1: \documentclass[prl,twocolumn]{revtex4}
2: \usepackage{array,amssymb,graphicx}
3:
4: %
5: % Definition of wide equation command
6: %
7: % Usage \wideeqn{arg}{Text}
8: % arg = t - used if the wide equation is at the top of the page.
9: % arg = b - used if the wide equation is at the bottom of the page.
10: % arg = m - otherwise.
11: %
12: \def\wideeqn#1#2{
13: \begin{widetext}
14: \noindent
15: \if#1t
16: \else
17: \raisebox{9pt}[0in][0.0in]
18: {$\rule{3.4in}{0.4pt}\rule{0.4pt}{6pt}$\hspace{3.6in}}
19: \fi
20: #2
21: \if#1b
22: \else
23: \raisebox{-9pt}[0in][0.0in]
24: {\hspace{3.55in}$\rule{0.4pt}{6pt}\rule[6pt]{3.5in}{0.4pt}$}
25: \fi
26: \end{widetext}
27: \noindent
28: }
29:
30: \begin{document}
31:
32: \title{Effective theory of incompressible quantum Hall liquid crystals}
33:
34: \author{Michael M. Fogler}
35:
36: \affiliation{Department of Physics, Massachusetts Institute of
37: Technology, 77 Massachusetts Avenue, Cambridge, Massachusetts 02139}
38:
39: %\date{\today}
40:
41: \begin{abstract}
42:
43: I propose an effective theory of zero-temperature phases of the quantum
44: Hall stipes: a smectic phase where the stripes are static and a novel
45: quantum nematic phase where the positional order is destroyed by
46: quantum fluctuations. The nematic is viewed as a Bose condensate
47: of dislocations whose interactions are mediated by a $U(1)$ gauge field.
48: Collective mode spectrum and the dynamical structure factor in the two
49: phases are calculated.
50:
51: \end{abstract}
52: \pacs{PACS numbers: 73.40.Hm}
53:
54: \maketitle
55:
56: Stripe phases are extremely common in nature. Once studied mainly in the
57: context of pattern formation and soft condensed-matter systems
58: (convection rolls, ferrofluids, diblock co-polymers, {\it
59: etc.\/}~\cite{Seul_95}), they are now recognized to be important in
60: venues ranging from neutron stars~\cite{Neutron_star} to neural networks
61: of the brain~\cite{Chklovskii}. The ``hard'' condensed-matter community
62: was alerted to the relevance of stripes after their discovery in
63: transitional metal oxides, especially, high-$T_c$
64: cuprates~\cite{High_Tc_stripes}. The subject of this Letter is the
65: stripe phase in another remarkable correlated system: a two-dimensional
66: (2D) electron liquid in a transverse magnetic field. This particular
67: phase is formed when the Landau level occupation factor $\nu$ is close
68: to a half-integer and is larger than some critical value
69: $\nu_c$~\cite{Fogler_96}. Unlike the still debated case of high-$T_c$,
70: the theory of the quantum Hall stripes rests on a much more solid
71: foundation~\cite{Fogler_96,Moessner_96}. The quantity $1 / \nu$ plays
72: the role of a small parameter controlling the strength of quantum
73: fluctuations beyond mean-field. Thus, the mean-field stripe solution is
74: stable and adequately captures the main properties of the ground state
75: at large $\nu$. Numerical calculations~\cite{Critical_nu,Scarola_01}
76: indicate that $\nu_c \sim 4$ for the physically relevant case of Coulomb
77: interaction. This provides a natural explanation of the magnetotransport
78: anisotropy observed in GaAs heterostructures~\cite{Experiments} at $\nu
79: \geq \frac92$.
80:
81: The states at $\nu \sim \nu_c$ are much less understood because the
82: fluctuations about the mean-field Hartree-Fock solution are large and
83: several almost degenerate states are in competition. Small changes in
84: microscopic parameters may therefore radically change the
85: nature of the ground state and bring to life novel quantum phases.
86:
87: %
88: % FIG. 1
89: %
90: \begin{figure}
91: \includegraphics[width=2.1in,bb=118 308 557 518]{qn_1.ps}
92: %\vspace{0.1in}
93: \setlength{\columnwidth}{3.2in}
94: \caption{
95: (a) Smectic (b) Nematic. One dislocation is encircled.
