1: \tolerance = 10000
2: \documentclass[prb,eqsecnum,twocolumn,floatfix]{revtex4}
3: \usepackage[dvips]{graphicx}
4: %\usepackage{showkeys}
5: \usepackage{latexsym}
6: \usepackage{amsmath}
7: \usepackage{amssymb}
8: %\preprint{cond-mat/04mmxxx}
9: \begin{document}
10: \title{Thermal melting of density waves on the square lattice}
11:
12:
13: \author{Adrian Del Maestro and Subir Sachdev}
14: \affiliation{Department of Physics, Yale University, P.O. Box
15: 208120, New Haven, CT 06520-8120, USA}
16:
17:
18:
19: \date{\today}
20:
21: \begin{abstract}
22: We present the theory of the effect of thermal fluctuations on
23: commensurate $p \times p$ density wave ordering on the square
24: lattice ($p \geq 3$, integer). For the case in which this order is
25: lost by a second order transition, we argue that the adjacent
26: state is generically an incommensurate striped state, with
27: commensurate $p$-periodic long range order along one direction,
28: and incommensurate quasi-long-range order along the orthogonal
29: direction. We also present the routes by which the fully
30: disordered high temperature state can be reached. For $p=4$, and
31: at special commensurate densities, the $4 \times 4$ commensurate
32: state can melt directly into the disordered state via a self-dual
33: critical point with non-universal exponents.
34: \end{abstract}
35:
36: \maketitle
37:
38: \section{Introduction}
39: \label{sec:intro}
40:
41: A variety of remarkable recent low temperature scanning tunneling
42: microscopy (STM) observations have revealed periodic modulations
43: in the local density of states of the cuprate superconductors
44: \cite{howald,mcelroy,hanaguri,fang}. Notably, the observations of
45: Hanaguri {\em et al} \cite{hanaguri} Ca$_{2-x}$Na$_x$CuO$_2$Cl$_2$
46: clearly show that the modulations have a commensurate period of 4
47: lattice spacings along both directions of the underlying square
48: lattice ($4 \times 4$ ordering). Closely related, but not
49: identical, modulations have been observed in higher temperature
50: STM \cite{ali}, and in neutron scattering
51: \cite{jtran,hinkov,buyers,hayden} observations. The differences
52: between the various experiments relate ({\em i\/}) to the period
53: of the ordering, which is incommensurate in Ref.~\onlinecite{ali},
54: and ({\em ii\/}) to whether the ordering extends along one or both
55: of the $x$ and $y$ axes---neutron scattering experiments are more
56: easily explained by anisotropic ordering along one of the axes
57: directions \cite{MVTU,GSU,vs}. An important open question is
58: whether these differences reflect a fundamental distinction in the
59: underlying electronic structure of the cuprates, or they can be
60: explained by the differing experimental parameters of temperature
61: and carrier concentration.
62:
63: This paper will begin with the assumption that at very low $T$,
64: for a small range of carrier densities, and in a sufficiently
65: clean sample, the system has perfect $4 \times 4$ density wave
66: order (and, more generally, $p \times p$ order with $p \geq 3$,
67: integer). We will then present a general phenomenological theory
68: of the effect of thermal fluctuations on such an ordered state,
69: describing the phases that appear at higher $T$ and for a wide
70: range of carrier concentration. Because we focus exclusively on
71: thermal fluctuations, our results are quite general and
72: independent of the precise microscopic nature of the density wave
73: ordering; in particular, the ordering could be a site charge
74: density wave, or a modulation in exchange or pairing energies due
75: to valence bond solid order. Our results depend only on the fact
76: that spin-singlet observables acquire a periodic modulation at low
77: $T$, and follow completely from the symmetry properties of such
78: states. The thermal fluctuations of such an ordered state can, and
79: will, be described by a purely classical theory of the density
80: wave order parameters.
81:
82: There is a actually large early literature on the melting of a
83: variety to two-dimensional solids on different substrates, and on
84: the commensurate-incommensurate transition
85: \cite{nh,apy,pt,ostlund,sue,schulz,haldane,huse}. However, most of
86: this work considered the case of the triangular solid on
87: substrates with six-fold symmetry, or the
88: commensurate-incommensurate transition on anisotropic solids. The
89: particular case of interest here, a square lattice substrate and
90: isotropic $p \times p$ ordering, appears not to have been
91: considered previously. We will therefore present an extension of
92: this earlier theory to the situation appropriate for the cuprate
93: superconductors. As we will show below, our results have a number
94: of novel features not found in the cases studied earlier.
95:
96: We will begin in Section~\ref{sec:mft} by defining the order
97: parameters of the commensurate $p \times p$ solid on the square
98: lattice. Symmetry considerations then allow us to obtain a
99: phenomenological free energy density controlling thermal
100: fluctuations of the order parameter, and a corresponding mean
101: field phase diagram shown in Fig.~\ref{figmft}.
102: \begin{figure}
103: \centering
104: \includegraphics[width=3in]{fig_mfpd.eps}
105: \caption{Mean field phase diagram of the free energy
106: $\mathcal{F}_{\Phi}$ in Eq.~(\ref{FP}). Although all parameters in
107: $\mathcal{F}_{\Phi}$ are functions of both temperature and carrier
108: concentration, the vertical axis above, $s$, is primarily a
109: function of temperature. The horizontal axis is tuned by
110: $\overline{\eta}$, which is mainly a measure of carrier
111: concentration away from the commensurate value appropriate to the
112: $p\times p$ ordered state. There are three stable mean field
113: phases separated by a first order transition (thick line), a
114: second order transition (solid line), and a second-order critical
115: point at $s=\overline{\eta}=0$. The phases A, B, C, are described
116: in the text. For phases A and B we show a schematic of the density
117: modulation $\delta \rho ({\bf r})$ for the special case $p=4$,
118: with the points representing the underlying lattice. Upon
119: including fluctuations, the transition between phases A and B can
120: become second-order; in this situation the phase adjacent to phase
121: A is a new phase B$_S$, the incommensurate striped phase. The
122: phase B$_S$ preempts a portion (or all) of phase B. See
123: Fig~\ref{figbs} for the fluctuation-corrected phase diagram of
124: phases A and B.} \label{figmft}
125: \end{figure}
126: At this stage, this phase diagram contains 3 phases whose
127: characteristics we summarize below:
128: \newline
129: (A) {\bf Commensurate solid:} This is the ground state with
130: long-range density wave order with period $p$ along both the $x$ and
131: $y$ axes.
132: \newline
133: (B) {\bf Incommensurate, floating solid:} In mean field theory, this
134: phase has density waves with the same incommensurate period along
135: both the $x$ and $y$ directions. This order is only quasi-long-range
136: {\em i.e.\/} density wave correlations decay with a power-law at
137: long distances.
138: \newline
139: (C) {\bf Liquid:} This is the high temperature phase in which all
140: density wave correlations decay exponentially with distance.
141:
142: The remainder of the paper will consider fluctuation corrections to
143: this mean field phase diagram.
144:
145: In mean field theory, the transition between phases A and B is
146: first order, with a jump in the period of the density wave from a
147: commensurate to an incommensurate value. We know from previous
148: work on anisotropic phases \cite{pt,sue,schulz,haldane} that such
149: commensurate-incommensurate transitions are driven by the
150: proliferation of domain walls, and we consider the domain wall
151: theory for such a transition in Section~\ref{sec:domains}. In our
152: case there are two sets of domain walls, running predominantly in
153: the $x$ and $y$ directions, and a key parameter will be the
154: intersection energy, $f_I$, of these domain walls. If $f_I < 0$,
155: then the mean field prediction of a first order transition to an
156: isotropic floating solid is maintained. However, for the case of a
157: positive intersection energy ($f_I > 0$), we will demonstrate that
158: there is a second order transition to an anisotropic,
159: incommensurate ``striped'' state, which we denote B$_S$. The B$_S$
160: state has commensurate long range order with a period of $p$
161: lattice along one axis, and incommensurate quasi-long-range order
162: along the orthogonal axis. Because B$_S$ has long-range order only
163: along one direction, it may also be labelled a `smectic'
164: \cite{chaikinlub,kfe}; however the presence of the underlying
165: lattice makes its fluctuations quite different from a conventional
166: smectic liquid crystal. The striped solid B$_S$ will replace a
167: portion of phase B, so that B$_S$ covers at least the entire
168: second-order phase boundary to phase A (see Fig~\ref{figbs}). The
169: remainder of phase B could remain an isotropic floating solid, or
170: it could be anisotropic in its entirety.
171:
172: Section~\ref{sec:floating} will consider the transitions from the
173: incommensurate states in phase B to the disordered liquid C. Two
174: distinct scenarios are possible, depending upon the whether the
175: transition takes place from an isotropic floating solid B, or from
176: the anisotropic solid (B$_S$) which has incommensurate order along
177: only one axis. In either case, the transition is driven by the
178: unbinding of dislocations \cite{nh,apy}, and the complete theory
179: for such transitions will be presented.
180:
181: Finally, Section~\ref{sec:dual} will present the theory of the
182: direct transition from the commensurate $p \times p$ solid A to
183: the liquid state C. From Fig.~\ref{figmft} it appears that such a
184: transition is only possible at special commensurate values of the
185: density. For the physically relevant case of $p=4$ we will find
186: that a direct second order transition is possible. It is described
187: by a self-dual theory with continuously varying exponents, and is
188: a generalization of the theory found in Ref.~\onlinecite{jkkn} for
189: the XY model.
190:
191: \section{Order parameters and mean field theory}
192: \label{sec:mft}
193:
194: As noted in Section~\ref{sec:intro}, all phases and transitions
195: examined here are associated with the order of a generic
196: `density', which could be any observable invariant under spin
197: rotations and time reversal. We represent this density by $\delta
198: \rho ({\bf r})$. At sufficiently low temperatures, in weak
199: disorder, and for some range of carrier concentration, we assume
200: that this density prefers to order with a period of $p$ lattice
201: spacings ($p=4$ is the case of interest for the cuprates). We can
202: therefore write
203: \begin{equation}
204: \delta \rho ({\bf r}) = \mbox{Re} \left[ \Phi_x e^{i {\bf K}_x
205: \cdot {\bf r}}\right] + \mbox{Re} \left[ \Phi_y e^{i {\bf K}_y
206: \cdot {\bf r}} \right], \label{defPhi}
207: \end{equation}
208: where ${\bf K}_x = (2 \pi /a) (1/p, 0)$, ${\bf K}_y = (2 \pi /a)
209: (0,1/p)$, and $\Phi_{x,y}$ are complex order parameters which vary
210: slowly the on the scale of a lattice spacing.
211:
212: We now want to write down the most general free energy for
213: $\Phi_{x,y}$ consistent with the symmetries of the underlying
214: lattice. Among these are $T_{x,y}$ which translate by a lattice
215: spacing in the $x,y$ directions, and $I_{x,y}$ which reflect the
216: $x,y$ axes. These operations lead to
217: \begin{eqnarray}
218: T_x &:& \Phi_x \rightarrow \Phi_x e^{2 i \pi/p}~~;~~\Phi_y
219: \rightarrow \Phi_y \nonumber \\
220: T_y &:& \Phi_x \rightarrow \Phi_x ~~;~~\Phi_y
221: \rightarrow \Phi_y e^{2 i \pi/p} \nonumber \\
222: I_x &:& \Phi_x \rightarrow \Phi_x^{\ast}~~;~~\Phi_y
223: \rightarrow \Phi_y \nonumber \\
224: I_y &:& \Phi_x \rightarrow \Phi_x~~;~~\Phi_y \rightarrow
225: \Phi_y^{\ast}. \label{t1}
226: \end{eqnarray}
227: We will also assume the symmetry of rotations by 90 degrees, $R$,
228: under which
229: \begin{equation}
230: R : \Phi_x \rightarrow \Phi_y~~;~~\Phi_y \rightarrow
231: \Phi_x^{\ast}. \label{t2}
232: \end{equation}
233: This symmetry is absent in some of the cuprate compounds (most
234: notably, in YBCO), and it is not difficult to extend our
235: considerations to include this case.
236:
237: We can now write down the most general free energy density,
238: expanded in powers of $\Phi_{x,y}$ and its gradients, consistent
239: with the symmetries in Eqs.~(\ref{t1}) and (\ref{t2}) (see also
240: Refs.~\onlinecite{zachar} and~\onlinecite{ying}). This is
241: \begin{eqnarray}
242: \mathcal{F}_{\Phi} &=& \int d^2 r \Biggl[ C_1 \left( \left|
243: \partial_x \Phi_x \right|^2 + \left| \partial_y \Phi_y \right|^2
244: \right) \nonumber \\
245: &~&~~~~~~+ C_2 \left( \left|
246: \partial_y \Phi_x \right|^2 + \left| \partial_x \Phi_y \right|^2
247: \right) \nonumber \\
248: &~&~~~~~~+ i \overline{\eta} \left( \Phi_x^{\ast} \partial_x
249: \Phi_x +
250: \Phi_y^{\ast} \partial_y \Phi_y \right) \nonumber \\
251: &~&~~~~~~+ s \left( |\Phi_x |^2 + |\Phi_y |^2 \right) +
252: \frac{u}{2} \left(|\Phi_x |^2 + |\Phi_y |^2 \right)^2 \nonumber
253: \\ &+& v |\Phi_x |^2 |\Phi_y |^2 + w \Phi_x^p + \mbox{c.c.} +
254: w \Phi_y^p + \mbox{c.c.} \Biggr]. \label{FP}
255: \end{eqnarray}
256:
257: Here $s$ is a parameter which we will tune to drive the
258: transition; it is assumed to contain the primary dependence on
259: temperature, and will also have some dependence on carrier
260: density.
261:
262: The term proportional to $\overline{\eta}$ is allowed by the
263: symmetries. It indicates that at sufficiently high temperature,
264: the density correlations are generically {\em incommensurate\/}.
265: This incommensurability is a consequence of the values of the
266: domain wall energies between the $p$ commensurate ordered states
267: \cite{ostlund,huse}: a domain wall between state 1 and state 2
268: (say) will generally have a different energy than a domain wall
269: between state 1 and state $p$. In other words, if we represent the
270: $p$ states along one direction by a $p$ state clock model, then
271: the interactions of the clock model are `chiral'. \cite{ostlund}
272:
273: The sign of the parameter $v$ implies a preference for either
274: isotropic order ($v<0$ prefers both $\langle \Phi_x \rangle$ and
275: $\langle \Phi_y \rangle$ non-zero), or anisotropic order ($v>0$
276: prefers only one of $\langle \Phi_x \rangle$ or $\Phi_y$
277: non-zero). We will assume throughout this paper that $v<0$, so
278: that $\langle \Phi_x \rangle$ and $\langle \Phi_y \rangle$ are
279: both non-zero at sufficiently low temperatures. Nevertheless, we
280: will find that thermal fluctuations can induce an anisotropic
281: striped state over an intermediate temperature range.
282:
283: The complex parameter $w$ accounts for the commensurability
284: lock-in energy. It ensures that at sufficiently low temperatures
285: (specifically, for $s$ sufficiently negative), the order has a
286: commensurate period of $p$ lattice spacings. For $v<0$ (assumed
287: throughout), the ordering will be $p \times p$. The phase of $w$
288: controls the phase of this ordering (`site-centered' or
289: `bond-centered'). None of our results will be sensitive to this
290: phase, and will apply equally to all of them.
291:
292: \subsection{Mean field theory}
293: \label{sec:mfta}
294:
295: Now we present the results of a simple mean field minimization of
296: $\mathcal{F}_\Phi$, leading to the phase diagram in
297: Fig.~\ref{figmft}. As described in the introduction
298: we find three distinct phases. The low temperature ground state, labelled by
299: A has long rage density wave order with period $p$ characterized by
300: \begin{equation}
301: \Phi_x = \Phi_y \propto e^{i \theta(p)}
302: \end{equation}
303: where $\theta(p) = (2n+1)\pi /p$ if the phase is bond-centered and $\theta(p)
304: = 2n\pi/p$ if the phase is site-centered with $n=0,1,\ldots,p-1$. The
305: inset in Fig.~\ref{figmft} A shows the locked in solid for the
306: special case of a bond-centered phase with $p=4$ where the positions of the
307: underlying atoms are indicated by the light gray circles. This commensurate phase
308: has a first order transition described by the line
309: \begin{equation}
310: s = -\left(\frac{\sqrt{2u+v-pw}}{\sqrt{2u+v}-\sqrt{2u+v-pw}}\right)
311: \frac{\overline{\eta}^2}{4C_1}
312: \end{equation}
313: to B which has incommensurate or floating density wave order
314: characterized by
315: \begin{equation}
316: \Phi_x \propto e^{i (\overline{\eta} / 2C_1)x} \quad \mbox{;}
317: \quad \Phi_y \propto e^{i (\overline{\eta} / 2C_1)y}.
318: \end{equation}
319: The magnitude of both $\Phi_x$ and $\Phi_y$ are equal and they have
320: the same incommensurate period along perpendicular directions. For
321: the special case of $p=4$ the floating state is shown as an inset in
322: Fig.~\ref{figmft} B. The floating solid can melt via a second
323: order phase transition defined by the line
324: \begin{equation}
325: s = \frac{\overline{\eta}^2}{4C_1}
326: \end{equation}
327: to a fully disordered state C with $\Phi_x=\Phi_y=0$. Finally, there is a
328: tri-critical point for $s=\overline{\eta}=0$ where the commensurate
329: solid A can melt directly to the liquid phase C.
330:
331: \section{Domain wall theory of the commensurate-incommensurate transition}
332: \label{sec:domains}
333:
334: In Section~\ref{sec:mfta} we found a mean-field first order
335: transition between the commensurate $p\times p$ solid A and the
336: incommensurate, isotropic floating solid B. This section will
337: examine fluctuations near this transition more carefully. We will
338: find that the transition can actually be second order under
339: suitable conditions, and the second-order order transition is to
340: an anisotropic, striped state with commensurate $p$ period order
341: along one direction, and incommensurate order along the orthogonal
342: direction.
343:
344: For our study of the initial melting of the commensurate ordered
345: state, we will assume that dislocations can be ignored. The effect
346: of dislocations will be considered in Section~\ref{sec:floating},
347: and we will then verify the self-consistency of this assumption.
348: Instead, the primary actors will be domain walls between the
349: commensurate states, as is also the case in previous theories
350: \cite{pt} of the commensurate-incommensurate transition in
351: anisotropic systems.
352:
353: In the absence of dislocations, we can focus on globally defined
354: single-valued angular variables $\theta_{x,y}$ with
355: \begin{equation}
356: \Phi_x \propto e^{i \theta_x}~~;~~ \Phi_y \propto e^{i \theta_y}.
357: \label{deftheta}
358: \end{equation}
359: Note that the values of $\theta_{x,y}$ span over all real numbers,
360: and not just modulo $2 \pi$. However, periodic boundary conditions
361: need only be satisfied modulo $2 \pi$. In the na\"ive continuum
362: limit, the free energy for $\theta_{x,y}$ can be expanded in
363: powers of the local ``strains'' $\nabla_{\bf r} \theta_{x,y}$:
364: \begin{eqnarray}
365: \mathcal{F}_\theta &=& \int d^2 r \Biggl[ \frac{K_1}{2} \left[
366: (\partial_x \theta_x)^2 + (\partial_y \theta_y )^2 \right]
367: \nonumber \\
368: &+& \frac{K_2}{2}\left[ (\partial_x \theta_y)^2 + (\partial_y
369: \theta_x )^2 \right] + K_3 (\partial_x \theta_y)(
370: \partial_y \theta_x)
371: \nonumber \\
372: &+& K_4 (\partial_x \theta_x) (\partial_y \theta_y) - \eta \left[
373: \partial_x \theta_x + \partial_y \theta_y \right] \nonumber \\
374: &-& h \left[ \cos ( p \theta_x ) + \cos (p \theta_y) \right] -
375: \ldots \Biggr]. \label{ftheta}
376: \end{eqnarray}
377: All terms above are, in principle, obtained from those in
378: Eq.~(\ref{FP}), but now we have retained terms up to second order
379: in spatial gradients. Now, the incommensuration is induced by the
380: total derivative terms proportional to $\eta$: these are non-zero
381: because the angular fields can accumulate an integer multiple of
382: $2 \pi$ even under periodic boundary conditions. Similarly, we do
383: not have the freedom to integrate by parts, and so combine the
384: terms proportional to $K_3$ and $K_4$. The commensurability energy
385: is now imposed by the $p$-fold field $h_p$.
386:
387: We will begin in Section~\ref{sec:solitons} by describing the
388: mean-field structure of the domain wall (or `soliton') excitations
389: of $\mathcal{F}_\theta$. Then, in Section~\ref{sec:wandering} we
390: will present the theory of the commensurate-incommensurate
391: transition driven by the proliferation of these domain walls.
392: Section~\ref{sec:spinwave} will address the nature of density
393: fluctuations within the incommensurate phases: we will estimate
394: the elastic constants of these phases using the domain wall theory
395: of Section~\ref{sec:wandering}.
396:
397:
398: \subsection{Energetics of domain walls}
399: \label{sec:solitons}
400:
401: Consider starting at low temperatures from a fully ordered
402: two-dimensional crystal. In this state $\theta_x = 2 \pi n/p$,
403: $\theta_y = 2\pi n' /p$, where $n$, $n'$ are integers, at all
404: points in space. We are now interested in the deviations from this
405: perfectly ordered state as measured by the continuum free energy
406: $\mathcal{F}_\theta$ in Eq.~(\ref{ftheta}).
407:
408: The simplest deviation is a single line domain wall in which
409: $\theta_x$ increases by $2 \pi/p$ across the domain wall. For
410: $\eta > 0$ (which we assume, without loss of generality), such a
411: domain wall will preferentially run in the $y$ direction {\em
412: i.e.\/} the domain wall has $\theta_y$ constant, and $\theta_x$ a
413: function of $x$ only. The $x$ dependence of $\theta_x$ can be
414: determined by the sine-Gordon saddle point equation
415: \begin{equation}
416: K_1 \partial_x^2 \theta_x(x) = ph\sin\left[p\theta_x (x)\right]
417: \end{equation}
418: subject to the boundary conditions
419: \begin{equation}
420: \theta_x (0) = 0 \quad \mbox{;} \quad \theta_x (L_x) =
421: \frac{2\pi}{p} \label{thetaBC}
422: \end{equation}
423: where $L_x$ is the size of the system in the $x$-direction. The
424: solution is a soliton with equation
425: \begin{equation}
426: \theta_x (x) =
427: \frac{4}{p}\tan^{-1}\left[e^{p\sqrt{h/K_1}(x-L_x/2)}\right]
428: \end{equation}
429: which allows us to determine the value of the domain wall free energy
430: per unit length (excluding the contribution of the $\eta$ term in
431: $\mathcal{F}_I$), which we represent by $\epsilon$,
432: \begin{equation}
433: \epsilon = \frac{8}{p}\sqrt{K_1 h}.
434: \end{equation}
435: Similarly, there is a corresponding domain wall in $\theta_y$,
436: with identical physical properties.
437:
438: Now we consider the interesting case with domain walls running in
439: both the $x$ and $y$ directions. These walls will intersect, and
440: we are interested in the nature of the intersection, and of the
441: intersection free energy $f_I$. We can focus in on a single domain wall
442: crossing by imposing the boundary conditions of Eq.~(\ref{thetaBC}) in both
443: the $x$ and $y$ directions. For the special case of perfectly straight
444: domain walls, $\theta_x(x,y) = \theta_x(x)$ and $\theta_y(x,y) =
445: \theta_y(y)$, the only interaction term in Eq.~(\ref{ftheta}) can be
446: integrated exactly to give
447: \begin{equation}
448: f_I = \left(\frac{2\pi}{p}\right)^2 K_4
449: \label{eqfi}
450: \end{equation}
451: indicating that the sign of the domain wall intersection energy is equal to
452: the sign of $K_4$. When the domain walls are not straight, the interaction
453: energy can be found by numerically minimizing $\mathcal{F}_\theta$ and
454: evaluating the interaction terms at the ground state field
455: configurations as seen in Fig.~\ref{figdomain}.
456: \begin{figure}
457: \centering
458: \includegraphics[width=3in]{fig_dwi.eps}
459: \caption{Eight possible configurations of the free energy density
460: for domain wall crossings depending on the numerical values of
461: $K_2$, $K_3$ and $K_4$ with $p=4$. The grey scale plots the size of the
462: local free energy density and all stiffnesses and energies are in units of
463: $K_1$. Panels a. through h. show decreasing domain wall
464: intersection energy $f_I$, and for straight walls, the numerically
465: computed value can be compared with $(\pi^2/4) K_4$.}
466: \label{figdomain}
467: \end{figure}
468: For the following discussion we refer to Fig.~\ref{figdomain} and it
469: is assumed that all stiffnesses are measured in units of $K_1$.
470: For positive values of $K_4$ and $K_2=1.0$ (panels a through
471: c) we observe straight domain walls for $K_3 < K_2$ and only
472: observe deviation when $K_3 \simeq K_2$.
473: In panel d, $K_3 \simeq K_2 < K_4$ and we find that the walls
474: wander quite significantly. Panels a through d all have $K_4 > 0$
475: and in agreement with our earlier discussion, $f_I > 0$ in all
476: four configurations. Panel e has no interaction terms ($K_3=K_4=0$)
477: and consequently $f_I = 0$. For negative values of $K_4$ we
478: find that the domain wall intersection energy changes to a
479: negative value. For $|K_3| < K_2$ the walls remain straight
480: (panels f and g) but as $K_3 \simeq K_4 \simeq -K_2$ the
481: intersection energy becomes large and negative and the system attempts
482: to extend the domain wall overlap over a finite region (panel h).
483:
484: Although we have only presented eight distinct configurations here,
485: all of the $K_j$ parameter space was examined. When the magnitude
486: of $K_3$ and $K_4$ are small with respect to $K_2$ and $K_1$ we always
487: find configurations with domain walls crossing at right angles where
488: the domain wall intersection energy can be calculated using
489: Eq.~(\ref{eqfi}).
490:
491: \subsection{Proliferation of domain walls}
492: \label{sec:wandering}
493:
494: Now we imagine increasing the parameter $\eta$ so that the total
495: free energy per unit length of a domain wall is eventually
496: negative. In such a situation we expect a proliferation of domain
497: walls, leading to the appearance of a floating solid with
498: incommensurate density correlations. This section will discuss the
499: theory of such a transition.
500:
501: Let the incommensurate state have domain walls in $\theta_x$ with
502: an average spacing $\ell_x$, and domain walls in $\theta_y$ with
503: an average spacing $\theta_y$. From the energy per unit length of
504: these domain walls, and their intersection energy, computed in
505: Section~\ref{sec:solitons}, there is a clearly a contribution to
506: the free energy per unit area, $\mathcal{F}_d$, given by (see
507: Fig.~\ref{figdomain})
508: \begin{equation}
509: \mathcal{F}_d^{(1)} (\ell_x , \ell_y) = \left( \epsilon - \frac{2
510: \pi \eta}{p} \right) \left( \frac{1}{\ell_x} + \frac{1}{\ell_y}
511: \right) + \frac{f_I}{\ell_x \ell_y}. \label{fd1}
512: \end{equation}
513:
514: However, in addition to the simple energetic contributions in
515: Eq.~(\ref{fd1}), we also have to consider the entropic
516: contribution of the wandering of the domain walls (Fig.~\ref{figdwnet}).
517: \begin{figure}
518: \centering
519: \includegraphics[width=3in,height=3in]{fig_dwnet.eps}
520: \caption{A net of wandering domain walls constructed using random walkers on
521: a lattice with both hard-core repulsion and restricted phase space. If
522: the domain walls are separated on average by a
523: distance $\ell_x$ in the $x$ direction and $\ell_y$ in the $y$
524: direction, then we observe collisions between domain walls running in the
525: same direction with separation on the order of $\ell_x^2$ or $\ell_y^2$
526: ($\ell_x = \ell_y$ here).}
527: \label{figdwnet}
528: \end{figure}
529: This can be computed using the elegant free fermion mapping of
530: Pokrovsky and Talapov \cite{pt}, which applies here (essentially unchanged)
531: separately to the domain walls in each direction. Briefly, the
532: argument runs as follows. Let $u_x (y)$ represent the $x$
533: co-ordinate of a domain wall in $\theta_x$. Any $y$ dependence in
534: $u_x$ increases the total length of the domain wall, and so leads
535: to a free energy cost
536: \begin{equation}
537: \mathcal{F}_w = \frac{1}{2} \epsilon \int dy\, \left(
538: \frac{du_x}{dy} \right)^2. \label{fwall}
539: \end{equation}
540: Now the
541: partition function at a temperature $T$ of such domain walls can
542: be mapped onto that for free fermions with a density $1/\ell_x$
543: and mass $\epsilon/T$. From the ground state energy of such a free
544: fermion system, we then obtain the contribution of the wandering
545: of the domain walls to the free energy density
546: \begin{equation}
547: \mathcal{F}_d^{(2)} (\ell_x, \ell_y) = \frac{\pi^2 T^2}{6
548: \epsilon} \left( \frac{1}{\ell_x^3} + \frac{1}{\ell_y^3} \right).
549: \end{equation}
550:
551: We are now faced with the simple problem of minimizing the free
552: energy density
553: \begin{equation}
554: \mathcal{F}_d (\ell_x, \ell_y) = \mathcal{F}_d^{(1)} (\ell_x,
555: \ell_y) + \mathcal{F}_d^{(2)} (\ell_x, \ell_y) \label{fd}
556: \end{equation}
557: as a function of $\ell_x$ and $\ell_y$ to determine the nature of
558: the commensurate-incommensurate transition. This simple
559: calculation turns out to have some interesting structure which we
560: will now describe. The results of such a minimization are
561: summarized in Fig~\ref{figbs}.
562: \begin{figure}
563: \centering
564: \includegraphics[width=3in]{fig_ABBs.eps}
565: \caption{Phase diagram of the commensurate-incommensurate
566: transition obtained by the minimization of Eq.~(\ref{fd}) over the
567: values of the mean domain wall spacing $\ell_x$ and $\ell_y$.}
568: \label{figbs}
569: \end{figure}
570: The nature of the phases appearing here depends upon the sign of
571: the domain wall intersection energy $f_I$.
572:
573: For negative $f_I$ corresponding to $K_4 < 0$ in
574: $\mathcal{F}_\theta$, as we increase the value of $\eta$ the mean
575: field prediction is preserved, and the ground state with long
576: range charge density wave order melts via a first order transition
577: to an isotropic incommensurate floating solid which we labelled as
578: B. This transition occurs when
579: \begin{equation}
580: \eta_{A B} = \frac{p\epsilon}{2\pi}\left[1 - \frac{3}{8}
581: \left(\frac{f_I}{\pi T}\right)^2 \right].
582: \end{equation}
583: That is, for $\eta < \eta_{A B}$ the system is commensurate with
584: $1/\ell_x = 1/\ell_y = 0$ but for $\eta > \eta_{A B}$ there is a
585: jump to the isotropic floating state B with domain wall densities
586: \begin{equation}
587: \frac{1}{\ell_x} = \frac{1}{\ell_y} = -\frac{f_I \epsilon}{\pi^2
588: T^2} \label{lxB} + \sqrt{ \left( \frac{ f_I \epsilon}{\pi^2 T^2}
589: \right)^2 - \frac{2 \epsilon (\epsilon - 2 \pi \eta/p)}{\pi^2
590: T^2}}.
591: \end{equation}
592:
593: If the domain wall intersection energy is positive, implying that
594: collisions are disfavored ($f_I > 0$, $K_4 > 0$), then as we
595: increase the value of $\eta$ from zero the state remains
596: commensurate with long range order and no domain walls ($1/\ell_x
597: = 1/\ell_y = 0$) until
598: \begin{equation}
599: \eta_{A B_S} = \frac{p\epsilon}{2\pi}
600: \end{equation}
601: where there is a second order phase transition to a state with
602: (say)
603: \begin{equation}
604: \frac{1}{\ell_y} = 0 \quad \mbox{;} \quad
605: \frac{1}{\ell_x} = \frac{1}{\pi T} \sqrt{2\epsilon\left( \frac{2\pi
606: \eta}{p} - \epsilon \right)}
607: \label{lxBI}
608: \end{equation}
609: which has domain walls running in \emph{either} the $x$ or $y$
610: direction but not \emph{both}. This anisotropic floating solid or
611: incommensurate striped state which we label B$_S$ has long range
612: charge density wave order with period $p$ in one direction and
613: quasi long range order with power law decay in the perpendicular
614: direction. Near the second-order transition between A and B$_S$,
615: insertion of Eq.~(\ref{lxBI}) into Eq.~(\ref{fd}) shows that
616: $\mathcal{F}_d \sim - (\epsilon - 2\pi \eta/p)^{3/2}$.
617:
618: As $\eta$ is increased further for $f_I >0$, the phase B$_S$
619: persists until
620: \begin{equation}
621: \eta_{B_S B} = \frac{p\epsilon}{2\pi}\left[1 +
622: \left(1.70968\ldots\right)\left(\frac{f_I}{\pi T}\right)^2 \right]
623: \end{equation}
624: where there is a first order transition to the isotropic
625: incommensurate floating solid B with $\ell_x$ and $\ell_y$ still
626: given by Eq.~(\ref{lxB}). In fact, the solution in Eq.~(\ref{lxB})
627: is a local minimum of the free energy for all $\eta > p
628: \epsilon/(2 \pi)$. However, its free energy behaves like
629: $\mathcal{F}_d \sim - (\epsilon - 2\pi \eta/p)^{2}$ upon
630: approaching phase A. Close enough to phase A, this free energy is
631: always larger than the free energy for phase B$_S$.
632:
633: Therefore, if the intersection energy for domain walls is positive,
634: the melting of the commensurate solid A always occurs via a
635: second-order transition to the striped anisotropic state B$_S$.
636:
637:
638: \subsection{Renormalization of elastic constants in incommensurate phases}
639: \label{sec:spinwave}
640:
641: Once the domain walls have proliferated in phases B and B$_S$, the
642: density correlations become incommensurate and appear at
643: wavevectors which are shifted from the commensurate values in
644: Eq.~(\ref{defPhi}). We are interested here in the long wavelength
645: ``spinwave'' density fluctuations about this incommensurate state.
646:
647: First, let us consider the phase B. Rather than defining the
648: spin-wave variables about the commensurate state as in
649: Eq.~(\ref{deftheta}), we now need to look at fluctuations above an
650: incommensurate ordered state. So now we write
651: \begin{equation}
652: \Phi_x \propto e^{i 2 \pi x/(p \ell_x)} e^{i \theta_x}~~;~~ \Phi_y
653: \propto e^{i 2 \pi y/(p \ell_y)} e^{i \theta_y}, \label{deftheta2}
654: \end{equation}
655: because the incommensurate ordering wavevectors are $ (2 \pi /a)
656: (1/p+a/\ell_x, 0)$ and $(2 \pi /a) (0,1/p+a/\ell_y)$. We are
657: interested in the effective action for these $\theta_{x,y}$ on a
658: coarse-grained scale much larger than $\ell_{x,y}$. This action
659: can be deduced by the same symmetry arguments made in
660: Section~\ref{sec:mft}. It is not difficult to see that this action
661: in the floating phase B has the same structure as
662: $\mathcal{F}_\theta$ in Eq.~(\ref{ftheta}), except that now
663: $\eta=h=0$:
664: \begin{eqnarray}
665: \mathcal{F}_{{\rm B}} &=& \int d^2 r \Biggl[ \frac{K_1}{2} \left[
666: (\partial_x \theta_x)^2 + (\partial_y \theta_y )^2 \right]
667: \nonumber \\
668: &+& \frac{K_2}{2}\left[ (\partial_x \theta_y)^2 + (\partial_y
669: \theta_x )^2 \right] + K_3 (\partial_x \theta_y)(
670: \partial_y \theta_x)
671: \nonumber \\
672: &+& K_4 (\partial_x \theta_x) (\partial_y \theta_y) \Biggr]
673: \label{ftheta1}
674: \end{eqnarray}
675:
676: For the phase B$_S$, the $\theta_y$ variables (say) are locked at
677: their commensurate value, and so only the $\theta_x$ variables
678: will contribute to the spin-wave theory:
679: \begin{eqnarray}
680: \mathcal{F}_{{\rm B}_S} &=& \int d^2 r \Biggl[ \frac{K_1}{2}
681: (\partial_x \theta_x)^2+ \frac{K_2}{2} (\partial_y
682: \theta_x )^2 \Biggr] \label{ftheta2}
683: \end{eqnarray}
684:
685: An important point is that the stiffnesses $K_{1-4}$ in
686: Eqs.~(\ref{ftheta1}) and (\ref{ftheta2}) be strongly renormalized
687: from the bare values in Eq.~(\ref{ftheta}). Following the analysis
688: of Ref.~\onlinecite{sue}, we will now estimate their renormalized
689: values $K_{1-4}^R$ in terms of the parameters appearing in the domain
690: wall free energies in Section~\ref{sec:wandering} (the superscript
691: $R$ will be used for clarity only in this subsection).
692:
693: First, imagine that we impose a uniform compressional strain
694: $\partial_x \theta_x$ on the floating solid. On the average, this
695: will cause the domain walls to move closer to each other, and
696: change the value of $\ell_x$ to $\ell_x - \delta \ell_x$. Because
697: $\theta_x$ changes by $2\pi/p$ across each domain wall, we
698: conclude that $\partial_x \theta_x = (2 \pi/p) \delta \ell_x
699: /\ell_x^2$. The change in $\ell_x$ will cause a change in free
700: energy that can be computed from Eq.~(\ref{fd}), and so we
701: conclude
702: \begin{equation}
703: K_1^R = \frac{p^2 \ell_x^4}{4 \pi^2} \frac{\partial^2
704: \mathcal{F}_d}{\partial \ell_x^2},
705: \end{equation}
706: where the derivative has to be computed at the equilibrium values
707: of $\ell_x$, $\ell_y$ which minimize Eq.~(\ref{fd}). This
708: determines
709: \begin{equation}
710: K_1^R = \frac{p^2 T^2}{4 \epsilon \ell_x},
711: \end{equation}
712: in both phases B and B$_S$. Similarly, we can apply a combined
713: compressional strain in $\theta_x$ and $\theta_y$ and conclude
714: \begin{equation}
715: K_4^R = \frac{p^2 \ell_x^2 \ell_y^2}{4 \pi^2} \frac{\partial^2
716: \mathcal{F}_d}{\partial \ell_x \partial \ell_y},
717: \end{equation}
718: This is non-zero only in the floating phase B, and we obtain
719: \begin{equation}
720: K_4^R =\frac{p^2 f_I}{4\pi^2}~~,~~\mbox{phase B only}.
721: \end{equation}
722:
723: To determine $K_2^R$, apply a small uniform shear strain to
724: $\theta_x$ by inducing a non-zero $\partial_y \theta_x$. This will
725: move the domain walls in $\theta_x$ such that the position of each
726: domain wall obeys $\partial_y u_x = (p \ell_x/(2 \pi)) \partial_y
727: \theta_x$. Inserting this into Eq.~(\ref{fwall}), we obtain
728: \begin{equation}
729: K_2^R = \frac{\epsilon}{\ell_x} \left( \frac{ p \ell_x}{2 \pi}
730: \right)^2 = \frac{ \epsilon p^2 \ell_x}{4 \pi^2},
731: \end{equation}
732: in both phases B and B$_S$.
733:
734: Finally, to determine $K_3^R$ we need to apply shear strains to
735: both $\theta_x$ and $\theta_y$. Neither of them causes a net
736: change in the density of domain walls, or in the number of
737: intersections between the domain walls. Consequently there is no
738: change to the free energy beyond that already accounted for by
739: $K_2^R$, and hence
740: \begin{equation}
741: K_3^R = 0,
742: \end{equation}
743: in both phases B and B$_S$.
744:
745: \section{Dislocation mediated melting of floating solids}
746: \label{sec:floating}
747:
748: We will now consider the transition from the isotropic floating
749: solid B and the striped floating solid B$_S$ to the disordered
750: liquid phase C. These transitions are driven by the unbinding of
751: dislocations.
752:
753: With the parameters of the ``spin-wave'' theory at hand in
754: Section~\ref{sec:spinwave}, we can address the energetics of the
755: the dislocations. We will derive the effective action for the
756: dislocations in Section~\ref{sec:disc} and then obtain the
757: renormalization group (RG) flow equations in
758: Section~\ref{sec:discrg}.
759:
760: \subsection{Dislocation interactions}
761: \label{sec:disc}
762:
763: For the most part, this section will be restricted to a discussion
764: of dislocations in phase B. The simpler case of the anisotropic
765: phase B$_S$ is easily obtained by only including those terms
766: arising from the fluctuations of $\theta_x$.
767:
768: Dislocations are simply `vortices' in the angular fields
769: $\theta_{x,y}$ under which
770: \begin{eqnarray}
771: \oint d{\bf r} \cdot \nabla_{\bf r} \theta_x &=& 2 \pi m_x ({\bf
772: r}_v) \nonumber \\
773: \oint d{\bf r} \cdot \nabla_{\bf r} \theta_y &=& 2 \pi m_y ({\bf
774: r}_v) \label{contour}
775: \end{eqnarray}
776: where $m_{x,y}$ are integers at the vortex ({\em i.e.\/}
777: dislocation) site ${\bf r}_v$, and the integral is over a contour
778: that encloses ${\bf r}_v$. Each dislocation is therefore
779: characterized by a doublet of integers $(m_x, m_y)$.
780:
781: To compute the interactions between these vortices, it is useful
782: to define continuum vortex densities by
783: \begin{equation}
784: m_i ({\bf r}) = \sum_v m ({\bf r}_v) \delta ({\bf r} - {\bf r}_v )
785: \end{equation}
786: where $i=x,y$. Then, after transforming to momentum (${\bf k}$)
787: space, the relation Eq.~(\ref{contour}) can be written simply as
788: \begin{equation}
789: k_i \theta_j = k_i \vartheta_j + \frac{2 \pi}{k^2}
790: \epsilon_{i\ell} k_{\ell} m_j \label{vartheta}
791: \end{equation}
792: where $\epsilon$ is the antisymmetric tensor, and $\vartheta_j$ is
793: an arbitrary smooth angular field which has no vortices. We now
794: insert Eq.~(\ref{vartheta}) into the free energy $\mathcal{F}_{\rm
795: B}$ in Eq.~(\ref{ftheta1}), and minimize with respect to
796: $\vartheta_j$. The result for the total free energy of the
797: vortices is then
798: \begin{eqnarray}
799: \mathcal{F}_v &=& \int \frac{d^2 k}{4 \pi^2} \frac{2 \pi^2}{D(k_x,
800: k_y)} \Biggl[ |m_x ({\bf k}) |^2 \left( k_x^2 K_1 (K_2^2 - K_3^2)
801: \right. \nonumber \\ &+& \left. k_y^2 K_2 (K_1^2 - K_4^2) \right)
802: + |m_y ({\bf k}) |^2 \left( k_y^2 K_1 (K_2^2 - K_3^2) \right.
803: \nonumber \\ &+& \left. k_x^2 K_2 (K_1^2 - K_4^2) \right) + m_x
804: ({\bf k}) m_y (- {\bf k}) k_x k_y \left(K_1^2 K_3 \right.
805: \nonumber \\ &+& \left. K_2^2 K_4 - K_3 K_4 (K_3 + K_4 ) \right)
806: \Biggr] \label{fv}
807: \end{eqnarray}
808: where
809: \begin{eqnarray}
810: && D (k_x, k_y) \equiv K_1 K_2 (k_x^2 + k_y^2)^2 \nonumber \\
811: &&~~~~~+\ k_x^2 k_y^2 \left( (K_1 - K_2)^2 - (K_3 + K_4)^2 \right).
812: \end{eqnarray}
813: Note that the values of $K_{1-4}^R$ from
814: Section~\ref{sec:spinwave} are to be inserted into the expressions
815: above; here, and henceforth, the superscript $R$ has been dropped.
816: The interaction between the vortices in now determined by
817: transforming Eq.~(\ref{fv}) back to real space. This takes the
818: form
819: \begin{eqnarray}
820: \mathcal{F}_v &=& E_c \sum_v \left[ m_x^2 ({\bf r}_v ) + m_y^2
821: ({\bf r}_v ) \right] \nonumber \\
822: &+& \sum_{v < v'} \Biggl[ m_x ({\bf r}_v) m_x ({\bf r}_{v'})
823: V\left(|{\bf r}_v - {\bf r}_{v'}|, \phi({\bf r}_v - {\bf
824: r}_{v'})\right) \nonumber \\
825: &+& m_y ({\bf r}_v) m_y ({\bf r}_{v'}) V\left(|{\bf r}_v - {\bf
826: r}_{v'}|, \phi({\bf r}_v - {\bf r}_{v'})+\pi/2\right)
827: \nonumber \\
828: &+& m_x ({\bf r}_v) m_y ({\bf r}_{v'}) W\left(|{\bf r}_v - {\bf
829: r}_{v'}|, \phi({\bf r}_v - {\bf r}_{v'})\right) \Biggr]
830: \end{eqnarray}
831: where $\phi ({\bf r}) = \arctan(y/x)$ is the angle the vector
832: ${\bf r}$ makes with the $x$ axis. We will not need the explicit
833: form of the interaction $W$ in our subsequent analysis, and so we
834: will not specify it explicitly; the interaction $V$ is given by
835: \begin{eqnarray}
836: && V(r, \phi) = \int_0^{\infty} \frac{e^{-ka} dk}{k} \int_0^{2 \pi} d \varphi \left[ e^{i k r
837: \cos(\phi - \varphi)} - 1 \right] \nonumber \\
838: &&\times \frac{ K_1 (K_2^2 - K_3^2) \cos^2 \varphi +
839: K_2 (K_1^2 - K_4^2) \sin^2 \varphi}{K_1 K_2 + \left( (K_1 -
840: K_2)^2 - (K_3 + K_4)^2 \right) \sin^2 \varphi \cos^2 \varphi }
841: \nonumber \\
842: &=& \frac{p^2 T^2}{4\pi^2} \int_0^{2 \pi} d \varphi
843: \int_0^{\infty} \frac{e^{-ka} dk}{k} \left[ \cos( k r
844: \cos(\phi - \varphi)) - 1 \right] \nonumber \\
845: &&~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~\times \Lambda (\widetilde{K}_i, \varphi)
846: \nonumber \\
847: &\equiv& - \frac{p^2 T^2}{2\pi} \widetilde{L}_0 \ln (r/a) + \widetilde{V} (\phi)
848: \label{vrphi}
849: \end{eqnarray}
850: where we have inserted a soft cutoff using the lattice spacing
851: $a$, and
852: \begin{eqnarray}
853: && \!\!\!\!\!\! \Lambda (\widetilde{K}_i, \varphi) \label{deflambda} \\
854: \!&&= \frac{ \widetilde{K}_2 \cos^2 \varphi + \widetilde{K}_1
855: \sin^2 \varphi}{\widetilde{K}_1 \widetilde{K}_2 + \left(
856: (\widetilde{K}_1 - \widetilde{K}_2)^2 - (\widetilde{K}_3 +
857: \widetilde{K}_4)^2 \right) \sin^2 \varphi \cos^2 \varphi }
858: \nonumber
859: \end{eqnarray}
860: is a function of the new couplings $\widetilde{K}_{1-4}$ defined
861: by
862: \begin{eqnarray}
863: \widetilde{K}_1 = \frac{p^2 T^2}{4\pi^2} \frac{K_2}{K_2^2 - K_3^2}
864: ~~&;&~~ \widetilde{K}_2 = \frac{p^2 T^2}{4\pi^2} \frac{K_1}{K_1^2
865: - K_4^2}
866: \nonumber \\
867: \widetilde{K}_3 = \frac{p^2 T^2}{4\pi^2} \frac{K_4}{K_1^2 - K_4^2}
868: ~~&;&~~ \widetilde{K}_4 = \frac{p^2 T^2}{4\pi^2} \frac{K_3}{K_2^2
869: -
870: K_3^2}. \nonumber \\
871: \label{kdual}
872: \end{eqnarray}
873: At this point these definitions of the $\widetilde{K}_{1-4}$ may
874: be viewed as arbitrary variables, but we will see later in
875: Section~\ref{sec:dual} and Appendix~\ref{app:dual} that these are
876: the couplings that appear in a self-dual mapping of the theory
877: $\mathcal{F}_\theta$. Also, from Appendix~\ref{app:dual}, note
878: that the $\widetilde{K}_{1-4}$ couplings appear upon taking the
879: inverse of the matrix of couplings between the strains in
880: $\mathcal{F}_{\rm B}$; consequently, the inverse expressions for
881: the $K_{1-4}$ in terms of the $\widetilde{K}_{1-4}$ have exactly
882: the same structure as in Eq.~(\ref{kdual}). The parameters in the
883: last line of Eq.~(\ref{vrphi}) are given by
884: \begin{eqnarray}
885: \widetilde{L}_0 &=& \int_0^{2 \pi} \frac{ d \varphi}{2 \pi} \Lambda
886: (\widetilde{K}_i, \varphi) \label{lkd} \\
887: &=& \frac{\widetilde{K}_1+\widetilde{K}_2}{\sqrt{\widetilde{K}_1 \widetilde{K}_2
888: \left[(\widetilde{K}_1+\widetilde{K}_2)^2 -(\widetilde{K}_3 +
889: \widetilde{K}_4)^2\right]}} \nonumber \\
890: \widetilde{V} (\phi) &=& -\frac{p^2 T^2}{4\pi^2} \int_0^{2 \pi} d
891: \varphi \ln \left(|\cos(\phi-\varphi)| \right) \Lambda
892: (\widetilde{K}_i, \varphi) \nonumber
893: \end{eqnarray}
894: where the calculation of $\widetilde{L}_0$ is described in
895: Appendix~\ref{app:L0}. For $\widetilde{K}_3 = \widetilde{K}_4=0$,
896: $\widetilde{V} (\phi)$ can be evaluated in closed form:
897: \begin{eqnarray}
898: \widetilde{V} (\phi) &=& \frac{p^2 T^2}{4 \pi (\widetilde{K}_1
899: \widetilde{K}_2)^{1/2}} \ln \left(
900: \frac{\left(\widetilde{K}_1^{1/2} +
901: \widetilde{K}_2^{1/2}\right)^2}{\widetilde{K}_1 \sin^2 \phi +
902: \widetilde{K}_2 \cos^2 \phi} \right). \nonumber \\ \label{wl0}
903: \end{eqnarray}
904:
905: \subsection{Renormalization group flows}
906: \label{sec:discrg}
907:
908: With the knowledge of the interactions between the dislocations,
909: the renormalization group equations can be derived by the methods
910: already described in Refs.~\onlinecite{nh} and~\onlinecite{apy}.
911: We introduce a vortex fugacity, $y = e^{-E_c/T}$ and examine the
912: effect of integrating out pairs of dislocations in an expansion in
913: powers of $y$. A simple and standard analysis shows that the flow
914: equation for the fugacity is
915: \begin{equation}
916: \frac{dy}{d \ell} = \left(2 - \frac{p^2 T}{4\pi} \widetilde{L}_0
917: \right) y. \label{rgy}
918: \end{equation}
919:
920: The renormalization of the $K_i$ from the vortices can be computed
921: by the method described in Appendix B of Ref.~\onlinecite{apy}. We
922: compute the renormalization of the elastic constants, $K_i$ by
923: determining the contribution of the vortices to a two-point
924: correlation of the strains. Such a procedure leads naturally to
925: flow equations for the `inverse' or `dual' $\widetilde{K}_i$
926: couplings, and we obtain
927: \begin{eqnarray}
928: \frac{d \widetilde{K}_1}{d\ell} &=& p^2
929: y^2 T\int_0^{2 \pi} d \phi \cos^2 \phi e^{{\widetilde{V}} (\phi)/T} \nonumber \\
930: \frac{d \widetilde{K}_2}{d\ell} &=& p^2
931: y^2 T\int_0^{2 \pi} d \phi \sin^2 \phi e^{\widetilde{V} (\phi)/T} \nonumber \\
932: \frac{d \widetilde{K}_3}{d\ell} &=& 0\nonumber \\
933: \frac{d \widetilde{K}_4}{d\ell} &=& 0 \label{rgd}
934: \end{eqnarray}
935: We can convert these into equations for the $K_i$ and obtain
936: \begin{eqnarray}
937: \frac{d {K}_1}{d\ell} &=& - \frac{4\pi^2 y^2}{T} (K_1^2 + K_4^2 )
938: \int_0^{2 \pi} d \phi \sin^2 \phi
939: e^{\widetilde{V} (\phi)/T } \nonumber \\
940: \frac{d {K}_2}{d\ell} &=& - \frac{4\pi^2 y^2}{T} (K_2^2 + K_3^2 )
941: \int_0^{2 \pi} d \phi \cos^2 \phi
942: e^{\widetilde{V} (\phi) /T} \nonumber \\
943: \frac{d {K}_3}{d\ell} &=& -\frac{8\pi^2 y^2}{T} K_2 K_3 \int_0^{2 \pi}
944: d \phi \cos^2 \phi e^{\widetilde{V} (\phi)/T} \nonumber \\
945: \frac{d {K}_4}{d\ell} &=& -\frac{8\pi^2 y^2}{T} K_1 K_4 \int_0^{2
946: \pi} d \phi \sin^2 \phi e^{\widetilde{V} (\phi)/T} \label{rgya}
947: \end{eqnarray}
948:
949: While complex in appearance, the flow equations Eq.~(\ref{rgy})
950: and (\ref{rgya}) predict a melting transition of the floating
951: phase B into the liquid phase C which is in a universality class
952: closely related to that of the Kosterlitz-Thouless (KT)
953: transition. The phase B is stable provided $\widetilde{L}_0 > 8
954: \pi/(p^2 T)$. Evaluating $\widetilde{L}_0$ at the values of the
955: $K_i^R$ determined in Section~\ref{sec:spinwave}, we find that
956: phase B is stable everywhere for $p=4$ towards an infinitesimal
957: vortex fugacity for the elastic constants in the domain wall
958: theory; this justifies the neglect of dislocations in our study of
959: domain wall proliferation in Section~\ref{sec:domains}. The
960: present equations Eq.~(\ref{rgya}) show how phase B will
961: ultimately become unstable to a vortex unbinding transition once
962: the vortex fugacity becomes larger. The flow equations in
963: Eq.~(\ref{rgya}) imply singularities in the elastic constants just
964: before the melting transition which can be computed as in
965: Ref.~\onlinecite{ssmelt}. Some of the universal amplitude ratios
966: here will be different from the KT transition, but all exponents
967: and critical singularities will be as in the KT transition.
968:
969: The flow equations for the melting of the anisotropic phase B$_S$
970: can be easily obtained from Eq.~(\ref{rgya}) simply by setting
971: $K_3=K_4=0$. In this case, the values of the elastic constants in
972: Section~\ref{sec:spinwave} and Eq.~(\ref{lkd}) imply that
973: \begin{equation} \widetilde{L}_0 =
974: \frac{4 \pi^2}{p^2 T^2} \left(K_1^R K_2^R \right)^{1/2} =
975: \frac{\pi}{T}.
976: \end{equation}
977: The equation for the vortex fugacity Eq.~(\ref{rgy}) implies then
978: that the vortex fugacity flows to zero as long as\cite{sue} $p^2 >
979: 8$, which is certainly the case for $p>3$. Hence the floating
980: phase B$_S$ is also initially stable to dislocation unbinding. At
981: higher temperatures, there can be a dislocation unbinding
982: transition in the Kosterlitz-Thouless universality class. Note
983: that in this sequence of transitions, even after such a transition
984: has occurred from the phase B$_S$, order has only been lost in $x$
985: direction, and commensurate long-range order remains in
986: $\theta_y$. So the system is still strongly ``striped'' and
987: anisotropic. Full isotropy will be restored only after a second
988: set of similar transitions in $\theta_y$, the first to floating
989: incommensurate order in $\theta_y$, and then a dislocation
990: unbinding transition to the liquid phase C.
991:
992: \section{Direct melting of the commensurate solid}
993: \label{sec:dual}
994:
995: As we noted in Section~\ref{sec:intro} and Fig~\ref{figmft}, it is
996: possible that the low temperature $4\times 4$ commensurate solid
997: can melt directly into the high temperature liquid state. In the
998: theories $\mathcal{F}_\Phi$ in Eq.~(\ref{FP}) and
999: $\mathcal{F}_\theta$ in Eq.~(\ref{ftheta}), such a transition
1000: appears possible at commensurate densities at which
1001: $\overline{\eta} = \eta =0$. So this section will neglect the
1002: influence of $\eta$ and perform a full renormalization group
1003: analysis of the theory $\mathcal{F}_\theta$, including both
1004: dislocations and the $p$-fold lock-in field $h$.
1005:
1006: A key property which aids the analysis is a self-duality of the
1007: action in which the dislocation fugacity, $y$, and the lock-in
1008: field $h$ are interchanged. The origin of this self-duality is
1009: similar to that discussed in Ref.~\onlinecite{jkkn}, and we
1010: present a brief derivation in Appendix~\ref{app:dual}. From this
1011: analysis we obtain the the partition function
1012: $\mathcal{F}_\theta$, including dislocations, obeys
1013: \begin{equation}
1014: \mathcal{Z}_\theta (K_1, K_2, K_3, K_4, h, y) = \mathcal{Z}_\theta
1015: (\widetilde{K}_1, \widetilde{K}_2, \widetilde{K}_3,
1016: \widetilde{K}_4, y, h), \label{dual}
1017: \end{equation}
1018: where the couplings $\widetilde{K}_i$ were defined in
1019: Eq.~(\ref{kdual}).
1020:
1021: Aided by this duality, we can now immediately deduce the full flow
1022: equations for all the elastic constants $K_i$, the vortex fugacity
1023: $y$ and the field $h$ from the results in
1024: Section~\ref{sec:discrg}. As always, these equations are valid for
1025: small $y$ and $h$ and are
1026: \begin{eqnarray}
1027: \frac{dh}{d \ell} &=& (2 - \frac{p^2 T}{4\pi} {L}_0 ) h \nonumber \\
1028: \frac{dy}{d \ell} &=& (2 - \frac{p^2 T}{4\pi} \widetilde{L}_0 ) y
1029: \nonumber \\
1030: \frac{d {K}_1}{d\ell} &=& \int_0^{2 \pi} d \phi \Biggl[ p^2 h^2 T
1031: \cos^2 \phi e^{{V} (\phi)/T} \nonumber \\
1032: &~&~~~~~~~~~~~~- \frac{4\pi^2 y^2}{T} (K_1^2 + K_4^2) \sin^2 \phi
1033: e^{\widetilde{V} (\phi)/T } \Biggr] \nonumber \\
1034: \frac{d {K}_2}{d\ell} &=& \int_0^{2 \pi} d \phi \Biggl[ p^2 h^2 T
1035: \sin^2 \phi e^{{V} (\phi)/T} \nonumber \\
1036: &~&~~~~~~~~~~~~- \frac{4\pi^2 y^2}{T} (K_2^2 + K_3^2) \cos^2 \phi
1037: e^{\widetilde{V} (\phi)/T } \Biggr] \nonumber \\
1038: \frac{d {K}_3}{d\ell} &=& -\frac{8 \pi^2
1039: y^2}{T} K_2 K_3 \int_0^{2 \pi} d \phi \cos^2 \phi e^{\widetilde{V} (\phi)/T} \nonumber \\
1040: \frac{d {K}_4}{d\ell} &=& -\frac{8 \pi^2 y^2}{T} K_1 K_4 \int_0^{2
1041: \pi} d \phi \sin^2 \phi e^{\widetilde{V} (\phi)/T} \label{rgall}
1042: \end{eqnarray}
1043: Here $L_0$ and $V(\phi)$ are defined as in Eq.~(\ref{lkd}), but
1044: with direct couplings:
1045: \begin{eqnarray}
1046: {L}_0 &=& \int_0^{2 \pi} \frac{ d \phi}{2 \pi} \Lambda ({K}_i, \phi) \\
1047: {V} (\phi) &=& -\frac{p^2 T^2}{4\pi^2} \int_0^{2 \pi} d \varphi
1048: \ln \left(|\cos(\phi-\varphi)| \right) \Lambda ({K}_i, \varphi)
1049: \nonumber \label{lk}
1050: \end{eqnarray}
1051:
1052: We will confine our analysis of Eq.~(\ref{rgall}) to the
1053: physically important case of $p=4$. For this case, we first
1054: searched for an intermediate phase with power-law correlations:
1055: such a phase would obtain if there was a set of values of
1056: $K_{1-4}$ for which {\em both\/} $y$ and $h$ flowed to zero. As
1057: shown in Appendix~\ref{app:L0}, there is no such regime of
1058: parameters.
1059:
1060: However, the flow equations in Eq.~(\ref{rgall}) do predict a
1061: direct second-order transition between the $4 \times 4$
1062: commensurate solid A and isotropic liquid C. This transition is
1063: described by a manifold of fixed points. In the 6-dimensional
1064: space of couplings, there is 2-dimensional manifold of fixed
1065: points specified by
1066: \begin{equation}
1067: y=h~~;~~\frac{K_1 K_2}{T^2} = \frac{4}{\pi^2}~~;~~K_3=K_4=0.
1068: \end{equation}
1069: In order to describe the flows near this manifold, it is
1070: convenient to make a change of variables from $y$, $h$, $K_1$, and
1071: $K_2$ to
1072: \begin{eqnarray}
1073: \alpha = y-h~~&,&~~\beta=y+h \nonumber \\
1074: K = \frac{\sqrt{K_1 K_2}}{T}-\frac{2}{\pi}~~&,&~~ \lambda =
1075: \sqrt{\frac{K_1}{K_2}}.
1076: \end{eqnarray}
1077: In these variables, the fixed point manifold is described by
1078: $\alpha=K=K_3=K_4=0$, while the values of $\lambda$ and $\beta$
1079: are arbitrary. All physical properties, including the exponents at
1080: the second-order critical point, will depend upon the bare values
1081: of $\lambda$ and $\beta$. We expanded Eqs.~(\ref{rgall}) to linear
1082: order in $\alpha$, $K$, $K_3$, and $K_4$, and after
1083: diagonalization of the flow equations, obtained the following
1084: renormalization group eigenvalues in this 4-dimensional subspace:
1085: \begin{eqnarray}
1086: && 2 \pi^2 \beta \left(\sqrt{\lambda} + 1/\sqrt{\lambda} \right)^4
1087: \nonumber \\
1088: &&~~~~~~~~\times \left( 1 \pm \left( 1+ \frac{4}{\pi^2 \beta
1089: \left(\sqrt{\lambda} +
1090: 1/\sqrt{\lambda} \right)^4}\right)^{1/2} \right),\nonumber \\
1091: &&~~~~- 4\pi^2 \beta^2 \left(\sqrt{\lambda} + 1/\sqrt{\lambda}
1092: \right)^4,
1093: \end{eqnarray}
1094: where the last eigenvalue is doubly degenerate. It is evident that
1095: for all $\lambda$, $\beta$ there are 3 negative eigenvalues and 1
1096: positive eigenvalue. This flow therefore describes a conventional
1097: second-order transition, with the correlation length exponent
1098: $\nu$ equal to the inverse of the positive eigenvalue.
1099:
1100: \section{Conclusions}
1101: \label{sec:conc}
1102:
1103: We have shown that thermal fluctuations on $4 \times 4$ ordered
1104: state lead generically to a rather rich phase diagram as a
1105: function of temperature and carrier concentration. This phase
1106: diagram generically must have regions with incommensurate and
1107: anisotropic ordering. Thus it is remains within the realm of
1108: possibility that the underlying physics of density wave ordering
1109: in all the cuprates is the same, and the distinctions between the
1110: experiments are entirely due to their distinct locations in our
1111: phase diagram.
1112:
1113: For easy reference, we conclude by listing the various routes the
1114: $4 \times 4$ solid A can melt into the liquid C. The list below is
1115: not exhaustive, and omits certain possibilities involving strong
1116: first order transitions between unrelated phases.
1117: \begin{enumerate}
1118: \item Phase A undergoes a second order Pokrovsky-Talapov \cite{pt}
1119: (PT) transition to the incommensurate striped phase B$_S$, as
1120: described in Section~\ref{sec:wandering}. This is followed by a
1121: first order transition into phase B as also described in
1122: Section~\ref{sec:wandering} and Fig~\ref{figbs}. Then phase B
1123: undergoes a dislocation mediated melting transition into phase C
1124: which is described by Eqs.~(\ref{rgy}) and (\ref{rgya}). This last
1125: transition is in a universality class which is nearly, but not
1126: exactly, KT. \item PT transition from phase A to phase B$_S$ as in
1127: 1. Phase B$_S$ then has a KT transition to a striped state which
1128: has long range order with period 4 along one direction, and
1129: exponentially decaying correlations along the other. This striped
1130: set melts into C after a second set of similar transitions: a PT
1131: transition into a state with incommensurate quasi-long range order
1132: in one direction only, and then finally a KT transition in C.
1133: \item First order transition from A to B as described in
1134: Section~\ref{sec:wandering} and Fig~\ref{figbs}. Then, a
1135: transition from B to C as in 2. \item At special commensurate
1136: densities, there is a direct, self-dual, second-order transition
1137: from A to C, described in Section~\ref{sec:dual}, with
1138: continuously varying exponents.
1139: \end{enumerate}
1140:
1141:
1142: \acknowledgments We thank E.~Fradkin, S.~Kivelson, and T.~Lubensky
1143: for valuable discussions. This research was supported by the
1144: National Science Foundation under grant DMR-0098226, and under
1145: grants DMR-0210790, PHY-9907949 at the Kavli Institute for
1146: Theoretical Physics. S.S. was also supported by the John Simon
1147: Guggenheim Memorial Foundation.
1148:
1149: \appendix
1150:
1151: \section{Evaluation of $\widetilde{L}_0$}
1152: \label{app:L0}
1153:
1154: This appendix will present the derivation of $\widetilde{L}_0$ in
1155: Eq.~(\ref{lkd}) and its application to the flow equations presented in
1156: Section~\ref{sec:dual}. We begin by writing the integral explicitly as
1157: \begin{eqnarray}
1158: \widetilde{L}_0 &=& \int_{0}^{2\pi} \frac{d \varphi}{2\pi} \label{intL0} \\
1159: &\times&\!\! \frac{ \widetilde{K}_2 \cos^2 \varphi + \widetilde{K}_1
1160: \sin^2 \varphi}{\widetilde{K}_1 \widetilde{K}_2 + \left[
1161: (\widetilde{K}_1 - \widetilde{K}_2)^2 - (\widetilde{K}_3 +
1162: \widetilde{K}_4)^2 \right] \sin^2 \varphi \cos^2 \varphi }
1163: \nonumber \\
1164: &=& \int_{0}^{2\pi} \frac{d \varphi}{\pi}\left[ \frac{\widetilde{K}_1 +
1165: \widetilde{K}_2}{\alpha^{+} - \alpha^{-}\cos^2 2\varphi} +
1166: \frac{(\widetilde{K}_2 - \widetilde{K}_1)\cos 2\varphi}
1167: {\alpha^{+} - \alpha^{-}\cos^2 2\varphi} \right]
1168: \nonumber
1169: \end{eqnarray}
1170: where
1171: \begin{equation}
1172: \alpha^{\pm} = (\widetilde{K}_1 \pm \widetilde{K}_2)^2 -
1173: (\widetilde{K}_3+\widetilde{K}_4)^2.
1174: \end{equation}
1175: The second term in the last line of Eq.~(\ref{intL0}) is identically zero and
1176: we are left with
1177: \begin{eqnarray}
1178: \widetilde{L}_0 &=& (\widetilde{K}_1 + \widetilde{K}_2)
1179: \int_{0}^{2\pi} \frac{d \varphi}{\pi} \frac{1}{\alpha^{+} - \alpha^{-}
1180: \cos^2 2\phi} \nonumber \\
1181: &=& 2\frac{\widetilde{K}_1 + \widetilde{K}_2} {\sqrt{\alpha^{+}(\alpha^{+} -
1182: \alpha^{-})}} \nonumber \\
1183: &=& \frac{\widetilde{K}_1+\widetilde{K}_2}{\sqrt{\widetilde{K}_1
1184: \widetilde{K}_2 \left[(\widetilde{K}_1+\widetilde{K}_2)^2
1185: -(\widetilde{K}_3 + \widetilde{K}_4)^2\right]}}, \label{L0dual}
1186: \end{eqnarray}
1187: which can be written in terms of our original coupling constants $K_i$ as
1188: \begin{equation}
1189: \widetilde{L}_0 = \frac{4\pi^2}{p^2T^2}
1190: \frac{K_1(K_2^2-K_3^2)+K_2(K_1^2-K_4^2)}
1191: {\sqrt{K_1K_2\left[(K_1+K_2)^2-(K_3+K_4)^2)\right]}}.
1192: \label{L0dualnt}
1193: \end{equation}
1194: This is not to be confused with $L_0$ from Eq.~(\ref{rgall}) which
1195: is given by
1196: \begin{equation}
1197: L_0 = \frac{K_1+K_2}{\sqrt{K_1K_2\left[(K_1+K_2)^2-(K_3+K_4)^2\right]}}.
1198: \label{L0reg}
1199: \end{equation}
1200:
1201: For the direct transition from A to C (Section~\ref{sec:dual}) for $p=4$ we wish
1202: to determine if there is any region close to the manifold of fixed points
1203: where both $h$ and $y$ are irrelevant. We do this by expanding
1204: Eqs.~(\ref{L0dualnt}) and (\ref{L0reg}) for small $K_3$ and $K_4$
1205: parameterized by
1206: \begin{equation}
1207: K_3 = \delta_3 K_2 \quad \mbox{;} \quad K_4 = \delta_4 K_1,
1208: \end{equation}
1209: where $|\delta_3| \mbox{ and } |\delta_4| < 1$ and substituting into
1210: the expressions for $dh/d\ell$ and $dy/d\ell$ in Eq.~(\ref{rgall}). Define
1211: \begin{equation}
1212: \beta(h) = \frac{1}{h}\frac{d h}{d\ell} \quad \mbox{;} \quad
1213: \beta(y) = \frac{1}{y}\frac{d y}{d\ell},
1214: \end{equation}
1215: it is then straightforward to show that $\beta(h) \gtrless 0$ if
1216: \begin{equation}
1217: K_1 \gtrless K_1^\ast\left[1 + \left(\frac{\pi^2 K_2^2
1218: \delta_3 + 4\delta_4T^2}{\pi^2 K_2^2+4T^2}\right)^2 \right] +
1219: \mathrm{O}\left(\delta_i^3\right)
1220: \label{K1h}
1221: \end{equation}
1222: and $\beta(y) \lessgtr 0$ if
1223: \begin{eqnarray}
1224: K_1 &\gtrless& K_1^\ast \left[1 + \left(\frac{\pi^2 K_2^2
1225: \delta_3 + 4\delta_4T^2}{\pi^2 K_2^2+4T^2}\right)^2 \right. \nonumber \\
1226: && \left.+ 2\left(\frac{2\pi K_2 T (\delta_4-\delta_3)}
1227: {\pi^2 K_2^2 + 4T^2}\right)^2 \right ] + \mathrm{O}\left(\delta_i^3\right)
1228: \label{K1y}
1229: \end{eqnarray}
1230: where $K_1^\ast = 4T^2/(\pi^2 K_2)$. The relative signs of $\beta(h)$ and
1231: $\beta(y)$ are shown in Fig.~\ref{figbetahy} and from
1232: Eqs.~(\ref{K1h}) and (\ref{K1y}) we have calculated that the two lines
1233: describing the zeros of $\beta(h)$ and $\beta(y)$ are always separated by
1234: \begin{equation}
1235: \frac{32T^4K_2(\delta_4-\delta_3)^2}{(\pi^2K_2^2+4T^2)^2} > 0
1236: \end{equation}
1237: and thus there are no values of $K_1$ and $K_2$ for small $K_3$
1238: and $K_4$ where the $4$-fold anisotropy and vortices are
1239: \emph{both} irrelevant.
1240: \begin{figure}
1241: \centering
1242: \includegraphics[width=3in]{fig_betahy.eps}
1243: \caption{The signs of $\beta(h)$ and $\beta(y)$ in the
1244: $K_1-K_2$ plane for $K_3 = -3K_2/5$ and $K_4 = 4K_1/5$. The line
1245: which indicates $\beta(h)=0$ always lies below the line describing $\beta(y)=0$
1246: and consequently $h$ and $y$ are never both irrelevant.}
1247: \label{figbetahy}
1248: \end{figure}
1249:
1250: \section{Duality mapping}
1251: \label{app:dual}
1252:
1253: This appendix will outline the derivation of Eq.~(\ref{dual}).
1254:
1255: It is useful to first consider $\mathcal{F}_\theta$ in
1256: Eq.~(\ref{ftheta}) with $h=\eta=0$, but to include the effect of
1257: vortices. In other words, we do wish the impose periodicity under
1258: $\theta_i \rightarrow \theta_i + 2 \pi$.
1259:
1260: First, we write the relevant portion of the action in the form
1261: \begin{equation}
1262: \frac{1}{2} \int d^2 r \partial_i \theta_a C_{ia,jb} \partial_j
1263: \theta_b
1264: \end{equation}
1265: where the indices $i$, $j$, $a$, $b$ extend over $x$, $y$, and the
1266: matrix of couplings $C$ can be easily related to the elastic
1267: constants $K_i$ in Eq.~(\ref{ftheta}). Now, we decouple this by a
1268: set of currents $J_{ia}$ and write the action as
1269: \begin{equation}
1270: \int d^2 r \left[ \frac{1}{2} J_{ia} C_{ia,jb}^{-1} J_{jb} + i
1271: J_{ia} \partial_i \theta_a \right]
1272: \end{equation}
1273: Imposing periodicity in $\theta_i \rightarrow \theta_i + 2 \pi$ is
1274: now equivalent to the requirement that the $J_{ia}$ are integers.
1275: We now integrate the $\theta_a$ out, and solve the resulting
1276: constraint equations by writing
1277: \begin{equation}
1278: J_{ia} = \frac{p}{2\pi} \epsilon_{ij} \partial_j
1279: \widetilde{\theta}_a.
1280: \end{equation}
1281: The constraints that the $J_{ia}$ are integers can now be imposed
1282: by demanding that the $\widetilde{\theta}_a$ take values which are
1283: integer multiples of $2\pi/p$. As usual, we can soften this
1284: constraint by introducing a vortex fugacity $y$, and so obtain the
1285: effective action
1286: \begin{equation}
1287: \int d^2 r \left[ \frac{p^2}{8 \pi^2} \epsilon_{ii'} \partial_{i'}
1288: \widetilde{\theta}_a C_{ia,jb}^{-1} \epsilon_{jj'} \partial_{j'}
1289: \widetilde{\theta}_b - y \sum_a \cos (p \widetilde{\theta}_a)
1290: \right]
1291: \end{equation}
1292: Upon explicitly working out the inverse of the matrix $C$, we find
1293: that this action has exactly the same form as the original
1294: $\mathcal{F}_{\theta}$, with the vortex fugacity $y$ playing the
1295: role of the lock-in field $h$, and the couplings $K_i$ replaced by
1296: the couplings $\widetilde{K}_i$ in Eq.~(\ref{kdual}).
1297:
1298:
1299: \begin{thebibliography}{99}
1300:
1301: \bibitem{howald} C. Howald, H. Eisaki, N. Kaneko, M. Greven, and A. Kapitulnik,
1302: Phys. Rev. B, {\bf 67}, 014533 (2003).
1303:
1304: \bibitem{mcelroy} K.~McElroy, D.-H.~Lee, J.~E.~Hoffman, K.~M.~Lang,
1305: E.~W.~Hudson, H.~Eisaki, S.~Uchida, J.~Lee, and J.~C.~Davis,
1306: cond-mat/0404005.
1307:
1308: \bibitem{hanaguri} T.~Hanaguri, C.~Lupien, Y.~Kohsaka, D.-H.~Lee,
1309: M.~Azuma, M.~Takano, H.~Takagi, and J.~C.~Davis, Nature {\bf 430}, 1001 (2004).
1310:
1311: \bibitem{fang} A.~Fang, C.~Howald, N.~Kaneko, M.~Greven, and
1312: A.~Kapitulnik, Phys. Rev. B {\bf 70}, 214514 (2004).
1313:
1314: \bibitem{ali} M.~Vershinin, S.~Misra, S.~Ono, Y.~Abe, Y.~Ando,
1315: and A.~Yazdani, Science {\bf 303}, 1995 (2004).
1316:
1317: \bibitem{jtran} J.~M.~Tranquada, H.~Woo, T.~G.~Perring, H.~Goka,
1318: G.~D.~Gu, G.~Xu, M.~Fujita, and K.~Yamada, Nature {\bf 429}, 534 (2004).
1319:
1320: \bibitem{hinkov}
1321: V.~Hinkov, S.~Pailhes, P.~Bourges, Y.~Sidis, A.~Ivanov, A.~Kulakov,
1322: C.~T.~Lin, D.~P.~Chen, C.~Bernhard, and B.~Keimer, Nature {\bf 430},
1323: 650 (2004).
1324:
1325: \bibitem{buyers}
1326: C.~Stock, W.~J.~L.~Buyers, R.~A.~Cowley, P.~S.~Clegg, R.~Coldea,
1327: C.~D.~Frost, R.~Liang, D.~Peets, D.~Bonn, W.~N.~Hardy, and
1328: R.~J.~Birgeneau, cond-mat/0408071.
1329:
1330: \bibitem{hayden}
1331: S.~M.~Hayden, H.~A.~Mook, P.~Dai, T.~G.~Perring, and F.~Dogan,
1332: Nature {\bf 429}, 531 (2004).
1333:
1334: \bibitem{MVTU}
1335: M. Vojta and T. Ulbricht, Phys. Rev. Lett. 93, 127002 (2004).
1336:
1337: \bibitem{GSU}
1338: G. S. Uhrig, K. P. Schmidt, and M. Gr\"uninger, cond-mat/0402659.
1339:
1340: \bibitem{vs} M.~Vojta and S.~Sachdev, cond-mat/0408461.
1341:
1342: \bibitem{nh} D.~R.~Nelson and B.~I.~Halperin, Phys. Rev. B {\bf 19}, 2457
1343: (1979).
1344:
1345: \bibitem{apy} A.~P.~Young, Phys. Rev. B {\bf 19}, 1855 (1979).
1346:
1347: \bibitem{pt} V.~L.~Pokrovsky and A.~L.~Talapov, Phys. Rev. Lett.
1348: {\bf 42}, 65 (1979); Zh. Eksp. Teor. Fiz. {\bf 78}, 269 (1980) [Sov.
1349: Phys. JETP {\bf 51}, 134 (1980)].
1350:
1351: \bibitem{ostlund} S.~Ostlund, Phys. Rev. B {\bf 24}, 398 (1981).
1352:
1353: \bibitem{sue} S.~N.~Coppersmith, D.~S.~Fisher, B.~I.~Halperin,
1354: P.~A.~Lee, and W.~F.~Brinkman, Phys. Rev. B {\bf 25}, 349 (1982).
1355:
1356: \bibitem{schulz} H.~J.~Schulz, Phys. Rev. B {\bf 28}, 2746 (1983).
1357:
1358: \bibitem{haldane} F.~D.~M.~Haldane, P.~Bak, and T.~Bohr,
1359: Phys. Rev. B {\bf 28}, 2743 (1983).
1360:
1361: \bibitem{huse} D.~A.~Huse and M.~E.~Fisher,
1362: Phys. Rev. B {\bf 29}, 239 (1984).
1363:
1364: \bibitem{chaikinlub} P.~M.~Chaikin and T.~C.~Lubensky, {\em The Principles of Condensed Matter
1365: Physics}, Cambridge University Press, Cambridge (2000).
1366:
1367: \bibitem{kfe} S.~A.~Kivelson, E.~Fradkin, and V.~J.~Emery, Nature {\bf 393},
1368: 550 (1998).
1369:
1370: \bibitem{jkkn} J.~V.~Jos\'e, L.~P.~Kadanoff, S.~Kirkpatrick, and
1371: D.~R.~Nelson, Phys. Rev. B {\bf 16}, 1217 (1977).
1372:
1373: \bibitem{zachar} O.~Zachar, S.~A.~Kivelson, and V.~J.~Emery, Phys. Rev.
1374: B {\bf 57}, 1422 (1998).
1375:
1376: \bibitem{ying} Y. Zhang, E. Demler and S. Sachdev, Phys. Rev. B {\bf 66}, 094501
1377: (2002).
1378:
1379: \bibitem{ssmelt} S. Sachdev, Phys. Rev. B {\bf 31}, 4476 (1985).
1380:
1381: \end{thebibliography}
1382:
1383: \end{document}
1384: