cond-mat0605103/vb.tex
1: \documentclass[aps,prb,twocolumn,groupedaddress,amsmath,amssymb,amsfonts,graphics]{revtex4}
2: 
3: \newcommand{\crit}[1]{ {#1}_{\text{c}} }
4: 
5: \DeclareMathOperator{\cycles}{\#}
6: \DeclareMathOperator{\tr}{tr}
7: \DeclareMathOperator{\arcsinh}{arcsinh}
8: 
9: \setcounter{MaxMatrixCols}{18}
10: 
11: \usepackage{graphicx}
12: \usepackage{bm}
13: \usepackage{bbold}
14: 
15: \begin{document}
16: 
17: \title{Some formal results for the valence bond basis}
18: 
19: \author{K.\ S.\ D.\ Beach}
20: \affiliation{Department of Physics, Boston University, 590 Commonwealth Avenue, Boston, MA 02215}
21: \author{Anders W.\ Sandvik}
22: \affiliation{Department of Physics, Boston University, 590 Commonwealth Avenue, Boston, MA 02215}
23: 
24: \date{May 3, 2006}
25: 
26: \begin{abstract}
27: In a system with an even number of SU(2) spins, there is an overcomplete set of 
28: states---consisting of all possible pairings of the spins into valence bonds---that spans
29: the $S=0$ Hilbert subspace. Operator expectation values in this basis are related to the 
30: properties of the closed loops that are formed by the overlap of valence bond states.
31: We construct a generating function for spin correlation functions of arbitrary order
32: and show that all nonvanishing contributions arise from configurations that are
33: topologically irreducible. We derive explicit formulas for the correlation functions
34: at second, fourth, and sixth order. We then extend the valence bond basis to include triplet bonds
35: and discuss how to compute properties that are related to operators acting outside the singlet sector.
36: These results are relevant to analytical calculations and to numerical valence bond
37: simulations using quantum Monte Carlo, variational wavefunctions, or exact diagonalization.
38: 
39: \end{abstract}
40: 
41: \maketitle
42: 
43: \section{Introduction \label{SEC:Introduction}}
44: 
45: The traditional way to represent the quantum states of a system of $S=1/2$ spins is to 
46: introduce a basis of $S^z$ eigenstates. Each state corresponds to a particular assignment 
47: of ``up'' or ``down'' to each spin in the lattice. This basis is complete and orthonormal. 
48: Another useful basis, dating back to Rumer and Pauling in the 1930s,\cite{Rumer32,Pauling33} 
49: is the so-called \emph{valence bond basis}, in which the states of the system are 
50: represented by partitions of the spins into pairs forming singlets. A valence bond (VB) state, 
51: \begin{equation} \label{EQ:vbstate}
52: \lvert v \rangle = \prod_{ij} [i,j],
53: \end{equation}
54: is a product of singlets,
55: \begin{equation}
56: [i,j] = \frac{1}{\sqrt{2}}(\lvert \uparrow_i \downarrow_j \rangle - \lvert \downarrow_i \uparrow_j \rangle),
57: \end{equation}
58: in which each site label appears only once. Pictorially, a VB state can be represented as 
59: a system of hard-core dimers, where each lattice site (spin) belongs to exactly one dimer,
60: as illustrated in Fig.~\ref{FIG:valence-Sz}. The VB basis spans the singlet sector of the 
61: Hilbert space, but is massively overcomplete. It is also highly nonorthogonal, having the 
62: unusual property that every two states in the  basis have nonzero overlap.
63: 
64: \begin{figure}
65: \begin{center}
66: \includegraphics{valence-Sz.eps}
67: \end{center}
68: \caption{ \label{FIG:valence-Sz}
69: The low-energy singlet sector of quantum antiferromagnets can be described
70: in either the $S^z$ or valence bond basis.
71: }
72: \end{figure}
73: 
74: The first important work in this basis was the calculation of the ground-state energy per 
75: spin of the infinite quantum Heisenberg chain by Hulth\'{e}n~\cite{Hulthen38} (building 
76: on the earlier work of Bethe~\cite{Bethe31}). Hulth\'{e}n also determined the eigenvalues 
77: and eigenstates of small finite chains of up to ten sites using the subset of  
78: ``noncrossing''~\cite{Rumer32} valence bond states (which form a complete basis). Majumdar and 
79: Ghosh later performed a similar calculation for the one-dimensional spin chain with nearest- and 
80: next-nearest-neighbour interactions.~\cite{Majumdar69} 
81: 
82: The valence bond basis has a special 
83: connection to the paramagnetic states of interacting antiferromagnets. Fazekas and Anderson 
84: introduced the resonating valence bond (RVB) picture to describe a possible spin liquid in 
85: frustrated magnets.~\cite{Fazekas74} Anderson later proposed that a doped RVB state may 
86: describe the cuprate superconductors.~\cite{Anderson87a,Anderson87b} This suggestion spurred 
87: wide interest in the VB description of antiferromagnets and possible exotic ground states 
88: that are naturally described in terms of VB states. In particular, VB states have proved 
89: useful as variational states, where the variational freedom lies in the distribution of 
90: valence bond lengths~\cite{Liang88,Kohmoto88,Lou06} or in a set of projected-BCS 
91: coefficients.~\cite{Poilblanc89,Gros88,Capriotti01} 
92: 
93: Since valence bond states have total 
94: spin invariance built in, they are also a natural choice for exact diagonalization within 
95: the low-energy $S=0$ matrix block. Such calculations have been used to study RVB states
96: with a restriction of the VB basis to include only a subset of states with short
97: dimers,~\cite{Lin90,Ramsesha90} which should be a good approximation for systems with
98: only short-range spin correlations. A generalization of this procedure---in which the valence 
99: bonds are antisymmetrized products of individual atomic eigenstates---has been widely adopted 
100: by computational chemists for use in molecular quantum mechanics.~\cite{Balint69,Soos90}
101: 
102: Variational calculations suggest that the collinear N\'{e}el ground state of the 
103: $d$-dimensional ($d>1$) Heisenberg antiferromagnet is well-described by a superposition 
104: of VB configurations whose distribution of bond lengths exhibits $1/r^p$ powerlaw 
105: behaviour, with $p < 5$ in two dimensions.\cite{Liang88} In two dimensions, it is also
106: believed that there are magnetically disordered states with energy very close to that 
107: of the ordered ground state.~\cite{Liang88} Thus, competing interactions may favour RVB 
108: states with short-ranged bonds and no long-range magnetic order.~\cite{Figueirido89,Bose91} 
109: They may also favour states with valence bond solid (VBS) order,~\cite{Read89,Read90}
110: in which the translational symmetry is broken but the spin-rotational symmetry remains intact.
111: To date, the only firm confirmation of an RVB state is in simplified quantum dimer 
112: models,~\cite{Kivelson87} in which only the dimer degrees of freedom are retained and the 
113: spin degrees of freedom associated with the dimers are neglected. 
114: 
115: In most cases, possible exotic spin states~\cite{Sondhi97,Sachdev03,Vishwanath04,Senthil04} 
116: simply cannot be studied in detail with unbiased methods such as quantum Monte Carlo 
117: (QMC), because of the infamous negative-sign problem. QMC simulations of quantum spin 
118: systems have traditionally been carried out in the standard $S^z$ basis. Although the 
119: feasibility of a Monte Carlo projection for improving a variational VB state was demonstrated 
120: more than fifteen years ago,\cite{Liang90} the use of the VB basis in QMC studies has been 
121: very limited so far.\cite{Santoro99} As it turns out, neither the overcompleteness nor the 
122: nonorthogonality of 
123: the valence bond basis is an impediment to carrying out QMC simulations in principle. In 
124: fact, even without a good variational state as a trial state, the ground state can be 
125: completely projected out starting with, \emph{e.g.}, an arbitrarily chosen basis state~\cite{Sandvik05}
126: using importance sampling and a simple local updating scheme. This method delivers 
127: performance comparable to the current state of the art for $S^z$-basis QMC calculations
128: (\emph{i.e.}, stochastic series expansion~\cite{Syluasen02} and world line~\cite{Evertz03} 
129: methods with loop-cluster updates). It was also noted in Ref.~\onlinecite{Sandvik05} 
130: that the VB projector method expands  the class of models that are sign-problem-free. 
131: A host of isotropic SU(2) invariant models with multi-spin interactions (involving 
132: four, six, \emph{etc.}~spins) can be studied, which in spite of not being frustrated in the standard 
133: sense (an odd number of antiferromagnetic interactions around a closed loop on the lattice) 
134: do give rise to a sign problem in standard methods. Thus, VB simulations open new 
135: opportunities to study ground state phases and transitions in quantum spin systems.
136: 
137: In addition to these new opportunities for QMC studies in the VB basis, we also 
138: believe that there is more to do variationally. There was not much follow-up on
139: the pioneering calculations by Liang, Doucot, and Anderson,\cite{Liang88} and with 
140: the increase in computer performance since that time, it is now feasible to consider more
141: complex wave-function optimizations.~\cite{Lou06} Although frustrated systems also 
142: cause sign problems in variational calculations, it may still be possible to extract useful 
143: information from them. It may also prove fruitful to study variational VB wave functions 
144: for nonfrustrated multi-spin interactions.
145: 
146: In both QMC and variational calculations, one would like to study a wide range of physical 
147: observables to characterize the ground state. Overlaps and matrix elements between two 
148: VB states $\lvert v\rangle$ and $\lvert v'\rangle$ can be related to the structure of the closed loops 
149: that are formed when their corresponding dimer configurations are superimposed.
150: There are two standard results:
151: (i) For a system of $2N$ spins,  $\langle v | v' \rangle = \pm 2^{N_{\circlearrowleft}-N}$
152: where $N_{\circlearrowleft} \le N$ is the number of loops. 
153: $\langle v | v' \rangle$ is unity when the states $\lvert v \rangle$ 
154: and $\lvert v' \rangle$ are identical and $N_{\circlearrowleft}$ is a maximum; its magnitude 
155: is halved each time a bond mismatch reduces the loop count by 
156: one.~\cite{Rokhsar88,Sutherland88}
157: (ii) $\langle v | \mathbf{S}_i\cdot\mathbf{S}_j | v' \rangle =
158: \pm \frac{3}{4} \langle v | v' \rangle$ if $i$ and $j$ belong to the same loop
159: and vanishes otherwise.~\cite{Liang88}
160: (In each case, the overall sign depends on the convention for assigning directions
161: to the bonds.)
162: We are not aware of expressions for more complicated matrix elements, \emph{e.g.}, 
163: those involving more than two spin operators or individual components of the spins.
164: Nor are we aware of expressions for quantities that are associated with triplet excitations
165: or for quantities, such as the spin stiffness, that cannot usually be written in terms of an 
166: equal-time correlation function. 
167: 
168: The goal of this paper is to provide a formal framework for organizing calculations in the 
169: valence bond basis and to present several new formulas 
170: relating the loop structure of overlapping valence bond states to a wider range of physical 
171: properties of the system. We do this in a formal way, using bond operators equivalent 
172: to those first derived by Sachdev and Bhatt,~\cite{Sachdev90} which in their usual context 
173: are associated with a fixed dimer configuration. Here, we generalize the creation and 
174: annihilation operators that describe the states of the two-spin system to the many-spin case, 
175: allowing the operators to act between arbitrary pairs of spins. With one additional anticommutator 
176: rule, we find that we can use these operators to organize calculations in the overcomplete 
177: valence bond basis. As a result, we are then able to address higher-order spin correlations, 
178: such as $(\mathbf{S}_i\cdot\mathbf{S}_j)(\mathbf{S}_k\cdot\mathbf{S}_l)$, and demonstrate that 
179: their matrix elements are topological in nature, depending not only on the loop membership of 
180: the various site labels (as is the case for $\mathbf{S}_i\cdot\mathbf{S}_j$) but also on the 
181: overall connectivity of the loops with respect to the $(i,j)$ and $(k,l)$ vertices. We introduce 
182: a generating function whose derivatives produce a related class of cumulant function. A 
183: diagrammatic expansion elucidates the structure of these functions and greatly simplifies 
184: their calculation. We extend the valence bond basis to include triplet as well as singlet 
185: bonds in an effort to access the full Hilbert space. This allows us to compute 
186: various spin correlation functions component-wise, including two-spin operators of the form 
187: $S_i^xS^x_j$ and four-spin operators such as $S_i^xS^x_jS^x_kS^x_l$ and $S_i^xS^x_jS^y_kS^y_l$. 
188: We also derive an expression for the singlet-triplet gap
189: and the spin stiffness. These results should be useful in the 
190: context of QMC, variational calculations, and in exact diagonalization.  Our formalism 
191: may also find use in approximate analytical calculations.
192: 
193: The organization of the paper is as follows. In Sec.~\ref{SEC:Formalism}, we introduce 
194: the valence bond operator formalism that is used throughout.
195: The three subsequent sections review how to construct 
196: $S=0$ valence bond states (Sec.~\ref{SEC:Basis}), how to describe their
197: evolution under action by a hamiltonian (Sec.~\ref{SEC:Evolution}),
198: and how to compute their overlaps (Sec.~\ref{SEC:Overlaps}).
199: In Sec.~\ref{SEC:Correlation}, we begin to derive formulas for the isotropic
200: spin correlation functions in a somewhat naive way. We reproduce these results in 
201: Sec.~\ref{SEC:Cumulant} using a more sophisticated diagrammatic approach and then carry 
202: the calculation to higher order. In Sec.~\ref{SEC:Extended}, we develop rules for
203: valence bond states with triplet excitations; these are used in Sec.~\ref{SEC:Components} 
204: to compute correlation functions of  components of the staggered magnetization.
205: In Sec.~\ref{SEC:Neel}, we discuss the N\'{e}el state and how it can be employed
206: as a reference state to compute the singlet-triplet gap (Sec.~\ref{SEC:Singlet-triplet})
207: and the spin stiffness (Sec.~\ref{SEC:Stiffness}).
208: 
209: \section{\label{SEC:Formalism} Valence bond operator formalism}
210: 
211: Two SU(2) spins labelled $i$ and $j$ have a total spin $\mathbf{S} = \mathbf{S}_i+\mathbf{S}_j$
212: satisfying 
213: $\tfrac{1}{2}\mathbf{S}^2 = \tfrac{3}{4} + \mathbf{S}_i\cdot\mathbf{S}_j$.
214: Thus, the usual Heisenberg interaction can be written as
215: \begin{equation} \label{EQ:Hij}
216: \hat{H}_{ij} = \mathbf{S}_i\cdot\mathbf{S}_j - \frac{1}{4} = \frac{1}{2}\mathbf{S}^2 - 1.
217: \end{equation}
218: The eigenstates of Eq.~\eqref{EQ:Hij} are states of well-defined total spin,
219: having eigenvalues $\tfrac{1}{2}S(S+1)-1$. We
220: enumerate them as follows: one ($S=0$)
221: singlet $\lvert 0 \rangle$ and three ($S=1$) triplets $\lvert 1 \rangle$, $\lvert 2 \rangle$, $\lvert 3 \rangle$
222: with energy $E^{\mu} = -\delta^{\mu0}$.
223: 
224: For bookkeeping purposes, it is helpful to introduce a boson representation for the spins, 
225: \begin{equation}
226: \mathbf{S}_i = \frac{1}{2}\sum_{ss'}b^{\dagger}_{is} \bm{\sigma}_{ss'} b_{is'}
227: \ \ \text{with} \ \ \sum_{s} b^{\dagger}_{is}b_{is} = 1
228: \end{equation}
229: (note the single-occupancy constraint),
230: and to define a valence bond operator
231: \begin{equation} \label{EQ:chi-def}
232: \chi^{\mu\dagger}_{ij} = \frac{1}{\sqrt{2}}\sum_{ss'}\tau^{\mu}_{ss'}
233: b_{is}^{\dagger} b_{js'}^{\dagger}
234: \end{equation}
235: that creates eigenstates of $\hat{H}_{ij}$ out of the bosonic vaccuum (\emph{i.e.},
236: $\lvert \mu \rangle = \chi_{ij}^{\mu\dagger} \lvert \text{vac} \rangle$
237: and $\chi_{ij}^{\mu}\lvert \text{vac} \rangle = 0$).
238: The eigenvalue equation $\hat{H}_{ij}\lvert \mu \rangle = E^{\mu}\lvert \mu \rangle$,
239: now written as
240: \begin{equation} \label{EQ:eigenequation}
241: \hat{H}_{ij}\chi^{\mu\dagger}_{ij} \lvert \text{vac} \rangle = 
242: - \delta^{\mu0} \chi^{0\dagger}_{ij}\lvert \text{vac} \rangle,
243: \end{equation}
244: determines the unknown coefficients $\tau^\mu$.
245: One possible solution to Eq.~\eqref{EQ:eigenequation} is
246: \begin{equation} \label{EQ:tau}
247: \tau^\mu = (\tau^0,{\bm \tau}) = (i\sigma^2, i\sigma^3, \mathbb{1}, -i\sigma^1),
248: \end{equation}
249: where $\sigma^{\mu} = (\mathbb{1},{\bm \sigma})$ denotes the four-vector
250: consisting of the unit matrix and the three Pauli matrices; the resulting states are
251: \begin{equation}
252: \lvert \mu \rangle = \chi^{\mu\dagger}_{ij}\lvert\text{vac}\rangle = \begin{cases}
253: \frac{1}{\sqrt{2}}(\lvert \uparrow_i\downarrow_j \rangle - \lvert \downarrow_i\uparrow_j \rangle )
254: & \text{if $\mu = 0$}\\
255: \frac{i}{\sqrt{2}}(\lvert \uparrow_i\uparrow_j \rangle - \lvert \downarrow_i\downarrow_j \rangle)
256: & \text{if $\mu = 1$}\\
257: \frac{1}{\sqrt{2}}(\lvert \uparrow_i\uparrow_j \rangle + \lvert \downarrow_i\downarrow_j \rangle)
258: & \text{if $\mu = 2$}\\
259: \frac{-i}{\sqrt{2}}(\lvert \uparrow_i\downarrow_j \rangle + \lvert \downarrow_i\uparrow_j \rangle)
260: & \text{if $\mu = 3$.}
261: \end{cases}
262: \end{equation}
263: 
264: Other linear combinations of the triplet states are equally valid, but this choice has the advantage 
265: that the labels correspond to real physical directions in the standard basis of $\mathbb{R}^3$. 
266: This is true in the sense that $S^a_i \lvert 0 \rangle \sim \lvert a \rangle$ for $a=1,2,3$.
267: The key is that Eq.~\eqref{EQ:tau} obeys ${\bm \sigma}\tau^0 = i{\bm \tau}$, 
268: which implies that
269: $\mathbf{S}_i\chi_{ij}^{0\dagger}\lvert \text{vac} \rangle
270: = (i/2) {\bm \chi}_{ij}^{\dagger}\lvert \text{vac} \rangle$.
271: 
272: We emphasize that the $\chi_{ij}^{\mu}$ operators completely describe the 
273: $\text{SU(2)}\otimes\text{SU(2)} \simeq \text{SO(4)}$ degrees of freedom
274: of the two-spin system. These operators
275: obey the completeness relation
276: \begin{equation} \label{EQ:chi-completeness}
277: \sum_{\mu} \chi_{ij}^{\mu\dagger}\chi_{ij}^{\mu} =\chi_{ij}^{0\dagger}\chi_{ij}^{0} + \bm{\chi}_{ij}^{\dagger}\cdot\bm{\chi}_{ij} = 1,
278: \end{equation}
279: which follows from 
280: $\sum_{\mu}\tau_{ss'}^\mu\tau^{\mu\dagger}_{r'r}
281:  = 2\delta_{sr}\delta_{s'r'}$ and the restriction to one boson per site.
