Transition from band insulator to Mott insulator and formation of local moment in half-filled ionic SU() Hubbard model
Abstract
We investigate the local moment formation in the half-filled SU() Hubbard model under a staggered ionic potential. As the Hubbard increases, the charge fluctuations are suppressed and eventually frozen when is above a critical value , marking the development of well-defined local moment with integer fermions on the A-sublattice and fermions on the B-sublattice, respectively. We obtain an analytical solution for for the paramagnetic ground state within the variational Gutzwiller approximation and renormalized mean field theory. For large , is found to depend on linearly with fixed , but sublinearly with fixed . 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() 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 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() case [4] with the spin operators satisfying the SU() algebra. The SU() 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() representations for the local spins, a conjugate representation with fermions on A-sublattice and fermions on B-sublattice [4] is mostly studied. It may support a generalized Neel order: for instance, the first and remaining flavors are occupied on the two sublattices, respectively. Hence, the SU() symmetry is broken into . The gapless fluctuations above this Neel order fall into the Grassmannian manifold [31, 32, 33, 34], which is reduced to the -component Ginzburg-Landau theory [35] or equivalently the famous CPN-1 model in the special case of .
One early motivation for doing the SU() generalization of Heisenberg model is to perform expansion around the saddle point at , providing a controllable way to reach the SU(2) model [36]. However, physically speaking, the SU() spin models should derive from the SU() Hubbard model in the limit that charge fluctuations are completely frozen. The SU() 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- model to life, but not just a gedanken model, and opens up a new field in the study of the finite but large 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() Hubbard model [65]. In this work, we will examine the specific conditions for the Hubbard and ionic potential under which the local moments can exist, with immediate relevance to experimental realization. We develop and apply an SU()-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() Hubbard model, we find the local moments, with quantized integer fermions on the A-sublattice and fermions on the B-sublattice, are well established when is above a critical value , which depends on , , and the ionic potential . For large , is found to depend on linearly with fixed , but sublinearly with fixed . 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() 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() Hubbard model in an ionic potential
Th ionic SU() Hubbard model we consider is described by the Hamiltonian , with
(1) |
and
(2) |
where labels the lattice site, denotes a nearest-neighbor bond, labels the flavor of the fermions, is the local density operator, denotes the Hubbard interaction, and is the staggered ionic potential. The model is clearly SU()-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 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 -term can be rewritten as
(3) |
where may be understood as the local ground charge tunable continuously by the staggered gate voltage . We ask whether can be quantized to integers () on the A-sublattice and on the B-sublattice, i.e., forming local moments, called -tuple moments, for large enough . The concept of local moment is a natural generalization of the SU(2) case for which only one kind of local moment with is possible. Here, however, the states with different -tuple moments should belong to different Mott insulating states. On the other hand, in the limit of , a nonzero always yields a band insulator. For large enough , the system is expected to enter different Mott insulating states labeled by . Whether these -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()-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()-symmetric case in which all the -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,
(4) |
where is the ground state of a free variational Hamiltonian to be specified, and is the Gutzwiller projection operator in the grand canonical ensemble
(5) |
where is the projection operator for the -tuple state (with -fermions),
(6) |
Clearly, in the absence of projection, we have . 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 -tuple is assumed to be
(7) |
where 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 to maintain the fermion density before and after projection in the grand canonical ensemble [70]. The fermion density on each site before projection is
(8) |
where is the average occupation number per flavor, and indicates the average performed with respect to . We have defined as the average of in the unprojected state,
(9) |
where is the combinatorial factor. After projection, we still require
(10) |
with
(11) |
where denotes average with respect to . 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
(12) |
Note the projection operator is simplified to . After substituting in Eq. 5, we obtain
(13) |
where we have used as a property of the projection operator . The fugacity (or in practice) is then tuned to satisfy the density restriction Eq. 10.
After obtaining all of and , 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 share the same local basis states as eigenstates,
(14) |
The kinetic energy is slightly more difficult to evaluate. Since the fermion hopping involves two sites, we need to keep two projectors, say, and in the hopping on the bond,
(15) |
Since the fermion operator is self-projective, we need to remove over projections before taking the quantum average. For a given site , we observe that
(16) | |||||
where we have defined a partial projection operator for fermions in the local Fock space excluding flavor ,
(17) |
Its average in is evaluated to be
(18) |
for any flavor in our SU()-symmetric case. Inserting the above relations in Eq. 15, we obtain
(19) |
where is the renormalization of the hopping by the projection, and
(20) | |||||
In the second line we have used Eq.(13) to trade for .
Combining the potential and kinetic energies, we obtain the total variational energy in the projected state,
(21) |
where is the average of hopping operator in the unprojected state. This energy is understood as a functional of (i) the fermion density which in turn depends on the trial wavefunction , and (ii) the Gutzwiller projection parameter . The fugacity parameters 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 with respect to , i.e., , with fixed fermion density , we obtain
(22) |
where is introduced as the Lagrange multiplier enforcing normalization of the wave function, and is a free Hamiltonian yet encoded with the renormalization effect from the Gutzwiller projection,
(23) |
where the variational local chemical potential is introduced to enforce . It can be shown that the single-particle spectrum of is just the quasipaticle excitation spectrum beyond the correlated variational ground state, with the quasiparticle weight renormalized by [70].
IV Application to the ionic SU() Hubbard model
In this section, we apply the variational Gutzwiller approximation and RMFT developed in the previous section to the ionic SU() 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) and the Gutzwiller parameter on the A-sublattice. Correspondingly, we can replace , , , etc., to obtain the relevant quantities on the B-sublattice, while remains the same. Under these simplifications, and become bond-independent, denoted as and , respectively. In particular, is now given by
(24) |
Due to the presence of an ionic potential, we choose in to write
(25) | |||||
Under such a parametrization, is a global factor renormalizing the effective bandwidth and quasiparticle excitations. The unprojected ground state , subsequently the fermion density , and the average hopping , only depend on . From and after some algebra, we obtain
(26) | |||
(27) |
Here is the coordination number, and is the density of states. As an illustrative example, we consider the Bethe lattice, for which , with the bandwidth, giving rise to
(28) | |||
(29) |
where and is the standard hypergeometric function.
IV.2 SU(10) case

Let us take SU(10) as an example. In Fig. 1(a), we present the hopping renormalization versus and . We find is suppressed by and drops to zero for large enough above a critical value . Interestingly, the regimes with form different Mott lobes, enclosed by the boundaries shown as white curves (to be calculated analytically in the next section).
For , we plot the fermion density on the A-sublattice (solid line, left scale) as a function of in Fig.1(b). For comparison, is also plotted (dashed line, right scale). As the ionic potential continuously varies, shows a staircase behavior. Within each plateau, is quantized to the nearest integer of and , where charge fluctuations are completely frozen. Between neighboring plateaus (Mott lobes), changes continuously between two neighboring integers and meanwhile is nonzero. In this region, the system is a band insulator with staggered charge density wave as long as , in which there is no gapless excitation but the charge density as a property of the ground state can be continuously tuned by the staggered ionic potential . In contrast, the uniform part of the charge density (averaged over both sublattices) does not change with , as indicated above.

In Fig. 2, we show the -dependence of , , and with a fixed corresponding to . It is seen that drops continuously with from at to at and maintains at zero for . The average fermion number is found to vary continuously with but quantized to when . The most direct way to see the local moment formation is through which increases with from the free limit value (given by Eq. 9) at to when . Meanwhile, drops to zero at (for clarity, only are plotted), which of course is a natural consequence of the normalization condition . Therefore, different -tuple moments are well established inside these Mott insulating phases.
The phase outside of the Mott lobes are characterized by nonzero , which in fact is a band insulator (except ) caused by the ionic potential in our bipartite lattice, although there is a renormalizaiton of the quasiparticle excitations. Therefore the phase transitions here from to 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 , 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 analytically for the Mott transitions. Near the Mott lobe labeled by , we have found approaches and all other are small and linearly drop to zero at as seen from Fig. 2. Therefore, it is reasonable to assume
(31) |
which satisfies the normalization condition and gives the fermion density . The hopping renormalization Eq. 24 in this approximation is given by
(32) |
such that the total energy per site in the projected state becomes
(33) |
where we defined
(34) |
as the charge frustration, or the deviation of away from an integer . Requiring in the limit of , we obtain , with a universal function independent of the details in the kinetic part of the Hamiltonian,
(35) |
Using the expressions for (Eq. 9) and (Eq. 18), it can be shown that is automatically invariant under the particle-hole transformation and . [We note that Eq. 35 can also be applied to the non-staggered case. The only exception is the value of (and hence ), to be obtained in a uniform potential which in turn describes a metal.] From Eq. 35, is found to depend strongly on the charge frustration . As (maximally charge frustrated), , as seen in Fig. 1(a). This means the Mott transition cannot be reached in this case. For , instead, we obtain finite , which we plot as a function of in Fig. 3(a). In the case of , the Brinkman-Rice result is recovered [2]. For larger , is reduced but always larger than . For a given , increases quickly as approaches or but is always smaller than .

To proceed, we also need to obtain . For the Bethe lattice, we plot the bare kinetic energy per site per flavor with respect to in Fig. 3(b). Clearly, as approaches or , drops to zero. Combining and , we obtain . For the cases of , and , respectively, we plot the results of versus in Fig. 4(a). The result of is also plotted in Fig. 1(a) for comparison. Similar calculations can be performed on odd as shown in Fig. 4(b) for , and , respectively. Note that for a fixed (e.g., ), drops slightly as approaches . This is the combined effect of the corresponding increase of [see Fig. 3(a)] and decrease of [see Fig. 3(b)].

For the -dependence of with , we show two representative results of and in the inset of Fig. 4(a). A perfect linear dependence versus for large is seen for (solid line) since approaches a constant from Fig. 3(a), and from Fig. 3(b). We also find the scaling of for any fixed (not shown). But for a fixed , e.g. as shown by the dashed lines in the inset of Fig. 4(a), does not linearly depend on any more. This is because decreases toward zero as increases to infinity, and thus does not maintain a fixed value but drops to zero. Therefore, the linear scaling of 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() “spins”.
Given the ground state with fermions on the A-sublattice and fermions on the B-sublattice, we can perform a second order perturbation with respect to the kinetic Hamiltonian , to obtain an effective Hamiltonian in the low energy sector,
(36) |
subject to on the A-sublattice and on the B-sublattice. This restriction suggests to define spin operators on site- as
(37) |
such that the traceless condition is satisfied. Further, it can be checked that these satisfy the SU() algebra:
(38) |
Using these spin operators, the above Hamiltonian can be rewritten as the SU() Heisenberg model
(39) |
where . Since here should be proportional to , the above Hamiltonian has a natural large- limit. The SU() 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() Hubbard model.
The above SU() 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,
(40) |
with for and for . The Neel order is then described by , with flavors of fermions on the A-sublattice and the remaining flavors on the B-sublattice.
Note that even if the local moments are ordered, the state still enjoys an internal symmetry group, , 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 [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()-symmetric fermionic systems. Applying to the ionic Hubbard model, we find the conjugate local moments, with fermions on the A-sublattice and fermions on the B-sublattice, are well established when the Hubbard is above a critical value . We obtained an analytical solution to wich depends on the bare kinematics and a universal function of , and the charge frustration . For large , is found to depend on linearly for fixed but sublinearly with fixed if 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() 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() case. It would be interesting to improve our study of the slightly more complicated ionic SU() 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- limit of SU() quantum spin chains, Phys. Rev. Lett. 54, 966 (1985).
- Affleck and Marston [1988] I. Affleck and J. B. Marston, Large- limit of the Heisenberg-Hubbard model: Implications for high- superconductors, Phys. Rev. B 37, 3774 (1988).
- Marston and Affleck [1989] J. B. Marston and I. Affleck, Large- 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() 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 -Models and CpN Models, and Instantons, Phys. Lett. B 82, 239 (1979).
- Hikami [1980] S. Hikami, Non-Linear 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 , 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 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 -flux 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 , 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)SU(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, 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 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 by the -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 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).