96: \label{Fig_real_space}
97: }
98: \end{figure}
99: %
100:
101: Particularly intriguing are proposed quantum {\it smectic\/} and quantum
102: {\it nematic\/} phases~\cite{Balents_96,Fradkin_99}. In the smectic
103: phase the system lacks translational symmetry (is periodically
104: modulated) along one of the spatial directions, say $\hat{\bf x}$, but
105: remains liquid-like along $\hat{\bf y}$. An example of the smectic phase
106: is a unidirectional charge-density-wave Hartree-Fock
107: state~\cite{Fogler_96}; however, the true quantum smectic must possess
108: certain amount of fluctuations beyond Hartree-Fock (shown pictorially in
109: Fig.~\ref{Fig_real_space}a) to prevent another periodic modulation along
110: $\hat{\bf
111: y}$~\cite{Fradkin_99,Yi_01,MacDonald_00,Fogler_in_preparation}. In
112: contrast, the nematic state is fully translationally invariant but lacks
113: the rotational invariance. It can be visualized as fluctuating stripes
114: preferentially aligned along the $\hat{\bf y}$-direction, see
115: Fig.~\ref{Fig_real_space}b. At present, concrete trial wavefunctions for
116: the quantum nematic are known only for specific filling
117: fractions~\cite{Musaelian_96}. This supports the idea that this phase is
118: physically realizable but forces us to seek approaches other than
119: ``wavefunction engineering'' to gain understanding. Below I construct
120: an effective long-wavelength low-frequency theory that applies to a
121: class of quantum Hall liquid crystals. It describes zero-temperature
122: properties and the quantum phase transition between these intriguing
123: states~\cite{Comment_on_1st_order}.
124:
125: {\it Smectic phase\/}.--- The smectic emerges from a uniform state when
126: a pair of collective modes with wavevectors ${\bf q} = \pm q_0 \hat{\bf
127: x}$ goes soft and condenses~\cite{Comment_on_soft_modes}. Hydrodynamic
128: fluctuations of the electron density projected onto the topmost Landau
129: level become of the form $n({\bf r}, t) + {\rm Re}\,\{\Psi({\bf r}, t)
130: e^{-i q_0 x}\}$, where $n$ is a smooth density component (with $q \ll
131: q_0$), $\Psi = \Psi_0 e^{i q_0 u}$ is the smectic order parameter, and
132: $u$ is the deviation of the stripes (density maxima) from uniformity,
133: see Fig.~\ref{Fig_real_space}a. To build an effective theory I make a
134: crucial assumption that the Goldstone mode associated with $u$ is {\it
135: the only\/} low-energy degree of freedom of the system. In general,
136: other low-energy excitations, e.g., those inherited from the parent
137: uniform state may be present. Thus, our theory describes only a certain
138: class of possible smectic states, those that are superstructures on top
139: of incompressible quantum Hall liquids~\cite{Comment_on_compressible}.
140: I proceed by writing down the effective Hamiltonian for $u$ and $n$,
141: containing only the most relevant terms consistent with the
142: symmetry~\cite{DeGennes_book,Fogler_00}:
143: %
144: \begin{equation}
145: H = \frac{Y}{2} (\partial_x u)^2
146: + \frac{K}{2} (\partial_y^2 u)^2
147: + \frac12 \delta n {\bf U} \delta n + C \delta n \partial_x u.
148: \label{H}
149: \end{equation}
150: %
151: Here $Y$ and $K$ are phenomenological compression and bending moduli,
152: $\delta n = n - n_0$ is the deviation of local density from the
153: equilibrium value $n_0$, and ${\bf U}$ is the integral operator with
154: kernel $U({\bf r})$, the electron-electron interaction potential. $U$ is
155: Coulombic at large $r$ but is modified by many-body screening, exchange
156: and correlation effects at short distances (see more details
157: in~\cite{Fogler_96}). The last term in Eq.~(\ref{H}) accounts for the
158: dependence of the smectic period on $n_0$, with $C = Y \partial \ln q_0
159: / \partial n_0$. It vanishes at the half-filling due to electron-hole
160: symmetry but is nonzero otherwise.
161:
162: One may wonder why include ``incompressible'' background $n$ in our
163: low-energy Hamiltonian~(\ref{H}). The reason is the important role $n$
164: plays in the dynamics of $u$. Indeed, $u$ determines the density
165: fluctuations near the star of the soft mode, e.g., $n_{{\bf q}_0 + {\bf
166: k}} = \frac12 (\Psi_0 e^{-i q_0 u})_{\bf k}$, while the density
167: operators $n_{\bf q}$ projected onto a single Landau level are
168: dynamically linked by the commutation relation~\cite{Girvin_86},
169: $[n_{\bf q}, n_{{\bf q}^\prime}] = 2 i \sin (\frac12 l^2 {\bf q} \wedge
170: {\bf q}^\prime) n_{{\bf q} + {\bf q}^\prime}$, where $l = \sqrt{\hbar /
171: m \omega_c}$ is the magnetic length and $\omega_c = e B / m c$ is the
172: cyclotron frequency for the external magnetic field $B$. The above
173: commutation relation is exact but difficult to deal with. I achieve a
174: simplification by replacing it with its expectation value expanded to
175: the lowest order in ${\bf q}$. This way I obtain the following
176: approximate commutation relation for use in our effective theory:
177: %
178: \begin{equation}
179: [n_{\bf q}, u_{{\bf q}^\prime}] = (2 \pi)^2 l^2 q_y
180: \delta({\bf q} + {\bf q}^\prime)
181: \label{commutator}
182: \end{equation}
183: %
184: (all other commutators vanish in the $q \to 0$ limit). Introducing the
185: canonical momentum $p$ by $\delta n = -\partial_y p$, I unify
186: Eqs.~(\ref{H}) and (\ref{commutator}) into the effective
187: action~\cite{Comment_on_commutator},
188: %
189: \begin{eqnarray}
190: A_{\rm sm} &=& \int\limits_0^\beta d \tau \int d^2 r \left\{
191: -\frac{i}{l^2} \hbar p \partial_\tau u + \frac{Y}{2} (\partial_x u)^2
192: + \frac{K}{2} (\partial_y^2 u)^2
193: \right.
194: \nonumber\\
195: \mbox{} &+& \left.\frac12 (\partial_y p) {\bf U} (\partial_y p)
196: - C \partial_x u \partial_y p
197: \right\}.
198: \label{A_sm}
199: \end{eqnarray}
200: %
201: This general form can be shown to pass two important tests. First,
202: after a change of variables it reproduces~\cite{Fogler_in_preparation}
203: an effective action derived for a specific model of the quantum Hall
204: smectic~\cite{MacDonald_00}. Second, the density structure factor,
205: easily calculated for the Gaussian theory~(\ref{A_sm}),
206: %
207: \begin{equation}
208: S({\bf q}, \omega) \simeq \frac{\hbar}{m}\, {\rm Im}\,
209: \frac{q_y^2 Q(q)}
210: {Q(q) \omega_p^2 (q_y / q)^2 - \omega^2 \omega_c^2 - i \omega \delta},
211: \label{Structure_smectic}
212: \end{equation}
213: %
214: coincides up to the Bose factor and dissipative terms with the finite
215: temperature result of Ref.~\cite{Fogler_00}. The notations used here are
216: $Q({\bf q}) = (\tilde{Y} q_x^2 + K q_y^4) / m n_0$, $\tilde{Y} = Y - C^2
217: / U$, and $\omega_p(q) = [n_0 U(q) q^2 / m]^{1/2}$. On the
218: $\omega > 0$ side, $S({\bf q}, \omega)$ consists of a single
219: $\delta$-function at the frequency of the magnetophonon mode,
220: %
221: \begin{equation}
222: \omega({\bf q}) = \frac{q_y}{q}\frac{\omega_p}{\omega_c}
223: \sqrt{Q({\bf q})}.
224: \label{omega_smectic}
225: \end{equation}
226: %
227:
228: {\it Dislocations and duality\/}.--- 2D smectics can exist only at zero
229: temperature. At $T > 0$ thermal fluctuations of the stripes restore the
230: translational symmetry, so that the highest possible degree of ordering
231: is that of the nematic~\cite{DeGennes_book}. The actual crossover of the
232: Hamiltonian from the short-distance smectic~(\ref{H}) to the
233: long-distance nematic form is quite nontrivial. It is driven
234: by thermally excited dislocations, which have an ability to screen the
235: compressional stress~\cite{Toner_81}. It is natural to assume then that
236: the smectic-nematic {\it quantum\/} phase transition is also driven by
237: topological defects. Pictorially, the difference between the smectic and
238: nematic can be represented as follows. The dislocations are viewed as
239: lines in the (2 + 1)D space. In the smectic phase, they form small
240: closed loops (Fig.~\ref{Fig_worldlines}a) that depict virtual pair
241: creation-annihilation events; in the nematic phase arbitrarily long
242: dislocation worldlines exist and may entangle
243: (Fig.~\ref{Fig_worldlines}b), similar to worldlines of particles in a
244: Bose superfluid~\cite{Feynman_book}.
245:
246: Now I will present a mathematical formalism supporting
247: these qualitative considerations. It is analogous to
248: the duality transformation introduced in Ref.~\cite{Fisher_89}.
249:
250: %
251: % FIG. 2
252: %
253: \begin{figure}
254: \includegraphics[width=2.1in,bb=122 265 568 513]{qn_2.ps}
255: \vspace{0.1in}
256: \setlength{\columnwidth}{3.2in}
257: \caption{
258: Worldlines of dislocations in (a) smectic (b) nematic.
259: \label{Fig_worldlines}
260: }
261: \end{figure}
262: %
263:
264: Our first step is to incorporate the dislocations into the effective
265: action~(\ref{A_sm}). This is accomplished by factorizing the smectic
266: order parameter, $\Psi = \Psi_0 e^{i q_0 u} \times \Psi_D$, where $u$ is
267: a regular single-valued function and $\Psi_D$ is the phase factor due to
268: the dislocations. The derivatives of $u$ in Eq.~(\ref{A_sm}) should now
269: be replaced by $\partial_\mu u - (i / q_0) \Psi_D^* \partial_\mu \Psi_D
270: \equiv (\partial_\mu u)_{\rm tot}$, $\mu = \tau, x, y$. Decoupling the
271: quadratic part of the action with a Hubbard-Stratanovich field
272: $\sigma_\mu$, I get
273: %
274: \begin{eqnarray}
275: A &=& A_D + \int\limits_0^\beta d \tau \int d^2 {\bf r}
276: [-i (\partial_\mu u)_{\rm tot} \sigma_\mu - H_a],
277: \label{A_Hubbard}
278: \\
279: H_a &=& -i \frac{C}{Y} l^2 \sigma_x \partial_y \sigma_\tau
280: + \frac{\sigma_x^2}{2 Y}
281: + \sigma_y (-2 K \partial_y^2)^{-1} \sigma_y
282: \nonumber\\
283: \mbox{} &+& \frac{l^4}{2} \partial_y \sigma_\tau
284: \left({\bf U} - \frac{C^2}{Y}\right)
285: \partial_y \sigma_\tau,
286: \label{H_a}
287: \end{eqnarray}
288: %
289: where $\hbar = 1$, $\sigma_\tau \equiv p / l^2$, and $A_D$ contains
290: terms describing dislocation cores (see below). Integration over $u$
291: gives the constraint $\partial_\mu \sigma_\mu = 0$, which I implement by
292: means of an auxillary $U(1)$ gauge field $a_\mu$,
293: %
294: \begin{equation}
295: \sigma_\mu = \epsilon_{\mu\nu\lambda}
296: \partial_\nu a_\lambda \equiv [\partial \times a]_\mu.
297: \label{a}
298: \end{equation}
299: %
300: This leads to the (dual) action of the form
301: %
302: \begin{equation}
303: A_{\rm dual} = A_D + \int\limits_0^\beta d \tau \int d^2 {\bf r}
304: (-i \Lambda a_\mu j_\mu + H_a),
305: \label{A_dual_first}
306: \end{equation}
307: %
308: where $\Lambda = 2 \pi / q_0$ and $j_\mu = (2 \pi i)^{-1}
309: \epsilon_{\mu\nu\lambda} \partial_\nu (\Psi_D^* \partial_\lambda
310: \Psi_D)$ is the dislocation 3-current. This current can be expressed in
311: terms of the second-quantized dislocation field $\Phi$,
312: %
313: \begin{equation}
314: j_\mu = t_\mu \Phi^* (-i \partial_\mu + \Lambda a_\mu) \Phi,
315: \label{j_mu}
316: \end{equation}
317: %
318: which is assumed to be bosonic. Other types of quantum statistics
319: for dislocations are exotic alternatives, which are
320: not investigated here.
321:
322: Making standard assumptions about $A_D$, I arrive at
323: %
324: \begin{eqnarray}
325: A_{\rm dual} &=& \int\limits_0^\beta d \tau \int d^2 {\bf r}
326: \biggl\{\frac{t_\mu}{2}|(-i \partial_\mu - \Lambda a_\mu
327: - e_D a_\mu^{\rm ext})
328: \Phi|^2\biggr.
329: \nonumber\\
330: \mbox{} &+& \biggl.V(\Phi) + H_a[a]\biggr\}.
331: \label{A_dual}
332: \end{eqnarray}
333: %
334: The obtained dual theory, Eqs.~(\ref{H_a}), (\ref{a}), and
335: (\ref{A_dual}) describes $\Phi$ bosons interacting with each other and
336: with the $U(1)$ gauge field $a_\mu$. The phenomenological parameters
337: introduced above are as follows. Parameter $t_\tau \sim \hbar^2 / E_c$,
338: where $E_c$ is the dislocation core energy. It was estimated within the
339: Hartree-Fock approximation in Refs.~\cite{Fogler_96} and
340: \cite{Wexler_xxx}. Near $\nu \sim \nu_c$ where the interstripe
341: separation $\Lambda$ approaches the quantum fluctuation scale $l$, $E_c$
342: is expected to be much smaller than the Hartree-Fock value, but any
343: precise estimate is challenging. Parameter $t_x$ of dimension of ${\rm
344: energy} \times {\rm (length)}^2$ is the hopping matrix element for
345: dislocation motion in the $\hat{\bf x}$-direction, i.e., dislocation
346: {\it glide\/}. Such a glide requires quantum tunneling and is
347: exponentially small unless $\Lambda \lesssim l$. Parameter $t_y$
348: describes the dislocation climb, which also originates from the dynamics
349: on the microscopic length scales. One may recall that the climb requires
350: mobile point defects. Although such defects are not among fundamental
351: low-energy excitations of the theory, composite point defects, in the
352: form of short stripe segments or dislocation pairs, do exist. They are
353: most likely short lived because dislocations are not expected to form
354: bound pairs in the nematic phase. Yet during their limited life span
355: such composite defects can perfectly well assist the climb by
356: micromotions of constituent electrons. In general, I am not aware of any
357: fundamental principle that would protect $t_y = 0$ value, although it
358: seems reasonable that $t_y \ll t_x$ in our case.
359:
360: Yet another phenomenological variable is the potential $V(\Phi) = m_\Phi
361: |\Phi|^2 + r_\Phi |\Phi|^4 + \ldots$, which accounts for a self-energy
362: and a short-range interaction between the dislocations; the scales of
363: $m_\Phi$ and $r_\Phi$ are set by $E_c$ and $E_c \Lambda^2$,
364: respectively. Finally, $e_D$ is electric charge of the dislocation that
365: couples to the external vector potential $a_\tau^{\rm ext} = a_x^{\rm
366: ext} = 0$, $a_y^{\rm ext} = B x$. This coupling is introduced only for
367: the sake of generality. Since I study electron liquid crystal phases
368: derived from incompressible liquids, I expect dislocations to be
369: electrically neutral, i.e., $e_D = 0$.
370:
371: Another few comments are in order. The derived theory is meant to
372: capture only the dynamics of neutral (dipolar) excitations of the
373: system. The underlying incompressible state has its own dynamics
374: characterized, most importantly, by the quantized Hall conductivity.
375:
376: There is a very strong similarity between our theory and the statistical
377: mechanics model of the 3D smectic-nematic transition studied
378: by Toner and others~\cite{Toner_82}.
379:
380: Integrating out the gauge field from $A_{\rm dual}$ leads to the model
381: of dislocation lines interacting via an effective Biot-Savart potential
382: (which in our case is short-range for Coulombic
383: $U$~\cite{Fogler_in_preparation}). One may argue~\cite{Helfrich_78} that
384: the unlimited expansion of the dislocation loops occurs when the energy
385: cost for creating a large loop of length $L$ is compensated by the
386: ``entropic gain'' $L / l_0$, where the persistence length $l_0$ is
387: determined by short-range physics. Despite the physical appeal of this
388: argument, the gauge theory~(\ref{A_dual}) is presumably better suited
389: for a quantitative analysis (perhaps, along the lines of
390: Ref.~\cite{Toner_82}).
391:
392: {\it Phases and their collective modes\/}.--- Let us now see how the
393: smectic and nematic states are reproduced in the dual theory. As
394: discussed above, the smectic phase corresponds to $\langle \Phi \rangle
395: = 0$. In this case $A_{\rm dual}$ reduces to $H_a$, which is quadratic.
396: The low-energy dynamics is that of a gas of noninteracting Goldstone
397: bosons, which are the aforementioned magnetophons. It is a simple
398: exercise to verify that their dispersion relation is given by
399: Eq.~(\ref{omega_smectic}).
400:
401: In the nematic phase dislocations have condensed, $\langle \Phi \rangle
402: = \Phi_0 \neq 0$. In conventional local $U(1)$ gauge theories,
403: the appearance of such an order parameter triggers the Anderson-Higgs
404: mechanism, eliminating the gapless Goldstone modes. This can not be the
405: case here because the nematic state does break the continuous symmetry
406: with respect to spatial rotations. The seeming paradox is resolved due
407: to the peculiar feature of the present gauge theory: the nonlocality of
408: the gauge-field strength term $H_a$, see Eq.~(\ref{H_a}). By virtue of
409: that, the condensation of $\Phi$ merely stiffens the collective mode,
410: leaving it gapless. Neglecting terms proportional to $C$, I find that
411: %
412: \begin{equation}
413: \omega_1({\bf q}) =
414: \left(\frac{m_x}{m_\tau} q_x^2 + m_x K q_y^2\right)^{1/2},
415: \:\: m_\mu \equiv t_\mu \Lambda^2 |\Phi_0|^2.
416: \label{omega_nematic_1}
417: \end{equation}
418: %
419: As for the magnetophonon mode of the smectic~(\ref{omega_smectic}), it
420: indeed acquires a small gap $\sqrt{m_y Y}$ at $q = 0$. It anti-crosses
421: with the acoustic branch~(\ref{omega_nematic_1}) near the point
422: $\omega_1^2({\bf q}) \sim m_y Y$, and at larger $q$ becomes the lowest
423: frequency collective mode with the dispersion relation
424: %
425: \begin{equation}
426: \omega_2({\bf q}) = \left[\frac{q^2_x q^2_y}{m^2 \omega_c^2} Y U(q)
427: + m_y Y\right]^{1/2}
428: \label{omega_nematic_2}
429: \end{equation}
430: %
431: only slightly different from~(\ref{omega_smectic}). At such $q$ the
432: structure factor of the nematic has two sets of $\delta$-functional
433: peaks,
434: %
435: %\begin{equation}
436: \[
437: S(\omega, {\bf q}) = \frac{\pi \hbar q_y^2}{m \omega_c^2}
438: \left[\frac{K q_y^4}{m n_0} \delta(\omega^2 - \omega_1^2) +
439: \frac{Y q_x^2}{m n_0} \delta(\omega^2 - \omega_2^2)\right],
440: \]
441: %\label{Structure_nematic}
442: %\end{equation}
443: %
444: which split between themselves the spectral weight of the single
445: collective mode of the smectic. The presence of the two modes can be
446: explained by the existence of two order parameters: a unit vector (more
447: precisely, director) ${\bf N}$ normal to the local stripe orientation
448: and the complex wavefunction $\Phi_0$ of the dislocation condensate.
449: Classical 2D nematics have two (overdamped) modes virtually for the same
450: reason~\cite{Toner_81}.
451:
452: In conclusion, let us address measurable properties of the novel quantum
453: Hall states considered here. At low temperature both the parent
454: incompressible state and its liquid crystal descendants will be
455: insulating. If $T$ is not too small, formation of stripe superstructures
456: can be deduced from the anisotropic
457: magnetoresistance~\cite{Experiments}. On the other hand, the microwave
458: absorption will be anisotropic even at low $T$ and would enable one to
459: further distinguish between the smectic and nematic phases: the nematic
460: will show two dispersing collective modes while the smectic will produce
461: a single one. To circumvent disorder pinning effects, such measurements
462: should be done at high enough $q$.
463:
464: {\it Acknowledgements}---This work is supported by MIT Pappalardo
465: Program in Physics. I wish to thank X.-G. Wen for useful discussions,
466: and also A.~Dorsey, L.~Radzihovsky, and C.~Wexler for valuable comments
467: on the manuscript.
468:
469: \begin{thebibliography}{29}
470:
471: \bibitem{Seul_95} M.~Seul and D.~Andelman,
472: Science, {\bf 267}, 476 (1995).
473:
474: \bibitem{Neutron_star} S.~Reddy, G.~Bertsch, and M.~Prakash,
475: Phys.\ Lett.\ B {\bf 475}, 1 (2000).
476:
477: \bibitem{Chklovskii} D.~B.~Chklovskii and A.~A.~Koulakov,
478: Physica A {\bf 284}, 318 (2000).
479:
480: \bibitem{High_Tc_stripes}
481: %J.~M.~Tranquada {\it et. al.\/},
482: %\prl {\bf 73}, 1003 (1994);
483: J.~M.~Tranquada {\it et. al.\/},
484: Nature {\bf 375}, 561 (1995);
485: S.~Mori, C.~H.~Chen, and S.-W.~Cheong,
486: Nature {\bf 392}, 473 (1998).
487:
488: \bibitem{Fogler_96} A.~A.~Koulakov, M.~M.~Fogler, and
489: B.~I.~Shklovskii,
490: \prl {\bf 76}, 499 (1996);
491: M.~M.~Fogler, A.~A.~Koulakov, and
492: B.~I.~Shklovskii,
493: \prb {\bf 54}, 1853 (1996).
494:
495: \bibitem{Moessner_96} R.~Moessner and J.~T.~Chalker,
496: \prb {\bf 54}, 5006 (1996).
497:
498: \bibitem{Critical_nu} M.~M.~Fogler and A.~A.~Koulakov,
499: \prb {\bf 55}, 9326 (1997);
500: E.~H.~Rezayi, F.~D.~M.~Haldane, and K.~Yang,
501: \prl {\bf 83}, 1219 (1999)
502: N.~Shibata and D.~Yoshioka,
503: cond-mat/0101401.
504:
505: \bibitem{Scarola_01} V.~W.~Scarola, K.~Park, J.~K.~Jain,
506: \prb {\bf 62}, R16259 (2001).
507:
508: \bibitem{Experiments} M.~P.~Lilly, K.~B.~Cooper, J.~P.~Eisenstein,
509: L.~N.~Pfeiffer, and K.~W.~West,
510: \prl {\bf 82}, 394 (1999);
511: R.~R.~Du, D.~C.~Tsui, H.~L.~St\"ormer, L.~N.~Pfeiffer, and K.~W.~West,
512: Solid\ State\ Commun. {\bf 109}, 389 (1999).
513:
514: \bibitem{Balents_96} L.~Balents,
515: Europhys.\ Lett. {\bf 33}, 291 (1996).
516:
517: \bibitem{Fradkin_99} E.~Fradkin and S.~A.~Kivelson,
518: \prb {\bf 59}, 8065 (1999).
519:
520: \bibitem{Yi_01} H.~Yi, H.~A.~Fertig, and R.~C\^ot\'e,
521: \prl (2001).
522:
523: \bibitem{MacDonald_00} A.~H.~MacDonald and M.~P.~A.~Fisher,
524: \prb {\bf 61}, 5724 (2000).
525:
526: \bibitem{Fogler_in_preparation} M.~M.~Fogler, in preparation.
527:
528: \bibitem{Musaelian_96} K.~Musaelian and R.~Joynt,
529: J.\ Phys.\ Cond.\ Mat. {\bf 8}, L105 (1996).
530:
531: \bibitem{Comment_on_1st_order} A finite-size study by Rezayi {\it et
532: al.\/}~\cite{Critical_nu} suggests that the transition from the stripe
533: phase to a uniform state as a function of the interaction parameters can
534: also occur via a first-order transition without the intermediate nematic
535: phase.
536:
537: \bibitem{Comment_on_soft_modes} I am interested in a regime
538: with very strong quantum fluctuations where the smectic order parameter
539: is small and only the main harmonic of the soft mode is
540: important.
541:
542: \bibitem{Comment_on_compressible} If the parent state is compressible,
543: the additional low-energy degrees of freedom would generate an effective
544: viscosity of Landau damping (or Caldeira-Leggett) type entering
545: Eq.~(\ref{H}). Its consequences are left for future study.
546:
547: \bibitem{Comment_on_commutator} In principle, $\delta n = -\partial_y p
548: + f(u)$ with some function $f$. In the absence of dislocations the most
549: relevant term is $f = A \partial_x u$, which leads again to
550: Eq.~(\ref{A_sm}). If dislocations are allowed, the situation is more
551: complicated, see later in the main text.
552:
553: \bibitem{DeGennes_book} P.~G.~de~Gennes and J.~Prost,
554: {\it The Physics of Liquid Crystals\/} (Oxford University Press, New York,
555: 1995).
556:
557: \bibitem{Fogler_00} M.~M.~Fogler and V.~M.~Vinokur,
558: \prl {\bf 84}, 5828 (2000).
559:
560: \bibitem{Girvin_86} S.~M.~Girvin, A.~H.~MacDonald, and P.~M.~Platzman,
561: \prb {\bf 33}, 2481 (1986).
562:
563: \bibitem{Toner_81} J.~Toner and D.~R.~Nelson,
564: \prb {\bf 23}, 316 (1981). For application of this
565: theory to the quantum Hall stripes, see Refs.~\cite{Fradkin_99}
566: and \cite{Wexler_xxx}.
567:
568: \bibitem{Wexler_xxx} C.~Wexler and A.~T.~Dorsey,
569: cond-mat/0009096.
570:
571: \bibitem{Feynman_book} R.~P.~Feynman and A.~R.~Hibbs,
572: {\it Quantum Mechanics and Path Integrals\/} (McGraw-Hill, New York,
573: 1965).
574:
575: \bibitem{Fisher_89} M.~P.~A.~Fisher and D.~H.~Lee,
576: \prb {\bf 39}, 2756 (1989).
577:
578: \bibitem{Toner_82} J.~Toner,
579: \prb {\bf 26}, 462 (1982);
580: A.~R.~Day, T.~C.~Lubensky, and A.~J.~McKane,
581: \pra {\bf 27}, 1461 (1983).
582:
583: \bibitem{Helfrich_78} W.~Helfrich,
584: J.\ Phys.\ (Paris) {\bf 39}, 1199 (1978).
585:
586: \end{thebibliography}
587: \end{document}
588:
589: ------------------------
590:
591: An example of the quantum nematic is the state
592: with the wavefunction~\cite{Musaelian_96}
593: %
594: \begin{equation}
595: \Psi = \prod\limits_{j < k} z_{j k} (z_{j k}^2 - a^2 l^2)
596: \exp \left(-\sum_j |z_j|^2 / 4 l_B^2\right),
597: \label{nematic_trial_wavefunction}
598: \end{equation}
599: %
600: where $z_{j k} = z_j - z_k$, $z_j$ is the complex coordinate of $j$th
601: electron, $z_j = x_j + i y_j$, $l_B$ is the magnetic length, and $a$ is
602: a complex parameter that determines the degree of orientational order
603: and the direction of the stripes (the rotational invariance is broken if
604: $|a|$ exceeds some critical value). This particular wavefunction
605: corresponds to $\nu = \frac13$ but its generalization to higher Landau
606: levels is straightforward.
607:
608: