Chinese Physics Letters, 2022, Vol. 39, No. 7, Article code 077502Express Letter Quantum Spin Liquid Phase in the Shastry–Sutherland Model Detected by an Improved Level Spectroscopic Method Ling Wang1*, Yalei Zhang2, and Anders W. Sandvik3,4* Affiliations 1Department of Physics, Zhejiang University, Hangzhou 310000, China 2Beijing Computational Science Research Center, Beijing 100193, China 3Department of Physics, Boston University, Boston, Massachusetts 02215, USA 4Beijing National Laboratory for Condensed Matter Physics and Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China Received 6 May 2022; accepted; published online 17 June 2022 *Corresponding authors. Email: lingwangqs@zju.edu.cn; sandvik@bu.edu Citation Text: Wang L, Zhang Y, and Sandvik A W 2022 Chin. Phys. Lett. 39 077502    Abstract We study the spin-$1/2$ two-dimensional Shastry–Sutherland spin model by exact diagonalization of clusters with periodic boundary conditions, developing an improved level spectroscopic technique using energy gaps between states with different quantum numbers. The crossing points of some of the relative (composite) gaps have much weaker finite-size drifts than the normally used gaps defined only with respect to the ground state, thus allowing precise determination of quantum critical points even with small clusters. Our results support the picture of a spin liquid phase intervening between the well-known plaquette-singlet and antiferromagnetic ground states, with phase boundaries in almost perfect agreement with a recent density matrix renormalization group study, where much larger cylindrical lattices were used [J. Yang et al., Phys. Rev. B 105, L060409 (2022)]. The method of using composite low-energy gaps to reduce scaling corrections has potentially broad applications in numerical studies of quantum critical phenomena.
cpl-39-7-077502-fig1.png
cpl-39-7-077502-fig2.png
cpl-39-7-077502-fig3.png
cpl-39-7-077502-fig4.png
cpl-39-7-077502-fig5.png
cpl-39-7-077502-fig6.png
cpl-39-7-077502-fig7.png
cpl-39-7-077502-fig8.png
cpl-39-7-077502-fig9.png
DOI:10.1088/0256-307X/39/7/077502 © 2022 Chinese Physics Society Article Text Quantum spin liquids (QSLs)[1] are some of the most intriguing phases of two-dimensional (2D) quantum matter, yet they have been experimentally elusive. The kagome Heisenberg antiferromagnet and the Kitaev honeycomb model are among the most well studied examples. The former hosts a QSL ground state whose nature was debated for years[2] but now is largely settled as a gapless variant.[3,4] The latter has exactly solvable gapped and gapless QSL phases.[5] Both models have attracted enormous attention because of their possible experimental realizations in layered quantum magnets.[6–8] Recent experiments support gapless QSLs in both kagome[9,10] and honeycomb systems.[11,12] Here it should be noted that various defects and disorder can drastically influence gapless excitations and drive quantum magnets to randomness-dominated quantum states completely different from the conjectured pristine gapless QSLs.[13–15] Experimentally, it is often difficult to distinguish between these states, as exemplified by contradictory studies of triangular-lattice systems[16–18] Another prominent quasi-2D frustrated quantum magnet is SrCu$_2$(BO$_3$)$_2$ (SCBO),[19–26] whose in-plane copper magnetic exchange and super-exchange integrals realize the inter-dimer ($J$) and intra-dimer ($J^{\prime}$) interactions of the spin-$1/2$ Shastry–Sutherland model (SSM),[27] as illustrated in Fig. 1(a). Under increasing hydrostatic pressure, the ratio $g\equiv J/J^{\prime}$ increases, and the material undergoes transitions among the three well established ground state phases of the SSM, i.e., the dimer singlet (DS) phase, a plaquette-singlet solid (PSS) phase, and an antiferromagnetic (AFM) phase.[19–26]
cpl-39-7-077502-fig1.png
Fig. 1. Illustration of the spin-$1/2$ models studied here. (a) The SSM, with blue and red lines indicating the Heisenberg AFM interactions $J$ and $J'$, respectively. (b) Heisenberg spin chain with long-range interactions, with all couplings of one spin (top of the ring) to all the other spins marked according to the type of coupling. The blue solid and dashed lines mark unfrustrated AFM (odd distances) and ferromagnetic (even distances) couplings, respectively, and the red lines show the frustrated AFM $J_2$ interactions.
Until recently, SCBO was not widely considered as a candidate for a QSL phase; instead the putative deconfined quantum critical point (DQCP) separating the PSS and AFM phases was the focus of theoretical studies of the SSM[28] and other models with PSS and AFM phases.[29,30] However, a recent density matrix renormalization group (DMRG) study detected a gapless QSL state intervening between the PSS and AFM phases of the SSM[31] within a narrow range of couplings, approximately $g \in (0.79,0.82)$. Subsequently, an intervening phase with similar boundaries was also indicated by a functional renormalization-group calculation.[32] If these results are correct, they open the interesting possibility of a QSL phase also between the PSS and AFM phases in SCBO, somewhere in the pressure range $2.6$ to $3.2$ GPa, where experiments so far[25,26] have not detected any conventional phase transitions or long-range order. This prospect of realizing a gapless QSL is especially important considering that SCBO can be synthesized with very low concentration of impurities, thus, it is free of the complicating disorder effects mentioned above. The aim of the present work is to further corroborate the QSL phase argued in Ref. [31], where excited-state gaps computed with the DMRG method were analyzed. Gap crossings associated with quantum phase transitions were identified, similar to the previously studied $J_1$-$J_2$ square-lattice Heisenberg model[33] (where several other works also agree on the existence of a QSL in roughly the same coupling range[34–38]). Crossing points flowing with increasing system size to two different points were found, $g_{\rm c1} \approx 0.79$ and $g_{\rm c2} \approx 0.82$, and these were associated with transitions out of the PSS phase and into the AFM phase, respectively. The gaps and correlation functions in the window $[g_{\rm c1},g_{\rm c2}]$ supported a gapless QSL phase between the PSS and AFM phases. Here we develop an improved level-spectroscopy method, using combinations of excitation energies beyond the gaps with respect to the ground state. By judicial choices of quantum numbers and identification of composite and elementary excitations on fully periodic lattices, spectral gap combinations can be defined whose crossing points exhibit only very weak dependence on the lattice size. Even with the small clusters accessible with exact diagonalization, we can confirm crossing points in excellent agreement with those extrapolated from the conventional gap crossings in much larger systems with cylindrical boundary conditions.[31] To further demonstrate the improved gap crossing method, we also consider a spin chain with long-range interactions, as illustrated in Fig. 1(b), which has a similar ground-state phase structure as a function of an exponent controlling the long-range couplings. We also study the relevant order parameters of both models. The results further demonstrate the utility of the level crossing method to detect quantum phase transitions when the system sizes are too small to reliably extrapolate the order parameters to the thermodynamic limit. Exact Diagonalization and Level Crossings. Exact diagonalization of the Hamiltonian is the most versatile numerical method for quantum lattice models, however strongly limited to small lattice sizes owing to the prohibitive exponential growth of the Hilbert space. Proper selection of cluster sizes and shapes, and thorough examination of their lattice symmetries (conserved quantum numbers for block-diagonalization), are the two most important steps for fully utilizing the power of the method.[39–43] The quantum numbers are also important for understanding and exploiting excitations, which are useful not only in their own right but also for detecting phase transitions of the ground state. The underlying assumption of the level spectroscopic method we will use here is that a change in the ground state at a quantum phase transition is also accompanied by a change in the elementary excitations, which can be reflected in a re-arrangement of energy levels with different quantum numbers. If that is the case, there will be real level crossings of excited states even when ground state transition takes place through an avoided level crossing (i.e., with the quantum numbers of the ground state on a finite cluster not changing versus the control parameter). The level crossing method is very well known in the context of 1D models, especially the frustrated $J_1$-$J_2$ Heisenberg chain where this approach originated.[44,45] The power of the method in this case lies in the fact that the crossing point between the lowest singlet and triplet excitations versus $J_2/J_1$ converges very rapidly to the critical point with increasing chain length $N$, with shifts proportional to $N^{-2}$. Subleading corrections are small, and the transition point can be obtained to precision $10^{-6}$[45] or even better[43] even with chain lengths only up to $N=32$, easily accessible with exact diagonalization. In other cases, e.g., the chain with long-range interactions that we will also consider here, the subleading corrections are more substantial but still reliable results can be obtained with relatively small chains.[46] More recently, the level-crossing approach has also been applied to 2D systems, in combination with a variety of methods for computing the relevant excited states, e.g., quantum Monte Carlo,[47] DMRG,[33] and sophisticated variational wave functions.[36,37] In the previous application to the SSM,[31] the DMRG method was used to generate excited states in several symmetry sectors on cylindrical lattices (i.e., with open boundaries in one lattice direction and periodic boundaries in the other direction). With fully periodic lattices, results converged to the degree necessary for reliable level-crossing studies are difficult to obtain with the DMRG method for system sizes much beyond those for which exact diagonalization (with, e.g., the Lanczos method) can be used. Periodic lattices are preferable, because of their higher symmetry, thus allowing access to additional quantum numbers beyond those used with the DMRG method. Here we will show that even very small periodic lattices already contain the spectral information pertaining to the ground state transitions of the SSM, but suitable gaps and combinations of gaps have to be identified.
cpl-39-7-077502-fig2.png
Fig. 2. The seven clusters studied in this work. For each system size $N$, a cut-out from the infinite lattice is indicated and periodic boundary conditions are applied to these finite clusters. The clusters are arranged according to their different symmetries and the sizes $N$ are (a) $16$, (b) $36$, (c) $32$, (d) $20$, (e) $40$, (f) $24$, and (g) $28$, with the sites included in each cluster marked by the black circles. The lattice symmetries are illustrated as follows: Gliding reflection operators $\mathcal{G}_x$ and $\mathcal{G}_y$ in (a) and (b), defined in Eq. (5), involve reflection with respect to the blue lines; analogous operations $\mathcal{G}'_x$ and $\mathcal{G}'_y$ are defined for the cluster in (c). Mirror reflections $\sigma_1$ and $\sigma_1$ are defined with respect to the red lines in (a)–(c). Rotation $R_\phi$ by an angle $\phi$ is defined with respect to the center of an empty or filled plaquette as indicated by the green semi-circles in (c)–(g).
The SSM Hamiltonian is $$ H=J\sum_{\langle ij\rangle}\boldsymbol{S}_i\cdot \boldsymbol{S}_{j} + J'\sum_{\langle ij\rangle'}\boldsymbol{S}_i\cdot \boldsymbol{S}_{j},~~ \tag {1} $$ where $\boldsymbol{S}_{i}$ are $S=1/2$ operators, $\langle ij\rangle$ in the $J$ sum represents all nearest-neighbor site pairs on a 2D square lattice, and $\langle ij\rangle'$ in the $J'$ sum indicates next-nearest neighbor sites belonging to one of the diagonal bonds in a given plaquette but only in every second plaquette, as illustrated in Fig. 1(a). As is customary, we will refer to the plaquettes with and without $J'$ interaction as filled and empty, respectively. We use the Lanczos method for periodic clusters with $N=16,20,24,28,32,36$; see Fig. 2. The $N=40$ system is beyond the reach of the Lanczos method within our computational resources, but some of its low-energy states can be completely converged by the implementation of the DMRG method described in Ref. [31]. For larger fully periodic clusters, convergence of excited states to the degree we demand here also becomes too challenging for DMRG (in contrast to the much larger cylindrical lattices studied previously[28,31]). The specific Heisenberg chain with long-range interactions, which we also study as a benchmark case, is defined by $$ H=\sum_{i=1}^N\sum_{r=1}^{N/2}J_r \boldsymbol{S}_i\cdot \boldsymbol{S}_{i+r},~~ \tag {2} $$ where the distance dependent couplings are given by[46] $$ J_2=g, \quad J_{r\neq 2}=\frac{(-1)^{r-1}}{r^{\alpha}}\Big(1+\sum_{r^{\prime}=3}^{N/2}\frac{1}{r^{\prime\alpha}}\Big)^{-1},~~ \tag {3} $$ with adjustable parameters $\alpha$ and $g$ and the normalization of $J_{r\not=2}$ chosen such that the sum of all nonfrustrated ($r\not=2$) interactions $|J_r|$ equals $1$. This model has been studied in previous works using the conventional level-crossing approach with energies computed with the Lanczos method for $N$ up to $32$[43] as well as with DMRG (in this case with fully periodic boundary conditions) with $N$ up to $48$.[33] The existence of a gapless QSL in this 1D model is not controversial, as even the Heisenberg chain with only nearest-neighbor interactions has a disordered ground state with algebraically decaying correlations. With the long-range un-frustrated interactions, long-range AFM order stabilizes when $\alpha$ is below a critical value close to $2$, with the exact value depending on short-distance details of $H$.[48] The third phase in this case is the same frustration-driven two-fold degenerate dimerized phase as in the $J_1$-$J_2$ chain. The QSL can be expected on general grounds for some range of the model parameters to be located between the AFM and dimer phases, and this was confirmed in Refs. [33,46]. Here we will show that the improved level crossing method that we developed for the SSM produces better results for the chain Hamiltonian Eq. (2) as well. The behavior of various gap crossing points, as the system transitions from dimerized to QSL and then to AFM, are very similar to those observed in the SSM. Symmetries of the SSM. The lattice symmetries exploited here are illustrated in Fig. 2 for all the SSM clusters used in our study. These symmetries are used to block diagonalize the Hamiltonian along with the conserved magnetization $S^z$ and the spin-inversion symmetry $Z$ (the latter only for $S^z=0$ states). We do not use the total spin $S$ for block diagonalization, because of the complicated basis vectors in this case, but we compute $S$ of the eigenstates after the diagonalization procedure. We first discuss the point-group symmetries of the standard $4\times 4$ ($N=16$) and $6\times 6$ ($N=36$) clusters; see Figs. 2(a) and 2(b). These clusters have translational symmetry in the ${x}$ and ${y}$ lattice directions, which we define using the operators $$ \mathcal{T}_x=T^2_x,~~~~ \mathcal{T}_y=T^2_y,~~ \tag {4} $$ where $T_x$ and $T_y$ denote the operations of translating by one lattice spacing in the respective directions. Periodic boundaries for an $L \times L$ cluster with even $L$ imply the conditions $\mathcal{T}^{L/2}_x=\mathcal{T}^{L/2}_y=1$. We use the gliding reflection symmetries defined by $$ \mathcal{G}_x=T_yP_x,~~~~ \mathcal{G}_y=T_xP_y,~~ \tag {5} $$ where $P_x$ and $P_y$ are mirror (reflection) operations with respect to vertical and horizontal lines passing through lattice sites. We also use diagonal mirror reflections $\sigma_1$ and $\sigma_2$, defined with respect to lines drawn through intra-dimer ($J'$) bonds. The $L \times L$ clusters are also invariant under the composite rotation defined as $$ \mathcal{R}=T_xT_yR_{\pi/2}=\mathcal{G}_x\sigma_1,~~ \tag {6} $$ where $R_{\pi/2}$ is the 90$^{\circ}$ rotation operation, but this composite symmetry does not further reduce the size of the Hamiltonian blocks after the other symmetries have been used. We nevertheless compute the eigenvalue of $\mathcal{R}$ using that of $\mathcal{G}_x$ and $\sigma_1$. The $N=32$ cluster is contained in a square that is $45^{\circ}$ rotated with respect to the lattice axes; see Fig. 2(c). Defining $T^{\prime}_x$ and $T^{\prime}_y$ as translations along the diagonal directions by one step, the cluster is invariant under the following operations: $\mathcal{T}^{\prime}_x=T^{\prime 2}_x$, $\mathcal{T}^{\prime}_y=T^{\prime 2}_y$, $\mathcal{G}^{\prime}_x=T^{\prime}_yP^{\prime}_x$, $\mathcal{G}^{\prime}_y=T^{\prime}_xP^{\prime}_y$, $\sigma_1$, $\sigma_2$, and $R_{\pi/2}$. Here we have defined $P^{\prime}_x$ and $P^{\prime}_y$ as mirror operations with respect to diagonal lines passing only through empty plaquettes. Imposing periodic boundary conditions corresponds to $\mathcal{T}^{\prime 2}_x=\mathcal{T}^{\prime 2}_y=1$. For this cluster, the rotation symmetry $R_{\pi/2}$ is also useful for block diagonalization. The $N=20$ and $N=40$ clusters, Figs. 2(d) and 2(e), are invariant under $\mathcal{T}_x$ and $\mathcal{T}_y$, and because of the tilting the periodicity implies $\mathcal{T}_x^2\mathcal{T}_y=1$ for $N=20$ and $\mathcal{T}_x^3\mathcal{T}_y=1$ for $N=40$. We also use the $90^\circ$ rotation symmetry, $R_{\pi/2}$, with respect to the center of an empty plaquette. Finally, the $N=24$ and $N=28$ clusters, Figs. 2(f) and 2(g), are similar, being symmetric with respect to a $180^\circ$ rotation $R_{\pi}$ about the center of a filled plaquette. The translational constraints are $\mathcal{T}_x\mathcal{T}_y^2=1$ and $\mathcal{T}_x^2\mathcal{T}_y=1$, respectively, for $N=24$ and $N=28$.
Table 1. Quantum numbers corresponding to the various point-group and spin symmetries for the investigated low-energy states of clusters with $N=16$, $N=32$, and $N=36$. All states have quantum number $+1$ (momentum zero) of the applicable translations $\mathcal{T}_x, \mathcal{T}_y$ or $\mathcal{T}^{\prime}_x, \mathcal{T}^{\prime}_y$. The spin inversion symmetry $Z$ is used only when $S^z=0$.
$\mathcal{G}_x$,$\mathcal{G}^{\prime}_x$ $\mathcal{G}_y$,$\mathcal{G}^{\prime}_y$ $\sigma_1$ $\sigma_2$ $\mathcal{R}$,$R_{\pi/2}$ $S^z$ $S$ $Z$
$S_1$ 1 1 1 1 1 0 0 1
$S_2$ $-1$ $-1$ $-1$ $-1$ 1 0 0 1
$T_1$ $-1$ $-1$ 1 1 $-1$ 0 1 $-1$
$T_2$ 1 1 $-1$ $-1$ $-1$ 0 1 $-1$
$Q_1$ 1 1 1 1 1 2 2
Table 2. Quantum numbers of the investigated state with respect to the applicable rotations for $N=20$, $N=24$, and $N=28$ clusters. All states have momentum zero.
$R_{\pi/2}$ $R_{\pi}$ $S^z$ $S$ $Z$
($N=20$) ($N=24,28$)
$S_1$ 1 1 0 0 1
$S_2$ 1 1 0 0 1
$T_1$ $-1$ 1 0 1 $-1$
$T_2$ $-1$ 1 0 1 $-1$
$Q_1$ 1 1 2 2
Characteristic SSM Eigenstates. Upon increasing $g$, the SSM undergoes a first-order quantum phase transition between the unique DS state and the two-fold degenerate PSS state by a true level crossing at $g\approx 0.685$.[49,50] We here focus solely on changes in the low-energy level spectrum for $g \ge 0.7$, excluding the well understood DS phase and the trivial transition out of it. We target the quantum phase transition from the PSS ground state to the putative QSL state at $g = g_{\rm c1} \approx 0.79$, followed by the transition from this state into the AFM state at $g = g_{\rm c2} \approx 0.82$.[31] Thus, we aim to understand how the low-energy spectrum changes as a function of $g$, as in Ref. [31] but with important differences because of the cylindrical boundary conditions used previously and the fully periodic clusters studied here. The two-fold degenerate singlet ground state is an essential and useful feature of the PSS phase of the SSM on the fully periodic clusters studied here. We label these states, whose degeneracy is lifted by finite-size effects, as $S_1$ and $S_2$. The characteristic Anderson rotor tower of states[51] is a hallmark of AFM order, and we consider the first two of these multiplets; the triplet excitation $T_1$ (which we compute in the $S^z=0$ sector) and the quintuplet $Q_1$ (for practical reasons computed in the $S^z=2$ sector). The intermediate QSL state of the SSM argued in Ref. [31] has not yet been fully characterized, and, thus, there are no rigorously known distinguishing spectral features of it. However, the results of Ref. [31] indicate that it should have gapless singlet and triplet excitations. Thus, all three phases under consideration should have gaps that vanish as the system size is increased, and we are interested in potential level crossings signaling the ground state phase transitions. In addition to the four low-energy states $S_1$, $S_2$, $T_1$, and $Q_1$, discussed above, we also study a triplet $T_2$ that can be regarded as an excitation above $S_2$ with the same relative quantum numbers as those of $T_1$ relative to $S_1$. All states studied here have momentum zero, i.e., the phase factor generated when applying the translation operators $\mathcal{T}_x$ and $\mathcal{T}_y$ in Eq. (4) to these states is $+1$. The absolute and relative lattice quantum numbers of interest here are therefore only the even ($+1$) and odd ($-1$) phases associated with the point-group symmetry operations. The absolute quantum numbers of the $N=16,32$ and $N=36$ clusters are listed in Table 1, and in Table 2 the applicable quantum numbers are similarly listed for $N=20$ and $N=24$, and $28$. For $N=40$, we have not been able to converge the target state $T_2$ with DMRG, but for all other states the quantum numbers are the same as those for $N=20$.
cpl-39-7-077502-fig3.png
Fig. 3. Cartoon picture of the $\pm$ superpositions of $\alpha$ type (bold squares in the left configuration) and $\beta$ type (right configuration) singlet plaquettes that form the two-fold degenerate ground states (quasi-degenerate for finite $N$) $S_1$ ($+$) and $S_2$ ($-$) of the PSS phase. Some of the symmetry operations used to understand (as explained in the text) the quantum numbers of the low-energy excitations $T_1$, $T_2$, and $Q_1$ are indicated with corresponding mirror lines.
The listed quantum numbers in Tables 1 and 2 can be understood with the aid of a cartoon picture of the two lowest singlet states in the PSS phase, illustrated in Fig. 3. These quasi-degenerate ground states of a finite cluster, which do not break the two-fold order-parameter symmetry, are even ($S_1$) and odd ($S_2$) superpositions of the two different plaquette tilings (with singlets on empty plaquettes, as is the case in the SSM[50]) that we refer to as $\alpha$ and $\beta$. In Fig. 3, only two singlet plaquettes on empty squares are highlighted for each case (i.e., those that fit within the small $4\times 4$ cluster). Though the SSM Hamiltonian is not bipartite, below we will also invoke the checkerboard sublattices A and B of the square-lattice sites. First consider operation on the $S_1$ or $S_2$ state by either $\mathcal{G}_x,\mathcal{G}_y,\sigma_1$, or $\sigma_2$ on the clusters in Fig. 2. All these operations effectively exchanges the $\alpha$ and $\beta$ sets of singlet plaquettes, therefore generate a phase (quantum number) $+1$ and $-1$ when acting on the $S_1$ and $S_2$, respectively, thus explaining the corresponding quantum numbers listed in Table 1. To understand the quantum numbers of the triplet excitations, $T_1$ and $T_2$, first note that a plaquette singlet can be regarded as a superposition of two parallel two-spin singlet bonds. Each singlet bond connects the A and B sublattices, thus, are odd with respect to exchanging A$\leftrightarrow$B of the two sublattices. For a system in which the total number of singlet bonds is even, i.e., for $N$ being an integer multiple of four (which is the case for all clusters studied here), the total product wave function of these singlets is even under A$\leftrightarrow$B. If one singlet is excited to a triplet, which is even under A$\leftrightarrow$B, such a state is anti-symmetric with respect to sublattice exchange. Note further that the operators $\mathcal{G}_x,\mathcal{G}_y$ involve A$\leftrightarrow$B site exchange while $\sigma_1$ and $\sigma_2$ do not. Thus, the quantum number $-1$ of $\mathcal{G}_x$ and $\mathcal{G}_y$ in the $T_1$ state arises from swapping A$\leftrightarrow$B because there is an odd number of remaining singlets pairs. Similarly, the quantum number $+1$ for $\sigma_1$ and $\sigma_2$ in $T_1$ follows because there is no sublattice swap. The same reasoning applies to the state $T_2$, i.e., the triplet excitation of $S_2$; the relative sign difference in the gliding and mirror quantum numbers with respect to $T_1$ (Table 1) arises from the odd superposition of the two sets $\alpha,\beta$ of plaquette tilings in $S_2$. The state $Q_1$ can be thought of as the result of exciting two singlet dimers of $S_1$ into triplets, and by applying symmetry operations as above, all reflection quantum numbers remain the same in $Q_1$ as in $S_1$ because of the even number of triplets. The quantum numbers of the rotation operators, $\mathcal{R}$, $R_{\pi}$, or $R_{\pi/2}$, depending on the cluster, can likewise be understood in light of Fig. 3 and how the symmetry operations correspond or not to sublattice and plaquette swaps. As an example, for the $N=32$ cluster the rotation operator $R_{\pi/2}$ swaps the A and B sublattices but not the $\alpha$ and $\beta$ singlet plaquettes. Therefore, for the states $S_1,S_2$ and $Q_1$, which contain an even number of singlet bonds, the quantum number is $+1$, while for $T_1$ and $T_2$, which contain an odd number of singlets, the rotation quantum number is $-1$. The above arguments apply to all clusters in Fig. 2 with their respective applicable symmetry operations. We have explained the quantum numbers by examining a simple picture of the singlets in the PSS phase, and when moving to other phases the energy levels for the finite systems evolve continuously. The states $\{S_1,S_2,T_1,T_2,Q_1\}$ are still defined according to their quantum numbers listed in Tables 1 and 2 and are always those evolving from the two lowest singlets, two lowest triplets, and lowest quintuplet in the PSS state. The state $S_1$ remains the ground state for all values of $g$ considered, and $S_2$, $T_1$, and $Q_1$ also remain the lowest states with their respective total spin. However, $T_2$ is not always the second lowest triplet in the QSL and AFM phases, though it is the first triplet with its full set of quantum numbers. Numerical SSM Results. We define the gaps $\varDelta(S_2)$, $\varDelta(T_1)$, $\varDelta(T_2)$, and $\varDelta(Q_1)$ relative to the ground state energy $E(S_1)$ and graph these versus $g$ in Fig. 4 for the clusters of size up to $N=36$. As explained above, our goal is to identify level (gap) crossings with the PSS–QSL and the QSL–AFM ground state transitions.
cpl-39-7-077502-fig4.png
Fig. 4. Energy gaps of the SSM vs the coupling for cluster sizes (a) $N=36$ (b) $32$, (c) $28$, (d) $24$, (e) $20$, (f) $16$. Conventional gaps defined relative to the ground state energy $E(S_1)$ are shown as follows: $\varDelta(S_2)$ (open red squares), $\varDelta(T_1)$ (open green circles), $\varDelta(T_2)$ (open blue up triangles), $\varDelta(Q_1)$ (open indigo down triangles). Triplet and quintuplet gaps defined with respect to other exited states are shown as follows: $\delta_T \equiv E(T_2)-E(S_2)$ (filled blue up triangles); $\delta_Q = E(Q_1)-E(T_1)$ (filled indigo down triangles). The kinks in the $\varDelta(Q_1)$ and $\delta_Q$ data between $g=0.7$ and $0.75$ are related to avoided level crossings close to the DS–PSS transition.
In Ref. [31], the extrapolated (with leading $1/N$ corrections) crossing point $g_{\rm c1} = 0.788 \pm 0.002$ between the lowest singlet and triplet excitation was identified as the PSS–QSL transition. Unlike the periodic clusters considered here, the cylindrical lattices studied in Ref. [31] break the asymptotic two-fold degeneracy of the PSS state because the boundaries favor one of the two singlet patterns. Thus, the first excited singlet was different from the quasi-degenerate ground state $S_2$ used here, and the level crossing studied previously is not a directly analogy to the singlet-triplet crossing accompanying the dimerization transition in the frustrated Heisenberg chain[44,45] (where the symmetry is not broken in periodic systems). An important aspect of the present work is that the crossing between the $S_2$ and $T_1$ levels is similar to the well understood 1D case, and a confirmation of the same asymptotic crossing point as in Ref. [31] will represent additional independent evidence for the correct identification of the quantum phase transition.
cpl-39-7-077502-fig5.png
Fig. 5. (a) SSM finite-size level crossing points obtained from the gaps $\varDelta(T_1)$, $\delta_T$, $\delta_Q$, and $\varDelta(Q_1)$, each crossing the singlet gap $\varDelta(S_2)$. The points are graphed vs the inverse system size according to the empirical linear scaling in $1/N$.[31,33] The underlying data are from Lanczos calculations such as those in Fig. 4, except for the largest cluster, $N=40$, for which the DMRG method was used. The two straight lines are fits to the $\varDelta(T_1)$ (red solid line) and $\delta_T$ (green solid line) points for $N\geq 20$ and extrapolate to $g_{\rm c1}=0.789$ and $g_{\rm c2}=0.824$, respectively. (b) Adjusted crossing points, $g_{\rm c2}^{\prime}$, Eq. (9), for which all points for given $N$ are shifted vertically by an equal amount so that the $\varDelta(T_1)$ points (red squares) fall exactly on the red fitted line from (a). A linear fit (green line) in $1/N$ is shown for the $\delta_T$ crossing points and extrapolates to $g_{\rm c2} = 0.826$. The form $g_{\rm c2}^{\prime}(N)=g_{\rm c2}+a/N+b/N^{3/2}$ was fitted to the other two data sets ($N\geq 20$) with $g_{\rm c2}$ constrained to the same value as above.
In Fig. 4, the crossing of the $\varDelta(S_2)$ and $\varDelta(T_1)$ gaps indeed are also close to the previous $g_{\rm c1}$ value for all clusters. Interpolated crossing $g$ values are graphed versus $1/N$ in Fig. 5(a) (red squares), where we include also the $N=40$ result obtained with the DMRG method. Here the overall size dependence is much weaker than in the cylindrical lattices,[31] though there is some un-smoothness as a consequence of the different cluster shapes. A line fit to all but the $N=16$ point gives $g_{\rm c1} = 0.789 \pm 0.004$ (where the estimated error, here and in other extrapolations reported below, was obtained from additional fits to all data sets excluding one of the points), in remarkable agreement with the value cited above from the much larger cylindrical lattices (up to $N=24\times 12$ spins). The weak size dependence of the crossing points and the consistency of the two calculations illustrate the advantage of periodic boundary conditions and also confirm the quantum-critical point with a different level crossing. The extrapolated crossing point between the lowest singlet and quintuplet excitations, $g_{\rm c2} = 0.820 \pm 0.002$, was identified as the QSL–AFM transition.[31] This crossing point had a much larger size dependence on the cylindrical lattices than the singlet-triplet crossing. The larger size dependence is also seen with our small periodic clusters, where the crossing points between $\varDelta(S_2)$ and $\varDelta(Q_1)$ are outside the range of Fig. 4. The crossing values, graphed in Fig. 5(a) (indigo down triangles), are consistent with the value of $g_{\rm c2}$ cited above but are too scattered for a meaningful extrapolation. Physically, the singlet-quintuplet crossing is motivated by the Anderson tower of rotor states in the AFM phase. The $S=0$ ground state $S_1$ is the lowest of these states, whose gaps with respect to $E(S_1)$ scale as $S(S+1)/N$ for $S>0$.[51] Other singlets, including $S_2$, have energies above these rotor states (for any $S>0$ and sufficiently large $N$). The triplet $T_1$, which becomes the $S=1$ rotor state in the AFM phase, already crosses from above to below $S_2$ at the PSS-QSL transition point $g_{\rm c1}$, as discussed above. There is no necessary reason why $Q_1$ should fall below $S_2$ in the QSL phase, e.g., in a scenario of a deconfined phase the quintuplet should contain four excited spinons, while $S_2$ and $T_1$ should be two-spinon excitations. However, being the $S=2$ rotor state in the AFM phase, $Q_1$ has to be below $S_2$ there. Thus, the $g$ value of the crossing between $\varDelta(Q_1)$ and $\varDelta(S_2)$ in the limit of infinite system size should coincide with the formation of AFM long-range order. The fact that the extrapolated crossing point $g_{\rm c2}$ indeed is larger than $g_{\rm c1}$ (in Ref. [31] and further below) supports an extended QSL phase instead of a direct transition point between the PSS and AFM phases. Here our aim is to identify other gap crossings associated with the QSL–AFM transitions, in particular with the hope of reducing the size dependence and allowing reliable extrapolation of $g_{\rm c2}$ even with small clusters. We note that the lower transition point $g_{\rm c1}$, as obtained in Ref. [31] and confirmed here, should not be controversial as it is close to other estimates of the end of the PSS phase,[28,50] in particular, in Ref. [28] the size dependence of the point marking the upper PSS bound is consistent with our $g_{\rm c1}$ value. To construct better $g_{\rm c2}$ estimators, we first observe that the second triplet gap $\varDelta(T_2)$ in Fig. 4 closely follows the singlet gap $\varDelta(S_2)$, reflecting the fact that $T_2$ can be regarded as a triplet excitation of $S_2$, in correspondence to the role of the first triplet $T_1$ with respect to the ground state $S_1$. Given that $S_1$ and $S_2$ are quasi-degenerate ground states in the PSS phase, the difference $$ \delta_T \equiv E(T_2)-E(S_2) \equiv \varDelta(T_2)-\varDelta(S_2)~~ \tag {7} $$ will also converge with increasing system size to the non-zero gap in this phase, and $\delta_T$ must then be above the singlet splitting $\varDelta(S_2)$ for sufficiently large $N$ (as is seen clearly in Fig. 4 for all clusters). As already discussed above, in the AFM phase $S_2$ must be above the low-lying Anderson $S>0$ rotor states. However, given that $S_2$ remains the lowest singlet excitation also in the AFM phase, it must also host long-range order and its own associated Anderson rotor tower. As $T_1$ is the lowest rotor excitation of $S_1$, the composite excitation $T_2$ is the lowest rotor state excited from $S_2$. Thus, in the AFM phase $\delta_T \propto 1/N$ and $\delta_T < \varDelta(S_2)$, which is also seen for larger $g$ values in Fig. 4. In the putative gapless QSL phase, we expect $S_2$ to still be the lowest excited singlet (which is also found numerically) and $\varDelta(S_2)$ should vanish with increasing $N$. Likewise, $\varDelta(T_1)$ should vanish as $N \to \infty$. Both the singlet and triplet gaps were found to scale as $N^{-1/2}$ on cylinders in Ref. [31]. We also expect such scaling of the gap of $T_2$ relative to $S_2$, i.e., $\delta_T \propto N^{-1/2} $. If $\delta_T$ remains larger than $\varDelta(T_1)$ and $\varDelta(S_2)$ also inside the QSL phase (as in the PSS phase), then the crossing point of $\varDelta(S_2)$ and $\delta_T$ will signal the QSL–AFM transition. While we have no formal proof of this behavior, on general grounds one can expect a composite excitation, such as $T_2$ excited from $S_2$, to be energetically more costly than its analogous elementary excitation, here $T_1$ obtained from the ground state $S_1$. These expectations are indeed borne out by the numerical crossing points between $\varDelta(S_2)$ and $\delta_T$ in Fig. 5(a) (green circles), where we observe a surprisingly weak size dependence. Fitting a line to the data graphed versus $1/N$ for $N \ge 20$, the extrapolated QSL–AFM transition point is at $g = 0.824 \pm 0.008$, fully consistent with $g_{\rm c2} = 0.820 \pm 0.002$ obtained previously with the larger cylindrical clusters. In this case, we do not have results for $N=40$, as the DMRG calculation for $T_2$ also demands calculation of several other triplets between $T_1$ and $T_2$ (with different quantum numbers that are not resolved in our DMRG implementation[52]). For yet another gap crossing corresponding to the QSL–AFM transition, we can construct a quantity similar to $\delta_T$, Eq. (7), based on the quintuplet state $Q_1$. In analogy with $T_2$ being an excitation of $S_2$, we can also regard $Q_1$ as a further excitation of $T_1$. Defining the corresponding relative gap as $$ \delta_Q\equiv E(Q_1)-E(T_1)\equiv\varDelta(Q_1)-\varDelta(T_1),~~ \tag {8} $$ we can make the same kind of arguments as in the case of $Q_1$ and $S_2$ in the AFM phase, now with $\varDelta(Q_1)=3\varDelta(T_1)$ asymptotically from the Anderson tower energies. Thus, asymptotically $\delta_Q \to 2\varDelta(T_1)$ and we must have $\delta_Q < \varDelta(S_2)$ in the AFM phase. Thus, we expect that the QSL–AFM transition is associated with the asymptotic crossing of $\delta_Q$ and $\varDelta(S_2)$. Such crossing points are indeed within the range of the graphs in Fig. 4, and in Fig. 5(a) the size dependence of the crossing $g$ values based on $\delta_Q$ (blue up triangles) is significantly reduced below that of $\varDelta(Q_1)$. Visually the points are consistent with an asymptotic flow to $g_{\rm c2}$, though the behavior is not smooth enough for extrapolating reliably. An interesting observation in Fig. 5(a) is that the conventional singlet-triplet crossing points (red open squares) and the crossing of the singlet and $\delta_T$ (green open circles) are highly correlated. Therefore, the distance between the points, i.e., asymptotically the size of the QSL phase, has much less size dependence than the individual crossing $g$ values. Upon close inspection, such correlations are also visible in the other $g_{\rm c2}$ estimates. In Fig. 5(b) we exploit these correlations (which should arise from the cluster shape affecting all low-energy excitations in a similar way) by plotting points that are shifted by equal amounts up or down for given $N$, so that the $\varDelta(T_1)$ points coincide exactly with the line fitted to those points in Fig. 5(a). In other words, we cancel out the cluster-dependent correlation effects by focusing on the relative crossing points but still taking into account the overall $g$ scale by adding the values corresponding to the line extrapolating to $g_{\rm c1}$. This procedure defines adjusted crossing points $$ g_{\rm c2}^{\prime}(N)=g_{\rm c2}(N)-g_{\rm c1}(N)+g^{\rm fit}_{\rm c1}(N),~~ \tag {9} $$ where $g_{\rm c1}(N)$ is the $\varDelta(T_1)$ crossing point, $g^{\rm fit}_{\rm c1}(N)$ the corresponding value from the line fit, and $g_{\rm c2}(N)$ is one of the other three crossing points. The so adjusted crossing points in Fig. 5(b) have much smoother size dependence. A line fit to all but the $N=16$ data points in the case of $\delta_T$ gives $g_{\rm c2} = 0.826 \pm 0.003$. This result is only slightly above the previous DMRG cylinder result; the calculations essentially agree within their estimated errors. The adjusted $g_{\rm c2}^{\prime}(N)$ data sets from the $\varDelta(Q_1)$ and $\delta_Q$ crossing points in Fig. 5(b) are also significantly smoothed compared to their original $g_{\rm c2}(N)$ values in Fig. 5(a), though the visibly large corrections to the linear form, in combination with the small number of points, still make independent extrapolations with these data sets difficult. However, by fixing the $N \to \infty$ value to that obtained from the $\delta_T$ linear fits and with $1/N^{3/2}$ corrections included (corresponding to $1/L^3$ when expressed in cluster length $L$), the adjusted $\varDelta(Q_1)$ and $\delta_Q$ data can both be fitted well, thus lending support to all three crossing points flowing to the QSL–AFM transition. The $g_{\rm c2}$ value extrapolated using $g^{\prime}_{\rm c2}(N)$ of course also depends on the line fit to the $\varDelta(T_1)$ data and its extrapolated $g_{\rm c1}$ value. However, the most important aspect of this analysis is that it leaves little doubt that there is a gap $g_{\rm c2}-g_{\rm c1} > 0$ between the two transition points, with the lines fitted to the $\varDelta(T_1)$ and $\delta_T$ data only approaching each other marginally with increasing system size. It should be noted that the estimated size of the QSL phase, $g_{\rm c2}-g_{\rm c1}= 0.037 \pm 0.003$, is not dependent on the two individual line fits but can be obtained by a single line fit to the difference between the $\varDelta(T_1)$ and $\delta_T$ points in Fig. 5(a). Numerical Results for the Spin Chain. Next we use the 1D frustrated spin chain Hamiltonian, Eq. (2), to further validate the conclusions drawn for the 2D SSM. In Ref. [46] the phase diagram was constructed based on Lanczos results for level crossings of gaps with respect to the ground state as well as correlation functions. In the limit of large decay exponent $\alpha$ of the long-range interaction, the model reduces to the well understood frustrated $J_1$-$J_2$ chain, where a quantum phase transition between the critical ground state (a QSL of the Luttinger-liquid type) and a two-fold degenerate dimerized ground state (which we now refer to as a dimer-singlet-solid, DSS) takes place at $J_2/J_1 \approx 0.2411$.[43–45] For $J_2=0$, the interactions are not frustrated, and a previous quantum Monte Carlo and field theory study of a similar model detected a transition between the critical state and an AFM state upon lowering the long range exponent $\alpha$.[48] Long-range AFM order is also intuitively expected when the interactions become strong enough at long distances so that the Mermin–Wagner theorem (which prohibits long-range AFM order for short-range interacting 1D Heisenberg systems) is no longer valid and mean field behavior sets in. Already based on these limiting behaviors, it is clear that the phase diagram in the full parameter space $(J_2,h)$, where for convenience we have defined $h=\alpha^{-1}$, contains DSS, QSL, and AFM phases. These phases can be traversed in said order, similar to the phases of the SSM versus $g$, by following appropriate paths in the parameter space. Based on the previous results for the phase diagram,[46] we here study the phase transitions along a vertical cut with $J_2=0.3$ fixed and $h$ varied. The same line in the phase diagram was also already studied with the DMRG method on periodic chains up to length $N=48$ in Ref. [33], where level crossings of the second singlet $S_2$ with the first triplet ($T_1$) and the first quintuplet ($Q_1$) were extrapolated to infinite size, resulting in $h_{\rm c1}=0.316$ and $h_{\rm c2}=0.476$. We here wish to demonstrate that the finite-size effects are reduced when instead using the composite gaps defined in Eqs. (7) and (8). We only present Lanczos calculations of chains of even length $N$ up to $N=32$.
cpl-39-7-077502-fig6.png
Fig. 6. Energy gaps of the spin chain Hamiltonian in Eq. (2) vs the inverse of the long-range exponent $h=\alpha^{-1}$: $\varDelta(S_2)$ (open red squares), $\varDelta(T_1)$ (open green circles), $\varDelta(T_2)$ (open blue up triangles), and $\varDelta(Q_1)$ (open indigo down triangles). The composite gaps are also shown; $\delta_T$ (filled blue up triangles) and $\delta_Q$ (filled indigo down triangles), as defined in Eqs. (7) and (8), respectively. The results in (a)–(h) were obtained with the Lanczos method for chains of length $N=32$ down to $N=18$ in steps of $2$.
Figure 6 shows the four gaps relative to the ground state $S_1$ for chain sizes from $N=18$ to $N=32$. The conventional gaps $\varDelta(S_2)$, $\varDelta(T_1)$, $\varDelta(T_2)$, and $\varDelta(Q_1)$ are shown along with the two composite gaps $\delta_T$ and $\delta_Q$ defined in the same way as in Eqs. (7) and (8). Here it should be noted that chains of length $N=4n$ (with integer $n$) have ground states with momentum $k=0$, while $N=4n+2$ chains have $k=\pi$ (defined with translation by one lattice spacing). The arguments that we made previously based on Fig. 3 regarding the quantum numbers and physical interpretations of the low-energy states of the SSM apply also to the 1D model with its two-fold degenerate DSS in place of the PSS of the SSM, including the dependence of the ground state momentum on the chain length (from even versus odd number of singlet bonds when $N=4n$ and $4n+2$, respectively). We therefore do not repeat the arguments for the level crossings flowing either to the DSS–QSL transition [$\varDelta(T_1)$ crossing $\varDelta(S_2)$] at $h=h_{\rm c1}$ or to the QSL–AFM transition [$\varDelta(Q_1)$, $\delta_T$, or $\delta_Q$ crossing $\varDelta(S_2)$] at $h=h_{\rm c2}$. In Fig. 7 we present the size dependence of the relevant crossing points, graphing them versus $1/N^2$ in which case we expect asymptotic linear behavior. A clear window $h_{\rm c2}-h_{\rm c1} >0$ between the extrapolated crossing points is apparent here, corresponding to the known QSL phase located between the DSS and AFM phases. Similar to the SSM, the $\varDelta(T_1)$ crossing points only exhibit weak size dependence in their flow to the DSS-QSL transition at $h_{\rm c1}$, here with smooth behavior as all system sizes correspond to the same shape of the lattice for all $N$. The extrapolated transition point is $h_{\rm c1} = 0.3177 \pm 0.0002$, in good agreement with the previous results,[33] where the same level crossing was used but with larger chains. Also similar to the SSM, the weakest size dependence of the three estimates for the QSL–AFM transition point $h_{\rm c2}$ is achieved with $\delta_T$, while $\varDelta(Q_1)$ exhibits the largest variations with $N$. Thus, the use of a composite gap indeed also reduces the finite-size effects in this case. In all cases, the data for the largest clusters can be fitted with lines versus $1/N^2$, with reasonable agreement between the different extrapolations and in good agreement with the previous $h_{\rm c2}$ result. A constrained fit to all data sets (the lines shown in Fig. 7) with a common infinite-$N$ point gives $h_{\rm c2}=0.4600 \pm 0.0005$, which is slightly lower (about $3\%$) than the previous DMRG result and likely more reliable.
cpl-39-7-077502-fig7.png
Fig. 7. Finite-size crossing points extracted from the gap data in Fig. 6. The lines show fits in which a common $N=\infty$ point was enforced in the case of $\delta_T$, $\delta_Q$, and $\varDelta(Q_1)$. Data for small $N$ were excluded to obtain acceptable fits.
Order Parameters. We define the squared AFM (staggered) magnetization for both the SSM and the spin chain in the standard way as $$ m_{\rm s}^2=\frac{1}{N^2}\sum_{ij}\phi_{ij}\langle\boldsymbol{S}_i\cdot\boldsymbol{S}_j\rangle,~~ \tag {10} $$ where $\phi_{ij}=+1$ and $-1$ for sites $i,j$ in the same and different sublattices, respectively. The squared dimer order parameter of the chain model is defined as $$ m_{\rm d}^2=\frac{1}{N^2}\sum_{i,j}(-1)^{i-j}\langle D_iD_j\rangle,~~ \tag {11} $$ where $D_i=\boldsymbol{S}_{i}\cdot\boldsymbol{S}_{i+1}$. In the case of the SSM, singlets forming on the empty plaquettes can be detected with various operators. Here we define a plaquette operator solely with diagonal spin operators; $\varPi_i=\sigma^z_{i}\sigma^z_{i+\hat{x}}\sigma^z_{i+\hat{y}}\sigma^z_{i+\hat{x}+\hat{y}}$, where $\sigma_i^z=2S^z_i$. Then $$ m_{\rm p}^2=\frac{4}{N^2}\sum_{i,j}\theta_{ij}\langle\varPi_i\varPi_j\rangle,~~ \tag {12} $$ where $i,j$ run only over the empty squares of the SSM lattice and $\theta_{ij}=+1$ and $-1$ for $i,j$ in the same row or different rows, respectively.
cpl-39-7-077502-fig8.png
Fig. 8. Squared order parameters of the SSM vs the coupling ratio for different system sizes. (a) Staggered magnetization, Eq. (10), (b) diagonal PSS order parameter, Eq. (12).
Results for both order parameters of the SSM are shown in Fig. 8. While the AFM order parameter increases with $g$ in Fig. 8(a) and the PSS order parameter $m^2_{\rm p}$ correspondingly shows an overall reduction with $g$ in Fig. 8(b), the signals are clearly very weak. It is not possible to extrapolate these order parameters to the thermodynamic limit, and we therefore do not show any such analysis here. In contrast, in the previous DMRG calculations[31] PSS and AFM order were clearly detected on the larger cylindrical lattices in the relevant windows of $g$ values, and inside the QSL phase a power-law behavior of the AFM order parameter was observed. The system sizes accessible to the Lanczos method are simply too small for detecting the phase boundaries, or even to extrapolate the order parameters deep inside the PSS and AFM phases that certainly exist. We have also tried other definitions of the PSS order parameter, e.g., using cyclic permutation operators on the plaquettes instead of the diagonal operators $\varPi_i$ in Eq. (12), but the $g$ dependence is always weak, similar to the data in Fig. 8(b). Results for the chain model are shown in Fig. 9. Here the trends versus the long-range parameter $\alpha^{-1}$ are clearer than in the SSM, but when contrasting the two sets of results it should be kept in mind that the system length is $N$ in the chain but $\sqrt{N}$ in the SSM, and the range of $N$ is similar in both cases. The length is of course what sets the cut-off for the correlation functions and needs to be taken large to reach the asymptotic forms of the squared order parameters. Even in the 1D case, it is not possible to reliably extract the boundaries of the critical phase using the order parameters. This problem is well known from studies of the dimerization transition in the simpler $J_1$-$J_2$ chain and was a strong impetus for the development and use of the level crossing method.[43–45]
cpl-39-7-077502-fig9.png
Fig. 9. Squared order parameters of the chain model vs the inverse of the long-range interaction exponent. (a) Staggered magnetization, Eq. (10), and (b) dimer order, Eq. (11).
Conclusions and Discussion. Our work presented here contributes to a growing sense that level spectroscopy is one of the most powerful generic methods for detecting quantum phase transitions, not only in the well-known context of 1D models[44–46,53,54] but also in 2D systems.[31,33,36,37,47,55–57] The main problem in 2D is that the accessible system sizes are typically small, and finite-size extrapolations of level crossings—the aspect of the level spectrum on which we have focused here—can be challenging. It is therefore important to extract the best possible information from the accessible level spectrum. The main conclusion on methods to draw from our study is that relative gaps between two excited states (composite excitation gaps) are useful alongside the conventional gaps relative to the ground state. In particular, the overall finite-size corrections of composite gap crossing points can be smaller. We have further shown that cluster-shape effects can be significantly removed by consider relative distances between crossing points of different gap combinations for the same system size. The power of the level crossing method to detect quantum phase transitions is further demonstrated by the very weak signals of the different phases in the order parameters. Though the order parameters computed on larger cylinders[31] fully support the level-crossing values of the phase boundaries, on the very small lattices accessible to Lanczos diagonalization they are not yet in the asymptotic regime where the size dependence can be reliably analyzed. The primary model studied here, the SSM, is one of the key models of quantum magnetism, yet its QSL phase situated between the PSS and AFM phases was only proposed very recently.[31] Our new results for the phase boundaries presented here are in remarkable agreement with the previous results, considering the very small clusters used, $N \le 40$, while in Ref. [31] different gap crossings of systems with up to $N=288$ spins were studied. Furthermore, the boundary conditions are different, fully periodic here versus cylindrical in Ref. [31]. This excellent agreement between these two different calculations suggests that the empirically found finite-size scaling behavior of gap crossing points, with leading $1/N$ corrections, is very robust and difficult to explain without the QSL phase located between the PSS and AFM phases. For comparison, we also studied a Heisenberg chain with long-range unfrustrated interactions and short-range frustration. The initial goal of this benchmark test was to investigate the same type of level crossings used for the SSM in a system where a QSL phase located between a two-fold degenerate singlet ground state and an AFM phase is not controversial. Beyond the technical confirmation of the method, including reduced finite-size corrections when a suitable composite gap is used instead of the conventional gaps relative to the ground state, the results also contain useful information pertaining to the nature of the QSL phase in the SSM. Very similar size dependence is apparent of all the crossing points of the 1D model in Fig. 7 and the analogous points for the SSM in Fig. 5, in particular the ordering of the crossing points obtained with the different crossing gaps. This correspondence of low-energy states indicates that the two QSL phases (and critical points) have similar deconfined spinon excitations despite the different dimensionalities, thus, they may have related field theory descriptions. We note that field theories of 2D gapless spin liquids are an active field of investigation,[58–63] and the spectral information obtained here may help to determine the exact nature of the QSL phase in the SSM. First-order direct PSS–AFM transitions have been studied in related 2D quantum spin models[30] and were previously expected also in the case of the SSM.[50,64] The most likely generic scenario for these 2D systems is a line of first-order transitions terminating at a multi-critical deconfined quantum-critical point, after which the QSL phase opens.[31,65] A given model may then either undergo a first-order PSS–AFM transition or cross the QSL phase, as we have argued here in the case of the SSM. We are currently exploring extended SSMs to further explore this scenario. Recent NMR experiments on SCBO have realized a PSS–AFM transition at low temperature, below $0.1$ K, driven by the strength of an external magnetic field at high hydrostatic pressure.[66] The low transition temperature and observed scaling behaviors suggest a nearby critical point at field strength close to $6$ T at a pressure slightly above $2.4$ GPa (the highest pressure studied). This critical point could possibly be the deconfined quantum critical point terminating the SSM QSL phase at a finite magnetic field,[31] thus motivating further studies of the QSL phase of the SSM with a magnetic field added to the Hamiltonian Eq. (1). The SSM QSL phase at zero field may possibly be realized in SCBO somewhere between pressures of $2.6$ and $3.2$ GPa, where heat capacity measurements[25,26] have not detected any phase transitions as a function of the temperature. In SCBO, a complicating factor is that the weak inter-layer couplings will also play some role,[25,29] especially when perturbing a 2D gapless phase. It would therefore also be important to study weakly coupled SSM layers. Acknowledgments. L.W. was supported by the National Natural Science Foundation of China (Grant Nos. 11874080 and 11734002). A.W.S. was supported as a Simons Investigator by the Simons Foundation (Grant No. 511064).
References Spin liquids in frustrated magnetsSpin-Liquid Ground State of the S = 1/2 Kagome Heisenberg AntiferromagnetGapless Spin-Liquid Ground State in the S = 1 / 2 Kagome AntiferromagnetSignatures of Dirac Cones in a DMRG Study of the Kagome Heisenberg ModelAnyons in an exactly solved model and beyondA Structurally Perfect S =1 /2 Kagomé AntiferromagnetSpin Dynamics of the Spin- 1 / 2 Kagome Lattice Antiferromagnet ZnCu 3 ( OH ) 6 Cl 2 Kitaev-Heisenberg Model on a Honeycomb Lattice: Possible Exotic Phases in Iridium Oxides A 2 IrO 3 Colloquium : Herbertsmithite and the search for the quantum spin liquidGapless ground state in the archetypal quantum kagome antiferromagnet ZnCu3(OH)6Cl2Gapless Spin Excitations in the Field-Induced Quantum Spin Liquid Phase of α RuCl 3 Giant phonon anomalies in the proximate Kitaev quantum spin liquid α-RuCl3Valence Bonds in Random Quantum Magnets: Theory and Application to YbMgGaO 4 Random-Singlet Phase in Disordered Two-Dimensional Quantum MagnetsNature of the randomness-induced quantum spin liquids in two dimensionsRare-Earth Triangular Lattice Spin Liquid: A Single-Crystal Study of YbMgGaO 4 Spin-Glass Ground State in a Triangular-Lattice Compound YbZnGaO 4 Scaling and data collapse from local moments in frustrated disordered quantum spin systemsExact Dimer Ground State and Quantized Magnetization Plateaus in the Two-Dimensional Spin System SrCu 2 ( BO 3 ) 2 A Novel Ordered Phase in SrCu2 (BO3 )2 under High PressureMagnetic nanopantograph in the SrCu2 (BO3 )2 Shastry–Sutherland latticeCrystallization of spin superlattices with pressure and field in the layered magnet SrCu2(BO3)24-spin plaquette singlet state in the Shastry–Sutherland compound SrCu2(BO3)2Sign switching of dimer correlations in SrCu 2 ( BO 3 ) 2 under hydrostatic pressureQuantum Phases of SrCu 2 ( BO 3 ) 2 from High-Pressure ThermodynamicsA quantum magnetic analogue to the critical point of waterExact ground state of a quantum mechanical antiferromagnetSignatures of a Deconfined Phase Transition on the Shastry-Sutherland Lattice: Applications to Quantum Critical SrCu 2 ( BO 3 ) 2 Emergent O(4) symmetry at the phase transition from plaquette-singlet to antiferromagnetic order in quasi-two-dimensional quantum magnets*Symmetry-enhanced discontinuous phase transition in a two-dimensional quantum magnetQuantum criticality and spin liquid phase in the Shastry-Sutherland modelRise and fall of plaquette order in the Shastry-Sutherland magnet revealed by pseudofermion functional renormalization groupCritical Level Crossings and Gapless Spin Liquid in the Square-Lattice Spin- 1 / 2 J 1 J 2 Heisenberg AntiferromagnetPlaquette Ordered Phase and Quantum Phase Diagram in the Spin- 1 2 J 1 J 2 Square Heisenberg ModelQuantum Spin Liquid in Spin 1/2 J1J2 Heisenberg Model on Square Lattice: Many-Variable Variational Monte Carlo Study Combined with Quantum-Number ProjectionsGapless spin liquid and valence-bond solid in the J 1 - J 2 Heisenberg model on the square lattice: Insights from singlet and triplet excitationsDirac-Type Nodal Spin Liquid Revealed by Refined Quantum Many-Body Solver Using Neural-Network Wave Function, Correlation Ratio, and Level SpectroscopyDeconfined criticality and a gapless Z 2 spin liquid in the square-lattice antiferromagnetLecture Notes in PhysicsDiagonalization- and Numerical Renormalization-Group-Based Methods for Interacting Quantum SystemsLecture Notes in PhysicsSpringer Series in Solid-State SciencesAIP Conference ProceedingsFluid-dimer critical point in S = antiferromagnetic Heisenberg chain with next nearest neighbor interactionsNumerical evidence for multiplicative logarithmic corrections from marginal operatorsGround States of a Frustrated Quantum Spin Chain with Long-Range InteractionsLevel spectroscopy in a two-dimensional quantum magnet: Linearly dispersing spinons at the deconfined quantum critical pointCritical phenomena and quantum phase transition in long range Heisenberg antiferromagnetic chainsQuantum Phase Transitions in the Shastry-Sutherland Model for SrCu 2 ( BO 3 ) 2 Tensor network study of the Shastry-Sutherland model in zero magnetic fieldNew Approach to the Theory of Superexchange InteractionsTricritical behavior in the extended Hubbard chainsGeneralized Moment Method for Gap Estimation and Quantum Monte Carlo Level SpectroscopyOrder versus disorder in the quantum Heisenberg antiferromagnet on the kagomé lattice using exact spectra analysisDetecting spontaneous symmetry breaking in finite-size spectra of frustrated quantum antiferromagnetsUniversal Signatures of Quantum Critical Points from Finite-Size Torus Spectra: A Window into the Operator Content of Higher-Dimensional Conformal Field TheoriesAlgebraic spin liquid as the mother of many competing ordersTopological Z 2 resonating-valence-bond spin liquid on the square latticeTransition between algebraic and Z 2 quantum spin liquids at large N Anomalous dimensions of monopole operators at the transitions between Dirac and topological spin liquidsThe emergence of gapless quantum spin liquid from deconfined quantum critical pointAnisotropic deconfined criticality in Dirac spin liquidsFirst-order transition between the plaquette valence bond solid and antiferromagnetic phases of the Shastry-Sutherland modelSelf-duality protected multicriticality in deconfined quantum phase transitionsDeconfined quantum criticality and emergent symmetry in SrCu2(BO3)2
[1] Balents L 2010 Nature 464 199
[2] Yan S, Huse D A, and White S R 2011 Science 332 1173
[3] Liao H J, Xie Z Y, Chen J, Liu Z Y, Xie H D, Huang R Z, Normand B, and Xiang T 2017 Phys. Rev. Lett. 118 137202
[4] He Y C, Zaletel M P, Oshikawa M, and Pollmann F 2017 Phys. Rev. X 7 031020
[5] Kitaev A 2006 Ann. Phys. 321 2
[6] Shores M P, Nytko E A, Bartlett B M, and Nocera D G 2005 J. Am. Chem. Soc. 127 13462
[7] Helton J S et al. 2007 Phys. Rev. Lett. 98 107204
[8] Chaloupka J, Jackeli G, and Khaliullin G 2010 Phys. Rev. Lett. 105 027204
[9] Norman M R 2016 Rev. Mod. Phys. 88 041002
[10] Khuntia P, Velazquez M, Barthélemy Q, Bert F, Kermarrec E, Legros A, Bernu B, Messio L, Zorko A, and Mendels P 2020 Nat. Phys. 16 469
[11] Zheng J, Ran K, Li T, Wang J, Wang P, Liu B, Liu Z X, Normand B, Wen J, and Yu W 2017 Phys. Rev. Lett. 119 227208
[12] Li H et al. 2021 Nat. Commun. 12 3513
[13] Kimchi I, Nahum A, and Senthil T 2018 Phys. Rev. X 8 031028
[14] Liu L, Shao H, Lin Y C, Guo W, and Sandvik A W 2018 Phys. Rev. X 8 041040
[15] Kawamura H and Uematsu K 2019 J. Phys.: Condens. Matter 31 504003
[16] Li Y, Chen G, Tong W, Pi L, Liu J, Yang Z, Wang X, and Zhang Q 2015 Phys. Rev. Lett. 115 167203
[17] Ma Z et al. 2018 Phys. Rev. Lett. 120 087201
[18] Kimchi I, Sheckelton J P, McQueen T M, and Lee P A 2018 Nat. Commun. 9 4367
[19] Kageyama H, Yoshimura K, Stern R, Mushnikov N V, Onizuka K, Kato M, Kosuge K, Slichter C P, Goto T, and Ueda Y 1999 Phys. Rev. Lett. 82 3168
[20] Waki T, Arai K, Takigawa M, Saiga Y, Uwatoko Y, Kageyama H, and Ueda Y 2007 J. Phys. Soc. Jpn. 76 073710
[21] Radtke G, Saul A, Dabkowska H A, Salamon M B, and Jaime M 2015 Proc. Natl. Acad. Sci. USA 112 1971
[22] Haravifard S et al. 2016 Nat. Commun. 7 11956
[23] Zayed M et al. 2017 Nat. Phys. 13 962
[24] Bettler S, Stoppel L, Yan Z, Gvasaliya S, and Zheludev A 2020 Phys. Rev. Res. 2 012010(R)
[25] Guo J, Sun G, Zhao B, Wang L, Hong W, Sidorov V A, Ma N, Wu Q, Li S, Meng Z Y, Sandvik A W, and Sun L 2020 Phys. Rev. Lett. 124 206602
[26] Jiménez J L et al. 2021 Nature 592 370
[27] Shastry B S and Sutherland B 1981 Physica B+C 108 1069
[28] Lee J Y, You Y Z, Sachdev S, and Vishwanath A 2019 Phys. Rev. X 9 041037
[29] Sun G, Ma N, Zhao B, Sandvik A W, and Meng Z Y 2021 Chin. Phys. B 30 067505
[30] Zhao B, Weinberg P, and Sandvik A W 2019 Nat. Phys. 15 678
[31] Yang J, Sandvik A W, and Wang L 2022 Phys. Rev. B 105 L060409
[32] Keleş A and Zhao E 2022 Phys. Rev. B 105 L041115
[33] Wang L and Sandvik A W 2018 Phys. Rev. Lett. 121 107202
[34] Gong S S, Zhu W, Sheng D N, Motrunich O I, and Fisher M P A 2014 Phys. Rev. Lett. 113 027201
[35] Morita S, Kaneko R, and Imada M 2015 J. Phys. Soc. Jpn. 84 024720
[36] Ferrari F and Becca F 2020 Phys. Rev. B 102 014417
[37] Nomura Y and Imada M 2021 Phys. Rev. X 11 031034
[38] Shackleton H, Thomson A, and Sachdev S 2021 Phys. Rev. B 104 045110
[39] Laflorencie N and Poilblanc D 2004 Quantum Magnetism 645 227
[40] Noack R M and Manmana S R 2005 AIP Conf. Proc. 789 93
[41] Weisse A and Fehske H 2008 Computational Many-Particle Physics 739 529
[42] Läuchli A M 2011 Numerical Simulations of Frustrated Systems, in Introduction to Frustrated Magnetism: Materials, Experiments, Theory. Springer Series in Solid-State Sciences edited by Lacroix C, Mendels P and Mila F (Berlin: Springer) vol 164 pp 481–511
[43] Sandvik A W 2010 AIP Conf. Proc. 1297 135
[44] Okamoto K and Nomura K 1992 Phys. Lett. A 169 433
[45] Eggert S 1996 Phys. Rev. B 54 R9612
[46] Sandvik A W 2010 Phys. Rev. Lett. 104 137204
[47] Suwa H, Sen A, and Sandvik A W 2016 Phys. Rev. B 94 144416
[48] Laflorencie N, Affleck I, and Berciu M 2005 J. Stat. Mech.: Theory Exp. 2005 P12001
[49] Koga A and Kawakami N 2000 Phys. Rev. Lett. 84 4461
[50] Corboz P and Mila F 2013 Phys. Rev. B 87 115144
[51] Anderson P W 1959 Phys. Rev. B 115 2
[52]Though these triplets with energy between $E(T_{1})$ and $E(T_{2})$ have lattice quantum numbers different from those of $T_{1}$ and $T_{2}$, the corresponding symmetries are not implemented in the DMRG calculation (but the total spin symmetry is implemented in this case), only computed as expectation values with the states obtained. The states have to be generated one-by-one starting from the lowest one and convergence of this procedure becomes increasingly challenging with the number of states computed (Refs.[31, 33]) and we have not been able to reach the state with quantum numbers corresponding to $T_{2}$ for $N=40$.
[53] Nakamura M 2000 Phys. Rev. B 61 16377
[54] Suwa H and Todo S 2015 Phys. Rev. Lett. 115 080601
[55] Lecheminant P, Bernu B, Lhuillier C, Pierre L, and Sindzingre P 1997 Phys. Rev. B 56 2521
[56] Misguich G and Sindzingre P 2007 J. Phys.: Condens. Matter 19 145202
[57] Schuler M, Whitsitt S, Henry L P, Sachdev S, and Läuchli A M 2016 Phys. Rev. Lett. 117 210401
[58] Hermele M, Senthil T, and Fisher M P A 2005 Phys. Rev. B 72 104404
[59] Chen J Y and Poilblanc D 2018 Phys. Rev. B 97 161107(R)
[60] Boyack R, Lin C H, Zerf N, Rayyan A, and Maciejko J 2018 Phys. Rev. B 98 035137
[61] Dupuis E, Boyack R, and Witczak-Krempa W 2021 arXiv:2108.05922 [cond-mat.str-el]
[62] Liu W Y, Hasik J, Gong S S, Poilblanc D, Chen W Q, and Gu Z C 2021 arXiv:2110.11138 [cond-mat.str-el]
[63] Shackleton H and Sachdev S 2022 arXiv:2203.01962 [cond-mat.str-el]
[64] Xi N, Chen H, Xie Z Y, and Yu R 2021 arXiv:2111.07368 [cond-mat.str-el]
[65] Lu D C, Xu C, and You Y Z 2021 Phys. Rev. B 104 205142
[66] Cui Y, Liu L, Lin H, Wu K H, Hong W, Liu X, Li C, Hu Z, Xi N, Li S, Yu R, Sandvik A W, and Yu W 2022 arXiv:2204.08133 [cond-mat.str-el]