermionic RVB state or RVB state generated from mean field ansatz to describe the ground state of the J1J4J_{1}-J_{4} model. For the general RVB state, there will be N(N1)2\frac{N(N-1)}{2} variational parameters to be optimized on a finite cluster with NN sites. For RVB state generated from mean field ansatz, we can choose either a U(1)U(1) ansatz, in which 𝚫=0\bm{\Delta}=0, or a Z2Z_{2} ansatz, in which both 𝝌\bm{\chi} and 𝚫\bm{\Delta} are nonzero. In the U(1)U(1) ansatz, there are 3N3N variational parameters to be optimized, which are the hopping amplitude χi,j\chi_{i,j} on the 3N3N nearest neighboring bonds. In the Z2Z_{2} ansatz, there are 6N+16N+1 variational parameters to be optimized, which are the hopping amplitude χi,j\chi_{i,j} and the pairing amplitude Δi,j\Delta_{i,j} on the 3N3N nearest neighboring bonds and the chemical potential setting the Fermi level of the spinon. The gRVB state represents the most general form of a Fermionic RVB state and it may not be generated by any short-ranged mean field ansatz. Correspondingly, it contains a much larger number of variational parameters. The optimization of large number of variational parameters calls for very efficient optimization algorithm, which we will now introduce.

IV Some new developments of variational optimization algorithm

The key step in the variational Monte Carlo optimization is the computation of the variational energy and its gradient. Suppose that the wave function of the variational state |Ψ|\Psi\rangle in a orthonormal basis |R|R\rangle is given by Ψ(R)\Psi(R), the variational energy is then given by

E=HΨ=Ψ|H|ΨΨ|Ψ=R|Ψ(R)|2Eloc(R)R|Ψ(R)|2E=\langle H\rangle_{\Psi}=\frac{\langle\Psi|H|\Psi\rangle}{\langle\Psi|\Psi\rangle}=\frac{\sum_{R}|\Psi(R)|^{2}E_{loc}(R)}{\sum_{R}|\Psi(R)|^{2}} (24)

in which the local energy Eloc(R)E_{loc}(R) is defined by

Eloc(R)=RR|H|RΨ(R)Ψ(R)E_{loc}(R)=\sum_{R^{\prime}}\langle R|H|R^{\prime}\rangle\frac{\Psi(R^{\prime})}{\Psi(R)} (25)

The gradient of the variational energy with respect to the variational parameters is given by

E=lnΨ(R)Eloc(R)ΨElnΨ(R)Ψ\nabla E=\langle\nabla\ln\Psi(R)E_{loc}(R)\rangle_{\Psi}-E\langle\nabla\ln\Psi(R)\rangle_{\Psi} (26)

Here we denotes the variational parameters as 𝜶\bm{\alpha} and abbreviate 𝜶\nabla_{\bm{\alpha}} as \nabla. These two quantities can be computed by standard Monte Carlo sampling on the distribution generated by |Ψ(R)|2|\Psi(R)|^{2}. In the calculation of the gradient, the key quantity is lnΨ(R)\nabla\ln\Psi(R). For the general RVB state, it is given by

lnΨ(R)=𝐓𝐫[𝐀𝐀1]\nabla\ln\Psi(R)=\mathbf{Tr}[\nabla\mathbf{A}\mathbf{A}^{-1}] (27)

Since we take the RVB amplitude a(ik,jk)a(i_{k},j_{k}) directly as the variational parameters, the matrix elements of 𝐀\nabla\mathbf{A} is either 1 or 0, depending on whether the gradient is taken on a given RVB amplitude. For RVB state generated from Gutzwiller projection of the ground state of a mean field ansatz, we have

lnΨ(R)=𝐓𝐫[𝚽𝚽1]\nabla\ln\Psi(R)=\mathbf{Tr}[\nabla\bm{\Phi}\bm{\Phi}^{-1}] (28)

The matrix elements of 𝚽\nabla\bm{\Phi} can be calculated from the first order perturbation theory as follows

ϕn=Em>0ϕm|HMF|ϕnEnEmϕm\nabla\phi_{n}=\sum_{E_{m}>0}\frac{\langle\phi_{m}|\nabla H_{MF}|\phi_{n}\rangle}{E_{n}-E_{m}}\phi_{m} (29)

Here |ϕn|\phi_{n}\rangle and EnE_{n} denote the nn-th eigenvector and eigenvalue of the mean field Hamiltonian HMFH_{MF}.

To proceed the optimization procedure, one need also the Hessian matrix of EE with respect to the variational parameters. However, the Hessian matrix is usually too expensive to be calculated numerically. Different variational optimization algorithms differ in their way to approximate the Hessian matrix, which we will now review briefly.

IV.1 The steepest descent

The steepest descent(SD) algorithm is the simplest optimization algorithm. It corresponds to setting the Hessian matrix proportional to the identity matrix. In the SD algorithm, the variational parameters are updated as follows

𝜶𝜶δE\bm{\alpha}\rightarrow\bm{\alpha}-\delta\nabla E (30)

in which δ\delta is the step length. The step length is usually adjusted by trial and error. A more intelligent way to adjust the step length is the following self-learning trick, in which the step length is updated according to the change in the direction of the gradient as follows

δδ×(1+ηEE|E||E|)\delta\rightarrow\delta\times(1+\eta\frac{\nabla E\cdot\nabla E^{\prime}}{|\nabla E||\nabla E^{\prime}|}) (31)

Here η(0,1)\eta\in(0,1) is an acceleration factor, E\nabla E and E\nabla E^{\prime} are the gradients of the energy in the current and the previous step. Suitable choice of η\eta can accelerate significantly the optimization procedure at the initial stage, when the energy gradient is large.

The SD algorithm usually works very well at the initial stage of the optimization procedure. However, it loses efficiency when the optimization procedure encounters long and narrow valley with flat bottom in the energy landscape. In such a situation, there will be large fluctuation in the eigenvalues of the Hessian matrix. Approximating the Hessian matrix with an identity matrix is then clearly not a wise choice.

IV.2 Stochastic reconfiguration

The stochastic reconfiguration algorithm is a widely used variational optimization algorithm. It mimics the effect of the Hessian matrix with a positive-definite Hermitian matrix 𝐒\mathbf{S} generated from the metric of the variational state in the variational spaceSeiji . More specifically, 𝐒\mathbf{S} is given by

𝐒=lnΨ(R)lnΨ(R)ΨlnΨ(R)ΨlnΨ(R)Ψ\mathbf{S}=\langle\nabla ln\Psi(R)\nabla ln\Psi(R)\rangle_{\Psi}-\langle\nabla ln\Psi(R)\rangle_{\Psi}\langle\nabla ln\Psi(R)\rangle_{\Psi} (32)

It is easy to show that 𝐒\mathbf{S} is just the Hessian matrix of the following quantity with respect to the change of variational parameters

ΔSR=22Ψ|ΨΨ|ΨΨ|Ψ\Delta^{\mathrm{SR}}=2-2\frac{\langle\Psi|\Psi^{\prime}\rangle}{\sqrt{\langle\Psi|\Psi\rangle\langle\Psi^{\prime}|\Psi^{\prime}\rangle}} (33)

Here we assume that |Ψ|\Psi\rangle is fixed and |Ψ|\Psi^{\prime}\rangle is varying. ΔSR\Delta^{\mathrm{SR}} defined in this way can be interpreted as the distance between |Ψ|\Psi\rangle and |Ψ|\Psi^{\prime}\rangle in the Hilbert space.

In the SR algorithm, the variational parameters are updated as follows

𝜶𝜶δ𝐒1E\bm{\alpha}\rightarrow\bm{\alpha}-\delta\ \mathbf{S}^{-1}\nabla E (34)

in which δ\delta is the step length. The self-learning acceleration trick is also applicable in the SR method.

The introduction of the 𝐒\mathbf{S} matrix in the SR method amounts to replace the naive distance in the Euclidean space of variational parameters with the distance in the Hilbert space. Such a regulation procedure can be very helpful when some variational parameters are nearly redundant. However, since 𝐒\mathbf{S} only depends on the variational state but not on the Hamiltonian, it can not approximate the effect of the true Hessian matrix correctly in certain situations. In practice, the SR method may still suffer slow convergence or even run away from true minimum. A better approximation of the effect of the Hessian matrix is needed.

IV.3 BFGS

The Broyden-Fletcher-Goldfarb-Shanno (BFGS) method is a quasi-Newton method. It generates an iterative approximation for the (inverse) of the Hessian matrixNocedal from the gradient. The approximate (inverse)Hessian matrix is updated as follows

𝐁k+1=(𝐈𝐬k𝐲kT𝐲kT𝐬k)𝐁k(𝐈𝐲k𝐬kT𝐲kT𝐬k)+𝐬k𝐬kT𝐲kT𝐬k\mathbf{B}_{k+1}=\left(\mathbf{I}-\frac{\mathbf{s}_{k}\mathbf{y}^{T}_{k}}{\mathbf{y}^{T}_{k}\mathbf{s}_{k}}\right)\mathbf{B}_{k}\left(\mathbf{I}-\frac{\mathbf{y}_{k}\mathbf{s}^{T}_{k}}{\mathbf{y}^{T}_{k}\mathbf{s}_{k}}\right)+\frac{\mathbf{s}_{k}\mathbf{s}^{T}_{k}}{\mathbf{y}^{T}_{k}\mathbf{s}_{k}} (35)

in which k=0,1,2..k=0,1,2.....

𝐬k\displaystyle\mathbf{s}_{k} =\displaystyle= 𝜶k+1𝜶k\displaystyle\bm{\alpha}_{k+1}-\bm{\alpha}_{k}
𝐲k\displaystyle\mathbf{y}_{k} =\displaystyle= Ek+1Ek\displaystyle\nabla E_{k+1}-\nabla E_{k} (36)

are the difference between successive variational parameters and energy gradient. 𝜶k=0\bm{\alpha}_{k=0} is the initial guess of the variational parameters and Ek=0\nabla E_{k=0} is the energy gradient at the starting point. The Hessian matrix is initially set to be the identity matrix, namely 𝐁k=0=𝐈\mathbf{B}_{k=0}=\mathbf{I}.

Using such an iterative approximation on the Hessian matrix, the variational parameters are updated as follows

𝜶k+1=𝜶k+δ𝐁kEk\bm{\alpha}_{k+1}=\bm{\alpha}_{k}+\delta\ \mathbf{B}_{k}\nabla E_{k} (37)

Here δ\delta is the step length. In principle the step length should be determined by a linear search in the direction of 𝐁kEk\mathbf{B}_{k}\nabla E_{k}. Such a linear search can be accomplished in principle in the variational Monte Carlo simulation by reweighting in the searching direction. However, to reduce computational expense, we choose a fixed step length by trial and error.

IV.4 Conjugate gradient

Another simpler method to go beyond the steepest descent method is the conjugate gradient(CG) methodNocedal . It corrects the searching direction as follows

𝐝k+1=Ek+1+βk+1𝐝k\mathbf{d}_{k+1}=-\nabla E_{k+1}+\beta_{k+1}\mathbf{d}_{k} (38)

in which k=0,1,2,.k=0,1,2,....

βk+1=Ek+1Ek+1EkEk\beta_{k+1}=\frac{\nabla E_{k+1}\cdot\nabla E_{k+1}}{\nabla E_{k}\cdot\nabla E_{k}} (39)

or

βk+1=max(Ek+1(Ek+1Ek)EkEk,0)\beta_{k+1}=max(\frac{\nabla E_{k+1}\cdot(\nabla E_{k+1}-\nabla E_{k})}{\nabla E_{k}\cdot\nabla E_{k}},0) (40)

Initially we set β0=0\beta_{0}=0.

The variational parameters are updated in the conjugate gradient method as follows

𝜶k+1=𝜶k+δ𝐝k\bm{\alpha}_{k+1}=\bm{\alpha}_{k}+\delta\ \mathbf{d}_{k} (41)

in which the step length δ\delta should be determined by linear search in the direction of 𝐝k\mathbf{d}_{k}. In practice, to reduce computational expense we choose a fixed step length by trial and error.

IV.5 Comparison of the performance of different optimization algorithms

We note that for both the BFGS and the CG method, the initial step of the optimization is just the SD update. As a result of the finite accuracy in the computation of the energy and the gradient in variational Monte Carlo simulation, we cut off the BFGS and the CG iteration at a finite depth. We find empirically that 10 steps of BFGS or CG iteration has the best balance between numerical efficiency and numerical stability, after which we restart the iteration by setting k=0k=0.

Fig.1 compares the performance of the four algorithms in a typical situation. Here we optimize the variational energy of the J1J4J_{1}-J_{4} model at J4=0.065J_{4}=0.065 with the U(1)U(1) mean field ansatz. We start from the same initial guess and use the same normalized step length. We find that the BFGS method has the best numerical efficiency and stability among the four algorithms. The CG algorithm exhibits a similar numerical efficiency as the BFGS algorithm but with a significantly larger fluctuation. Both the steepest descent and the stochastic reconfiguration method fail to escape from the the local minimum around E0.4976E\approx-0.4976 within 2000 optimization steps. In the following, we will mainly use the BFGS method. However, we note that the conjugate gradient method has the advantage that it does not need to store the approximate Hessian matrix, which is huge when the number of variational parameters is large.

Refer to caption
Figure 1: The comparison of the performance of the four optimization algorithms in a typical situation. Shown here is the variational energy of the J1J4J_{1}-J_{4} model at J4=0.065J_{4}=0.065 with a U(1)U(1) mean field ansatz, starting from the same initial guess and using the same normalized step length.

V The variational phase diagram of the J1J4J_{1}-J_{4} model

We have performed variational optimization of the J1J4J_{1}-J_{4} model on a L×LL\times L cluster with periodic boundary condition in both the 𝐚1\mathbf{a}_{1} and the 𝐚2\mathbf{a}_{2} direction(see Fig.2). We have adopted both the general RVB state and the RVB state generated from a mean field ansatz of either the U(1)U(1) or the Z2Z_{2} form. No further assumption is made on the form of the RVB state. Fig.3 shows the variational phase diagram we get from the optimization. For clearness we only present the result for the Z2Z_{2} RVB state. The results for the gRVB state and the U(1)U(1) RVB state are qualitatively similar, with the variational energies satisfying EgRVBEZ2EU(1)E_{gRVB}\leq E_{Z_{2}}\leq E_{U(1)}.

We find that there are three phases in the phase diagram of the J1J4J_{1}-J_{4} model. For 0J40.0450\leq J_{4}\leq 0.045, the optimized RVB state is the well known π\pi-flux phase on the triangular latticeSeiji . The optimized variational energy exhibits only a tiny curvature in this regime, indicating that the π\pi-flux phase, the best representative of the 120 degree ordered phase in the space of Fermionic RVB state, enjoys a finite range of stability when we increase the ring exchange coupling. For J40.09J_{4}\geq 0.09, the variational energy also has a very small curvature. We find that the optimized RVB state in this regime breaks the translational symmetry and exhibits a 4×64\times 6 periodicity in its local spin correlation pattern. This phase will thus be called the 4×64\times 6 phase in the following. For 0.045J40.090.045\leq J_{4}\leq 0.09, we find another symmetry breaking phase with zigzag spin correlation pattern. This intermediate phase will be called the zigzag phase in the following. Notably, we find that the optimized variational energy is significantly lower than that of the SFS state. In fact, the SFS state is found to be never a good approximation of the ground state of the J1J4J_{1}-J_{4} model. This is the most important finding of this work.

Refer to caption
Figure 2: The triangular lattice and its reciprocal vectors. Our calculation is done on a L×LL\times L cluster with periodic boundary condition in both the 𝐚1\mathbf{a}_{1} and the 𝐚2\mathbf{a}_{2} direction.
Refer to caption
Figure 3: The variational phase diagram of the J1J4J_{1}-J_{4} model on the triangular lattice. Shown here is the result obtained from the Z2Z_{2} RVB state. The computation is done on a 12×1212\times 12 cluster with periodic boundary condition. The results for the gRVB state and the U(1)U(1) RVB state are qualitatively similar. Here the thick blue line and thick red line represent the variational energies of the SFS state and the π\pi-flux phase. The 4×64\times 6 phase in the large J4J_{4} region exhibits 4×64\times 6 periodicity in its local spin correlation pattern. The dashed line within the zigzag phase separates two phases with different translational symmetries.

Below we will present the results for each of the three phases in more detail.

Refer to caption
Figure 4: The variational energies of the three kinds of RVB state as functions of J4J_{4} in the range 0J40.0450\leq J_{4}\leq 0.045, computed on a 12×1212\times 12 cluster with periodic boundary condition. The variational state is essentially independent of J4J_{4} in this regime and corresponds to the π\pi-flux phase on the triangular lattice.

V.1 The π\pi-flux phase

Fig.4 plots the variational energies computed from the three kinds of RVB state as a function of J4J_{4} for 0J40.0450\leq J_{4}\leq 0.045. The variational energies of all these three kinds of RVB state are very close to each other and are within statistical error perfect linear functions of J4J_{4}. Such a nearly zero curvature behavior in the variational energy indicates that the variational ground state is almost independent of J4J_{4}. In the case of the U(1)U(1) RVB state, the optimized variational state reduces to the well known π\pi-flux phase on the triangular lattice. This can be explicitly seen from the optimized variational parameters of the U(1)U(1) ansatz, which indeed encloses a gauge flux of π\pi around every elementary rhombi of the triangular lattice. We find that the small difference in the optimized variational energies between the Z2Z_{2} and the U(1)U(1) RVB state can be attributed to finite size effect. On a 24×2424\times 24 cluster, the optimized variational energy of the two states become indistinguishable within the statistical error of the variational Monte Carlo simulation. The equivalence between the optimized U(1)U(1) and Z2Z_{2} RVB state in this low J4J_{4} regime can also be seen from the static spin structure factor S(𝐪)S(\mathbf{q}), which is defined as

S(𝐪)=1Ni,j𝐒i𝐒jei𝐪(𝐑i𝐑j)S(\mathbf{q})=\frac{1}{N}\sum_{i,j}\mathbf{S}_{i}\cdot\mathbf{S}_{j}\ e^{i\mathbf{q}\cdot(\mathbf{R}_{i}-\mathbf{R}_{j})} (42)

Here we define the components of the wave vector as follows

𝐪=q1𝐛1+q2𝐪2\mathbf{q}=q_{1}\mathbf{b}_{1}+q_{2}\mathbf{q}_{2} (43)

in which 𝐛1,2\mathbf{b}_{1,2} are the two reciprocal vectors of the triangular lattice. In Fig.5, we present the static spin structure factor of the optimized U(1)U(1) and Z2Z_{2} RVB state at J4=0J_{4}=0. We find that S(𝐪)S(\mathbf{q}) is within statistical error independent of J4J_{4} in the whole 0J40.0450\leq J_{4}\leq 0.045 regime and features pronounced peaks at the wave vector corresponding to the 120 degree ordered phase, namely 𝐪=(2π3,4π3)\mathbf{q}=(\frac{2\pi}{3},\frac{4\pi}{3}) and 𝐪=(4π3,2π3)\mathbf{q}=(\frac{4\pi}{3},\frac{2\pi}{3}).

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 5: The static spin structure factor S(𝐪)S(\mathbf{q}) of the optimized (a)U(1)U(1), (b)Z2Z_{2} and (c)gRVB states for J4=0J_{4}=0. S(𝐪)S(\mathbf{q}) is essentially independent of J4J_{4} for 0J40.0450\leq J_{4}\leq 0.045. (d)Comparison of the static spin structure factor of the three kind of RVB states along the path (0,2π)(2π,0)(0,2\pi)\rightarrow(2\pi,0).

The energy difference between the general RVB state and the U(1)U(1) or the Z2Z_{2} RVB state is more tricky. We find that the optimized gRVB state can not be generated by any short-ranged mean field ansatz. However, the static spin structure factor of the optimized gRVB state is essentially indistinguishable from that of the optimized U(1)U(1) or Z2Z_{2} RVB state(see Fig5c and 5d). It is currently impossible to conduct the optimization of the gRVB state on a cluster significantly larger than the 12×1212\times 12 cluster. For example, on a 24×2424\times 24 cluster, the number of variational parameters in the gRVB state becomes N(N1)2=165600\frac{N(N-1)}{2}=165600, which is too large to be reliably optimized.

In all, we find that the optimized RVB state in the whole 0J40.0450\leq J_{4}\leq 0.045 regime represents a state essentially equivalent to the π\pi-flux phase on the triangular lattice, which is the closest counterpart of the 120 degree ordered phase in the space of Fermionic RVB state.

Refer to caption
Refer to caption
Figure 6: The convergence of the variational energy(a) and the variational parameters(b) towards those of the SFS state. Here we set the distance of the initial guess from the SFS ansatz to be η=0.75\eta=0.75.

V.2 The large J4J_{4} phase

It is generally believed that in the large J4J_{4} regime the SFS state is the most favorable variational ground state. Indeed, we find that the SFS state is locally extremely stable in this regime. The SFS state is generated by the following mean field ansatz

χi,j\displaystyle\chi_{i,j} =\displaystyle= {1,nearestneighbor0,otherwise\displaystyle\Big{\{}\begin{array}[]{c}1,\ nearest\ neighbor\\ 0,\ \ otherwise\end{array} (46)
Δi,j\displaystyle\Delta_{i,j} =\displaystyle= 0\displaystyle 0 (47)

As a result of the translational and rotational symmetry of the mean field ansatz, the energy gradient must be also translational and rotational symmetric. The energy gradient in the SFS phase thus must vanish since the Gutzwiller projected state is invariant under a uniform rescaling of all variational parameters. In other words, the SFS phase is an exact saddle point in the space of Fermionic RVB state. To illustrate the local stability of the SFS state, we have performed variational optimization at J4=0.15J_{4}=0.15 starting from the following initial guess of the mean field ansatz

χi,j\displaystyle\chi_{i,j} =\displaystyle= {1±ηr,nearestneighbor0,otherwise\displaystyle\Big{\{}\begin{array}[]{c}1\pm\eta r,\ nearest\ neighbor\\ 0,\ \ otherwise\end{array} (50)
Δi,j\displaystyle\Delta_{i,j} =\displaystyle= 0\displaystyle 0 (51)

Here rr is a random number distributed uniformly in the range of r(0,1)r\in(0,1), η\eta is a constant measuring the distance of the initial guess from the SFS ansatz. As is shown in Fig.6, even if we choose η\eta as large as η=0.75\eta=0.75, the variational energy still converges to a value very close to that of the SFS state, which is E0.576E\approx-0.576 at J4=0.15J_{4}=0.15. The corresponding variational parameters also converge to that of the SFS ansatz.

However, the SFS state is only locally stable. To illustrate this point, we have performed variational optimization starting from fully random initial guess of the variational parameters. The optimization is done for J4=0.15J_{4}=0.15 on a 12×1212\times 12 cluster with periodic boundary condition. A preliminary trial shows that the optimized variational ground state exhibits an approximate 4×64\times 6 modulation in its local spin correlation pattern. This is illustrated in Fig.7. We note that the modulation in the local spin correlation is very strong, ranging from almost pure spin singlet correlation to pure spin triplet correlation between nearest neighboring spins. The translational symmetry is seriously broken.

We then refine the optimization by assuming a 4×64\times 6 periodicity in the variational parameters. We find that for both the U(1)U(1) and the Z2Z_{2} mean field ansatz, the optimized variational energy converge to values much lower than that of the SFS state. Fig.8 and Fig.9 illustrate the convergence of the variational energies for randomly chosen initial guess of the U(1)U(1) and the Z2Z_{2} mean field ansatz. For U(1)U(1) mean field ansatz, we find that about 1/31/3 of the variational optimization trials converge to the lowest variational energy of E0.603E\approx-0.603. For Z2Z_{2} mean field ansatz, about 1/21/2 of the variational optimization trials converge to the lowest variational energy of E0.606E\approx-0.606. The optimized variational energy for both types of RVB states are about 5%5\% lower than the energy of the SFS state, which is E0.576E\approx-0.576 at J4=0.15J_{4}=0.15.

Refer to caption
Figure 7: The optimized variational ground state at J4=0.15J_{4}=0.15 exhibits 4×64\times 6 modulation in its local spin correlation pattern. Shown here is the spin correlation between nearest neighboring sites. Note that the magnitude of the modulation in the local spin correlation is very large, ranging from nearly pure spin singlet correlation to almost pure spin triplet correlation between nearest neighboring spins.
Refer to caption
Refer to caption
Figure 8: (a)The convergence of the variational energy at J4=0.15J_{4}=0.15 for a U(1)U(1) mean field ansatz. Shown here is the results from 82 optimization trials starting from randomly chosen initial guess of the U(1)U(1) mean field ansatz with 4×64\times 6 periodicity. The calculation is done on a 12×1212\times 12 cluster with periodic boundary condition. About 1/31/3 of the optimization trials converge to the lowest variational energy of E0.603E\approx-0.603. This is much lower than the energy of the SFS state, which is E0.576E\approx-0.576. (b)The histogram of the optimized variational energies from the 82 optimization trials.
Refer to caption
Refer to caption
Figure 9: (a)The convergence of the variational energy at J4=0.15J_{4}=0.15 for a Z2Z_{2} mean field ansatz. Shown here is the results from 57 optimization trials starting from randomly chosen initial guess of the Z2Z_{2} mean field ansatz with 4×64\times 6 periodicity. The calculation is done on a 12×1212\times 12 cluster with periodic boundary condition. About 1/21/2 of the optimization trials converge to the lowest variational energy of E0.606E\approx-0.606. This is also much lower than the energy of the SFS state, which is E0.576E\approx-0.576. (b)The histogram of the optimized variational energies from the 57 optimization trials.

On the 12×1212\times 12 cluster studied here, the most general periodicity of the mean field ansatz that is compatible with the periodic boundary condition of the cluster is C1×C2C_{1}\times C_{2}, in which C1,2=1,2,3,4,6,12C_{1,2}=1,2,3,4,6,12 denotes the period in the 𝐚1\mathbf{a}_{1} and 𝐚2\mathbf{a}_{2} direction. We have performed refined variational optimization assuming each of such periodicity. We find that the lowest variational energy is always obtained with the 4×64\times 6 periodicity at J4=0.15J_{4}=0.15.

We find that the optimized variational energy as a function of J4J_{4} exhibits a very small curvature in the large J4J_{4} regime(see Fig.10). This implies that the variational ground state is almost independent of J4J_{4} in this part of the phase diagram also. A naive linear extrapolation of the variational energy implies that the SFS state can not be the true ground state of the J1J4J_{1}-J_{4} model for J40.5J_{4}\leq 0.5. To check if the SFS state can be stabilized at still larger J4J_{4}, we have performed variational optimization for J4=10J_{4}=10, a value that is too large to be realistic for real materials. We find that a symmetry breaking phase with the same 4×64\times 6 periodicity in its spin correlation pattern is still significantly lower in energy than the SFS state. More specifically, we find that the optimized variational energy is E16.09E\approx-16.09 at J4=10J_{4}=10. This is again about 5%5\% lower than the energy of the SFS state, which is E15.33E\approx-15.33 at J4=10J_{4}=10. Fig.11 illustrates the local spin correlation pattern in the optimized state, which is similar to that at J4=0.15J_{4}=0.15. We thus conclude that the SFS state is never the best RVB state for the J1J4J_{1}-J_{4} model and that the 4×64\times 6 state is the best variational ground state in the whole range of J40.09J_{4}\geq 0.09. Such a state breaks both the translational and the point group symmetry of the model and exhibits 4×64\times 6 modulation in its local spin correlation pattern.

Refer to caption
Figure 10: The variational energies of the three kinds of RVB state in the range of 0.09J40.150.09\leq J_{4}\leq 0.15. The computation is done on a 12×1212\times 12 cluster with periodic boundary condition. The optimized variational ground state is found to be almost independent of J4J_{4} in this regime, as is evident from the smallness of the curvature in the optimized variational energy. Such a state is found to exhibit a 4×64\times 6 modulation in its local spin correlation pattern.
Refer to caption
Figure 11: The optimized variational ground state at J4=10J_{4}=10 exhibits the same 4×64\times 6 modulation in its local spin correlation pattern as that for J4=0.15J_{4}=0.15. Shown here is the spin correlation between nearest neighboring sites.
Refer to caption
Refer to caption
Refer to caption
Figure 12: The local spin correlation pattern of the optimized variational ground state for (a)0.045J40.0550.045\leq J_{4}\leq 0.055 and (b)0.055J40.090.055\leq J_{4}\leq 0.09. Both states exhibit zigzag spin correlation pattern, with the antiferromagnetic correlated backbones(plotted here in blue) running through (a)the 2𝐚1𝐚22\mathbf{a}_{1}-\mathbf{a}_{2} direction and (b)the 2𝐚2𝐚12\mathbf{a}_{2}-\mathbf{a}_{1} direction. (c)The local spin correlation pattern in a classical zigzag state, in which the neighboring spin correlation takes the value of 1-1, +14+\frac{1}{4} and 14-\frac{1}{4} on the blue, red and green bonds. The zigzag pattern shown in (a) differs from that shown in (b) by an additional two-fold translational symmetry breaking.

V.3 The intermediate phase(s)

The optimized variational ground state in the intermediate regime of 0.045J40.090.045\leq J_{4}\leq 0.09 is characterized by a zigzag modulation in its local spin correlation pattern. As is illustrated in Fig.12, the zigzag pattern manifests itself most evidently in the antiferromagnetic correlated backbones running through the 2𝐚1𝐚22\mathbf{a}_{1}-\mathbf{a}_{2} or equivalent directions of the triangular lattice. The translational symmetry and the rotational symmetry are thus spontaneously broken. Looking more closely, we find that the zigzag pattern in the 0.045J40.0550.045\leq J_{4}\leq 0.055 regime differs from that in the 0.055J40.090.055\leq J_{4}\leq 0.09 regime by an additional two-fold translational symmetry breaking. Fig.13 shows the optimized variational energy for the zigzag phase. The energy difference between the SFS state and the zigzag state is even more significant than that in the 4×64\times 6 phase. A closer inspection also shows that the energy difference between the U(1)U(1) and the Z2Z_{2} state is more significant in the zigzag regime than that in the π\pi-flux and the large J4J_{4} regime.

Refer to caption
Figure 13: The variational energies of the three kinds of RVB state as functions of J4J_{4} in the range of 0.045J40.090.045\leq J_{4}\leq 0.09. The computation is done on a 12×1212\times 12 cluster with periodic boundary condition. The optimized variational ground state in this regime is found to exhibit a zigzag local spin correlation pattern. The zigzag pattern in the 0.045J40.0550.045\leq J_{4}\leq 0.055 regime differs from that in the 0.055J40.090.055\leq J_{4}\leq 0.09 regime by an additional two-fold translational symmetry breaking.

Instead of the zigzag phase found here, several translational invariant phases have been proposed in the intermediate J4J_{4} regime in the literature. In Ref.[Grover, ], a nodal dd-wave Z2Z_{2} state with nematic spinon dispersion is proposed to be the variational ground state in the intermediate regime. It is later found that another Z2Z_{2} state with d+idd+id-wave spinon pairing and quadratic band touching(QBT) in the spinon dispersion is more favorable than the nodal d-wave state in a tiny range of J4J_{4}Xu . The QBT state hosts gapless spinon excitation with a finite density of state and a gapped gauge fluctuation spectrum as a result of the d+idd+id-wave spinon pairing. Such an excitation characteristics is argued to be helpful to resolve the puzzle related to the anomalous gauge fluctuation correction to the specific heat in the SFS state. While we think that the anomalous gauge fluctuation correction in the SFS state is only a theoretical artifact of the conventional gauge theory argumentLi2 , it is nevertheless interesting to compare the variational energies of these novel states to our variational result. In Fig.14, we plot the variational energies of the major candidates of the variational ground state of the J1J4J_{1}-J_{4} model in the intermediate J4J_{4} regime. From this figure we see that the energy advantage of the QBT state over the nodal dd-wave state is meaningless when compared to the huge energy difference between these states and the zigzag state we find from variational optimization. This establishes firmly the zigzag nature of the variational ground state in the intermediate regime.

A zigzag phase has also been reported in a similar parameter regime in a recent DMRG study of the J1J2J4J_{1}-J_{2}-J_{4} modelMoore , in which J2J_{2} denotes the exchange coupling between next nearest neighboring spins. Different from what we found here, the zigzag phase reported in Ref.[Moore, ] exhibits magnetic long range order. The Fermionic RVB state we adopted in this study is not expected to describe accurately such magnetic long range ordered phases. However, as we have seen in the case of the π\pi-flux phase for 0J40.0450\leq J_{4}\leq 0.045, the Fermionic RVB state can nevertheless reproduce correctly the qualitative feature of the spin structure factor of the magnetic ordered phase. We find that this is also the case in the zigzag phase. In Fig.15 we plot the spin structure factor of the model at J4=0.06J_{4}=0.06 calculated from the optimized gRVB statenote1 . The spin structure factor is characterized by the prominent peaks at 𝐪=(±π2,π)\mathbf{q}=(\pm\frac{\pi}{2},\pi) and the weaker peak at 𝐪=(π,0)\mathbf{q}=(\pi,0). These are exactly the positions of the magnetic Bragg peaks in the classical zigzag phase(illustrated in Fig15c).

With these considerations in mind, it is better to interpret the π\pi-flux phase in the small J4J_{4} regime and the zigzag phase in the intermediate J4J_{4} regime both as the closest approximation of the corresponding magnetic ordered phases, namely the 120 degree ordered phase and the zigzag ordered phase. Thus, the transition at J4=0.045J_{4}=0.045 should be better understood as a first order transition between two magnetic ordered phases, rather than the transition between a spin liquid phase and a valence bond solid phase. According to Ref.[Xu, ], the 120 degree ordered phase becomes degenerate with the nodal d-wave state at J4=0.0525J_{4}=0.0525. Since the energy of the zigzag phase found here is lower than that of the nodal d-wave state at J4=0.0525J_{4}=0.0525, we expect that the transition between the 120 degree ordered phase and the zigzag phase to occur at a smaller value of J4J_{4} than 0.05250.0525. This is indeed the case in our calculation.

Refer to caption
Figure 14: The variational energies of the major candidates of the variational ground state of the J1J4J_{1}-J_{4} model in the intermediate regime. The inset shows an enlarged view of the crossing region of the nodal dd-wave and the d+idd+id QBT state. The energy of the nodal dd-wave and the QBT state are taken from Ref.[Xu, ]. The results of this work are computed on a 12×1212\times 12 cluster with periodic boundary condition.
Refer to caption
Refer to caption
Refer to caption
Figure 15: (a)The spin structure factor of the J1J4J_{1}-J_{4} model at J4=0.06J_{4}=0.06 calculated from the optimized gRVB state. (b)The spin structure factor in a classical zigzag phase. (c)Illustration of the classical zigzag phase on the triangular lattice. The thick blue bonds highlight the antiferromagnetic correlated backbones in the zigzag phase. The spin structure factor shown in (b) is calculated assuming that cotϕ=13\cot\phi=\frac{1}{3} and that the spin has a unit length.

V.4 Comparison with the phase diagram obtained from exact diagonalization

To provide further support to the variational results, we have performed exact diagonalization(ED) calculation on a 6×66\times 6 cluster with periodic boundary condition. Here we focus on the identity representation. Using translational and point group symmetry, together with the spin rotational symmetry along the zz-axis, we can reduce the number of symmetrized basis to 31,554,903. The Hamiltonian matrix is then diagonalized by the Lanczos method to obtain its lowest few eigenvalues. Fig.16 shows the ground state energy as a function of J4J_{4} obtained from the ED calculation. Two features of the ground state energy curve are of particular interest to us. First, the curvature of the ground state energy is very small in both the small J4J_{4} and the large J4J_{4} regime, a trend in good agreement with the variational result presented above. Second, in the intermediate J4J_{4} regime, level crossings between the ground state and the first excited state are observed, implying potential first order transition in the thermodynamic limit. Such transitions may be closely related to the appearance of the zigzag phase we obtained in the variational study. Clearly, to make more solid conclusion on this issue, more systematic ED calculation is needed.

Refer to caption
Refer to caption
Figure 16: (a)The ground state energy of the J1J4J_{1}-J_{4} model calculated from exact diagonalization on a 6×66\times 6 cluster with periodic boundary condition(black dots). This is compared with the variational energies of the π\pi-flux phase(green dots), the SFS phase(blue dots) and the optimized Z2Z_{2} RVB state(red dots) we obtained on a 12×1212\times 12 cluster with periodic boundary condition. (b)The lowest two eigenvalues in the identity representation of the 6×66\times 6 cluster cross at both J4=0.049J_{4}=0.049 and J4=0.059J_{4}=0.059, implying potential first order transition there in the thermodynamic limit.

We note that the finite size effect on a 6×66\times 6 cluster can be very significant, especially when the considered state is gapless. For example, if we perform variational optimization on a 6×66\times 6 cluster with periodic boundary condition, what we would obtain at J4=0.15J_{4}=0.15 is a symmetry broken state with a 3×3\sqrt{3}\times\sqrt{3} modulation in its local spin correlation pattern( as is illustrated in Fig.17), rather the 4×64\times 6 state we find on the 12×1212\times 12 cluster. The modulation of the local spin correlation in the 3×3\sqrt{3}\times\sqrt{3} state is found to be much weaker than that in the 4×64\times 6 state. Such a result is partially expected since the 4×64\times 6 state is incompatible with the periodic boundary condition of the 6×66\times 6 cluster. However, when we extend the optimized ansatz of such a 3×3\sqrt{3}\times\sqrt{3} state to a 12×1212\times 12 cluster, we find that the variational energy is very bad. On the other hand, when we extend the optimized ansatz of the 4×64\times 6 state on the 12×1212\times 12 cluster to a 24×2424\times 24 cluster, the variational energy is essentially unchanged. Fig.18 illustrates such contrasting behavior of the 3×3\sqrt{3}\times\sqrt{3} state and the 4×64\times 6 state. Thus, the dominance of the nearly uniform 3×3\sqrt{3}\times\sqrt{3} state on the 6×66\times 6 cluster is purely a finite size effect. This indicates that a sufficiently large cluster is needed to draw conclusion on the phase diagram of model.

Refer to caption
Figure 17: The local spin correlation pattern of the optimized variational ground state at J4=0.15J_{4}=0.15 on a 6×66\times 6 cluster with periodic boundary condition. This state exhibits a 3×3\sqrt{3}\times\sqrt{3} local spin correlation pattern, rather than the 4×64\times 6 pattern we find above on the 12×1212\times 12 cluster. The magnitude of the modulation in the local spin correlation is also much smaller than that in the 4×64\times 6 state.
Refer to caption
Refer to caption
Figure 18: The contrasting behavior of the 3×3\sqrt{3}\times\sqrt{3} state and the 4×64\times 6 state during the optimization procedure at J4=0.15J_{4}=0.15. (a)While the 3×3\sqrt{3}\times\sqrt{3} state is the best variational state at J4=0.15J_{4}=0.15 on a 6×66\times 6 cluster, it generates a very bad variational energy on a 12×1212\times 12 cluster. (b)On the other hand, the variational energy of the 4×64\times 6 state is essentially unchanged when the calculation is extended from the 12×1212\times 12 cluster to the much larger 24×2424\times 24 cluster. This contrasting behavior indicates that the dominance of the 3×3\sqrt{3}\times\sqrt{3} state on the 6×66\times 6 cluster is purely a finite size effect.

VI Discussions and conclusions

While it is generally believed that the long-sought U(1)U(1) spin liquid with a large spinon Fermi surface can be realized in the large J4J_{4} regime of the J1J4J_{1}-J_{4} model on the triangular lattice, recent DMRG simulation indicates that the expected spin liquid phase may be replaced by some symmetry breaking phase in the real phase diagram of this model. The tension between such a DMRG result and the abundant variational results on this model calls for a systematic variational study of the J1J4J_{1}-J_{4} model without assuming any symmetry a prior. This is a formidable numerical task. With the increase of the number of variational parameters, the variational optimization procedure becomes increasingly tricky as a result of the abundance of local minimum in the energy landscape and/or the large fluctuation in the eigenvalues of the Hessian matrix. The J1J4J_{1}-J_{4} model is a typical example in this regard. As we have shown in the last section, the SFS state is an extremely stable local minimum of the variational energy of this model as a result of its special symmetry properties.

To map out the genuine ground state phase diagram of the J1J4J_{1}-J_{4} model on the triangular lattice, we have proposed several improvements on the variational optimization algorithm. The key to such improvements is a better approximation of the Hessian matrix with the gradient information. We find that the finite depth BFGS algorithm has the best balance between numerical efficiency and stability. It can often proceed further the optimization procedure when the conventional steepest descent and the stochastic reconfiguration algorithm get stuck.

We have used the improved algorithms to optimize three kinds of Fermionic RVB state for the J1J4J_{1}-J_{4} model on the triangular lattice, namely, a general RVB state whose RVB amplitudes are treated directly as variational parameters, a U(1)U(1) RVB state generated from Gutzwiller projection of the ground state of a mean field Hamiltonian with only Fermion hopping terms, and a Z2Z_{2} RVB state generated from Gutzwiller projection of the ground state of a BCS-type mean field Hamiltonian. We get consistent results from all these three kinds of RVB state on the ground state phase diagram of the J1J4J_{1}-J_{4} model on the triangular lattice.

From our variational optimization, we find that for 0J40.0450\leq J_{4}\leq 0.045, the best variational state of the model is the well known π\pi-flux phase on the triangular lattice. This state is the closest approximation of the 120 degree ordered phase on the triangular lattice within the subspace of the Fermionic RVB state, as can be seen from its static spin structure factor. Indeed, such a Dirac spin liquid state can be tuned continuously into a Bosonic RVB stateSeiji , which provides an extremely accurate description of the ground state of the antiferromagnetic Heisenberg model on the triangular latticeZhangQ . We find that the variational state for 0J40.0450\leq J_{4}\leq 0.045 is almost independent of J4J_{4}, consistent with the observation that the variational energy has nearly zero curvature as a function of J4J_{4}. Such an observation agrees well with the result obtained from ED calculation on a 6×66\times 6 cluster. Thus, while the π\pi-flux phase does not posses true magnetic long range order, it nevertheless captures correctly the spin correlation pattern in the 120 degree ordered phase. We thus expect that an extended variational calculation involving magnetic long range order will not change the structure of the phase diagram qualitatively.

For J40.09J_{4}\geq 0.09, a regime which is thought to be the most favorable for the SFS state, we find that the optimized variational ground state exhibits a 4×64\times 6 modulation in its local spin correlation pattern. The magnitude of the modulation is found to be very large, ranging from nearly pure spin singlet correlation to nearly pure spin triplet correlation between nearest neighboring spins. The energy gain related to such a symmetry breaking is found to be quite large. More specifically, the energy gain of the 4×64\times 6 state over the SFS state is found to be about 5%5\% in the whole J40.09J_{4}\geq 0.09 regime. We find that this conclusion is robust on larger clusters. In addition, we find that the curvature in the variational energy is also very small in the J40.09J_{4}\geq 0.09 regime, implying the nearly J4J_{4}-independent nature of the variational ground state. These results indicate collectively that the 4×64\times 6 phase is the variational ground state in the whole J40.09J_{4}\geq 0.09 regime.

The variational ground state of the intermediate regime of 0.045J40.090.045\leq J_{4}\leq 0.09 is found to exhibit zigzag modulation in its local spin correlation pattern. Depending on if an additional translational symmetry is broken or not, the intermediate regime can be further divided into a type-I zigzag phase for 0.045J40.0550.045\leq J_{4}\leq 0.055 and a type-II zigzag phase for 0.055J40.090.055\leq J_{4}\leq 0.09. All these different phases are found to be connected by first order transitions. We find that the variational energy of the zigzag phase is not only much lower than that of the SFS state, but is also much lower than that of other previously proposed variational states, in particular, the nodal dd-wave phase and the d+idd+id-wave phase with quadratic band touching in spinon dispersion.

Taken all these results together, we conclude that while the SFS state is locally extremely stable, it is never the true variational ground state of the J1J4J_{1}-J_{4} model on the triangular lattice. In addition, no translational symmetric spin liquid state can be stabilized in the J1J4J_{1}-J_{4} model. We find that the energy advantage by breaking the translational symmetry can be very significant. Compared to such large energy gain, the tiny energy difference between the nodal dd-wave state and the d+idd+id-wave state with quadratic band touching in the spinon dispersion appears meaningless. Thus, any serious variational study of the ground state phase diagram of the J1J4J_{1}-J_{4} model should take into account the possibility of translational symmetry breaking. We think that this is true not only for the particular model studied in this paper, but also applies in the variational study of general highly frustrated quantum antiferromagnetic models. The BFGS algorithm proposed in this paper thus has much broader range of applications in future study of quantum spin liquids.

Another lesson that we can learn from our result is that the finite size effect may significantly distort the phase diagram of a highly frustrated spin model on a finite cluster, since the model may develop symmetry breaking pattern with a rather large unit cell in its ground state. To extract the genuine behavior of the model in the thermodynamic limit, sufficiently large system(or sufficiently wide system in the case of DMRG simulation) is needed. This calls for new developments in the variational optimization algorithm and the DMRG algorithm.

Finally, while the SFS state is unstable against the 4×64\times 6 state in the large J4J_{4} regime, it may still be stabilized in real materials by additional couplings that frustrate the 4×64\times 6 modulation pattern. A thorough study involving other major spin couplings that can be generated from the strong coupling expansion of the Hubbard model on the triangular latticeSchmidt is necessary to fully address this problem. We leave such a study to future works.

Acknowledgements.
We acknowledge the support from the National Natural Science Foundation of China(Grant No.12274457).

References

  • (1) P. A. Lee, Science 321, 1306 (2008).
  • (2) L. Balents, Nature 464, 199(2010).
  • (3) B. Dalla Piazza, M. Mourigal, N. B. Christensen, G. J. Nilson, P. Tregenna-Piggott, T. G. Perring, M. Enderle, D. F. McMorrow, D. A. Ivanov and H. M. Rø{\o}nnow, Nat. Phys. 11, 62(2015).
  • (4) S. Ito, N. Kurita, H. Tanaka, S. Ohira-Kawamura, K. Nakajima, S. Itoh, K. Kuwahara and K. Kakurai, Nat. Commun. 8, 235 (2017).
  • (5) F. Ferrari and F. Becca, Phys. Rev. X 9, 031026 (2019).
  • (6) Chun Zhang and Tao Li, Phys. Rev. B 102, 075108 (2020).
  • (7) P. A. Lee, N. Nagaosa, Phys. Rev. B 46 5621 (1992).
  • (8) Y. Shimizu, K. Miyagawa, K. Kanoda, M. Maesato, G. Saito, Phys. Rev. Lett. 91, 107001 (2003).
  • (9) S. Yamashita, Y. Nakazawa, M. Oguni, Y. Oshima, H. Nojiri, Y. Shimizu, K. Miyagawa, and K. Kanoda, Nature Phys. 4, 459 (2008).
  • (10) Y. Zhou, K. Kanoda, T.-K. Ng, Rev. Mod. Phys. 89, 025003 (2017).
  • (11) M. Yamashita, N. Nakata, Y. Senshu, M. Nagata, H. M Yamamoto, R. Kato, T. Shibauchi, Y. Matsuda, Science 328,1246, (2010).
  • (12) P. Bourgeois-Hope, F. Laliberté, E. Lefrançois, G. Grissonnanche, S. Renéde Cotret, R. Gordon, S. Kitou, H. Sawa, H. Cui, R. Kato, L. Taillefer, and N. Doiron-Leyraud, Phys. Rev. X 9, 041051 (2019).
  • (13) J. M. Ni, B. L. Pan, Y. Y. Huang, J. Y. Zeng, Y. J. Yu, E. J. Cheng, L. S. Wang, R. Kato, and S. Y. Li, Phys. Rev. Lett. 123, 247204 (2019).
  • (14) K. T. Law and P. A. Lee, PNAS 114, 6996 (2017).
  • (15) A. Ribak, I. Silber, C. Baines, K. Chashka, Z. Salman, Y. Dagan, and A. Kanigel, Phys. Rev. B 96, 195131 (2017).
  • (16) M. Klanjšek, A. Zorko, R. Žitko, J. Mravlje, Z. Jagličić, P. K. Biswas, P. Prelovšek, D. Mihailovic, and D. Arčon, Nat. Phys. 13, 1130 (2017).
  • (17) Y. Li, H. Liao, Z. Zhang, S. Li, F. Jin, L. Ling, L. Zhang, Y. Zou, L. Pi, Z. Yang, J. Wang, Z. Wu, and Q. Zhang, Sci. Rep. 5, 16419 (2015).
  • (18) O. I. Motrunich Phys. Rev. B 72, 045105 (2005).
  • (19) S.-S. Lee, P. A. Lee, Phys. Rev. Lett. 95, 036403 (2005).
  • (20) T. Grover, N. Trivedi, T. Senthil, and P. A. Lee, Phys. Rev. B 81, 245121 (2010).
  • (21) R. V. Mishmash, J. R. Garrison, S. Bieri, C. Xu, Phys. Rev. Lett. 111, 157203 (2013).
  • (22) Q. R. Zhao and Z. X. Liu, Phys. Rev. Lett. 127, 127205 (2021).
  • (23) L. W. He and J. X. Li, arXiv:2306.03478.
  • (24) W. Y. He, X. Y. Xu, G. Chen, K. T. Law, and P. A. Lee, Phys. Rev. Lett. 121, 046401 (2018).
  • (25) T. Cookmeyer, J. Motruk, and J. E. Moore, Phys. Rev. Lett. 127, 087201 (2021).
  • (26) W. LiMing, G. Misguich, P. Sindzingre, and C. Lhuillier, Phys. Rev. B 62, 6372 (2000).
  • (27) H. Y. Yang, A. M. Läuchli, F. Mila, and K. P. Schmidt, Phys. Rev. Lett. 105, 267204 (2010).
  • (28) K. Seki and S.Yunoki, Phys. Rev. B 101, 235115 (2020).
  • (29) Qiu Zhang and Tao Li, J. Phys.: Condens. Matter 33, 375601(2021).
  • (30) S.Yunoki and S. Sorella, Phys. Rev. B 74, 014408 (2006).
  • (31) P. W. Anderson, Mater. Res. Bull. 8, 153(1973).
  • (32) M. Barkeshli, H. Yao, S. A. Kivelson, Phys. Rev. B 87, 140402(R) (2013).
  • (33) Tao Li, Phys. Rev. B 104, 165123(2021).
  • (34) J. Nocedal and S. J. Wright, NumericalOptimizationNumerical\ Optimization, Springer(2008).
  • (35) We note that the spin correlation function in the zigzag phase is not translational invariant. Thus its Fourier transform is not diagonal in wave vector. The spin structure factor we plot in Fig.15 corresponds to the diagonal part of the Fourier component of the spin correlation function.