282: The anticommutation relation
283: \begin{equation} \label{EQ:chi2-commutator}
284: [\chi^{\mu}_{ij},\chi^{\nu\dagger}_{ij}]\lvert \text{vac} \rangle = \delta^{\mu\nu}\lvert \text{vac} \rangle
285: \end{equation}
286: is inherited from the properties 
287: $[b_{is},b_{js'}^\dagger] = \delta_{ij}\delta_{ss'}$ and $b_{is}\lvert \text{vac} \rangle = 0$.
288: 
289: Any valid (\emph{i.e.}, $\sum_s b_{is}^\dagger b_{is}=$1 preserving) operation on one or both of the spins has an SO(4) representation.
290: By construction, the operator equivalence
291: \begin{equation} \label{EQ:SdotS-chi}
292: \mathbf{S}_i \cdot \mathbf{S}_j = -\frac{3}{4} \chi_{ij}^{0\dagger}\chi_{ij}^{0}
293: + \frac{1}{4} \bm{\chi}_{ij}^{\dagger}\cdot\bm{\chi}_{ij}
294: \end{equation}
295: holds, and elimination of ${\bm \chi}$ via Eq.~\eqref{EQ:chi-completeness} yields
296: \begin{equation} \label{EQ:SdotS-quarter-chi}
297: -\hat{H}_{ij} = \frac{1}{4} - \mathbf{S}_i \cdot \mathbf{S}_j = \chi_{ij}^{0\dagger}\chi_{ij}^{0}.
298: \end{equation}
299: For an arbitrary bilinear operator $\hat{O} = \sum_{\mu\nu} O^{\mu\nu}\chi^{\mu\dagger}_{ij}
300: \chi^{\nu}_{ij}$,
301: two applications of Eq.~\eqref{EQ:chi2-commutator}
302: will coax out the coefficient matrix,
303: $O^{\mu\nu} = \langle\text{vac} \lvert [ \chi^\mu, [\hat{O},\chi^{\nu\dagger}]]
304: \lvert \text{vac} \rangle$.
305: 
306: In the case of the spin operators themselves, one finds
307: $(S^a_i)^{\mu\nu} = \frac{1}{2}\tr(\tau^{\mu\dagger}\sigma^a\tau^\nu)$
308: and
309: $(S^a_j)^{\mu\nu} = \frac{1}{2}\tr(\tau^{\nu}\sigma^{a*}\tau^{\mu\dagger})$.
310: Evaluation of the traces leads to
311: \begin{align} \label{EQ:SiminusSj}
312: S^a_i - S^a_j &= i\bigl( \chi^{a\dagger}_{ij} \chi^0_{ij} - 
313: \chi^{0\dagger}_{ij}\chi^a_{ij}\bigr),\\ \label{EQ:SiplusSj}
314: S^a_i + S^a_j &= i\epsilon^{abc}\chi^{b\dagger}_{ij} \chi^c_{ij}.
315: \end{align}
316: Equations~\eqref{EQ:SiminusSj} and \eqref{EQ:SiplusSj} turn out to be 
317: very useful, but we should not regard them as the fundamental operator equivalence 
318: rules. They seem to suggest that two-spin operations are always quartic in $\chi^{\mu}_{ij}$,
319: which is clearly not true [\emph{cf.}\ Eq.~\eqref{EQ:SdotS-chi}].
320: 
321: There is an alternative to computing the coefficient matrix directly.
322: For operators that transform in a known way, we can induce the same
323: transformation in $\chi^{\mu}_{ij}$ and equate the corresponding terms.
324: Consider a spin rotation $S^a \to R^{ab}(\theta\mathbf{n})S^b$
325: of $\theta$ radians about the axis $\mathbf{n}$ ($\lvert \mathbf{n} \rvert = 1$).
326: In the boson language, this rotation is equivalent to 
327: the unitary transformation $b \to Ub$ with
328: $U(\theta\mathbf{n}) = e^{(i\theta/2)\mathbf{n}\cdot\sigma} = \mathbb{1}\cos(\theta/2) + i\mathbf{n}
329: \cdot {\bm \sigma} \sin(\theta/2)$.
330: Now suppose that we rotate $\mathbf{S}_i$ and $\mathbf{S}_j$ by two different angles
331: about the same axis: putting $b_i \to U(\theta_i\mathbf{n})b_i$ 
332: and $b_j \to U(\theta_j\mathbf{n})b_j$
333: into Eq.~\eqref{EQ:chi-def}, 
334: we find that  the rotation 
335: is equivalent to the valence bond operator transformation
336: \begin{equation} \label{EQ:rottransform}
337: \chi^0_{ij} \to \chi^0_{ij}\cos(\theta_{ij}/2) + \mathbf{n}\cdot{\bm \chi}_{ij}\sin(\theta_{ij}/2),
338: \end{equation}
339: where $\theta_{ij} = \theta_i - \theta_j$ is the relative rotation angle.
340: 
341: If $\mathbf{n}$ is directed along the 3 axis, the rotation matrix
342: $R^{ab} = \frac{1}{2}\tr U^\dagger \sigma^a U \sigma^b$
343: has the form
344: \begin{equation}
345: R(\theta\mathbf{e}^3) = \begin{pmatrix} \cos\theta & \sin\theta & 0 \\
346: -\sin\theta & \cos\theta & 0 \\
347: 0 & 0 & 0
348: \end{pmatrix}.
349: \end{equation}
350: Described in terms of raising and lowering operators, the transformation amounts to
351: \begin{equation}
352: \begin{split}
353: S^{+} & \to e^{-i\theta}S^{+}\\
354: S^{-} & \to e^{i\theta}S^{-}\\
355: S^3 & \to S^3,
356: \end{split}
357: \end{equation}
358: and in particular,
359: \begin{equation}
360: S^+_iS^-_j + S^-_iS^+_j \to e^{-i\theta_{ij}} S_i^+S_j^-
361: + e^{i\theta_{ij}} S_i^+S_j^-.
362: \end{equation}
363: Consequently, the isotropic Heisenberg interaction, expanded in powers of $\theta_{ij}$, behaves as
364: \begin{multline} \label{EQ:SidotSjrotate}
365: \mathbf{S}_i\cdot\mathbf{S}_j \to \mathbf{S}_i\cdot\mathbf{S}_j -\frac{i}{2}\theta_{ij} \Bigl( S_i^+S_j^-
366: - S_i^+S_j^-\Bigr)\\ - \frac{1}{4}\theta_{ij}^2\Bigl( S_i^+S_j^-
367: + S_i^-S_j^+\Bigr).
368: \end{multline}
369: 
370: According to Eq.~\eqref{EQ:rottransform}, 
371: the same transformation applied to the valence bond operators gives
372: \begin{multline} \label{EQ:chidaggerchirotate}
373: \chi^{0\dagger}_{ij}\chi^{0}_{ij} \to \chi^{0\dagger}_{ij}\chi^{0}_{ij} +
374: \frac{\theta_{ij}}{2}\bigl( \chi^{0\dagger}_{ij}\chi^{3}_{ij}
375: + \chi^{3\dagger}_{ij}\chi^{0}_{ij} \bigr)\\
376: + \frac{\theta_{ij}^2}{4}\bigl( -\chi^{0\dagger}_{ij}\chi^{0}_{ij} 
377: +\chi^{3\dagger}_{ij}\chi^{3}_{ij} \bigr).
378: \end{multline}
379: Comparing Eqs.~\eqref{EQ:SidotSjrotate} and \eqref{EQ:chidaggerchirotate}
380: allows us to make the identification
381: \begin{equation}
382: \begin{split}
383: i\bigl( S_i^+S_j^- - S_i^+S_j^-\bigr) &= \chi^{0\dagger}_{ij}\chi^{3}_{ij}
384: + \chi^{3\dagger}_{ij}\chi^{0}_{ij}, \\
385: S_i^+S_j^- + S_i^+S_j^- &= -\chi^{0\dagger}_{ij}\chi^{0}_{ij}
386: + \chi^{3\dagger}_{ij}\chi^{3}_{ij}.
387: \end{split}
388: \end{equation}
389: 
390: \begin{table}
391:  \caption{\label{TAB:Opequiv} Operator equivalence rules}
392: \begin{ruledtabular}
393: \begin{tabular}{@{\qquad\qquad}c@{\ \ }|c@{\qquad\qquad}}
394: spin basis & valence bond basis
395:  \tabularnewline \hline
396: $\mathbf{S}_i\cdot\mathbf{S}_j$ & $-\frac{3}{4}\chi_{ij}^{0\dagger}\chi_{ij}^0
397: + \frac{1}{4}{\bm \chi}_{ij}^\dagger\cdot{\bm \chi}_{ij}$
398:  \tabularnewline  
399: $\mathbf{S}_i\cdot\mathbf{S}_j-\frac{1}{4}$ & $-\chi_{ij}^{0\dagger}\chi_{ij}^0$
400:  \tabularnewline  
401: $\mathbf{S}_i - \mathbf{S}_j$ & $i( \chi_{ij}^{0\dagger}{\bm \chi}_{ij}
402: -{\bm \chi}_{ij}^{\dagger}\chi_{ij}^0)$
403:  \tabularnewline  
404: $\mathbf{S}_i + \mathbf{S}_j$ & $i {\bm \chi}_{ij}^{\dagger}\times {\bm \chi}_{ij}$
405:  \tabularnewline  
406:  $S^+_iS^-_j + S^-_iS^+_j$ & $-\chi_{ij}^{0\dagger}\chi_{ij}^0 
407:  + \chi_{ij}^{3\dagger}\chi_{ij}^3$
408:  \tabularnewline  
409:  $i(S^+_iS^-_j - S^-_iS^+_j)$ & $\chi_{ij}^{0\dagger}\chi_{ij}^3 
410:  + \chi_{ij}^{3\dagger}\chi_{ij}^0$
411: \end{tabular}
412: \end{ruledtabular}
413: \end{table}
414: 
415: We have so far confined our discussion to a system of two spins.
416: Nonetheless, everything derived up to this point applies equally well to \emph{any} two spins of
417: a many-spin system. In that sense, the results summarized in Table~\ref{TAB:Opequiv} 
418: are valid generally: we simply assert that there are operators $\chi^{\mu}_{ij}$ associated with
419: \emph{every} pair of sites in the lattice. The only remaining step is to determine what 
420: the appropriate anticommutator algebra is for operators with one site index in 
421: common. A straightforward calculation shows that
422: \begin{equation} \label{EQ:chi3-commutator}
423: [\chi^{\rho}_{ij},\chi^{\mu\dagger}_{kj}\chi^{\nu\dagger}_{il}]\lvert \text{vac} \rangle = \frac{1}{2}\,T^{\lambda\mu\rho\nu}\,\chi^{\lambda\dagger}_{kl}
424: \lvert \text{vac} \rangle,
425: \end{equation}
426: where 
427: $T^{\lambda\mu\rho\nu} = \frac{1}{2}\tr \tau^{\lambda\dagger} \tau^{\mu} \tau^{\rho\dagger} \tau^\nu$
428: and summation over the repeated index $\lambda$ is implied.
429: 
430: \section{\label{SEC:Basis} $S=0$ valence bond basis}
431: 
432: For a system of many SU(2) spins, the structure of the Hilbert space
433: can be determined from the product rule
434: $\tfrac{1}{2} \otimes S = (S-\tfrac{1}{2})\oplus(S+\tfrac{1}{2})$, where $S$
435: denotes the $(2S+1)$-degenerate spin-$S$ state.
436: Thus,
437: \begin{equation}
438: \begin{split}
439: \tfrac{1}{2}\otimes\tfrac{1}{2} &= 0\oplus 1 \\
440: \tfrac{1}{2}\otimes\tfrac{1}{2}\otimes\tfrac{1}{2} & = 
441: \tfrac{1}{2}\oplus\tfrac{1}{2}\oplus \tfrac{3}{2}\\
442: \tfrac{1}{2}\otimes\tfrac{1}{2}\otimes\tfrac{1}{2}\otimes\tfrac{1}{2} &= 0 \oplus 0 \oplus 1 \oplus 1 \oplus 1 \oplus 2\\
443: &\,\,\,\vdots \\
444: \underbrace{\tfrac{1}{2}\otimes\tfrac{1}{2}\otimes\tfrac{1}{2}\otimes\cdots\otimes\tfrac{1}{2}}_{2N\ \text{times}} &=
445: \prod_{S=0}^N \underbrace{S\oplus\cdots\oplus S}_{C^{(N)}_S\ \text{times}}
446: \end{split}
447: \end{equation}
448: The number, $C^{(N)}_{S}$, of $S$-blocks that appears in the Hilbert space of $2N$ 
449: spins is given in Table~\ref{EQ:block-count}. The number of states in the
450: singlet sector, $C^{(N)}_{0} = \frac{1}{N+1}{2N \choose N} = \frac{(2N)!}{N!(N+1)!}$,
451: represents a small fraction of the total number of states,
452: $\sum_{S=0}^{N} (2S+1) C^{(N)}_{S} = 2^{2N}$.
453: 
454: \begin{table}[h]
455:  \caption{ \label{EQ:block-count} Values of the coefficient $C^{(N)}_S$,
456:  representing the number of spin-$S$ blocks in
457:  the Hilbert space of $N$ pairs of SU(2) spins.}
458: \begin{ruledtabular}
459: \begin{tabular}{@{\quad}l|cccccccccc}
460:  & 0 & 1/2 & 1 & 3/2 & 2 & 5/2 & 3 & 7/2
461:  \tabularnewline \hline
462:  & & 1 \tabularnewline
463: 1 &  1 & &  1 \tabularnewline
464:  &   & 2 & & 1  \tabularnewline
465: 2 &  2 & & 3 & & 1 \tabularnewline
466:  & & 5  & & 4 & & 1\tabularnewline
467: 3 & 5 && 9  && 5 && 1 \tabularnewline
468:  &&  14 &&  14 && 6 && 1 \tabularnewline
469: 4 & 14 && 28 && 20  && 7 && 1 \tabularnewline
470:  &&  42 &&  48 && 27 && 8 && 1 \tabularnewline
471: 5 & 42 && 90 && 75  && 35 && 9 \tabularnewline
472:   && 132 && 165 && 120  && 24 && 10 \tabularnewline
473: 6 & 132 && 297 && 285  && 144 && 34 \tabularnewline  
474: \end{tabular}
475: \end{ruledtabular}
476: \end{table}
477: 
478: To construct a valence bond state in the singlet sector
479: [as per Eq.~\eqref{EQ:vbstate}], we simply
480: group the spins into $N$ pairs and, starting from the bosonic vacuum,
481: act with $\chi^{0\dagger}_{ij}$ for each pair $(i,j)$.
482: The number of such states is
483: \begin{equation}
484: T_N = \frac{1}{N!} {2N \choose 2}{2N-2 \choose 2}\cdots{2 \choose 2} = \frac{(2N)!}{2^N(N!)}.
485: \end{equation}
486: This dwarfs the number of independent states in the singlet sector
487: for large $N$ ($T_N \gg C^{(N)}_0$).
488: 
489: As a consequence of the overcompleteness, linear combinations of valence bond 
490: states are generally not  unique---an ambiguity related to the fact that linear independence fails 
491: even at the level of two singlets:
492: \begin{equation} \label{EQ:linearly-dependent}
493: \chi^{0\dagger}_{il}\chi^{0\dagger}_{jk}
494: + \chi^{0\dagger}_{ij}\chi^{0\dagger}_{kl}
495: + \chi^{0\dagger}_{lj}\chi^{0\dagger}_{ik} = 0.
496: \end{equation}
497: In the case of two spins ($N=1$), there is a single tiling ($T_1 = 1$) corresponding to the
498: unique singlet state ($C^{(1)}_0 = 1$). In the case of four spins ($N=2$), the number of 
499: tilings ($T_2 = 3$) exceeds the number of physical singlet states ($C^{(2)}_0 = 2$) by one.
500: It is useful to eliminate this superfluous state. For concreteness, 
501: let us suppose that the lattice of $2N$ spins is described as the union
502: $A\cup B$ of sites $A = \{ i_1, i_2, \ldots, i_N\}$
503: and $B = \{ j_1, j_2, \ldots, j_N\}$.
504: This is simply a labelling trick and does not depend on the
505: lattice being bipartite. We restrict the basis to include only
506: those valence bonds that connect A sites to B sites.
507: This is possible since any unwanted AA and BB bonds can be replaced 
508: by two AB bonds via Eq.~\eqref{EQ:linearly-dependent}.
509: 
510: \begin{figure}
511: \includegraphics{tilings.eps}
512: \caption{\label{FIG:tilings}The set of all possible bond tilings,
513: $T_N = (2N)!/2^NN!$ in number, is shown for $N=1,2,3$.
514: The restricted set of AB tilings (blue) numbers $\tilde{T}_N = N!$,
515: which still exceeds the number of true singlet states.
516: A and B sites are shown as filled and open circles, respectively.
517: }
518: \end{figure}
519: 
520: The AB valence bond basis is intimately related to $\mathcal{S}_N$, the
521: symmetric group of degree $N$. There is a one-to-one correspondence between the bond
522: tilings and the permutations that are the elements of $\mathcal{S}_N$.
523: This correspondence holds because
524: every bond tiling can be thought of as a rearrangement of the $B$ labels.
525: That is, for each $P \in \mathcal{S}_N$, there is a state
526: \begin{equation} \label{EQ:operator-string}
527: \lvert P \rangle = \hat{P}^\dagger\lvert \text{vac}\rangle = \prod_{n=1}^N\chi^{0\dagger}_{i_n,j_{Pn}}\lvert \text{vac} \rangle.
528: \end{equation}
529: It follows that there are $\tilde{T}_N = N!$ such states.
530: For example, Fig.~\ref{FIG:tilings} shows the $3! = 6$ AB bond configurations
531: for $N=3$. These states are indexed by the permutations
532: \begin{align*}
533: P &=  (1)(2)(3) & P &=  (1)(2\ 3) & P &=  (1\ 2)(3) \\
534: P &= (1\ 2\ 3) &  P &= (3\ 2\ 1) & P &= (1\ 3)(2),
535: \end{align*}
536: written here in cycle notation.
537: 
538: The restriction to the AB basis is useful in that it gives us a
539: systematic way to fix the phase of each state.
540: This is an issue because the singlet is a directed bond: $\chi^{0}_{ji} = -\chi^{0}_{ij}$
541: and ${\bm \chi}_{ji} = {\bm \chi}_{ij}$.
542: We adopt the convention that the canonical form of the operator 
543: $\chi^\mu_{ij}$ has $i\in A$
544: and $j \in B$, as in Eq.~\eqref{EQ:operator-string}.
545: As we shall see in Sect.~\ref{SEC:Overlaps}, all overlaps of AB valence bond states are
546: positive definite, which is not true in the unrestricted basis.
547: 
548: One disadvantage to the AB basis is that the identity operator
549: in the $S=0$ subspace has a complicated form.
550: In the unrestricted basis, it is diagonal,
551: \begin{equation} \label{EQ:identityop}
552: \hat{1}_0 = \frac{2^N}{(N+1)!}\sum_{v} \lvert v \rangle \langle v \rvert,
553: \end{equation}
554: with a  normalization $T_N/C^{(N)}_0$\!, but the process of eliminating all AA and BB 
555: bonds from Eq.~\eqref{EQ:identityop} introduces offdiagonal terms.
556: For $N=2$ and $N=3$, the identity operators are
557: \begin{equation}
558: \hat{1}_0 = \frac{2}{3} \sum_{P,Q}\lvert P \rangle
559: \begin{pmatrix} \phantom{+}2 & -1 \\ -1 & \phantom{+}2 \end{pmatrix}_{\!\!P,Q}
560: \!\!\!\!\!\!\langle Q \rvert
561: \end{equation}
562: and
563: \begin{equation}
564: \hat{1}_0 = \frac{1}{3}\sum_{P,Q}\lvert P \rangle
565: \begin{pmatrix} \phantom{+}4 & -1 & -1 & \phantom{+}0 & \phantom{+}0 & -1 \\
566: -1 & \phantom{+}4 & \phantom{+}0 & -1 & -1 & \phantom{+}0 \\ 
567: -1 & \phantom{+}0 & \phantom{+}4 & -1 & -1 & \phantom{+}0 \\ 
568: \phantom{+}0 & -1 & -1 & \phantom{+}4 & \phantom{+}0 & -1 \\ 
569: \phantom{+}0 & -1 & -1 & \phantom{+}0 & \phantom{+}4 & -1 \\
570: -1 & \phantom{+}0 & \phantom{+}0 & -1 & -1 & \phantom{+}4
571: \end{pmatrix}_{\!\!P,Q}
572: \!\!\!\!\!\!\langle Q \rvert. 
573: \end{equation}
574: In these examples, the rows and columns of the coefficient matrices are arranged to match
575: the ordering of the permutations in Fig.~\ref{FIG:tilings}. The entries are 
576: $(2N)!/2^{N-1}(N!)^2-1$ along the main diagonal and $0$ or $-1$ elsewhere, which is
577: a consequence of there being at most one AA/BB set to unravel. For $N>3$,
578: the situation is more complicated, and we do not know of a simple general 
579: expression for $\hat{1}_0$. In most practical applications, however, such an 
580: expression is not needed.
581: 
582: It is possible to construct basis sets that are still more restrictive
583: (\emph{i.e.}, having fewer states than the AB basis).
584: There is a hierarchy of linear dependence relations 
585: [similar to Eq.~\eqref{EQ:linearly-dependent}] for groups
586: of three, four, and higher numbers of singlet bonds, which can
587: be used to further eliminate unwanted states.
588: When carried out to its fullest extent, this elimination procedure
589: leaves a set of valence bond configurations equal in number to the 
590: actual number of $S=0$ states. Explicit construction of the basis
591: in this limit can be accomplished by arranging the lattice sites 
592: $i_1,j_1,i_2,j_2,\cdots,i_N,j_N$ on a ring and keeping
593: only the AB tilings that produce no bond crossings.
594: Note that even this minimal set contains long bonds
595: connecting sites that are macroscopically separated, which
596: suggests that bonds on all length scales are required for completeness.
597: 
598: \section{\label{SEC:Evolution} Evolution of valence bond states}
599: 
600: Now suppose that we have a hamiltonian of the form
601: \begin{equation} \label{EQ:modelH}
602: \hat{H} = \sum_{ij} J_{ij}\hat{H}_{ij} 
603: - \sum_{ijkl} K_{ijkl} \hat{H}_{ij} \hat{H}_{kl} + \cdots
604: \end{equation}
605: with second-, forth-, and potentially higher-order spin interations.~\cite{Raman05}
606: Here, $\hat{H}_{ij}$ is defined as in Eq.~\eqref{EQ:SdotS-quarter-chi}.
607: In order to understand how an arbitrary state evolves under $\hat{H}$,
608: we need to know how a given valence bond state evolves under 
609: repeated applications of $-\hat{H}_{ij}$.
610: 
611: Depending on the circumstances, $-\hat{H}_{ij}$
612: is either a diagonal operation that  leaves the overall configuration unchanged
613: or an off-diagonal operation that maps two bonds to their complementary tiling:
614: \begin{align} \label{EQ:updaterule1}
615: \Bigl(\frac{1}{4}-\mathbf{S}_i\cdot\mathbf{S}_j\Bigr)[i,j] &= [i,j],\\ \label{EQ:updaterule2}
616: \Bigl(\frac{1}{4}-\mathbf{S}_i\cdot\mathbf{S}_j\Bigr)[l,i][j,k] &= [i,j][k,l].
617: \end{align}
618: This result, which is true in the unrestricted basis, has long been known.~\cite{Hulthen38}
619: It will be instructive to see how it emerges within our formalism
620: and how it is modified to accommodate the restriction to AB bonds.
621: 
622: Let us start with the state $\lvert P \rangle = \hat{P}^\dagger\lvert \text{vac}\rangle$,
623: where $\hat{P}^\dagger$ is the operator string defined in Eq.~\eqref{EQ:operator-string}.
624: Then,
625: \begin{equation} \label{EQ:updatePhat}
626: -\hat{H}_{ij}\lvert P \rangle = 
627: \chi_{ij}^{0\dagger}\chi_{ij}^{0}\hat{P}^\dagger\lvert \text{vac}\rangle
628: = \chi_{ij}^{0\dagger}[\chi_{ij}^{0}, \hat{P}^{\dagger}] \lvert\text{vac}\rangle.
629: \end{equation}
630: There are two possibilities 
631: to consider in evaluating the anticommutator: 
632: (i) If there is already a bond connecting the active sites 
633: (\emph{i.e.}, $\hat{P}^\dagger = \cdots \chi^{0\dagger}_{ij} \cdots$),
634: then only two operators play a role in the anticommutator. In this case,
635: Eq.~\eqref{EQ:chi2-commutator} is the appropriate rule to apply, and 
636: Eq.~\eqref{EQ:updatePhat} simplifies to
637: \begin{equation} \label{EQ:nonupdated-op-seq}
638: \chi_{ij}^{0\dagger}\chi_{ij}^{0}\lvert P \rangle = \lvert P \rangle.
639: \end{equation}
640: (ii)
641: If the active sites are each connected elsewhere 
642: (\emph{i.e.}, $\hat{P}^\dagger = \cdots \chi^{0\dagger}_{il}\cdots\chi^{0\dagger}_{kj} \cdots$ 
643: for some $k\in A$, $l \in B$), then
644: three operators are involved. This necessitates the use
645: of Eq.~\eqref{EQ:chi3-commutator} and leads to 
646: \begin{equation} \label{EQ:updated-op-seq}
647: \chi_{ij}^{0\dagger}\chi_{ij}^{0}\lvert P \rangle = 
648: \frac{1}{2} \chi_{i_1j_{P1}}^{0\dagger}\cdots
649: \chi_{ij}^{0\dagger}\cdots\chi^{0\dagger}_{kl} \cdots
650:  \chi_{i_Nj_{PN}}^{0\dagger}
651:  \lvert \text{vac} \rangle.
652: \end{equation}
653: The new state on the right-hand side is itself a valence bond state
654: (different from $\lvert P \rangle$), since it consists of $N$ operators and 
655: none of its $2N$ indices are repeated. If $i\in A$ and $j\in B$,
656: then this state contains only AB bonds.
657: 
658: A reconfiguration of the bonds manifests itself as change in the permutation
659: indexing the state. If $i = i_n$ and $j = j_m$ are in opposite sublattices,
660: then Eqs.~\eqref{EQ:nonupdated-op-seq} and \eqref{EQ:updated-op-seq} 
661: are summarized by the compact update rule
662: \begin{equation} \label{EQ:updateinjm}
663: -\hat{H}_{i_nj_m} \lvert P \rangle = \biggl(\frac{1}{2}\biggr)^{\!1-\delta_{Pn,m}}\lvert (Pn\  m) P \rangle,
664: \end{equation}
665: where the term $(Pn\ m)$ inside the ket is a 2-cycle acting on $P$
666: (the effect of which is to swap the labels $j_{Pn}$ and $j_m$).
667: Equation~\eqref{EQ:updateinjm} makes clear that the hamiltonian is an identity
668: operation whenever it acts on a preexisting bond; otherwise it induces a single 
669: transposition and a multiplicative factor $\frac{1}{2}$. Note that the transposition
670: itself depends on $P$.
671: 
672: If, however, $i$ and $j$ are in the same sublattice, then the 
673: $\chi_{ij}^{0\dagger}\chi^{0\dagger}_{kl}$ operators in
674: Eq.~\eqref{EQ:updated-op-seq}
675: have to be eliminated in favour of AB bonds.
676: This leads to
677: \begin{equation} \label{EQ:updateinim}
678: -\hat{H}_{i_ni_m} \lvert P \rangle = \frac{1}{2}\lvert P \rangle - \biggl(\frac{1}{2}\biggr)^{\!1-\delta_{n,m}}
679: \lvert (Pn\  Pm) P \rangle
680: \end{equation}
681: and
682: \begin{equation} \label{EQ:updatejnjm}
683: -\hat{H}_{j_nj_m} \lvert P \rangle = \frac{1}{2}\lvert P \rangle - \biggl(\frac{1}{2}\biggr)^{\!1-\delta_{n,m}}
684: \lvert (n\  m) P \rangle.
685: \end{equation}
686: These ``frustrating'' interactions create a linear superposition of states, rather
687: than just a rearrangement of bonds. They also introduce bond configurations
688: with negative weight, violating the conditions of the Marshall sign theorem.~\cite{Marshall55}
689: Figure~\ref{FIG:updates} illustrates the frustrating and nonfrustrating bond flips that can occur.
690: 
691: If there are no frustrating interactions in the hamiltonian then the
692: ground state can be written as a superposition of valence bond states
693: $\lvert \psi \rangle = \sum_P c_P \lvert P \rangle$ with all $c_P \ge 0$.
694: In that special case, Monte Carlo simulation is sign-problem-free. 
695: 
696: \begin{figure}
697: \begin{center}
698: \includegraphics{updates.eps}
699: \end{center}
700: \caption{ \label{FIG:updates}
701: AB valence bonds obey a simple set of update rules when acted on with an isotropic
702: spin interaction term. Here the update rules for
703: $1/4-\mathbf{S}_i \cdot \mathbf{S}_j$ are summarized.
704: The red bar denotes the interaction between the spins at sites $i$ and $j$.
705: The site labels follow the notation in Eqs.~\eqref{EQ:updateinjm}
706: and \eqref{EQ:updateinim}. Of the terms that involve a reconfiguration
707: of the bonds (dashed outlines indicate the previous bond locations),
708: one corresponds to an exchange of $j_{Pn}$ and $j_m$ and the 
709: other $j_{Pn}$ and $j_{Pm}$.
710: }
711: \end{figure}
712: 
713: \section{\label{SEC:Overlaps} Overlaps and matrix elements of valence bond states}
714: 
715: It follows from Eq.~\eqref{EQ:operator-string} that the overlap of any two valence bond states
716: is equal to the vacuum-state expectation value of a length-$2N$ operator string:
717: \begin{multline}
718: \langle Q \vert P \rangle = \langle\text{vac}\rvert \chi^{0}_{i_N,j_{QN}}\cdots
719:  \chi^{0}_{i_2,j_{Q2}}\chi^{0}_{i_1,j_{Q1}}\\
720: \times \chi^{0\dagger}_{i_1,j_{P1}}\chi^{0\dagger}_{i_2,j_{P2}}\cdots
721: \chi^{0\dagger}_{i_N,j_{PN}}\lvert \text{vac} \rangle.
722: \end{multline}
723: The creation and annihilation operators can be shuffled past one another
724: so long as they share no indices in common. The maximal such rearrangement
725: leaves the string grouped into several linked chains of operators; these
726: are related to the disjoint cycles of the composite permutation $\bar{Q}P$
727: (we use the notation $\bar{Q} = Q^{-1}$ to denote the inverse permutation of $Q$)
728: and
729: define a set of directed loops (with $\bar{Q}P$ defining the proper order), \emph{e.g.},
730: \begin{multline}
731: i_1 \to j_{P1} \to i_{\bar{Q}P1} \to j_{P\bar{Q}P1}\\
732:  \to i_{\bar{Q}P\bar{Q}P1}
733: \to \cdots \to i_{(\bar{Q}P)^k1} = i_1.
734: \end{multline}
735: Figure~\ref{FIG:overlap} illustrates the construction.
736: Each of these loops is even in length---constituting a chain with equal numbers of
737: links from $P$ and $Q$. The cycle decomposition is related to the loop membership
738: of the site indices by 
739: \begin{equation}\label{EQ:sameloop}
740: \begin{split} 
741: n \overset{\bar{Q}P}{\sim} \bar{P}m &\ \Leftrightarrow\ \text{$i_n$ and $j_m$ in same loop}\\ 
742: n \overset{\bar{Q}P}{\sim} m &\ \Leftrightarrow\ \text{$i_n$ and $i_m$ in same loop}\\ 
743: \bar{P}n \overset{\bar{Q}P}{\sim} \bar{P}m &\ \Leftrightarrow\ \text{$j_n$ and $j_m$ in same loop}
744: \end{split}
745: \end{equation}
746: We use the notation $x \overset{\bar{Q}P}{\sim} y$ to indicate that $x$ and $y$ are in the same cycle of $\bar{Q}P$.
747: All the loops taken together form 
748: a set of closed paths covering the lattice.
749: Hence,
750: \begin{equation} \label{EQ:loop-length-sum}
751: \sum_{k=1}^{N} 2k\cdot n_{k} = 2N,
752: \end{equation}
753: where $n_k$ is the number of loops of length $2k$ (or 
754: the number of $k$-cycles in $\bar{Q}P$).
755: 
756: The value of the overlap can be computed by way of a simple decimation
757: scheme.~\cite{Rokhsar88,Sutherland88} Each loop can be eliminated by
758: repeated application of
759: Eqs.~\eqref{EQ:chi2-commutator} and \eqref{EQ:chi3-commutator}, specialized
760: to the case of $\rho = \nu = \mu = 0$:
761: \begin{equation} \label{EQ:chi02-commutator}
762: [\chi^{0}_{ij},\chi^{0\dagger}_{ij}]\lvert \text{vac} \rangle = \lvert \text{vac} \rangle
763: \end{equation}
764: and
765: \begin{equation} \label{EQ:chi03-commutator}
766: [\chi^{0}_{ij},\chi^{0\dagger}_{kj}\chi^{0\dagger}_{il}]\lvert \text{vac} \rangle = \frac{1}{2}\chi^{0\dagger}_{kl}
767: \lvert \text{vac} \rangle.
768: \end{equation}
769: For a loop of length $2k$, $k-1$ applications of Eq.~\eqref{EQ:chi03-commutator}
770: removes $2k-2$ links and yields $k-1$ factors of 1/2. The final two operators are removed
771: with a single application of Eq.~\eqref{EQ:chi02-commutator}, leaving 
772: $\langle \text{vac} | \text{vac} \rangle  = 1$.
773: Hence,
774: \begin{equation} \label{EQ:QPoverlap}
775: \langle Q | P \rangle = \prod_{l=1}^N
776: \biggl(\frac{1}{2^{k-1}}\biggr)^{n_{k}}
777: = \frac{2^{\sum_{l=1}^N n_{k}}}{2^{\sum_{l=1}^N k\cdot n_{k}}}
778: = 2^{N_{\circlearrowleft}-N},
779: \end{equation}
780: where we have used Eq.~\eqref{EQ:loop-length-sum}
781: and defined the total number of loops $N_{\circlearrowleft} = \sum_{k=1}^N n_{k}$.
782: 
783: \begin{figure}
784: \includegraphics{overlap.eps}
785: \caption{ \label{FIG:overlap}
786: The top-left and top-right panels illustrate valence bond states
787: $\lvert P \rangle$ and $\lvert Q \rangle$ corresponding to the
788: permutations
789: $P = (1\ 2)(3)(4)(5)\cdots$ and
790: $Q = (1)(2\ 4\ 5\ 3)\cdots$;
791: the sites $i_1,i_2,\ldots,i_5$ (filled circles) and $j_1,j_2,\ldots,j_5$ (open circles)
792: are numbered accordingly (in no particular order).
793: The overlap $\langle Q | P \rangle$ between these states
794: is literally that: a superposition of the two bond configurations. 
795: Since the end point of one bond is always the starting point of another,
796: a set of closed bond paths is formed. These are in one-to-one correspondence
797: with the disjoint cycles of $\bar{Q}P$.
798: In this example, $\bar{Q}P = (1\ 2\ 4\ 5\ 3)\cdots$
799: has a cycle of length 5 and a corresponding bond loop of length 10.
800: }
801: \end{figure}
802: 
803: Similar arguments can be used to compute the matrix elements of
804: a spin-rotation-invariant operator $\hat{O}$.
805: In general, such an operator acting on a valence bond state 
806: produces a linear superposition of states with modified bond configurations:
807: \begin{equation} \label{EQ:OhatP}
808: \hat{O}\lvert P \rangle = \sum_{P'} O_{P,P'}\lvert P' \rangle.
809: \end{equation}
810: (The situation is more complicated if $\hat{O}$ generates states outside 
811: the $S=0$ sector; this is discussed further in Sect.~\ref{SEC:Extended}.)
812: Since the basis is overcomplete, the coefficients $O_{P,P'}$ are not unique and
813:  $O_{P,P'} \neq \langle P \lvert \hat{O} \rvert P' \rangle$, as would be
814: the case in an orthonormal basis.
815: Equations~\eqref{EQ:QPoverlap} and \eqref{EQ:OhatP} imply that 
816: matrix elements are related to changes in the loop number:
817: \begin{equation} \label{EQ:QOhatP}
818: \frac{\langle Q \lvert \hat{O} \rvert P \rangle}{\langle Q | P \rangle}
819: =  \sum_{P'} O_{P,P'} 2^{N_{\circlearrowleft}'-N_{\circlearrowleft}}.
820: \end{equation}
821: Here, $N_{\circlearrowleft}$ is the number of cycles in $\bar{Q}P$ and
822: $N_{\circlearrowleft}'$ the number in $\bar{Q}P'$.
823: 
824: Evaluation of Eq.~\eqref{EQ:QOhatP} requires some general rules for how the number of cycles changes
825: as $P \to P'$. Since the permutation $P'\bar{P}$ can always be decomposed into a product of transpositions, it 
826: suffices to consider the situation where the two states differ by a single transposition.
827: In that case, $N_{\circlearrowleft}' - N_{\circlearrowleft} = \pm 1$ (see the Appendix~\ref{SEC:Permutation}), since a transposition $(n\ m)$ either merges the two cycles that each contain one of
828: $\bar{P}n$ and $\bar{P}m$ or splits the single cycle that contains them both.
829: This can be expressed formally as
830: \begin{equation} \label{EQ:overlapratio}
831: \frac{\langle Q \lvert (n\ m) P\rangle}{\langle Q | P \rangle} \biggl(\frac{1}{2}
832: \biggr)^{1-\delta_{n,m}} 
833: = \begin{cases}
834: 1 & \text{if $\bar{P}n \overset{\bar{Q}P}{\sim} \bar{P}m$ } \\
835: \frac{1}{4} & \text{otherwise}.
836: \end{cases}
837: \end{equation}
838: Here, the delta function takes care of the possibility that $n=m$, which corresponds
839: to no transposition at all.
840: 
841: Applying this result to Eqs.~\eqref{EQ:updateinjm}--\eqref{EQ:updatejnjm}
842: [with the right-hand side of Eq.~\eqref{EQ:overlapratio} interpreted according 
843: to \eqref{EQ:sameloop}] gives 
844: \begin{equation} \label{EQ:chidaggerchi}
845: \frac{\langle Q \lvert \chi^{0\dagger}_{ij} \chi^{0}_{ij} \rvert P \rangle}
846: { \langle Q | P \rangle } = \frac{1}{4}\bigl( 1 - 3\epsilon_{ij}\delta^{\alpha_{i}\alpha_{j}} \bigr),
847: \end{equation}
848: where $\alpha_i$ and $\alpha_j$ are unique labels for the loops 
849: passing through sites $i$ and $j$. We have introduced the notation
850: \begin{equation}
851: \epsilon_{ij} = \begin{cases}
852: -1 & \text{if $i,j$ are in different sublattices}\\
853: +1 & \text{otherwise}.
854: \end{cases}
855: \end{equation}
856: 
857: \begin{figure}
858: \includegraphics{Cij.eps}
859: \caption{ \label{FIG:Cij} 
860: The operator
861: $\chi_{ij}^{0\dagger}\chi_{ij}^0 = 1/4-\mathbf{S}_i \cdot \mathbf{S}_j$
862: causes a rearrangement of bonds, as shown in Fig.~\ref{FIG:updates}.
863: Thus, its matrix element can be understood in terms of joining
864: or splitting loops.
865: }
866: \end{figure}
867: 
868: To clarify, let us rederive Eq.~\eqref{EQ:chidaggerchi} by considering explicitly
869: the effect of $\chi^{0\dagger}_{ij}\chi^{0}_{ij}$ acting to the right on $\lvert P \rangle$
870: and how the overlap of the resulting state with $\langle Q \rvert$ differs from
871: the loop structure of $\langle Q | P \rangle$.
872: We distinguish between the AB case, 
873: in which the sites $i$ and $j$ are in different sublattices, and the AA/BB case, 
874: in which they are in the same sublattice.
875: For AB operators, the possibilities are illustrated in Fig.~\ref{FIG:Cij}(a--c).
876: If the two sites belong to the same loop then either an
877: offdiagonal operation splits the loop, giving a contribution $(1/2)(2^1) = 1$ [following 
878: Eq.~\eqref{EQ:QOhatP}]
879: or a diagonal operation leaves the loop unchanged, also giving $(1)(2^0) = 1$.
880: If the two sites belong to different loops then the offdiagonal
881: operation joins the two loops, giving $(1/2)(2^{-1})=1/4$. Combining
882: these results gives
883: \begin{equation} \label{EQ:chiijchiij}
884: \begin{split}
885: \frac{\langle Q \lvert \chi^{0\dagger}_{ij}\chi^{0}_{ij}
886: \rvert P \rangle}
887: {\langle Q | P \rangle}
888: &= \delta^{\alpha_i\alpha_j} + \frac{1}{4}\Bigl(1-\delta^{\alpha_i\alpha_j}\Bigr)\\
889: &= \frac{3}{4}\delta^{\alpha_i\alpha_j} + \frac{1}{4},
890: \end{split}
891: \end{equation}
892: For AA/BB operators, the possible bond reconfigurations are illustrated in
893: Figs.~\ref{FIG:Cij}(d) and \ref{FIG:Cij}(e).
894: The two sites are either in the same loop, giving
895: $(1/2)(2^0-2^1) = -1/2$, or in different loops, giving  $(1/2)(2^0-2^{-1}) = 1/4$. The result,
896: \begin{equation}
897: \begin{split}
898: \frac{\langle Q \lvert \chi^{0\dagger}_{ij}\chi^{0}_{ij}
899: \rvert P \rangle}
900: {\langle Q | P \rangle}
901:  &= -\frac{1}{2}\delta^{\alpha_i\alpha_j}+ \frac{1}{4}\Bigl(1-\delta^{\alpha_i\alpha_j}\Bigr)\\
902: &= -\frac{3}{4}\delta^{\alpha_i\alpha_j} + \frac{1}{4},
903: \end{split}
904: \end{equation}
905: differs from Eq.~\eqref{EQ:chiijchiij} by only a sign. Using $\epsilon_{ij}$
906: to account for the difference in sign yields Eq.~\eqref{EQ:chidaggerchi}.
907: 
908: The same formal manipulations that lead from Eq.~\eqref{EQ:overlapratio} to 
909: Eq.~\eqref{EQ:chidaggerchi} can be chained together to evaluate a long 
910: operator sequence one transposition at a time. For example,
911: taking $i = i_n \in A$ and $j = j_m \in B$, we get
912: \begin{equation}
913: \begin{split}
914: \frac{\langle Q \lvert \hat{O} \chi^{0\dagger}_{ij}\chi^0_{ij} \rvert P \rangle}
915: {\langle Q | P \rangle}
916: &= \frac{\langle Q \lvert \hat{O} \rvert (Pn\ m)P \rangle}
917: {\langle Q | P \rangle} \biggl(\frac{1}{2}\biggr)^{1-\delta_{Pn,m}} \\
918: &= \frac{\langle Q \lvert \hat{O} \rvert P' \rangle}
919: {\langle Q | P' \rangle}
920: \frac{\langle Q | P' \rangle}
921: {\langle Q | P \rangle} \biggl(\frac{1}{2}\biggr)^{1-\delta_{Pn,m}} \\
922: &= \frac{\langle Q \lvert \hat{O} \rvert P' \rangle}
923: {\langle Q | P' \rangle}
924: \frac{1}{4}\bigl( 1 + 3\delta^{\alpha_i \alpha_j}\bigr),
925: \end{split}
926: \end{equation}
927: where $P' = (Pn\ m)P$ describes the new bond configuration after one transposition.
928: For arbitrary site indices, the expression reads
929: \begin{multline} \label{EQ:Ochiijmatrixelement}
930: \frac{\langle Q \lvert \hat{O} \chi^{0\dagger}_{ij}\chi^0_{ij} \rvert P \rangle}
931: {\langle Q | P \rangle}
932: = \frac{1}{4}(1+\epsilon_{ij})
933: \frac{\langle Q \lvert \hat{O} \rvert P \rangle}
934: {\langle Q | P \rangle}\\
935: - \epsilon_{ij} \frac{\langle Q \lvert \hat{O} \rvert P' \rangle}
936: {\langle Q | P' \rangle} \frac{1}{4}\bigl( 1 + 3\delta^{\alpha_i \alpha_j}\bigr).
937: \end{multline}
938: 
939: \section{\label{SEC:Correlation} Spin Correlation Functions}
940: 
941: So long as the ground state of the spin system is a global singlet, its wavefunction can be written
942: as a linear superposition of valence bond states: $\lvert \psi \rangle = \sum_P c_P\lvert P \rangle$.
943: Accordingly, operator expectation values have the form
944: \begin{equation} \label{EQ:psiOpsi}
945: \langle \hat{O} \rangle = \frac{\langle \psi \lvert \hat{O} \rvert \psi \rangle}{\langle \psi | \psi \rangle}
946: = \frac{\sum_{P,Q} W(P,Q)
947: \frac{\langle Q \lvert \hat{O} \rvert P \rangle}{\langle Q | P \rangle}
948:  }{\sum_{P,Q} W(P,Q) },
949: \end{equation}
950: with $P,Q \in \mathcal{S}_N$ and $\hat{O}$ some operator of interest.
951: The quantity $W(P,Q) = c_Q c_P \langle Q | P \rangle$ may be determined 
952: variationally~\cite{Lou06} or it may arise as a sampling weight in the context
953: of a Monte Carlo projection scheme.~\cite{Sandvik05}
954: The matrix element  $\langle Q | \hat{O} | P \rangle/\langle Q | P \rangle$
955: is related to the properties of the loops that are formed when the singlet tilings of 
956: the two states are superimposed. In fact, since its value depends only on the properties
957: of the loops, the individual valence bond states can be abstracted away entirely,
958: leaving what is essentially a loop estimator for $\hat{O}$; we denote this function $O_{\mathcal{L}}$.
959: In this way of thinking, Eq.~\eqref{EQ:psiOpsi} can best be understood as 
960: $\langle \hat{O} \rangle = \langle O_{\mathcal{L}} \rangle_W$, 
961: an ensemble average of $O_{\mathcal{L}}$ in the gas of fluctuating 
962: loops.\cite{Sutherland88b, Kohmoto88b}
963: 
964: In this section, we want to determine the loop estimators for the spin operators $\mathbf{S}_i\cdot\mathbf{S}_j$
965: and $(\mathbf{S}_i\cdot\mathbf{S}_j)(\mathbf{S}_k\cdot\mathbf{S}_l)$.
966: All the work to compute the second order result has already been done.
967: Since $\mathbf{S}_i\cdot\mathbf{S}_j = \frac{1}{4} - \chi^{0\dagger}_{ij}\chi^0_{ij}$,
968: comparison with Eq.~\eqref{EQ:chidaggerchi} immediately yields
969: \begin{equation} \label{EQ:SidotSj}
970: \bigl(\mathbf{S}_i \cdot\mathbf{S}_j\bigr)_{\mathcal{L}}
971: = \frac{3}{4}\epsilon_{ij} \delta^{\alpha_i\alpha_j}.
972: \end{equation}
973: For the fourth order result, we begin by specializing Eq.~\eqref{EQ:Ochiijmatrixelement}
974: to the case $\hat{O} = \chi^{0\dagger}_{kl}\chi^{0}_{kl}$, which gives
975: \begin{multline}
976: \frac{\langle Q \lvert \chi^{0\dagger}_{ij}\chi^{0}_{ij}\chi^{0\dagger}_{kl}\chi^{0}_{kl} \rvert P\rangle}
977: {\langle Q | P \rangle}
978: = \frac{1}{16}\bigl( 1 - \epsilon_{ij} 3\delta^{\alpha'_i\alpha'_j}\bigr)\bigl( 1 - \epsilon_{kl} 3\delta^{\alpha_k\alpha_l}\bigr)\\
979: + \frac{3}{16}(1+\epsilon_{ij})\epsilon_{kl} \bigl( \delta^{\alpha'_i\alpha'_j} - \delta^{\alpha_i\alpha_j} \bigr).
980: \end{multline}
981: Here, $\alpha$ labels the loops in $\bar{Q}P$ and $\alpha'$ labels the
982: loops in $\bar{Q}P'$ where $P'$ is the modified configuration after
983: $\chi^{0\dagger}_{kl}\chi^{0}_{kl}$ has acted on $\lvert P \rangle$.
984: Making use of
985: \begin{multline} \label{EQ:chi4toS4}
986: \chi^{0\dagger}_{ij}\chi^{0}_{ij}\chi^{0\dagger}_{kl}\chi^{0}_{kl}
987: = \frac{1}{16} - \frac{1}{4}\mathbf{S}_i\cdot\mathbf{S}_j
988: - \frac{1}{4}\mathbf{S}_k\cdot\mathbf{S}_l\\
989: + (\mathbf{S}_i\cdot\mathbf{S}_j)(\mathbf{S}_k\cdot\mathbf{S}_l)
990: \end{multline}
991: and Eq.~\eqref{EQ:SidotSj}, we arrive at
992: \begin{multline} \label{EQ:SidotSjSkdotSlalphaprime}
993: \bigl[ \bigl(\mathbf{S}_i\cdot\mathbf{S}_j\bigr)\bigl(\mathbf{S}_k\cdot\mathbf{S}_l\bigr)\bigr]_{\mathcal{L}}\\
994: = \epsilon_{ij}\epsilon_{kl} \biggl[\frac{3}{16}\bigl( \delta^{\alpha'_i\alpha'_j} - \delta^{\alpha_i\alpha_j} \bigr)
995: + \frac{9}{16} \delta^{\alpha'_i\alpha'_j} \delta^{\alpha_k\alpha_l} \biggr].
996: \end{multline}
997: There is no contribution when $\delta^{\alpha_i'\alpha_j'} = \delta^{\alpha_i\alpha_j}=0$.
998: Hence, the estimator is nonzero only if all four site indices belong 
999: to the same loop or if there are two indices in each of two loops. The possible 
1000: configurations are shown in Fig.~\ref{FIG:Cijkl}.
1001: Note that if the vertices $(i,j)$ and $(k,l)$ reside in different loops 
1002: or if they are in the same loop but remain unlinked (\emph{i.e.}, there is a path
1003: along the loop from $i$ to $j$ that does not encounter $k$ or $l$) then
1004: $\delta^{\alpha_i'\alpha_j'} = \delta^{\alpha_i\alpha_j}$. 
1005: This equality holds because the operation on $(k,l)$
1006: does not disrupt the loop structure at sites $i$ and $j$.
1007: In this case, the bracketed term on the right-hand side is $\frac{9}{16} \delta^{\alpha_i \alpha_j }$.
1008: Otherwise, $\delta^{\alpha_i'\alpha_j'} = 1-\delta^{\alpha_i\alpha_j}$
1009:  and the bracketed term is $ \frac{3}{16} (1-2\delta^{\alpha_i \alpha_j })$.
1010: 
1011: \begin{figure}
1012: \includegraphics{Cijkl.eps}
1013: \caption{ \label{FIG:Cijkl}
1014: The expectation value of
1015: $\epsilon_{ij}\epsilon_{kl}\bigl(\mathbf{S}_i\cdot\mathbf{S}_j\bigr)\bigl(\mathbf{S}_k\cdot\mathbf{S}_l\bigr)$ is nonzero only if $(i,j)$ and $(k,l)$ connect one or two loops.
1016: For one loop, the value is either $9/16$ or $-3/16$ depending on whether
1017: $k$ and $l$ are in the (a) same or (b) different loop segments
1018: connecting $i$ and $j$. For two loops, the value is either $9/16$ or $3/16$
1019: depending on whether the interactions leave the loops (c) disjoint or (d) connected.
1020: }
1021: \end{figure}
1022: 
1023: The undesirable aspect of Eq.~\eqref{EQ:SidotSjSkdotSlalphaprime} is that it is history-dependent. 
1024: There are several cases to consider because $\alpha'$ references the loop configuration
1025: as it exists after application of the $(k,l)$ vertex. Ideally, we want to reexpress the estimator
1026: in a way that eliminates the $\alpha'$ labels. This can be accomplished---at the expense
1027: of introducing a new quantity $A$---as follows:
1028: \begin{multline} \label{EQ:SidotSjSkdotSl}
1029: \bigl[ \bigl(\mathbf{S}_i\cdot\mathbf{S}_j\bigr)\bigl(\mathbf{S}_k\cdot\mathbf{S}_l\bigr)\bigr]_{\mathcal{L}}\\
1030: \begin{split}
1031: = \epsilon_{ij}\epsilon_{kl} \biggl[&
1032:  -\frac{3}{8}\bigl(1+2A\bigr)
1033: \delta^{\alpha_i\alpha_j}\delta^{\alpha_j\alpha_k}\delta^{\alpha_k\alpha_l}\\
1034: &+ \frac{9}{16}\delta^{\alpha_i\alpha_j}\delta^{\alpha_k\alpha_l}\\
1035: &+ \frac{3}{16}\bigl(\delta^{\alpha_i\alpha_k}\delta^{\alpha_j\alpha_l}
1036: + \delta^{\alpha_i\alpha_l}\delta^{\alpha_j\alpha_k}\bigr)\biggr].
1037: \end{split}
1038: \end{multline}
1039: $A = 0,1$ is a topological term whose value is nonzero when
1040: traversal of the loop reveals an antisymmetric permulation of the 
1041: site labels $i,j,k,l$. It distinguishes the Fig.~\ref{FIG:Cijkl}(b) configuration ($A=1$) from 
1042: that of Fig.~\ref{FIG:Cijkl}(a) ($A=0)$. For spin correlations of order four and above, 
1043: the loop estimator is no longer simply a function of the loop labels $\alpha_i$.
1044: 
1045: We can make use of Eqs.~\eqref{EQ:SidotSj} and \eqref{EQ:SidotSjSkdotSl}
1046: to calculate powers of $\hat{\mathbf{M}} = \sum_{i\in A} \mathbf{S}_i - \sum_{j\in B} \mathbf{S}_j$,
1047: which on a bipartite lattice corresponds to the staggered magnetization.
1048: At second order, we have $\hat{\mathbf{M}}^2 = \sum_{ij}\epsilon_{ij} \mathbf{S}_i\cdot
1049: \mathbf{S}_j$ and hence, via
1050: Eq.~\eqref{EQ:SidotSj}, 
1051: \begin{equation} \label{EQ:M2Lalpha}
1052: \mathbf{M}^2_{\mathcal{L}} = \frac{\langle Q \lvert \hat{\mathbf{M}}^2 \rvert P \rangle}{\langle Q | P \rangle}
1053: = \frac{3}{4}\sum_{\alpha=1}^{N_{\circlearrowleft}} L_{\alpha}^2.
1054: \end{equation}
1055: This follows because 
1056: there are $L_{\alpha}^2$ ways to choose two sites from a loop of length
1057: $L_{\alpha}$.
1058: In the same way,
1059: $\hat{\mathbf{M}}^4
1060: = \sum_{ijkl} \epsilon_{ij}\epsilon_{kl}
1061: \bigl(\mathbf{S}_i\cdot\mathbf{S}_j\bigr)\bigl(\mathbf{S}_k\cdot\mathbf{S}_l\bigr)$
1062: can be computed using Eq.~\eqref{EQ:SidotSjSkdotSl}.
1063: The only complication is the case where $i,j,k,l$ are all in the same loop $\alpha$.
1064: Of the $L_\alpha^4$ such configurations, 
1065: \begin{equation} \label{EQ:Aeq1count}
1066: L_{\alpha}\sum_{j=1}^{L_\alpha} \bigl[ 2j(L_\alpha-j)-1 \bigr] = \frac{1}{3}L_\alpha^4 - \frac{4}{3}L_\alpha^2
1067: \end{equation}
1068: have weight $-3/16$. The counting reflects the fact that there are $L_\alpha$ ways to fix $i$ and, 
1069: as $j$ ranges over the loop, $2j(L_\alpha-j)-1$ ways to place $k$ and $l$ in opposite loop segments.
1070: (This includes mixed cases of the form $i=k$, $j\neq l$ which are $A=1$ in nature,
1071: but omits those of the form $i=k$, $j=l$, which are $A=0$.)
1072: The remainder have weight $9/16$.
1073: Thus, the total contribution is
1074: \begin{multline}
1075: \sum_\alpha \biggl[ \frac{9}{16}\biggl(\frac{2}{3}L_\alpha^4 + \frac{4}{3}L_\alpha^2
1076: \biggr) - \frac{3}{16}\biggl(\frac{1}{3}L_\alpha^4 - \frac{4}{3}L_\alpha^2\biggr) \biggr]\\
1077: + \biggl(2\times\frac{3}{16}+\frac{9}{16}\biggr)\sum_{\alpha\neq\beta} L_\alpha^2L_\beta^2,
1078: \end{multline}
1079: which simplifies to
1080: \begin{equation}
1081: \begin{split} \label{EQ:M4Lalpha}
1082: \mathbf{M}^4_{\mathcal{L}}
1083: &= \sum_\alpha \biggl( -\frac{5}{8}L_\alpha^4 + L_\alpha^2
1084: \biggr) + \frac{15}{16}\biggl(\sum_\alpha L_\alpha^2\biggr)^2.
1085: \end{split}
1086: \end{equation}
1087: 
1088: \section{\label{SEC:Cumulant} Cumulant generating function}
1089: 
1090: Calculating spin correlation functions as we did in the previous section
1091: involves keeping track of all possible ways that a given number of
1092: site index pairs, $(i,j),(k,l),\ldots,$ can be assigned to a set of valence bond loops. 
1093: For each such assignment, we then have to determine the contribution 
1094: to the correlation function based on how the loops are reconfigured
1095: by the valence bond operators $\chi^{0\dagger}_{ij} \chi^{0}_{ij}, \chi^{0\dagger}_{kl} \chi^{0}_{kl},
1096: \ldots.$
1097: This brute force approach 
1098: becomes increasingly 
1099: cumbersome (and errorprone)
1100: as the order of the correlation 
1101: function becomes large. At second and fourth order, the problem is manageable.
1102: For $\mathbf{S}_i \cdot \mathbf{S}_j$, there are are two distinct configurations---one 
1103: that contributes (when $i$ and $j$ are in the same loop) and one that
1104: does not (when $i$ and $j$ are in different loops)---and for
1105: $(\mathbf{S}_i \cdot \mathbf{S}_j)(\mathbf{S}_j \cdot \mathbf{S}_l)$,
1106: there are eight configurations---four that contribute (those shown in Fig.~\ref{FIG:Cijkl}) 
1107: and four that do not. But at sixth order, there are already 33 configurations to consider---an 
1108: intractable nightmare.
1109: 
1110: In this section, we present a vastly more simple approach that emerges from a deeper 
1111: understanding of the relationship between correlation functions and loops.
1112: To start, we introduce a new operator
1113: \begin{equation}
1114: \hat{\gamma}_{ij} = \frac{1}{4} + \epsilon_{ij}\mathbf{S}_i\cdot\mathbf{S}_j
1115: = \frac{1}{4}\bigl(1+\epsilon_{ij}\bigr) - \epsilon_{ij} \chi^{0\dagger}_{ij}
1116: \chi^{0}_{ij},
1117: \end{equation}
1118: which is designed to have a positive definite matrix element, irrespective of the
1119: sublattice membership of $i$ and $j$. 
1120: What motivates this definition is a desire to compensate for the asymmetry between
1121: Eq.~\eqref{EQ:updateinjm} and Eqs.~\eqref{EQ:updateinim} and \eqref{EQ:updatejnjm},
1122: which describe the effect of $\chi^{0\dagger}_{ij}\chi^{0}_{ij}$ on a valence bond state.
1123: In the AB case ($\epsilon_{ij} = -1$), the operator $\hat{\gamma}_{ij} = \chi^{0\dagger}_{ij}\chi^{0}_{ij}$
1124: produces only a rearrangement of bonds and leaves the phase of the state 
1125: unchanged, which is the desired state of affairs;
1126: in the AA/BB case ($\epsilon_{ij} = 1$), the operator 
1127: $\hat{\gamma}_{ij} = \tfrac{1}{2}-\chi^{0\dagger}_{ij} \chi^{0}_{ij}$
1128: is engineered to behave in exactly the same way, by cancelling the first term
1129: and changing the sign of the second term on the right-hand side of 
1130: Eqs.~\eqref{EQ:updateinim} and \eqref{EQ:updatejnjm}.
1131: Pictorially, this amounts to removing the identity part 
1132: and changing the sign of the
1133: transposition part from $-\tfrac{1}{2}$ to $\tfrac{1}{2}$
1134: on the right-hand side of the third (bottom) update rule in Fig.~\ref{FIG:updates}.
1135: The rearrangement of individual singlets is governed by
1136: \begin{align}
1137: \hat{\gamma}_{ij}[i,j] &= \hat{\gamma}_{ii}[i,j] = [i,j] \\
1138: \hat{\gamma}_{ij}[i,l][k,j] &= \hat{\gamma}_{ik}[i,l][k,j] = \frac{1}{2}[i,j][k,l]
1139: \end{align}
1140: for sites $i,k \in A$ and $j,l \in B$. Hence, 
1141: \begin{align} \label{EQ:gammainjnP}
1142: \hat{\gamma}_{i_nj_m} \lvert P \rangle &= 2^{\delta_{Pn,m}-1}\lvert (Pn\ m) P \rangle,\\
1143: \hat{\gamma}_{i_ni_m} \lvert P \rangle &= 2^{\delta_{n,m}-1}\lvert (Pn\ Pm) P \rangle,\\
1144: \hat{\gamma}_{j_nj_m} \lvert P \rangle &= 2^{\delta_{n,m}-1}\lvert (n\ m) P \rangle.
1145: \end{align}
1146: 
1147: \begin{figure}
1148: \begin{center}
1149: \includegraphics{Cijklmn.eps}
1150: \end{center}
1151: \caption{ \label{FIG:Cijklmn}
1152: Two examples of the correlation function defined in Eq.~\eqref{EQ:correlation-gamma}
1153: acting at third order.}
1154: \end{figure}
1155: 
1156: Since $\hat{\gamma}_{ij}$ involves only the transposition of bond pairs,
1157: its effect on loops is limited to splitting one loop into two (when $i$ and $j$ belong
1158: to the same loop) and joining two loops into one (when $i$ and $j$ belong to different loops).
1159: Its correlation function obeys the simple rule
1160: \begin{equation} \label{EQ:correlation-gamma}
1161: C_{\underbrace{ijkl\cdots}_{\text{$2n$ indices} }} 
1162: = \frac{ \langle Q \lvert \hat{\gamma}_{ij}\hat{\gamma}_{kl}\cdots \rvert P \rangle }
1163: { \langle Q | P \rangle }
1164: = 2^{\Delta N_{\circlearrowleft} - n} > 0,
1165: \end{equation}
1166: where $\Delta N_{\circlearrowleft}$ represents the change in loop number
1167: after the loops have been split or joined at ($i,j$), ($k,l$), \emph{etc}.
1168: Two examples are given in Fig.~\ref{FIG:Cijklmn}.
1169: 
1170: \begin{figure*}
1171: \begin{center}
1172: \includegraphics{correlations1.eps}
1173: \end{center}
1174: \caption{ \label{FIG:correlations1}
1175: The diagrammatic expansion of the cumulant generating function 
1176: $F$ [from Eq.~\eqref{EQ:Fgenerating}] is shown to third order in the coupling $a_{ij}$.
1177: The black lines denote valence bond loops and the red lines $\hat{\gamma}$ operators.
1178: }
1179: \end{figure*}
1180: 
1181: We now introduce a generating function
1182: \begin{equation} \label{EQ:Fgenerating}
1183: F[a] = \log \langle Q \lvert \exp \biggl( \sum_{ij} a_{ij} \hat{\gamma}_{ij} \biggr) \rvert P \rangle.
1184: \end{equation}
1185: Its diagramatic expansion in powers of the coupling $a_{ij}$, shown in Fig.~\ref{FIG:correlations1},
1186: is completely analogous to the Goldstone diagrams familiar from standard
1187: many-body theory. (There are only connected diagrams because of the linked-cluster theorem.)
1188: The $n^{\text{th}}$ order derivatives of $F$ are the cumulants of the correlators defined in
1189: Eq.~\eqref{EQ:correlation-gamma},
1190: \begin{equation}
1191: \tilde{C}_{\underbrace{ijk\cdots}_{\text{$2n$ indices} }} =  \frac{ \partial^n F}{\partial a_{ij}
1192: \partial a_{kl} \cdots }\biggr\rvert_{a=0},
1193: \end{equation}
1194: and are related to them by
1195: \begin{align}
1196: \tilde{C}_{ij} &= C_{ij}, \\
1197: \tilde{C}_{ijkl} &= C_{ijkl} - \tilde{C}_{ij}\tilde{C}_{kl}, \\
1198: \begin{split}
1199: \tilde{C}_{ijklmn} &= C_{ijklmn} - \tilde{C}_{ij}\tilde{C}_{klmn} \\
1200:                                &\quad - \tilde{C}_{kl}\tilde{C}_{ijmn} - \tilde{C}_{mn}\tilde{C}_{ijkl}\\
1201:                                 &\quad - \tilde{C}_{ij}\tilde{C}_{kl}\tilde{C}_{mn}.
1202: \end{split}
1203: \end{align}
1204: What is important about the cumulants is that their subtractive terms lead to perfect
1205: cancellation (\emph{e.g.}, $C_{ijkl} = C_{ij}C_{kl}$ so that $\tilde{C}_{ijkl} = 0$) 
1206: whenever the $\hat{\gamma}$-vertices connect the loops into a network that 
1207: cannot be disentangled by cutting it along two loop segments.
1208: The exact statement is that a cumulant vanishes unless its
1209: configuration is irreducible in the standard many-body diagram sense.
1210: See Fig.~\ref{FIG:correlations2}.
1211: 
1212: \begin{figure}
1213: \begin{center}
1214: \includegraphics{correlations2.eps}
1215: \end{center}
1216: \caption{ \label{FIG:correlations2}
1217: At each order, the contributions to $F$ can be classified according to their irreducibility.
1218: Nonirreducible diagrams are those that can be stitched together from self-energy parts of
1219: lower order in $\gamma$. The self-energy parts, shown as shaded arcs, are built
1220: by removing a single loop segment from the diagrams in Fig.~\ref{FIG:correlations1}. 
1221: The values listed next to each diagram represent the $\tilde{C}$ cumulant value
1222: associated with that configuration. $\tilde{C} = 0$ for the composite terms.
1223: }
1224: \end{figure}
1225: 
1226: The cumulants, in turn, are related to the physical spin correlation functions by
1227: \begin{align} \label{EQ:CijtoS}
1228: \tilde{C}_{ij} &= S_{ijkl}+\tfrac{1}{4} \\ \label{EQ:CijkltoS}
1229: \tilde{C}_{ijkl} &= S_{ijkl} - S_{ij}S_{kl} \\ \label{EQ:CijklmntoS}
1230: \begin{split}
1231: \tilde{C}_{ijklmn} &= S_{ijklmn} - S_{ij}S_{klmn} \\
1232:                                &\quad - S_{kl}S_{ijmn} - S_{mn}S_{ijkl}\\
1233:                                 &\quad + 2S_{ij}S_{kl}S_{mn}.
1234: \end{split}
1235: \end{align}
1236: Here, we have used the shorthand 
1237: \begin{equation} \label{EQ:Sshorthand}
1238: S_{ijkl\cdots} = (\epsilon_{ij}\epsilon_{kl}\cdots)\frac{\langle Q \lvert (\mathbf{S}_i\cdot\mathbf{S}_j)
1239:  (\mathbf{S}_k\cdot\mathbf{S}_l) \cdots
1240: \rvert P \rangle}{\langle Q | P \rangle}.
1241: \end{equation}
1242: Thus, computing spin correlation functions at any order is just a matter of determining
1243: the relevant irreducible diagrams, computing the corresponding cumulants, and 
1244: then solving for the desired term in Eqs.~\eqref{EQ:CijtoS}--\eqref{EQ:CijklmntoS}.
1245: By this means, we recover the loop estimators Eq.~\eqref{EQ:SidotSj}
1246: and Eq.~\eqref{EQ:SidotSjSkdotSl}.
1247: 
1248: It is also possible to compute even powers of the staggered magnetization by
1249: summing over all the site indices:
1250: \begin{align}
1251: \sum_{ij} \tilde{C}_{ij} &= \mathbf{M}^2_{\mathcal{L}} + N^2,\\
1252: \sum_{ijkl} \tilde{C}_{ijkl} &= \mathbf{M}^4_{\mathcal{L}} - \bigl( \mathbf{M}^2_{\mathcal{L}} \bigr)^2, \\
1253: \sum_{ijklmn}\!\! \tilde{C}_{ijklmn} &= \mathbf{M}^6_{\mathcal{L}}
1254: - 3 \mathbf{M}^2_{\mathcal{L}} \mathbf{M}^4_{\mathcal{L}} + 2\bigl( \mathbf{M}^2_{\mathcal{L}} \bigr)^3.
1255: \end{align}
1256: At second order in $\hat{\mathbf{M}}$ (first order in $\hat{\gamma}_{ij}$), there
1257: are two irreducible diagrams: a one-loop configuration of weight 1 and a two-loop
1258: configuration of weight 1/4. Since there are $L_{\alpha}^2$ ways to select two sites in
1259: loop $\alpha$ and $L_{\alpha}L_{\beta}$ ways to select one site in loop $\alpha$ 
1260: and one in loop $\beta$, we have
1261: \begin{equation}
1262: \begin{split}
1263: \mathbf{M}^2_{\mathcal{L}} &= 1\sum_\alpha L_\alpha^2 + \frac{1}{4} \sum_{\alpha\neq\beta} L_\alpha L_\beta - N^2 \\
1264: &= \frac{3}{4}\sum_{\alpha} L_\alpha^2 + \frac{1}{4} \biggl( \sum_\alpha L_\alpha \biggr)^2
1265: - N^2 \\
1266: &= \frac{3}{4}\sum_{\alpha} L_\alpha^2.
1267: \end{split}
1268: \end{equation}
1269: A glance at Eq.~\eqref{EQ:M2Lalpha} confirms that this method produces the correct result.
1270: 
1271: At fourth order in $\hat{\mathbf{M}}$ (second order in $\hat{\gamma}_{ij}$), there are again
1272: only two irreducible diagrams: a one-loop (cross) configuration of weight $-3/4$ and a two-loop
1273: configuration of weight $3/16$. We have already worked out the counting 
1274: in Eq.~\eqref{EQ:Aeq1count}. Hence,
1275: \begin{equation}
1276: \begin{split}
1277: \mathbf{M}^4_{\mathcal{L}} &= -\frac{3}{4}\sum_\alpha \Bigl(\frac{1}{3}L_\alpha^4 - \frac{4}{3}L^2_\alpha\Bigr) + \frac{3}{16} \sum_{\alpha\neq\beta} 2L_\alpha^2 L_\beta^2 
1278: +\bigl(\mathbf{M}^2_{\mathcal{L}}\bigr)^2 \\
1279: &= \sum_{\alpha} \Bigl(-\frac{5}{8}L_\alpha^4 + L_{\alpha}^2\Bigr) + \frac{3}{8} \biggl( \sum_\alpha L_\alpha^2 \biggr)^2 +  \biggl(\frac{3}{4} \sum_\alpha L_\alpha^2 \biggr)^2 \\
1280: &= \sum_{\alpha} \Bigl(-\frac{5}{8}L_\alpha^4 + L_{\alpha}^2\Bigr) + \frac{15}{16} \biggl( \sum_\alpha L_\alpha^2 \biggr)^2,
1281: \end{split}
1282: \end{equation}
1283: which is the result of Eq.~\eqref{EQ:M4Lalpha}. 
1284: 
1285: At sixth order in $\hat{\mathbf{M}}$ (third order in $\hat{\gamma}_{ij}$), there are five
1286: irreducible diagrams. The diagrams of weight $3/32$ and $-3/32$ both number
1287: $\sum_{\alpha\neq\beta}2L_{\alpha}^3L_{\beta}^3$ and thus cancel each other,
1288: so we only need to consider the other three: a one-loop (asterisk) configuration
1289: of weight $3/2$, a one-loop (railway tie) configuration of weight $3/4$, and
1290: a two-loop configuration of weight $-3/16$. Counting the number of site index 
1291: arrangements for these three cases requires some work, but it is no different in principle from 
1292: what we did in Eq.~\eqref{EQ:Aeq1count}. We skip the details and simply
1293: report the result:
1294: \begin{equation}
1295: \begin{split}
1296: \mathbf{M}^6_{\mathcal{L}}
1297: &= \frac{3}{2}\sum_{\alpha} \Bigl( \frac{1}{15}L_{\alpha}^6 - \frac{16}{15}L_{\alpha}^2\Bigr)\\
1298: &\quad+ \frac{3}{4}\sum_{\alpha} \Bigl( \frac{1}{5}L_{\alpha}^6 - \frac{4}{3}L_{\alpha}^4 
1299: + \frac{32}{15}L_{\alpha}^2\Bigr)\\
1300: &\quad- \frac{3}{16} 2\sum_{\alpha\neq\beta} L_{\alpha}^2L_{\beta}^2\bigl(
1301: L_{\alpha}^2+L_{\beta}^2-4\bigr)\\
1302: &\quad+ 3\mathbf{M}^2_{\mathcal{L}}\mathbf{M}^4_{\mathcal{L}} - 2\bigl(\mathbf{M}^2_{\mathcal{L}}
1303: \bigr)^3.
1304: \end{split}
1305: \end{equation}
1306: This simplifies to
1307: \begin{equation}
1308: \begin{split}
1309: \mathbf{M}^6_{\mathcal{L}}
1310: &= \sum_{\alpha} \Bigl(L_{\alpha}^6 - \frac{5}{2}L_{\alpha}^4\Bigr)
1311: +\frac{81}{64}\biggl( \sum_\alpha L_\alpha^2 \biggr)^3\\
1312: &\quad -\frac{69}{32}\biggl( \sum_\alpha L_\alpha^4 \biggr)\biggl( \sum_\alpha L_\alpha^2 \biggr).
1313: \end{split}
1314: \end{equation}
1315: 
1316: In terms of the $\bar{Q}P$ cycle lengths ($L_\alpha \to 2k_\alpha$), the 
1317: magnetization formulas are
1318: \begin{align} \label{EQ:M2kalpha}
1319: \mathbf{M}^2_{\mathcal{L}} &= 3\sum_{\alpha} k^2_{\alpha}, \\ \label{EQ:M4kalpha}
1320: \mathbf{M}^4_{\mathcal{L}} &= 
1321: \sum_{\alpha} \bigl(-10k^4_{\alpha}+4k^2_{\alpha}\bigr)
1322: +15 \biggl(\sum_{\alpha} k^2_{\alpha}\biggr)^2, \\
1323: \begin{split}
1324: \mathbf{M}^6_{\mathcal{L}}  &= \sum_{\alpha} \bigl(64k^6_{\alpha}-40k^4_{\alpha}\bigr)
1325: + 81\biggl(\sum_{\alpha} k^2_{\alpha}\biggr)^3\\
1326: &\quad- 138 \biggl(\sum_{\alpha} k^4_{\alpha}\biggr)\biggl(\sum_{\alpha} k^2_{\alpha}\biggr).
1327: \end{split}
1328: \end{align}
1329: Recall that in Sect.~\ref{SEC:Overlaps} we used 
1330: $n_k = \sum_{\alpha=1}^{N_{\circlearrowleft}}\delta(k-k_{\alpha})$,
1331: the cycle length distribution function, to derive an expression for the overlap $\langle Q | P \rangle$.
1332: It is easy to see that the second moment of $n_k$ is related to the presence
1333: of long-range antiferromagnetic order in the system:
1334: a loop average of Eq.~\eqref{EQ:M2kalpha} gives
1335: $\langle \hat{\mathbf{M}}^2\rangle = \sum_k 3k^2 \langle n_k \rangle_W$.
1336: Similarly, the fourth and sixth powers of the staggered magnetization have terms
1337: involving the cycle-length correlations $\langle n_kn_{k'} \rangle_W$ and
1338: $\langle n_kn_{k'}n_{k''} \rangle_W$.
1339: 
1340: The key feature is the tail of the cycle-length distribution. If it is too weak, then the
1341: loop gas is characterized by a macroscopic number of small loops.
1342: In the opposite limit, the system has a vanishing number density of 
1343: loops, although each loop contains a nonvanishing fraction of all the spins.
1344: These system-spanning loops are the foundation of the long-range order.
1345: As an example, suppose that $\langle n_k \rangle_W \sim k^{-p}$, with the 
1346: normalization set by the constraint $N = \sum_{k=1}^N k\langle n_k \rangle_W$.
1347: The standard large-$N$ summation rules are
1348: \begin{equation}
1349: \sum_{k=1}^N k^{-p} = \begin{cases}
1350: \frac{1}{1-p}N^{1-p} + \mathcal{O}(N^{-p}) & \text{if $p<1$} \\
1351: \log N + \mathcal{O}(1) & \text{if $p=1$}\\
1352: \zeta(p) + \mathcal{O}(N^{1-p}) & \text{if $p > 1$},
1353: \end{cases}
1354: \end{equation}
1355: where $\zeta(p)$ is the Reimann Zeta function.
1356: These can be used to derive the asymptotic behaviour listed in Table~\ref{TAB:thermlimit}.
1357: Figure~\ref{FIG:lrorder} depicts the $N\to\infty$ behaviour of the
1358: loop number density and the average staggered moment:
1359: \begin{align}
1360: \frac{\langle N_{\circlearrowleft} \rangle}{N} &= \frac{1}{N} \sum_{k=1}^N \langle n_k \rangle_W 
1361: \to \frac{\zeta(p)}{\zeta(p-1)}\theta(p-2), \\
1362: \frac{\langle \hat{\mathbf{M}}^2 \rangle}{(2N)^2} &= \frac{3}{4N^2}\sum_{k=1}^N k^2 \langle n_k \rangle_W
1363: \to \frac{3(2-p)}{4(3-p)}\theta(2-p).
1364: \end{align}
1365: Clearly, long-range order exists for a range of exponents $0 \le p < 2$.
1366: The phase transition at the critical value $p_{\text{c}}=2$ should be visible as a sharp step
1367: in the function $\langle \hat{\mathbf{M}}^2 \rangle/\langle N_{\circlearrowleft}\rangle^2$
1368: and in the Binder ratios~\cite{Binder81} $\langle \hat{\mathbf{M}}^4 \rangle/\langle \hat{\mathbf{M}}^2\rangle^2$
1369: and $\langle \hat{\mathbf{M}}^6 \rangle/\langle \hat{\mathbf{M}}^2\rangle^3$.
1370: 
1371: \begin{table}
1372:  \caption{\label{TAB:thermlimit} Asymptotic behaviour of the loop density and staggered magnetization 
1373:  in the thermodynamic limit as a function of the loop distribution tail.}
1374: \begin{ruledtabular}
1375: \begin{tabular}{@{\quad}c|c|c@{\quad}}
1376: $\langle n_k \rangle_W \sim k^{-p}$ & $\mathcal{O}(\langle N_{\circlearrowleft} \rangle/N)$ & 
1377: $\mathcal{O}(\langle \hat{\mathbf{M}}^2 \rangle/N^2)$  \tabularnewline \hline
1378: $0 \le p < 1$ & $\frac{1}{N}$ & $1$ \tabularnewline
1379: $p = 1$ & $\frac{\log N}{N}$ & $1$ \tabularnewline
1380: $1 < p < 2$ & $\frac{1}{N^{2-p}}$ & $1$ \tabularnewline
1381: $p = 2$ & $\frac{1}{\log N}$ & $\frac{1}{\log N}$ \tabularnewline
1382: $2 < p < 3$ & $1$ & $\frac{1}{N^{p-2}}$ \tabularnewline
1383: $p \ge 3$ & $1$ & $\frac{1}{N}$ \tabularnewline
1384: short-ranged & $1$ & $\frac{1}{N}$
1385: \end{tabular}
1386: \end{ruledtabular}
1387: \end{table}
1388: 
1389: \begin{figure}
1390: \begin{center}
1391: \includegraphics{lrorder.eps}
1392: \end{center}
1393: \caption{ \label{FIG:lrorder}
1394: The squared staggered magnetization and the loop density are plotted 
1395: as a function of the exponent $p$
1396: for a cycle-length distribution of the form $\langle n_k \rangle_W \sim k^{-p}$.
1397: The two quantities are complementary indicators of the long range order below $p=2$.
1398:  }
1399: \end{figure}
1400: 
1401: In the $p\to \infty$ limit, the cycle-length distribution becomes sharply peaked 
1402: at the minimum length, $\langle n_k \rangle_W = N\delta_{k,1}$, which is the result 
1403: for a fixed dimer configuration.
1404: More generally, a generic VBS state will have $\langle n_k \rangle_W 
1405: \sim \theta(N_0-k)$, where $N_0$ is the size of the plaquet (the basic unit 
1406: of translational symmetry breaking). When the distribution is short-ranged---having
1407: either an upper cutoff or an exponentially suppressed tail---there is no magnetic order.
1408: 
1409: \section{\label{SEC:Extended} Valence bond coverings of the full Hilbert space}
1410: 
1411: In previous sections, we discussed the properties of the $S=0$ valence bond states.
1412: We now turn our attention to the full Hilbert space and consider 
1413: \emph{extended} valence bond states of the form
1414: \begin{equation} 
1415: \lvert P; \vec{\mu} \rangle = \prod_{n=1}^N\chi^{\mu_n\dagger}_{i_n,Pj_n}\lvert \text{vac} \rangle.
1416: \end{equation}
1417: Here, $\vec{\mu} = (\mu_1,\mu_2, \ldots, \mu_N) \in \mathbb{Z}_4^N$
1418: is a vector of indices describing the singlet/triplet character of each bond.
1419: The set of such states is overcomplete with respect to the full Hilbert space
1420: ($4^NN! \gg 2^{2N}$).
1421: The restriction to AB bonds is still meaningful in the extended basis 
1422: since for every index pair $\mu'',\nu''$ there is at least one set of indices $\mu,\nu,\mu',\nu'$
1423: (in fact there are exactly four) such that the equation
1424: \begin{equation}
1425: c_1 \chi^{\mu\dagger}_{il}\chi^{\nu\dagger}_{jk}
1426: + c_2 \chi^{\mu'\dagger}_{ij}\chi^{\nu'\dagger}_{kl}
1427: + c_3 \chi^{\mu''\dagger}_{lj}\chi^{\nu''\dagger}_{ik} = 0
1428: \end{equation}
1429: has a nontrivial solution. This is the singlet/triplet generalization of Eq.~\eqref{EQ:linearly-dependent}.
1430: 
1431: To understand how triplet states evolve under the Heisenberg hamiltonian, 
1432: we must reproduce the analysis of Sec.~\ref{SEC:Evolution} but now with
1433: Eqs.~\eqref{EQ:chi2-commutator} and \eqref{EQ:chi3-commutator}
1434: specialized to 
1435: \begin{equation} 
1436: [\chi^{0}_{ij},\chi^{\mu\dagger}_{ij}]\lvert \text{vac} \rangle = \delta^{0\mu}\lvert \text{vac} \rangle
1437: \end{equation}
1438: and
1439: \begin{equation} 
1440: [\chi^{0}_{ij},\chi^{\mu\dagger}_{kj}\chi^{\nu\dagger}_{il}]\lvert \text{vac} \rangle = \frac{1}{2}\,T^{\lambda\mu 0\nu}\,\chi^{\lambda\dagger}_{kl}
1441: \lvert \text{vac} \rangle.
1442: \end{equation}
1443: As before, the operator $-\hat{H}_{ij}$ has both a diagonal and offdiagonal part.
1444: If there is a pre-existing singlet bond across the active sites, the bond configuration
1445: is left unchanged. If there is a pre-existing triplet bond, however, the state is annihilated.
1446: Otherwise, the bonds are rearranged and the singlet/triplet labels reassigned
1447: in a spin-conserving fashion. The update rule for $-\hat{H}_{ij}$ with $i\in A$ and $j\in B$,
1448: illustrated in Fig.~\eqref{FIG:updates2}, is given by
1449: \begin{widetext}
1450: \begin{equation}
1451: -\hat{H}_{i_nj_m} \lvert P;\vec{\mu} \rangle = \begin{cases}
1452: \delta^{\mu_n 0} \lvert P; \vec{\mu} \rangle & \text{if $Pm = n$}\\
1453: \frac{1}{2}T^{\lambda \mu_n 0\mu_{\bar{P}m} }\lvert (Pn\ m) P; (\mu_1,\ldots,
1454: \mu_{n-1},0,\mu_{n+1},\ldots,\mu_{\bar{P}m-1},
1455: \lambda,\mu_{\bar{P}m+1},\ldots,\mu_N)\rangle & \text{otherwise}\\
1456: \end{cases}
1457: \end{equation}
1458: \end{widetext}
1459: 
1460: There are some important differences from the purely singlet-bond case.
1461: There is now the possibility of anihilating a state
1462: by acting directly on a triplet bond, 
1463: unlike in the singlet sector, where $(\hat{H}_{ij}\hat{H}_{i'j'} \hat{H}_{i''j''}\cdots) \lvert P \rangle \neq 0$
1464: always. Also, when bonds of different species interact,
1465: the weight $T^{\lambda\mu 0\nu}$ is potentially negative:
1466: the $\tau$ operators obey the identity
1467: \begin{equation}
1468: \tau^\mu\tau^{0\dagger}\tau^\nu = \begin{pmatrix}
1469: \tau^0 & \tau^1 & \tau^2 & \tau^3\\ 
1470: \tau^1 & \tau^0 & \tau^3 & -\tau^2\\
1471: \tau^2 & -\tau^3 & \tau^0 & \tau^1\\
1472: \tau^3 & \tau^2 & -\tau^1 & \tau^0
1473: \end{pmatrix}_{\mu\nu}
1474: \end{equation}
1475: and hence
1476: \begin{equation}
1477: T^{\lambda\mu 0 \nu} = \delta^{\lambda 0}\delta^{\mu\nu} +  \epsilon^{\mu\nu\lambda},
1478: \end{equation}
1479: where $\epsilon^{\mu\nu\lambda}$ is the alternating symbol.
1480: 
1481: \begin{figure}
1482: \begin{center}
1483: \includegraphics{updates2.eps}
1484: \end{center}
1485: \caption{ \label{FIG:updates2}
1486: Summary of the update rules for triplet bonds.
1487: The number of cross hatches on each bond indicates its triplet character.
1488:  }
1489: \end{figure}
1490: 
1491: Overlaps of extended valence bond states can be computed as before, only now we must use
1492: the most general form of the anticommutation relationship in Eq.~\eqref{EQ:chi3-commutator}.
1493: We find that
1494: \begin{equation} \label{EQ:QvuPmuoverlap}
1495: \langle Q; \vec{\nu} | P; \vec{\mu} \rangle = 
1496: 2^{N_{\circlearrowleft}-N}
1497: \prod_{\alpha=1}^{N_{\circlearrowleft}} K_{\alpha}.
1498: \end{equation}
1499: Here, $N_{\circlearrowleft}$ is the number of cycles in $\bar{Q}P$
1500: and $K_{\alpha}$ is a chirality factor whose value is given by the trace of
1501: the product of $\tau^{\nu_n\dagger}\tau^{\mu_n}$ in proper order around each cycle.
1502: The form of $K_{\alpha}$ follows from the
1503: ``boxcar'' property
1504: \begin{equation}
1505: \sum_\lambda\tr\bigl( \cdots \tau^{\nu\dagger}\tau^{\lambda}
1506: \bigr)T^{\lambda\mu\nu'\mu'} = \tr\bigl(\cdots\tau^{\nu\dagger}\tau^{\mu}
1507: \tau^{\nu'\dagger}\tau^{\mu'}\bigr).
1508: \end{equation}
1509: For example, the loop $\bar{Q}P = (1\ 2\ 4\ 5\ 3)\cdots$, 
1510: highlighted in Fig.~\ref{FIG:overlap},
1511: has associated with it a factor
1512: \begin{equation}
1513: K_{\alpha} = \frac{1}{2}\tr \bigl(\tau^{\nu_1\dagger}\tau^{\mu_1}
1514: \tau^{\nu_2\dagger}\tau^{\mu_2}
1515: \tau^{\nu_4\dagger}\tau^{\mu_4}
1516: \tau^{\nu_5\dagger}\tau^{\mu_5}
1517: \tau^{\nu_3\dagger}\tau^{\mu_3}\bigr).
1518: \end{equation}
1519: 
1520: Since $K_{\alpha}$ takes the values $0$ or $\pm 1$, the overlap is no longer
1521: guaranteed to be positive definite, as it was in the case of only singlet bonds. 
1522: If we do not mix triplet species, however, positive definiteness is preserved.
1523: In that case, $K_{\alpha} = 1$ if the loop contains an even number of triplets
1524: and $K_{\alpha} = 0$ if the loop contains an odd number. More generally,
1525: if we move around a given loop in proper order, encountering triplets
1526: of species $a,b,c,\ldots$ along the way, then $K_{\alpha}$ takes the value
1527: \begin{equation}
1528: \begin{split}
1529: 0 &\quad \text{one triplet} \\
1530: (-1)^{1+|\vec{\mu}|}\delta^{ab} &\quad \text{two triplets} \\
1531: (-1)^{1+|\vec{\mu}|}\epsilon^{abc} &\quad \text{three triplets} \\
1532: (-1)^{|\vec{\mu}|}(\delta^{ab}\delta^{cd} + \delta^{ad}\delta^{bc} - \delta^{ac}\delta^{bd})
1533: &\quad \text{four triplets}
1534: \end{split}
1535: \end{equation}
1536: where $|\vec{\mu}|$ counts the number of nonzero entries (triplets) 
1537: in $\vec{\mu}$ that belong to the loop in question.
1538: In other words, the overall sign depends on how many triplets arise
1539: from $\lvert P; \vec{\mu} \rangle$ and how many from $\lvert Q; \vec{\nu} \rangle$.
1540: 
1541: \section{\label{SEC:Components} Staggered Magnetization Components}
1542: 
1543: The staggered magnetization operator has the special property that it can always be 
1544: written so as to induce no rearrangement of bonds. This is simply a matter of grouping its
1545: terms to match the particular bond tiling of the state it is acting on.
1546: That is, for any $P \in \mathcal{S}_N$,
1547: \begin{equation} \label{EQ:Manorearrange}
1548: \hat{M}^a = \sum_{i\in A} S_i^a - \sum_{j \in B} S_j^a
1549: = \sum_{n=1}^N \bigl(S_{i_n}^a - S_{j_{Pn}}^a\bigr).
1550: \end{equation}
1551: Since the bracketed term behaves according to Eq.~\eqref{EQ:SiminusSj},
1552: we see that the effect of $\hat{M}^a$ on an $S=0$ valence bond state is to create
1553: a superposition of states, each of which has one bond promoted to an $a$-triplet:
1554: \begin{equation}
1555: \hat{M}^a \lvert P \rangle = i \sum_{n=1}^N \lvert P; a_n \rangle.
1556: \end{equation}
1557: Here, $\lvert P; a_n \rangle$ denotes the state $\lvert P; \vec{\mu} \rangle$
1558: with $\vec{\mu}$ having all zero entries except for $a$ in the $n^{\text{th}}$ slot.
1559: Since lone triplets always produce zero-contribution loops, we find that
1560: \begin{equation}
1561: M^a_{\mathcal{L}} = \frac{\langle Q \rvert \hat{M}^a \lvert P \rangle}{\langle Q | P \rangle}
1562:  =  i \sum_{n=1}^N \frac{\langle Q | P; a_n \rangle}{\langle Q | P \rangle} = 0,
1563: \end{equation}
1564: which is the expected result for rotationally invariant states $\lvert P \rangle$
1565: and $\lvert Q \rangle$.
1566: Two triplets of the same species, however, if they originate in different states,
1567: give unit weight whenever they lie in the same
1568: loop. Thus,
1569: \begin{equation} \label{EQ:MaMb}
1570: (M^aM^b)_{\mathcal{L}}
1571:  = \sum_{m=1}^N \sum_{n=1}^N \frac{\langle Q; a_m | P; b_n \rangle}{\langle Q | P \rangle}
1572:  = \delta^{ab}\sum_{\alpha} k_{\alpha}^2.
1573: \end{equation}
1574: The final equality holds because, in a cycle of length $k_\alpha$, there are $k_\alpha$ ways 
1575: to place the $m^{\text{th}}$ $Q$ link and $k_\alpha$ ways to place the $n^{\text{th}}$ $P$ link.
1576: 
1577: Correlation functions at fourth order can be computed in much the same way.
1578: Two applications of $\hat{M}^a$ to a valence bond state yields
1579: \begin{equation}
1580: (\hat{M}^a)^2\lvert P \rangle 
1581: = N\lvert P \rangle - \sum_{m\neq n} \lvert P; a_m,a_n \rangle,
1582: \end{equation}
1583: and the overlap of two such states is
1584: \begin{widetext}
1585: \begin{equation} \label{EQ:Ma2Mb2uneval}
1586: \frac{\langle Q \lvert (\hat{M}^a)^2(\hat{M}^b)^2 \rvert P \rangle}{\langle Q | P \rangle}
1587: = N^2 - \sum_{m\neq n}\biggl( \frac{\langle Q; a_m,a_n | P \rangle}{\langle Q | P \rangle}
1588: + \frac{\langle Q | P; b_m,b_n \rangle}{\langle Q | P \rangle}\biggr)
1589:  + \sum_{m\neq n}\sum_{m'\neq n'} 
1590: \frac{\langle Q; a_m,a_n | P; b_{m'},b_{n'} \rangle}{\langle Q | P \rangle}.
1591: \end{equation}
1592: \end{widetext}
1593: 
1594: The bracketed terms in Eq.~\eqref{EQ:Ma2Mb2uneval} each take the
1595: value $-1$ if the two triplets are in the same loop and vanish otherwise.
1596: In the case of one triplet species ($a=b$), the final term contributes if there is either
1597: a single loop containing all four triplets or a pair of loops containing 
1598: two triplets each. These configurations all have weight +1.
1599: Applying the appropriate counting arguments yields
1600: \begin{multline} \label{EQ:Ma4count}
1601: (M^a)^4_{\mathcal{L}} = N^2 + 2N\sum_{\alpha}k_\alpha(k_\alpha-1)
1602: + \sum_\alpha k_\alpha^2(k_\alpha-1)^2\\
1603: + \sum_{\alpha\neq\beta}k_\alpha(k_\alpha-1)k_\beta(k_\beta-1)
1604: + 2\sum_{\alpha\neq\beta}k_\alpha^2k_\beta^2.
1605: \end{multline}
1606: Note that the $m\neq n$ and $m' \neq n'$ constraints have no effect when
1607: $m$ and $m'$ ($m$ and $n'$) are in one cycle and $n$ and $n'$ ($n$ and $m'$) 
1608: are in another. Equation~\eqref{EQ:Ma4count}
1609: simplifies considerably to give
1610: \begin{equation} \label{EQ:Ma4}
1611: (M^a)^4_{\mathcal{L}}  = -2\sum_\alpha k_\alpha^4 + 3\biggl(\sum_{\alpha} k_\alpha^2\biggr)^2.
1612: \end{equation}
1613: 
1614: If there are two triplet species $(a\neq b)$, the situation is slightly more complicated.
1615: When all four triplets are in the same loop, there are 
1616: \begin{equation} \label{EQ:4tripletsw1}
1617: 2k_\alpha \sum_{n=3}^{k_\alpha} (n-1)(n-2) = \frac{2}{3}k_\alpha^4 - 2k_\alpha^3 + \frac{4}{3}k_\alpha^2 
1618: \end{equation}
1619: configurations, having weight $+1$, in which the $a$ and $b$ triplets
1620: appear in consecutive order around the loop ($aabb$) and
1621: \begin{equation} \label{EQ:4tripletswm1}
1622: 2k_\alpha \sum_{n=2}^{k_\alpha} (n-1)(k-n+1) = -\frac{1}{3}k_\alpha^2 + \frac{1}{3}k_\alpha^4
1623: \end{equation}
1624: configurations, having weight $-1$, in which they appear in alternating order ($abab$).
1625: [Note that Eqs.~\eqref{EQ:4tripletsw1} and \eqref{EQ:4tripletswm1} sum to 
1626: $k_\alpha^2(k_\alpha-1)^2$, as required.]
1627: There is also a positive contribution when the $a$ and $b$ triplets are paired 
1628: $(aa)(bb)$ in two different loops, but none when they are paired $(ab)(ab)$.
1629: Hence,
1630: \begin{multline} \label{Ma2Mb2count}
1631: \bigl[ (M^a)^2(M^b)^2\bigr]_{\mathcal{L}} = N^2 + 2N\sum_{\alpha}k_\alpha(k_\alpha-1)\\
1632: + \sum_\alpha \Bigl(\frac{2}{3}k_\alpha^4 - 2k_\alpha^3 + \frac{4}{3}k_\alpha^2  \Bigr)
1633: - \sum_\alpha \Bigl(\frac{1}{3}k_\alpha^4-\frac{1}{3}k_\alpha^2 \Bigr)\\
1634: + \sum_{\alpha\neq\beta}k_\alpha(k_\alpha-1)k_\beta(k_\beta-1),
1635: \end{multline}
1636: which simplifies to
1637: \begin{equation} \label{EQ:Ma2Mb2}
1638: \bigl[ (M^a)^2(M^b)^2\bigr]_{\mathcal{L}}
1639: = \sum_\alpha \Bigl(\frac{2}{3}k_\alpha^2 - \frac{2}{3}k_\alpha^4\Bigr)
1640: + \biggl(\sum_{\alpha} k_\alpha^2\biggr)^2.
1641: \end{equation}
1642: 
1643: It is important to verify that these expressions are compatible with those of
1644: Sects.~\ref{SEC:Correlation} and \ref{SEC:Cumulant}.
1645: First, spin isotropy demands that 
1646: $\langle (\hat{M}^a)^2 \rangle = \tfrac{1}{3}\langle \hat{\mathbf{M}}^2 \rangle$,
1647: which is confirmed by comparing Eqs.~\eqref{EQ:M2kalpha} and \eqref{EQ:MaMb}.
1648: Second, the identity
1649: \begin{equation}
1650: \hat{\mathbf{M}}^4
1651: = \sum_a (\hat{M}^a)^4 + \sum_{a\neq b} (\hat{M}^a)^2(\hat{M}^b)^2
1652: \end{equation}
1653: requires that 3~$\times$~Eq.~\eqref{EQ:Ma4} and 6~$\times$~Eq.~\eqref{EQ:Ma2Mb2}
1654: sum to Eq.~\eqref{EQ:M4kalpha}, which is indeed true.
1655: In summary, the loops estimators for components of the staggered magnetization
1656: at second and fourth order are
1657: \begin{align}
1658: (M^aM^b)_{\mathcal{L}} &= \delta^{ab}\sum_{\alpha} k^2_{\alpha}, \\
1659: \begin{split}
1660: \bigl[ (M^a)^2(M^b)^2 \bigr]_{\mathcal{L}}  &= \frac{2}{3}\sum_{\alpha} \bigl[
1661: \bigl(1-\delta^{ab}\bigr)k^2_{\alpha}-\bigl(1+2\delta^{ab}\bigr)k^4_{\alpha}\bigr)\bigr]\\
1662: &\quad +\bigl(1+2\delta^{ab}\bigr)\biggl(\sum_{\alpha} k^2_{\alpha}\biggr)^2.
1663: \end{split}
1664: \end{align}
1665: 
1666: \section{\label{SEC:Neel}The N\'{e}el State}
1667: 
1668: For a bipartite spin system, in which the $A$ and $B$ site labels designate
1669: genuine sublattices, the N\'{e}el state can be written in terms of any AB valence 
1670: bond configuration. This freedom exists because the staggered moments can be 
1671: organized into spin-up/spin-down pairs such that
1672: \begin{equation}
1673: \lvert R \rangle = \prod_{i \in A}b^\dagger_{i\uparrow}\prod_{j \in B}b^\dagger_{j\downarrow}\lvert
1674: \text{vac} \rangle = \prod_{n=1}^N b^\dagger_{i_n\uparrow}b^\dagger_{j_{Pn}\downarrow}\lvert
1675: \text{vac} \rangle
1676: \end{equation}
1677: for any $P \in \mathcal{S}_N$. Since
1678: $b^\dagger_{i\uparrow}b^\dagger_{j\downarrow} = \frac{1}{\sqrt{2}}(\chi^{0\dagger}_{ij}
1679: + i\chi^{3\dagger}_{ij})$, the N\'{e}el state can be viewed as a superposition of
1680: states with a fixed (but arbitrary) bond configuration and total triplet number ranging from $0$ to $N$:
1681: \begin{equation} \label{EQ:Neelstate}
1682: \lvert R \rangle = \frac{1}{2^{N/2}} \Bigl[ \lvert P \rangle
1683: + i \sum_{n=1}^N \lvert P; 3_n \rangle - \sum_{m<n}\lvert P; 3_m,3_n \rangle 
1684: + \cdots \Bigr].
1685: \end{equation}
1686: If we define $t_a(\vec{\mu}) = \sum_{n=1}^N \delta_{\mu_n,a}$, which counts the number of
1687: $a$-triplets in the index vector $\vec{\mu}$, then the overlap of $\lvert R \rangle$ 
1688: with an extended valence bond state $\lvert P; \vec{\mu} \rangle$ can be written as
1689: \begin{equation} \label{EQ:Neeloverlap}
1690: \langle R | P; \vec{\mu} \rangle = \delta_{t_1,0}\delta_{t_2,0}
1691: e^{i(\pi/2) t_3} 2^{-N/2}.
1692: \end{equation}
1693: The overlap vanishes if the state $\lvert P; \vec{\mu} \rangle$ contains any $1$- or $2$-triplets.
1694: Otherwise it is independent of $P$ and constant up to an overall phase factor with an
1695: angle given by $\pi/2 \times \text{the number of $3$-triplets}$.
1696: The derivation of Eq.~\eqref{EQ:Neeloverlap} relies on nothing more than the 
1697: (species) orthogonality of bonds that connect the same two sites
1698: [expressed in Eq.~\eqref{EQ:chi2-commutator}]. There is no need to account
1699: for bond-configurational mismatch because the permutation $P$ in Eq.~\eqref{EQ:Neelstate} 
1700: can be chosen equal to the permutation in $\lvert P; \vec{\mu} \rangle$.
1701: 
1702: Matrix elements taken between the N\'{e}el state and a valence bond singlet state are particularly
1703: easy to compute. For example, applying Eq.~\eqref{EQ:gammainjnP}, we find that
1704: \begin{equation}
1705: \frac{\langle R \rvert \hat{\gamma}_{i_nj_m} \lvert P \rangle}{\langle R | P \rangle}
1706: = \biggl(\frac{1}{2}\biggr)^{1-\delta_{Pn,m}}
1707: \frac{\langle R | (Pn\ m) P \rangle}{\langle R | P \rangle},
1708: \end{equation}
1709: where the ratio of overlaps on the right-hand side is 1 since both 
1710: $\langle R | P \rangle$ and $\langle R | (Pn\ m) P \rangle$ have the value $2^{-N/2}$.
1711: In general, the matrix element picks up a factor of $\tfrac{1}{2}$ for each offdiagonal
1712: $\hat{\gamma}$ operation (keeping in mind that subsequent operations act on the
1713: reconfigured bonds). The evaluation rule can be expressed as
1714: \begin{equation}
1715: \frac{\langle R \rvert \overbrace{\hat{\gamma}_{ij}\hat{\gamma}_{kl}\cdots}^{\text{$n$ operators}} \lvert P \rangle}{\langle R | P \rangle}
1716: = 2^{N_{\circlearrowleft}-n},
1717: \end{equation}
1718: where $N_{\circlearrowleft}$ now refers to the number of closed loops that are formed
1719: by the bonds and $\hat{\gamma}$ vertices.
1720: 
1721: \begin{figure}
1722: \begin{center}
1723: \includegraphics{reference-overlap.eps}
1724: \end{center}
1725: \caption{\label{FIG:referenve-overlap}
1726: Matrix elements of $\hat{\gamma}$ taken between a valence bond state 
1727: and the reference N\'{e}el state. Closed loops containing one triplet
1728: do not contribute, whereas open strings with one triplet contribute 
1729: the same value as their singlet-bond-only counterpart.}
1730: \end{figure}
1731: 
1732: In analogy with Eq.~\eqref{EQ:QvuPmuoverlap}, the rules can be extended to cover
1733: states with arbitrary numbers of triplets: 
1734: \begin{equation}
1735: \frac{\langle R \rvert \overbrace{\hat{\gamma}_{ij}\hat{\gamma}_{kl}\cdots}^{\text{$n$ operators}} \lvert P; \vec{\mu} \rangle}{\langle R | P \rangle}
1736: = 2^{N_{\circlearrowleft}-n} \prod_{\alpha=1}^{N_{\circlearrowleft}+N_{\looparrowleft}} K_\alpha.
1737: \end{equation}
1738: In this case, the index $\alpha$ ranges over both the closed loops and the open strings ($N_{\looparrowleft}$
1739: in number) that are formed. 
1740: A closed loop involving $k \le n$ operators and bonds numbered $1,2,  \ldots, k$ has associated with it
1741: a factor
1742: \begin{equation}
1743: K_{\alpha} = \frac{1}{2}\tr \bigl( \tau^{0\dagger}\tau^{\mu_1}\tau^{0\dagger}\tau^{\mu_2}
1744: \cdots \tau^{0\dagger}\tau^{\mu_k}\bigr).
1745: \end{equation}
1746: An open string involving $k \le n$ operators and bonds numbered $1,2,\ldots,k+1$ has a factor
1747: \begin{equation}
1748: K_{\alpha} = \frac{i}{2}\tr \bigl( \tau^{3\dagger}\tau^{\mu_1}\tau^{0\dagger}\tau^{\mu_2}
1749: \cdots \tau^{0\dagger}\tau^{\mu_{k+1}}\bigr).
1750: \end{equation}
1751: Note that any lone triplet in a closed loop gives $K_{\alpha}=0$, whereas 
1752: a lone 3-triplet in an open string gives $K_{\alpha}=i$. See Fig.~\ref{FIG:referenve-overlap}.
1753: 
1754: \section{\label{SEC:Singlet-triplet}Singlet-triplet Gap}
1755: 
1756: Consider a quantum spin system whose hamiltonian $\hat{H}$ has a
1757: global singlet ground state 
1758: $\lvert \psi_0 \rangle$.
1759: Starting from an arbitrary singlet trial state 
1760: \begin{equation}
1761: \lvert \psi^{\text{trial}}_{0} \rangle = \sum_P c_P^{\text{trial}} \lvert P \rangle,
1762: \end{equation}
1763: the true ground state wavefunction can be obtained by projection:
1764: \begin{equation} \label{EQ:projtrial0}
1765: \lvert \psi_{0} \rangle = \lim_{\tau\to\infty} e^{-\tau\hat{H}} \lvert \psi^{\text{trial}}_{0} \rangle
1766: = \sum_P c_P \lvert P \rangle.
1767: \end{equation}
1768: From the eigen-equation $\hat{H}\lvert \psi_{0} \rangle = E_0\lvert \psi_{0} \rangle$,
1769: the ground state energy can be isolated by acting from the left with an appropriate reference state:
1770: \begin{equation}
1771: E_0 = \frac{\langle R \rvert \hat{H} \lvert \psi_{0}\rangle}{\langle R | \psi_{0}\rangle}.
1772: \end{equation}
1773: For our purposes, it is most convenient to use a N\'{e}el reference state $\lvert R \rangle$, since
1774: it has a constant-magnitude overlap with all states in the full Hilbert space.
1775: Thus, 
1776: \begin{equation} \label{EQ:E0_W}
1777: E_0 = \frac{\langle R \rvert \hat{H} \lvert \psi_{0}\rangle}{\langle R | \psi_{0}\rangle}
1778: = \frac{ \sum_P W(P) \frac{\langle R \rvert \hat{H} \lvert P\rangle}{\langle R | P\rangle}}{\sum_P W(P)}
1779: \end{equation}
1780: with $W(P) = c_P \langle R | P \rangle = 2^{-N/2} c_P$.
1781: This formulation is similar to that of Eq.~\eqref{EQ:psiOpsi}, except that here
1782: the basic objects of interest are bonds rather than loops.
1783: We can view the ground state energy $E_0 = \langle H_\mathcal{B} \rangle_W$
1784: as the expectation value of the bond estimator for the hamiltonian, $H_\mathcal{B}$, 
1785: in a fluctuating gas of valence bonds.
1786: 
1787: Similarly, by constructing a trial wavefunction in the triplet sector,
1788: \begin{equation}
1789: \lvert \psi^{\text{trial}}_{1}\rangle = \hat{M}^3 \lvert \psi^{\text{trial}}_{0} \rangle = i\sum_{n=1}^N \sum_P c_P^{\text{trial}} \lvert P; 3_n \rangle,
1790: \end{equation}
1791: we can project onto the lowest-energy triplet state:
1792: \begin{equation}
1793: \lvert \psi_{1} \rangle = \lim_{\tau\to\infty} e^{-\tau\hat{H}} \hat{M}^3\lvert \psi^{\text{trial}}_{0} \rangle.
1794: \end{equation}
1795: The structure of this projection is closely related to that of Eq.~\eqref{EQ:projtrial0}. In the singlet case,
1796: the evolution operator consists of a long string of $-\hat{H}_{ij}$ operators that successively
1797: reconfigure the bond configuration $\lvert P \rangle$ of the trial state.
1798: Here, we simply need to compute the result of the same strings operating
1799: on $\lvert P; 3_n \rangle$. We know, however, that the update rules for singlet
1800: and triplet bonds are identical except that a direct diagonal operation kills a triplet.
1801: It is straightforward to reinterpret the operator string by designating one
1802: bond of the trial state as a triplet and tracing its evolution. 
1803: As shown in Fig.~\ref{FIG:gauntlet}, the triplet will either run the gauntlet or die in the attempt.
1804: Accordingly, we can write
1805: \begin{equation}
1806: \lvert \psi_{1} \rangle = i \sum_{n=1}^N \sum_P  g_n(P)\,c_P \lvert P; 3_n\rangle,
1807: \end{equation}
1808: where $g_n(P)$ counts the number of surviving triplets whose final destination
1809: is the $n^{\text{th}}$ bond of configuration $P$. By definition,
1810: \begin{equation}
1811: 0 \le \sum_{n=1}^N g_n(P) \le N.
1812: \end{equation}
1813: 
1814: \begin{figure}
1815: \begin{center}
1816: \includegraphics{gauntlet.eps}
1817: \end{center}
1818: \caption{\label{FIG:gauntlet} (Top) The evolution operator (in series expansion)
1819: involves long series of
1820: $-\hat{H}_{ij}$ operator strings that map VB configurations in the trial
1821: state to VB configurations in the true ground state through
1822: sequential reordering of the bonds. In the figure, the propagation index~\cite{Sandvik97}
1823: increases from right to left along the horizontal axis. A set of singlet valence
1824: bonds are arranged along the vertical spatial axis.
1825: The red bars denote nearest neighbour interactions.
1826: (Middle) The same operator string can be reinterpreted as acting on a trial
1827: state with one triplet, marked here in purple. 
1828: (Bottom) A triplet state that is acted on directly by a diagonal operation is annihilated.
1829:  }
1830: \end{figure}
1831: 
1832: The energy of the low-lying triplet state is
1833: \begin{equation} 
1834: E_1 = \frac{\langle R \rvert \hat{H} \lvert \psi_{1}\rangle}{\langle R | \psi_{1}\rangle}
1835: = \frac{ \sum_P W(P) \sum_{n=1}^Ng_n(P) \frac{\langle R \rvert \hat{H} \lvert P; 3_n\rangle}{\langle R | P;3_n\rangle}}{\sum_P W(P) \sum_{n=1}^Ng_n(P)}.
1836: \end{equation}
1837: Here we have made use of the fact that $\langle R | P; 3_n \rangle = i\langle R | P \rangle$.
1838: Alternatively, $E_1$ can be expressed as
1839: \begin{equation} \label{EQ:E1_W}
1840: E_1 =   \frac{\bigl\langle \sum_n g_n H_{\mathcal{B}} \bigr\rangle_W}
1841: {\bigl\langle \sum_n g_n \bigr\rangle_W}
1842: \end{equation}
1843: since $\langle R \rvert \hat{H} \lvert P; 3_n\rangle/\langle R | P;3_n\rangle$
1844: and $\langle R \rvert \hat{H} \lvert P \rangle/\langle R | P \rangle$ differ
1845: only when $g_n = 0$.
1846: Combining Eqs.~\eqref{EQ:E0_W} and \eqref{EQ:E1_W} gives
1847: \begin{equation} \label{EQ:gap}
1848: E_1 - E_0 = \frac{\bigl\langle \sum_n g_n H_{\mathcal{B}} \bigr\rangle_W
1849: -\bigl\langle \sum_n g_n \bigr\rangle_W \langle H_{\mathcal{B}} \rangle_W}
1850: {\bigl\langle \sum_n g_n \bigr\rangle_W}.
1851: \end{equation}
1852: 
1853: A numerical measurement of the singlet-triplet gap via Eq.~\eqref{EQ:gap}
1854: has the advantage that it can be carried out during a simulation of the
1855: $S=0$ ground state properties. This gives a substantial error cancellation
1856: in stochastic methods over techniques where $E_0$ and $E_1$ are 
1857: computed individually and subtracted. In some systems, however, the
1858: lowest-energy singlet and triplet states may differ so substantially that
1859: the reweighting of the operator string becomes inefficient. In that case,
1860: so few triplets survive the projection that $\langle \sum_n g_n \rangle \approx 0$.
1861: 
1862: \section{\label{SEC:Stiffness} Spin stiffness }
1863: 
1864: Again, let us imagine that the $S=0$ ground state is obtained via projection:
1865: \begin{equation}
1866: \lvert \psi \rangle = \lim_{\tau\to\infty} e^{-\tau\hat{H}} \lvert \psi^{\text{trial}} \rangle.
1867: \end{equation}
1868: Now suppose that we introduce a twist field $\phi(\mathbf{r})$ 
1869: representing a local rotation of the spins about the 3 axis.
1870: We will be concerned with the limit in which the field gradient 
1871: $\nabla \phi(\mathbf{r})$ is small with respect to the lattice spacing.
1872: For concreteness, we endow the field with a single long-wavelength mode
1873: $\phi(\mathbf{r}) = \phi_0 + \mathbf{r}\cdot\mathbf{q}$. As a consequence, 
1874: the relative spin rotation angles satisfy $\theta_{ij} = (\mathbf{r}_i - \mathbf{r}_j)\cdot\mathbf{q}$.
1875: 
1876: Equation~\eqref{EQ:chidaggerchirotate} then suggests that the hamiltonian---assuming it has the form 
1877: of Eq.~\eqref{EQ:modelH}---transforms as
1878: \begin{equation} \label{EQ:H-twisted}
1879: -\hat{H} \to -\hat{H}[\mathbf{q}] = -\hat{H} + q^a\hat{T}^a + \frac{1}{2}q^aq^b\hat{G}^{ab},
1880: \end{equation}
1881: where the gradient and Hessian are given by
1882: \begin{equation}\label{EQ:Ta}
1883: \hat{T}^a = \frac{1}{2}\sum_{\langle ij \rangle}J_{ij} (\mathbf{r}_i-\mathbf{r}_j)\cdot\mathbf{e}^a
1884: \Bigl(\chi^{0\dagger}_{ij}\chi^{3}_{ij}+\chi^{3\dagger}_{ij}\chi^{0}_{ij}\Bigr) + \cdots
1885: \end{equation}
1886: and
1887: \begin{multline} \label{EQ:Gab}
1888: \hat{G}^{ab} = \frac{1}{2}\sum_{\langle ij \rangle}J_{ij} (\mathbf{r}_i-\mathbf{r}_j)\cdot\mathbf{e}^a
1889: (\mathbf{r}_i-\mathbf{r}_j)\cdot\mathbf{e}^b\\
1890: \times\Bigl(-\chi^{0\dagger}_{ij}\chi^{0}_{ij}
1891: + \chi^{3\dagger}_{ij}\chi^{3}_{ij}\Bigr) + \cdots
1892: \end{multline}
1893: Here, $+ \cdots$ represents additional interaction terms in $K_{ijkl}$ and beyond.
1894: Equation~\eqref{EQ:Ta} is a spin current operator that changes the number of 
1895: triplet bonds by $\pm1$. (See Fig.~\ref{FIG:updates3}.) It does so in proportion to the 
1896: interaction strength and the projection of the bond length onto the axis of propagation.
1897: Equation~\eqref{EQ:Gab}, on the other hand, changes the triplet count by 
1898: $0$ or $\pm 2$.
1899: 
1900: \begin{figure*}
1901: \includegraphics{updates3.eps}
1902: \caption{ \label{FIG:updates3}
1903: Several possible bond reconfigurations are shown, corresponding
1904: to action by each term in $\hat{T}^a$. The relevant prefactors are suppressed.}
1905: \end{figure*}
1906: 
1907: Under the influence of the twist field, the ground state energy is
1908: \begin{equation}
1909: E[\mathbf{q}] = \lim_{\tau\to\infty} \frac{ \langle R \lvert \hat{H}[\mathbf{q}] e^{-\tau \hat{H}[\mathbf{q}]} \rvert \psi_0 \rangle }{ \langle R \lvert e^{-\tau \hat{H}[\mathbf{q}]} \rvert \psi_0 \rangle }.
1910: \end{equation}
1911: By virtue of the variational principle, $E[\mathbf{q}]$ must be stationary
1912: with respect to variation.
1913: Its convexity at $\mathbf{q}=0$ is a measure of the spin stiffness.
1914: Derivatives $\partial_a =  \partial/\partial q^a$ can be handled
1915: using the identity
1916: \begin{equation}
1917: \int_0^\tau \! d\tau' \, e^{-(\tau-\tau')\hat{H}} \partial_a \hat{H}
1918: e^{-\tau'\hat{H}}
1919: = - \partial_a\bigl( e^{-\tau\hat{H}} \bigr).
1920: \end{equation}
1921: Hence,
1922: \begin{equation} \label{EQ:Eqdqa}
1923: \frac{\partial E[\mathbf{q}]}{\partial q^a} \biggr\rvert_{\mathbf{q}=0}
1924: = \frac{\langle R \lvert \hat{T}^a \rvert \psi \rangle}
1925: {\langle R | \psi \rangle}
1926: + \frac{\langle R \lvert (\hat{H}-E) \rvert \psi^a \rangle}
1927: {\langle R | \psi \rangle} = 0,
1928: \end{equation}
1929: with $\lvert \psi^a \rangle = \partial_a \lvert \psi \rangle\bigr\rvert_{\mathbf{q}=0}$
1930: given by
1931: \begin{equation}
1932: \lvert \psi^a \rangle =
1933: \int_0^\tau \! d\tau' \, e^{-(\tau-\tau')\hat{H}} \hat{T}^a e^{-\tau'\hat{H}}\lvert \psi^{\text{trial}} \rangle.
1934: \end{equation}
1935: The state $\lvert \psi^a \rangle$ evolves from the
1936: singlet trial state but differs from the projected
1937: ground state in that it is constructed with an $\mathbf{e}^a$-directed triplet injected
1938: at all times $\tau' < \tau$.
1939: 
1940: Making use of Eq.~\eqref{EQ:Eqdqa}, we can 
1941: show that the spin stiffness is given by
1942: \begin{widetext}
1943: \begin{equation} \label{EQ:spinstiffness}
1944: \rho^{ab}_{\text{s}} = \frac{\partial^2 E[\mathbf{q}]}{\partial q^a\partial q^b}
1945: \biggr\rvert_{\mathbf{q}=0}
1946: = \frac{\langle R \lvert \hat{G}^{ab} \rvert \psi \rangle}
1947: {\langle R | \psi \rangle}
1948: + \frac{\langle R \lvert \hat{T}^{a} \rvert \psi^b \rangle}
1949: {\langle R | \psi \rangle}
1950: + \frac{\langle R \lvert \hat{T}^{b} \rvert \psi^a \rangle}
1951: {\langle R | \psi \rangle}
1952: + \frac{\langle R \lvert (\hat{H}-E) \rvert \psi^{ab} \rangle}
1953: {\langle R | \psi \rangle},
1954: \end{equation}
1955: where
1956: \begin{equation}
1957: \begin{split}
1958: \lvert \psi^{ab} \rangle &= 
1959: \int_0^\tau \! d\tau' \! \int_{\tau'}^\tau \! d\tau'' \, e^{-(\tau-\tau'')\hat{H}} \hat{T}^a e^{-(\tau''-\tau')\hat{H}} \hat{T}^b e^{-\tau'\hat{H}}\lvert \psi^{\text{trial}} \rangle \\
1960: &+\int_0^\tau \! d\tau' \! \int_{\tau'}^\tau \! d\tau'' \, e^{-(\tau-\tau'')\hat{H}} \hat{T}^b e^{-(\tau''-\tau')\hat{H}} \hat{T}^a e^{-\tau'\hat{H}}\lvert \psi^{\text{trial}} \rangle \\
1961: & + \int_0^\tau \! d\tau' \, e^{-(\tau-\tau')\hat{H}} \hat{G}^{ab} e^{-\tau'\hat{H}}\lvert \psi^{\text{trial}} \rangle.
1962: \end{split}
1963: \end{equation}
1964: \end{widetext}
1965: The state $\lvert \psi^{ab} \rangle$ evolves from the
1966: singlet trial state $\lvert \psi^{\text{trial}} \rangle$ with two $\mathbf{e}^a$- and
1967: $\mathbf{e}^b$-directed triplets injected at times $\tau'$ and $\tau''$.
1968: 
1969: Equation~\eqref{EQ:spinstiffness} can be evaluated using the reweighting
1970: trick introduced in Sect.~\ref{SEC:Singlet-triplet}. 
1971: The same operator sequence used to generate the singlet ground state can
1972: be reinterpreted to include all possible triplet pairs (at all possible
1973: starting locations, not simply at the $\tau = 0$ end as was the case
1974: for the singlet-triplet gap measurement). The spin stiffness is related
1975: to the fluctuations in energy of the final bond configuration arising
1976: from operator sequences in which neither triplet is annihilated.
1977: 
1978: \section{Summary}
1979: 
1980: We have presented in detail a formal framework for organizing calculations in the
1981: valence bond basis. This approach is based on manipulations of valence bond 
1982: creation and annihilation operators, $\chi^{\mu\dagger}_{ij}$ and $\chi^{\mu}_{ij}$,
1983: that act between any two sites of the spin lattice. 
1984: These operators are similar to those introduced in Ref.~\onlinecite{Sachdev90} for fixed dimer configurations but are endowed with an expanded operator algebra that is
1985: compatible with the overcomplete basis of all possible dimer configurations.
1986: 
1987: We have focused on what we call the AB valence bond basis, in 
1988: which half the lattice sites are assigned the label A and the other half B
1989: and only dimers connecting sites with different labels are allowed.
1990: (The use of AB bonds only is routine in the literature, especially when
1991: the hamiltonian of interest does not contain explicitly frustrating AA/BB interactions.)
1992: The states in this restricted basis are uniquely characterized by 
1993: permutations of the B site labels. Overlaps and matrix elements of such
1994: states are related in a systematic way to the cycle structure of the permutations.
1995: We have shown, for example, that the presence of antiferromagnetism in a bipartite system
1996: is related to the long tail behaviour of the cycle length distribution $\langle n_k \rangle_W$.
1997: 
1998: Correlation functions of the isotropic spin interaction 
1999: $\mathbf{S}_i \cdot \mathbf{S}_j$ are related 
2000: to the cumulants of the operator
2001: $\hat{\gamma}_{ij} = \frac{1}{4}\bigl(1+\epsilon_{ij}\bigr) 
2002: - \epsilon_{ij} \chi^{0\dagger}_{ij} \chi^{0}_{ij}$.
2003: The various contributions can thus be computed in a straightforward way
2004: from the set of connected Goldstone diagrams---albeit with the diagrams
2005: interpreted quite differently than they are in the usual quantum many-body physics context.
2006: (Moreover, there is no need to worry about which of $i$ and $j$ are A or B sites.
2007: The different cases are all handled automatically by the sign factor $\epsilon_{ij}$.)
2008: As an example of this approach, we have derived explicit expressions for second-, 
2009: fourth-, and sixth-order expectation values of the staggered magnetization.
2010: 
2011: We have also emphasized that, when triplet bonds are fully accounted for,
2012: the valence bond basis spans the entire Hilbert space, not merely the $S=0$ subspace.
2013: The valence bond operators carry an index $\mu$ that specifies the singlet ($\mu = 0$) or 
2014: triplet ($\mu = 1,2,3$) nature of each bond. We have extended the rules for computing
2015: overlaps and matrix elements to include states with triplet bonds. 
2016: As an example of how this can be useful, we have derived
2017: expressions for the singet-triplet gap and the spin stiffness
2018: that can be interpreted a reweightings of the
2019: operator string in a projected singlet ground state.
2020: The reweighting formulation is ideally suited for use in valence bond projector Monte Carlo.
2021: 
2022: This material is based upon work supported by the National Science 
2023: Foundation under Grant No.\ DMR-0513930.
2024: 
2025: \appendix
2026: 
2027: \section{\label{SEC:Permutation}}
2028: A permutation $P \in \mathcal{S}_N$ is a bijective map that takes $n$ to $Pn$
2029: (and $\bar{P}n$ to $n$) for every $1 \le n \le N$:
2030: \begin{equation} \label{EQ:permPexplicit}
2031: P = \begin{pmatrix}
2032: 1 & P1\\
2033: 2 & P2\\
2034: \vdots & \vdots \\
2035: N & PN
2036: \end{pmatrix}.
2037: \end{equation}
2038: An alternative notation for Eq.~\eqref{EQ:permPexplicit}
2039: involves cycles of the form $(1\ 2\ 3)$, which expresses
2040: the functional relationship $1\to2$, $2\to 3$, and $3 \to 1$.
2041: Every permutation can be decomposed into a product of disjoint cycles
2042: \begin{equation} \label{EQ:cycledecomp}
2043: P = \prod_{\alpha=1}^{\cycles(P)}
2044: ( \xi_\alpha\ P \xi_\alpha\ P^2 \xi_\alpha\ \cdots \ P^{k_\alpha-1} \xi_\alpha ),
2045: \end{equation}
2046: where $\cycles(P)$ is the number of cycles in $P$, $ \xi_\alpha$ is 
2047: a representative number (the lowest, say) in a particular cycle, and $k_\alpha$ is
2048: its length. Two numbers $n$ and $m$ are in the same cycle of $P$ iff
2049: there is some integer power $k \ge 0$ such that $P^kn = m$, which we denote
2050: $n \overset{P}{\sim} m$. It follows that $n$ is in the cycle labelled $\alpha$
2051: iff $n \overset{P}{\sim} \xi_{\alpha}$.
2052: 
2053: The cycle decomposition given in Eq.~\eqref{EQ:cycledecomp} is unique
2054: and requires the fewest possible number of cycles. One can, however,
2055: further decompose each term into several smaller cycles that are no longer disjoint
2056: (\emph{i.e.}, they may have elements in common).
2057: This process can always be carried further until the permutation
2058: is expressed entirely as a product of transpositions (cycles of length 2).
2059: Accordingly, to understand how the cycle structure of $\bar{Q}P'$ differs from that
2060: of $\bar{Q}P$, we need only consider the sequence of changes
2061: induced by the chain of transpositions that separates $P'$ from $P$.
2062: 
2063: The result of a single transposition acting on $P$ is
2064: \begin{align}
2065: (n\ m)P &= \begin{pmatrix}
2066: 1 & P1\\
2067: 2 & P2\\
2068: \vdots & \vdots \\
2069: \bar{P}n & m \\
2070: \vdots & \vdots \\
2071: \bar{P}m & n \\
2072: \vdots & \vdots \\
2073: N & PN
2074: \end{pmatrix}.
2075: \intertext{Hence,}
2076: \bar{Q}(n\ m)P &= \begin{pmatrix}
2077: 1 & \bar{Q}P1\\
2078: 2 & \bar{Q}P2\\
2079: \vdots & \vdots \\
2080: \bar{P}n & \bar{Q}m \\
2081: \vdots & \vdots \\
2082: \bar{P}m & \bar{Q}n \\
2083: \vdots & \vdots \\
2084: N & P\bar{Q}N
2085: \end{pmatrix}.
2086: \end{align}
2087: As to how this differs from $\bar{Q}P$, there are three possibilities to consider.
2088: First, when $n=m$, the 2-cycle is just an identity element; the equalities
2089: $\bar{Q}P = \bar{Q}(n\ m)P$ and $\cycles(\bar{Q}P) = \cycles(\bar{Q}(n\ m)P)$
2090: follow trivially.
2091: Otherwise, the outcome depends on whether $\bar{P}n$
2092: and $\bar{P}m$ are in the same cycle. If they are,
2093: the transposition splits one cycle into two,
2094: \begin{align}
2095: \bar{Q}P &= (1\ 8\  3\ \bar{P}n\ 2\ 7\ 4\ \bar{P}m\ 6\ 5)\cdots \\
2096: \bar{Q}(n\ m)P &= (1\ 8\ 3\ \bar{P}m\ 6\ 5)(2\ 7\ 4\ \bar{P}n)\cdots,
2097: \end{align}
2098: and $\cycles(\bar{Q}(n\ m)P) - \cycles(\bar{Q}P) = +1$.
2099: If $\bar{P}n$ and $\bar{P}m$ are in different cycles,
2100: the transposition merges two cycles into one,
2101: \begin{align}
2102: \bar{Q}P &= (1\ 3\ 7\ \bar{P}n\ 4\ 6)(2\ 8\ \bar{P}m\ 5)\cdots \\
2103: \bar{Q}(n\ m)P &= (1\ 3\ 7\ \bar{P}m\ 5\ 2\ 8\ \bar{P}m\ 4\ 6)\cdots,
2104: \end{align}
2105: and $\cycles(\bar{Q}(n\ m)P) - \cycles(\bar{Q}P) = -1$.
2106: These three cases can be summarized by
2107: \begin{equation}
2108: \cycles(\bar{Q}(n\ m)P) - \cycles(\bar{Q}P) + \delta_{n,m} 
2109: = \begin{cases}
2110: +1 & \text{if $\bar{P}n \overset{\bar{Q}P}{\sim} \bar{P}m$ } \\
2111: -1 & \text{otherwise}
2112: \end{cases}
2113: \end{equation}
2114: or, equivalently,
2115: \begin{equation}
2116: \frac{\langle Q \lvert (n\ m) P\rangle}{\langle Q | P \rangle} \biggl(\frac{1}{2}
2117: \biggr)^{1-\delta_{n,m}} 
2118: = \begin{cases}
2119: 1 & \text{if $\bar{P}n \overset{\bar{Q}P}{\sim} \bar{P}m$ } \\
2120: \frac{1}{4} & \text{otherwise}.
2121: \end{cases}
2122: \end{equation}
2123: 
2124: \begin{thebibliography}{99}
2125: 
2126: \bibitem{Rumer32} G.\ Rumer, Gottingen Nachr.\ Tech.\ {\bf 1932}, 377 (1932).
2127: \bibitem{Pauling33} L.\ Pauling, J.\ Chem.\ Phys.\ {\bf 1}, 280 (1933).
2128: \bibitem{Hulthen38} L.\ Hulth\'{e}n, Arkiv Mat.\ Astron.\ Fysik {\bf 26A}, No.\ 11 (1938).
2129: \bibitem{Bethe31} H.\ Bethe, Z.\ Phys.\ {\bf 71}, 205 (1931).
2130: \bibitem{Majumdar69} C.\ K.\ Majumdar and D.\ K.\ Ghosh, J.\ Math.\ Phys.\ {\bf 10}, 1388 (1969).
2131: \bibitem{Fazekas74} P.\ Fazekas and P.\ W.\ Anderson, Philos.\ Mag.\ {\bf 30}, 23 (1974).
2132: \bibitem{Anderson87a} P.\ W.\ Anderson, Science {\bf 235}, 1196 (1987).
2133: \bibitem{Anderson87b} P.\ W.\ Anderson, G.\ Baskaran, Z.\ Zou, and T.\ Hsu, Phys.\ Rev.\ Lett.\ {\bf 58},
2134: 2790 (1987).
2135: \bibitem{Liang88} S.\ Liang, B.\ Doucot, and P.\ W.\ Anderson, Phys.\ Rev.\ Lett.\ {\bf 61}, 365 (1988).
2136: \bibitem{Kohmoto88} M.\ Kohmoto, Phys.\ Rev.\ B {\bf 37}, 3812 (1988).
2137: \bibitem{Lou06} J.\ Lou and A.\ W.\ Sandvik, to be published; {\tt cond-mat/0605034}.
2138: \bibitem{Poilblanc89} D.\ Poilblanc, Phys.\ Rev.\ B {\bf 39}, 140 (1989).
2139: \bibitem{Gros88} C.\ Gros, Phys.\ Rev.\ B {\bf 38}, 931 (1988).
2140: \bibitem{Capriotti01} L.\ Capriotti, F.\ Becca, A.\ Parola, and S.\ Sorella, Phys.\ Rev.\ Lett.\ {\bf 87}, 097201 (2001).
2141: \bibitem{Lin90} H.\ Q.\ Lin, Phys.\ Rev.\ B {\bf 42}, 6561 (1990).
2142: \bibitem{Ramsesha90} S.\ Ramasesha and Krishna Das, Phys.\ Rev.\ B {\bf 42}, 10682 (1990).
2143: \bibitem{Balint69} G.\ G.\ Balint-Kurti and M.\ Karplus, J.\ Chem.\ Phys.\ {\bf 50}, 478 (1969).
2144: \bibitem{Soos90} Z.\ G.\ Soos and S.\ Ramasesha, in \emph{Valence Bond Theory and Chemical Structure}, edited by D.\ J.\ Klein and N.\ Trinajstic, Elsevier, New York (1990).
2145: \bibitem{Figueirido89} F.\ Figueirido \emph{et al}., Phys.\ Rev.\ B {\bf 41}, 4619 (1989).
2146: \bibitem{Bose91} I.\ Bose and P.\ Mitra, Phys.\ Rev.\ B {\bf 44}, 443 (1991).
2147: \bibitem{Read89} N.\ Read and S.\ Sachdev, Phys.\ Rev.\ Lett.\ {\bf 62}, 1694 (1989).
2148: \bibitem{Read90} N.\ Read and S.\ Sachdev, Phys.\ Rev.\ B {\bf 42}, 4568 (1990).
2149: \bibitem{Kivelson87} S.\ A.\ Kivelson, D.\ S.\ Rokhsar, and J.\ P.\ Sethna, Phys. Rev. B {\bf 35}, 8865 (1987).
2150: \bibitem{Sondhi97} S.\ L.\ Sondhi \emph{et al}., Rev.\ Mod.\ Phys.\ {\bf 69}, 315 (1997).
2151: \bibitem{Sachdev03} S.\ Sachdev, Rev.\ Mod.\ Phys.\ {\bf 75}, 913 (2003).
2152: \bibitem{Vishwanath04} A.\ Vishwanath, L.\ Balents, and T.\ Senthil, Phys.\ Rev.\ B {\bf 69}, 224416 (2004).
2153: \bibitem{Senthil04} T.\ Senthil \emph{et al.}, Science {\bf 303}, 1490 (2004).
2154: \bibitem{Liang90} S. Liang, Phys.\ Rev.\ Lett.\ {\bf 42}, 6555 (1990).
2155: \bibitem{Santoro99} G.\ Santoro, S.\ Sorella, L.\ Guidoni, A.\ Parola, and E.\ Tosatti, 
2156: Phys.\ Rev.\ Lett.\ {\bf 83}, 3065 (1999).
2157: \bibitem{Sandvik05} A.\ W.\ Sandvik, Phys.\ Rev.\ Lett.\ {\bf 95}, 207203 (2005).
2158: \bibitem{Syluasen02} O.\ F.\ Sylju\aa sen \emph{et al.}, Phys.\ Rev.\ E {\bf 66}, 046701 (2002). 
2159: \bibitem{Evertz03} H.\ G.\ Evertz, Adv.\ Phys.\ {\bf 52}, 1 (2003).
2160: \bibitem{Rokhsar88} D.\ S.\ Rokhsar and S.\ A.\ Kivelson, Phys.\ Rev.\ Lett.\ {\bf 61}, 2376 (1988).
2161: \bibitem{Sutherland88} B.\ Sutherland, Phys.\ Rev.\ B {\bf 37}, 3786 (1988). 
2162: \bibitem{Sachdev90} S.\ Sachdev and R.\ N.\ Bhatt, Phys.\ Rev.\ B 41, 9323 (1990).
2163: \bibitem{Raman05} K.\ S.\ Raman, R.\ Moessner, and S.\ L.\ Sondhi, Phys.\ Rev.\ B {\bf 72},
2164: 064413 (2005).
2165: \bibitem{Marshall55} W.\ Marshall, Proc.\ Roy.\ Soc.\ (London) {\bf A232}, 48 (1955).
2166: \bibitem{Sutherland88b} B.\ Sutherland. Phys.\ Rev.\ B {\bf 38}, 7192 (1988).
2167: \bibitem{Kohmoto88b} M.\ Kohmoto and Y.\ Shapir, Phys.\ Rev.\ B {\bf 37}, 9439 (1988).
2168: \bibitem{Binder81} K.\ Binder, Z.\ Phys.\ B {\bf 43}, 119 (1981).
2169: \bibitem{Sandvik97} A.\ W.\ Sandvik, R.\ R.\ P.\ Singh, and D.\ K.\ Campbell, Phys.\ Rev.\ B {\bf 56},
2170: 14510 (1997).
2171: 
2172: \end{thebibliography}
2173: 
2174: \end{document}
2175: