Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2025 Aug 11;122(33):e2426111122. doi: 10.1073/pnas.2426111122

Quantum spin liquid from electron–phonon coupling

Xun Cai a, Zhaoyu Han b, Zi-Xiang Li a,1, Steven A Kivelson b,1, Hong Yao c,1
PMCID: PMC12377744  PMID: 40789028

Significance

The search for quantum spin liquids—an exotic phase of matter that features fractionalization and topological order—has primarily focused on frustrated antiferromagnets. In this work, we reveal an alternative route to its physical realization in systems with strong coupling of electrons to relatively high-frequency phonons. We obtained unambiguous numerical evidence of this phase from large-scale sign-free quantum Monte Carlo simulations of a simple and physical model on a triangular lattice.

Keywords: quantum spin liquids, electron–phonon coupling, quantum Monte Carlo

Abstract

A quantum spin liquid (QSL) is an exotic insulating phase with emergent gauge fields and fractionalized excitations. However, the unambiguous demonstration of the existence of a QSL in a “nonengineered” microscopic model (or in any material) remains challenging. Here, using numerically exact sign-problem-free quantum Monte Carlo simulations, we show that a QSL arises in a nonengineered electron–phonon model. Specifically, we investigate the ground-state phase diagram of the bond Su–Schrieffer–Heeger model on a 2D triangular lattice at (one electron per site), which we show includes a QSL phase which is fully gapped, exhibits no symmetry-breaking order, and supports deconfined fractionalized holon excitations. This suggests promising routes for finding QSLs in realistic materials and high-Tc superconductivity by lightly doping them.


One of the central focuses of modern condensed matter physics has been the quest for a physical system with a quantum spin liquid (QSL) ground state (1), an exotic insulating state harboring fractionalized excitations, deconfined emergent gauge fields, and topological order. (For reviews, see, e.g., refs. 210.) Research on QSLs has been largely motivated by their intriguing properties such as fractionalized excitations, potential relevance to high-Tc superconductivity (SC) (1114), and their potential applications in topological quantum computation (15, 16). Despite tremendous experimental efforts in the past decades (see, e.g., refs. 1727), unambiguous experimental evidence establishing the existence of a QSL in any real material remains elusive. Consequently, analytically or numerically exact solutions of “natural” or “nonengineered” microscopic models that establish the existence of a QSL phase are of fundamental importance and can potentially provide useful guidance in searching for QSLs in realizable materials.

To date, most analytical and numerical studies have been devoted to searching for QSLs in frustrated quantum magnets where only local repulsion between electrons are considered (see e.g., refs. 2866). Since electron–phonon coupling (EPC) is also ubiquitous in materials, it is natural to ask (12) if EPC can be the primary microscopic mechanism of QSL formation. Nonetheless, exploration of QSLs induced by pure EPC has been rare partly because EPC generates attractions which tend to favor pairing between itinerant electrons instead of spin interactions between local magnetic moments. A recent development is the establishment of resonating-valence-bond (RVB) states from strong-coupling analyses of two special electron-phonon models defined on a (generalized) Lieb lattice (67, 68). Moreover, it was recently shown that a Su–Schrieffer–Heeger (SSH) type EPC on the bipartite square lattice can induce antiferromagnetic (AF) ordering as well as valence bond solid (VBS) phases (6973), which partly motivates us to ask whether this type of EPC can induce a QSL on a nonbipartite lattice.

Here, we show that a QSL can indeed be found in such a nonengineered EPC model on the triangular lattice in a broad, intermediate coupling parameter regime. Specifically, we investigated a prototypical microscopic model featuring an SSH-type coupling between optical phonons and electrons on a triangular lattice. This model is sign-problem-free (7479), so we have been able to study it using large-scale numerically exact determinant quantum Monte Carlo (QMC) simulations (8083). The ground-state phase diagram inferred from the state-of-the-art QMC is shown in Fig. 1: The ground state possesses s-wave SC long-range order at weak EPC or whenever the phonon frequency is sufficiently high. VBS order is dominant for intermediate to strong EPC whenever the phonon frequency is sufficiently small.

Fig. 1.

Fig. 1.

Zero-temperature phase diagram of the SSH electron-phonon model at half-filling on triangular lattice with varying EPC strength λ and phonon frequency ω0, obtained from state-of-the-art QMC simulations. Here, SC, SVBS, IVBS, and QSL denote s-wave superconductivity, staggered VBS, incommensurate VBS, and quantum spin liquid, respectively. The QSL to SC transition is shown to be continuous and consistent with XY* universality. The red and blue stars indicate the values of the parameters adopted for the calculations reported in Fig. 3 to establish that holons are deconfined/confined in the QSL and SVBS phases, respectively.

Most significantly, a gapped QSL phase emerges between the SC and VBS phases. Several features accessible to large-scale QMC simulations are used to establish this: 1) The QSL has a finite gap to all excitations. 2) There are no spontaneously broken symmetries—at least none with an order parameter magnitude large enough to be detectable. 3) There exists a deconfined (fractionalized) holon excitation that has charge e and no spin. (Modulo only the possibility that we could be fooled were the confinement scale larger than our system sizes, this constitutes smoking-gun evidence (30) that the indicated phase is a QSL.) 4) We observe anomalous power-law correlations at the apparently continuous quantum phase transition between the QSL and SC phases, which is consistent with the XY universality, expected (84) in this circumstance. We, moreover, provide a heuristic understanding of the emergence of QSL phases in this model employing a strong-coupling perspective. Our results unveil the emergence of a QSL in a nonengineered EPC model, an occurrence that, to the best of our knowledge, has not been previously reported.

The present results should serve as an inspiration to substantially broaden the search for QSL candidate materials—to include materials without any obvious magnetism but with strong EPC involving relatively high-frequency phonons, especially in cases in which the electronic bands are relatively flat. The proximity of the QSL to an SC phase is also suggestive that high-Tc SC might occur in proximity to such a QSL phase.

Model.

We consider the bond SSH model on a triangular lattice described by the Hamiltonian:

H^=ij[t+gX^ijB^ij+P^ij22M+K2X^ij2], [1]

where B^ij=σ(ciσcjσ+h.c.) is the bond-density operator in which ciσ creates an electron on site i with spin polarization σ=,, and X^ij and P^ij are, respectively, the displacement and momentum operators of an optical phonon residing on the bond between nearest-neighbor (NN) sites ij. Here, t is the bare nearest-neighbor (NN) hopping amplitude and g is the EPC strength. The relevant dimensionless measure of the EPC strength is λzg2KW, where z=6 is the number of nearest-neighbors and W=9t is the electron’s bare bandwidth on triangular lattice (85). For simplicity we have assumed Einstein phonons with bare frequency ω0=K/M. We set ħ=1 throughout this paper.

Because this model is free from the notorious fermion minus sign problem, we have been able to perform numerically exact QMC simulations to access the ground-state (zero-temperature) properties up to system sizes 18×18. To do this, we implemented projective determinant QMC simulations (86). We have focused on the model at half-filling by fixing the electron number rather than the chemical potential. The DQMC simulations on the model with strong electron–phonon coupling are typically computationally more demanding than purely electronic models due to the longer autocorrelation times. We have adopted various strategies to reduce the autocorrelation time; nonetheless heavy computational resources have still been required to achieve reliable results with high accuracy (details are in SI Appendix).

Ground-State Phase Diagram.

The ground-state phase diagram in the extreme adiabatic limit, ω0=0, is shown along the Bottom line of Fig. 1. In this limit, quantum fluctuations of the phonon fields vanish and the phonon configurations {Xij} are static, so ground-state properties can be accessed straightforwardly by finding the phonon configurations {Xij} that minimize the adiabatic energy. For small λ, because the Fermi surface is not perfectly nested, a symmetry-preserving metallic ground-state is stable for λ smaller than a finite critical value. For stronger coupling, incommensurate VBS (IVBS) long-range order is found for λc1<λ<λc2 (λc10.42 and λc21.62) with a λ-dependent ordering vector that is not far from the optimal Fermi surface nesting vector. Even in the presence of IVBS order, ungapped pockets of Fermi surface persist for λc1<λ<λFS (λFS0.96), while the fermionic spectrum is fully gapped for λFS<λ<λc2. In stronger coupling, λ>λc2, commensurate VBS order emerges with the staggered pattern shown in the Inset of Fig. 1. The transitions at λc1 and λFS are continuous while the transition at λc2 is first order, accompanied by a discontinuous jump in the ordering vector. (See SI Appendix for details.)

The phase diagram for small but nonzero ω0 is readily inferred by continuity from ω0=0, as we have verified for finite ω0 using QMC. Specifically, VBS persists to nonzero ω0. Moreover, with or without VBS order, the presence of a Fermi surface at ω0=0 for all λ<λFS implies the existence of a Cooper instability and hence the emergence of superconducting order at small but nonzero ω0 - an expectation that is consistent with our QMC results in the studied range ω0/t0.5. In the regime close to the onset of IVBS order, SC should coexist with IVBS. Generally, the SC phase expands and the VBS phases are pushed to higher λ with increasing ω0. Most significantly, in an intermediate regime of ω0 and λ, we find a fully gapped Z2 QSL in the ground state.

Broken Symmetries.

To determine where in the phase diagram the ground-state has various patterns of spontaneous symmetry-breaking, we have computed the structure factor S(q,L)=1L4ijeiq·(rirj)O^iO^j and the associated correlation-length ratio

Rc(L)1S(Q+δq,L)S(Q,L)1|δqξ|2, [2]

where O^j is one of a number of possible local order-parameter fields, Q is the momentum at which the structure factor peaks and |δq|=4π/3L is the minimum crystalline momentum on a lattice with system size L×L. As indicated in the equations, in a broken symmetry phase, S(Q,L)|Φ|2 and Rc(L)1 as L, where |Φ| is the magnitude of the order parameter, while with only short-range order, it follows that S(Q,L)0 and that Rc(L)(ξ/L)20 as L, where ξ is the correlation length. As a function of a control parameter, a symmetry-breaking transition can be identified, as in Fig. 2A, with a crossing point at which Rc(L) transitions from being a decreasing to an increasing function of L, i.e., a point at which Rc(L) is independent of L for large L. We also consider the Binder ratio Rb(L) to identify the phase transition to a state with spontaneous symmetry breaking, as shown in Fig. 2B. Details of correlation-length and Binder ratios are in SI Appendix.

Fig. 2.

Fig. 2.

QMC results on the triangular lattice with size L×L, where the EPC strength is fixed at λ=2.16 (corresponding to g=1.8). (A) The correlation-length ratio (defined in Eq. 2) for onsite s-wave superconducting order as a function of ω0. The crossing point for different system sizes indicates that the transition between the SC and non-SC (QSL) phases occurs at ω01.2. (B) The Binder ratio (defined in SI Appendix, Eq. S5) for SVBS order as a function of ω0. The transition between SVBS and the uniform (QSL) phase occurs at ω00.7. (C) Finite-size-scaling results of the single-particle and spin gaps, Δsingle and Δspin, in the QSL phase for ω0=1.0. Both Δsingle and Δspin extrapolate to finite values as L, confirming the fully gapped character of the QSL phase.

An example of this analysis is shown in Fig. 2, where for a fixed value of λ=2.16 (corresponding to g=1.8) we have computed the superconductivity correlation-length ratio RSCc (i.e. with O^i corresponding to an onsite s-wave pair-field creation operator cici and Q=0) and the staggered VBS Binder ratio RVBSb [i.e. with Oi corresponding to a valence bond density operator and Q=(π,0)] as a function of ω0 for L up to L=16. The SC correlation-length ratio RSCc(L) (shown in Fig. 2A) increases with system size when ω01.2, indicating that the ground state possesses long-ranged SC order, and decreases for ω0<1.2, implying only short-range SC correlations. At sufficiently low frequency ω0<0.7, the order in the ground state is a staggered VBS, as evidenced by the results of VBS Binder ratios (shown in Fig. 2B). Consequently, in the intermediate frequency regime 0.7<ω0<1.2, the ground state has neither SC nor VBS order. We have repeated the same analysis for λ=2.94 (i.e. g=2.1), and again found an intermediate range of ω0 with neither SC nor VBS ordering, as shown in Fig. 1. We further checked for and found no evidence indicating other possible spontaneously broken symmetries in the intermediate phase, including other forms of VBS ordering (other values of Q) such as columnar VBS and 12×12 plaquette VBS, and loop current ordering (details are included in SI Appendix). Thus, there apparently exists, in a range of intermediate λ and ω0, a symmetric phase without any spontaneous symmetry breaking, consistent with a QSL phase.

Spectral Features of the QSL.

We have also obtained information concerning the single-particle and neutral spin spectra by analyzing the corresponding imaginary-time correlators in the intermediate phase with no symmetry breaking order. In both cases, we find that the local two-time correlator falls exponentially with time, implying a gap in the spectrum. (Details of the fitting procedure are presented in SI Appendix.) The single-particle and spin gap inferred in this way for representative couplings, λ=2.16 and ω0=1.0, in the intermediate phase are shown for various system sizes, L, in Fig. 2C, from which it is apparent that they both extrapolate to a finite value in the limit L. The bond–bond correlations obtained from QMC are short-ranged with a short correlation length (less than one lattice constant); the singlet gap is smaller but also finite (see SI Appendix for details). Consequently, the symmetric intermediate phase is fully gapped. According to the celebrated Lieb-Schultz-Mattis-Oshikawa-Hastings theorem (8789), any symmetric phase with a finite excitation gap on a triangular lattice at half-filling of spin-12 electrons cannot be a trivial phase; it must be a gapped QSL with accompanying topological order.

Fractionalized Excitations.

To corroborate its existence, it is desirable to establish direct signatures of a QSL, for instance, the existence of fractionalized excitations such as deconfined holons. The energy cost of creating two spatially separated holons in a QSL approaches a finite constant even when they are far from each other. In contrast, in any topologically trivial phase such as a VBS, holons are confined in the sense that the energy cost of creating two far separated holons increases linearly with separation. (Since the spin gap is finite, the energy cost of two far-separated e charges in a confining phase ceases to grow linearly beyond a critical separation determined by the spin-gap. For larger separations, it is energetically favorable to create two accompanying charge 0 spin 1/2 “spinons,” so that the charge excitations become a pair of ordinary charge e spin 1/2 holes.)

To investigate holon deconfinement, we perform QMC simulations for the same model but with two electrons (with opposite Sz) removed from the half-filled system. Moreover, we localize the associated excitations by adding two “impurity” potentials to the Hamiltonian,

H^H^+α=1,2Vpn^rα [3]

which couple to the local charge (relative to half-filling) at two separate sites r1 and r2, respectively, as schematically shown as red points in the Inset of Fig. 3A. Simulations have been performed on lattices of size Lx×Ly, with Ly=8, and with the two impurity sites separated in the x direction by r12=|r1r2|=Lx/2. We focus particularly on the energy difference between doped systems with and without impurity potential V, and the two-point correlations Q^1Q^2 and S^1zS^2z, where Q^αiRα(n^i1) and z-component of spin, S^αziRαS^iz in each of two spatially separated regions Rα (Rα consisting of the site rα and its first and second neighbors, as illustrated by the circles in the Inset of Fig. 3A) for different values of Lx. The on-site impurity potential is fixed at Vp=30.

Fig. 3.

Fig. 3.

QMC results for two holes doped away from half filling in terms of quantities defined in Eq. 3 and the following paragraph. The doped holes are localized in the neighborhood of two “impurities,” α=1 & 2 as indicated by the red dots in the Inset of (A), on which there is an onsite potential Vp=30; they are separated along the x direction by r12=Lx/2. The EPC strength is fixed at λ=2.94 and two representative values of ω0 are treated: ω0=0.6 (in the SVBS) and ω0=1.8 (in the QSL). The interimpurity charge–charge (Q^1Q^2) and spin–spin (S^1zS^2z) correlations are shown as a function of 1/Lx in (A and B), respectively. The energy difference between the states with (Vp=30) and without (Vp=0) the impurity potential is shown in (C) as a function of r12.

The results of charge and spin correlations are shown in Fig. 3A and B, respectively. In the QSL regime, with increasing Lx the charge correlation grows to a value close to one, while the spin correlation vanishes, implying that the excitation trapped by each impurity potential is a holon with charge-e but spin-zero. The energy difference ΔE between the states of the doped holes with and without impurity potentials is shown in Fig. 3C; ΔE approximately saturates to a constant for r126, as expected for deconfined holons. For comparison, we performed similar calculations in the regime of VBS. The behaviors of charge and spin correlations in the VBS phase are qualitatively similar to those in the QSL regime, as presented in Fig. 3A and B, implying that the excitations trapped by impurity potentials in the VBS are also holons; however, the energy cost ΔE increases roughly linearly with r12, as expected in a (weakly) confining phase. The results are consistent with the expectation of deconfined holons only in the QSL phase.

It is in principle possible for the holons to be fermionic or bosonic—a feature determined by energetics rather than by the topology of the state (90). This could ideally be measured by determining the adiabatic phase under exchange, but this is not possible with our present approach (91). However, the fact that the QSL to SC phase transition (see below) is consistent with condensation of bosonic holons (30, 84) suggests that the holons are bosonic.

Quantum Phase Transitions.

Although the QSL phase is fully gapped, on approach to any continuous transition from this to another phase, one or more gaps are expected to vanish. (If the transition is, instead, weakly first order, the appropriate gap should get small but not quite vanish—a subtle distinction that we might not be able to distinguish without more extensive studies than we have so far carried out.) Given that the QSL to SC transition appears to be continuous, we would expect the spin and quasiparticle gaps to remain finite at and on both sides of the transition, but the charge gap should vanish at the transition and in the SC phase. As shown in SI Appendix, Fig. S7, we have verified that indeed the charge gap vanishes at two representative points, one on the QSL-SC phase boundary and one in the SC phase, while the single-particle and spin gaps are nonzero. At a continuous QSL to VBS transition, the spin, quasi-particle, and charge gaps should remain finite at and on both sides of the transition, but the gap to spin 0 neutral excitations—which is finite in both phases—should vanish at the transition. We have not explicitly tested this at present.

The quantum phase transition between the insulating state and the neighboring SC phase is expected to exhibit additional distinct features if the insulating phase supports fractionalized excitations. Specifically, the quantum phase transition from a QSL with Z2 topological order to a conventional (s-wave) SC can proceed through the condensation of holons/doublons (“chargeons”); the corresponding critical properties (84) are characterized by 3d XY* universality, reflecting the existence of fractionalized excitations and exhibiting power-law correlations at criticality with a very large anomalous dimension ηXY*1.49 (92). In contrast, the transition between a conventional insulator and an SC, or between a QSL and an exotic SC phase (with deconfined visons) is expected to exhibit conventional 3d XY criticality, i.e. with ηXY0.034. By numerical analysis of pair–pair correlations at the QSL-SC transition in the present model, as detailed in SI Appendix, we found a η1.45±0.08, consistent with an XY transition within the error bar. This further confirms that the symmetry-preserving insulating phase in the triangular SSH model is a Z2 QSL.

Strong Coupling Analysis.

We can get a feeling for why a QSL phase arises and what its character is, by approaching the problem from the theoretically tractable strong coupling limit, where tg2/K (i.e. λ1) and ω0g2/K.

In the adiabatic limit ω0=0, complete identification of the degenerate ground-states can be obtained straightforwardly from a combination of analytic analysis and numerical solution of large finite size systems; the electronic Hamiltonian is noninteracting, so all that is required is to minimize the adiabatic energy with respect to the phonon coordinates, {Xij}. Notably, to the first order in t, there is an extensive number of degenerate ground-states, which we will refer to as hard-core dimer states. They can be identified with fully packed configurations C of hard-core dimers, where each dimer labels a nearest-neighbor bond and where exactly one dimer touches each site of the lattice, as shown in Fig. 4. In the state |C, two electrons occupy the bonding orbital on each dimer-covered bond so that the expectation value of B^ij=2 and X^ij=2g/K; on the remaining bonds, B^ij=0 and X^ij=0.

Fig. 4.

Fig. 4.

A schematic representation of the RK wave function on the triangular lattice. Each dimer represents two electrons occupying the bonding orbital between two neighboring sites (see the dimer bond creation operator bij after Eq. 4).

Restricting attention to the dimer states, we note that they remain degenerate with each other to first order in both t and ω0.* To summarize, ignoring terms of higher than first order in t and ω0, there is a manifold of degenerate ground states spanned by a set of linearly independent dimer states, |C, in one-to-one correspondence with the set of hard-core dimer coverings of the lattice. The electron wavefunction of each state is

|Ce={ij}Cb^ij|vac, [4]

where b^ij12(ci+cj)(ci+cj)=12[(cicj+cjci)+(cici+cjcj)] is the valence bond creation operator and |vac is the empty state. The phonon wave functions are certain Gaussian states for the phonon modes with Xij=2g/K on occupied bonds and Xij=0 on other bonds (details are given in SI Appendix).

Working in a higher order in t and ω0, we could, in principle, obtain, in terms of the parameters of the original model, an effective model that operates in the hard-core dimer subspace. The form of the resulting effective hard-core quantum dimer Hamiltonian, H^dimer=V^+J^, is highly constrained, and indeed of precisely the same form as has been analyzed in various earlier studies in which the origin of the effective model is entirely different (30). There are two types of terms: interaction terms, V^, that are diagonal in the dimer basis, and kinetic terms, J^, that are off diagonal. To low order in the small parameters, both terms are short-range, so for convenience, we will explicitly consider only those involving pairs of nearest-neighbor dimers, i.e.

graphic file with name pnas.2426111122inline01.jpg [5]

where the thick lines represent dimers and the summation is over all possible four-sided plaquettes, and

graphic file with name pnas.2426111122inline02.jpg [6]

(Further range terms that are similar in structure occur as well, although they are still smaller than these most local terms in the strong-coupling limit.)

The interaction terms are relatively simple to compute. For instance, to second order in t and first order in ω0, the interaction can be shown explicitly to be V=t2K/g2212ω0 (details of derivations in SI Appendix). The kinetic term J necessarily vanishes for any t in the ω00 limit as they are tunneling processes—akin to small bipolaron hopping—involving rearrangements of the phonon coordinates. For 0<ω0t/λ, where the potential energy still dominates the kinetic energy, the dimer repulsion (V>0) favors a staggered VBS ground state, consistent with the results obtained from the large-scale QMC calculations for large values of λ and small ω0/t. However, even if we restrict attention to the strong coupling limit, the distinct dependence of J and V on t and ω0 implies that as a function of ω0/t one can access the situation in which JV, in which case it has been shown that (on a nonbipartite lattice) there exists a perturbatively stable, gapped QSL phase with Z2 topological order (33, 49). In particular, at the special RK point (30), the ground state can be expressed as an equal superposition of all possible dimer covering configurations:

|RK=N1/2C|C, [7]

where N is the number of hard-core dimer states. A schematic representation of |RK is shown in Fig. 4. This is a short-range RVB state with Z2 topological order (15, 3133), that we believe is a representative state of the QSL phase we demonstrated in the present QMC study.

Concluding Remarks.

Using state-of-the-art QMC, we have established the existence of a QSL phase in the ground state of SSH electron-phonon models on a triangular lattice on the basis of multiple, independent criteria. This is a distinctive case for a nonengineered model, at least for one dominated by EPC. As SSH-type phonons exist commonly in quantum materials, our study points out a promising direction for the search for QSLs in systems, including twisted moire systems (93), with such phonon couplings.

As the SSH electron-phonon model in Eq. 1 is free from sign problems even for finite doping away from half-filling, its physical properties can be accurately studied by large-scale QMC simulations over a broader range of conditions, which is currently one of our ongoing efforts. In the near future, we intend to further explore the nature of the SC state proximate to the QSL to gain further insight into the nature of possible SC emerging from a lightly doped QSL (9498), a problem that has been conjectured to have relevance to the cuprate superconductors (1114).

Supplementary Material

Appendix 01 (PDF)

Acknowledgments

We would like to thank Eduardo Fradkin, David Huse, Roger Melko, Subir Sachdev, T. Senthil, and Zhouquan Wan for helpful discussions; Z.H. and S.A.K. thank Kyung-Su Kim and John Sous for collaborations on related topics. This work was supported in part by the Ministry of Science & Technology of China under Grant No. 2021YFA1400100 (H.Y.), by National Natural Science Foundation of China under Grant Nos. 12347107 (X.C., Z.-X.L., and H.Y.) and 12334003 (H.Y.), by the New Cornerstone Science Foundation through the Xplorer Prize (H.Y.), and by NSF grant No. DMR-2310312 (Z.H. and S.A.K.).

Author contributions

X.C., Z.H., Z.-X.L., S.A.K., and H.Y. designed research; performed research; and wrote the paper.

Competing interests

The authors declare no competing interest.

Footnotes

Reviewers: D.A.H., Princeton University; P.A.L., Massachusetts Institute of Technology; and M.P.Z., University of California Berkeley.

*We have also verified the robustness of the nearly degenerate ground state manifold of the dimer states in the presence of various additional terms, e.g. onsite Hubbard repulsion and Holstein-type EPC.

From the estimates of V and J as a function of t and ω0 in the strong coupling limit, as shown in SI Appendix, one would conclude that JV only if ω0t2K/g2. Since the arguments we have presented so far rely not only on λ1, and ω0K/g21, this requires us to extrapolate the strong coupling results to intermediate values of λ and ω0K/g2 to access the region studied by QMC. For this reason, the strong coupling analysis of the QSL phase just presented must be viewed as heuristic rather than controlled for the original EPC problem.

Contributor Information

Zi-Xiang Li, Email: zixiangli@iphy.ac.cn.

Steven A. Kivelson, Email: kivelson@stanford.edu.

Hong Yao, Email: yaohong@tsinghua.edu.cn.

Data, Materials, and Software Availability

Some study data are available. The data files are huge. The relevant data are included in the manuscript. Upon request, within reason, we can provide more data to those who ask.

Supporting Information

References

  • 1.Anderson P., Resonating valence bonds: A new kind of insulator? Mater. Res. Bull. 8, 153 (1973). [Google Scholar]
  • 2.Lee P. A., An end to the drought of quantum spin liquids. Science 321, 1306–1307 (2008). [DOI] [PubMed] [Google Scholar]
  • 3.Balents L., Spin liquids in frustrated magnets. Nature 464, 199–208 (2010). [DOI] [PubMed] [Google Scholar]
  • 4.Savary L., Balents L., Quantum spin liquids: A review. Rep. Prog. Phys. 80, 016502 (2016). [DOI] [PubMed] [Google Scholar]
  • 5.Norman M. R., Colloquium: Herbertsmithite and the search for the quantum spin liquid. Rev. Mod. Phys. 88, 041002 (2016). [Google Scholar]
  • 6.Wen X. G., Colloquium: Zoo of quantum-topological phases of matter. Rev. Mod. Phys. 89, 041004 (2017). [Google Scholar]
  • 7.Zhou Y., Kanoda K., Ng T. K., Quantum spin liquid states. Rev. Mod. Phys. 89, 025003 (2017). [Google Scholar]
  • 8.Knolle J., Moessner R., A field guide to spin liquids. Ann. Rev. Condens. Matter Phys. 10, 451–472 (2019). [Google Scholar]
  • 9.Broholm C., et al. , Quantum spin liquids. Science 367, eaay0668 (2020). [DOI] [PubMed] [Google Scholar]
  • 10.Kivelson S. A., Sondhi S., 50 years of quantum spin liquids. Nat. Rev. Phys. 5, 368–369 (2023). [Google Scholar]
  • 11.Anderson P. W., The resonating valence bond state in La2CuO4 and superconductivity. Science 235, 1196–1198 (1987). [DOI] [PubMed] [Google Scholar]
  • 12.Kivelson S. A., Rokhsar D. S., Sethna J. P., Topology of the resonating valence-bond state: Solitons and high-Tc superconductivity. Phys. Rev. B 35, 8865–8868 (1987). [DOI] [PubMed] [Google Scholar]
  • 13.Lee P. A., Nagaosa N., Wen X. G., Doping a Mott insulator: Physics of high-temperature superconductivity. Rev. Mod. Phys. 78, 17–85 (2006). [Google Scholar]
  • 14.Anderson P. W., et al. , The physics behind high-temperature superconducting cuprates: The ‘plain vanilla’ version of RVB. J. Phys. Condens. Matter 16, R755 (2004). [Google Scholar]
  • 15.Kitaev A., Fault-tolerant quantum computation by anyons. Ann. Phys. 303, 2–30 (2003). [Google Scholar]
  • 16.Nayak C., Simon S. H., Stern A., Freedman M., Das Sarma S., Non-abelian anyons and topological quantum computation. Rev. Mod. Phys. 80, 1083–1159 (2008). [Google Scholar]
  • 17.Y. Okamoto, M. Nohara, H. Aruga-Katori, H. Takagi, Spin-liquid state in the s=1/2 hyperkagome antiferromagnet Na4Ir3O8. Phys. Rev. Lett. 99, 137207 (2007). [DOI] [PubMed]
  • 18.Yamashita M., et al. , Highly mobile gapless excitations in a two-dimensional candidate quantum spin liquid. Science 328, 1246–1248 (2010). [DOI] [PubMed] [Google Scholar]
  • 19.Pratt F. L., et al. , Magnetic and non-magnetic phases of a quantum spin liquid. Nature 471, 612–616 (2011). [DOI] [PubMed] [Google Scholar]
  • 20.Han T. H., et al. , Fractionalized excitations in the spin-liquid state of a Kagome-lattice antiferromagnet. Nature 492, 406–410 (2012). [DOI] [PubMed] [Google Scholar]
  • 21.Fu M., Imai T., Han T. H., Lee Y. S., Evidence for a gapped spin-liquid ground state in a Kagome Heisenberg antiferromagnet. Science 350, 655–658 (2015). [DOI] [PubMed] [Google Scholar]
  • 22.Shen Y., et al. , Evidence for a spinon fermi surface in a triangular-lattice quantum-spin-liquid candidate. Nature 540, 559–562 (2016). [DOI] [PubMed] [Google Scholar]
  • 23.Banerjee A., et al. , Proximate Kitaev quantum spin liquid behaviour in a honeycomb magnet. Nat. Mater. 15, 733–740 (2016). [DOI] [PubMed] [Google Scholar]
  • 24.Banerjee A., et al. , Neutron scattering in the proximate quantum spin liquid α-RuCl3. Science 356, 1055–1059 (2017). [DOI] [PubMed] [Google Scholar]
  • 25.Kasahara Y., et al. , Majorana quantization and half-integer thermal quantum hall effect in a Kitaev spin liquid. Nature 559, 227–231 (2018). [DOI] [PubMed] [Google Scholar]
  • 26.Miksch B., et al. , Gapped magnetic ground state in quantum spin liquid candidate κ-(BEDT-TTF)2Cu2(CN)3. Science 372, 276 (2021). [DOI] [PubMed] [Google Scholar]
  • 27.Scheie A., et al. , Proximate spin liquid and fractionalization in the triangular antiferromagnet KYbSe2. Nat. Phys. 20, 74–81 (2024). [Google Scholar]
  • 28.Kalmeyer V., Laughlin R. B., Equivalence of the resonating-valence-bond and fractional quantum hall states. Phys. Rev. Lett. 59, 2095–2098 (1987). [DOI] [PubMed] [Google Scholar]
  • 29.Baskaran G., Zou Z., Anderson P., The resonating valence bond state and high-Tc superconductivity - a mean field theory. Solid State Commun. 63, 973–976 (1987). [Google Scholar]
  • 30.Rokhsar D. S., Kivelson S. A., Superconductivity and the quantum hard-core dimer gas. Phys. Rev. Lett. 61, 2376–2379 (1988). [DOI] [PubMed] [Google Scholar]
  • 31.Read N., Sachdev S., Large-n expansion for frustrated quantum antiferromagnets. Phys. Rev. Lett. 66, 1773–1776 (1991). [DOI] [PubMed] [Google Scholar]
  • 32.Wen X. G., Mean-field theory of spin-liquid states with finite energy gap and topological orders. Phys. Rev. B 44, 2664–2672 (1991). [DOI] [PubMed] [Google Scholar]
  • 33.Moessner R., Sondhi S. L., Resonating valence bond phase in the triangular lattice quantum dimer model. Phys. Rev. Lett. 86, 1881–1884 (2001). [DOI] [PubMed] [Google Scholar]
  • 34.Balents L., Fisher M. P. A., Girvin S. M., Fractionalization in an easy-axis Kagome antiferromagnet. Phys. Rev. B 65, 224412 (2002). [Google Scholar]
  • 35.Wen X. G., Quantum orders in an exact soluble model. Phys. Rev. Lett. 90, 016803 (2003). [DOI] [PubMed] [Google Scholar]
  • 36.Hermele M., et al. , Stability of u(1) spin liquids in two dimensions. Phys. Rev. B 70, 214437 (2004). [Google Scholar]
  • 37.Wang F., Vishwanath A., Spin-liquid states on the triangular and Kagomé lattices: A projective-symmetry-group analysis of Schwinger boson states. Phys. Rev. B 74, 174423 (2006). [Google Scholar]
  • 38.Kitaev A., Anyons in an exactly solved model and beyond. Ann. Phys. 321, 2–111 (2006). [Google Scholar]
  • 39.Yao H., Kivelson S. A., Exact chiral spin liquid with non-abelian anyons. Phys. Rev. Lett. 99, 247203 (2007). [DOI] [PubMed] [Google Scholar]
  • 40.Ran Y., Hermele M., Lee P. A., Wen X. G., Projected-wave-function study of the spin-1/2 Heisenberg model on the Kagomé lattice. Phys. Rev. Lett. 98, 117205 (2007). [DOI] [PubMed] [Google Scholar]
  • 41.Jackeli G., Khaliullin G., Mott insulators in the strong spin-orbit coupling limit: From Heisenberg to a quantum compass and Kitaev models. Phys. Rev. Lett. 102, 017205 (2009). [DOI] [PubMed] [Google Scholar]
  • 42.Yao H., Lee D. H., Fermionic magnons, non-abelian spinons, and the spin quantum hall effect from an exactly solvable spin-1/2 Kitaev model with SU(2) symmetry. Phys. Rev. Lett. 107, 087205 (2011). [DOI] [PubMed] [Google Scholar]
  • 43.Isakov S. V., Hastings M. B., Melko R. G., Topological entanglement entropy of a Bose-Hubbard spin liquid. Nat. Phys. 7, 772–775 (2011). [Google Scholar]
  • 44.Yan S., Huse D. A., White S. R., Spin-liquid ground state of the s=1/2 Kagome Heisenberg antiferromagnet. Science 332, 1173–1176 (2011). [DOI] [PubMed] [Google Scholar]
  • 45.Corboz P., Lajkó M., Läuchli A. M., Penc K., Mila F., Spin-orbital quantum liquid on the honeycomb lattice. Phys. Rev. X 2, 041013 (2012). [Google Scholar]
  • 46.Depenbrock S., McCulloch I. P., Schollwöck U., Nature of the spin-liquid ground state of the s=1/2 Heisenberg model on the Kagome lattice. Phys. Rev. Lett. 109, 067201 (2012). [DOI] [PubMed] [Google Scholar]
  • 47.Jiang H. C., Yao H., Balents L., Spin liquid ground state of the spin-12 square J1-J2 Heisenberg model. Phys. Rev. B 86, 024424 (2012). [Google Scholar]
  • 48.Jiang H. C., Wang Z., Balents L., Identifying topological order by entanglement entropy. Nat. Phys. 8, 902–905 (2012). [Google Scholar]
  • 49.Yao H., Kivelson S. A., Exact spin liquid ground states of the quantum dimer model on the square and honeycomb lattices. Phys. Rev. Lett. 108, 247206 (2012). [DOI] [PubMed] [Google Scholar]
  • 50.Wang L., Poilblanc D., Gu Z. C., Wen X. G., Verstraete F., Constructing a gapless spin-liquid state for the spin-1/2 J1J2 heisenberg model on a square lattice. Phys. Rev. Lett. 111, 037202 (2013). [DOI] [PubMed] [Google Scholar]
  • 51.Mishmash R. V., Garrison J. R., Bieri S., Xu C., Theory of a competitive spin liquid state for weak Mott insulators on the triangular lattice. Phys. Rev. Lett. 111, 157203 (2013). [DOI] [PubMed] [Google Scholar]
  • 52.He Y. C., Sheng D. N., Chen Y., Chiral spin liquid in a frustrated anisotropic Kagome Heisenberg model. Phys. Rev. Lett. 112, 137202 (2014). [DOI] [PubMed] [Google Scholar]
  • 53.Rau J. G., Lee E. K. H., Kee H. Y., Generic spin model for the honeycomb iridates beyond the Kitaev limit. Phys. Rev. Lett. 112, 077204 (2014). [DOI] [PubMed] [Google Scholar]
  • 54.Gong S. S., Zhu W., Balents L., Sheng D. N., Global phase diagram of competing ordered and quantum spin-liquid phases on the Kagome lattice. Phys. Rev. B 91, 075112 (2015). [Google Scholar]
  • 55.Liao H. J., et al. , Gapless spin-liquid ground state in the s=1/2 Kagome antiferromagnet. Phys. Rev. Lett. 118, 137202 (2017). [DOI] [PubMed] [Google Scholar]
  • 56.He Y. C., Zaletel M. P., Oshikawa M., Pollmann F., Signatures of Dirac cones in a DMRG study of the Kagome Heisenberg model. Phys. Rev. X 7, 031020 (2017). [Google Scholar]
  • 57.Wang Y. C., Zhang X. F., Pollmann F., Cheng M., Meng Z. Y., Quantum spin liquid with even ising gauge field structure on Kagome lattice. Phys. Rev. Lett. 121, 057202 (2018). [DOI] [PubMed] [Google Scholar]
  • 58.Zhu W., Chen X., He Y. C., Witczak-Krempa W., Entanglement signatures of emergent Dirac fermions: Kagome spin liquid and quantum criticality. Sci. Adv. 4, eaat5535 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Song X. Y., Wang C., Vishwanath A., He Y. C., Unifying description of competing orders in two-dimensional quantum magnets. Nat. Commun. 10, 4254 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Hu S., Zhu W., Eggert S., He Y. C., Dirac spin liquid on the spin-1/2 triangular Heisenberg antiferromagnet. Phys. Rev. Lett. 123, 207203 (2019). [DOI] [PubMed] [Google Scholar]
  • 61.Szasz A., Motruk J., Zaletel M. P., Moore J. E., Chiral spin liquid phase of the triangular lattice Hubbard model: A density matrix renormalization group study. Phys. Rev. X 10, 021042 (2020). [Google Scholar]
  • 62.Chen B. B., et al. , Quantum spin liquid with emergent chiral order in the triangular-lattice Hubbard model. Phys. Rev. B 106, 094420 (2022). [Google Scholar]
  • 63.Jin H. K., Sun R. Y., Tu H. H., Zhou Y., Unveiling a critical stripy state in the triangular-lattice SU(4) spin-orbital model. Sci. Bull. 67, 918–923 (2022). [DOI] [PubMed] [Google Scholar]
  • 64.Liu W. Y., et al. , Emergence of gapless quantum spin liquid from deconfined quantum critical point. Phys. Rev. X 12, 031039 (2022). [Google Scholar]
  • 65.Jiang Y. F., Jiang H. C., Nature of quantum spin liquids of the s=12 Heisenberg antiferromagnet on the triangular lattice: A parallel DMRG study. Phys. Rev. B 107, L140411 (2023). [Google Scholar]
  • 66.Yu X. J., Shi S. H., Xu L., Li Z. X., Emergence of competing orders and possible quantum spin liquid in SU(n) fermions. Phys. Rev. Lett. 132, 036704 (2024). [DOI] [PubMed] [Google Scholar]
  • 67.Han Z., Kivelson S. A., Resonating valence bond states in an electron-phonon system. Phys. Rev. Lett. 130, 186404 (2023). [DOI] [PubMed] [Google Scholar]
  • 68.Han Z., Kivelson S. A., Emergent gauge fields in band insulators. Proc. Natl. Acad. Sci. U.S.A. 122, e2421778122 (2025). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Cai X., Li Z. X., Yao H., Antiferromagnetism induced by bond Su-Schrieffer-Heeger electron-phonon coupling: A quantum Monte Carlo study. Phys. Rev. Lett. 127, 247203 (2021). [DOI] [PubMed] [Google Scholar]
  • 70.Cai X., Li Z. X., Yao H., Robustness of antiferromagnetism in the Su-Schrieffer-Heeger Hubbard model. Phys. Rev. B 106, L081115 (2022). [Google Scholar]
  • 71.Götz A., Beyl S., Hohenadler M., Assaad F. F., Valence-bond solid to antiferromagnet transition in the two-dimensional Su-Schrieffer-Heeger model by Langevin dynamics. Phys. Rev. B 105, 085151 (2022). [Google Scholar]
  • 72.Feng C., Xing B., Poletti D., Scalettar R., Batrouni G., Phase diagram of the Su-Schrieffer-Heeger-Hubbard model on a square lattice. Phys. Rev. B 106, L081114 (2022). [Google Scholar]
  • 73.Götz A., Hohenadler M., Assaad F. F., Phases and exotic phase transitions of a two-dimensional Su-Schrieffer-Heeger model. Phys. Rev. B 109, 195154 (2024). [Google Scholar]
  • 74.Troyer M., Wiese U. J., Computational complexity and fundamental limitations to fermionic quantum Monte Carlo simulations. Phys. Rev. Lett. 94, 170201 (2005). [DOI] [PubMed] [Google Scholar]
  • 75.Wu C., Zhang S. C., Sufficient condition for absence of the sign problem in the fermionic quantum Monte Carlo algorithm. Phys. Rev. B 71, 155115 (2005). [Google Scholar]
  • 76.Berg E., Metlitski M. A., Sachdev S., Sign-problem–free quantum Monte Carlo of the onset of antiferromagnetism in metals. Science 338, 1606–1609 (2012). [DOI] [PubMed] [Google Scholar]
  • 77.Li Z. X., Jiang Y. F., Yao H., Solving the fermion sign problem in quantum Monte Carlo simulations by majorana representation. Phys. Rev. B 91, 241117 (2015). [Google Scholar]
  • 78.Li Z. X., Jiang Y. F., Yao H., Majorana-time-reversal symmetries: A fundamental principle for sign-problem-free quantum Monte Carlo simulations. Phys. Rev. Lett. 117, 267002 (2016). [DOI] [PubMed] [Google Scholar]
  • 79.Wei Z. C., Wu C., Li Y., Zhang S., Xiang T., Majorana positivity and the fermion sign problem of quantum Monte Carlo simulations. Phys. Rev. Lett. 116, 250601 (2016). [DOI] [PubMed] [Google Scholar]
  • 80.Blankenbecler R., Scalapino D. J., Sugar R. L., Monte Carlo calculations of coupled boson-fermion systems. I. Phys. Rev. D 24, 2278–2286 (1981). [Google Scholar]
  • 81.Hirsch J. E., Two-dimensional Hubbard model: Numerical simulation study. Phys. Rev. B 31, 4403–4419 (1985). [DOI] [PubMed] [Google Scholar]
  • 82.F. Assaad, H. Evertz, World-Line and Determinantal Quantum Monte Carlo Methods for Spins, Phonons and Electrons, H. Fehske, R. Schneider, A. Weiße, Eds. (Springer, Berlin Heidelberg, Berlin, Heidelberg, 2008), pp. 277–356.
  • 83.J. Gubernatis, N. Kawashima, P. Werner, Quantum Monte Carlo Methods (Cambridge University Press, 2016).
  • 84.Senthil T., Fisher M. P. A., Z2 gauge theory of electron fractionalization in strongly correlated systems. Phys. Rev. B 62, 7850–7881 (2000). [Google Scholar]
  • 85.Tanjaroon Ly A., Cohen-Stead B., Malkaruge Costa S., Johnston S., Comparative study of the superconductivity in the Holstein and optical Su-Schrieffer-Heeger models. Phys. Rev. B 108, 184501 (2023). [Google Scholar]
  • 86.Feldbacher M., Assaad F. F., Efficient calculation of imaginary-time-displaced correlation functions in the projector auxiliary-field quantum Monte Carlo algorithm. Phys. Rev. B 63, 073105 (2001). [Google Scholar]
  • 87.Lieb E., Schultz T., Mattis D., Two soluble models of an antiferromagnetic chain. Ann. Phys. 16, 407–466 (1961). [Google Scholar]
  • 88.Oshikawa M., Topological approach to Luttinger’s theorem and the fermi surface of a Kondo lattice. Phys. Rev. Lett. 84, 3370–3373 (2000). [DOI] [PubMed] [Google Scholar]
  • 89.Hastings M. B., Lieb-Schultz-Mattis in higher dimensions. Phys. Rev. B 69, 104431 (2004). [Google Scholar]
  • 90.Kivelson S., Statistics of holons in the quantum hard-core dimer gas. Phys. Rev. B 39, 259–264 (1989). [DOI] [PubMed] [Google Scholar]
  • 91.Arovas D., Schrieffer J. R., Wilczek F., Fractional statistics and the quantum hall effect. Phys. Rev. Lett. 53, 722–723 (1984). [Google Scholar]
  • 92.Isakov S. V., Melko R. G., Hastings M. B., Universal signatures of fractionalized quantum critical points. Science 335, 193–195 (2012). [DOI] [PubMed] [Google Scholar]
  • 93.Yu J., et al. , Non-trivial quantum geometry and the strength of electron-phonon coupling. Nat. Phys. 20, 1262–1268 (2024). [Google Scholar]
  • 94.Jiang H. C., Kivelson S. A., High temperature superconductivity in a lightly doped quantum spin liquid. Phys. Rev. Lett. 127, 097002 (2021). [DOI] [PubMed] [Google Scholar]
  • 95.Li Z. X., Louie S. G., Lee D. H., Emergent superconductivity and non-fermi liquid transport in a doped valence bond solid insulator. Phys. Rev. B 107, L041103 (2023). [Google Scholar]
  • 96.Jiang Y. F., Jiang H. C., Topological superconductivity in the doped chiral spin liquid on the triangular lattice. Phys. Rev. Lett. 125, 157002 (2020). [DOI] [PubMed] [Google Scholar]
  • 97.Jiang Y. F., Yao H., Yang F., Possible superconductivity with a Bogoliubov Fermi surface in a lightly doped Kagome u(1) spin liquid. Phys. Rev. Lett. 127, 187003 (2021). [DOI] [PubMed] [Google Scholar]
  • 98.Peng C., Jiang Y. F., Sheng D. N., Jiang H. C., Doping quantum spin liquids on the Kagome lattice. Adv. Quantum Technol. 4, 2000126 (2021). [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Appendix 01 (PDF)

Data Availability Statement

Some study data are available. The data files are huge. The relevant data are included in the manuscript. Upon request, within reason, we can provide more data to those who ask.


Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES