1: \documentclass[prl,aps,epsfig,twocolumn,showpacs]{revtex4}
2: \usepackage{epsfig,amssymb}
3:
4: \newcommand\beq{\begin{equation}}
5: \newcommand\eeq{\end{equation}}
6: \newcommand\bea{\begin{eqnarray}}
7: \newcommand\eea{\end{eqnarray}}
8: \newcommand\non{\nonumber}
9: \newcommand\noi{\noindent}
10: \newcommand\al{\alpha}
11: \newcommand\be{\beta}
12: \newcommand\de{\delta}
13: \newcommand\ep{\epsilon}
14: \newcommand\ka{\kappa}
15: \newcommand\La{\Lambda}
16: \newcommand\si{\sigma}
17: \newcommand\dg{\dagger}
18: \newcommand\up{\uparrow}
19: \newcommand\dn{\downarrow}
20: \newcommand\pa{\partial}
21: \newcommand\bib{\bibitem}
22:
23: \begin{document}
24:
25: \textheight=23.8cm
26:
27: \title{\Large Gapless points of dimerized quantum spin chains: analytical and
28: numerical studies}
29: \author{\bf V. Ravi Chandra, Diptiman Sen and Naveen Surendran}
30: \affiliation{\it Centre for High Energy Physics, Indian Institute of Science,
31: Bangalore 560012, India}
32:
33: \date{\today}
34: \pacs{~75.10.Jm, ~75.10.Pq, ~73.43.Nq}
35:
36: \begin{abstract}
37: We study the locations of the gapless points which occur for quantum spin
38: chains of finite length (with a twisted boundary condition) at particular
39: values of the nearest neighbor dimerization, as a function of the spin $S$
40: and the number of sites. For strong dimerization and large values of $S$, a
41: tunneling calculation reproduces the same results as those obtained from more
42: involved field theoretic methods using the non-linear $\sigma$-model approach.
43: A different analytical calculation of the matrix element between the two N\'eel
44: states gives a set of gapless points; for strong dimerization, these differ
45: significantly from the tunneling values. Finally, the exact diagonalization
46: method for a finite number of sites yields a set of gapless points which are
47: in good agreement with the N\'eel state calculations for all values of the
48: dimerization, but the agreement with the tunneling values is not very good
49: even for large $S$. This raises questions about possible corrections to the
50: tunneling results.
51: \end{abstract}
52: \maketitle
53: \vskip .6 true cm
54:
55: \section{\bf I. Introduction}
56:
57: One-dimensional quantum spin systems have been studied extensively for many
58: years, particularly after Haldane predicted theoretically that Heisenberg
59: integer spin chains should have a gap between the ground state and the first
60: excited state \cite{haldane}, and this was then observed experimentally in a
61: spin-1 system \cite{buyers}. Haldane's analysis used a non-linear
62: $\sigma$-model (NLSM) which is a field theoretic description of the
63: long-distance and low-energy modes of the spin system
64: \cite{affleck1,fradkin,auerbach,sierra,naveen}.
65:
66: Although the NLSM approach is supposed to be accurate only for large values
67: of the spin $S$, it is found to be qualitatively correct even for small values
68: of $S$. For instance, if there is a dimerization in the nearest
69: neighbor Heisenberg couplings, taken to be 1 and $\ka$ alternately, the NLSM
70: predicts that there is a discrete set of values of $\ka$ lying in the range
71: $0 \le \ka \le 1$ for which the spin chain is gapless; these correspond
72: to quantum phase transitions. Further, the
73: number of gapless points is predicted to be the number of integers $\le S +
74: 1/2$; in particular, the undimerized chain (with $\ka =1$) is a gapless point
75: if $S$ is a half-odd-integer. Numerical analysis shows this to be true for
76: values of $S$ up to 2 \cite{kato,pati,kitazawa1,nakamura}; however, the
77: numerically obtained values of $\ka$ at the gapless points do not agree well
78: with the NLSM values. It therefore appears that there must be corrections to
79: the NLSM analysis for small values of $S$.
80:
81: The NLSM approach that has been used so far to find the gapless points is
82: based on certain properties of a field theory in two dimensions (one space and
83: one time); the gapless points occur when the coefficient of a topological term
84: is given by $\pi$ modulo $2 \pi$ (as will be discussed in Sec. II). Although
85: there are arguments justifying this criterion \cite{shankar}, there does not
86: seem to be a simple physical picture behind it. One of the aims of our paper
87: will be to provide a picture based on tunneling between two classical
88: ground states.
89:
90: Numerically, there are different ways of finding the gapless points for a
91: dimerized spin chain. One of the most accurate ways is based on exact
92: diagonalization studies of a finite spin chain with a twisted boundary
93: condition, to be specified more precisely in Sec. II
94: \cite{kitazawa1,kitazawa2}. In the presence of the twist, it is found that
95: the gap between the ground state and first excited state vanishes at a value
96: of $\ka$ which is a function of the number of sites $2N$ (we use this notation
97: since the number of sites will always be taken to be even). We will be mainly
98: interested in finite system sizes in our work; however, it is known from
99: conformal field theory \cite{kitazawa1,kitazawa2,affleck2,okamoto,nomura}
100: that the locations of gapless points for the infinite system can be found very
101: accurately by finding the locations of those points for finite systems, and
102: then extrapolating to $N \to \infty$.
103:
104: In this paper, we will use three different approaches to study finite systems
105: with a twisted boundary condition, in order to find the gapless points as a
106: function of $S$ and $N$. The first two approaches will be analytical; they will
107: be based on the idea that in the presence of a twist, the system has two
108: classical ground states, called N\'eel states, which are degenerate. The
109: degeneracy may be broken in quantum mechanics by tunneling. However, if the
110: tunneling amplitude is zero, we get a gapless point in the sense that the
111: lowest two states have the same energy. The third approach will be numerical
112: and will be based on exact diagonalization of finite systems using various
113: symmetries.
114:
115: In Sec. II, we will first define the dimerized quantum spin chain and review
116: how it can be described using a NLSM field theory. We will then describe the
117: twisted boundary condition for a finite system. In Sec. III, we will describe
118: our first analytical approach. This is based on a tunneling calculation for a
119: chain with a finite number of sites. For reasons explained below, this method
120: is limited to small values of $\ka$. We will see that the expressions for the
121: scale of the gap and the locations of the gapless points agree with those
122: obtained from the NLSM field theory; this is remarkable because our analysis
123: will be based only on quantum mechanical tunneling in a finite system, while
124: the field theoretic analysis is based on a renormalization group equation and
125: the presence of a topological term. Our second approach, described in Sec. IV,
126: is based on a direct quantum mechanical calculation of the matrix element
127: between the two N\'eel states to lowest order in the Hamiltonian; this method
128: works for all values of $\ka$. We will see that the locations of the gapless
129: points obtained by the two methods differ substantially for small $\ka$. We
130: make some speculations about how the tunneling results may be corrected. In
131: Sec. V, we use exact diagonalization to find the gaps and the locations of the
132: gapless points as functions of $\ka$ for different values of $S$ and $N$. We
133: find that the numerical results for the locations of the gapless points agree
134: quite well with those found by the second analytical method, and therefore
135: disagree with those found by the tunneling approach. In Sec. VI, we will
136: summarize our results.
137:
138: \section{\bf II. Dimerized quantum spin chain}
139:
140: \subsection{A. Field theoretic description}
141:
142: In this subsection, we will briefly review the NLSM field theory for a
143: dimerized spin chain with an infinite number of sites. The Hamiltonian is
144: given by
145: \beq
146: H ~=~ \sum_i ~[ {\vec S}_{2i-1} \cdot {\vec S}_{2i} ~+~ \ka ~{\vec S}_{2i}
147: \cdot {\vec S}_{2i+1}] ~,
148: \label{ham1}
149: \eeq
150: where we have spin $S$ at every site, and $\ka \ge 0$. This describes an
151: Heisenberg antiferromagnetic spin chain since all the couplings are positive.
152: We will set $\hbar = 1$, so that ${\vec S}_i^2 = S(S+1)$. (We will specify the
153: appropriate boundary conditions when we discuss finite systems below). In many
154: papers, the nearest neighbor couplings are taken to be $1 + \de$ and $1 - \de$,
155: instead of 1 and $\ka$. After a re-scaling of the Hamiltonian, we see that the
156: two parameters are related as
157: \beq
158: \ka ~=~ \frac{1 - \de}{1 + \de} ~.
159: \eeq
160:
161: In the classical limit $S \to \infty$, the ground state of Eq. (\ref{ham1}) is
162: given by a configuration in which all the spins at odd sites point in the same
163: direction while all the spins at even sites point in the opposite direction.
164: This motivates us to define a variable
165: \beq
166: {\vec \phi} (x) = \frac{{\vec S}_{2n-1} ~-~ {\vec S}_{2n}}{2S} ~,
167: \label{phi}
168: \eeq
169: where $x=2na$ denotes the spatial coordinate and $a$ is the lattice spacing;
170: $x$ becomes a continuous variable in the limit $a \to 0$. In the
171: classical limit, $\vec \phi$ becomes a unit vector in three dimensions; the
172: model is called the NLSM because of this non-linear constraint. One can
173: then derive an action in terms of the field ${\vec \phi} (x,t)$; this takes
174: the form \cite{haldane,affleck1}
175: \bea
176: {\cal S} &=& \int ~dt dx ~[~ \frac{1}{2cg} (\frac{\pa {\vec \phi}}{\pa t} )^2 ~
177: -~ \frac{c}{2g} ( \frac{\pa {\vec \phi}}{\pa x} )^2 ~]~ \non \\
178: & & + ~\frac{\theta}{4 \pi} ~\int ~dt dx ~{\vec \phi} ~\cdot ~
179: \frac{\pa {\vec \phi}}{\pa t} ~\times ~\frac{\pa {\vec \phi}}{\pa x} ~,
180: \label{action}
181: \eea
182: where
183: \bea
184: c &=& 2 a S ~\sqrt{\ka} ~, \non \\
185: g &=& \frac{1}{S} ~\frac{1 + \ka}{\sqrt{\ka}} ~, \non \\
186: {\rm and} \quad \theta &=& 4 \pi S ~\frac{\ka}{1 + \ka} ~.
187: \label{para}
188: \eea
189: The parameters $c$, $g$ and $\theta$ denote the spin wave velocity, the
190: strength of the interactions between the spin waves (even though the first two
191: terms in (\ref{action}) are quadratic, it describes an interacting theory
192: because of the non-linear constraint on $\vec \phi$), and the coefficient of a
193: topological term respectively. One can show that the term multiplying $\theta$
194: in (\ref{action}) is topological in the sense that its value is always an
195: integer.
196:
197: It is known that the system governed by Eq. (\ref{action}) is gapless if
198: $\theta = \pi$ modulo $2 \pi$ and $g$ is less than a critical value
199: \cite{haldane,affleck1,shankar}. This
200: implies that the theory is gapless if $4S \ka/(1+\ka) = 1, 3, \cdots$ and $g$
201: is small enough. In particular, this means that in the range $0 \le \ka \le 1$,
202: there are a discrete set of values of $\ka$ for which the system is gapless;
203: the number of such values is given by the number of integers $\le S +1/2$. For
204: all values of $\theta \ne \pi$ modulo $2 \pi$, the system is gapped. For
205: $\theta = 0$ modulo $2 \pi$, the gap is given by $\exp (-2 \pi /g)$. This
206: follows from the fact that the interaction $g$ effectively becomes a function
207: of the length scale $L$ and satisfies a renormalization group equation of
208: the form $dg_{eff} /d \ln L = g_{eff}^2 /(2\pi)$. This implies that
209: $g_{eff} (L)$ becomes very large at a length scale given by $L_0 \sim ~a
210: \exp (2\pi /g)$, where $g = g_{eff} (a)$ is given in Eq. (\ref{para}). This
211: is the correlation length of the system; the energy gap is related to the
212: inverse of this length, namely, $\Delta E \sim ~c \exp (-2 \pi /g)$.
213:
214: \subsection{B. Twisted boundary condition}
215:
216: In this subsection, we will study the same model as in Eq. (\ref{ham1}) but
217: with a finite number of sites going from 1 to $2N$. Although a periodic
218: boundary condition would appear to be the simplest, it turns out that a more
219: useful boundary condition is one with a twist \cite{kitazawa1,kitazawa2,kolb}.
220: We define the Hamiltonian to be
221: \bea
222: H &=& \sum_{n=1}^N ~[ S_{2n-1}^x S_{2n}^x + S_{2n-1}^y S_{2n}^y + S_{2n-1}^z
223: S_{2n}^z] \non \\
224: &+& \ka ~\sum_{n=1}^{N-1} ~[ S_{2n}^x S_{2n+1}^x + S_{2n}^y S_{2n+1}^y +
225: S_{2n}^z S_{2n+1}^z] \non \\
226: &+& \ka ~[ - S_{2N}^x S_1^x - S_{2N}^y S_1^y + S_{2N}^z S_1^z] ~.
227: \label{ham2}
228: \eea
229: Note that the bond going from site $2N$ to site $1$, to be called the bond
230: $(2N,1)$ for short, has a minus sign for the $xx$ and $yy$ couplings. We will
231: call this a twisted boundary condition; it is equivalent to rotating the $x$
232: and $y$ components of ${\vec S}_1$ by $\pi$ about the $z$ axis just for that
233: bond.
234:
235: An advantage of the twisted boundary condition is that classically, the
236: Hamiltonian in (\ref{ham2}) has exactly two ground states, namely,
237: (i) ${\vec S}_{2n-1} = (0,0,S)$ and ${\vec S}_{2n} = (0,0,-S)$ for all $n$,
238: and (ii) ${\vec S}_{2n-1} = (0,0,-S)$ and ${\vec S}_{2n} = (0,0,S)$ for all
239: $n$. We will call these two N\'eel states $N_1$ and $N_2$ respectively.
240: (This is in contrast to the case of periodic boundary conditions where there
241: is an infinite family of classical ground states because ${\vec S}_{2n-1} =
242: - {\vec S}_{2n}$ can point in any direction). If the classical degeneracy is
243: broken quantum mechanically in any way, there will be a gap between the lowest
244: two states of the system, while if the degeneracy remains unbroken, the system
245: will be gapless. In Secs. III and IV, we will describe two ways of analytically
246: studying whether the degeneracy is broken.
247:
248: Let us now describe the various symmetries of the Hamiltonian in Eq.
249: (\ref{ham2}). Although the total spin is not a good quantum number because of
250: the twist on one bond, $S_{tot}^z = \sum_n (S_{2n-1}^z + S_{2n}^z)$ is a good
251: quantum number.
252:
253: Eq. (\ref{ham2}) satisfies the duality property $H (\ka) = \ka {\tilde H}
254: (1/\ka)$, where $\tilde H$ is related to $H$ by an unitary transformation.
255: [The unitary transformation is required because the twist only exists at the
256: bond $(2N,1)$ whose strength is $\ka$. By applying a rotation $S_1^x \to -
257: S_1^x$ and $S_1^y \to - S_1^y$, we can move the twist from the bond $(2N,1)$
258: to the bond $(1,2)$.] Since a unitary transformation does not affect the
259: spectrum, we conclude that if there is a gapless point at a value $\ka$,
260: there must also be a gapless point at $1/\ka$.
261:
262: Next, we define the parity transformation $P$ as reflection of the system
263: about the bond $(2N,1)$, namely,
264: \beq
265: {\vec S}_i \leftrightarrow {\vec S}_{2N+1-i} ~.
266: \label{parity}
267: \eeq
268: This is a discrete symmetry of the Hamiltonian $H$, and all eigenstates of $H$
269: will be eigenstates of $P$ with eigenvalue $\pm 1$. For any value of $\ka$,
270: we find that the ground and first excited states always have opposite values
271: of the parity. We will show below that the relative parity of the ground
272: states in the limits $\ka \to 0$ and $\ka \to \infty$ is given by
273: $(-1)^{2S}$. For integer $S$, this will imply that in the range $0 < \ka <
274: \infty$, the number of crossings between the ground state and the first
275: excited state (and hence the number of gapless points) must be even. But for
276: half-odd-integer $S$, the number of such crossings must be odd; combined with
277: the duality $\ka \to 1/\ka$, this implies that there must be a crossing
278: and therefore a gapless point at $\ka = 1$ (this is a self-dual point).
279:
280: To prove the statement about the relative parities of the ground states at
281: $\ka \to 0$ and $\ka \to \infty$ being given by $(-1)^{2S}$, we proceed as
282: follows. We first observe that if there are only two spins 1 and 2 governed
283: by the untwisted Hamiltonian $h_u = S_1^x S_2^x + S_1^y S_2^y + S_1^z
284: S_2^z$, then under the reflection $1 \leftrightarrow 2$, the ground state
285: has the parity $(-1)^{2S}$, while the first excited state has the parity
286: $(-1)^{2S+1}$. But for the twisted Hamiltonian $h_t = - S_1^x S_2^x -
287: S_1^y S_2^y + S_1^z S_2^z$, the ground state and first excited
288: state have parities equal to $1$ and $-1$ respectively. (This can be proved
289: using the Perron-Frobenius theorem for a real symmetric matrix). We now
290: consider the entire system with $2N$ sites. In the limit $\ka \to 0$,
291: the ground state of the system is given by a direct product of ground states
292: over the bonds $(1,2)$, $(3,4)$, ..., $(2N-1,2N)$. Since there are $N$ dimers,
293: the parity of this state is $(-1)^{2SN}$, while the parity of the first
294: excited state is $(-1)^{2SN+1}$. On the other hand, in the limit $\ka \to
295: \infty$, the ground state of the system is given by a direct product of ground
296: states over the bonds $(2,3)$, $(4,5)$, ..., $(2N,1)$. Under parity,
297: the parity of this state is $(-1)^{2SN+2S}$, while the parity of the first
298: excited state is $(-1)^{2SN+2S+1}$. A comparison between the ground states
299: in the two limits shows that they have a relative parity of $(-1)^{2S}$.
300:
301: \section{\bf III. Tunneling approach to the finite spin chain}
302:
303: In this section, we will study the model defined in Eq. (\ref{ham2}) using a
304: tunneling approach. We are interested in the limit $S \to \infty$ and $\ka \to
305: 0$, such that $\ka S$ is of order 1. We will compute the action of the system
306: and use that to compute the tunneling amplitude between the two N\'eel states.
307:
308: In the limit $\ka \to 0$, the system consists of decoupled dimers whose energy
309: levels are easy to compute. (We will assume in this section that there are at
310: least three dimers, i.e, $N \ge 3$). For the dimer on the bond $(2n-1,2n)$, we
311: define the variables
312: \bea
313: {\vec \phi}_n &=& \frac{{\vec S}_{2n-1} ~-~ {\vec S}_{2n}}{2S} ~,
314: \non \\
315: {\vec l}_n &=& {\vec S}_{2n-1} ~+~ {\vec S}_{2n} ~.
316: \label{phil1}
317: \eea
318: These variables satisfy the relations
319: \bea
320: {\vec \phi}_n^2 &=& 1 ~+~ \frac{1}{S} ~-~ \frac{{\vec l}_n^2}{4S^2} ~, \non \\
321: {\vec \phi}_n ~\cdot ~{\vec l}_n &=& 0 ~, \non \\
322: \left[ l_m^a ~,~ \phi_n^b \right] &=& i ~\de_{mn} ~\sum_c ~\ep^{abc} ~
323: \phi_n^c ~, \non \\
324: \left[ l_m^a ~,~ l_n^b \right] &=& i ~\de_{mn} ~\sum_c ~\ep^{abc} ~l_n^c ~,
325: \non \\
326: {\rm and} \quad \left[ \phi_m^a ~,~ \phi_n^b \right] &=& \frac{i}{4S^2} ~
327: \de_{mn} ~\sum_c ~\ep^{abc} ~l_n^c ~.
328: \label{phil2}
329: \eea
330:
331: Several simplifications occur in the classical limit $S \to \infty$. Firstly,
332: in the N\'eel state, $\vec l$ is equal to zero while $\vec \phi$ is an unit
333: vector; the latter is clear from the first equation in (\ref{phil2}). We will
334: therefore set ${\vec \phi}_n^2 = 1$ exactly. Secondly, we will take the
335: commutator $[ \phi_m^a , \phi_n^b ] = 0$ due to the fourth equation in
336: (\ref{phil2}). Finally, given the second and third equations in (\ref{phil2}),
337: we can obtain the momentum which is canonically conjugate to $\phi_n$, namely,
338: ${\vec l}_n = {\vec \phi}_n ~\times ~ {\vec \Pi}_n$, which satisfies the
339: commutation relation
340: \beq
341: \left[ \phi_m^a ~,~ \Pi_n^b \right] ~=~ i ~\de_{mn} ~\de_{ab} ~.
342: \eeq
343:
344: Using Eq. (\ref{phil1}), we can write the Hamiltonian in (\ref{ham2}) in
345: terms of $\vec \phi$ and $\vec \Pi$, and then obtain the Lagrangian as
346: \beq
347: L ~=~ \sum_{n=1}^N ~\frac{d {\vec \phi}_n}{d t} ~\cdot ~
348: {\vec \Pi}_n ~-~ H ~.
349: \eeq
350: We eventually find that
351: \bea
352: L &=& \sum_{n=1}^N ~\frac{1}{2} ~( \frac{d {\vec \phi}_n}{d t} )^2 ~
353: +~ \ka S^2 ~\sum_{n=1}^{N-1} ~{\vec \phi}_n \cdot {\vec \phi}_{n+1}
354: \non \\
355: & & +~ \ka S^2 [~- \phi_N^x \phi_1^x ~-~ \phi_N^y \phi_1^y ~+~ \phi_N^z
356: \phi_1^z ~] \non \\
357: & & + ~\frac{\ka S}{2} ~\sum_{n=1}^{N-1} ~(\frac{d{\vec \phi}_n}{dt} ~+~
358: \frac{d{\vec \phi}_{n+1}}{dt}) \cdot {\vec \phi}_n \times {\vec \phi}_{n+1}
359: \non \\
360: & & + ~\frac{\ka S}{2} ~[~~ (\frac{d \phi_N^x}{dt} ~-~ \frac{d \phi_1^x}{dt})
361: (\phi_N^y \phi_1^z ~+~ \phi_N^z \phi_1^y) \non \\
362: & & ~~~~~~~ + ~(\frac{d \phi_N^y}{dt} ~-~ \frac{d \phi_1^y}{dt})
363: (-\phi_N^z \phi_1^x ~-~ \phi_N^x \phi_1^z) \non \\
364: & & ~~~~~~~ + ~(\frac{d \phi_N^z}{dt} ~+~ \frac{d \phi_1^z}{dt})
365: (-\phi_N^x \phi_1^y ~+~ \phi_N^y \phi_1^x) ~] \non \\
366: & & +~ \ka^2 S^2 ~[ {\rm terms ~of ~fourth ~order ~in ~{\vec \phi}}_n ]~.
367: \label{lag}
368: \eea
369: The terms in the third line of this Lagrangian are what give rise to the
370: topological term in Eq. (\ref{action}) in the continuum limit. In the last
371: line of (\ref{lag}), the terms of fourth order in ${\vec \phi}_n$ are chosen
372: in such a way that when we compute the Hamiltonian from it, it agrees with Eq.
373: (\ref{ham2}). We will now see why these fourth order terms are not important.
374:
375: In the limit $\ka \to 0$, $S \to \infty$ and $\ka S$ of order 1, we can scale
376: the time $t$ by a factor of $\sqrt S$ to show that only the first two lines of
377: Eq. (\ref{lag}) contribute to the Euler-Lagrange equations of motion (EOM);
378: this is a major simplification. To compute the tunneling amplitude between the
379: two N\'eel states, we will find the solutions of the EOM in imaginary time.
380: We then find that the tunneling amplitude comes with a phase which arises from
381: the third through sixth lines of Eq. (\ref{lag}); thus these terms are
382: important even though they do not directly contribute to the EOM. The fourth
383: order terms in the last line of (\ref{lag}) do not contribute to either the
384: EOM or the phase, and we will therefore ignore them henceforth.
385:
386: In imaginary time (denoted by the symbol $\tau$), the action takes the form
387: \bea
388: {\cal S}_I &=& \int d\tau ~[ \sum_{n=1}^N ~\frac{1}{2} ~( \frac{d
389: {\vec \phi}_n}{d \tau} )^2 + \ka S^2 (N - \sum_{n=1}^{N-1} ~{\vec \phi}_n
390: \cdot {\vec \phi}_{n+1}) \non \\
391: & & ~~~~~~~ +~ \ka S^2 [~ \phi_N^x \phi_1^x ~+~ \phi_N^y \phi_1^y ~-~ \phi_N^z
392: \phi_1^z ~] \non \\
393: & & -i ~\frac{\ka S}{2} ~\sum_{n=1}^{N-1} ~(\frac{d{\vec \phi}_n}{d\tau} ~+~
394: \frac{d{\vec \phi}_{n+1}}{d\tau}) \cdot {\vec \phi}_n \times {\vec \phi}_{n+1}
395: \non \\
396: & & -i ~\frac{\ka S}{2} ~[~~ (\frac{d \phi_N^x}{d\tau} ~-~ \frac{d
397: \phi_1^x}{d\tau}) (\phi_N^y \phi_1^z ~+~ \phi_N^z \phi_1^y) \non \\
398: & & ~~~~~~~ + ~(\frac{d \phi_N^y}{d\tau} ~-~ \frac{d \phi_1^y}{d\tau})
399: (-\phi_N^z \phi_1^x ~-~ \phi_N^x \phi_1^z) \non \\
400: & & ~~~~~~~ + ~(\frac{d \phi_N^z}{d\tau} ~+~ \frac{d \phi_1^z}{d\tau})
401: (-\phi_N^x \phi_1^y ~+~ \phi_N^y \phi_1^x) ~] ~] ~. \non \\
402: & &
403: \label{si}
404: \eea
405: (We have introduced a constant $\ka S^2 N$ so that the action vanishes for
406: each of the two N\'eel states). The tunneling amplitude will be given by the
407: sum of $\exp (-{\cal S}_I)$ along all the paths of extremal action which join
408: the N\'eel states. We will now determine these extremal paths.
409:
410: Let us use polar angles to write the variables ${\vec \phi}_n =
411: (\sin \al_n \cos \be_n, \sin \al_n \sin \be_n , \cos \al_n)$. The N\'eel
412: states 1 and 2 are given by $\al_n = 0$ for all $n$ and $\al_n = \pi$ for
413: all $n$ respectively. We now solve the EOM following from the first two lines
414: of Eq. (\ref{si}) in order to obtain the paths going from state 1 to state 2.
415: We will not write the EOM explicitly here, but directly present the solutions.
416: We find that the two paths which have the least action are given by
417:
418: \noi A: ~$\al_n (\tau) = \al (\tau)$ and $\be_n = \be_0 + (n \pi /N)$ for
419: all $n$, and
420:
421: \noi B: ~$\al_n (\tau) = \al (\tau)$ and $\be_n = \be_0 - (n \pi /N)$ for
422: all $n$,
423:
424: \noi where $\be_0$ is an arbitrary angle.
425: In both cases, the function $\al (\tau)$ satisfies the EOM
426: \beq
427: \frac{d^2 \al}{dt^2} ~=~ \ka S^2 ~\sin (2 \al) ~(1 ~-~ \cos \frac{\pi}{N}) ~,
428: \eeq
429: with the boundary conditions $\al (-\infty) =0$, $\al (\infty) =\pi$, and
430: $d \al (\pm \infty) /dt = 0$. This implies that
431: \beq
432: \frac{d\al}{dt} ~=~ 2 S \sqrt{\ka} ~\sin \al ~\sin \frac{\pi}{2N} ~.
433: \eeq
434: Using this we can evaluate the contribution of the first two lines of
435: (\ref{si}) along either one of the paths joining the N\'eel states. We
436: find that the real part of the action is given by
437: \bea
438: {\rm Re} ~{\cal S}_I &=& N \int_{-\infty}^\infty d\tau ~[\frac{1}{2} ~
439: (\frac{d\al}{d\tau})^2 + \ka S^2 \sin^2 \al ~(1 - \cos \frac{\pi}{N})] \non \\
440: &=& 4 \sqrt{\ka} S N ~\sin \frac{\pi}{2N} ~.
441: \label{real}
442: \eea
443:
444: We can now evaluate the contribution of the imaginary terms in Eq. (\ref{si})
445: to the action. We find that for path A, they contribute $-i2\ka S N \sin(\pi
446: /N)$, while for path B, they contribute $i2\ka S N \sin(\pi /N)$. Hence, the
447: total contribution of the two paths to $\exp (-{\cal S}_I)$ is given by
448: \beq
449: \Delta ~\sim~ \cos (2\ka S N \sin \frac{\pi}{N}) ~\exp (-4 \sqrt{\ka} S N ~
450: \sin \frac{\pi}{2N}) ~,
451: \label{tungap}
452: \eeq
453: up to some pre-factors which are determined by fluctuations about the
454: classical paths. Since $\Delta$ is the matrix element between two classically
455: degenerate states, the energy gap between the two states is given by $2
456: |\Delta|$. We thus see that the gap vanishes if $2\ka S N \sin (\pi /N)$
457: is an odd multiple of $\pi /2$, i.e., if
458: \beq
459: 4\ka S N \sin \frac{\pi}{N} ~=~ \pi {\rm ~modulo~} 2 \pi ~.
460: \label{gapless}
461: \eeq
462: This is the same condition as the one satisfied by the parameter $\theta$ in
463: Sec. II A in the limits $S, N \to \infty$ and $\ka \to 0$. Further, if $4\ka
464: S N \sin (\pi /N) = 0$ modulo $2 \pi$, the gap is given by $\exp (-{\rm
465: Re} ~ {\cal S}_I)$ which agrees with the expression $\exp (-2\pi /g)$ given in
466: Sec. II A for $S, N \to \infty$ and $\ka \to 0$. Thus, a simple quantum
467: mechanical tunneling calculation seems to reproduce the same conditions as
468: those obtained earlier by more complex field theoretic calculations involving
469: topological terms and a renormalization group analysis.
470:
471: Before ending this section, we should note that there are other pairs of paths
472: with extremal action, which have the form
473:
474: \noi A: ~$\al_n (\tau) = \al (\tau)$ and $\be_n = \be_0 + (pn \pi /N)$ for
475: all $n$, and
476:
477: \noi B: ~$\al_n (\tau) = \al (\tau)$ and $\be_n = \be_0 - (pn \pi /N)$ for
478: all $n$,
479:
480: \noi where $p = 3, 5, \cdots$ (going up to the largest odd integer $\le N-1$)
481: labels the different pairs of paths. However, the real part of the action of
482: these paths is given by $4 \sqrt{\ka} S N ~\sin (p\pi /2N)$, which, for large
483: $S$, is much larger than the expression given in Eq. (\ref{real}); their
484: contributions to the tunneling amplitude are therefore much smaller.
485:
486: Finally, we would like to note that it is important that the twist
487: in the boundary condition should be by $\pi$, and not by any other angle.
488: Even though any non-zero twist would lead to two N\'eel ground states
489: classically, the pairs of tunneling paths between those two ground states
490: would not have the same real part of the action if the twist angle was
491: different from $\pi$. The pairs of paths would therefore not cancel each
492: other no matter what the imaginary parts of their actions are. (Two complex
493: numbers cannot add up to zero, no matter what their phases are, if their
494: magnitudes are not equal).
495:
496: \section{\bf IV. A second approach to the matrix element between N\'eel
497: states}
498:
499: In the previous section we argued that the gapless points of the Hamiltonian
500: in Eq. (\ref{ham2}) can be identified with the values of $\ka$ for which
501: the tunneling amplitude between the two classical ground states vanish. In
502: this section we will calculate this amplitude using an alternate method.
503:
504: The twist on the edge bond $(2N,1)$ breaks the global $SU(2)$ symmetry and thus
505: lifts the continuous degeneracy of the classical ground states. With the twist,
506: the two degenerate ground states of the Hamiltonian are the N\'eel states
507: which are connected to each other by rotation by $\pi$ about the $y$ axis,
508: \bea
509: | N_1 \rangle &=& |S, -S, S, \cdots, -S, S, -S \rangle ~, \non \\
510: {\rm and} \quad | N_2 \rangle &=& |-S, S, -S, \cdots, S, -S, S \rangle ~,
511: \label{neel1}
512: \eea
513: where $| \{m_i\} \rangle $ denotes the state with $S^z_i$ eigenvalue $m_i$.
514:
515: We are interested in the zeroes of the quantity
516: \beq
517: T = \langle N_2 | e^{-\beta H} |N_1 \rangle ~,
518: \label{transamp}
519: \eeq
520: as a function of $\ka$. Though the calculation of the above matrix element
521: is an interacting many-body problem, we can obtain its zeroes `exactly' in
522: the thermodynamic limit.
523:
524: First we note that, in the expansion of the exponential,
525: \beq
526: e^{-\beta H} = \sum_{n=0}^\infty ~\frac{(-\beta H)^n}{n!}~,
527: \eeq
528: the first term which makes a non-zero contribution to $T$ has $n=2SN$. This
529: is because to take $|N_1 \rangle$ to $|N_2 \rangle$, spins belonging to the
530: $A$-sublattice have to be flipped from $|S \rangle$ to $|-S\rangle$, and
531: this requires the action of $(S^-)^{2S}$ for each spin. Similarly, the action
532: of $(S^+)^{2S}$ will take spins in the $B$-sublattice from $|-S\rangle$ to
533: $|S\rangle$.
534:
535: Next, we will calculate the values of $\ka$ for which
536: \beq
537: \langle N_2 | H^{2SN} |N_1 \rangle = 0 ~.
538: \label{leadmat}
539: \eeq
540: Then we will show that as $N \to \infty$, Eq. (\ref{leadmat}) implies
541: that
542: \beq
543: \langle N_2 | H^{2SN+k} |N_1 \rangle = 0
544: \label{subleadmat}
545: \eeq
546: for any finite $k$. This in turn will imply that $T$ is zero.
547:
548: The only term in $H^{2SN}$ which makes a non-zero contribution to the left
549: hand side of Eq. (\ref{leadmat}) is
550: \[ \prod_{i=1}^N (S_{2i}^-)^{2S}(S_{2i+1}^+)^{2S}. \]
551: We need to count the number of ways in which such a term can arise. The
552: contribution comes from terms of the following type,
553: \beq
554: \prod_{i=1}^N (S_{2i}^-~S_{2i+1}^+)^{m}(S_{2i-1}^+ ~ S_{2i}^-)^{2S-m} ~,
555: \label{nonzero}
556: \eeq
557: where $0 \le m \le 2S$. The above term can be obtained in
558: \[\frac{(2SN)!}{(m!)^N \big((2S-m)!\big)^N} \] ways and comes with a
559: weight $(-1)^m \ka^{mN}$. Here we have neglected an overall
560: $m$-independent factor due to the Clebsch-Gordon coefficients arising
561: from the repeated application of $S^+$ and $S^-$ operators. The
562: condition in Eq. (\ref{leadmat}) then becomes,
563: \bea
564: \sum_{m=0}^{2S} ~(-1)^m ~\Big(a_m(\ka) \Big)^N &=& 0 ~, \non \\
565: {\rm where} \quad a_m(\ka) &=& \frac{\ka^m}{m!(2S-m)!} ~.
566: \label{roots}
567: \eea
568: Before proceeding further, we note that the above condition preserves the
569: duality symmetry under $\ka \to 1/\ka$. This is because
570: Eq. (\ref{roots}) can be written as,
571: \beq
572: \ka^{2SN} (-1)^{2S} \sum_m (-1)^m \Big(a_m(1 / \ka)\Big)^N = 0 ~.
573: \label{dualroots}
574: \eeq
575: Hence, if $\ka^*$ is a solution, so is $1/\ka^*$. Thus we can restrict
576: ourselves to the range $0 \le \ka \le 1$.
577:
578: Eq. (\ref{roots}) determines the roots of a polynomial of order $2SN$, which,
579: in general, cannot be solved analytically. But it turns out that we can obtain
580: the roots in the limit $N \to \infty$. In this limit, depending on the value
581: of $\ka$, one particular term in the sum is predominant, and the rest of the
582: terms can be neglected compared to this. The dominant term is determined by,
583: \beq
584: \max_m ~a_m \equiv a_{m^*} ~.
585: \label{dominant}
586: \eeq
587: As $\ka$ varies from $0$ to $1$, $m^*$ successively takes values
588: $m^* = 0, 1, \cdots , S$ for integer $S$ and $m^* = 0, 1,
589: \cdots , S+1/2$ for half-odd-integer $S$. Noting that neighboring terms in
590: the sum have opposite signs, Eq. (\ref{roots}) can be satisfied only when
591: \beq
592: a_m = a_{m+1} ~.
593: \label{roots1}
594: \eeq
595: Thus the gapless points are given by
596: \beq
597: \ka^*_m = \frac{m+1}{2S-m} ~,
598: \label{gaplesspts}
599: \eeq
600: where, $m = 0, 1, \cdots , S-1$ for integer $S$ and $m = 0, 1, \cdots , S-1/2$
601: for half-odd-integer $S$.
602:
603: To complete our argument that Eq. (\ref{gaplesspts}) gives the gapless points,
604: we need to show that Eq. (\ref{leadmat}) implies Eq. (\ref{subleadmat}). To
605: this end, we first note that the non-zero contributions to the matrix element
606: from $H^{2SN+k}$ can be obtained from the contributing terms in $H^{2SN}$,
607: given in Eq. (\ref{nonzero}), by adding terms of the form $S_i^z S_{i+1}^z$,
608: $S_i^+ S_{i+1}^-$ or $S_i^- S_{i+1}^+$. Now the weight coming from the
609: Clebsch-Gordon coefficients depend on the order in which the terms appear.
610: But formally one can write that, Eq. (\ref{subleadmat}) implies
611: \beq
612: \sum_m (-1)^m b_{m,k} \Big(a_m(\ka) \Big)^N = 0 ~,
613: \label{nonzerok}
614: \eeq
615: where $b_{m,k}$ are finite undetermined constants independent of $N$. As
616: before, in the large-$N$ limit, the left hand side of Eq. (\ref{nonzerok})
617: can be zero only through the mutual cancellation of a pair of neighboring
618: terms, {\it i.e.}, when
619: \beq
620: \frac{a_m}{a_{m+1}} = \left( \frac{b_{m+1}}{b_{m}} \right)^{1/N} ~.
621: \eeq
622: As $N \to \infty$, this reduces to the condition in Eq. (\ref{roots1}).
623: In other words, the vanishing of $\langle N_2 | H^{2SN} | N_1 \rangle$ is a
624: sufficient condition for the vanishing of $\langle N_2 | H^{2SN+k} | N_1
625: \rangle$ for any finite $k$ as $N \to \infty$.
626:
627: For half-odd-integer spins, $\ka^* = 1$ is a solution of Eq. (\ref{gaplesspts})
628: for $m = S-1/2$, but it is {\it not} a solution for integer spins. This is
629: consistent with Haldane's conjecture \cite{haldane} that the uniform chain is
630: gapless for half-odd-integer spins and gapped for integer spins.
631:
632: Since the identification of gapless points with the zeroes of the transition
633: amplitudes between the two N\'eel states is essentially a semi-classical
634: approximation, we expect the formula given by Eq. (\ref{gaplesspts}) to get
635: better for larger values of $S$.
636:
637: As with the tunneling calculation in Sec III, here also one can see that it
638: is crucial to have a twist by $\pi$ and not any other angle. Let us suppose
639: that the twist angle is $\chi$. Then the Hamiltonian for the bond between the
640: spins at sites $2N$ and $1$ will be
641: \[ \ka (S_{2N}^z S_1^z + e^{i\chi}~ S_{2N}^+ S_1^- + e^{-i\chi}~
642: S_{2N}^- S_1^+). \]
643: Then the equivalent of the condition in Eq. (\ref{roots1}) will read
644: \beq
645: a_m = -e^{i\chi} a_{m+1}.
646: \label{roots2}
647: \eeq
648: Since $a_m$'s are all real and positive, such a condition can be satisfied only when
649: $\chi = \pi$, in which case Eq. (\ref{roots2}) becomes
650: Eq. (\ref{roots1}).
651:
652: Finally, let us compare the gapless values of $\ka$ given in Eq.
653: (\ref{gaplesspts}) with those given in Eq. (\ref{gapless}) in the limit
654: $N \to \infty$, namely,
655: \beq
656: \ka^*_m = \frac{m+1/2}{2S} ~,
657: \label{gapless2}
658: \eeq
659: where, $m = 0, 1, \cdots$. [Since Eq. (\ref{gapless}) was derived under the
660: assumption that $S \to \infty$ and $\ka S$ is of order 1, we must restrict $m$
661: to be much less than $S$ in Eq. (\ref{gapless2}).] We see that for large
662: values of $S$ and $m \ll S$, the values of $\ka^*_m$ in Eqs. (\ref{gaplesspts})
663: and (\ref{gapless2}) are related by a shift of $1/(4S)$. It
664: would be useful to understand more deeply why this is so. Heuristically, this
665: discrepancy can be explained by postulating that the phase difference between
666: the actions for the two paths discussed in Sec. III has an additional factor
667: of $\pi$ for some reason. Eq. (\ref{gapless}) would then change to $4\pi \ka
668: S = 0$ modulo $2 \pi$ for $N \to \infty$; this condition would be equivalent
669: to Eq. (\ref{gaplesspts}) for $m \ll S$. In a different problem (tunneling
670: of a charged particle in two dimensions in the presence of a large magnetic
671: field), it was empirically found that an additional factor of $\pi$ appears
672: due to fluctuations about the tunneling paths \cite{jain}. It may
673: be worth studying if something similar happens in our problem.
674:
675: \section{V. Numerical results for finite systems}
676:
677: In this section, we numerically determine the values of $\ka$ for which the
678: Hamiltonian becomes gapless using exact diagonalization
679: of finite systems. These results, being a direct calculation of the gapless
680: points, will give us information about the physical regimes in which the
681: analytical methods outlined in Secs. III and IV are valid.
682:
683: We begin with a brief outline of the method used to find the gapless points.
684: As we vary $\ka$ in Eq. (\ref{ham1}), it is known that the gapless points
685: separate various phases whose ground states are represented approximately by
686: different valence bond states (see Refs. \cite{kitazawa1,oshikawa} and
687: references therein). Specifically, for a given spin $S$, there are $2S+1$
688: different phases separated by the $2S$ gapless points for $\ka$ between $0$
689: and $\infty$. The lowest two eigenstates of an untwisted Hamiltonian never
690: cross each other in energy, even at the transition from one phase to the other;
691: also, the ground state always has the same eigenvalue $(-1)^{2SN}$ for the
692: parity $P$. It is here that we make use of the twisted boundary condition in
693: Eq. (\ref{ham2}). It has been shown \cite{kitazawa1} that the lowest two
694: eigenstates of this Hamiltonian (both of which lie in the $S_{tot}^z=0$ sector)
695: have different parity eigenvalues, and they cross at certain points which, in
696: the limit $N \to \infty$, become the gapless points of the Hamiltonian in
697: Eq. (\ref{ham1}). This means
698: that one can locate the gapless points of the Hamiltonian in Eq. (\ref{ham1})
699: by studying the crossing of the two lowest eigenvalues of the Hamiltonian in
700: Eq. (\ref{ham2}) in {\it different parity sectors}. This enables us to
701: find the gapless points without having to consider two very closely spaced
702: eigenvalues lying within a single symmetry sector.
703:
704: We study the Hamiltonian in Eq. (\ref{ham2}) numerically. Because of the
705: duality between $\ka$ and $1/\ka$ we restrict our studies to $0~\leq~\ka~
706: \leq 1$. We use the Lanczos algorithm to diagonalize finite systems from $N=3$
707: to $N=5$. Spins from $S=2$ to 3 are studied for $N=3$ to $N=5$, but for
708: $S = 3.5$ to 7 we restrict ourselves to $N=3$ and $N=4$ due to computational
709: limitations.
710:
711: \begin{figure}[htb]
712: \begin{center}
713: \epsfig{figure=fig1.ps,width=8cm}
714: \vskip .2 true cm
715: \epsfig{figure=fig2.ps,width=8cm}
716: \end{center}
717: \caption{Energy difference between the two lowest energies in the $P=-1$ and
718: $P=+1$ sectors as a function of $\ka$, for $N=3$. The upper figure is for
719: $S=3.5$, and the lower figure is for $S=4$. The locations of the gapless
720: points are shown.}
721: \end{figure}
722:
723: In Fig. 1, we present two representative data plots, for $N=3$.
724: The upper plot is for $S=3.5$ and the lower one is
725: for $S=4$. Plotted on the $y$ axis in both figures is $E(P=-1) - E(P=+1)$,
726: the energy difference between the two lowest energy eigenstates in the
727: two parity sectors. Wherever the plot crosses the $x$ axis we have a gapless
728: point of the Hamiltonian in Eq. (\ref{ham1}). The values of the crossing
729: points are indicated in the plots. From the discussion in Sec. II, we know
730: that the ground state for $N=3$, $S=3.5$ should have a parity $-1$, and for
731: $N=3$, $S=4$ should have a parity $+1$. This is indeed borne out by the plots
732: in the figure. Moreover, the gapless point at $\ka=1$ is also present for
733: $S=3.5$ as expected. The nature of these plots for other spins and different
734: lattice sizes is similar. Since our emphasis in this work is on the locations
735: of the gapless points, we now turn to analyzing those points more closely.
736: [Incidentally, we observe in Fig. 1 that the envelope of the magnitude of the
737: gap is rapidly decreasing with increasing $\ka$; this is in accordance with
738: the exponential factor for the tunneling amplitude in Eq. (\ref{tungap}).]
739:
740: \begin{figure}[htb]
741: \begin{center}
742: \epsfig{figure=fig3.ps,width=8cm}
743: \end{center}
744: \caption{The location $\ka^1$ of the gapless point closest to zero, as a
745: function of the spin, for $N=3$. The results from numerical calculations
746: (dots), N\'eel state calculations (crosses) and tunneling calculations
747: (squares) are shown. The joining lines are guides for the eye. The inset shows
748: the percentage variation of the N\'eel state calculations when compared with
749: the numerical results.}
750: \end{figure}
751:
752: \begin{figure}[htb]
753: \begin{center}
754: \epsfig{figure=fig4.ps,width=8cm}
755: \end{center}
756: \caption{The location $\ka^*$ of the gapless point closest to (but less than)
757: 1, as a function of the spin, for $N=3$. The results from numerical
758: calculations (dots) and N\'eel state calculations (crosses) are shown. The
759: top and bottom parts of the figure are for integer and half-odd-integer
760: values of the spin respectivly. The joining lines are guides for the eye. The
761: insets shows the percentage variation of the N\'eel state calculations when
762: compared with the numerical results.}
763: \end{figure}
764:
765: Figure 2 shows a comparison of the different methods used to calculate the
766: gapless points. Plotted on the $y$ axis is $\ka^1$, the gapless point closest
767: to the origin for various values of the spin, for $N=3$. The topmost plot
768: (marked by dots) is for values obtained from the numerical calculations,
769: called $\ka^1_{num}$. The next plot (crosses) is for values
770: obtained from the N\'eel state calculation in Eq. (\ref{roots}),
771: $\ka^1_{Neel}$. The plot at the bottom (squares) is for values
772: obtained from the tunneling expression in Eq. (\ref{gapless}), $\ka^1_{tun}$.
773: Clearly, the values of $\ka^1$ obtained using the N\'eel state calculation
774: are in closer agreement with the numerical values than the tunneling values.
775: The tunneling values do not converge with increasing $S$, even
776: though one is looking at data for spin values as large as $S=7$.
777: On the other hand, the plot of values using
778: the N\'eel state calculations converges much faster. This is made clearer
779: in the plot by the inset where on the $y$ axis we have plotted the percentage
780: deviation, ($(\ka^{1}_{num}-\ka^{1}_{Neel})/\ka^{1}_{num}) \times 100$,
781: of the values of $\ka^1_{Neel}$ from the numerically obtained values.
782:
783: Figure 3 shows a comparison between the numerical results (dots)
784: and those obtained from the N\'eel state calculation (crosses) for
785: $\ka^{*}$, the gapless point closest to $\ka=1$. [Unlike Fig. 2, we have not
786: shown the tunneling results based on Eq. (\ref{gapless})) because that
787: formula for the gapless points is not valid when $\ka$ is close to 1.]
788: The data sets for integer and half-odd-integer spins have
789: been plotted separately. The plot and the inset at the top are for integer
790: spins, and the ones at the bottom are for half-odd-integer spins. As before,
791: we see that the agreement with the numerical results improves as we go to
792: larger spins. We also see from the insets that for a given spin, the agreement
793: is much better near $\ka=1$ than it was near $\ka=0$. This suggests that the
794: N\'eel state calculation gets better as we increase $\ka$ from 0 to 1.
795: This is something which is seen very clearly in the following tables.
796:
797: \begin{center}
798: \begin{tabular}{|c|c|c|c|c|} \hline
799: $\ka_{num}$ & $\ka_{Neel}$ & \% deviation& $\ka_{tun}$ & \% deviation \\
800: \hline
801: 0.083 & 0.080 & 3.6\% &0.047 &45.4\% \\
802: 0.190 & 0.180 & 5.2\% & 0.140 &26.3\% \\
803: 0.306 & 0.293 & 4.2\% &0.232 & 24.2\% \\
804: 0.438 & 0.424 & 3.19\% &0.326 &25.6\% \\
805: 0.773 & 0.765 & 1.03\%&0.512 &33.8\% \\
806: 1& 1& 0\%& & \\ \hline
807: \end{tabular}
808: \end{center}
809:
810: \begin{center}
811: \begin{tabular}{|c|c|c|c|c|} \hline
812: $\ka_{num}$ & $\ka_{Neel}$ & \% deviation& $\ka_{tun}$ &
813: \% deviation \\ \hline
814: 0.077 & 0.074 & 3.89\% &0.043 &44.2\% \\
815: 0.174 & 0.166 & 4.02\% & 0.130 &25.3\% \\
816: 0.281 & 0.269 & 4.98\% &0.252 & 23.1\% \\
817: 0.400 & 0.387 & 3.25\% &0.302 &24.5\% \\
818: 0.536 & 0.525 & 2.05\%&0.389 &27.4\% \\
819: 0.695& 0.685& 1.29\%&0.475 & 31.7\% \\
820: 0.887& 0.884& 0.34\%&0.561& 36.8\% \\\hline
821: \end{tabular}
822: \end{center}
823:
824: \noi Table 1. Comparison of the values of all the gapless points obtained
825: using the three methods of calculating them, for $S=6.5$ (top) and 7 (bottom).
826: The data presented is for $N=3$.
827: \vspace{0.25cm}
828:
829: Table 1 shows how the numerical results, the N\'eel state calculations and the
830: tunneling results compare for all the gapless points. The table shows the
831: values of $\ka$ at which the Hamiltonian in Eq. (\ref{ham1}) is gapless for
832: $S=6.5$ (top) and $S=7$ (bottom), for $N=3$. The percentage deviations as
833: defined earlier are also shown for the N\'eel state
834: and tunneling calculations. As conjectured after looking at Fig. 3, we
835: see that the N\'eel state calculation gives successively better approximations
836: to the actual gapless points as we go further from the origin $\ka = 0$. We
837: have shown the tunneling values for all the gapless points only for
838: completion; the formula given by Eq. (\ref{gapless}) is valid only
839: for $\ka S$ of order 1. We again see that these values have much
840: larger percentage deviations from the values obtained from the other two
841: methods, even though the values of the spins considered are quite large.
842:
843: \begin{figure}[htb]
844: \begin{center}
845: \epsfig{figure=fig5.ps,width=8cm}
846: \end{center}
847: \caption{Variation of $\ka^1$ with $N$ for $S=2.5$, $3$, $3.5$, and $4$. The
848: numbers at the left of each graph are the $N \to \infty$ extrapolated values.}
849: \end{figure}
850:
851: We now look at how the values of $\ka$ at the gapless points change with $N$
852: and see how the $N \rightarrow \infty$ values compare with those given by Eq.
853: (\ref{gaplesspts}). We take $\ka^1$ as an example. Figure 3 shows the behavior
854: for $S=2.5, 3, 3.5$, and 4. We find the $N \rightarrow \infty$ values by
855: extrapolating the best fits obtained by fitting the data to even polynomials
856: in $1/(2N)^2$ following Ref. \cite {nakamura}. All the data are for $N=2,3,4$
857: and 5 (for $S=2.5$ and 3).
858:
859: The extrapolated values of $\ka^1$ in the $N \to \infty$ limit and the
860: corresponding values obtained from Eq. (\ref{gaplesspts}) (in parentheses)
861: for $S=2.5$, $3.0$, $3.5$ and $4.0$ are given by $0.223 ~(0.200)$,
862: $0.179 ~(0.167)$, $0.146 ~(0.143)$ and $0.124 ~(0.125)$ respectively.
863: Clearly, the agreement between the two sets of values gets better for
864: larger values of the spin.
865:
866: \section{\bf VI. Conclusions}
867:
868: We have used three different techniques to find the gapless points of a
869: dimerized spin-$S$ chain with a finite number of sites and
870: with a twisted boundary
871: condition. The first technique uses a tunneling approach which is expected to
872: be valid in the limit $S \to \infty$, $\ka \to 0$ and $\ka S$ of order 1.
873: Remarkably, we find that a quantum mechanical tunneling calculation reproduces
874: the same expressions for the locations of the gapless points and the gap as
875: those obtained by more involved field theoretic techniques.
876:
877: However, a direct numerical study of the gapless points shows a systematic
878: deviation from the tunneling results in the limit discussed above. It would be
879: useful to know why the tunneling results differ systematically from the
880: numerical results in this limit. One possible idea is to examine if an
881: additional factor of $\pi$ appears in the fluctuation pre-factor of the
882: tunneling amplitude as mentioned at the end of Sec. IV.
883:
884: In view of the discrepancy between the tunneling and numerical results, we have
885: presented a second analytical derivation of the gapless points which is based
886: on a calculation of the matrix element between the two N\'eel states to lowest
887: order in powers of the Hamiltonian; this derivation is expected to become more
888: accurate as the number of sites becomes large. We find that the results
889: obtained by this approach agree much better with the numerical results than
890: the tunneling results, even in the limit $\ka \to 0$. It may be instructive to
891: understand in more detail why there is such a good agreement between this
892: relatively simple analytical calculation and the numerical results.
893:
894: One of the features of the numerical results shown above is that the N\'eel
895: state calculation always underestimates the values of $\ka$ which correspond to
896: gapless points, while all the time getting closer to the actual values with
897: increasing $S$ (given $N$) or increasing $N$ (given $S$). This may indicate
898: positive corrections of order $1/N$ and $1/S$ to the formula obtained from the
899: N\'eel state calculation.
900:
901: \begin{acknowledgments}
902: V.R.C. thanks CSIR, India for financial support. We thank DST, India for
903: financial support under the project SP/S2/M-11/2000.
904: \end{acknowledgments}
905:
906: \begin{thebibliography}{99}
907:
908: \bib{haldane} F. D. M. Haldane, Phys. Lett. {\bf 93A}, 464 (1983); Phys. Rev.
909: Lett. {\bf 50}, 1153 (1983).
910:
911: \bib{buyers} W. J. L. Buyers, R. M. Morra, R. L. Armstrong, M. J. Hogan, P.
912: Gerlach and K. Hirakawa, Phys. Rev. Lett. {\bf 56}, 371 (1986); J. P. Renard,
913: M. Verdaguer, L. P. Regnault, W. A. C. Erkelens, J. Rossat-Mignod and W. G.
914: Stirling, Europhys. Lett. {\bf 3}, 945 (1987); S. Ma, C. Broholm, D. H. Reich,
915: B. J. Sternlieb and R. W. Erwin, Phys. Rev. Lett. {\bf 69}, 3571 (1992).
916:
917: \bib{affleck1} I. Affleck, in {\it Fields, Strings and Critical Phenomena},
918: edited by E. Brezin and J. Zinn-Justin (North-Holland, Amsterdam, 1989).
919:
920: \bib{fradkin} E. Fradkin, {\it Field Theories of Condensed Matter Systems}
921: (Addison-Wesley, Reading, 1991).
922:
923: \bib{auerbach} A. Auerbach, {\it Interacting electrons and quantum magnetism}
924: (Springer-Verlag, New York, 1994).
925:
926: \bib{sierra} G. Sierra, in {\it Strongly Correlated Magnetic and
927: Superconducting Systems}, edited by G. Sierra and M. A. Martin-Delgado,
928: Lecture Notes in Physics 478 (Springer, Berlin, 1997).
929:
930: \bib{naveen} A. M. M. Pruisken, R. Shankar, and N. Surendran, Phys. Rev. B
931: {\bf 72}, 035329 (2005).
932:
933: \bib{kato} Y. Kato and A. Tanaka, J. Phys. Soc. Jpn. {\bf 63}, 1277 (1994).
934:
935: \bib{pati} S. Pati, R. Chitra, D. Sen, H. R. Krishnamurthy, and S. Ramasesha,
936: Europhys. Lett. {\bf 33}, 707 (1996).
937:
938: \bib{kitazawa1} A. Kitazawa and K. Nomura, J. Phys. Soc. Jpn. {\bf 66}, 3379
939: (1997).
940:
941: \bib{nakamura} M. Nakamura and S. Todo, Phys. Rev. Lett. {\bf 89}, 077204
942: (2002).
943:
944: \bib{shankar} R. Shankar and N. Read, Nucl. Phys. B {\bf 336}, 457 (1990);
945: I. Affleck, Phys. Rev. Lett. {\bf 56}, 408 (1986).
946:
947: \bib{kitazawa2} A. Kitazawa, J. Phys. A {\bf 30}, L285 (1997).
948:
949: \bib{affleck2} I. Affleck, D. Gepner, H. J. Schulz and T. Ziman, J. Phys. A
950: {\bf 22}, 511 (1989).
951:
952: \bib{okamoto} K. Okamoto and K. Nomura, Phys. Lett. A {\bf 169}, 433 (1992).
953:
954: \bib{nomura} K. Nomura, J. Phys. A {\bf 28}, 5451 (1995).
955:
956: \bib{kolb} M. Kolb, Phys. Rev. B {\bf 31}, 7494 (1985).
957:
958: \bib{jain} J. K. Jain and S. Kivelson, Phys. Rev. B {\bf 37}, 4111 (1988).
959:
960: \bib{oshikawa} M. Oshikawa, J. Phys. Condens. Matter {\bf 4}, 7469 (1992).
961:
962: \end{thebibliography}
963:
964: \end{document}
965:
966: