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

Transition from band insulator to Mott insulator and formation of local moment in half-filled ionic SU(NN) Hubbard model

Shan-Yue Wang National Laboratory of Solid State Microstructures and School of Physics, Nanjing University, Nanjing, 210093, China    Da Wang National Laboratory of Solid State Microstructures and School of Physics, Nanjing University, Nanjing, 210093, China Collaborative Innovation Center of Advanced Microstructures, Nanjing University, Nanjing 210093, China    Qiang-Hua Wang National Laboratory of Solid State Microstructures and School of Physics, Nanjing University, Nanjing, 210093, China Collaborative Innovation Center of Advanced Microstructures, Nanjing University, Nanjing 210093, China
Abstract

We investigate the local moment formation in the half-filled SU(NN) Hubbard model under a staggered ionic potential. As the Hubbard UU increases, the charge fluctuations are suppressed and eventually frozen when UU is above a critical value UcU_{c}, marking the development of well-defined local moment with integer mm fermions on the A-sublattice and (Nm)(N-m) fermions on the B-sublattice, respectively. We obtain an analytical solution for UcU_{c} for the paramagnetic ground state within the variational Gutzwiller approximation and renormalized mean field theory. For large NN, UcU_{c} is found to depend on NN linearly with fixed m/Nm/N, but sublinearly with fixed mm. The local moment formation is accompanied by a peculiar phase transition from the band insulator to the Mott insulator, where the ionic potential and quasiparticle weight are renormalized to zero simultaneously. Inside the Mott phase, the low energy physics is described by the SU(NN) Heisenberg model with conjugate representations, which is widely studied in the literature.

I Introduction

Quantum spin models are a class of physical models describing “spins” or “local moments” which originate from strong correlations between fermions (e.g. electrons or cold atoms) such that the “charge” (fermion number) degrees of freedom are frozen [1]. For instance, the Heisenberg model is a low energy description of the Hubbard model only when the Hubbard UU is large enough to drive the system into the Mott insulating phase [2, 3]. In the literature, the SU(2) Heisenberg model has been generalized to the SU(NN) case [4] with the spin operators satisfying the SU(NN) algebra. The SU(NN) Heisenberg models provide a vast area to explore many new phenomena [5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30]. Among different SU(NN) representations for the local spins, a conjugate representation with mm fermions on A-sublattice and (Nm)(N-m) fermions on B-sublattice [4] is mostly studied. It may support a generalized Neel order: for instance, the first mm and remaining (Nm)(N-m) flavors are occupied on the two sublattices, respectively. Hence, the SU(NN) symmetry is broken into SU(m)×SU(Nm)\mathrm{SU}(m)\times\mathrm{SU}(N-m). The gapless fluctuations above this Neel order fall into the Grassmannian manifold Gr(N,m)=U(N)/[U(m)×U(Nm)]\mathrm{Gr}(N,m)=\mathrm{U}(N)/[\mathrm{U}(m)\times\mathrm{U}(N-m)] [31, 32, 33, 34], which is reduced to the NN-component Ginzburg-Landau theory [35] or equivalently the famous CPN-1 model in the special case of m=1m=1.

One early motivation for doing the SU(NN) generalization of Heisenberg model is to perform 1/N1/N expansion around the saddle point at N=N=\infty, providing a controllable way to reach the SU(2) model [36]. However, physically speaking, the SU(NN) spin models should derive from the SU(NN) Hubbard model in the limit that charge fluctuations are completely frozen. The SU(NN) Hubbard model is widely studied,[37, 38, 39, 40, 41, 42, 43, 44, 45, 46, 47, 48, 49, 50], and is now within experimental reach, thanks to the fast technical development, mostly in the field of cold atoms [51, 52, 53, 54, 55, 56, 57, 58, 59, 60, 61, 62, 63, 64]. This brings the large-NN model to life, but not just a gedanken model, and opens up a new field in the study of the finite but large NN versions of such models, in the search for novel quantum spin states.

However, it is important to ask under what condition is the system aptly described by the quantum spin model for which the local moments have to be well established. In our previous work, we have proposed to add a staggered ionic potential to the SU(NN) Hubbard model [65]. In this work, we will examine the specific conditions for the Hubbard UU and ionic potential VV under which the local moments can exist, with immediate relevance to experimental realization. We develop and apply an SU(NN)-symmetric renormalized mean field theory (RMFT) based on the variational approach of the Gutzwiller projection approximation [66, 67, 2, 68, 69, 70, 71]. The RMFT developed here may be applied or extended straightforwardly for general models with a large number of fermion flavors subject to any internal symmetry. For the ionic SU(NN) Hubbard model, we find the local moments, with quantized integer mm fermions on the A-sublattice and (Nm)(N-m) fermions on the B-sublattice, are well established when UU is above a critical value UcU_{c}, which depends on NN, mm, and the ionic potential VV. For large NN, UcU_{c} is found to depend on NN linearly with fixed m/Nm/N, but sublinearly with fixed mm. In addition, the local moment formation is accompanied by a peculiar transition from the band insulator to the Mott insulator [72, 73], at which the ionic potential and quasiparticle weight are renormalized to zero simultaneously. Finally, we show that the low energy physics of local moments is described by the widely studied SU(NN) quantum spin model inside the Mott insulating phase. Our results shed light on the realization of such models in, e.g., cold atoms.

II SU(NN) Hubbard model in an ionic potential

Th ionic SU(NN) Hubbard model we consider is described by the Hamiltonian H=Ht+HUH=H_{t}+H_{U}, with

Ht\displaystyle H_{t} =\displaystyle= tij,a[ca(i)ca(j)+H.c.],\displaystyle-t\sum_{\langle ij\rangle,a}[c_{a}^{\dagger}(i)c_{a}(j)+{\rm H.c.}], (1)

and

HU\displaystyle H_{U} =\displaystyle= U2i[n^(i)N2]2+Vi(1)in^(i),\displaystyle\frac{U}{2}\sum_{i}\left[\hat{n}(i)-\frac{N}{2}\right]^{2}+V\sum_{i}(-1)^{i}\hat{n}(i), (2)

where ii labels the lattice site, ij\langle ij\rangle denotes a nearest-neighbor bond, a=1,2,,Na=1,2,\cdots,N labels the flavor of the fermions, n^(i)=an^a(i)=aca(i)ca(i)\hat{n}(i)=\sum_{a}\hat{n}_{a}(i)=\sum_{a}c_{a}^{\dagger}(i)c_{a}(i) is the local density operator, UU denotes the Hubbard interaction, and VV is the staggered ionic potential. The model is clearly SU(NN)-symmetric in the internal flavor space. In real space, it breaks the translational symmetry, since A- and B-sublattices are distinct. However, a particle-hole transformation cia(1)iciac_{ia}\to(-1)^{i}c_{ia}^{\dagger} interchanges these sublattices, so the system is invariant under A-B sublattice transformation combined with the particle-hole transformation. As a result, the charge density is exactly staggered and the system is at half filling on average. Such a symmetry can be used to reduce the variational parameters, as can be seen in later discussions.

The HUH_{U}-term can be rewritten as

HU=U2i[n^(i)n0(i)]2,\displaystyle H_{U}=\frac{U}{2}\sum_{i}\left[\hat{n}(i)-n_{0}(i)\right]^{2}, (3)

where n0(i)=N2(1)iVUn_{0}(i)=\frac{N}{2}-(-1)^{i}\frac{V}{U} may be understood as the local ground charge tunable continuously by the staggered gate voltage VV. We ask whether n^(i)\hat{n}(i) can be quantized to integers mm (0<m<N0<m<N) on the A-sublattice and (Nm)(N-m) on the B-sublattice, i.e., forming local moments, called mm-tuple moments, for large enough UU. The concept of local moment is a natural generalization of the SU(2) case for which only one kind of local moment with m=1m=1 is possible. Here, however, the states with different mm-tuple moments should belong to different Mott insulating states. On the other hand, in the limit of U=0U=0, a nonzero VV always yields a band insulator. For large enough UU, the system is expected to enter different Mott insulating states labeled by mm. Whether these mm-tuple moments exist and how to describe these band-to-Mott insulator transitions are the main concerns of the present work. To answer these questions clearly, and for simplicity, we shall focus on the paramagnetic case in the following.

III SU(NN)-symmetric Gutzwiller projection approximation and RMFT

Local moment formation is beyond any Hartree-Fock mean field description. We employ the standard Gutzwiller projection approximation to treat the correlation effect. The SU(2) version of such a theory has been applied widely [67, 2, 68, 70, 71], and will be extended here to the SU(NN)-symmetric case in which all the NN-flavors are equivalent. For sufficient generality, we present the theory for an arbitrary case of the applied potential in this section, and will specify the ionic potential in the next section.

Specifically, we consider a variational theory with the following trial wave function,

|ψ=𝒫^|ψ0,\displaystyle|\psi\rangle=\hat{{\cal P}}|\psi_{0}\rangle, (4)

where |ψ0|\psi_{0}\rangle is the ground state of a free variational Hamiltonian to be specified, and 𝒫^\hat{{\cal P}} is the Gutzwiller projection operator in the grand canonical ensemble

𝒫^=Πi𝒫^i,𝒫^i=k=0Nηk(i)yikQ^k(i),\displaystyle\hat{{\cal P}}=\Pi_{i}\hat{{\cal P}}_{i},\ \ \hat{{\cal P}}_{i}=\sum_{k=0}^{N}\eta_{k}(i)y_{i}^{k}\hat{Q}_{k}(i), (5)

where Q^k(i)\hat{Q}_{k}(i) is the projection operator for the kk-tuple state (with kk-fermions),

Q^k(i)=S={a|=1,,k}bSn^b(i)bS[1n^b(i)].\displaystyle\hat{Q}_{k}(i)=\sum_{S=\{a_{\ell}|\ell=1,\cdots,k\}}\prod_{b\in S}\hat{n}_{b}(i)\prod_{b\notin S}[1-\hat{n}_{b}(i)]. (6)

Clearly, in the absence of projection, we have 𝒫^i=kQ^k(i)=1\hat{{\cal P}}_{i}=\sum_{k}\hat{Q}_{k}(i)=1. The idea of the Gutzwiller projection is to reassign weights to the basis states, and this is how correlations (at least the local ones) can be captured. Here, the weight for the kk-tuple is assumed to be

ηk(i)yik=exp(gik2+klnyi)\displaystyle\eta_{k}(i)y_{i}^{k}=\exp(-g_{i}k^{2}+k\ln y_{i}) (7)

where gig_{i} is the site-dependent Gutzwiller projection parameter to punish multi-fermion occupations (but can be chosen to be uniform in our case). In the spirit of density functional theory [74], the ground state energy is a unique functional of the density distribution. Therefore we have introduced a fugacity yiy_{i} to maintain the fermion density before and after projection in the grand canonical ensemble [70]. The fermion density on each site before projection is

Nfi=n^(i)0=kkqk0(i),\displaystyle Nf_{i}=\langle\hat{n}(i)\rangle_{0}=\sum_{k}kq_{k0}(i), (8)

where fif_{i} is the average occupation number per flavor, and 0\langle\cdot\rangle_{0} indicates the average performed with respect to |ψ0|\psi_{0}\rangle. We have defined qk0(i)q_{k0}(i) as the average of Q^k(i)\hat{Q}_{k}(i) in the unprojected state,

qk0(i)=Q^k(i)0=CNkfk(i)[1f(i)]Nk,\displaystyle q_{k0}(i)=\langle\hat{Q}_{k}(i)\rangle_{0}=C_{N}^{k}f^{k}(i)[1-f(i)]^{N-k}, (9)

where CNkC_{N}^{k} is the combinatorial factor. After projection, we still require

Nfi=n^(i)=kkqk(i),\displaystyle Nf_{i}=\langle\hat{n}(i)\rangle=\sum_{k}kq_{k}(i), (10)

with

qk(i)=Q^k(i)=𝒫^Q^k(i)𝒫^0𝒫^𝒫^0,\displaystyle q_{k}(i)=\langle\hat{Q}_{k}(i)\rangle=\frac{\langle\hat{{\cal P}}\hat{Q}_{k}(i)\hat{{\cal P}}\rangle_{0}}{\langle\hat{{\cal P}}\hat{{\cal P}}\rangle_{0}}, (11)

where \langle\cdot\rangle denotes average with respect to |ψ|\psi\rangle. Exact evaluation on the right hand side is difficult. To make analytical progress, we resort to the usual Gutzwiller approximation [67]: the projection operator unrelated to the target operator under average can be Wick-contracted separately. This approximation can be shown to be exact in infinite dimensions [75] for general on-site interactions [69] , and turns out to work satisfactorily in finite dimensions [68, 69, 71]. Under the Gutzwiller approximation, we have

qk(i)=𝒫^iQ^k(i)𝒫^i0𝒫^i𝒫^i0.\displaystyle q_{k}(i)=\frac{\langle\hat{{\cal P}}_{i}\hat{Q}_{k}(i)\hat{{\cal P}}_{i}\rangle_{0}}{\langle\hat{{\cal P}}_{i}\hat{{\cal P}}_{i}\rangle_{0}}. (12)

Note the projection operator 𝒫{\cal P} is simplified to 𝒫i{\cal P}_{i}. After substituting 𝒫^i\hat{{\cal P}}_{i} in Eq. 5, we obtain

qk(i)=ηk2(i)yi2kqk0(i)𝒟i,𝒟i=kηk2(i)yi2kqk0(i),\displaystyle q_{k}(i)=\frac{\eta_{k}^{2}(i)y_{i}^{2k}q_{k0}(i)}{{\cal D}_{i}},\,\,{\cal D}_{i}=\sum_{k}\eta_{k}^{2}(i)y_{i}^{2k}q_{k0}(i), (13)

where we have used Q^k2(i)=Q^k(i)\hat{Q}_{k}^{2}(i)=\hat{Q}_{k}(i) as a property of the projection operator Q^k(i)\hat{Q}_{k}(i). The fugacity yiy_{i} (or lnyi\ln y_{i} in practice) is then tuned to satisfy the density restriction Eq. 10.

After obtaining all of qk0(i)q_{k0}(i) and qk(i)q_{k}(i), we are in a position to evaluate the variational energy in the projected state under the Gutzwiller approximation. The local charging energy is obtained most straightforwardly, given the fact that the total charge operator and the projectors Q^k(i)\hat{Q}_{k}(i) share the same local basis states as eigenstates,

EU=HU=U2ik=0N[kn0(i)]2qk(i).\displaystyle E_{U}=\langle H_{U}\rangle=\frac{U}{2}\sum_{i}\sum_{k=0}^{N}\left[k-n_{0}(i)\right]^{2}q_{k}(i). (14)

The kinetic energy is slightly more difficult to evaluate. Since the fermion hopping involves two sites, we need to keep two projectors, say, 𝒫^i\hat{{\cal P}}_{i} and 𝒫^j\hat{{\cal P}}_{j} in the hopping on the ij\langle ij\rangle bond,

ca(i)ca(j)=𝒫^ica(i)𝒫^i𝒫^jca(j)𝒫^j0𝒫^i2𝒫^j20.\displaystyle\langle c_{a}^{\dagger}(i)c_{a}(j)\rangle=\frac{\langle\hat{{\cal P}}_{i}c_{a}^{\dagger}(i)\hat{{\cal P}}_{i}\hat{{\cal P}}_{j}c_{a}(j)\hat{{\cal P}}_{j}\rangle_{0}}{\langle\hat{{\cal P}}_{i}^{2}\hat{{\cal P}}_{j}^{2}\rangle_{0}}. (15)

Since the fermion operator is self-projective, we need to remove over projections before taking the quantum average. For a given site ii, we observe that

𝒫^ica(i)𝒫^i\displaystyle\hat{{\cal P}}_{i}c_{a}(i)\hat{{\cal P}}_{i} =\displaystyle= k[ηk(i)yikQ^k(i)]ca(i)[ηk+1(i)yik+1Q^k+1(i)]\displaystyle\sum_{k}[\eta_{k}(i)y_{i}^{k}\hat{Q}_{k}(i)]c_{a}(i)[\eta_{k+1}(i)y_{i}^{k+1}\hat{Q}_{k+1}(i)] (16)
\displaystyle\equiv k[ηk(i)ηk+1(i)yi2k+1Q^ka^(i)]ca(i),\displaystyle\sum_{k}[\eta_{k}(i)\eta_{k+1}(i)y_{i}^{2k+1}\hat{Q}_{k}^{\hat{a}}(i)]c_{a}(i),

where we have defined a partial projection operator Qka^(i)Q_{k}^{\hat{a}}(i) for kk fermions in the local Fock space excluding flavor aa,

Q^ka^(i)=S={a|=1,,k;aa}bSn^b(i)bS,ba[1n^b(i)].\displaystyle\hat{Q}_{k}^{\hat{a}}(i)=\sum_{S=\{a_{\ell}|\ell=1,\cdots,k;a_{\ell}\neq a\}}\prod_{b\in S}\hat{n}_{b}(i)\prod_{b\notin S,b\neq a}[1-\hat{n}_{b}(i)]. (17)

Its average in |ψ0|\psi_{0}\rangle is evaluated to be

qk0a^(i)=Q^ka^(i)0=CN1kfik(1fi)N1k,\displaystyle q_{k0}^{\hat{a}}(i)=\langle\hat{Q}_{k}^{\hat{a}}(i)\rangle_{0}=C_{N-1}^{k}f_{i}^{k}(1-f_{i})^{N-1-k}, (18)

for any flavor aa in our SU(NN)-symmetric case. Inserting the above relations in Eq. 15, we obtain

ca(i)ca(j)=gt(i,j)ca(i)ca(j)0,\displaystyle\langle c_{a}^{\dagger}(i)c_{a}(j)\rangle=g_{t}(i,j)\langle c_{a}^{\dagger}(i)c_{a}(j)\rangle_{0}, (19)

where gt(i,j)=z(i)z(j)g_{t}(i,j)=z(i)z(j) is the renormalization of the hopping by the projection, and

z(i)\displaystyle z(i) =\displaystyle= kηk(i)ηk+1(i)yi2k+1qk0a^(i)𝒟i\displaystyle\frac{\sum_{k}\eta_{k}(i)\eta_{k+1}(i)y_{i}^{2k+1}q_{k0}^{\hat{a}}(i)}{{\cal D}_{i}} (20)
=\displaystyle= k=0N1qk0a^(i)qk(i)qk+1(i)qk0(i)qk+1,0(i).\displaystyle\sum_{k=0}^{N-1}q_{k0}^{\hat{a}}(i)\sqrt{\frac{q_{k}(i)q_{k+1}(i)}{q_{k0}(i)q_{k+1,0}(i)}}.

In the second line we have used Eq.(13) to trade ηk(i)yik/𝒟i\eta_{k}(i)y_{i}^{k}/\sqrt{{\cal D}_{i}} for qk(i)/qk0(i)\sqrt{q_{k}(i)/q_{k0}(i)}.

Combining the potential and kinetic energies, we obtain the total variational energy EE in the projected state,

E=tijagt(i,j)χij0+U2ik=0N[kn0(i)]2qk(i),\displaystyle E=-t\sum_{\langle ij\rangle a}g_{t}(i,j)\chi_{ij0}+\frac{U}{2}\sum_{i}\sum_{k=0}^{N}\left[k-n_{0}(i)\right]^{2}q_{k}(i), (21)

where χij0=ca(i)ca(j)+H.c.0\chi_{ij0}=\left\langle c_{a}^{\dagger}(i)c_{a}(j)+{\rm H.c.}\right\rangle_{0} is the average of hopping operator in the unprojected state. This energy is understood as a functional of (i) the fermion density fif_{i} which in turn depends on the trial wavefunction |ψ0|\psi_{0}\rangle, and (ii) the Gutzwiller projection parameter gig_{i}. The fugacity parameters yiy_{i} are taken as Lagrange multipliers that are eliminated by forcing the invariance of the local fermion density under the projection. The variational Gutzwiller approximation is closely related to the RMFT. Minimizing EE with respect to |ψ0|\psi_{0}\rangle, i.e., δE/δψ0|=0\delta E/\delta\langle\psi_{0}|=0, with fixed fermion density NfiNf_{i}, we obtain

HRMFT|ψ0=|ψ0,\displaystyle H_{\rm RMFT}|\psi_{0}\rangle={\cal E}|\psi_{0}\rangle, (22)

where {\cal E} is introduced as the Lagrange multiplier enforcing normalization of the wave function, and HRMFTH_{\rm RMFT} is a free Hamiltonian yet encoded with the renormalization effect from the Gutzwiller projection,

HRMFT=tijagt(i,j)[ca(i)ca(j)+H.c.]iμin^(i),\displaystyle H_{\rm RMFT}=-t\sum_{\langle ij\rangle a}g_{t}(i,j)[c_{a}^{\dagger}(i)c_{a}(j)+{\rm H.c.}]-\sum_{i}\mu_{i}\hat{n}(i), (23)

where the variational local chemical potential μi\mu_{i} is introduced to enforce Nfi=n^(i)0Nf_{i}=\langle\hat{n}(i)\rangle_{0}. It can be shown that the single-particle spectrum of HRMFTH_{\rm RMFT} is just the quasipaticle excitation spectrum beyond the correlated variational ground state, with the quasiparticle weight renormalized by gtg_{t} [70].

IV Application to the ionic SU(NN) Hubbard model

In this section, we apply the variational Gutzwiller approximation and RMFT developed in the previous section to the ionic SU(NN) Hubbard model in our interest.

IV.1 General formalism

Due to the particle-hole and sublattice symmetries, and without involving further symmetry breaking, we only have to specify the fermion density (per flavor) ff and the Gutzwiller parameter gg on the A-sublattice. Correspondingly, we can replace f1ff\to 1-f, kNkk\to N-k, n0Nn0n_{0}\to N-n_{0}, etc., to obtain the relevant quantities on the B-sublattice, while gg remains the same. Under these simplifications, χij0\chi_{ij0} and gt(i,j)g_{t}(i,j) become bond-independent, denoted as χ0\chi_{0} and gtg_{t}, respectively. In particular, gtg_{t} is now given by

gt=(kqk0a^qkqk+1qk0qk+1,0)2.\displaystyle g_{t}=\left(\sum_{k}q_{k0}^{\hat{a}}\sqrt{\frac{q_{k}q_{k+1}}{q_{k0}q_{k+1,0}}}\right)^{2}. (24)

Due to the presence of an ionic potential, we choose μi=(1)igtΔc\mu_{i}=-(-1)^{i}g_{t}\Delta_{c} in HRMFTH_{\rm RMFT} to write

HRMFT\displaystyle H_{\rm RMFT} =\displaystyle= gttija[ca(i)ca(j)+H.c.]\displaystyle-g_{t}t\sum_{\langle ij\rangle a}[c_{a}^{\dagger}(i)c_{a}(j)+{\rm H.c.}] (25)
+gtΔci()in^(i).\displaystyle+g_{t}\Delta_{c}\sum_{i}(-)^{i}\hat{n}(i).

Under such a parametrization, gtg_{t} is a global factor renormalizing the effective bandwidth and quasiparticle excitations. The unprojected ground state |ψ0|\psi_{0}\rangle, subsequently the fermion density ff, and the average hopping χ0\chi_{0}, only depend on Δc\Delta_{c}. From HRMFTH_{\rm RMFT} and after some algebra, we obtain

f=12𝑑ερ(ε)(1Δcε2+Δc2).\displaystyle f=\frac{1}{2}\int d\varepsilon\rho(\varepsilon)\left(1-\frac{\Delta_{c}}{\sqrt{\varepsilon^{2}+\Delta_{c}^{2}}}\right). (26)
ζtχ0=𝑑ερ(ε)ε2ε2+Δc2.\displaystyle\zeta t\chi_{0}=\int d\varepsilon\rho(\varepsilon)\frac{\varepsilon^{2}}{\sqrt{\varepsilon^{2}+\Delta_{c}^{2}}}. (27)

Here ζ\zeta is the coordination number, and ρ(ε)\rho(\varepsilon) is the density of states. As an illustrative example, we consider the Bethe lattice, for which ρ(ε)=(4/πW)14ε2/W2\rho(\varepsilon)=(4/\pi W)\sqrt{1-{4\varepsilon^{2}}/{W^{2}}}, with WW the bandwidth, giving rise to

f=1212Δ~Δ~2+12F1(12,32;2;1Δ~2+1),\displaystyle f=\frac{1}{2}-\frac{1}{2}\frac{\tilde{\Delta}}{\sqrt{\tilde{\Delta}^{2}+1}}\,_{2}F_{1}\left(\frac{1}{2},\frac{3}{2};2;\frac{1}{\tilde{\Delta}^{2}+1}\right), (28)
ζtχ0=W42F1(12,32;3;1Δ~2+1),\displaystyle\zeta t\chi_{0}=\frac{W}{4}\,_{2}F_{1}\left(\frac{1}{2},\frac{3}{2};3;\frac{1}{\tilde{\Delta}^{2}+1}\right), (29)

where Δ~=2Δc/W\tilde{\Delta}=2\Delta_{c}/W and F12{}_{2}F_{1} is the standard hypergeometric function.

In practice, for a given ff, we construct qk0q_{k0} (Eq. 9), qk0a^q_{k0}^{\hat{a}} (Eq. 18), and qkq_{k} (Eq. 10 by tuning yy), and hence gtg_{t} (Eq. 24). Then together with χ0\chi_{0}, we obtain the total variational energy (Eq. 21) per site explicitly given by

E=gtEK0+U2k(kN2+VU)2qk.\displaystyle E=g_{t}E_{K}^{0}+\frac{U}{2}\sum_{k}\left(k-\frac{N}{2}+\frac{V}{U}\right)^{2}q_{k}. (30)

where EK0=Nζtχ0/2E_{K}^{0}=-N\zeta t\chi_{0}/2. Finally, the energy EE needs to be optimized with respect to (f,g)(f,g) or equivalently (Δc,g)(\Delta_{c},g).

IV.2 SU(10) case

Refer to caption
Figure 1: Results of the ionic SU(10) Hubbard model. (a) Band renormalization factor gtg_{t} versus Hubbard UU and ionic potential VV. The color encodes the value of gtg_{t}. The white curves enclose Mott lobes with gt=0g_{t}=0. (b) Fermion density nAn_{A} on the A-sublattice (solid, left scale) and gtg_{t} (dashed, right scale) versus V/UV/U for U/W=10U/W=10.

Let us take SU(10) as an example. In Fig. 1(a), we present the hopping renormalization gtg_{t} versus UU and VV. We find gtg_{t} is suppressed by UU and drops to zero for large enough UU above a critical value UcU_{c}. Interestingly, the regimes with gt=0g_{t}=0 form different Mott lobes, enclosed by the boundaries shown as white curves (to be calculated analytically in the next section).

For U/W=10U/W=10, we plot the fermion density nAn_{A} on the A-sublattice (solid line, left scale) as a function of V/UV/U in Fig.1(b). For comparison, gtg_{t} is also plotted (dashed line, right scale). As the ionic potential VV continuously varies, nAn_{A} shows a staircase behavior. Within each plateau, nA=mn_{A}=m is quantized to the nearest integer of n0=N2VUn_{0}=\frac{N}{2}-\frac{V}{U} and gt=0g_{t}=0, where charge fluctuations are completely frozen. Between neighboring plateaus (Mott lobes), nAn_{A} changes continuously between two neighboring integers and meanwhile gtg_{t} is nonzero. In this region, the system is a band insulator with staggered charge density wave as long as mN/2m\neq N/2, in which there is no gapless excitation but the charge density nAn_{A} as a property of the ground state can be continuously tuned by the staggered ionic potential VV. In contrast, the uniform part of the charge density (averaged over both sublattices) does not change with VV, as indicated above.

Refer to caption
Figure 2: Band renormalization factor gtg_{t}, average of the kk-tuple projection operator qkq_{k} and fermion number nAn_{A} versus UU for V/U=1.8V/U=1.8, corresponding to m=3m=3 and δ=0.2\delta=0.2, in the case of SU(1010). For clarity, only qmq_{m} and qm±1q_{m\pm 1} are plotted.

In Fig. 2, we show the UU-dependence of gtg_{t}, nAn_{A}, qmq_{m} and qm±1q_{m\pm 1} with a fixed V/U=1.8V/U=1.8 corresponding to m=3m=3. It is seen that gtg_{t} drops continuously with UU from 11 at U=0U=0 to 0 at UcU_{c} and maintains at zero for U>UcU>U_{c}. The average fermion number nAn_{A} is found to vary continuously with U<UcU<U_{c} but quantized to mm when UUcU\geq U_{c}. The most direct way to see the local moment formation is through qmq_{m} which increases with UU from the free limit value (given by Eq. 9) at U=0U=0 to 11 when UUcU\geq U_{c}. Meanwhile, qkmq_{k\neq m} drops to zero at UcU_{c} (for clarity, only qm±1q_{m\pm 1} are plotted), which of course is a natural consequence of the normalization condition kqk=1\sum_{k}q_{k}=1. Therefore, different mm-tuple moments are well established inside these Mott insulating phases.

The phase outside of the Mott lobes are characterized by nonzero gtg_{t}, which in fact is a band insulator (except V=0V=0) caused by the ionic potential in our bipartite lattice, although there is a renormalizaiton of the quasiparticle excitations. Therefore the phase transitions here from gt0g_{t}\neq 0 to gt=0g_{t}=0 are not the usual metal-insulator transitions but from the band insulator to the Mott insulator. It is an interesting question to ask whether the band gap closes to generate a metallic phase at or near the phase transition [72, 73, 76]. Our answer to this question is actually bilateral: the effective excitation gap for the quasiparticles (under projection) is given by gtΔcg_{t}\Delta_{c}, which vanishes as the Mott limit is approached, but at the same time the quasiparticle weight also vanishes.

IV.3 Mott transitions

We now try to obtain the critical UcU_{c} analytically for the Mott transitions. Near the Mott lobe labeled by mm, we have found qmq_{m} approaches 11 and all other qkmq_{k\neq m} are small and linearly drop to zero at UcU_{c} as seen from Fig. 2. Therefore, it is reasonable to assume

qk=(1ϵϵ+)δkm+ϵδk,m1+ϵ+δk,m+1,\displaystyle q_{k}=(1-\epsilon_{-}-\epsilon_{+})\delta_{km}+\epsilon_{-}\delta_{k,m-1}+\epsilon_{+}\delta_{k,m+1}, (31)

which satisfies the normalization condition kqk=1\sum_{k}q_{k}=1 and gives the fermion density Nf=kkqk=m+(ϵ+ϵ)Nf=\sum_{k}kq_{k}=m+(\epsilon_{+}-\epsilon_{-}). The hopping renormalization Eq. 24 in this approximation is given by

gt=1ϵϵ+qm0(qm1,0a^ϵqm1,0+qm0a^ϵ+qm+1,0)2,\displaystyle g_{t}=\frac{1-\epsilon_{-}-\epsilon_{+}}{q_{m0}}\left(q_{m-1,0}^{\hat{a}}\sqrt{\frac{\epsilon_{-}}{q_{m-1,0}}}+q_{m0}^{\hat{a}}\sqrt{\frac{\epsilon_{+}}{q_{m+1,0}}}\right)^{2}, (32)

such that the total energy per site in the projected state becomes

E=gt|EK0|+U2(ϵ+ϵ+)+U4δ(ϵ+ϵ),\displaystyle E=-g_{t}|E_{K}^{0}|+\frac{U}{2}(\epsilon_{-}+\epsilon_{+})+\frac{U}{4}\delta(\epsilon_{+}-\epsilon_{-}), (33)

where we defined

δ=n0m=N2VUm\displaystyle\delta=n_{0}-m=\frac{N}{2}-\frac{V}{U}-m (34)

as the charge frustration, or the deviation of N/2V/UN/2-V/U away from an integer mm. Requiring E/ϵ±=0\partial E/\partial\epsilon_{\pm}=0 in the limit of ϵ±0\epsilon_{\pm}\to 0, we obtain Uc=uc|EK0|U_{c}=u_{c}|E_{K}^{0}|, with a universal function ucu_{c} independent of the details in the kinetic part of the Hamiltonian,

uc=212δqm0a^2qm0qm+1,0+21+2δqm1,0a^2qm1,0qm0.\displaystyle u_{c}=\frac{2}{1-2\delta}\frac{q_{m0}^{\hat{a}2}}{q_{m0}q_{m+1,0}}+\frac{2}{1+2\delta}\frac{q_{m-1,0}^{\hat{a}2}}{q_{m-1,0}q_{m0}}. (35)

Using the expressions for qk,0q_{k,0} (Eq. 9) and qk,0a^q_{k,0}^{\hat{a}} (Eq. 18), it can be shown that UcU_{c} is automatically invariant under the particle-hole transformation mNmm\leftrightarrow N-m and f1ff\leftrightarrow 1-f. [We note that Eq. 35 can also be applied to the non-staggered case. The only exception is the value of χ0\chi_{0} (and hence EK0E_{K}^{0}), to be obtained in a uniform potential which in turn describes a metal.] From Eq. 35, ucu_{c} is found to depend strongly on the charge frustration δ\delta. As δ±1/2\delta\to\pm 1/2 (maximally charge frustrated), ucu_{c}\to\infty, as seen in Fig. 1(a). This means the Mott transition cannot be reached in this case. For δ=0\delta=0, instead, we obtain finite ucu_{c}, which we plot as a function of m/Nm/N in Fig. 3(a). In the case of N=2N=2, the Brinkman-Rice result uc=8u_{c}=8 is recovered [2]. For larger NN, ucu_{c} is reduced but always larger than 44. For a given NN, ucu_{c} increases quickly as mm approaches 11 or N1N-1 but is always smaller than 88.

Refer to caption
Figure 3: (a) The universal function uc=Uc/|Ek0|u_{c}=U_{c}/|E_{k}^{0}| for δ=0\delta=0 versus m/Nm/N for a series of NN up to 100100. (b) The kinetic energy per site per flavor EK/NE_{K}/N versus m/Nm/N.

To proceed, we also need |EK0||E_{K}^{0}| to obtain UcU_{c}. For the Bethe lattice, we plot the bare kinetic energy EK0E_{K}^{0} per site per flavor with respect to m/Nm/N in Fig. 3(b). Clearly, as m/Nm/N approaches 0 or 11, |EK0||E_{K}^{0}| drops to zero. Combining ucu_{c} and |EK0||E_{K}^{0}|, we obtain UcU_{c}. For the cases of N=2N=2, 66 and 1010, respectively, we plot the results of UcU_{c} versus V/UV/U in Fig. 4(a). The result of N=10N=10 is also plotted in Fig. 1(a) for comparison. Similar calculations can be performed on odd NN as shown in Fig. 4(b) for N=3N=3, 77 and 1111, respectively. Note that for a fixed NN (e.g., N=10N=10), UcU_{c} drops slightly as mm approaches 11. This is the combined effect of the corresponding increase of ucu_{c} [see Fig. 3(a)] and decrease of |EK0||E_{K}^{0}| [see Fig. 3(b)].

Refer to caption
Figure 4: (a) The critical value UcU_{c} versus V/UV/U for N=2N=2 (solid), 66 (dashed) and 1010 (dash-dotted), respectively. The Mott lobes are labeled by mm. (b) Uc/WU_{c}/W for δ=0\delta=0 versus NN for m=N/2m=N/2 (solid) and m=1m=1 (dashed).

For the NN-dependence of UcU_{c} with δ=0\delta=0, we show two representative results of m=N/2m=N/2 and m=1m=1 in the inset of Fig. 4(a). A perfect linear dependence UcU_{c} versus NN for large NN is seen for m=N/2m=N/2 (solid line) since ucu_{c} approaches a constant 44 from Fig. 3(a), and |EK0|N|E_{K}^{0}|\propto N from Fig. 3(b). We also find the scaling of UcNU_{c}\propto N for any fixed m/Nm/N (not shown). But for a fixed mm, e.g. m=1m=1 as shown by the dashed lines in the inset of Fig. 4(a), UcU_{c} does not linearly depend on NN any more. This is because m/Nm/N decreases toward zero as NN increases to infinity, and thus |EK0|/N|E_{K}^{0}|/N does not maintain a fixed value but drops to zero. Therefore, the linear scaling of Uc=uc|EK0|NU_{c}=u_{c}|E_{K}^{0}|\propto N breaks down to a sublinear behavior.

V spin description of the Mott insulating states

We have found the conditions for different Mott insulating states in which different types of local moments are formed and charge degrees of freedom are frozen. The low energy effective theory inside these Mott lobes should be described by these local moments, or equivalently the SU(NN) “spins”.

Given the ground state with mm fermions on the A-sublattice and (Nm)(N-m) fermions on the B-sublattice, we can perform a second order perturbation with respect to the kinetic Hamiltonian HtH_{t}, to obtain an effective Hamiltonian in the low energy sector,

H=4t2(1+δ2)Uijabca(i)cb(i)cb(j)ca(j),\displaystyle H=\frac{4t^{2}}{(1+\delta^{2})U}\sum_{\langle ij\rangle}\sum_{ab}c_{a}(i)^{\dagger}c_{b}(i)c_{b}^{\dagger}(j)c_{a}(j), (36)

subject to ni=aca(i)ca(i)=mn_{i}=\sum_{a}c_{a}(i)^{\dagger}c_{a}(i)=m on the A-sublattice and ni=Nmn_{i}=N-m on the B-sublattice. This restriction suggests to define spin operators Sab(i)S_{ab}(i) on site-ii as

Sab(i)=ca(i)cb(i)niNδab,\displaystyle S_{ab}(i)=c_{a}^{\dagger}(i)c_{b}(i)-\frac{n_{i}}{N}\delta_{ab}, (37)

such that the traceless condition TrS(i)=0{\rm Tr}~{}S(i)=0 is satisfied. Further, it can be checked that these SabS_{ab} satisfy the SU(NN) algebra:

[Sab,Scd]=δbcSadδadScb.\displaystyle[S_{ab},S_{cd}]=\delta_{bc}S_{ad}-\delta_{ad}S_{cb}. (38)

Using these spin operators, the above Hamiltonian can be rewritten as the SU(NN) Heisenberg model

H=JijabSab(i)Sba(j),\displaystyle H=J\sum_{\langle ij\rangle ab}S_{ab}(i)S_{ba}(j), (39)

where J=4t2(1+δ2)UJ=\frac{4t^{2}}{(1+\delta^{2})U}. Since UUcU\sim U_{c} here should be proportional to NN, the above Hamiltonian has a natural large-NN limit. The SU(NN) Heisenberg model has been widely studied in the literature, as a mathematical generalization of the SU(2) Heisenberg model [4]. In this work, we have shown its relation to the ionic SU(NN) Hubbard model.

The above SU(NN) Hubbard or Heisenberg model supports an antiferromagnetic ground state with the Neel order. To represent the Neel order, we may select a specific spin axis, e.g., one of the diagonal Cartan base,

Lm(i)=amaca(i)ca(i),\displaystyle L_{m}(i)=\sum_{a}\ell_{m}^{a}c_{a}^{\dagger}(i)c_{a}(i), (40)

with ma=1m\ell_{m}^{a}=\frac{1}{m} for ama\leq m and 1Nm-\frac{1}{N-m} for a>ma>m. The Neel order is then described by Lm(i)(1)i𝒩\langle L_{m}(i)\rangle\sim(-1)^{i}{\cal N}, with mm flavors of fermions on the A-sublattice and the remaining (Nm)(N-m) flavors on the B-sublattice.

Note that even if the local moments LmL_{m} are ordered, the state still enjoys an internal symmetry group, SU(m)×SU(Nm)\mathrm{SU}(m)\times\mathrm{SU}(N-m), which becomes of merely gauge degrees of freedom if the charge is fully quantized. The Goldstone modes above the Neel ordered state fall into the Grassmannian manifold Gr(N,m)=U(N)/[U(m)×U(Nm)]\mathrm{Gr}(N,m)=\mathrm{U}(N)/[\mathrm{U}(m)\times\mathrm{U}(N-m)] [31]. Such fluctuations exchange the flavor content of the local moments without affecting the charge, in analogy to the spin rotation in the SU(2) system.

VI Conclusion

In summary, we have developed a Gutzwiller approximation and RMFT for the SU(NN)-symmetric fermionic systems. Applying to the ionic Hubbard model, we find the conjugate local moments, with mm fermions on the A-sublattice and (Nm)(N-m) fermions on the B-sublattice, are well established when the Hubbard UU is above a critical value UcU_{c}. We obtained an analytical solution to UcU_{c} wich depends on the bare kinematics and a universal function of mm, NN and the charge frustration δ\delta. For large NN, UcU_{c} is found to depend on NN linearly for fixed m/Nm/N but sublinearly with fixed mm if NN is fixed. The local moment formation is accompanied by a peculiar band-insulator to Mott-insulator transition, where the ionic potential and quasiparticle weight are renormalized to zero simultaneously. Inside the Mott insulating phase, the system is effectively described by the SU(NN) Heisenberg model which is widely studied previously in the literature. Our results shed light on the realization of such models in cold atom systems.

Finally, several remarks on the Gutzwiller projection are in order. First, it can be improved by including additional Jastrow factors [71]. Second, in one dimension, the Gutzwiller projection is inaccurate or even fails, while a long-range Jastrow factor alone (without Gutzwiller projection) turns out to be able to capture the Mott insulating state correctly [77]. Third, even in infinite dimensions, the metal-Mott insulator transition is better described by the Gutzwiller projection followed by a partial Schrieffer-Wolff unitary transformation [78]. The latter two directions are intriguing and even challenge the notion of the Mott state defined by the absence of double occupancy, in the SU(22) case. It would be interesting to improve our study of the slightly more complicated ionic SU(NN) Hubbard model along similar lines. However, we believe our results for two and higher dimensional ionic models should provide a qualitatively correct picture regarding the multiple transitions from the band insulator to Mott insulator, as well as the order of magnitude of the critical interactions.

Acknowledgements.
This work is supported by the National Natural Science Foundation of China (under Grant No. 11874205, No. 12274205 and No. 11574134).

References

  • Anderson [1961] P. W. Anderson, Localized magnetic states in metals, Phys. Rev. 124, 41 (1961).
  • Brinkman and Rice [1970] W. F. Brinkman and T. M. Rice, Application of Gutzwiller’s Variational Method to the Metal-Insulator Transition, Phys. Rev. B 2, 4302 (1970).
  • Imada et al. [1998] M. Imada, A. Fujimori, and Y. Tokura, Metal-insulator transitions, Rev. Mod. Phys. 70, 1039 (1998).
  • Affleck [1985] I. Affleck, Large-nn limit of SU(nn) quantum spin chains, Phys. Rev. Lett. 54, 966 (1985).
  • Affleck and Marston [1988] I. Affleck and J. B. Marston, Large-nn limit of the Heisenberg-Hubbard model: Implications for high-TcT_{c} superconductors, Phys. Rev. B 37, 3774 (1988).
  • Marston and Affleck [1989] J. B. Marston and I. Affleck, Large-nn limit of the Hubbard-Heisenberg model, Phys. Rev. B 39, 11538 (1989).
  • Arovas and Auerbach [1988] D. P. Arovas and A. Auerbach, Functional integral theories of low-dimensional quantum Heisenberg models, Phys. Rev. B 38, 316 (1988).
  • Read and Sachdev [1989a] N. Read and S. Sachdev, Valence-bond and spin-Peierls ground states of low-dimensional quantum antiferromagnets, Phys. Rev. Lett. 62, 1694 (1989a).
  • Read and Sachdev [1989b] N. Read and S. Sachdev, Some features of the phase diagram of the square lattice SU(N) antiferromagnet, Nucl. Phys. B 316, 609 (1989b).
  • Read and Sachdev [1990] N. Read and S. Sachdev, Spin-Peierls, valence-bond solid, and Neel ground states of low-dimensional quantum antiferromagnets, Phys. Rev. B 42, 4568 (1990).
  • Harada et al. [2003] K. Harada, N. Kawashima, and M. Troyer, Néel and Spin-Peierls Ground States of Two-Dimensional SU(N) Quantum Antiferromagnets, Phys. Rev. Lett. 90, 117203 (2003).
  • Assaad [2005] F. Assaad, Phase Diagram of the Half-Filled Two-Dimensional SU(n) Hubbard-Heisenberg Model: A Quantum Monte Carlo Study, Phys. Rev. B 71, 075103 (2005).
  • Kawashima and Tanabe [2007] N. Kawashima and Y. Tanabe, Ground States of the SU(N) Heisenberg Model, Phys. Rev. Lett. 98, 057202 (2007).
  • Arovas [2008] D. Arovas, Simplex solid states of SU(N) quantum antiferromagnets, Phys. Rev. B 77, 104404 (2008).
  • Xu and Wu [2008] C. Xu and C. Wu, Resonating plaquette phases in su(4) heisenberg antiferromagnet, Phys. Rev. B 77, 134449 (2008).
  • Wu [2010] C. Wu, Exotic many-body physics with large-spin fermi gases, Physics 3, 92 (2010).
  • Beach et al. [2009] K. Beach, F. Alet, M. Mambrini, and S. Capponi, SU(N) heisenberg model on the square lattice: A continuous-N quantum monte carlo study, Phys. Rev. B 80, 184401 (2009).
  • Lou et al. [2009] J. Lou, A. W. Sandvik, and N. Kawashima, Antiferromagnetic to valence-bond-solid transitions in two-dimensional SU(N) Heisenberg models with multispin interactions, Phys. Rev. B 80, 180414 (2009).
  • Rachel et al. [2009] S. Rachel, R. Thomale, M. Fuehringer, P. Schmitteckert, and M. Greiter, Spinon confinement and the Haldane gap in SU(n) spin chains, Phys. Rev. B 80, 180420 (2009).
  • Kaul and Sandvik [2012] R. K. Kaul and A. W. Sandvik, Lattice Model for the SU(N) Néel to Valence-Bond Solid Quantum Phase Transition at Large N, Phys. Rev. Lett. 108, 137201 (2012).
  • Corboz et al. [2012] P. Corboz, K. Penc, F. Mila, and A. M. Läuchli, Simplex solids in SU(N{N}) Heisenberg models on the kagome and checkerboard lattices, Phys. Rev. B 86, 041106 (2012).
  • Harada et al. [2013] K. Harada, T. Suzuki, T. Okubo, H. Matsuo, J. Lou, H. Watanabe, S. Todo, and N. Kawashima, Possibility of Deconfined Criticality in SU(N) Heisenberg Models at Small N, Phys. Rev. B 88, 220408 (2013).
  • Nataf and Mila [2014] P. Nataf and F. Mila, Exact Diagonalization of Heisenberg SU(N) Models, Phys. Rev. Lett. 113, 127204 (2014).
  • Dufour et al. [2015] J. Dufour, P. Nataf, and F. Mila, Variational Monte Carlo investigation of SU(N) Heisenberg chains, Phys. Rev. B 91, 174427 (2015).
  • Okubo et al. [2015] T. Okubo, K. Harada, J. Lou, and N. Kawashima, SU(N) Heisenberg model with multi-column representations, Phys. Rev. B 92, 134404 (2015).
  • Suzuki et al. [2015] T. Suzuki, K. Harada, H. Matsuo, S. Todo, and N. Kawashima, Thermal phase transition of generalized Heisenberg models for SU(N) spins on square and honeycomb lattices, Phys. Rev. B 91, 094414 (2015).
  • Nataf and Mila [2016] P. Nataf and F. Mila, Exact diagonalization of Heisenberg SU(N) chains in the fully symmetric and antisymmetric representations, Phys. Rev. B 93, 155134 (2016).
  • D’Emidio and Kaul [2016] J. D’Emidio and R. K. Kaul, First-order superfluid to valence bond solid phase transitions in easy-plane SU(N) magnets for small-n, Phys. Rev. B 93, 054406 (2016).
  • Nataf and Mila [2018] P. Nataf and F. Mila, Density matrix renormalization group simulations of SU(N) Heisenberg chains using standard Young tableaus: Fundamental representation and comparison with a finite-size Bethe ansatz, Phys. Rev. B 97, 134420 (2018).
  • Kim et al. [2019] F. H. Kim, F. F. Assaad, K. Penc, and F. Mila, Dimensional crossover in the SU(4) Heisenberg model in the six-dimensional antisymmetric self-conjugate representation revealed by quantum Monte Carlo and linear flavor-wave theory, Phys. Rev. B 100, 085103 (2019).
  • MacFarlane [1979] A. J. MacFarlane, Generalizations of σ\sigma-Models and CpN Models, and Instantons, Phys. Lett. B 82, 239 (1979).
  • Hikami [1980] S. Hikami, Non-Linear σ\sigma Model of Grassmann Manifold and Non-Abelian Gauge Field with Scalar Coupling, Prog. Theor. Phys. 64, 1425 (1980).
  • Duerksen [1981] G. Duerksen, Dynamical Symmetry Breaking in Supersymmetric U(N+m)/U(n)U(m) Chiral Models, Phys. Rev. D 24, 926 (1981).
  • Maharana [1983] J. Maharana, The canonical structure of generalized non-linear sigma models in constrained hamiltonian formalism, Ann. IHP Phys. Théorique 39, 193 (1983).
  • Halperin et al. [1974] B. I. Halperin, T. C. Lubensky, and S.-K. Ma, First-order phase transitions in superconductors and smectic-A liquid crystals, Phys. Rev. Lett. 32, 292 (1974).
  • Auerbach [1994] A. Auerbach, Interacting electrons and quantum magnetism (1994).
  • Lu [1994] J. P. Lu, Metal-insulator transitions in degenerate Hubbard models and AxC60A_{x}C_{60}Phys. Rev. B 49, 5687 (1994).
  • Wu et al. [2003] C. Wu, J.-P. Hu, and S.-C. Zhang, Exact SO(5) Symmetry in the Spin-3/2 Fermionic System, Phys. Rev. Lett. 91, 186402 (2003).
  • Honerkamp and Hofstetter [2004] C. Honerkamp and W. Hofstetter, Ultracold fermions and the su(n) hubbard model, Phys. Rev. Lett. 92, 170403 (2004).
  • Buchta et al. [2007] K. Buchta, o. Legeza, E. Szirmai, and J. Solyom, Mott transition and dimerization in the one-dimensional SU(n) Hubbard model, Phys. Rev. B 75, 155108 (2007).
  • Hung et al. [2011] H.-H. Hung, Y. Wang, and C. Wu, Quantum magnetism in ultracold alkali and alkaline-earth fermion systems with symplectic symmetry, Phys. Rev. B 84, 054406 (2011).
  • Cai et al. [2013a] Z. Cai, H.-H. Hung, L. Wang, D. Zheng, and C. Wu, Pomeranchuk Cooling of SU(2N) Ultracold Fermions in Optical Lattices, Phys. Rev. Lett. 110, 220401 (2013a).
  • Cai et al. [2013b] Z. Cai, H.-H. Hung, L. Wang, and C. Wu, Quantum magnetic properties of the SU(2N){{SU}}({{2N}}) Hubbard model in the square lattice: A quantum Monte Carlo study, Phys. Rev. B 88, 125108 (2013b).
  • Lang et al. [2013] T. C. Lang, Z. Y. Meng, A. Muramatsu, S. Wessel, and F. F. Assaad, Dimerized Solids and Resonating Plaquette Order in SU(N)-Dirac Fermions, Phys. Rev. Lett. 111, 066401 (2013).
  • Wang et al. [2014] D. Wang, Y. Li, Z. Cai, Z. Zhou, Y. Wang, and C. Wu, Competing Orders in the 2D Half-Filled SU(2N) Hubbard Model through the Pinning-Field Quantum Monte Carlo Simulations, Phys. Rev. Lett. 112, 156403 (2014).
  • Zhou et al. [2014] Z. Zhou, Z. Cai, C. Wu, and Y. Wang, Quantum monte carlo simulations of thermodynamic properties of SU(2N) ultracold fermions in optical lattices, Phys. Rev. B 90, 235139 (2014).
  • Zhou et al. [2016] Z. Zhou, D. Wang, Z. Y. Meng, Y. Wang, and C. Wu, Mott insulating states and quantum phase transitions of correlated SU(2N) Dirac fermions, Phys. Rev. B 93, 245157 (2016).
  • Zhou et al. [2017] Z. Zhou, D. Wang, C. Wu, and Y. Wang, Finite-temperature valence-bond-solid transitions and thermodynamic properties of interacting SU(2N) Dirac fermions, Phys. Rev. B 95, 085128 (2017).
  • Zhou et al. [2018] Z. Zhou, C. Wu, and Y. Wang, Mott transition in the π\pi-flux SU(4)SU(4) Hubbard model on a square lattice, Phys. Rev. B 97, 195122 (2018).
  • Wang et al. [2019a] D. Wang, L. Wang, and C. Wu, Slater and Mott insulating states in the SU(6) Hubbard model, Phys. Rev. B 100, 115155 (2019a).
  • Bloch and Zwerger [2008] I. Bloch and W. Zwerger, Many-body physics with ultracold gases, Rev. Mod. Phys. 80, 885 (2008).
  • Bloch et al. [2012] I. Bloch, J. Dalibard, and S. Nascimbène, Quantum simulations with ultracold quantum gases, Nat. Phys. 8, 267 (2012).
  • DeSalvo et al. [2010] B. J. DeSalvo, M. Yan, P. G. Mickelson, Y. N. Martinez de Escobar, and T. C. Killian, Degenerate Fermi Gas of Sr87{}^{87}\mathrm{Sr}Phys. Rev. Lett. 105, 030402 (2010).
  • Taie et al. [2010] S. Taie, Y. Takasu, S. Sugawa, R. Yamazaki, T. Tsujimoto, R. Murakami, and Y. Takahashi, Realization of a SU(2)×\timesSU(6) System of Fermions in a Cold Atomic Gas, Phys. Rev. Lett. 105, 190401 (2010).
  • Krauser et al. [2012] J. S. Krauser, J. Heinze, N. Flaschner, S. Gotze, O. Jurgensen, D.-S. Luhmann, C. Becker, and K. Sengstock, Coherent multi-flavour spin dynamics in a fermionic quantum gas, Nature Phys. 8, 813 (2012).
  • Taie et al. [2012] S. Taie, R. Yamazaki, S. Sugawa, and Y. Takahashi, An SU(6) Mott insulator of an atomic Fermi gas realized by large-spin Pomeranchuk cooling, Nature Phys. 8, 825 (2012).
  • Zhang et al. [2014] X. Zhang, M. Bishof, S. L. Bromley, C. V. Kraus, M. S. Safronova, P. Zoller, A. M. Rey, and J. Ye, Spectroscopic observation of SU(N)-symmetric interactions in Sr orbital magnetism, Science 345, 1467 (2014).
  • Cazalilla and Rey [2014] M. A. Cazalilla and A. M. Rey, Ultracold Fermi gases with emergent SU(N) symmetry, Rep. Prog. Phys. 77, 124401 (2014).
  • Hart et al. [2015] R. A. Hart, P. M. Duarte, T.-L. Yang, X. Liu, T. Paiva, E. Khatami, R. T. Scalettar, N. Trivedi, D. A. Huse, and R. G. Hulet, Observation of antiferromagnetic correlations in the Hubbard model with ultracold atoms, Nature 519, 211 (2015).
  • Laflamme et al. [2016] C. Laflamme, W. Evans, M. Dalmonte, U. Gerber, H. Mejía-Díaz, W. Bietenholz, U. J. Wiese, and P. Zoller, CPN1\mathrm{CP}^{N-1} quantum field theories with alkaline-earth atoms in optical lattices, Ann. Phys. 370, 117 (2016).
  • Hofrichter et al. [2016] C. Hofrichter, L. Riegger, F. Scazza, M. Höfer, D. R. Fernandes, I. Bloch, and S. Fölling, Direct Probing of the Mott Crossover in the SU(N)\mathrm{SU}(N) Fermi-Hubbard Model, Phys. Rev. X 6, 021030 (2016).
  • Ozawa et al. [2018] H. Ozawa, S. Taie, Y. Takasu, and Y. Takahashi, Antiferromagnetic Spin Correlation of SU(N) Fermi Gas in an Optical Superlattice, Phys. Rev. Lett. 121, 225303 (2018).
  • Sonderhouse et al. [2020] L. Sonderhouse, C. Sanner, R. B. Hutson, A. Goban, T. Bilitewski, L. Yan, W. R. Milner, A. M. Rey, and J. Ye, Thermodynamics of a deeply degenerate SU(N)-symmetric Fermi gas, Nat. Phys. 16, 1216 (2020).
  • Taie et al. [2022] S. Taie, E. Ibarra-Garcia-Padilla, N. Nishizawa, Y. Takasu, Y. Kuno, H.-T. Wei, R. T. Scalettar, K. R. A. Hazzard, and Y. Takahashi, Observation of antiferromagnetic correlations in an ultracold SU(N) Hubbard model, Nature Physics 18, 1356 (2022).
  • Wang et al. [2019b] S.-Y. Wang, D. Wang, and Q.-H. Wang, Critical exponents of the nonlinear sigma model on a Grassmann manifold U(N)/U(m)U(Nm)U(N)/U(m)U(N-m) by the 1/N1/{N}-expansion, Phys. Rev. B 99, 165142 (2019b).
  • Gutzwiller [1963] M. C. Gutzwiller, Effect of Correlation on the Ferromagnetism of Transition Metals, Phys. Rev. Lett. 10, 159 (1963).
  • Gutzwiller [1965] M. C. Gutzwiller, Correlation of Electrons in a Narrow ss Band, Phys. Rev. 137, A1726 (1965).
  • Vollhardt [1984] D. Vollhardt, Normal He3: an almost localized Fermi liquid, Rev. Mod. Phys. 56, 99 (1984).
  • Bunemann et al. [1998] J. Bunemann, W. Weber, and F. Gebhard, Multiband Gutzwiller wave functions for general on-site interactions, Phys. Rev. B 57, 6896 (1998).
  • Wang et al. [2006] Q.-H. Wang, Z. Wang, Y. Chen, and F. Zhang, Unrestricted renormalized mean field theory of strongly correlated electron systems, Phys. Rev. B 73, 092507 (2006).
  • Edegger et al. [2007] B. Edegger, V. N. Muthukumar, and C. Gros, Gutzwiller–RVB theory of high-temperature superconductivity: Results from renormalized mean-field theory and variational Monte Carlo calculations, Adv. Phys. 56, 927 (2007).
  • Garg et al. [2006] A. Garg, H. R. Krishnamurthy, and M. Randeria, Can Correlations Drive a Band Insulator Metallic?, Phys. Rev. Lett. 97, 046403 (2006).
  • Kancharla and Dagotto [2007] S. S. Kancharla and E. Dagotto, Correlated Insulated Phase Suggests Bond Order between Band and Mott Insulators in Two Dimensions, Phys. Rev. Lett. 98, 016402 (2007).
  • Hohenberg and Kohn [1964] P. Hohenberg and W. Kohn, Inhomogeneous electron gas, Phys. Rev. 136, B864 (1964).
  • Gebhard [1990] F. Gebhard, Gutzwiller correlated wave functions in finite dimensions d: A systematic expansion in 1/d, Phys. Rev. B 41, 9452 (1990).
  • Chattopadhyay et al. [2019] A. Chattopadhyay, S. Bag, H. R. Krishnamurthy, and A. Garg, Phase diagram of the half-filled ionic hubbard model in the limit of strong correlations, Phys. Rev. B 99, 155127 (2019).
  • Capello et al. [2005] M. Capello, F. Becca, M. Fabrizio, S. Sorella, and E. Tosatti, Variational description of mott insulators, Phys. Rev. Lett. 94, 026406 (2005).
  • Wysokiński and Fabrizio [2017] M. M. Wysokiński and M. Fabrizio, Mott physics beyond the brinkman-rice scenario, Phys. Rev. B 95, 161106 (2017).