1: \documentclass[floatfix,preprintnumbers,amssymb]{revtex4}
2: %\documentclass[12pt]{iopart}
3:
4: \usepackage{graphicx}
5:
6: \usepackage{epsfig,psfrag,amsmath,amssymb,float,psfrag}
7: \newcommand{\be}{\begin{equation}}
8:
9: \newcommand{\ee}{\end{equation}}
10:
11: \newcommand{\bea}{\begin{eqnarray}}
12:
13: \newcommand{\eea}{\end{eqnarray}}
14:
15:
16: \newcommand{\bc}{\begin{center}}
17:
18: \newcommand{\ec}{\end{center}}
19:
20: \begin{document}
21:
22: \title{Critical phenomena and quantum phase transition in long range
23: Heisenberg antiferromagnetic chains}
24:
25: \author{Nicolas Laflorencie, Ian Affleck and Mona Berciu}
26:
27: \affiliation{Department of Physics \& Astronomy, University of British
28: Columbia, Vancouver, B.C., Canada, V6T 1Z1}
29:
30: \begin{abstract}
31: Antiferromagnetic Hamiltonians with short-range, non-frustrating
32: interactions are well-known to exhibit long range magnetic order in
33: dimensions, $d\geq 2$ but exhibit only quasi long range order, with
34: power law decay of correlations, in $d=1$ (for half-integer spin). On
35: the other hand, non-frustrating long range interactions can induce
36: long range order in $d=1$. We study Hamiltonians in which the long
37: range interactions have an adjustable amplitude $\lambda$, as well as
38: an adjustable power-law $1/|x|^\alpha$, using a combination of quantum
39: Monte Carlo and analytic methods: spin-wave, large-$N$ non-linear
40: $\sigma$ model, and renormalization group methods. We map out the
41: phase diagram in the $\lambda$-$\alpha$ plane and study the nature of
42: the critical line separating the phases with long range and quasi long
43: range order. We find that this corresponds to a novel line of critical
44: points with continuously varying critical exponents and a dynamical
45: exponent, $z<1$.
46: \end{abstract}
47:
48: \maketitle
49:
50: \section{Introduction}
51:
52: The ground-state (GS) of the nearest neighbor antiferromagnetic (AF)
53: Heisenberg model on a bipartite lattice:
54: \begin{equation}
55: {\cal{H}}=\sum_{\langle i,j\rangle}\vec S_i\cdot \vec S_j,
56: \end{equation}
57: is generally expected to have long range order (LRO):
58: \begin{equation}
59: \label{Neel}
60: \langle \vec S_{0}\cdot \vec S_r\rangle \to \pm m^{2}_{\rm{AF}},\ \
61: (r\to \infty),
62: \end{equation}
63: for any spin magnitude, $S$ and any dimension $d\geq
64: 2$~\cite{Ian94}. On the other hand, in dimension $d=1$, the behavior
65: depends on whether $S$ is integer or
66: half-integer~\cite{Haldane83-87}. In the half-integer case the
67: spin-spin correlation function
68: \begin{equation}
69: \langle\vec S_0\cdot \vec S_r\rangle\propto \frac{(-1)^{r}{\sqrt{\ln
70: r}}}{r},
71: \label{corr}
72: \end{equation}
73: is expected~\cite{Giamarchi89}, characteristic of a {\it quasi}-long
74: range order (QLRO). (In the integer spin, Haldane gap case,
75: correlations decay exponentially.) This behavior in $d=1$ for
76: half-integer $S$ is believed to be universal, not depending on the
77: magnitude of $S$ nor on the details of the Hamiltonian as long as it
78: is short range and not too frustrating. Long range interactions
79: (i.e. power-law decaying with the relative distance between
80: interacting moments) can be introduced in spin models either for some
81: experimental reasons like dipolar or RKKY interactions or simply
82: because of some theoretical relevance. For instance a famous example
83: is the Haldane-Shastry model~\cite{HS88} with AF frustrating $1/r^2$
84: interaction which exhibits an exact RVB GS. Another
85: theoretical interest comes from the possibility to interpolate
86: between discrete dimensions by tuning continuously the exponent that
87: governs the decay of the interaction with the distance. Indeed, the
88: possibility for true LRO to occur in $d=1$ with long range
89: interactions has motivated many studies during the last
90: decades~\cite{MW66,68-69,Fisher72,Kosterlitz76,Brezin76,Cardy81,Frolich82,Luijten01,Bruno01,Yusuf04}
91: and is the subject of the present paper. While a long standing debate
92: about the critical behavior of the Ising model in $d$ dimensions with
93: long range ferromagnetic interaction decaying like $r^{-d-\sigma}$
94: has been quite intense during the last thirty
95: years~\cite{68-69,Frolich82,Luijten01}, the N-vector model
96: has also been a subject of interest for many
97: authors~\cite{Fisher72,Kosterlitz76,Brezin76,Cardy81}. Concerning
98: the Heisenberg model with long range interaction $\sim r^{-\alpha}$,
99: the seminal paper of Mermin and Wagner~\cite{MW66} proving the
100: absence of LRO at finite temperature $T$ in $d\le 2$ for $\alpha>d+2$
101: has been recently reconsidered by Bruno~\cite{Bruno01} who gave
102: stronger conditions for the absence of spontaneous magnetic order at
103: $T>0$ in $d\le 2$. For instance, he proved that the AF non
104: frustrating one dimensional model \be {\cal{H}} = \sum_i\left[ \vec
105: S_i\cdot \vec S_{i+1} -\lambda \sum_{j=2}^\infty (-1)^{j}{\vec
106: S_i\cdot \vec S_{i+j} \over j^{\alpha}}\right].
107: \label{QLRO}
108: \ee does not have N\'eel order for any temperature $T>0$ if $\alpha
109: \geq 2$.
110: Actually, much less
111: is known about the $T=0$ case, except the work of Parreira {\it et
112: al.}~\cite{Parreira97} where the authors signaled the existence of the
113: bound $\alpha=3$ over which $T=0$ LRO is ruled out~\cite{note1}. A
114: particular case, the model (\ref{QLRO}) with $\lambda=1$ was recently
115: analyzed at both $T=0$ and $T>0$ in Ref.~\cite{Yusuf04}, using the
116: lowest order spin-wave (SW) approximation, expected to be valid for
117: large enough $S$ and small enough $\alpha$. There it was shown that
118: the SW dispersion relation takes the {\it sublinear} form, at low
119: $k$:
120: \be
121: \label{SL}
122: \omega (k) \propto |k|^{(\alpha -1)/2}, \ee for $\alpha
123: <3$. Consequently the quantum $1/S$ reduction of the order parameter:
124: \be \Delta m_q \propto \int {dk\over \omega (k)},\ee is finite for any
125: $\alpha <3$. By requiring that $\Delta m_q < S$, a consistency
126: condition on the SW approximation, it is concluded that LRO occurs for
127: any $S$ at sufficiently small $\alpha$. (However, such an estimate is
128: presumably only reliable for $S\gg1$.) After correcting a numerical
129: error in Ref.~\cite{Yusuf04}, the SW
130: prediction for the $S=1/2$, $\lambda =1$ case is existence of N\'eel
131: order at $T=0$ for $\alpha < \alpha^{sw}_c= 2.46$.
132:
133: In this work, we extend the results of Yusuf {\it et al.} in several
134: ways, focusing on the zero temperature behavior of the non frustrating
135: spin $1/2$ Hamiltonian (\ref{QLRO}) with long range interaction of
136: adjustable strength $\lambda$ and exponent $\alpha$. In Sec. II, we
137: consider the relevance of the long range term as a
138: perturbation to the
139: nearest neighbor interaction, using a simple heuristic
140: argument of mean-field type as well as the power-counting of the
141: scaling dimension of the perturbation. For
142: $\lambda\ll 1$, we find that the long range perturbation is marginal
143: if $\alpha=2$ and relevant (irrelevant) for $\alpha<2$ ($\alpha>2$). We then
144: investigate the $\alpha$- and $\lambda$-dependence of
145: the critical behavior using various techniques. We begin, in Sec. III, with
146: semi-classical calculations: the SW expansion and a large-N
147: approximation based on the non-linear $\sigma$ model. Both
148: approximations give qualitatively similar phase boundaries,
149: and sublinear dispersion like in Eq.~(\ref{SL}) in the ordered
150: phase. Some of the critical exponents can also be estimated within
151: these approximations. However, the results obtained in the SW or
152: large-N approximations are not
153: quantitatively correct. We therefore use large scale numerical
154: simulations to investigate more precisely the phase diagram of this
155: model in Secs. IV and V. We study systems of up to
156: $L=4000$ sites using quantum Monte Carlo (QMC) methods, based on a
157: stochastic series expansion (SSE) of the partition
158: function~\cite{Sandvik02,Sandvik03}. We verify that for $S=1/2$, there
159: are indeed stable phases with both QLRO given by Eq. (\ref{corr}) and
160: with true N\'eel LRO [Eq.~(\ref{Neel})]. We accurately
161: determine the phase boundary, as well as some of the
162: critical exponents which are found to vary continuously along the
163: critical line. In Sec. VI, we also apply analytic renormalization
164: group (RG) methods to
165: investigate the case $\lambda \ll 1$. Sec. VII contains
166: conclusions. In two appendices we
167: gives further details on the spin-wave theory and large-N calculations.
168:
169: \section{Relevance of the perturbation: mean field and scaling
170: arguments}
171: \label{sec:heur}
172: Let us consider a short range spin $1/2$ chain with an
173: additional long range perturbation of the form
174: \be
175: \label{pert}
176: \sum_{r,r'}J(r,r'){\vec S}_{r}\cdot {\vec S}_{r'},
177: \ee
178: with
179: \be
180: \label{jex}
181: J(r,r')=-\frac{(-1)^{|r-r'|}}{|r-r'|^{\alpha}}.
182: \ee
183: Following an argument given by Cardy~\cite{CardyBook} for the
184: relevance of a long range perturbation, we can in first approximation
185: look at the mean field correction to the free energy coming
186: from this long range term (\ref{pert}):
187: \be
188: \delta F =\sum_{r,r'}J(r,r')\langle {\vec S}_{r}\cdot {\vec
189: S}_{r'}\rangle,
190: \ee
191: where $\langle ... \rangle$ is evaluated in the unperturbed system
192: where we know the behavior of the correlation function
193: \be
194: \langle {\vec S}_{r}\cdot {\vec S}_{r'}\rangle \sim
195: \frac{(-1)^{|r-r'|}}{|r-r'|^{z+\eta-1}}.
196: \ee
197: In a finite system of length $L$, the change
198: in the free energy per site $\delta f$ thus scales like
199: \be
200: \delta f \sim \int_{1}^{L}\frac{dr}{r^{\alpha+z+\eta-1}}.
201: \ee
202: The integral above will give a constant term and a size dependent term
203: \be
204: \delta f(L)\sim L^{2-\alpha-z-\eta}\sim L^{-\alpha},
205: \ee
206: where we have used the fact that $z=\eta=1$ in the short range QLRO
207: regime of the spin $1/2$ chain. Then, we
208: can compare this with the usual
209: finite size corrections to the free energy of the conformally
210: invariant short range $S=1/2$ chain which are known to scale like $L^{-2}$
211: to lowest order~\cite{Ian86}. This tells us that (to first order
212: perturbation) if $\alpha<2$ the long range perturbation creates a
213: correction which dominates the $L^{-2}$ correction of the unperturbed
214: fixed point and is probably a relevant perturbation for the short
215: range model.
216:
217: Another way of deriving this result is to compute the scaling
218: dimension of the perturbation, based on the usual continuum
219: formulation of the short range model in which uniform and staggered
220: magnetization density operators, $(\vec J_L +\vec J_R)$ and $\vec n $
221: are introduced:
222: \be
223: \vec S(x) \approx (\vec J_L+\vec J_R)+(-1)^x\vec
224: n(x).\label{cont_lim}
225: \ee
226: Only slowly varying Fourier modes of the fields $\vec
227: J_{L/R}(x)$ and $\vec n(x)$ are present in the low energy effective
228: Hamiltonian. $\vec J_{L/R}$ are the conserved left/right-moving spin
229: densities. Ignoring a marginally irrelevant interaction, the staggered
230: magnetization field, $\vec n$, has the Green's function:
231: \be
232: \langle n^a(z)n^b(0)\rangle = {\delta^{ab}\over |z|},
233: \label{ncorr}
234: \ee
235: with $z\equiv \tau + ix$.
236: The long range perturbation adds to the low energy, continuum
237: limit of the imaginary
238: time action, a term of the form:
239: \be
240: \delta S[\vec n]\sim -\lambda\int {d\tau dx dy\over
241: |x-y|^{\alpha}}\vec n(\tau ,x)\cdot \vec n(\tau ,y).
242: \label{Seff}
243: \ee
244: Utilizing the fact that, from Eq. (\ref{ncorr}),
245: $\vec n$ has a scaling dimension of 1/2, a simple power counting tells
246: us that the perturbation is irrelevant for $\alpha >2$, relevant for
247: $\alpha <2$, and
248: marginal for $\alpha =2$.
249: Also note that $\lambda >0$ corresponds to
250: non-frustrating interactions which favor the N\'eel state with
251: $\langle n^z\rangle\neq 0$.
252: \section{Spin-wave expansion and Large-N approximation}
253: \label{sec:SW}
254: \subsection{Spin-wave expansion}
255: This calculation simply generalizes that of Yusuf {\em et al.} in
256: Ref. \cite{Yusuf04}, to $\lambda \ne 1$. We summarize here the main
257: steps. Some further results are given in Appendix A. On the LRO side of
258: the transition, we use the
259: Holstein-Primakoff approximation~\cite{Holstein}:
260: $$
261: S_{i}^z = S-a^{\dagger}_{i}a_{i}, S_{i}^{+}\approx
262: \sqrt{2S} a_{i}, S_{i}^{-}\approx
263: \sqrt{2S} a^{\dagger}_{i},
264: $$
265: for $i$ odd and
266: $$
267: S_{j}^z = b^{\dagger}_{j}b_{j}-S, S_{j}^{+}\approx
268: \sqrt{2S} b^{\dagger}_{j}, S_{j}^{-}\approx
269: \sqrt{2S} b_{j},
270: $$
271: for $j$ even, and retain only the quadratic terms in the Hamiltonian (\ref{QLRO}).
272: After a Fourier transform over the reduced Brillouin
273: zone $k\in (-{\pi\over 2a}, {\pi\over 2a})$, we find:
274: \begin{equation}
275: {\cal{H}}_{SW} \approx S \sum_{k}^{}\Bigl[ \left(\gamma - f(k)
276: \right)\left(a_k^{\dagger} a_k + b_k^{\dagger} b_k \right)
277: + g(k) \left(a_k^{\dagger} b_{-k}^{\dagger} +b_{-k} a_k\right)\Bigr]+\dots
278: \ee
279: where, for an infinite chain \cite{note2}:
280: $$
281: \gamma = 2 + 2 \lambda \sum_{n=2}^{\infty} \frac{1}{(2n-1)^{\alpha}}
282: $$
283: $$
284: f(k) = 2 \lambda \sum_{n=1}^{\infty} \frac{ \cos(2kna)-1}{(2n)^{\alpha}}
285: $$
286: $$
287: g(k) = 2 \cos(ka) + 2 \lambda \sum_{n=2}^{\infty}
288: \frac{\cos[k(2n-1)a]}{(2n-1)^\alpha}
289: $$
290: This quadratic Hamiltonian can be diagonalized with a
291: Bogoliubov transformation to:
292: \be
293: {\cal H}_{SW} \approx S \sum_{k}^{} \omega_k
294: \left(\chi_{k,1}^{\dagger}\chi_{k,1}+\chi_{k,2}^{\dagger}\chi_{k,2}
295: \right)
296: \label{hsw}
297: \ee
298: with a SW spectrum
299: \be
300: \omega_k = \sqrt{\left[\gamma - f(k)
301: \right]^2 + [g(k)]^2 }\stackrel{k\rightarrow 0}{\longrightarrow}
302: k^{\alpha-1\over 2} ,
303: \ee
304: as discussed above. At $T=0$, the correction
305: to the staggered magnetization at any site is
306: $$
307: \Delta m_q = \langle a^{\dagger}_{i} a_{i}\rangle = \langle
308: b^{\dagger}_{j} b_{j}\rangle = {a\over 2\pi} \int_{-{\pi\over
309: 2a}}^{\pi\over 2a} dk \left[\frac{\gamma - f(k)}{\omega_k} -1\right]
310: $$
311: The consistency condition $\Delta m_q
312: < S$ then allows us to find the SW approximation for the value
313: of $\alpha_{c}^{sw}$ below which long range N\'eel order is established. As
314: already stated, for $S=1/2$ and $\lambda =1$, we find $\alpha_{c}^{sw} =
315: 2.46$. A plot of $\alpha_{c}^{sw}$ vs. $\lambda$, for $S=1/2$, is shown in
316: Fig.~\ref{fig:SW}.
317:
318: \begin{figure}[!ht]
319: \bc
320: \includegraphics[width=8cm,clip]{Figures/PhDgSW.eps}
321: \caption{Spin-wave approximation prediction for the value
322: $\alpha_{c}^{sw}$ below which long range N\'eel order is expected at
323: $T=0$, as a function of $\lambda$ and for $S=1/2$. The critical
324: curve saturates at $\alpha \simeq 2.9032$ when $\lambda\to\infty$.}
325: \label{fig:SW}
326: \ec
327: \end{figure}
328: This phase boundary given by the lowest order SW approximation turns
329: out not to be quantitatively correct, as we are going to see with the QMC
330: results presented in section~\ref{sec:QMC1}. In particular it happens that SW
331: predictions miss the fact that the critical line goes to $\alpha=2$
332: when $\lambda \to 0$. Nevertheless SW predictions are, to some
333: extent, shared by large-$N$ calculations as we are going to see
334: below.
335:
336: \subsection{Large-$N$ approximation}
337: The details of this calculation are given in
338: Appendix~\ref{app:LN}. Here we present the main steps and discuss the
339: results which come from this approximation. We generalize the
340: N\'eel order parameter field, $\vec n(\tau,x)$ appearing in Eq. (\ref{cont_lim}),
341: to an $N$-component field and take the limit of large $N$. In this
342: approximation 2 phases occur in the $\lambda$ -$\alpha$ plane.
343: The critical line terminates at $\alpha =1$, as in spin-wave approximation.
344: These two phases are
345: a phase with N\'eel order and a disordered phase with a finite correlation length.
346: (The unusual quasi long range ordered phase is special to the case $N=3$
347: and is not captured by the large-$N$ approximation.) Along the critical line
348: separating these two phases the mean field result, $\eta =3-\alpha$ is obtained.
349: The dynamical exponent takes the value $z=(\alpha -1)/2$, corresponding
350: to the dispersion relation $\omega \propto |k|^{(\alpha -1)/2}$ also
351: obtained in spin-wave theory. The correlation length diverges with an exponent, $\nu$
352: defined by:
353: \be \xi \propto |\lambda_c-\lambda|^{-\nu},\ee
354: with
355: \bea \nu &=& 1/(\alpha -1),\ \ (1<\alpha <5/3) \nonumber \\
356: \nu &=& 2/(3-\alpha ),\ \ (5/3<\alpha <3).\eea
357:
358:
359:
360: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
361: \section{Quantum Monte Carlo results I: Finite size effects}
362: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
363: \label{sec:QMC1}
364: In this section, we present results obtained using the QMC SSE method
365: based on directed loop
366: updates~\cite{Sandvik02}. This algorithm, used here to investigate the
367: model (\ref{QLRO}), has been proposed recently by Sandvik~\cite{Sandvik03}
368: to study spin Hamiltonians with non-frustrating long range interactions.
369: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
370: \subsection{Finite size corrections}
371: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
372:
373: We first focus on the $\lambda=1$ case, studied by SW in Ref.~\cite{Yusuf04}, governed by the following
374: Hamiltonian
375: \be
376: {\cal H}=-\sum_{i,j\geq 1}\frac{(-1)^j}{j^{\alpha}}{\vec S}_{i}\cdot{\vec S}_{i+j}.
377: \label{eq:iso}
378: \ee
379: %
380: In order to detect a N\'eel instability at the thermodynamic limit,
381: we compute the staggered structure factor, normalized per site, on finite length
382: spin $S=1/2$ chains, defined by
383: \be
384: S_{\pi}(L)=\frac{1}{L^2}\sum_{i,j}(-1)^{i-j}\langle{{\vec S}}_{i}\cdot{{\vec S}}_{j}\rangle=\frac{3}{L^2}\langle (\sum_{i=1}^{L}(-1)^i S_{i}^{z})^2\rangle.
385: \ee
386: We have performed SSE simulations for different system sizes, up to
387: $L_{\rm{max}}=4096$, at temperatures $\beta^{-1}=1/2L$
388: low enough to get the GS properties.
389: Results for $S_{\pi}(L)$ are shown vs $1/L$ in the left panel of
390: Fig.~\ref{fig:spi} for
391: different values of the power-law exponent $\alpha$. The staggered
392: structure factor displays two types of behavior: for small values of
393: $\alpha$ it saturates to a finite non-zero number whereas for large
394: enough $\alpha$, $S_{\pi}(L)$ vanishes when $L\to\infty$.
395: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
396: \begin{figure}[!ht]
397: \bc
398: \psfrag{B}{\Huge{N\'eel}}
399: \psfrag{K}{\Large QLRO}
400: \begin{minipage}{\columnwidth}
401: \includegraphics[width=0.46\columnwidth,clip]{Figures/Spi_Lambda1_QMC.eps}
402: \includegraphics[width=0.44\columnwidth,clip]{Figures/mAF_Lambda1_QMC.eps}
403: \end{minipage}
404: \caption{Left panel: $T=0$ QMC results for the staggered structure factor per
405: site $S_{\pi}(L)$ computed in the GS for the Hamiltonian
406: (\ref{eq:iso}) and plotted vs the inverse system size
407: $L^{-1}$ in a log-log scale.
408: Different symbols are used for different values of the
409: power-law exponent $\alpha$, as indicated on the plot. The case
410: with only nearest neighbor interactions is also shown (N.N green crosses) for
411: comparison. Right panel: Infinite size AF order
412: parameter $m_{\rm{AF}}$ plotted vs $\alpha$, obtained using finite size scaling of
413: $S_{\pi}(L)$ (shown on the left panel). The quantum
414: phase transition between the N\'eel
415: phase ($m_{\rm{AF}}\ne 0$) and the QLRO phase ($m_{\rm{AF}}=0$)
416: occurs at $\alpha_c=2.225\pm 0.025$. The SW estimate
417: ($\alpha_{c}^{\rm{sw}}\simeq 2.46$) is indicated by the arrow.}
418: \label{fig:spi}
419: \ec
420: \end{figure}
421: %
422: Then, in order to extract the thermodynamic limit behavior of
423: $S_{\pi}$, we perform a finite size analysis in order to get the AF
424: order parameter, given by
425: \be
426: \sqrt{S_{\pi}(L)}\to m_{\rm{AF}},\ \ (L\to \infty).
427: \label{maf1}
428: \ee
429: Utilizing the fact that in the QLRO regime the spin-spin correlation
430: functions decay as stated in Eq.~(\ref{corr}), we therefore expect in
431: this regime the following behavior for the staggered structure factor
432: per site
433: \be
434: S_{\pi}(L)= \frac{1}{L}\int_{1}^{L}(-1)^r\langle {\vec{S}}_0 \cdot
435: {\vec{S}}_r \rangle dr\sim \frac{\left(\ln L\right)^{3/2}}{L},\ \ (L\to \infty).
436: \ee
437: On the other hand, in the N\'eel phase, the finite size scaling of the
438: order parameter can be evaluated using the small $k$ SW spectrum (see appendix~\ref{app:SW}), leading to
439: \be
440: S_{\pi}(L)-m_{\rm{AF}}^2\sim L^{\frac{\alpha-3}{2}}+O(L^{{\alpha-3}}).
441: \label{eq:fss}
442: \ee
443: We used second order polynomial fits in $L^{\frac{\alpha-3}{2}}$ to
444: extrapolate the finite size data to their thermodynamic limit
445: values, shown in the right panel of Fig.~\ref{fig:spi}.
446: The quantum phase transition between the AF N\'eel order and the
447: QLRO phase is clearly visible for a critical value
448: $2.2<\alpha_c<2.25$. It is also interesting to compare this estimate
449: with the one from SW approximation giving $\alpha_{c}^{\rm{sw}}\simeq 2.46$.
450:
451:
452: Let us now concentrate on the $\alpha-$ and $\lambda-$dependent
453: Hamiltonian (\ref{QLRO}) by keeping a fixed value for $\alpha$ while
454: varying $\lambda$. We first focus
455: on the case with $\alpha=2.1$ which is expected to display a
456: transition for a non-zero value of $\lambda$.
457: As pointed out by Reger and Young studying finite size AF clusters in
458: $d=2$~\cite{Reger88}, the sublattice (infinite size) magnetization can be obtained
459: either from the staggered structure factor [Eq.~(\ref{maf1})] or from
460: the correlation functions at the largest separation
461: \be
462: \label{maf2}
463: C(L)=\langle \vec{S}_i\cdot\vec{S}_{i+L/2}\rangle \to
464: \pm m_{\rm{AF}}^{2},\ \ (L\to \infty).
465: \ee
466: In the N\'eel phase, both estimators $S_{\pi}(L)$ and $C(L)$ are
467: expected to converge to $m_{\rm{AF}}^2$ with a similar
468: power-law behavior but with different pre-factors. This feature is
469: illustrated by the computation of $C(L)$ and $S_{\pi}(L)$ for
470: $\alpha=2.1$ and four different value of the long range term strength
471: $\lambda=3,~2~,0.9,~0.6$, as shown in Fig.~\ref{fig:2.1}.
472: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
473: \begin{figure}[!ht]
474: \bc
475: \psfrag{Z}{\tiny{\bf{(a) $\lambda=3$}}}
476: \psfrag{Y}{\tiny{\bf{(b) $\lambda=2$}}}
477: \psfrag{X}{\tiny{\bf{(c) $\lambda=0.9$}}}
478: \psfrag{W}{\tiny{\bf{(d) $\lambda=0.6$}}}
479: \includegraphics[width=10cm,clip]{Figures/Extrapolation_alpha2.1_QMC.eps}
480: \caption{Staggered structure factor per site
481: $S_{\pi}(L)$ [Eq.~(\ref{maf1})] and mid-chain correlation function
482: $C(L)$ [Eq.~(\ref{maf2})] computed at $T=0$ with QMC for $\alpha=2.1$
483: and 4 different
484: values of $\lambda$, as shown on the plot (a)-(d). Finite size
485: scaling has been performed on finite systems up to
486: $L_{\rm{max}}=128$ for the largest values of $\lambda$ ((a) and
487: (b)) and up to $L_{\rm{max}}=256$ for the smallest ((c) and
488: (d)). The dashed lines are polynomial fits of the form
489: $m_{\rm{AF}}^2+a_1 L^{\frac{\alpha-3}{2}}+a_2 L^{\alpha-3}$.}
490:
491: \label{fig:2.1}
492: \ec
493: \end{figure}
494: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
495: Using second order polynomial fits in $L^{\frac{\alpha-3}{2}}$, we can
496: obtain infinite size extrapolated values for $m_{\rm{AF}}(\lambda)$
497: from $S_{\pi}(L)$ or $C(L)$, as reported in table~\ref{table1}.
498: Finite size effects are more pronounced for the staggered structure
499: factor than for the mid-chain correlation function because
500: $S_{\pi}(L)$ is the result of the integration of the staggered
501: correlation function along the entire chain and therefore is sensitive
502: to short distance terms. However, the estimates for the sublattice
503: magnetization obtained from $C(L)$ and $S_{\pi}(L)$ (see
504: table~\ref{table1}) are both in good agreement, especially when the
505: system is deeply in the N\'eel regime (large values of $\lambda$). On
506: the other hand, when the system is approaching the quantum critical
507: point (QCP) where $m_{\rm{AF}}\to 0$, the finite size effects are
508: significant enough to prevent us from obtaining a very precise
509: estimate of the critical coupling $\lambda_c$ where the AF LRO
510: vanishes.
511:
512: \begin{table}
513: \label{table1}
514: \begin{tabular}{r|r|r|}
515: $\lambda$&$m_{\rm{AF}}$ from $S_{\pi}(L)$&$m_{\rm{AF}}$ from $C(L)$\\
516: \hline
517: 3&0.353&0.356\\
518: 2&0.295&0.301\\
519: 0.9&0.106&0.091\\
520: 0.6&0&0.02\\
521: \end{tabular}
522: \caption{Infinite size extrapolated values of the
523: sublattice magnetization $m_{\rm{AF}}$ obtained for $\alpha=2.1$ and
524: $\lambda=3,~2,~0.9,~0.6$ from power-law fits of the staggered
525: structure factor $S_{\pi}(L)$ and the mid-chain correlation function $C(L)$ (see Fig.~\ref{fig:2.1}).}
526: \end{table}
527: %
528: %\begin{table}
529: %\begin{indented}
530: %\item[]\begin{tabular}{@{}lll}
531: %\br
532: %$\lambda$&$m_{\rm{AF}}$ from $S_{\pi}(L)$&$m_{\rm{AF}}$ from $C(L)$\\
533: %\mr
534: %3&0.353&0.356\\
535: %2&0.295&0.301\\
536: %0.9&0.106&0.091\\
537: %0.6&0&0.02\\
538: %\br
539: %\end{tabular}
540: %\caption{\label{table1}Infinite size extrapolated values of the
541: %sublattice magnetization $m_{\rm{AF}}$ obtained for $\alpha=2.1$ and
542: %$\lambda=3,~2,~0.9,~0.6$ from power-law fits of the staggered
543: %structure factor $S_{\pi}(L)$ and the mid-chain correlation function $C(L)$ (see Fig.~\ref{fig:2.1}).}
544: %\end{indented}
545: %\end{table}
546: %
547: Of course, in principle it is possible to perform
548: very large scale numerical SSE simulations on the largest reachable system
549: sizes, as we did for the $\lambda=1$ case with
550: $L_{\rm{max}}=4096$. However, since our goal here is to investigate
551: the quantum critical phenomena in the $\lambda-\alpha$ plane, we need
552: a good sampling of this parameter space and we therefore restrict
553: the simulations over systems of maximum size $L_{\rm{max}}\le 1024$.
554: We then use another strategy, based on scaling
555: arguments, to perform a better data analysis
556: close to criticality. This is described next.
557: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
558: \subsection{Scaling analysis}
559: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
560: As previously discussed, the finite size effects are bigger for the
561: staggered structure factor than for the mid-chain correlation
562: function. Therefore, we now focus on $C(L)$ which is
563: expected to saturate to a constant value in the N\'eel phase, whereas in the
564: QLRO regime, the behavior $C(L)\to \sqrt{\ln (L/a)}/L$ is
565: expected, $a$ being a non-universal constant.
566:
567:
568: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
569: \begin{figure}[!ht]
570: \bc
571: \psfrag{W}{$C(L)$}
572: \psfrag{X}{$L$}
573: \psfrag{Y}{$\Psi_{\lambda}C(L)$}
574: \psfrag{Z}{$L/\xi_{\lambda}$}
575: \begin{minipage}{\columnwidth}
576: {\hskip -0.2cm{\includegraphics[width=0.49\columnwidth,clip]{Figures/Corr_alpha2.1_QMC.eps}}}
577: {\hskip 0.2cm{\includegraphics[width=0.49\columnwidth,clip]{Figures/Collapse_Corr_alpha2.1_QMC.eps}}}
578: \end{minipage}
579: \caption{$T=0$ QMC results for the mid-chain correlation function $C(L)$
580: computed for $\alpha=2.1$ and different values of $\lambda$, as
581: indicated on the plot. (a) $C(L)$ is plotted vs the system size $L$ for
582: $0\le \lambda\le 1.15$. (b)
583: Both x- and y-axis are rescaled using two parameters:
584: the crossover
585: length scale $\xi_{\lambda}$ and $\Psi_{\lambda}$. The data collapse results in two
586: universal curves: one for the N\'eel ordered phase (top one) and
587: one for the QLRO regime (lower one). Note that for clarity, the
588: QLRO universal curve has been shifted downwards. The red dashed-line
589: materializes the critical separatrix between the two regimes, decaying with
590: an exponent $\simeq 0.63$. Inset: Crossover
591: length scale $\xi_{\lambda}$ extracted from the data collapse in the
592: QLRO (full circles) and N\'eel regime (open squares). The dashed lines
593: are power-law fits of the form $|\lambda-\lambda_c|^{-\nu}$
594: with $\lambda_c\simeq 0.45$ (indicated by vertical
595: dotted line) and $\nu\sim 15$.}
596: \label{fig:cL}
597: \ec
598: \end{figure}
599: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
600: In order to illustrate the scaling analysis, let us continue to study
601: the case with $\alpha=2.1$, as in the previous
602: subsection. We have computed $C(L)$ for several values of the long range coupling
603: strength $\lambda$ in the range $[0,1.15]$, for sizes up to $L=256$. The results,
604: shown in Fig.~\ref{fig:cL} (a), clearly show the existence of a
605: finite critical value
606: $\lambda_c$ which separates the QLRO and the N\'eel regimes. In order to
607: locate precisely this QCP, let us assume that
608: a typical length
609: scale $\xi_{\lambda}$ governs a crossover from the QCP to the N\'eel
610: phase if $\lambda > \lambda_c$ and from the QCP to the QLRO regime if
611: $\lambda < \lambda_c$. Precisely at the critical point, the spin-spin
612: correlation function decay like a power-law
613: %
614: \be
615: C_{\rm{QCP}}(L)\sim L^{1-z-\eta},
616: \label{eq:QCP}
617: \ee
618: %
619: thus defining the critical exponents $\eta$ and $z$, the critical
620: dynamical exponent. Without making any
621: assumption about the values of the aforementioned critical exponents,
622: let us now define scaling
623: functions $f_{\pm}(x)$, with $x=L/\xi_{\lambda}$,
624: for $\lambda>\lambda_c$ and $\lambda<\lambda_c$, respectively, by
625: %
626: \be
627: f_{\pm}(x)=\frac{C(L)}{C_{\rm{QCP}}(L)}
628: \ee
629: %
630: Hence, the scaling functions obey
631: %
632: \bea
633: \label{eq:scaling}
634: f_-(x)&\propto&
635: x^{-2+z+\eta}{\sqrt{\ln x}}{\rm{~if~}} x\gg 1 {\rm{~and~}} \lambda
636: < \lambda_c {\rm{~~~~(QLRO)}}\nonumber \\
637: f_+(x)&\propto&x^{-1+z+\eta}{\rm{~if~}} x\gg 1 {\rm~{and~}} \lambda > \lambda_c {\rm{~~~~(NEEL)}}\nonumber \\
638: f_{\pm}(0)&=&1 {\rm{~if~}} \lambda \simeq \lambda_c {\rm{~~~~(QCP)}}
639: \eea
640: %
641: It is convenient to also rescale the y-axis with the unknown function
642: $\Psi_{\lambda}$ in order to get
643: $C(L)\times
644: \Psi_{\lambda}=f(x)\times x^{1-z-\eta}$.
645: We then expect $\Psi_{\lambda}$ to be proportional to $\xi_{\lambda}^{z+\eta-1}$.
646: Using such scaling forms, we have obtained the collapse of the data
647: shown in Fig.~\ref{fig:cL}(a) into two universal curves shown in
648: Fig.~\ref{fig:cL}(b). The parameters $\xi_{\lambda}$ and
649: $\Psi_{\lambda}$ have been chosen to give the best data collapses.
650: Using such a scaling analysis, we find a critical coupling
651: $\lambda_c=0.45\pm 0.05$ that we can compare to the overestimated value
652: $\lambda_c=0.6$ previously found using the more simple finite size
653: scaling Eq.~(\ref{eq:fss}). The critical correlation (given by the separatrix between the two
654: regimes in Fig.~\ref{fig:cL}(b)) is characterized here by a power-law
655: decay with an exponent $(z+\eta-1)_{\rm{QCP}}\simeq 0.63$.
656: Note also that the crossover length scale $\xi_{\lambda}$, plotted in the inset of
657: Fig.~\ref{fig:cL}(b), diverges on both sides of the
658: transition with
659: a large exponent $\nu\sim 15$~\cite{notexi}.
660: These and other issues related to the critical exponents will be discussed in
661: detail in Sec.~\ref{sec:crit}.
662: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
663: \section{Quantum Monte Carlo II: Phase diagram and Critical behavior}
664: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
665: The scaling analysis described above has been repeated for several values
666: of $\alpha$ in order to explore and construct the phase diagram of the
667: model (\ref{QLRO}) in the $\lambda-\alpha$ plane.
668: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
669: \subsection{$\alpha=2$: Marginal case}
670: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
671: Let us first focus on the marginal case with
672: $\alpha=2$ for which a similar data collapse analysis is performed and
673: shown in Fig.~\ref{fig:cL1} for the mid-chain correlation
674: function.
675:
676: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
677: \begin{figure}[!ht]
678: \bc
679: \psfrag{Y}{$\Psi_{\lambda}C(L)$}
680: \psfrag{Z}{$L/\xi_{\lambda}$}
681: \includegraphics[width=8cm,clip]{Figures/Collapse_Corr_alpha2_QMC.eps}
682: \ec
683: \caption{$T=0$ QMC results at $\alpha=2$ for the mid-chain correlation
684: functions $C(L)$. As in
685: Fig.~\ref{fig:cL}, both x- and y-axis have been rescaled in order
686: to get the best data collapse. For the different values of $\lambda$
687: indicated on the plot, the data collapse on a unique crossover
688: curve towards the N\'eel ordered phase. Inset: Crossover
689: length scale $\xi_{\lambda}$ plotted in a linear-log scale vs
690: $\lambda^{-0.4}$. The dashed line is a fit of the form
691: Eq.~(\ref{eq:expfit}) with $\sigma=0.4$.}
692: \label{fig:cL1}
693: \end{figure}
694: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
695: Using QMC simulations results for chains up to $L=512$ sites, with
696: $\lambda \in [0,3]$, we have been able to get an universal curve (see
697: Fig.~\ref{fig:cL1}) which shows a crossover towards a N\'eel order
698: phase (i.e. $C(L)\to{\rm{constant}}$ if $L\gg \xi_{\lambda}$).
699: Note that for $\lambda < 0.1$, the typical
700: length scale necessary to get a good collapse becomes very large
701: so that there is no overlap between our different curves and the
702: data collapse analysis is impossible to achieve.
703: Nevertheless, the crossover length scale
704: $\xi_{\lambda}$, plotted in the inset of Fig.~\ref{fig:cL1}, displays
705: an exponential divergence when $\lambda\to 0$. Guided by the RG calculations presented in
706: section~\ref{sec:RG}, we can fit the $\lambda$-dependence of the crossover length scale by
707: \be\label{eq:expfit}
708: \xi_{\lambda}\sim \exp(C/\lambda^{\sigma}),
709: \ee
710: with $\sigma=0.4$ and $C$ being a free parameter.
711: It is however important to note
712: that since $\xi_{\lambda}$ suffers from large error bars, and so does the
713: fitting parameter, we have
714: forced $\sigma$ to its value found in Eq.~(\ref{eq:exp}).
715:
716: Unambiguously, $\xi_{\lambda}$ is found to diverge when $\lambda\to 0$
717: which means that at the marginal point
718: $\alpha=2$, any $\lambda >0$ will drive the system towards the N\'eel
719: phase. In other words, the
720: long range interaction perturbation of strength $\lambda$ is
721: {\it{marginally relevant}} at $\alpha=2$. This result agrees
722: with the RG calculations presented
723: in section~\ref{sec:RG}.
724: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
725: \subsection{Phase Diagram}
726: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
727: As previously stated, when $\alpha\le 2$ the long range interaction is a
728: relevant perturbation and any $\lambda>0$ will drive the QLRO
729: phase towards a AF ordered N\'eel phase with
730: $m_{\rm{AF}}\ne 0$. On the other hand, when $\alpha> 2$ a simple
731: power-counting tells us that the perturbation is irrelevant which
732: should imply that the QLRO is stable against a small perturbation
733: $\lambda>0$.
734: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
735: \begin{figure}[!ht]
736: \bc
737: \psfrag{NEEL}{\Huge N\'eel}
738: \psfrag{QLRO}{\Huge QLRO}
739: \includegraphics[width=10cm,clip]{Figures/PhaseDiag_QMC.eps}
740: \caption{$T=0$ phase diagram of the long range $S=1/2$ model
741: [Eq.~(\ref{QLRO})] computed by large scale QMC
742: simulations and plotted in the $\lambda-\alpha$ plane. A line of
743: critical points (circles) separates a long range ordered phase
744: (N\'eel) and a quasi long range ordered phase (QLRO). The error
745: bars, due to some uncertainties in the finite size scaling analysis of
746: the numerical data, are explicitly show on the plot. The dashed line
747: is a guide for the eyes.}
748: \label{fig:phasediag}
749: \ec
750: \end{figure}
751: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
752: It turns out that such a simple argument is not sufficient to provide
753: a correct description of the quantum critical behavior of the
754: system (see the next section for a more advanced field theoretical
755: description). Based on large scale numerical simulations, we
756: provide hereafter a picture which is consistent with the existence of a
757: non-trivial line of fixed points in the
758: $\lambda-\alpha$ plane.
759:
760: Using QMC simulations on systems of up to $L=1000$, we
761: performed the scaling analysis for the mid-chain correlation
762: function $C(L)$ as well as for the staggered susceptibility (see
763: below) and computed the phase diagram for $2\le \alpha\le
764: 2.7$. For each value of $\alpha$, the QCP $\lambda_c$ is found
765: by the separatrix between the two crossover functions (see
766: Fig.~\ref{fig:cL}(b)) with some error bars due to the discrete
767: sampling in the $\lambda$ space as well as the strong divergence of
768: the crossover length scale close to the critical point which makes
769: the data collapse delicate.
770: We present in Fig.~\ref{fig:phasediag} the QMC phase diagram in the
771: $\lambda-\alpha$ plane.
772: As discussed, $\lambda$ is marginally relevant at $\alpha =2$, driving
773: the system towards a N\'eel phase with LRO. At small $\lambda$, the critical line
774: increases sharply from $\alpha=2$ and displays a
775: negative curvature. By contrast, spin-wave theory (see
776: Fig.~\ref{fig:SW}) and large-$N$
777: approximation predict
778: that $\alpha_c(\lambda\rightarrow 0 )=1$. In the range of $\lambda$ considered here
779: ($\lambda<8$), the critical line stays well below the value
780: $\alpha=3$ and we expect this feature to remain true for all
781: $\lambda$. This behavior is
782: consistent with the proof of absence of LRO at $T=0$ for $\alpha >3$
783: \cite{Parreira97,note1}.
784:
785: %
786: %Let us give a few words about this computed phase boundary in the
787: %$\lambda-\alpha$ plane which is quantitatively different from the one
788: %computed with SW (see Fig.~\ref{fig:SW}).
789: %First of all, $\lambda$ is marginally relevant at $\alpha =2$, driving
790: %the system towards a N\'eel phase with LRO. At small $\lambda$, the critical line
791: %increases sharply from $\alpha=2$ and displays a
792: %negative curvature. By contrast, spin-wave theory and large-$N$
793: % approximation predict
794: %that $\alpha_c(\lambda )$ hits the $\alpha$ axis
795: %at $\alpha =1$. In the range of $\lambda$ considered here
796: %($\lambda<8$), the critical line stays well below the upper bound
797: %$\alpha=3$ and regarding the shape of
798: %the curve, we expect this feature to remain true.
799: %Also, we
800: %could perhaps expect the critical line to eventually saturate asymptotically
801: %towards a value $\alpha^{*}<3$. This issue is of course only
802: %speculative and not conclusive at all.
803:
804:
805: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
806: \subsection{Critical exponents}
807: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
808: \label{sec:crit}
809: The transition line between N\'eel LRO and QLRO is a
810: non trivial line which displays continuously varying critical
811: exponents, as we show now.
812:
813: \subsubsection{Divergence of the crossover length scale}
814: \label{sec:nu}
815: The standard theory of quantum phase transitions involves a set of
816: critical exponents which govern the universal behavior of various
817: quantities close to or at the QCP. One of them is
818: $\nu$ which tells us how does the correlation length diverge in the
819: real space direction close to the critical point. Usually this
820: correlation length is defined in the disordered phase by the
821: exponential decay $\sim
822: \exp(-r/\xi)$ of the correlation function associated with the order
823: parameter. In our case, the non ordered regime $\lambda<\lambda_c$ is already
824: critical and thus the correlation length is intrinsically
825: infinite. Nevertheless, the typical length scale $\xi_\lambda$ which
826: governs the crossover phenomenon, diverges at the QCP (on both sides) with an exponent which we call $\nu$ by analogy:
827: %
828: \be
829: \xi_{\lambda} \propto |\lambda -\lambda_c|^{-\nu}.
830: \ee
831: %
832: As already discussed, an accurate numerical evaluation of the
833: exponent $\nu$ is difficult, because of some intrinsic
834: uncertainties in the data collapse procedure.
835: Nevertheless, at our
836: level of precision we observe this crossover
837: length scale exponent increasing when $\alpha\to 2^+$.
838: In particular,
839: at the marginal point $\alpha=2$, we find an exponential divergence
840: of $\xi_{\lambda}$ near $\lambda=0$ [Eq.~(\ref{eq:expfit})], formally
841: corresponding to $\nu=\infty$. This divergence of $\nu$ when
842: $\alpha$ approaches 2 is actually in good agreement with the results of
843: field theory and RG calculations presented in Sec.~\ref{sec:RG}.
844:
845:
846: \subsubsection{Staggered magnetization exponent and hyperscaling relation}
847: The other scaling parameter $\Psi_{\lambda}$, used for the collapse of
848: the correlation functions data, also contains some
849: information. First, if the scaling hypothesis used above with the help
850: of the crossover functions is correct, we expect $\Psi_{\lambda}\sim
851: \xi_{\lambda}^{z+\eta-1}$ which gives another estimate for the critical
852: exponent of the decay of the
853: correlation function [Eq.~({\ref{eq:QCP})]. This is illustrated in Fig.~\ref{fig:PsiXi}
854: where $\Psi_{\lambda}$ is plotted vs $\xi_{\lambda}$ for
855: $\alpha=2.1$. Data, presented for both sides of the transition in
856: Fig.~\ref{fig:PsiXi}, clearly display power-law dependences with an
857: exponent $\eta+z-1$ in very good agreement with the value of $0.63$
858: previously found along the separatrix in Fig.~\ref{fig:cL}(b). Note
859: also that the agreement is even better when getting closer to the QCP.
860: %
861: \begin{figure}[!ht]
862: \bc
863: \includegraphics[width=8cm,clip]{Figures/Psi-Xi.alpha2.1.eps}
864: \caption{Log-log plot of the scaling parameters $\Psi_{\lambda}$ and $\xi_{\lambda}$
865: obtained from the collapses shown in Fig.~\ref{fig:cL}(b) for
866: $\alpha=2.1$ in both phases: N\'eel ({\Large$\circ$}) and QLRO
867: ($\square$). Within error bars (explicitly shown on the plot), data
868: are fitted by power-laws (dotted lines) of the form
869: $\xi_{\lambda}^{0.617}$ for the N\'eel regime and
870: $\xi_{\lambda}^{0.636}$ for the QLRO.}
871: \label{fig:PsiXi}
872: \ec
873: \end{figure}
874: %
875:
876: In the ordered phase, according to Eq.~(\ref{maf2}), we expect for the
877: AF order parameter
878: %
879: \be
880: m_{\rm{AF}}\propto \xi_{\lambda}^{\frac{1-z-\eta}{2}}\propto
881: (\lambda-\lambda_c)^{\frac{\nu(1-z-\eta)}{2}}.
882: \ee
883: %
884: This implies the usual {\it hyperscaling} relation involving the critical
885: exponent $\beta$ governing the onset of the order parameter
886: %
887: \be
888: m_{\rm{AF}}\propto (\lambda-\lambda_c)^{\beta},
889: \ee
890: which must therefore satisfy:
891: \be
892: 2\beta=\nu(z+\eta-1),
893: \ee
894: {\em i.e.} the usual hyperscaling relation in $d=1$.
895: %
896:
897: \subsubsection{Analytical estimate of the exponent $\eta$}
898: Following the same philosophy as in the mean field argument given in
899: section~\ref{sec:heur}, we can calculate the expectation value of
900: the long range perturbation
901: \be
902: J(r,r')\langle {\vec S}_{r}\cdot {\vec
903: S}_{r'}\rangle\sim \frac{1}{|r-r'|^{\alpha+z+\eta-1}},
904: \ee
905: at the QCP, with some unknown critical exponents
906: $z$ and $\eta$. The finite size correction to the free energy density
907: now scales like $L^{2-\alpha-z-\eta}$. The
908: singular part of the free energy at some non
909: trivial QCP is expected to scale like $L^{-1-z}$ for a finite
910: size system. Thus the two corrections will scale in a similar way if
911: %
912: \be
913: \label{MFeta}
914: \eta=3-\alpha.
915: \ee
916: %
917: The same condition is also obtained by
918: demanding that the long range interaction be invariant under the RG
919: transformation involving a scale factor $s$:
920: \be
921: \vec n(\tau ,x) \to s^{-(z-1+\eta )/2}\vec n(\tau /s^z,x/s).
922: \ee
923: (The rescaling factor of $s^{-(z-1+\eta )/2}$ for $\vec n$ implies the equal
924: time correlation exponent of $z-1+\eta$.)
925: Rescaling $x$ and $\tau$ inside the integral Eq.~(\ref{Seff}) (which
926: represents the contribution of the long range term in the action), the
927: condition for invariance under this RG
928: transformation leads to $2+z-\alpha-(z-1+\eta)=0$, which gives the
929: expression (\ref{MFeta}) for $\eta$.
930: As already shown in section~\ref{sec:SW},
931: Large-$N$ calculations also give the same value for $\eta$~\cite{noteta}.
932: Let us mention that the RG analysis, presented below in section~\ref{sec:RG}, also agree
933: with such an estimate, up to order $(\alpha-2)^2$.
934:
935: \subsubsection{Numerical determination of the exponent $\eta$: scaling of
936: the staggered susceptibility}
937: %
938: \begin{figure}[!ht]
939: \bc
940: \includegraphics[width=10cm,clip]{Figures/ChiStag.alpha2.2.eps}
941: \caption{QMC results for the $T=0$ staggered susceptibility
942: $\chi(\pi)$ [Eq.~(\ref{def:stachi})] computed for $\alpha=2.2$ on
943: systems up to $L=1024$ spins. The upper inset shows $\chi(\pi)$ vs $L$ for
944: various $\lambda\in[0.1,1.9]$. The main plot shows the results of a
945: data collapse onto two universal curves, after a rescaling of both
946: $x$- and $y$-axis using two parameters $\xi_{\lambda}$ and
947: $\Theta_{\lambda}$.
948: Asymptotically, the LRO curve (top one with data
949: for $0.95\le \lambda\le 1.9$) saturates
950: towards a constant whereas the QLRO one (lower one with data for
951: $0.1\le\lambda\le 0.85$) displays a
952: $L^{-1}$ behavior, characteristic of $\eta=1$. Between them the
953: separatrix shows the critical behavior around the transition at
954: $\lambda_c\simeq 0.9$, decaying like $L^{-0.8\pm 0.01}$. The lower
955: inset shows a log-log plot of the scaling parametrs $\Theta_{\lambda}$ and
956: $\xi_{\lambda}$ used to achieve the collapses in both phases,
957: N\'eel ({\Large$\circ$}) and QLRO ($\square$).
958: Within error bars, data
959: can be fitted for the entire range by power-laws (dotted lines) of
960: the form $\xi_{\lambda}^{0.785}$ for the N\'eel regime and
961: $\xi_{\lambda}^{0.815}$ for the QLRO.}
962: \label{fig:chi}
963: \ec
964: \end{figure}
965: %
966:
967: The $T=0$ staggered susceptibility, defined on a finite ring of size
968: $L$ by
969: %
970: \be
971: \label{def:stachi}
972: \chi(\pi)=\frac{1}{L}\sum_{ij}(-1)^{|i-j|}\int_{0}^{\infty}\langle{\vec{S}}_{i}(0)\cdot{\vec{S}}_{j}(\tau)\rangle
973: d\tau,
974: \ee
975: %
976: obeys the standard finite size scaling at the QCP:
977: %
978: \be
979: \chi(\pi)\propto L^{2-\eta}.
980: \ee
981: %
982: Also we know, for instance from SW calculation (see
983: appendix~\ref{app:SW}), that in a N\'eel ordered state the staggered
984: susceptibility will scale quadratically with the size
985: $L$. On the other hand, in the QLRO characterized by $\eta=1$, we rather expect a linear
986: scaling of $\chi(\pi)$ with $L$. Consequently, there are
987: three distinct regimes for the staggered susceptibility:
988: %
989: \be
990: \label{eq:chi}
991: \chi(\pi)\times L^{-2}\sim\left\{
992: \begin{array}{lr}
993: {\rm{constant}}&{\rm{~if~}} \lambda > \lambda_c {\rm{~~~~(NEEL)}}\\
994: L^{-1} &{\rm{~if~}} \lambda < \lambda_c {\rm{~~~~(QLRO)}}\\
995: L^{-\eta} &{\rm{~if~}} \lambda = \lambda_c {\rm{~~~~(QCP).}}
996: \end{array}
997: \right.
998: \ee
999: %
1000: We use the same scaling procedure as for the
1001: correlation functions, to obtain data collapses onto two different
1002: curves, as illustrated in Fig.~\ref{fig:chi} for
1003: $\alpha=2.2$.
1004: Indeed $\chi(\pi)$,
1005: computed with QMC on chains of up to $L=1000$
1006: sites, displays clearly a crossover phenomenon on both sides of the
1007: transition, also characterized by a crossover length scale which is
1008: directly proportional to the one previously extracted in the analysis
1009: of the correlation functions. Note also that such an analysis provides
1010: a second physical observable way to locate the QCP: in fact, the analysis of $C(L)$
1011: and $\chi(\pi)$ both agree (within the error bars) on the value of
1012: $\lambda_c$.
1013: Moreover, we expect the scaling hypothesis to be
1014: valid if $\Theta_{\lambda}\sim \xi_{\lambda}^{\eta}$.
1015: This is actually the case, as illustrated in the lower inset of Fig.~\ref{fig:chi}
1016: where we find $\eta=0.8\pm 0.015$.
1017: %
1018: \begin{figure}[!ht]
1019: \bc
1020: \includegraphics[width=8cm,clip]{Figures/eta.eps}
1021: \caption{Numerical estimate for the critical exponent $\eta$ along the
1022: critical line obtained using the critical behavior of $\chi(\pi)$
1023: computed with QMC (red
1024: symbols). The red dashed line is a guide to the eyes. The blue full
1025: line is the analytical estimate $\eta=3-\alpha$.}
1026: \label{fig:eta}
1027: \ec
1028: \end{figure}
1029: %
1030: We can also obtain the quantum critical exponent
1031: $\eta$ from the separatrix between the N\'eel and QLRO
1032: regimes (see Fig.~\ref{fig:chi}), which is expected to decay as
1033: $L^{-\eta}$. For $\alpha=2.2$, we find $\eta=0.8\pm 0.01$.
1034: We have repeated this computation of $\chi(\pi)$ for several other values of $\alpha
1035: \in [2.1, 2.7]$ to calculate the corresponding $\eta(\alpha)$. The results are
1036: plotted in Fig.~\ref{fig:eta}, and compared to the previously
1037: discussed estimate
1038: $\eta=3-\alpha$. It is very remarkable to see
1039: how this rough estimate reproduces quite well the actual value. Only
1040: for $\alpha>2.3$ a deviation starts to appear.}
1041: %
1042: \subsubsection{Dynamical exponent $z<1$}
1043: %
1044: The dynamical critical exponent $z$, involved for instance in the
1045: critical decay of the correlation function Eq.~(\ref{eq:QCP}),
1046: can be evaluated from the spin-spin correlation function at
1047: the QCP. From the fit of the separatrix between the two data
1048: collapses (see for instance Fig.~\ref{fig:cL} where for $\alpha=2.1$
1049: we estimated $\eta+z-1=0.63\pm 0.03$) we obtain an estimate for
1050: $\eta+z-1$. Then, using the estimates of $\eta$, determined separately with the
1051: staggered susceptibility, we obtain a numerical evaluation of $z$.
1052: Results are shown in Fig.~\ref{fig:z} for $2\le\alpha\le2.7$. For
1053: $\alpha=2$, the QCP at $\lambda=0$ displays the critical behavior of
1054: the short range model, with $\eta=z=1$. Surprisingly, when moving
1055: from $\alpha=2$ along the transition line, $z$ becomes very rapidly
1056: $<1$ and, within the error bars, seems to saturate around a value
1057: $\sim 0.75$. It is actually natural to expect $z\neq 1$ since the long
1058: range interaction breaks Lorentz invariance. However, unlike for the
1059: estimate of $\eta$, the dynamical
1060: exponent obtained within the large-$N$ expansion $z=(\alpha-1)/2$ does
1061: not agree with the QMC results. As we discuss in the next section,
1062: using a ``RG improved'' perturbation theory, $z$ is
1063: found to be $<1$ but does not display such a big reduction.
1064: %
1065: \begin{figure}[!ht]
1066: \bc
1067: \includegraphics[width=8cm,clip]{Figures/z.eps}
1068: \caption{Numerical estimates for the dynamical critical exponent $z$
1069: along the critical line obtained using the critical behavior of the
1070: correlation function [Eq.~(\ref{eq:QCP})] computed with QMC,
1071: and the numerical estimate of $\eta$ (see Fig.~\ref{fig:eta}). The
1072: numerical results (open circles) suffer from large error bars, as
1073: shown on the plot. The green dashed line is a guide to the eyes.}
1074: \label{fig:z}
1075: \ec
1076: \end{figure}
1077: %
1078: %
1079: \section{Field Theory/Renormalization Group Results}
1080: \label{sec:RG}
1081: The low energy, continuum limit, imaginary
1082: time action takes the form: \be S[\vec n]=S_0[\vec n]-g\int d\tau
1083: dx \phi (\tau ,x) -\lambda a^{\alpha -2}\int {d\tau dx dy\over
1084: |x-y|^{\alpha}}\vec n(\tau ,x)\cdot \vec n(\tau ,y).
1085: \label{Seff2}\ee
1086: $\vec n(\tau ,x)$ is the antiferromagnetic order parameter field
1087: defined by the continuum limit expression of Eq. (\ref{cont_lim}).
1088: Here $S_0$ is the action for a free massless relativistic boson, in
1089: terms of which $\vec n$ may be represented in a non-linear way.
1090: Equivalently, we may regard it as the action of the $k=1$
1091: Wess-Zumino-Witten non-linear $\sigma$ model. The field $\phi$ is
1092: defined as: \be \phi = 2\pi\vec J_L\cdot \vec
1093: J_R,\ee and is normalized so: \be <\phi (z)\phi (0)>={3\over
1094: 16\pi^2|z|^4}.\ee The corresponding coupling constant, $g$ has a bare value
1095: of order 1 for the short range AF chain and is
1096: marginally irrelevant. It is responsible for various logarithmic
1097: corrections such as the one in the correlation function of
1098: Eq. (\ref{corr}). Note that the
1099: dimensionless coupling constant for the long range interaction
1100: $\lambda$ is only proportional to the one used before for the lattice
1101: microscopic model in Eq.~(\ref{QLRO}), and $a$
1102: is a short distance cut-off with dimensions of length. As already
1103: noted in section~\ref{sec:heur},
1104: since $\vec n$ has a scaling dimension of 1/2 from Eq. (\ref{ncorr}),
1105: $\lambda$ is irrelevant for $\alpha >2$, relevant for $\alpha <2$, and
1106: marginal for $\alpha =2$. Also note that $\lambda >0$ corresponds to
1107: non-frustrating interactions which favor the N\'eel state where
1108: $\langle n^z\rangle\neq 0$. Our strategy is to study this model when $0<\alpha -
1109: 2\ll 1$ and $0<\lambda \ll 1$ using perturbation theory in $g$ and
1110: $\lambda$. Since $g$ renormalizes to $0$ at large length scales, when
1111: $\lambda =0$ (i.e. in the short range model) this can give useful
1112: results, for small bare $\lambda$,
1113: even when the bare value of $g$ is O(1).
1114: We will find that an interplay between the local marginal
1115: coupling constant $g$ and the irrelevant non-local coupling constant
1116: $\lambda$ governs the critical behavior in this regime.
1117:
1118: We now consider the low energy effective field theory for the long
1119: range model in Eq. (\ref{Seff2}), in the limit of small $g$ and
1120: $\lambda$, using RG methods. When $\lambda =0$, the
1121: RG equations reduce to the standard ones for
1122: the short range model. These take the form: \be {dg\over d\ln a}
1123: =-g^2 -(1/2)g^3+\ldots .\label{RGg}\ee Here we define our RG
1124: transformation by increasing the short distance cut-off $a$. The bare
1125: value of $g$ is positive for any
1126: non-frustrated
1127: short range model and is typically O(1). The basin of
1128: attraction of the $g=0$ fixed point is known to extend to such large
1129: bare values of $g$ so that $g=0$ is the universal stable fixed point
1130: for short range models. The flow of $g$ towards zero at long length
1131: scales is controlled by the quadratic term in the $\beta$-function of
1132: Eq. (\ref{RGg}), giving: \be g(a)\to {1\over \ln
1133: (a/a_0)},\ee where $a_0$ is the original cut-off and $a$ is a larger
1134: value obtained from integrating out modes with wave-lengths between
1135: $a_0$ and $a$. This logarithmically slow flow of $g(a)$ to zero is
1136: responsible for logarithmic corrections to the correlation function
1137: and other properties of the short-range models. A linear term in the
1138: $\beta$-function for $\lambda$ follows immediately from the factor of
1139: $a^{\alpha -2}$ in Eq. (\ref{Seff2}) which in turn is a consequence of
1140: the fact that $n$ has scaling dimension $1/2$: \be {d\lambda \over
1141: d\ln a}= (2-\alpha )\lambda +\ldots \label{RGl0}\ee So, ignoring the
1142: effects of $g$, $\lambda$ grows larger at long length scales for
1143: $\alpha <2$ but smaller for $\alpha >2$. Long range interactions are
1144: irrelevant for $\alpha >2$. However, it is necessary to consider
1145: higher order terms in the $\beta$-functions for both $g$ and $\lambda$
1146: to understand the phase diagram, even at $\alpha \approx 2$.
1147:
1148: To calculate additional terms in the $\beta$-functions, we define our
1149: ultra-violet cut-off by forbidding any 2 points in space-imaginary
1150: time from getting closer than $a$, in a perturbative calculation of
1151: the partition function (or long distance Green's functions). In
1152: particular, this means that the long range term in the action is
1153: cut-off as: \be S_{\lambda}=-\lambda a^{\alpha -2}\int_{|x-y|>a}
1154: {d\tau dx dy\over |x-y|^{\alpha}}\vec n(\tau ,x)\cdot \vec n(\tau ,y).
1155: \label{Sl}\ee
1156: When the cut-off is increased from $a_0$ to $a=a_0+\delta a$, there is
1157: an additional change in $S$ of first order in $\delta a$, which comes
1158: from the change in the integration region: \be \delta S = -\lambda
1159: a_0^{\alpha -2}\int dx d\tau \left[\int_{-a}^{-a_0} + \int_{a_0}^{a}
1160: \right] {du\over |u|^{\alpha}}\vec n(\tau ,x)\cdot \vec n(\tau ,x+u)
1161: .\label{deltaS}\ee Since both factors of $\vec n$ are very close
1162: together, we may use the operator product expansion. This follows
1163: from the 3-point Green's function: \be <n^a(z_1)\phi
1164: (z_2)n^b(z_3)>={1\over 8\pi}{|z_{13}|\over
1165: |z_{12}|^2|z_{23}|^2}.\ee This implies the operator product expansion
1166: (OPE): \be n^a(z)n^b(0)\to {\delta^{ab}2\pi \over 3}|z|\phi (0) +
1167: \ldots \label{OPEnn}\ee
1168: Using this in Eq. (\ref{deltaS}), gives: \be \delta S =
1169: -\lambda a_0^{\alpha -2}\int d\tau dx \phi (\tau ,x)
1170: 4\pi \int_{a_0}^a{du\over |u|^{\alpha -1}} \approx -{\delta a\over
1171: a}4\pi \lambda \int d\tau dx \phi (\tau ,x).\ee This corresponds to
1172: a renormalization of $g$: \be \delta g = 4\pi {\delta a\over
1173: a}\lambda \ee and hence to another term in the $\beta$-function for
1174: $g$: \be {dg\over d\ln a} = 4\pi \lambda -g^2
1175: +\ldots \label{RGp}\ee There is one more term in the RG equations that
1176: is important at small $\alpha -2$, corresponding to a correction to
1177: $\lambda$ of order $\lambda g$. This can be calculated from the OPE:
1178: \be \phi (z)n^a(0)\to {1\over 8\pi |z|^2}n^a(0) + \ldots ,\ee
1179: giving: \be d\lambda /d\ln a =-(\alpha -2)\lambda +(1/2)\lambda g + \ldots
1180: \label{RGl}\ee
1181: The RG equations, Eq. (\ref{RGp}) and (\ref{RGl}) have an unstable
1182: fixed point for $\alpha >2$, at: \bea g_c&\approx&2
1183: (\alpha -2)\nonumber \\ \lambda_c &\approx& {1\over \pi}(\alpha
1184: -2)^2.\eea For $\alpha <2$, a positive $\lambda$ always runs away to
1185: large values as we lower the cut off (i.e. increase $a$),
1186: corresponding to LRO. On the other hand, for $\alpha
1187: >2$, a small enough positive bare $\lambda$ flows to zero while a
1188: larger bare value flows to large values (see Fig.~\ref{fig:RGflow}).
1189: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1190: \begin{figure}[!ht]
1191: \bc
1192: \includegraphics[width=8cm,clip]{Figures/RGflow.eps}
1193: \caption{Renormalization group flow of Eqs. (\ref{RGp}) and (\ref{RGl})
1194: in the case $\alpha=2.3$.
1195: The dotted line represents, schematically, the values of the bare
1196: couplings in the field theory as the parameter lambda in the lattice model
1197: is varied. The unstable fixed point at $\lambda_c=0.0286$, $g_c=0.6$
1198: separate the flow to the stable fixed point at $\lambda=g=0$ which
1199: represents the quasi-long-range ordered phase and the flow to infinite
1200: $\lambda$, $g$, which represents the long range ordered phase. The black
1201: lines with double arrows denote the separatrixes between these two phases.
1202: (The corrections to the flow equations are presumably significant for this
1203: large a value of $\alpha -2$, but we graph this case for ease of
1204: visualization.}
1205: \label{fig:RGflow}
1206: \ec
1207: \end{figure}
1208: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1209: These statements remain true
1210: even when the bare value of $g$ is O(1) as we expect it to be in
1211: general for a short range spin chain. For a small bare $\lambda$, $g$
1212: initially renormalizes towards small values as it would in the short
1213: range chain until eventually Eqs. (\ref{RGp}), (\ref{RGl}) becomes
1214: valid. The stable $\lambda =g =0$ fixed points corresponds to the
1215: standard QLRO phase of the short range spin chain.
1216: The non-trivial unstable fixed point separates the ordered and quasi
1217: long range ordered phases.
1218: Of course there are higher order terms in
1219: both RG equations, but they do not invalidate our conclusions on the
1220: location of the fixed point, for small enough $\alpha -2$. Both terms
1221: on the right hand side of Eq. (\ref{RGp}) are $O[(\alpha -2)^2]$ at
1222: the fixed point; any possible higher order terms such as $g^3$ or
1223: $\lambda^4$ are at least of $O[(\alpha -2)^3]$. Similarly, both
1224: terms on the right hand side of Eq. (\ref{RGl}) are $O[(\alpha
1225: -2)^3]$ at the fixed point; higher order terms are at least of
1226: $O[(\alpha -2)^4]$. To reach this conclusion it is important to
1227: realize that there cannot be any terms in $d\lambda /d\ln a$ which
1228: contain no factors of $\lambda$; a purely short range interaction
1229: cannot generate a long range one although the reverse is not true.
1230:
1231: Thus we appear to
1232: have a rare example of a non-trivial fixed point which can be
1233: accessed perturbatively. (But see the discussion
1234: below of potential problems with this approach.)
1235: We note that a similar expansion
1236: for long range classical spin models was introduced in \cite{Fisher72}.
1237: See also \cite{Sak,Bhatt}.
1238: Our quantum spin model, in the continuum limit, non-linear $\sigma$
1239: model approximation, appears rather similar, in the
1240: imaginary time path integral formulation. An important
1241: difference, however, is that our model has an action which
1242: is long range in the space direction but short-range in the time
1243: direction. Thus it corresponds to a classical model in
1244: 2 space dimensions with short range interactions in one
1245: direction and long range interactions in the other. It is
1246: this asymmetry which leads to a dynamical critical exponent $z<1$.
1247: Another important difference from a 2-dimensional Heisenberg
1248: model is the topological term in the short-range part of
1249: the action which is responsible for the quasi-long-range order.
1250: We remark that an integer-spin quantum Heisenberg chain
1251: with long range interactions could be expected to have
1252: identical critical behavior to a classical Heisenberg model
1253: in two dimensions with interactions which are long range in one dimension.
1254: We also remark that an xxz quantum spin chain with long
1255: range interactions could be expected to have the
1256: same critical behavior as a two-dimensional classical xy
1257: model with interactions which are long range in one dimension.
1258:
1259: The phase boundary (or separatrix) can be found by determining the
1260: line in the $g-\lambda$ plane which renormalizes to the critical
1261: point. Combining Eqs. (\ref{RGp}) and (\ref{RGl}) gives: \be
1262: \int_{\lambda_0}^{\lambda_c}{d\lambda \over
1263: \lambda}=(1/2)\int_{g_c}^{g_0}{dg(g-g_c)\over g^2-g_c^2(\lambda
1264: /\lambda_c)},
1265: \label{flow}\ee
1266: where $\lambda = \lambda (g)$ is a function of $g$ along the RG flow
1267: in the integral on the right hand side of Eq. (\ref{flow}). $g_0$ and
1268: $\lambda_0$ are the values at arbitrary points on the separatrix.
1269: Since $\lambda$ increases monotonically to the value $\lambda_c$ with
1270: increasing $a$, we may obtain an upper and lower bound on the right
1271: hand side by replacing $\lambda (g)$ by $\lambda_c$ and $0$
1272: respectively inside the integral on the right hand side of
1273: Eq. (\ref{flow}): \be (1/2)\ln (g_0/eg_c)+g_c/2g_0<\ln
1274: (\lambda_c/\lambda_0)<(1/2)\ln [(g_0+g_c)/2g_c].\ee Now using the fact
1275: that $g_c<<g_0$, this becomes: \be \sqrt{g_0/eg_c}<\lambda_c/\lambda_0
1276: <\sqrt{g_0/2g_c},\ee that is: \be .637 (\alpha
1277: -2)^{5/2}/\sqrt{g_0}<\lambda_0<.743 (\alpha
1278: -2)^{5/2}/\sqrt{g_0}.\label{bound}\ee Thus on the separatrix
1279: $\lambda_0$ is $O[(\alpha -2)^{5/2}]$. [A numerical solution of the
1280: RG equations indicates that $\lambda_0$ is very close to the lower
1281: bound in Eq. (\ref{bound}).] Assuming a bare $g$ of O(1), it is then
1282: possible to predict the shape of the phase boundary in the $\lambda -
1283: \alpha $ plane close to $\alpha =2$. There is an unknown
1284: multiplicative factor relating the lattice coupling $\lambda$
1285: to the continuum coupling $\lambda$. However, we can predict that \be
1286: \label{eq:alphac}
1287: \alpha_c(\lambda )\to 2+C\lambda^{2/5},
1288: \ee for some unknown constant
1289: factor, $C$, as $\alpha \to 2$. As mentioned above, we expect that
1290: $2<\alpha _c(\lambda )<3$ for all $\lambda$ and all $S$.
1291:
1292: While our QMC results also predict
1293: that $\alpha_c\to 2$ when $\lambda \to 0$,
1294: but otherwise Eq.~(\ref{eq:alphac}) does not agree well
1295: with the QMC result as shown in fig.~\ref{fig:zoom}.
1296: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1297: \begin{figure}[!ht]
1298: \bc
1299: \psfrag{A}{\tiny{$2+0.14\lambda^{2/5}$}}
1300: \psfrag{B}{\tiny{$2+0.11\lambda^{2/5}$}}
1301: \includegraphics[width=8cm,clip]{Figures/ZoomPhDg.eps}
1302: \caption{QMC phase diagram at small $\lambda$ compared
1303: to the RG prediction of Eq. (\ref{eq:alphac}) with $C=0.14$ and $C=0.11$.}
1304: \label{fig:zoom}
1305: \ec
1306: \end{figure}
1307: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1308: It is interesting to note that lowest order SW theory and the large-$N$ approximation
1309: make the mean-field prediction
1310: that $\alpha_c \to 1$ as $\lambda \to 0$, in clear disagreement with
1311: our RG and QMC results.
1312:
1313: Linearizing the $\beta$-functions at the critical point gives: \be
1314: {d\over d\ln a}\left(\begin{array}{c}\lambda -\lambda_c\\
1315: g-g_c\end{array}\right)\approx \left(\begin{array}{cc}
1316: 0&(\alpha -2)^2/(2\pi)\\ 4\pi&-4(\alpha -2)\end{array}\right)
1317: \left(\begin{array}{c}\lambda -\lambda_c\\ g-g_c\end{array}\right) \ee
1318: This matrix has one positive (unstable) right eigenvalue,
1319: $(\sqrt{6}-2)(\alpha -2)$, implying a crossover length scale: \be
1320: \xi \propto |\lambda -\lambda_c|^{-\nu},\ee with a critical exponent:
1321: \be \nu = {1\over (\sqrt{6}-2)(\alpha -2)},\ee which diverges as
1322: $\alpha \to 2$. The corresponding unstable direction is: \be
1323: \lambda -\lambda_c={(\alpha -2)\over 2\pi(\sqrt{6}-2)}(g -g_c).\ee
1324: For $\alpha =2$ exactly, LRO occurs for any $\lambda >0$
1325: but behavior characteristic of the quasi long range ordered
1326: fixed point occurs out to a cross over length scale:
1327: \be
1328: \xi \propto \exp [C/\lambda_0^{2/5}],
1329: \label{eq:exp}
1330: \ee
1331: for a constant factor $C$.
1332:
1333: We may also determine the critical exponent, $\eta +z-1$, controlling
1334: the equal-time correlation function at the non-trivial critical
1335: exponent: \be <S_0^aS_j^b>\propto {\delta^{ab}\over |j|^{\eta
1336: +z-1}},\label{eta}\ee when $\alpha -2<<1$. Since $\lambda$ and $g$ are
1337: both small at the critical point, for $\alpha -2<<1$, we may simply do
1338: ``RG improved'' perturbation theory. That is we
1339: calculate the correlation function to first order perturbation theory
1340: in $g$, replace $g$ by $g_c$, and interpret the result as the
1341: expression Eq. (\ref{eta}) with \be \eta + z = 2+ O(g_c).\ee [To
1342: lowest order in $(\alpha -2)$ we only need consider $g$, not
1343: $\lambda$, since $\lambda_c \propto g_c^2$.] The Green's function, up
1344: to first order in perturbation theory in $g$, the bare coupling, is:
1345: \bea <n^a(z)n^b(0)>&=&{\delta^{ab}\over |z|}+g\int
1346: d^2z'<n^a(z)n^b(0)\phi (z')>\nonumber \\ &=&{\delta^{ab}\over
1347: |z|}+{g\over 2\sqrt{3}}\delta^{ab}\int d^2z'{|z|\over
1348: |z'|^2|z-z'|^2}+\ldots \label{corr2}\eea The integral in
1349: Eq. (\ref{corr2}) must be restricted to the region $|z'|>a$,
1350: $|z-z'|>a$. For $|z|>>a$, the integral is dominated by the two
1351: regions, $|z'|<<|z|$ and $|z'-z|<<|z|$, giving \bea
1352: <n^a(z)n^b(0)>&\approx &{\delta^{ab}\over |z|}\left[1+{g\over
1353: 2\sqrt{3}}2\int {d^2z'\over |z'|^2}\right] \nonumber \\ &\approx
1354: &{\delta^{ab}\over |z|}\left[ 1+{g\over 2\sqrt{3}}4\pi \ln
1355: (|z|/a)\right] .\eea Now replacing $g$ by $g_c$, its fixed point value
1356: we obtain: \be <n^a(z)n^b(0)>={\delta^{ab}\over |z|}\left\{1+(\alpha
1357: -2)\ln (|z|/a)+O[(\alpha -2)^2]\right\} \ee Now, using the fact that
1358: the correlation function should have a pure power-law form at the
1359: fixed point, we may interpret this result as the leading term in the
1360: expansion of: \be <n^a(z)n^b(0)>={\delta^{ab} A\over |z|^{1-(\alpha
1361: -2)}}.\ee To this order in $(\alpha -2)$ we obtain the same exponent
1362: for $z=ix$ or $z=\tau$, implying that the corrections to the dynamical
1363: exponent, $z=1$, are higher order in $(\alpha -2)$. Thus: \be \eta =
1364: 1-(\alpha -2) + O[(\alpha -2)^2].\ee
1365: This is the same value of $\eta$ found in the large $N$ approximation
1366: and by a simple scaling argument in sub-section (V-C-3). It
1367: agrees well with QMC results for $\alpha <2.3$ as shown in
1368: fig.~\ref{fig:eta}.
1369: It is natural to expect that
1370: $z\neq 1$ since the long range interaction breaks Lorentz invariance.
1371: Since $\omega \propto |k|^{(\alpha -1)/2}$ in the ordered phase it is
1372: natural to expect that $z<1$ at the critical point, as
1373: is found in the large $N$ approximation. However, clearly
1374: $1-z$ must be at least of $O[(\alpha -2)^2]$ since the long range
1375: coupling constant, $\lambda$ is of that order. In fact, we suspect that
1376: $1-z$ is even higher order than quadratic in $(\alpha -2)$.
1377: This conclusion does not fit well with the QMC results
1378: for $z$, presented in fig.~\ref{fig:z}. There it was found
1379: (although with large error bars) that $z$ appears to have
1380: a nearly constant value, $z\approx .75$, for $\alpha \geq 2.1$.
1381: As $\alpha$ is further decreased $z$ appears to rise very rapidly
1382: towards $1$.
1383:
1384: So far, we have ignored another possible interaction:
1385: \be S \to S -(g'\pi /3)\int d\tau dx (\vec J_L^2+\vec J_R^2).
1386: \label{g'def}\ee
1387: This interaction is, in fact, present for the short-range spin-chain
1388: with a large coefficient. Since the Hamiltonian for the
1389: $k=1$ WZW model can be written quadratic in currents, this
1390: ``interaction'' term can be regarded as simply shifting the
1391: velocity, which we have so far set equal to $1$, to:
1392: \be v \to 1-g'/2.\ee
1393: The RG equations, for the short range model, including $g'$ take
1394: the form, to cubic order:
1395: \bea
1396: {dg\over d\ln a}&=& -g^2-(1/2)g(g^2+g'^2)\label{RGg'1} \\
1397: {dg'\over d\ln a}&=& (3/4)g^3.\label{RGg'2}\eea
1398: Starting with $g$, $g'>0$ and O(1), these equations
1399: predict that $g\to 0$ and $g'$ flows to a value of O(1).
1400: The large value of $g'$ at the fixed point can simply
1401: be interpreted as a large renormalization of the velocity,
1402: provided that $g'<2$. In fact, this is what happens,
1403: for example in the Hubbard model at half-filling. The
1404: spin velocity is reduced by the Hubbard interactions.
1405: An alternative approach is to adjust $v$ to the
1406: correct value and drop $g'$ completely from the RG
1407: equations. In fact, Eq. (\ref{RGg'2}) depends strongly
1408: on the renormalization and cut-off scheme. With
1409: a Lorentz invariant cut-off and renormalization procedure,
1410: the non Lorentz-invariant term, proportional to $g'$ will
1411: not be generated under the RG if it is initially absent.
1412: Breaking of Lorentz invariance in this problem at
1413: low energies, just means shifting the velocity. If
1414: we work directly with the exact velocity, then it is
1415: apparently permissible to set $g'=0$ and use
1416: a Lorentz invariant renormalization procedure so that
1417: $g'$ remains zero under renormalization. In fact,
1418: using this procedure in Eq. (\ref{RGg'1})
1419: leads to various predictions of
1420: logarithmic corrections which are in good agreement
1421: with Bethe ansatz and numerical results for the S=1/2 chain.
1422: If we set $g'$ equal to some arbitrary non-zero value
1423: in Eq. (\ref{RGg'1}) the coefficients and powers of
1424: log corrections would change, resulting in worse
1425: agreements with numerical results. Thus, this
1426: procedure of ignoring $g'$ seems to be a valid and
1427: useful one.
1428:
1429: We now consider the interplay of the long range coupling
1430: constant, $\lambda$, with the non-Lorentz invariant
1431: local coupling, $g'$. The needed OPE's can be
1432: obtained from the general conformal field theory
1433: result for the 3-point Green's function of
1434: the energy momentum operator with a primary field of left-dimension $1/4$:
1435: \be <T(z)n^a(z_1)n^b(z_2)>={1\over 2\pi}
1436: \sum_{i=1}^2
1437: \left[{1/4\over (z-z_i)^2}+{1\over z-z_i}{\partial \over \partial z_i}
1438: \right]{\delta^{ab}\over |z_1-z_2|}.\ee
1439: Here $T=(2\pi /3)\vec J_L^2$, is the left-moving part of
1440: the Hamiltonian. This gives:
1441: \be
1442: <T(z)n^a(z_1)n^b(z_2)>={1\over8\pi}{\delta^{ab}(z_1-z_2)^2
1443: \over |z_1-z_2|(z-z_1)^2(z-z_2)^2}.\ee
1444: Also using the 2-point function of $T$:
1445: \be <T(z_1)T(z_2))>={1\over 2(2\pi)^2(z_1-z_2)^4},\ee
1446: we can deduce the OPE:
1447: \be n^a(z)n^b(0)\to {\delta^{ab}\pi z^2\over |z|}T(0).\ee
1448: Now consider the case where the separation is in the space direction,
1449: $z=ix$
1450: \be n^a(x)n^b(0)\to \delta^{ab}\pi |x|[(2/3)\phi (0)
1451: -T(0)-\bar T(0)]+\ldots .\ee
1452: Here we have included the identical OPE coefficient for
1453: $\bar T \equiv (2\pi /3)\vec J_R^2$ and also the coefficient
1454: deduced earlier, in Eq. (\ref{OPEnn}).
1455: Using the same cut-off and RG transformation procedure as above,
1456: this implies a term in the $\beta$-function:
1457: \be {dg'\over d\ln a}=-6\pi \lambda .\ee
1458: This drives $g'$ towards negative values, corresponding to
1459: increasing the velocity. If we continue to use
1460: our previous RG transformation, we find that the
1461: $d\lambda /d\ln a$ {\it does not} pick up an term
1462: $\propto \lambda g'$. The difference from the
1463: non-zero $\lambda g$ term in Eq. (\ref{RG3}) arises from
1464: the fact that the OPE is now:
1465: \be T(z)n^a(0)\to {1\over 8\pi z^2}n^a(0).\ee
1466: The RG transformation gives:
1467: \be \delta \lambda \propto \int {d^2z\over z^2},\ee
1468: where, as before, the integral is over a circular shell
1469: with radius between $a$ and $a+\delta a$. This integral
1470: vanishes by rotational invariance. Thus the
1471: complete set of RG equations to low order is:
1472: \bea {d\lambda \over d\ln a}&=&-(\alpha -2)\lambda +\lambda g/2\nonumber \\
1473: {dg\over d\ln a}&=& 4\pi \lambda -g^2-(g+g')g^2/2\nonumber \\
1474: {dg'\over d\ln a}&=&-6\pi \lambda .\label{RG3}\eea
1475: These equations have an unstable fixed point at:
1476: \bea \lambda_c&=&0\nonumber \\
1477: g_c&=&2(\alpha -2)\nonumber \\
1478: g'_c&=&-2-2(\alpha -2)\approx -2.\eea
1479: In this approximation, we obtain the same prediction for $\eta
1480: \approx 1-(\alpha -2)$, as before and still get $z\approx 1$.
1481: There is a shift in the velocity of O(1).
1482:
1483: The large value of $g'$ at the fixed point makes the predictions
1484: of this RG analysis intrinsically suspect. As in the short-range
1485: case, we might agree to set $v$ equal to its renormalized
1486: value and then drop $g'$ from the RG equations. This leads
1487: to the same predictions about the value of $\lambda$ on
1488: the separatrix, and $\nu$ as obtained above.
1489:
1490: However, there are some worrisome features of this RG analysis
1491: which arise from the long range interaction. We expect
1492: a dynamical exponent $z<1$
1493: at the critical point. It then does not make sense to use
1494: a rotationally invariant (i.e. Lorentz invariant) RG
1495: transformation. We would then get back a $g^3$ term
1496: in $dg'/d\ln a$ and, very importantly, a $\lambda g'$ term
1497: in $d\lambda /d\ln a$. We would then generally find that
1498: $g$ is not small, O($\alpha -2$), at the fixed point.
1499: $\lambda$ would also not be small at the fixed point.
1500: In this case we would lose all perturbative control
1501: over the critical behavior even for $\alpha$ only
1502: slightly larger than $2$. In this case, the unstable critical
1503: point would not be close to the QLRO fixed point, for
1504: $\alpha$ close to $2$. One possibility is that the
1505: effects associated with $z<1$ can be ignored
1506: to lowest non-trivial order $\alpha -2$. Then our
1507: use of a Lorentz invariant RG transformation
1508: may be justified. In this case,
1509: the fixed point really is close to the QLRO critical point
1510: for $\alpha$ sightly larger than 2 and our predictions
1511: for $\eta$ and $\nu$ are correct in this limit.
1512:
1513: The QMC results seem to give at least
1514: partial confirmation of the validity of an RG
1515: approach based on the $\alpha -2$ expansion. Most
1516: importantly, $\alpha_c\to 2$ as $\lambda \to 0$ and
1517: the critical exponents $\eta$ and $z$ appear to approach their values
1518: in the quasi long range ordered phase ($\eta =z=1$) in this limit,
1519: with $\nu$ diverging. Furthermore,
1520: excellent agreement with the prediction for $\eta$ was obtained,
1521: over a rather large range of $\alpha$ (up to $2.2$), as shown
1522: in fig.~\ref{fig:eta}. We were not
1523: able to obtain accurate estimates of $\nu$ from QMC to
1524: test the RG prediction. On
1525: the other hand, $z$ showed rather surprising behavior, in fig.~\ref{fig:z},
1526: dropping rapidly from $1$ to about $.75$ as $\alpha$ is increased
1527: from $2$ to $2.1$. This suggests that the asympotic, small
1528: $\alpha -2$ behavior may only occur for very small values of
1529: $\alpha -2<<.1$. Numerical difficulties preclude obtaining
1530: QMC data in this region. Furthermore, the phase boundary
1531: as determined by QMC, $\alpha_c(\lambda )$
1532: could not be fit well to the RG prediction of Eq.~(\ref{eq:alphac}),
1533: except possibly at very small $\alpha_c$ (where we have no data)
1534: as seen in fig.~\ref{fig:zoom}. This could be interpreted
1535: as meaning that our RG approach based on an $\alpha -2$ expansion
1536: is correct in principle but is only valid in practice for
1537: extremely small values of $\alpha -2$. (The good agreement
1538: for $\eta$ then appears fortuitous.) Alternatively,
1539: the discrepancies may indicate a problem with our RG approach,
1540: perhaps resulting from our cavalier treatment of the
1541: non-Lorentz invariant interaction in Eq.~(\ref{g'def}).
1542:
1543:
1544:
1545: \section{Conclusions}
1546: We have studied long range non-frustrating $S=1/2$ antiferromagnetic chains
1547: using spin-wave theory, large-$N$ approximation, quantum Monte Carlo and
1548: analytic renormalization group methods based on an expansion
1549: in $\alpha -2$. All methods predict to a line of critical
1550: points in the $\lambda$-$\alpha$ plane with continuously varying
1551: critical exponents. This critical line separates phases with true
1552: N\'eel long range order and quasi long range order.
1553: Quantum Monte Carlo and renormalization group methods
1554: indicate that this critical line terminates at $\lambda =0$, $\alpha
1555: =2$ and suggest that,
1556: along the critical line as $\alpha \to 2^+$,
1557: $\eta \approx 3-\alpha$, while $\nu$ diverges and $z\to 1^-$.
1558: \begin{acknowledgements}
1559: We would like to thank J. Cardy for very helpful discussions.
1560: The research of N.L., I.A. and M.B. was supported by NSERC of Canada.
1561: The research of I.A. was supported by the Canadian Institute for
1562: Advanced Research. The numerical simulations were carried out on the
1563: WestGrid network, funded in part by the Canada Foundation for
1564: Innovation.
1565:
1566:
1567: \end{acknowledgements}
1568: \appendix
1569: \section{Calculation of finite size corrections from SW:
1570: contributions from the $k=0$ and finite $k$ modes.}
1571: \label{app:SW}
1572:
1573: \subsection{General method}
1574:
1575: Let ${\cal H}(h)={\cal H}-h\hat{O}$,
1576: where $\hat{O}$ is an operator and $h$ is a field. If we denote
1577: by $|h\rangle$ the GS of ${\cal H}(h)$, it follows that the
1578: GS energy is $E_{GS}(h) = \langle h | {\cal H}-h {\hat O}|
1579: h\rangle$. Since $\langle h
1580: | h\rangle =1$, it is straightforward to show (in direct analogy to
1581: Feynman's theorem) that:
1582: \begin{equation}
1583: \label{M1}
1584: \langle \hat{O} \rangle= \langle h | {\hat O} |h\rangle = -
1585: \frac{\partial E_{GS}(h)}{\partial h}
1586: \end{equation}
1587:
1588: In general we need the expectation values of
1589: various operators ${\hat O}$ in the
1590: unperturbed GS, i.e. in the limit $h \rightarrow 0$. It
1591: follows that all we have to do is to compute the change in the ground-state
1592: energy, due to the perturbation $-h\hat{O}$, to first order in $h$.
1593:
1594: In the remainder of this Appendix, the Hamiltonian ${\cal H}$ is that
1595: of Eq. (\ref{QLRO}). We are interested in finite-size chains with an
1596: even number of sites $L$, and periodic boundary conditions.
1597:
1598:
1599: \subsection{Staggered susceptibility}
1600:
1601: Let $\hat{O}=\sum_{i}^{}(-1)^i S^z_i$. Then, according to
1602: Eq. (\ref{M1}), the staggered magnetization at $T=0$ is:
1603: $$
1604: M_{\pi}= \langle \sum_{i}^{}(-1)^i S^z_i \rangle = -\left.{d
1605: E_{GS}\over dh}\right|_{h\rightarrow 0}
1606: $$
1607: and therefore the staggered susceptibility is:
1608: $$
1609: \chi(\pi) = {1 \over L} \left.{d M_{\pi} \over
1610: dh}\right|_{h\rightarrow 0} = - {1\over L} \left. {d^2
1611: E_{GS} \over d h^2}\right|_{h\rightarrow 0}.
1612: $$
1613:
1614:
1615: \subsubsection{The $k=0$ contribution}
1616: We Fourier transform the spin operators, $\vec{S}_{2n}= 2/L
1617: \sum_{k}^{} \exp{(ik(2na))}
1618: \vec{S}^{e}_k$, $\vec{S}_{2n+1}= 2/L \sum_{k}^{} \exp{(ik(2n+1)a))}
1619: \vec{S}^{o}_k$, and collect only the $k=0$
1620: components. Let us denote $\vec{S}_{1} = \vec{S}^{e}_{k=0}=\sum_{n}^{}\vec{S}_{2n}
1621: $ and $\vec{S}_{2} = \vec{S}^{o}_{k=0}=\sum_{n}^{}\vec{S}_{2n+1}$ the
1622: total spins of the two magnetic sublattices, of $L/2$ spins
1623: each. Since we are in a N\'eel
1624: ordered state, $\vec{S}_1$ and $\vec{S}_2$ are spins of total
1625: magnitude $S_L=LS/2=L/4$ for spins
1626: $S=1/2$. Then, up to some constants that do not depend on $h$:
1627: \be
1628: {\cal H}_{k=0}(h) = {\cal H}_{k=0} -{\cal H}_1
1629: \ee
1630: where
1631: \be
1632: {\cal H}_{k=0} = j \vec{S}_1\cdot \vec{S}_2,
1633: \ee
1634: \be
1635: {\cal H}_1=h \left( S_1^z-S_2^z\right)
1636: \ee
1637: and
1638: \be
1639: j = {2 \over L} \left[1+\lambda \sum_{n\ge 1} {1 \over
1640: (2n+1)^\alpha} \right]= {2J_{eff}\over L}
1641: \ee
1642:
1643: We need to do perturbation theory to second order in $h$, to find the
1644: staggered susceptibility. The
1645: ground-state of ${\cal H}_{k=0}$ is the state:
1646: \be
1647: |0\rangle = |S_T=0, M_T=0, S_L, S_L\rangle
1648: = {1 \over \sqrt{2S_L+1}}
1649: \sum_{m=-S_L}^{S_L}(-1)^m |m, -m\rangle
1650: \ee
1651: where $\vec{S}_T = \vec{S}_1 + \vec{S}_2$.
1652: The perturbation links this only to other states with $M_T=0$ (see below). Let
1653: us denote \be
1654: |S_T\rangle = |S_T,0,S_L,S_L\rangle
1655: \ee
1656: where $S_T=0,1,...,2S_L$. With this notation, the
1657: second-order correction to the ground-state energy is:
1658: \be
1659: \Delta E_{\rm GS}^{(2)} = \sum_{n=1}^{2S_L} \frac{|\langle n | {\cal H}_1| 0\rangle|^2}{E_0-E_n}
1660: \ee
1661:
1662: However,
1663: \be
1664: {\cal H}_1 |0\rangle = {2h\over \sqrt{2S_L+1}} \sum_{m=-S_L}^{S_L} (-1)^m m
1665: |m, -m\rangle
1666: \ee
1667: Interestingly enough, one can show that:
1668: \be
1669: |1\rangle = \alpha_1 \sum_{m=-S_L}^{S_L} (-1)^m m
1670: |m, -m\rangle
1671: \ee
1672: where the normalization constant is:
1673: \be
1674: \alpha_1=\sqrt{3\over S_L(S_L+1)(2S_L+1)}.
1675: \ee
1676: It follows:
1677: \be
1678: \langle 1 |{\cal H}_1|0\rangle = {2h\over \sqrt{2S_L+1}\alpha_1}
1679: \ee
1680: and $\langle n |{\cal H}_1|0\rangle = 0$, $\forall n \ge 2$. By
1681: direct calculation, we find:
1682: \be
1683: E_0 = -jS_L(S_L+1); E_1 = j - jS_L(S_L+1)
1684: \ee
1685: and therefore:
1686: \be
1687: \Delta E_{\rm GS}^{(2)} = \frac{|\langle 1 | {\cal H}_1|
1688: 0\rangle|^2}{E_0-E_1} = -{4h^2\over (2S_L+1)\alpha_1^2 j}
1689: = -{4h^2\over 3j}S_L(S_L+1)
1690: \ee
1691: Since $j=2J_{eff}/L$, $S_L=LS/2$, we find:
1692: \be
1693: \Delta E_{\rm GS}^{(2)} = -{h^2\over 3J_{eff}}L^2S\left({L\over 2}S+1\right).
1694: \ee
1695: As a result, the contribution of the $k=0$ modes to the staggered susceptibility
1696: is:
1697: \be
1698: \chi_{k=0}(\pi) = - {1\over L} \left.{d^2 E_{GS}\over d h^2}\right|_{h=0}
1699: = {2 \over 3}{LS\over J_{eff}}\left({L\over 2}S+1\right)\sim L^2
1700: \ee
1701:
1702: \subsubsection{The contribution of finite $k$ modes}
1703:
1704: Within the spin-wave approximation, the contribution of the $k\ne 0$
1705: modes to the ground-state energy of ${\cal H} - h \hat{O}$ is the
1706: zero-mode energy:
1707: \be
1708: E_{GS} \sim
1709: \sum_{k\ne 0 }^{}\omega_k=\sum_{k}^{}\sqrt{(\gamma-f(k)+h)^2-g^2(k)}
1710: \ee
1711: Here, the spin-wave dispersion is changed by the addition of the
1712: perturbation $-h \sum_{i}^{}(-1)^i S^z_i$. The second derivative of
1713: $E_{GS}$ with respect to $h$ can now be calculated trivially, and in
1714: the limit $h\rightarrow 0$ we find:
1715: \be
1716: \chi_{k \ne 0}(\pi)\sim {1\over L}\sum_{k}{g^2(k)\over \omega_k^{3}}
1717: \ee
1718: where $g(k)$ was defined before Eq. (\ref{hsw}).
1719: As $k \rightarrow 0$, $g(k)\rightarrow \gamma=const$, $\omega_k
1720: \sim k^{\alpha-1\over 2}$, and therefore
1721: \be
1722: \chi_{k \ne 0}(\pi)\sim{1\over L}^{1-3{\alpha-1\over2}}\sim L^{3\alpha-5\over 2}
1723: \ee
1724: If $\alpha <3$, this is a smaller power than the $L^2$ contribution obtained from the
1725: $k=0$ mode. It follows that within the N\'eel ordered state, the
1726: staggered susceptibility scales like $L^2$ (at least within the
1727: spin-wave approximation).
1728:
1729:
1730: \subsection{Transverse correlation function}
1731:
1732: We now choose
1733: \be
1734: \hat{O} = \sum_{i}^{}S^+_{i+n} S^-_{i}
1735: \ee
1736: so that its expectation values $\langle \hat{O}\rangle$ is the
1737: transverse contribution to C(n).
1738: For simplicity, we assume $n$ to be even (calculations can be done similarly
1739: for odd $n$). Since this is not a Hermitian operator, let:
1740: \be
1741: \hat{O}_A = \sum_{i}^{}\left(S^+_{i+n} S^-_{i}+S^+_{i-n} S^-_{i}\right)
1742: \ee
1743: \be
1744: \hat{O}_B = i\sum_{i}^{}\left(S^+_{i+n} S^-_{i}-S^+_{i-n} S^-_{i}\right)
1745: \ee
1746: Both these operators are hermitian. According to Eq. (\ref{M1}) and
1747: using invariance to translations, we
1748: then have
1749: \be
1750: L \langle S^+_{n} S^-_{0}\rangle = \langle \sum_{i}^{}S^+_{i+n} S^-_{i}\rangle
1751: = \langle \hat{O}_A\rangle - i \langle \hat{O}_B \rangle \rightarrow
1752: \langle S^+_{n} S^-_{0}\rangle = -{1 \over L}
1753: \left[{dE_{GS}^{(A)}\over dh} -i {dE_{GS}^{(B)}\over dh} \right]
1754: \ee
1755: where $E_{GS}^{(A/B)}$ are the ground-states energies in the presence
1756: of perturbations $-h \hat{O}_{A/B}$.
1757:
1758: \subsubsection{The contribution of finite $k$ modes}
1759:
1760: After a Fourier transform, we use the Holstein-Primakoff
1761: representation
1762: for all $k \ne 0$ modes. Keeping only quadratic terms, we find:
1763: \be
1764: \hat{O}_A = 4S \sum_{k}^{}\cos(nka) \left(b_k^{\dagger} b_k + a_k^{\dagger} a_k \right)
1765: \ee
1766: \be
1767: \hat{O}_B = 4S \sum_{k}^{}\sin(nka) \left(b_k^{\dagger} b_k - a_k^{\dagger} a_k \right)
1768: \ee
1769:
1770: After adding this to the unperturbed Hamiltonian (in the SW approximation)
1771: and diagonalizing, we find the ground-state energies to be:
1772: \be
1773: E_{GS}^{(A)} = JS \sum_{k \ne 0 }^{}\omega_{k, A} +4sh \sum_{k\ne 0}^{} \cos(nka)
1774: \ee
1775: \be
1776: E_{GS}^{(B)} = JS \sum_{k \ne 0 }^{}\omega_{k, B} - 4sh \sum_{k\ne 0}^{} \sin(nka)
1777: \ee
1778: where
1779: \be
1780: \omega_{k, A} = \sqrt{[\gamma - f(k) -{4h\over J} \cos(nka)]^2-[g(k)]^2}
1781: \ee
1782: \be
1783: \omega_{k, B} = \sqrt{[\gamma - f(k) +{4h\over J} \sin(nka)]^2-[g(k)]^2}
1784: \ee
1785:
1786: After taking the first derivatives and setting $h=0$, we find the
1787: $k\ne 0$ modes'
1788: contribution to the transverse correlation to be:
1789: \be
1790: \langle S^+_n S^-_0\rangle = {2S\over L} \sum_{k \ne 0}^{}\frac{(\gamma
1791: - f(k)) e^{ikna}}{\omega_k} -{2S\over L} \sum_{k \ne 0}^{} e^{ikna}
1792: \ee
1793: In the limit $k \rightarrow 0$, $f(k) \rightarrow 0$,
1794: $\omega_k\rightarrow k^{\alpha-1\over 2}$ and therefore:
1795: \be
1796: \langle S^+_n S^-_0\rangle= a_1 L^{\alpha-3\over 2}( 1 + ...) +2S/L
1797: \ee
1798: The second term is the second sum ( $\sum_{k \ne 0}^{} e^{ikna} =
1799: \delta_{n,0}L/2 - 1 =-1$, since $n >0$).
1800: For $\alpha > 2$, the $L^{\alpha-3\over 2}$ term is dominant.
1801:
1802: \subsubsection{The $k=0$ mode}
1803:
1804: Keeping the full $k=0$ contributions, we find
1805: \be
1806: {\cal H}_{k=0}(h) = j \vec{S}_1\cdot \vec{S}_2 - {2h \over L} \left(S_1^+S_1^- +
1807: S_2^+ S_2^-\right)
1808: \ee
1809: The notation has been introduced in the previous section. The
1810: ground-state $|0\rangle$ of ${\cal H}_{k=0}$ is known (see section on
1811: staggered susceptibility), so the first order contribution to $E_{GS}(h)$
1812: can be evaluated directly:
1813: \be
1814: L \langle S^+_{n} S^-_0\rangle = -\left.{d E_{GS} \over
1815: dh}\right|_{h=0} \rightarrow
1816: \langle S^+_{n} S^-_0\rangle = {2 \over L^2} \langle 0 | S_1^+S_1^- +
1817: S_2^+ S_2^-|0\rangle
1818: \ee
1819: The calculation is trivial, and we find:
1820: \be
1821: \langle S^+_{n} S^-_0\rangle = {2 S^2\over 3}\left[1 + {2\over LS}
1822: \dots \right] \sim {1 \over L}
1823: \ee
1824: It follows that for this correlation function, the finite $k$ modes
1825: give the dominant $L$ dependence, which is $L^{\alpha-3\over 2}$.
1826:
1827: This calculation can be repeated for the parallel contribution to the
1828: correlation, $\langle S^z_{n} S^z_0\rangle$. The $L$ dependence
1829: remains the same, so we conclude that in the N\'eel state and within
1830: SW approximation, $C(L) \sim L^{\alpha-3\over 2}$.
1831:
1832:
1833:
1834:
1835:
1836: \section{Large-$N$ calculation}
1837: \label{app:LN}
1838: Considering the order parameter $\vec \phi$ is a $N$-component unit vector
1839: field
1840: \be |\vec \phi (\tau ,x)|^2=1,\label{constraint}\ee
1841: the action can be written like
1842: \be
1843: S={N\over 2g}\int d\tau dx [(\partial \vec \phi /\partial \tau)^2
1844: +(\partial \vec \phi /\partial x)^2]
1845: -\lambda N
1846: \int d\tau dx dy \vec \phi (\tau ,x)\cdot \vec \phi (\tau ,y)/|x-y|^{\alpha}.
1847: \ee
1848: We have set $v=1$. $g\propto 1/s$ is a coupling constant, not related to what we
1849: called $g$ in other sections. $g$ and $\lambda$ are scaled
1850: by $N$ in order to have a smooth large-$N$ limit. Inside a path integral, we may integrate over all
1851: fields $\vec \phi (\tau ,x)$, without the constraint of Eq. (\ref{constraint})
1852: provided that we introduce a Lagrange multiplier field, $\sigma (\tau ,x)$:
1853: \be S \to S+{iN\over 2g}\int d\tau dx \sigma (\vec \phi^2-1).\ee
1854: The action is now quadratic in unconstrained fields, so that we may, in
1855: principle, do the Gaussian integration over $\vec \phi$. For this purpose
1856: it is convenient to write the long range term in $\omega$-$k$ space
1857: using:
1858:
1859: \be
1860: \int_a^\infty dx{e^{ikx}\over |x|^\alpha}\approx {2\over (\alpha -1)}
1861: \left(a^{-(\alpha -1)}-|k|^{\alpha -1}\Gamma (2-\alpha )\sin [(2-\alpha )\pi /2]\right).
1862: \label{int}
1863: \ee
1864:
1865: Here $a$ is a short distance cut-off and this equation is valid for
1866: $|k|a\ll 1$.
1867: $\Gamma$ is Euler's Gamma function. Note that the prefactor blows up, $\propto 1/(\alpha -1)$
1868: as $\alpha \to 1$.
1869: Note also that for $1<\alpha <2$, both $\Gamma (2-\alpha )$ and
1870: $\sin [\pi (2-\alpha )/2]$ are $>0$, so that the second term in Eq. (\ref{int}) is $>0$.
1871: As $\alpha \to 2$, $\Gamma (2-\alpha )\to 1/(2-\alpha)$ so that $\Gamma (2-\alpha )
1872: \sin [\pi (2-\alpha )/2]\to \pi /2$. For $2<\alpha <3$, both $\Gamma (2-\alpha )$
1873: and $\sin [\pi (2-\alpha )/2]<0$ so that their product is positive, blowing up as $\alpha \to 3$.
1874:
1875: The first term in Eq. (\ref{int}) can be eliminated by shifting $\sigma$ by a (imaginary) constant.
1876: Thus we may write:
1877: \be
1878: S={NV\over 2g}\int {d\omega dk\over (2\pi )^2} \vec \phi (\omega ,k)\cdot \vec \phi (-\omega ,-k)
1879: [\omega^2+k^2+C\lambda |k|^{\alpha -1}] +{iN\over 2g}\int d\tau dx \sigma (\vec \phi^2-1).
1880: \ee
1881: Here
1882: \be C\equiv {4\over \alpha -1}\Gamma (2-\alpha )\sin [\pi (2-\alpha )/2]\ee
1883: and $V$ is the space-time volume.
1884: Now we do the functional integral over $\vec \phi$. This gives an effective action for the
1885: field $\sigma$, which has an overall factor of $N$ in front of it, and no further dependence on $N$.
1886: At large $N$ the functional integral over $\sigma$ is dominated by a saddle point corresponding
1887: to a constant and purely imaginary value of $\sigma $. Assuming $\sigma$ is constant,
1888: this effective action is:
1889: \be S_{eff}/V = {N\over 2}\left\{ \int {d\omega dk\over (2\pi )^2}\ln [\omega^2+k^2+C|k|^{\alpha -1}+i\sigma]
1890: -{i\sigma \over g}\right\}\ee
1891: The saddle point is found by looking for a stationary point of $S_{eff}$. Setting
1892: $i\sigma =m^2$ at the saddle point gives the self-consistent equation which
1893: determines $m^2$:
1894: \bea
1895: {1\over g}&=&\int {d\omega dk\over (2\pi )^2}{1\over \omega^2+k^2+C\lambda |k|^{\alpha -1}+m^2}\nonumber\\
1896: &=&\int^\Lambda_{-\Lambda} {dk\over 4\pi }{1\over \sqrt{k^2+C\lambda
1897: |k|^{\alpha -1}+m^2}}
1898: \eea
1899: Here we have introduced an ultra-violet cut-off, $\Lambda$ of O($a^{-1}$).
1900: For any $\alpha$, $\lambda$ and $g$ for which this equation has a solution with $m^2>0$,
1901: the system is in the disordered phase, with a finite gap, $m$. Note that, for $\alpha >3$,
1902: the integral diverges as $k\to 0$ when $m=0$. For small finite $m$ it behaves as
1903: $\ln (\Lambda /m)$. Since this diverges as $m\to 0$, there will always be a solution for $m$,
1904: no matter how small is $g$. On the other hand, for $1<\alpha <3$, there will be a
1905: solution for $\lambda <\lambda_c$ only. At the critical value of $\lambda$, $m=0$, so
1906: $\lambda_c$ is determined by:
1907: \be {1\over g}=\int^\Lambda_{-\Lambda} {dk\over 4\pi }{1\over \sqrt{k^2+C\lambda_c |k|^{\alpha -1}}}\ee
1908: As $\alpha \to 1$, $C\to 4/(\alpha -1)$ and this becomes:
1909: \be {1\over g}=\int^\Lambda_{-\Lambda} {dk\over 4\pi }{1\over \sqrt{k^2+4\lambda_c /(\alpha -1)}}\ee
1910: This we see that $\lambda_c\propto (\alpha -1)$, as $\alpha \to 1$. $\alpha =1$ is the critical
1911: value of $\alpha$, for which $\lambda_c\to 0$. The behavior of $\lambda_c(\alpha )$
1912: is qualitatively similar to what is obtained from SW theory, including the
1913: behavior near $\alpha =1$. Right at the critical point, the AF spin-correlation function
1914: is determined by the effective action with $m=0$, and $\vec \phi$
1915: treated as a
1916: non-interacting, free field,
1917: in the large $N$ approximation. This implies a dispersion relation:
1918: \be \omega =\sqrt{k^2+C\lambda |k|^{\alpha -1}},\ee
1919: and hence a dynamical exponent:
1920: \be z=(\alpha -1)/2.\ee
1921: This dispersion relation is the same as in SW theory, but there is no long range
1922: order at the critical point and hence no ambiguity in the value of $z$. The spin correlation function
1923: is given by:
1924: \be <\phi^a(\tau ,x)\phi^b(0,0)>=\delta^{ab}{g\over N}\int {d\omega dk\over (2\pi )^2}
1925: {e^{i(\omega \tau +kx)}\over \omega^2+k^2+C\lambda |k|^{\alpha -1}}.\ee
1926: In particular, the equal-time correlation function is:
1927: \be <\phi^a(0,x)\phi^b(0,0)>=\delta^{ab}{g\over N}\int {dk\over (4\pi )}
1928: {e^{ikx}\over \sqrt{k^2+C\lambda |k|^{\alpha -1}}}.\ee
1929: At large $x$, we may approximate this by dropping the $k^2$ term. It then follows from
1930: a rescaling that this decays as:
1931: \be <\phi^a(0,x)\phi^b(0,0)>\propto {\delta^{ab}\over |x|^{(3-\alpha )/2}}.\ee
1932: The standard definition of the critical exponent $\eta$ then implies:
1933: \be z-1+\eta = (3-\alpha )/2,\ee
1934: giving:
1935: \be \eta =3-\alpha .\ee
1936: Note that this is simply the behavior of the transverse correlation function in the
1937: ordered phase, according to SW theory.
1938:
1939: To calculate $\nu$, we need to calculate how $m$ vanishes with $\lambda_c-\lambda$
1940: as $\lambda \to \lambda_c$, from below. It turns out that there are two
1941: different behaviors, depending on whether $1<\alpha \leq 5/3$ or $5/3\leq \alpha <3$. A small change in $\lambda$ leads to a change in $1/g$
1942: which is linear in $\delta \lambda$.
1943: Consider the effect of a small non-zero $m$ on the gap equation. For $1<\alpha \leq 5/3$,
1944: we may Taylor expand the gap equation:
1945: \be
1946: {1\over g}\approx
1947: \int^\Lambda_{-\Lambda} {dk\over 4\pi }\left[{1\over \sqrt{k^2+C\lambda |k|^{\alpha -1}}}-(m^2/2){1\over [k^2+C\lambda |k|^{\alpha -1}]^{3/2}}\right]
1948: \ee
1949: (Note that this integral is finite at $k\to 0$ for $(3(\alpha -1)/2<1$
1950: only, implying $\alpha <5/3$.)
1951: Setting $\lambda =\lambda_c$, we see that:
1952: \be 1/g-1/g_c \propto -m^2.\ee
1953: Thus $m^2\propto -\delta \lambda$ in this case. Now consider
1954: the equal time correlation function for a small $m^2$.
1955: \be <\phi^a(\tau ,x)\phi^b(0,0)>=\delta^{ab}{g\over N}\int { dk\over (4\pi )}
1956: {e^{ikx}\over \sqrt{k^2+C\lambda |k|^{\alpha -1}+m^2}}.\ee
1957: At large distances and small $m^2$ we should be able to drop the $k^2$ term
1958: and ignore the ultra-violet cut-off. A rescaling of the $k$-integration
1959: variable then implies a correlation length:
1960: \be \xi \propto m^{-2/(\alpha -1)}\propto (\delta \lambda )^{-1/(\alpha -1)}
1961: \ee
1962: and hence an exponent:
1963: \be \nu = 1/(\alpha -1),\ \ (1<\alpha \leq 5/3 ).\ee
1964:
1965: Now consider the case $5/3\leq \alpha <3$. Keeping a small non-zero
1966: $m^2$, we find:
1967: \be {d\over dm^2}{\left( 1\over g\right) }={-1\over 2}
1968: \int {dk\over 4\pi}{1\over [\lambda C|k|^{\alpha -1}+m^2]^{3/2}}.\ee
1969: Since the integral is dominated by $|k|$ of $O(m^{2/(\alpha -1)}$,
1970: we have taken the cut-off to infinity and dropped the $k^2$ term.
1971: By scaling, we see that:
1972: \be {d\over dm^2}{\left( 1\over g\right) }\propto m^{(5-3\alpha )/(\alpha -1)}
1973: .\ee
1974: Integrating with respect to $m^2$ gives:
1975: \be { 1\over g}\approx {1\over g_c}-Am^{(3-\alpha )/(\alpha -1)},\ee
1976: for a constant, $A$. Thus we see that:
1977: \be m\propto (-\delta \lambda )^{(\alpha -1)/(3-\alpha )}.\ee
1978: This gives:
1979: \be \xi \propto m^{-2/(\alpha -1)}\propto (-\delta \lambda )^{-2/(3-\alpha )}.
1980: \ee
1981: Thus:
1982: \be \nu = {2\over 3-\alpha},\ \ (5/3 \leq \alpha <3).\ee
1983:
1984: \begin{thebibliography}{999}
1985:
1986: \bibitem{Ian94} I. Affleck, M. P. Gelfand and R. R. P. Singh,
1987: J. Phys. A {\bf 27}, 7313 (1994).
1988:
1989: \bibitem{Haldane83-87} F. D. M. Haldane, Phys. Rev. Let. {\bf 50},
1990: 1153 (1983); I. Affleck and F. D. M. Haldane, Phys. Rev. B {\bf 36},
1991: 5291 (1987).
1992:
1993: \bibitem{Giamarchi89} I. Affleck, D. Gepner, H. J. Schulz and T. Ziman,
1994: J. Phys. A {\bf 22}, 511 (1989); T. Giamarchi and H. J. Schulz, Phys. Rev. B {\bf
1995: 39}, 4620 (1989).
1996:
1997: \bibitem{HS88} F. D. M. Haldane, Phys. Rev. Lett. {\bf 60}, 635 (1988);
1998: S. Shastry, {\it ibid} {\bf 60}, 639 (1988).
1999:
2000: \bibitem{MW66} N. D. Mermin and H. Wagner, Phys. Rev. Lett. {\bf 17},
2001: 1133 (1966).
2002:
2003: \bibitem{68-69} D. Ruelle, Commun. Math. Phys. {\bf 9}, 267 (1968);
2004: F. J. Dyson, {\it ibid} {\bf 12}, 212 (1969); D. J. Thouless,
2005: Phys. Rev. {\bf 187}, 732 (1969).
2006: \bibitem{Fisher72} M. E. Fisher, S-K. Ma and B. G. Nickel,
2007: Phys. Rev. Lett. {\bf 29}, 917 (1972).
2008:
2009: \bibitem{Kosterlitz76} J. M. Kosterlitz, Phys. Rev. Lett. {\bf 37},
2010: 1577 (1976).
2011:
2012: \bibitem{Brezin76} E. Brezin, J. Zinn-Justin and J-C. Le Guillou,
2013: J. Phys. A {\bf 9}, L119 (1976).
2014:
2015: \bibitem{Cardy81} J. L. Cardy, J. Phys. A {\bf 14}, 1407 (1981).
2016:
2017: \bibitem{Frolich82} J. Fr\"ohlich and T. Spencer,
2018: Commun. Math. Phys. {\bf 84}, 87 (1982).
2019:
2020: \bibitem{Luijten01} E. Luijten and H. Messingfeld,
2021: Phys. Rev. Lett. {\bf 86}, 5305 (2001).
2022:
2023: \bibitem{Bruno01} P. Bruno, Phys. Rev. Lett. {\bf 87}, 137203 (2001).
2024:
2025: \bibitem{Yusuf04}E. Yusuf, A. Joshi and K. Yang, Phys. Rev. B {\bf 69},
2026: 144412 (2004).
2027:
2028: \bibitem{Parreira97} J. R. Parreira, O. Bolina and J. F. Perez,
2029: J. Phys. A {\bf 30}, 1095 (1997).
2030:
2031: \bibitem{note1} The proof for non-existence of LRO at $T=0$ for
2032: $\lambda =1$, if $\alpha
2033: >3$ \cite{Parreira97}, can be trivially extended and
2034: shown to hold for all $\lambda \ne 1$.
2035:
2036: \bibitem{Sandvik02} O.~F.~Syljuasen and A.~W.~Sandvik, Phys. Rev. E
2037: {\bf 66}, 046701 (2002).
2038:
2039: \bibitem{Sandvik03} A.~W.~Sandvik, Phys. Rev. E {\bf 68}, 056701 (2003).
2040:
2041: \bibitem{CardyBook} J. Cardy, {\it Scaling and Renormalization in
2042: Statistical Physics} (Cambridge University Press, Cambridge, 1996),
2043: Chap. 4.
2044:
2045: \bibitem{Ian86} H. W. J. Bl\"ote, J. L. Cardy,
2046: and M. P. Nightingale, Phys. Rev. Lett. {\bf 56}, 742 (1986); I. Affleck,
2047: Phys. Rev. Lett. {\bf 56}, 746 (1986).
2048:
2049: \bibitem{Holstein} T. Holstein and H. Primakoff, Phys. Rev. {\bf 58},
2050: 1098 (1940).
2051:
2052: \bibitem{note2} These equations are the analogs of equations (6) of
2053: Ref. \cite{Yusuf04}. We use $\gamma$ instead of $\alpha$ since
2054: in our work $\alpha$ defines the spatial variation of the exchange
2055: (the parameter called $\beta$ in Ref. \cite{Yusuf04}). The
2056: definition of $f(k)$ is corrected by a factor of $2$.
2057:
2058:
2059: \bibitem{Reger88} J. D. Reger and A. P. Young, Phys. Rev. B {\bf{37}}, 5978 (1988).
2060:
2061: \bibitem{notexi}
2062: A precise determination of
2063: the exponent $\nu$ is actually impossible to achieve
2064: because of large error bars for $\xi_{\lambda}$ due to some natural
2065: uncertainties in the data collapse procedure.
2066: \bibitem{noteta}
2067: It is important to note that, up to small corrections, this
2068: estimate Eq.~(\ref{MFeta}) was
2069: also found 3 decades ago by several authors for the related $n$-vector
2070: model with long range interaction~\cite{Fisher72,Brezin76,Ma73}.
2071:
2072: \bibitem{Ma73} S-K. Ma, Phys. Rev. A {\bf 7}, 2172 (1973).
2073:
2074: \bibitem{Sak} J. Sak, Phys. Rev. {\bf B15}, 4344 (1977).
2075:
2076: \bibitem{Bhatt} J. Bhattacharjee, J.L. Cardy and D.J. Scalapino,
2077: Phys. Rev. {\bf B25}, 1681 (1982).
2078:
2079: \end{thebibliography}
2080: \end{document}
2081:
2082: