cond-mat0703261/DBL.tex
1: %\documentclass[prb,preprint]{revtex4}
2: \documentclass[prb,twocolumn]{revtex4}
3: \usepackage[dvips]{graphicx}
4: \usepackage{latexsym,amsmath,amssymb,bm,euscript}
5: \usepackage[dvips]{color}
6: 
7: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8: %%% some defs
9: \def\etal{~\textit{et~al.}} % etal
10: \def\ra{\rangle} % bra
11: \def\la{\langle} % ket
12: \def\up{\uparrow}
13: \def\dn{\downarrow}
14: \def\Hc{{\rm H.c.}}
15: 
16: 
17: \begin{document}
18: 
19: \title{D-wave correlated Critical Bose Liquids in two dimensions}
20: 
21: \author{Olexei I. Motrunich}
22: \affiliation{Department of Physics, California Institute of Technology,
23: Pasadena, CA 91125}
24: \author{Matthew P. A. Fisher}
25: \affiliation{Kavli Institute for Theoretical Physics, 
26: University of California, Santa Barbara, CA 93106}
27: 
28: \date{March 9, 2007}
29: 
30: \begin{abstract}
31: 
32: We develop a description of a new quantum liquid phase of interacting 
33: bosons confined in two dimensions which possesses relative D-wave 
34: two-body correlations.  We refer to this stable quantum phase as a 
35: D-wave Bose Liquid (DBL).  The DBL has no broken symmetries, 
36: supports gapless boson excitations which reside on ``Bose surfaces" 
37: in momentum space, and exhibits power law correlation functions 
38: characterized by a manifold of continuously variable exponents.
39: While the DBL can be constructed for bosons moving in the 2d continuum,
40: the state only respects the point group symmetries of the square lattice.
41: On the square lattice the DBL respects all symmetries and does not 
42: require a particular lattice filling.  But lattice effects do allow for 
43: the possibility of a second distinct phase, a quasi-local variant which 
44: we refer to as a D-wave Local Bose Liquid (DLBL).
45: Remarkably, the DLBL has short-range boson correlations and hence no 
46: Bose surfaces, despite sharing gapless excitations and other critical 
47: signatures with the DBL.  Moreover, both phases are metals with a 
48: resistance that vanishes as a power of the temperature.
49: We establish these results by constructing a class of many-particle 
50: wavefunctions for the DBL, which are time reversal invariant analogs of 
51: Laughlin's quantum Hall wavefunction for bosons in a half-filled 
52: Landau level.
53: A gauge theory formulation leads to a simple mean field theory and a 
54: suitable $N-$flavor generalization enables incorporation of gauge field
55: fluctuations to deduce the properties of the DBL/DLBL in a controlled 
56: and systematic fashion.
57: Various equal time correlation functions thereby obtained are in 
58: qualitative accord with the properties inferred from the variational 
59: wavefunctions.  
60: We also identify a promising microscopic Hamiltonian which might manifest
61: the DBL or DLBL, and perform a variational energetics study comparing to 
62: other competing phases including the superfluid.
63: We suggest how the D-wave Bose Liquid wavefunction can be suitably 
64: generalized to describe an itinerant non-Fermi liquid phase of electrons 
65: on the square lattice with a no double occupancy constraint, 
66: a D-wave metal phase.
67: 
68: \end{abstract}
69: 
70: \maketitle
71: 
72: 
73: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
74: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
75: \section{Introduction}
76: \label{sec:intro}
77: 
78: The principal roadblock impeding progress in disentangling the physics 
79: of the cuprate superconductors is arguably our inability to access 
80: quantum ground states of two-dimensional itinerant electrons which are 
81: qualitatively distinct from a Landau Fermi liquid.
82: Overcoming this obstruction is of paramount importance, indispensable 
83: in explaining the strange metal behavior observed near optimal doping 
84: and a likely requisite to account for the emergent pseudo-gap at lower 
85: energies.
86: A putative underlying paramagnetic Mott insulator provides the 
87: scaffolding for a popular class of theories, which view the pseudo-gap 
88: as a lightly doped spin liquid.
89: \cite{Anderson, Baskaran, KotliarLiu, IoffeLarkin, LeeNagaosaWen}  
90: Significant progress has been made in developing the groundwork and 
91: there now exists a well established theoretical framework to describe a 
92: myriad of distinct spin liquids.  For the cuprates the most promising
93: spin liquids are described in terms of fermionic spinons minimally 
94: coupled to a compact $U(1)$ gauge field and moving in various 
95: background fluxes.
96: Upon doping formidable challenges arise.  Bosonic holons carrying the 
97: electron charge become mobile carriers and lead to electrical conduction.
98: But at low temperatures Bose condensation appears inevitable and this 
99: leads to Fermi liquid behavior, either a metal with conventional Landau 
100: quasiparticles or a  BCS superconductor if the spinons are paired.
101: Accessing a pseudo-gap or a strange metal which conduct electricity 
102: despite the absence of long-lived Landau quasiparticles requires doped 
103: holons that form an uncondensed quantum Bose fluid rather than a 
104: condensed superfluid.  But is this possible, even in principle?
105: If possible, what properties would such a putative ``Bose metal'' 
106: exhibit?\cite{Feigelman, Dalidovich, Galitski, Alicea}
107: And what theoretical framework is appropriate?
108: The slave-particle gauge theory approach has been stymied by this 
109: stumbling block for over 15 years.  
110: In this paper we provide (some) answers to these questions by 
111: constructing explicit examples of such unusual phases of bosons
112: that may offer some routes out of the conundrum.
113: 
114: Our goal, then, is to access and explore uncondensed quantum phases of
115: 2d bosons which are conducting fluids but not superfluids.
116: Specifically we have in mind hard core bosons moving on a 2d square 
117: lattice, but seek to construct states which do not require particular 
118: commensurate densities.  While our construction can be implemented for 
119: bosons moving in the 2d continuum Euclidean plane, the states will only 
120: possess the reduced point group symmetry of the square lattice.  
121: Despite our interest in time reversal invariant quantum ground states, 
122: our technical approach will be strongly informed by theories of the 
123: fractional quantum Hall effect (FQHE).  In some regards, the quantum 
124: phases that we construct are time reversal invariant analogs of the 
125: Laughlin state for bosons in a half-filled Landau level.  
126: But the physical properties of the phases will be dramatically different 
127: from the incompressible FQHE states, and will have gapless excitations 
128: and metallic transport for example.
129: 
130: To motivate and illustrate our approach, it will be helpful to briefly 
131: revisit the bosonic FQHE.  Consider the Laughlin wavefunction for bosons 
132: in a half-filled Landau level,
133: \begin{equation}
134: \Psi_{\nu = 1/2} (z_1, z_2, \dots, z_N) = \prod_{i<j} (z_i - z_j)^2 ~.
135: \label{Psi_nu_1/2}
136: \end{equation}
137: Much of the physics of the Laughlin state has its origin in the structure
138: of zeroes, which reveals that any two particles upon close approach in 
139: real space are in a relative two-body $d+id$ state, $(z_i-z_j)^2$.  
140: Our first objective is to construct a time reversal invariant (real) 
141: wavefunction in which particle pairs are similarly in a relative 
142: $d_{xy}$ state.  
143: A clue is offered by noting that the Laughlin wavefunction is the square 
144: of a Vandermonde determinant, $\Psi_{\nu = 1/2} = (\det {\bf V})^2$.  
145: In the Vandermonde determinant, all pairs of particles are in a 
146: relative $p+ip$ state, $(z_i-z_j)$.  Upon squaring, the two $p+ip$ states
147: combine to form a single $d+id$ state, which is essentially just 
148: addition of angular momentum.
149: 
150: This suggests constructing a time reversal invariant boson wavefunction 
151: in zero magnetic field by simply squaring a determinant constructed from 
152: momentum states within a Fermi sea,
153: \begin{equation}
154: \Psi({\bf r}_1, {\bf r}_2, \dots, {\bf r}_N) 
155: = ( \det e^{i {\bf k}_i \cdot {\bf r}_j} )^2 ~, \quad
156: \text{(S-type)}.
157: \label{Psi_det-squared}
158: \end{equation}
159: Let's examine the nodal structure.  The generic behavior of each 
160: fermion determinant when any two particles are taken close together 
161: is dictated by Fermi statistics and reality of the wavefunction
162: and has the functional form $({\bf r}_i -{\bf r}_j) \cdot \hat{\bm l}$.
163: The unit vector $\hat{\bm l}$ will depend in a complicated way on the 
164: location of all the other particles 
165: (see Ref.~\onlinecite{Ceperley91} for a discussion and illustrations of 
166: the free fermion nodes).
167: When $\hat{\bm l} = \hat{\bf y}$, this is a $p_x$ form vanishing along a 
168: nodal line (in the relative coordinate) parallel to the $x$-axis.
169: Unfortunately, in contrast with the Laughlin case, 
170: squaring the determinant leads here to an ``extended" $s$-wave form 
171: with a quadratic nodal line rather than the desired $d$-wave.
172: 
173: But consider instead multiplying together two fermion determinants, 
174: each constructed by filling up a Fermi sea of momentum states,
175: but with different Fermi surfaces.  
176: A wavefunction with $d_{xy}$ two-particle correlations
177: can be constructed by choosing two elliptical Fermi surfaces, one with 
178: its long axis along the $x$-axis and the other rotated by 90 degrees,
179: as illustrated in Fig.~\ref{fig:introFS},
180: \begin{equation}
181: \Psi({\bf r}_1, {\bf r}_2, \dots, {\bf r}_N) = 
182: (\det)_x \times (\det)_y ~, \quad \text{(D$_{xy}$-type)} ~,
183: \label{Psi_DBL}
184: \end{equation}
185: where the short-hands $(\det)_x$ and $(\det)_y$ represent the 
186: corresponding Slater determinants.
187: In the limit of extreme eccentricity of the ellipses, the wavefunction 
188: will have the desired $d_{xy}$ form when two particles are brought 
189: close together, $(x_i-x_j)(y_i-y_j)$.  Away from this limit, the two 
190: nodal lines will not align precisely with the $x$ and $y$ axes,
191: but upon taking one particle around the other the wavefunction will 
192: exhibit the same sign structure as a $d_{xy}$ orbital, $+ - + -$,
193: changing sign twice. 
194: A picture of such sign structure as seen by a test particle is shown in 
195: Fig.~\ref{fig:nodes}.
196: 
197: 
198: \begin{figure}
199: \centerline{\includegraphics[width=\columnwidth]{figures/introFS.eps}}
200: \vskip -2mm
201: \caption{DBL wavefunction Eq.~(\ref{Psi_DBL}) is obtained by multiplying 
202: two Slater determinants corresponding to distinct but symmetry related
203: Fermi surfaces, which can be viewed as Fermi surfaces of $d_1$ and 
204: $d_2$ slave fermions.
205: The left panel shows an example with closed elliptical Fermi surfaces, 
206: while the right panel is a case with open Fermi surfaces in the first 
207: Brillouin zone of the square lattice.
208: }
209: \label{fig:introFS}
210: \end{figure}
211: 
212: 
213: \begin{figure}
214: \centerline{\includegraphics[width=3.0in]{figures/nodes.eps}}
215: \vskip -2mm
216: \caption{``Nodal picture'' of the continuum DBL wavefunction as probed 
217: by moving a single particle (whose initial location is marked by the 
218: filled circle) while keeping the rest of the particles fixed
219: (open circles).
220: The signs of the wavefunction are indicated in several places to 
221: bring out the $+ - + -$ pattern upon encircling a target particle.
222: }
223: \label{fig:nodes}
224: \end{figure}
225: 
226: 
227: Our principal thesis is that this wavefunction captures qualitative 
228: features of a new quantum liquid phase of bosons, which we call
229: ``D-wave-correlated Bose Liquid'' (DBL).
230: But a variational wavefunction does not provide a complete 
231: characterization of a quantum phase, and cannot be used to address its 
232: stability.  As for the FQHE, a field theoretic approach, such as 
233: Chern-Simons gauge theory, is both desirable and ultimately necessary.
234: Since we require time reversal invariance, Chern-Simons is inappropriate,
235: but a tractable field theoretic framework for the DBL will nevertheless
236: turn out to be a gauge theory.
237: 
238: Indeed, we follow closely the gauge theory approaches to spin liquids in 
239: quantum antiferromagnets,\cite{LeeNagaosaWen, WenPSG} which after all are 
240: lattice bosonic systems.  But there are some notable and important differences 
241: for the DBL which we discuss below.  The variational wavefunction in 
242: Eq.~(\ref{Psi_DBL}), being a product of two fermion determinants, naturally 
243: suggests expressing the boson creation operator as a product of two fermion 
244: operators,
245: \begin{equation}
246: b^\dagger ({\bf r}) = d_1^\dagger ({\bf r}) d_2^\dagger ({\bf r}) ~,
247: \end{equation}
248: with the 2d position ${\bf r}$ either continuous or denoting the 
249: discrete sites of a square lattice.
250: In a general case, this decomposition introduces a local $SU(2)$ gauge 
251: redundancy well-known in the slave-fermion treatments of the 
252: spin-1/2 antiferromagnet 
253: (see Ref.~\onlinecite{WenPSG} for a recent comprehensive discussion).
254: The S-type wavefunction corresponds to a so-called $SU(2)$ liquid,
255: while the DBL wavefunctions are $U(1)$ liquids and require a $U(1)$ 
256: gauge field which is minimally coupled to the two fermions with 
257: opposite gauge charges (the product $d_1^\dagger d_2^\dagger$
258: is then a gauge neutral composite, the physical boson).  
259: 
260: Within slave particle theory a mean field state with a fixed background 
261: gauge ``magnetic" flux is chosen.
262: This choice is constrained by symmetry:  All the physical symmetries 
263: of the boson model, even if not respected by the mean field ansatz, 
264: must be present in the full gauge theory once gauge transformations are 
265: allowed.
266: The collection of these symmetry transformations in the gauge theory
267: is called the Projective Symmetry Group (PSG).\cite{WenPSG}  
268: For lattice boson theories many different mean field states with
269: different PSG's are possible, while little is offered to guide
270: which mean field state is the correct one.
271: 
272: In the continuum, the symmetries of the Euclidean plane together with 
273: time reversal invariance are more restrictive and we can only offer 
274: two PSG mean field ansatze in this case.
275: One is a mean field state with zero gauge flux and spherical Fermi seas
276: for both fermions; this is an $SU(2)$ ansatz and corresponds to the 
277: S-type wavefunction considered earlier.
278: A different $U(1)$ ansatz is obtained by considering $d_1$ and $d_2$ 
279: fermions each moving in a uniform field but of opposite signs for the 
280: two; in this case, the boson wavefunction is schematically
281: $\Psi = (\det) \times (\det)^*$,
282: e.g., $\Psi = |\det {\bf V}|^2 = |\Psi_{\nu=1/2}|$ to give a concrete 
283: example;
284: this wavefunction is again non-negative and has off-diagonal
285: quasi-long-range order.\cite{GirvinMacDonald, Kane}
286: 
287: For bosons moving in continuous space but with only the square lattice 
288: point group symmetry, we can offer only four PSG mean fields; 
289: all of them have zero flux, but differ in the way the 90 degree rotation 
290: and mirror symmetries are realized.
291: One example has both $d_1$ and $d_2$ Fermi surfaces individually
292: respecting the symmetries but otherwise independent of each other.  
293: Two more examples are the already introduced D$_{xy}$ state,
294: Eq.~(\ref{Psi_DBL}) and Fig.~\ref{fig:introFS}, with elliptical 
295: Fermi surfaces elongated along the $x$ and $y$ axes, 
296: and a similarly constructed D$_{x^2-y^2}$ state with ellipses along 
297: $\hat{\bf x} \pm \hat{\bf y}$;
298: in both cases, the $d_1$ and $d_2$ Fermi surfaces are related by
299: the 90 degree rotation, but the two states differ in the way the mirror 
300: symmetries are realized.
301: In the fourth example, each Fermi surface is invariant under the rotation
302: but not under the mirrors, which instead are realized by interchanging 
303: the two fermions.
304: 
305: Each of the four presented ansatze can be loosely called a D-wave Bose 
306: liquid as far as the nodal pictures like that in Fig.~\ref{fig:nodes}
307: are concerned; the gauge theory analysis would also be rather similar.
308: From now on, we will focus on the D$_{xy}$ boson liquid.
309: The associated mean field wavefunction once Gutzwiller projected into
310: the physical Hilbert space, $d_1^\dagger d_1 = d^\dagger_2 d_2$, 
311: is precisely the variational wavefunction in Eq.~(\ref{Psi_DBL}).
312: A full incorporation of gauge fluctuations about the mean field 
313: performs, in principle, this projection exactly.  But in practice the gauge theory approach allows incorporation of 
314: slowly varying gauge fluxes and is different in this respect from
315: the wavefunction.  
316: Provided the gauge theory is not in a confined state and no physical 
317: symmetries are broken, an exact description of the putative DBL phase 
318: is thereby obtained.  
319: Herein, we introduce a large $N$ generalization of the gauge 
320: theory which allows a controlled and systematic treatment of the gauge 
321: fluctuations order by order in powers of $1/N$.
322: We can then perturbatively extract physical properties of the DBL phase.
323: Whether the DBL phase survives down to $N=1$ is not something that 
324: we can reliably address.
325: 
326: In describing the D$_{xy}$ liquid phase for lattice bosons there is 
327: considerable freedom when choosing the precise form of the slave fermion 
328: hopping in the mean field Hamiltonian.  The simplest choice is near 
329: neighbor hopping with different amplitudes in the $x$ and $y$ directions.
330: But even with this restriction, there are two different possible 
331: Fermi surface topologies, closed or open, as depicted in 
332: Fig.~\ref{fig:introFS}.  
333: The Fermi surface topology has rather dramatic consequences 
334: for the nature of the associated D-wave Bose fluid.
335: When the Fermi surfaces are open and have no parallel Fermi surface 
336: tangents, the resulting phase has a quasi-local character -- 
337: the boson Green's function is found to fall off exponentially in space.
338: This is a distinct quantum phase which we refer to as a 
339: D-Wave Local Bose liquid (DLBL).
340: The DLBL is intrinsically very stable against gauge fluctuations and 
341: we are fairly confident that when present it will be a stable quantum 
342: phase.  The effect of gauge fluctuations on the DBL phase, however, 
343: are more subtle.   Within our systematic large $N$ approach we do 
344: find a regime where the DBL is stable, but it is less clear that 
345: this will survive down to $N=1$.
346: The Gutzwiller wavefunction does appear to describe a DBL phase with 
347: properties that are consistent with those inferred from the gauge theory.
348: This provides some support for the stability of the DBL phase.
349: 
350: Together the gauge theory and the variational wavefunction provide a 
351: consistent and rather complete picture of both the DBL and the DLBL.
352: Here we briefly highlight some of the important characteristics.
353: In the DBL, the single particle boson Green's function, 
354: $G_b({\bf r}, \tau) \equiv \la b^\dagger ({\bf r},\tau) b({\bf 0},0) \ra$,
355: decays as an oscillatory power law at equal times,
356: \begin{eqnarray}
357: G_b({\bf r},0) &\sim & 
358: \frac{\cos[({\bf k}_{F_1} - {\bf k}_{F_2}) \cdot {\bf r}] }
359:      {|{\bf r}|^{4-\eta}} \nonumber \\
360: &+& 
361: \frac{\cos[({\bf k}_{F_1} + {\bf k}_{F_2}) \cdot {\bf r} - 3\pi/2]} 
362:      {|{\bf r}|^4} ~,
363: \label{Gb_final}
364: \end{eqnarray}
365: where the two wave vectors ${\bf k}_{F_{1,2}}$ depend on the observation 
366: direction $\hat{\bf r}$, as does the anomalous exponent $\eta > 0$.
367: As the direction of $\hat{\bf r}$ is rotated in real space, the 
368: wave vectors ${\bf k}_{F_{1,2}}$ and ${\bf k}_{F_1} \pm {\bf k}_{F_2}$ 
369: trace out closed momentum space curves, and for the DBL in the continuum 
370: these curves will have the topology of a circle.
371: 
372: If one measures the boson Green's function and finds it to be of the
373: above form, it is natural to refer to the 
374: ${\bf k}_{F_1} \pm {\bf k}_{F_2}$ as ``Bose surfaces.''
375: Within the gauge theory description the origin of these singular surfaces
376: can be traced to the Fermi surfaces of the constituent fermions.
377: Specifically, ${\bf k}_{F_{1,2}}$ are locations on the two Fermi 
378: surfaces where the surface normals are along the observation 
379: direction $\hat{\bf r}$, and the momentum space areas enclosed by 
380: each of these surfaces will equal $(2\pi)^2 \rho_b$.
381: Note that the ${\bf k}_{F_{1,2}}$ surfaces can be uniquely
382: reconstructed from the measured Bose surfaces 
383: ${\bf k}_{F_1} \pm {\bf k}_{F_2}$.
384: In this sense, the Bose surfaces in the DBL contain information
385: analogous to Luttinger's Fermi surface volume theorem in the 
386: Fermi liquid.  The Bose surface can also be extracted directly 
387: from the DBL wavefunction by using variational Monte Carlo to compute 
388: the equal time boson Green's function, and can also be indirectly 
389: inferred by using the shape of the two Fermi surfaces input into the 
390: $(\det)_x$ and $(\det)_y$ factors.  It is both reassuring and quite 
391: remarkable that these two coincide.
392: 
393: % The spatially local Green's function in the DBL falls off in time as,
394: % \begin{equation}
395: % G_b({\bf 0}, \tau) \sim \frac{1}{|\ln(\tau)|^{1/2}} 
396: % |\tau|^{-(2-3\eta_0/2)} ~.
397: % \end{equation}
398: % Here $\eta_0$ is the maximum value of the direction dependent exponent
399: % $\eta(\hat{\bf r})$.  The local boson tunneling density of states 
400: % vanishes as a power of energy, up to logarithmic factors, 
401: % $A_b(E) \sim E^{1 - 3\eta_0/2}$.
402: 
403: The spatially local Green's function, $G_b({\bf 0}, \tau)$, in the DBL 
404: falls off in time as a power law which is at least as slow as 
405: $1/\tau^2$ corresponding to the local boson tunneling density of states 
406: $A_b(E) \sim E$;
407: these are mean field results obtained by combining two fermions each 
408: with a finite density of states, but we suspect the time decay may
409: actually be slower upon including gauge fluctuations.
410: 
411: 
412: The behavior of the boson Green's function in the DLBL is quite 
413: different.
414: The time decay at ${\bf r}=0$ is a power law, 
415: $G_b({\bf 0}, \tau) \sim \tau^{-2}$,
416: which is an exact result insensitive to lattice scale details.
417: But the equal time Green's function falls off exponentially in space
418: in the DLBL, $G_b({\bf r}, 0) \sim \exp(-r/\xi)$.
419: Despite this, a two boson box correlator,
420: \begin{equation}
421: {\cal B}_b (x) \equiv
422: \la b^\dagger(0,0) b^\dagger(x,x) b(x,0) b(0,x) \ra ~,
423: \label{Bb_def}
424: \end{equation}
425: falls off as a (non-oscillatory) power law at large distances in the 
426: DLBL, ${\cal B}(x) \sim -x^{-8}$, with both the sign (negative) and the 
427: exponent being universal and insensitive to lattice scale physics.
428: Paradoxically, this seems to imply that a pair of bosons injected into 
429: the DLBL on opposite corners of a square box can move more readily 
430: than a single injected boson.
431: But in a strongly interacting quantum state such a single-particle 
432: interpretation can be misleading -- the dynamics of an injected boson 
433: is not that of a weakly interacting quasiparticle.
434: 
435: In both the DBL and the DLBL the density-density correlator
436: $D_b({\bf r}) \equiv
437:  \la \hat{\rho}_b({\bf r}) \hat{\rho}_b({\bf 0}) \ra$ 
438: behaves at large distances as,
439: \begin{equation}
440: D_b({\bf r}) \sim - \sum_{\alpha = 1,2} 
441: \frac{\cos[2{\bf k}_{F_\alpha} \cdot {\bf r} - 3\pi/2]}
442:       {|{\bf r}|^{4-\gamma_\alpha}}
443: + \frac{1}{|{\bf r}|^4} ~,
444: \label{Db_final}
445: \end{equation}
446: where again both the wave vectors ${\bf k}_{F_{1,2}}$ and the anomalous 
447: exponent $\gamma_\alpha$ depend on the observation direction 
448: $\hat{\bf r}$.   Measurement of such correlation can be used to extract 
449: ${\bf k}_{F_{1,2}}$ in the DLBL, which are not accessible from the boson 
450: Green's function here.
451: Upon rotating the unit vector $\hat{\bf r}$ in real space, 
452: ${\bf k}_{F_{1,2}}$ will again trace out the Fermi surfaces of the 
453: underlying fermions which for the DLBL will be open.
454: Both the DBL and the DLBL are conductors with a resistance varying with 
455: temperature as $R \sim T^{4/3}$.  This is particularly striking for the
456: DLBL where the boson Green's function is short-ranged.  
457: Evidently, it is not possible to understand the metallic transport in 
458: terms of the motion of weakly interacting bosons.  In a sense it is the 
459: fermionic constituents which transport the charge, but even this is not 
460: quite right since the $d_1,d_2$ fermions do not exist as well defined
461: quasiparticle excitations, being strongly scattered by the gauge 
462: fluctuations.
463: 
464: The lattice gauge theory formulation can be used to motivate a lattice
465: boson Hamiltonian which might plausibly have the DBL/DLBL as a ground 
466: state.  The simplest such model consists of a near neighbor boson 
467: hopping term supplemented by a four-site ``ring" exchange term:
468: \begin{eqnarray}
469: \label{Hring}
470: H &=& H_J + H_4 ~,\\
471: H_J &=& -J \sum_{{\bf r};\, \hat{\mu} = \hat{\bf x}, \hat{\bf y}}
472: (b^\dagger_{\bf r} b_{{\bf r} + \hat{\mu}} + h.c.) ~,\\
473: H_4 &=& K_4 \sum_{\bf r} 
474: (b^\dagger_{\bf r} b_{{\bf r} + \hat{\bf x}}
475: b^\dagger_{{\bf r} + \hat{\bf x} + \hat{\bf y}} b_{{\bf r} + \hat{\bf y}}
476: + h.c.) ~,
477: \end{eqnarray}
478: with $J,K_4 \ge 0$.  This model is fully specified by two dimensionless 
479: numbers, the ratio $J/K$ and the boson filling $0 \le \rho_b \le 1$.
480: This ring Hamiltonian with $J>0$ and $K_4<0$ was introduced and studied 
481: by Paramekanti\etal,\cite{Paramekanti} and later studied extensively
482: with quantum Monte Carlo by Sandvik\etal,\cite{Sandvik}
483: Melko\etal,\cite{Melko} and Rousseau\etal\cite{Rousseau}
484: There is no sign problem in this regime, and it was possible to
485: access large system sizes and low temperatures.
486: But for $J>0$ and $K_4>0$, there is a sign problem since the Hamiltonian 
487: does not satisfy the Marshall sign conditions, and one expects the 
488: ground state wavefunction to take both positive and negative values.
489: For this ring model we have evaluated the variational energetics
490: for the DBL/DLBL wavefunctions and compared these to the energy of a 
491: superfluid wavefunction of the usual Jastrow form.
492: Within this necessarily limited energetics study we do find a region 
493: of parameter space with small $J/K_4$ and near half filling where the 
494: DLBL wavefunction has the lowest energy.  In view of its quasi-local 
495: character,  we suspect that DMRG studies could be fruitful in helping 
496: establish whether or not the DLBL is present in the phase diagram of 
497: this (or related) ring Hamiltonian.
498: 
499: 
500: The paper is organized as follows.
501: In Sec.~\ref{sec:motiv_Wavefnc} we discuss in more detail the 
502: wavefunction motivation for considering the DBL states.
503: In Sec.~\ref{sec:MFT} we introduce the lattice gauge theory
504: description and study properties of the DBL state in the 
505: mean field theory that ignores gauge fluctuations.
506: This provides an initial guide to the singular (Bose) surfaces,
507: and is followed in Sec.~\ref{sec:Wavefnc_props} with numerical
508: characterization of the properties of the actual DBL wavefunctions.
509: In Sec.~\ref{sec:Gauge_fluct}, we consider the full gauge theory
510: description, focusing on the effects of the gauge fluctuations on
511: the singularities across the Bose surfaces, and in particular
512: obtain the long-distance properties such as
513: Eq.~(\ref{Gb_final}) and Eq.~(\ref{Db_final}) to order $1/N$ in our systematic
514: large $N$ approach.
515: In Sec.~\ref{sec:eff_FT} we address the issue of the stability
516: of the DBL by considering a putative fixed point theory.
517: We conclude in Sec.~\ref{sec:concl} with a discussion of physical
518: properties and possible future directions.
519: 
520: 
521: 
522: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
523: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
524: \section{Boson Wavefunctions and Nodes}
525: \label{sec:motiv_Wavefnc}
526: 
527: In this section, we expand on the initial motivation for considering
528: the DBL states as a way to perform (a kind of) flux attachment 
529: transformation in a time reversal invariant manner.
530: We also discuss the nodal structure of the bosonic wavefunctions,
531: providing some justification for the qualifier ``D-wave'' in the 
532: suggested wavefunction names in this work.  
533: To this end, we consider a ``relative single particle wavefunction"
534: -- more precisely, a cross-section of the many-body wavefunction --
535: defined as follows.
536: Fixing the positions of all the particles except one, we define a 
537: function which depends explicitly on the coordinates of the one 
538: ``test" particle, and implicitly on the coordinates of all the other 
539: particles,
540: \begin{equation}
541: \Phi_{{\bf r}_2, \dots, {\bf r}_N}({\bf r}) \equiv
542: \Psi({\bf r}, {\bf r}_2, \dots, {\bf r}_N) ~.
543: \label{Phi_b}
544: \end{equation}
545: For notational ease we will henceforth drop the implicit dependence 
546: on the $2(N-1)$ spatial coordinates and just use the notation 
547: $\Phi_b({\bf r})$.
548: 
549: 
550: \subsection{Laughlin $\nu=1/2$ revisited}
551: \label{subsec:Laughlin}
552: Consider the Laughlin wavefunction for bosons in a half-filled Landau 
553: level, Eq.~(\ref{Psi_nu_1/2}).  In this case, the relative single
554: particle wavefunction $\Phi(z)$ is complex and has double strength 
555: zeroes at the positions of all the other $N-1$ particles.
556: Upon close approach to a specific particle, say $z_i$, the function
557: $\Phi(z) \sim (z-z_i)^2$ is of a $d+id$ form.  It is in this sense
558: that the $\nu=1/2$ Laughlin state can be viewed as a D-wave fluid:
559: All pairs of particles, upon close approach in real space, are in a 
560: two-body $d+id$ state.
561: Our goal is to construct a time reversal invariant analog of the 
562: $\nu=1/2$ Laughlin state in which particle pairs are similarly in
563: a relative $d_{xy}$ or $d_{x^2-y^2}$ state.
564: 
565: In thinking about Laughlin states, it has been particularly instructive 
566: to view them in terms of composite particles created by 
567: ``flux attachment."
568: For example, if one re-expresses the $\nu=1/2$ Laughlin state as,
569: \begin{equation}
570: \Psi_{\nu=1/2} = \prod_{i<j} (z_i - z_j) \, \Psi_{cf} ~,
571: \end{equation}
572: the ``composite fermion" wavefunction is simply a (Vandermonde) 
573: determinant, $\det {\bf V}$, of filled Landau level orbitals,
574: \begin{equation}
575: \Psi_{cf} = \det {\bf V} = \prod_{i<j} (z_i - z_j) ~,
576: \end{equation}
577: with $V_{ij} = (z_i)^j$.
578: In a second quantized framework, flux attachment can be achieved by 
579: introducing a Chern-Simons gauge field\cite{Zhang_CS}.
580: Chern-Simons gauge theory has been a useful tool to describe the full 
581: Haldane-Halperin hierarchy of fractional quantum Hall states,
582: encoding the fractional charge and statistics of the quasiparticles as
583: well as the structure of the edge excitations\cite{MPAF_Lee, Wen_edge}.  But in discussing bosons
584: in a time reversal invariant setting (zero magnetic field), 
585: flux-attachment techniques are problematic.  For instance, the usual 
586: flux smearing mean field approximation is likely to inadvertently
587: break time reversal invariance.  
588: 
589: In seeking to avoid Chern-Simons theory, it is worth noting that
590: the $\nu=1/2$ state is a perfect square -- a square of the Vandermonde 
591: determinant,
592: \begin{equation}
593: \Psi_{\nu =1/2} = (\det {\bf V})^2 ~.
594: \label{detV2}
595: \end{equation}
596: This suggests that the $\nu=1/2$ Laughlin state can be fruitfully 
597: described within a slave particle framework.\cite{Wen_proj4FQHE}
598: Consider decomposing the boson creation operator as a product of two 
599: fermions,
600: $b^\dagger ({\bf r}) = d_1^\dagger ({\bf r}) d_2^\dagger ({\bf r})$.
601: As discussed in the Introduction, there is a local gauge redundancy,
602: and a correct treatment requires the presence of a gauge field 
603: minimally coupled to both fermions.
604: In the slave particle mean field approach to the $\nu=1/2$ boson problem,
605: one simply drops the gauge field, obtaining a problem of two fermion 
606: flavors each moving in a magnetic field.
607: If the electrical charge of the boson is divided equally, each fermion 
608: flavor is effectively in a full Landau level.
609: The mean field wavefunction is,
610: \begin{equation}
611: \Psi^{MF}_{\nu = 1/2} = \prod_{i<j} (z^{(1)}_i - z^{(1)}_j) \times 
612: \prod_{i<j} (z^{(2)}_i - z^{(2)}_j) ~.
613: \end{equation}
614: To obtain a wavefunction for the bosons it is necessary to project
615: into the physical Hilbert space, and this is achieved here by simply 
616: setting $z^{(1)}_i = z^{(2)}_i = z_i$.
617: One thereby recovers the Laughlin state as a square of the 
618: Vandermonde determinant, Eq.~(\ref{detV2}).
619: 
620: 
621: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
622: \subsection{Time Reversal Invariant Wavefunctions}
623: \label{subsec:det-squared}
624: 
625: Within a first quantized framework, the relative simplicity of a time 
626: reversal invariant bosonic superfluid as compared to the Laughlin state 
627: is the nodelessness of the ground state wavefunction.  
628: As Feynman argued many years ago, for non-relativistic bosons moving 
629: in the continuum with an interaction only depending on the particle 
630: positions, the kinetic energy of any bosonic wavefunction which has 
631: sign changes could be reduced by making all the signs positive while 
632: keeping the magnitude of the wavefunction fixed to leave the potential 
633: energy unchanged.
634: The ground state wavefunction should thus be everywhere non-negative.  
635: For the non-interacting Bose gas the ground state wavefunction is just
636: $\Psi =1$, but in the presence of interactions a popular variational
637: wavefunction is of the Jastrow form,\cite{Kane}
638: \begin{eqnarray}
639: \Psi_{\rm Jastrow}({\bf r}_1, {\bf r}_2, \dots, {\bf r}_N) 
640: \;\propto\; e^{-\sum_{i<j} u({\bf r}_i - {\bf r}_j)} ~,
641: \label{Psi_Jastrow}
642: \end{eqnarray}
643: with variational freedom in the two-particle pseudopotential 
644: $u({\bf r}_i - {\bf r}_j)$, which is usually taken to approach zero as 
645: $1/|{\bf r}|^p$ at large separations.
646: 
647: As in the case with a magnetic field present, we again consider
648: the relative single particle wavefunction.
649: For a Bose condensed  superfluid in a time reversal invariant system 
650: $\Phi_b({\bf r})$ can be taken as real and is  
651: everywhere non-negative.  If there are repulsive interactions between
652: the bosons in the superfluid, the amplitude of $\Phi_b({\bf r})$ will be 
653: reduced when the test particle is taken nearby another particle, 
654: which is implemented by the Jastrow pseudopotential in 
655: Eq.~(\ref{Psi_Jastrow}).
656: But the sign of $\Phi_b({\bf r})$ will remain positive, so in some sense 
657: all of the particle pairs are in a relative S-type state.
658: In the special case of a hard core interaction,
659: $\Phi_b({\bf r}) \to 0$ for ${\bf r} \to {\bf r}_i$, 
660: so one can then view this as an ``extended" S-type wavefunction.
661: 
662: Motivated by the preceding discussion of the $\nu=1/2$ Laughlin state,
663: in Eq.~(\ref{Psi_det-squared}) we introduced a simple time reversal 
664: invariant wavefunction for hardcore bosons which is the square of a 
665: determinant for free fermions filling a Fermi sea.
666: By construction this wavefunction is non-negative.  
667: Moreover, the wavefunction will have zeroes which coincide with the 
668: nodes of the filled Fermi sea,
669: $\Psi_{FS}({\bf r}_1, {\bf r}_2, \dots, {\bf r}_N) 
670: = \det e^{i {\bf k}_i \cdot {\bf r}_j}$.  
671: With time reversal invariance, a relative single fermion wavefunction, 
672: $\Phi_f({\bf r}) \equiv \Psi_{FS}({\bf r}, {\bf r}_2, \dots, {\bf r}_N)$, 
673: can be taken as real.
674: As a result, the nodal structure of $\Phi_f({\bf r})$ will be 
675: qualitatively different than for the filled Landau level state,
676: vanishing along nodal lines rather than at isolated points.  
677: These nodal lines will pass through the positions of all of the other 
678: fermions.\cite{Ceperley91}  
679: Upon taking the ``test" particle across a nodal line, the function 
680: $\Phi_f({\bf r})$ changes sign, vanishing linearly upon approaching the 
681: nodal line.
682: If one takes the position of the test particle close to another particle,
683: $\Phi_f({\bf r})$ will have a $p$-wave character -- in particular a 
684: $p_x$ form if we define the $x$-axis as being parallel to the 
685: nodal line.
686: 
687: Despite the $p$-wave character of the free fermion determinant, the 
688: determinant-squared wavefunction will not have a $d$-wave character. 
689: Rather, the relative single boson wavefunction will vanish 
690: quadratically upon crossing the nodal lines of the fermion determinant.
691: When the test particle is taken near another particle ${\bf r}_i$, 
692: the function vanishes quadratically,
693: $\Phi_b({\bf r}) \sim (x - x_i)^2$.  
694: The ``pair" wavefunction is thus of an ``extended S-type" form, 
695: vanishing along the residual nodal line.
696: 
697: Let us now consider the DBL many-body wavefunction Eq.~(\ref{Psi_DBL}), 
698: which is the product of two different Slater determinants for fermions 
699: that fill elliptical Fermi surfaces as shown in Fig.~\ref{fig:introFS}.
700: The nodal structure is revealed by exploring $\Phi_b({\bf r})$.
701: This function will have two sets of nodal lines, 
702: one from each of the determinants.  Due to the elliptical nature of the 
703: respective Fermi surfaces, the two sets of lines, both of which pass 
704: through all of the particles, will generally not coincide with one 
705: another.  Indeed, the fermion determinant coming from an elliptical 
706: Fermi surface will have nodal lines running preferentially perpendicular 
707: to the long axis of the ellipse.  Focusing on the behavior of 
708: $\Phi_b({\bf r})$ near a target particle ${\bf r}_i$, one anticipates a 
709: behavior of the $d_{xy}$ form,
710: $\Phi_b({\bf r}) \sim (x - x_i)(y - y_i)$.
711: Here we have assumed that the two nodal lines actually coincide with the
712: $x$ and $y$ axes.  In general, for a typical configuration of fixed 
713: particle coordinates and a given target particle, this will not 
714: precisely be the case.  More generically, the two nodal lines will 
715: intersect a particle at two angles which are not aligned with the axes.
716: But the sign of the relative wavefunction when the test particle 
717: encircles the target particle will still behave as $+ - + -$, 
718: the same sign structure as a D$_{xy}$ or D$_{x^2-y^2}$ orbital.  
719: This is illustrated in Fig.~\ref{fig:nodes}.
720: 
721: Since the DBL many-body wavefunction is not nodeless, it cannot be the 
722: ground state of a continuum Hamiltonian of bosons.  
723: If we put the coordinates on the sites of a 2d square lattice, 
724: the ground state wavefunction is only assured to be non-negative if the 
725: sign and form of the hopping matrix in the lattice tight binding 
726: Hamiltonian is such that it satisfies the Marshall sign conditions -- 
727: the requirement that a choice of gauge is possible to make all of the 
728: off-diagonal matrix elements negative.  
729: In Sec.~\ref{sec:Energetics} we consider a particular Marshall sign 
730: violating lattice boson Hamiltonian which might exhibit a ground state 
731: of the proposed D-wave form.
732: 
733: 
734: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
735: \subsection{Precedents of $(\det)_1 \times (\det)_2$ wavefunctions for 
736: spin liquids}
737: Before focusing solely on the DBL states, 
738: we want to mention that our construction of time reversal invariant 
739: bosonic wavefunctions as a product of two distinct determinants has 
740: nice precedents in the studies of spin liquids on the triangular 
741: lattice.  One can view the triangular Heisenberg antiferromagnet as a 
742: system of hard-core bosons at half-filling in the background field of 
743: flux $\pi$ through each triangle.  
744: Kalmeyer and Laughlin\cite{KL} proposed to view this in the continuum, 
745: obtaining a boson system at $\nu=1/2$, and their chiral spin liquid
746: wavefunction is precisely the lattice analog of the $\nu=1/2$ state 
747: Eq.~(\ref{Psi_nu_1/2}).  Alternatively, using a slave fermion 
748: approach,\cite{WenWilczekZee, LaughlinZou}
749: $b^\dagger = d_1^\dagger d_2^\dagger$,
750: we divide the boson charge equally between the two fermions,
751: so each sees on average a flux of $\pi/2$ per triangle;
752: the mean field where the $d_1$ and $d_2$ see the same static flux of 
753: $\pi/2$ per triangle gives a filled Landau level for each fermion
754: and reproduces precisely the Laughlin-Kalmeyer chiral spin liquid.
755:   
756: However, we can be more creative about the fluxes seen by the slave 
757: fermions while maintaining the average of $\pi/2$ flux per triangle.  
758: One example is to take different flux patterns for the two species as 
759: follows:  For the $d_1$ fermions, put $0$ flux through all up-pointing 
760: triangles and $\pi$ flux through all down-pointing triangles, 
761: while for the $d_2$ fermions interchange the locations of the $0$ and 
762: $\pi$ fluxes.
763: This state is in fact identical to the so-called 
764: $U1B \tau^1 \tau_-^0 \tau_+^1$ spin liquid found in 
765: Ref.~\onlinecite{ZhouWen}.
766: It is a time reversal invariant gapless algebraic spin liquid (ASL)
767: with Dirac nodes in the spinon spectrum;
768: it has very good energetics for the nearest neighbor triangular 
769: antiferromagnet, starting from the isotropic lattice and all the way to 
770: the limit of weakly coupled chains
771: (e.g., Ref.~\onlinecite{Yunoki06} found a different gauge-equivalent 
772: formulation of this state without realizing its ASL character).
773: 
774: Another example is obtained by taking yet different flux patterns:
775: Select one lattice direction -- chain direction in the anisotropic
776: lattice case.  For the $d_1$ fermions, put $0$ flux through triangles 
777: siding even chains and $\pi$ flux through triangles siding odd chains, 
778: while for the $d_2$ fermions interchange the locations of the $0$ 
779: and $\pi$ fluxes.  This is again a time reversal invariant ASL and is 
780: listed as $U1C \tau_+^0 \tau_-^0 \tau^1$ in 
781: Ref.~\onlinecite{ZhouWen}.
782: Unlike the $U1B$ state, there is no isotropic $U1C$ liquid, but
783: the energetics performance of $U1C$ and $U1B$ spin liquids is almost 
784: indistinguishable for weakly coupled chains and matches that of competing
785: magnetically ordered states; in this context, the time reversal 
786: invariant $U1B$ and $U1C$ liquids are significantly better than the 
787: chiral Laughlin-Kalmeyer variant.
788: 
789: Our construction of the DBL states differs in that we do not require
790: any special filling for the bosons and there is no special Heisenberg
791: spin symmetry.  Neither of these special conditions of the spin model 
792: setting are needed for the construction and subsequent gauge theory 
793: analysis to go through.  
794: Given the growing belief\cite{Rantner, WenPSG, Hermele} 
795: that critical spin liquids do exist, we do not see any reasons why the 
796: situation should be any different for our boson liquids at arbitrary 
797: incommensurate densities.  In our construction, such liquids will 
798: generically have some underlying partially filled bands and therefore 
799: Fermi surfaces of slave fermions.  
800: Of course, whether a particular state is realized in a given model 
801: requires a detailed case by case study, and in Sec.~\ref{sec:Energetics} 
802: we suggest some frustrated boson models that may stabilize the DBL phase.
803: Our primary goal in the next Secs.~\ref{sec:MFT}-\ref{sec:eff_FT}
804: will be to characterize the DBL states without worrying where to find 
805: them.
806: \\
807: 
808: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
809: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
810: \section{Mean Field Theory for the D-wave Bose Liquid}
811: \label{sec:MFT}
812: 
813: \subsection{Gauge Theory formulation}
814: \label{subsec:Gauge_formulation}
815: 
816: We next consider the challenge of constructing a field theory which 
817: can access such a $D_{xy}$-Bose liquid (DBL) state.
818: Due to our inability to implement flux attachment in a tractable 
819: time reversal invariant manner, we follow instead the slave particle 
820: approach.  As above, we decompose the hard core lattice boson as a 
821: product of two fermions, $b^\dagger = d_1^\dagger d_2^\dagger$.
822: Consider then a lattice $U(1)$ gauge theory on the square lattice 
823: in terms of these slave particles:
824: \begin{equation}
825: H_{U(1)} = H_t + H_a ~,
826: \label{HU1}
827: \end{equation}
828: with the fermion hopping Hamiltonian of the form,
829: \begin{widetext}
830: \begin{eqnarray}
831: \label{Ht}
832: H_t &=& -\sum_{\bf r}
833: \left[ 
834: t_\parallel e^{i a_x({\bf r})} d_1^\dagger({\bf r}) d_1({\bf r} + \hat{\bf x}) + 
835: t_\perp e^{i a_y({\bf r})} d_1^\dagger({\bf r}) d_1({\bf r} + \hat{\bf y})+ h.c. \right]\\
836: && -\sum_{\bf r}
837: \left[
838: t_\perp e^{-i a_x({\bf r})} d_2^\dagger({\bf r}) d_2({\bf r} + \hat{\bf x}) + 
839: t_\parallel e^{-i a_y({\bf r})} d_2^\dagger({\bf r}) d_2({\bf r} + \hat{\bf y})+ h.c. \right] ~.
840: \end{eqnarray}
841: \end{widetext}
842: The $d_1$ fermion hopping amplitudes in the $\hat{\bf x}$ and 
843: $\hat{\bf y}$ directions are $t_\parallel$ and $t_\perp$ respectively,
844: while the two amplitudes are interchanged for the $d_2$ fermions.
845: In the following, we take $t_\parallel \ge t_\perp$ for concreteness.
846: Note also that the two fermion species carry opposite gauge charges.
847: The gauge field Hamiltonian is simply
848: \begin{equation}
849: H_a = h \sum_{\bf r} \sum_{\mu = x, y} e^2_\mu({\bf r}) 
850: - K \sum_{\bf r} \cos[ (\nabla \times a)_{\bf r}  ] ~,
851: \label{Hgauge}
852: \end{equation}
853: where the lattice ``magnetic" field is
854: \begin{equation}
855: (\nabla \times a)_{\bf r} = a_x({\bf r}) + a_y({\bf r} + \hat{\bf x}) 
856: - a_x({\bf r}+ \hat{\bf y}) - a_y({\bf r}) ~.
857: \end{equation}
858: The integer-valued ``electric" field $e_\mu({\bf r})$ is canonically 
859: conjugate to the compact gauge field $a_\mu({\bf r})$ on the same link.  
860: The above Hamiltonian is supplemented by a gauge constraint on the 
861: physical states, which must satisfy Gauss' law,
862: \begin{equation}
863: (\nabla \cdot e)_{\bf r} = 
864: d^\dagger_1({\bf r}) d_1({\bf r}) - d^\dagger_2({\bf r}) d_2({\bf r}) ~,
865: \end{equation} with
866: \begin{equation}
867: (\nabla \cdot e)_{\bf r} = e_x({\bf r}) - e_x({\bf r}-\hat{\bf x})
868:  + e_y({\bf r}) - e_y({\bf r}-\hat{\bf y}) ~.
869: \end{equation} 
870: 
871: In the limit $h \gg K, t_\parallel, t_\perp$, the electric field 
872: vanishes, and Gauss' law reduces to $d^\dagger_1 d_1 = d^\dagger_2 d_2$ 
873: which projects back into the physical boson Hilbert space.  
874: In this strong coupling limit, it is possible to perturbatively 
875: eliminate the gauge field to obtain a Hamiltonian for hard core bosons 
876: hopping on the square lattice with additional ring exchange terms.
877: We pursue this in Section~\ref{sec:Energetics} where we compare the 
878: energetics of the DBL wavefunction with other states.  
879: Here we instead focus on the weak coupling limit, $K \gg h$, 
880: which suppresses the magnetic flux.
881: The usual slave particle mean field treatment corresponds to simply 
882: setting $(\nabla \times a)_{\bf r}$ equal to a constant.  
883: The simplest mean field state, and the one which should correspond to 
884: the wavefunctions in the previous section, is with zero flux through all 
885: plaquettes.  The corresponding mean field Hamiltonian describes 
886: non-interacting slave fermions.  
887: Each species has anisotropic near-neighbor hopping amplitudes,
888: but the two are related under the 90 degree rotation,
889: thus producing a boson liquid that respects the symmetries of the square 
890: lattice.
891: 
892: 
893: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
894: \subsection{Mean field results for the DBL}
895: \label{subsec:MF_props}
896: 
897: In order to focus on the effects of the underlying Fermi surfaces
898: without the complications of lattice physics such as Brillouin zone 
899: folding, we first consider fermions in the 2d continuum with anisotropic 
900: effective masses.
901: The resulting elliptical Fermi surfaces are,
902: \begin{eqnarray}
903: \label{wparam}
904: d_1: \quad (w k_x)^2 + (k_y/w)^2 &=& k_F^2 ~,\\
905: d_2: \quad (k_x/w)^2 + (w k_y)^2 &=& k_F^2 ~,
906: \end{eqnarray}
907: where the parameter $w \ge 1$ characterizes the degree of eccentricity
908: (the conventional eccentricity of the ellipses is given by 
909: $\epsilon = \sqrt{1 - 1/w^2}$).
910: In the D-wave Bose liquid, $w$ signifies the mismatch between the
911: two Fermi surfaces, and we will refer to this measure as
912: ``D-eccentricity".
913: We study equal time correlation functions since these can be compared 
914: directly with the properties of the wavefunctions, which is done in 
915: Sec.~\ref{sec:Wavefnc_props};
916: we also consider temporal dependencies within the mean field as a
917: measure of spectral properties.
918: It is useful to have in mind that much of the following analysis of 
919: long-distance properties needs only the knowledge of relevant Fermi
920: surface patches and not of the full surfaces.
921: 
922: Consider first the one-particle off-diagonal density matrix 
923: (or Green's function) for the boson,
924: $G_b({\bf r}, \tau) = 
925:  \la b^\dagger({\bf r}, \tau) b({\bf 0}, \tau) \ra$, with
926: $G_b({\bf 0}, 0) = \bar\rho$ -- the average boson density.
927: It will also be of interest to consider the momentum occupation 
928: probability,
929: \begin{eqnarray}
930: \la n_{\bf k} \ra = \la b^\dagger_{\bf k} b_{\bf k} \ra =
931: \int d{\bf r}\; G_b({\bf r}) e^{-i {\bf k} \cdot {\bf r}} ~. 
932: \end{eqnarray}
933: Within the mean field theory, the natural approximation for the order 
934: parameter correlation is
935: \begin{equation}
936: \label{Gb_MFdef}
937: G_b^{MF}({\bf r}, \tau) = 
938: G^{MF}_{d_1}({\bf r}, \tau) G^{MF}_{d_2}({\bf r}, \tau) / \bar\rho ~,
939: \end{equation}
940: where $G^{MF}_{d_\alpha}$ are the mean field (bare) fermion Green's 
941: functions.  This approximation satisfies 
942: $G_b({\bf 0}) = G_{d_\alpha}({\bf 0}) = \bar\rho$.
943: Here and below, the imaginary time $\tau$ is understood to be zero
944: if not explicitly present.
945: 
946: The fermion Green's functions are readily calculated.
947: Thus, at long distances $r \gg k_F^{-1}$, the main contribution to 
948: $G^{MF}_{d_\alpha}({\bf r})$ comes from the Fermi surface patches where 
949: the group velocity is parallel or antiparallel to the observation 
950: direction $\hat{\bf r} = {\bf r}/|{\bf r}|$.  
951: With inversion symmetry, we can denote the corresponding patch 
952: locations as $\pm {\bf k}_{F_\alpha}$ and the Fermi surface 
953: curvature as $c_\alpha$, and obtain
954: \begin{equation}
955: G^{MF}_{d_\alpha}({\bf r}) \approx \frac{1}{2^{1/2} \pi^{3/2}}
956: \frac{\cos({\bf k}_{F_\alpha} \cdot {\bf r} - 3\pi/4)} 
957:      {c_\alpha^{1/2} |{\bf r}|^{3/2}} ~.
958: \label{Gd_MF}
959: \end{equation}
960: It is important to remember that ${\bf k}_{F_1}, {\bf k}_{F_2}$, 
961: and $c_1, c_2$ depend implicitly on the direction $\hat{\bf r}$ and are 
962: different in general when the $d_1$ and $d_2$ Fermi surfaces do not 
963: coincide.
964: 
965: For the equal-time boson Green's function we thus get
966: \begin{eqnarray}
967: G_b^{MF}({\bf r}) &\sim& 
968: \frac{\cos[({\bf k}_{F_1} - {\bf k}_{F_2}) \cdot {\bf r}]}
969:      {c_1^{1/2} c_2^{1/2} |{\bf r}|^3} \nonumber \\
970: &+& 
971: \frac{\cos[({\bf k}_{F_1} + {\bf k}_{F_2}) \cdot {\bf r} - 3\pi/2]}
972:      {c_1^{1/2} c_2^{1/2} |{\bf r}|^3} ~,
973: \label{Gb_MF}
974: \end{eqnarray}
975: which decays algebraically while oscillating with 
976: $\hat{\bf r}$-direction-dependent wave vectors 
977: ${\bf k}_{F_1} + {\bf k}_{F_2}$ and ${\bf k}_{F_1} - {\bf k}_{F_2}$.
978: Such wave vectors, which are constructed by considering patches on the 
979: two Fermi surfaces that are parallel to each other, will span some 
980: new loci in the momentum space as illustrated in Fig.~\ref{fig:loci2kF}.
981: The above large-distance behavior corresponds to singularities 
982: $n_{\bf k} \sim |\delta k|^{3/2}$ across these lines.
983: In the zero eccentricity limit (i.e., when $w=1$), the two Fermi 
984: surfaces coincide, giving
985: \begin{equation}
986: G^{MF}_b (r)  \sim \frac{1 + \cos[2 k_F r - 3\pi/2]}{r^3}~,
987: \quad\quad \text{(S-type)} ~.
988: \end{equation}
989: Compared to the general case with $w>1$,
990: the ${\bf k}_{F_1} - {\bf k}_{F_2}$ locus shrinks here to zero momentum, 
991: while the singularity in $n_{\bf k}$ is hardened to $|k|$.
992: 
993: As we show in Sec.~\ref{sec:Gauge_fluct}, the power law decay of the 
994: mean field boson Green's function survives in the presence of gauge 
995: fluctuations (but with modified exponents).
996: The algebraic decay is a consequence (and a measurable indication) 
997: of the ``criticality'' of the DBL and its gapless excitations.
998: Of course, by the very construction we expect many gapless excitations
999: because of the underlying Fermi surfaces.  
1000: Some measure of the low-energy spectrum is contained in the time 
1001: dependence of the local boson Green's function,
1002: \begin{equation}
1003: G_b^{MF}({\bf 0}, \tau) = \frac{\nu_0^2}{\tau^2} ~,
1004: \label{Gtau}
1005: \end{equation}
1006: where $\nu_0$ is the density of states at the Fermi energy for each
1007: species.  The corresponding local boson spectral function is
1008: \begin{equation}
1009: A_b^{MF}({\bf r}={\bf 0}, E) = 
1010: \int_{\bf k} A_b^{MF}({\bf k}, E) = \nu_0^2 E ~.
1011: \label{Aloc}
1012: \end{equation}
1013: In the discussion of the DBL phase on the lattice in 
1014: Sec.~\ref{subsec:openFS}, we will encounter a situation where the 
1015: boson Green's function decays exponentially in space because of the 
1016: topology of the Fermi surfaces, while the above spectral signature 
1017: of the low-energy excitations depends only on the density of states
1018: which is a property of the entire Fermi surface and is insensitive 
1019: to the topology otherwise.
1020: 
1021: The most natural instability of the DBL is towards a superfluid state.  
1022: As we discuss in Section~\ref{sec:eff_FT}, in the limit of 
1023: vanishing D-eccentricity ($w=1)$, the resulting S-type Bose Liquid phase 
1024: obtained in mean field theory is, in the presence of gauge fluctuations,
1025: most probably unstable to superfluidity.  
1026: Moreover, our Sec.~\ref{sec:Wavefnc_props} analysis of the 
1027: determinant squared wavefunction appropriate to the S-Bose Liquid
1028: is consistent with off-diagonal long-range order.  
1029: These results strongly suggest that the mean field S-type Bose liquid 
1030: phase probably can not exist as a stable quantum phase.  
1031: However, with non-vanishing D-eccentricity, both the gauge theory 
1032: analysis in Sec.~\ref{sec:eff_FT} and the properties of 
1033: the corresponding $(\det)_x (\det)_y$ wavefunction which we explore
1034: in Sec.~\ref{sec:Wavefnc_props} suggest that the DBL is a stable 
1035: critical quantum phase.
1036: 
1037: It is also instructive to examine several other correlation functions 
1038: within the present mean field treatment.  
1039: Specifically, consider the boson density-density correlation function,
1040: \begin{eqnarray}
1041: D_b({\bf r}) &=&
1042: \la :\! \delta\hat\rho(0) \delta\hat\rho({\bf r}) \!: \ra  \\
1043: &=& N(N-1) \la \delta^d({\bf r}_1) \delta^d({\bf r}_2 - {\bf r}) \ra
1044: - \bar\rho^2 ~,
1045: \label{densitycorr:def}
1046: \end{eqnarray}
1047: where $N$ is the total number of particles.
1048: As defined, $D_b({\bf r})$ approaches zero for large separations, while 
1049: negative values at small distances signify a correlation hole.
1050: The density structure factor can be calculated as
1051: \begin{equation}
1052: D_b({\bf k}) = \int d{\bf r}\; D_b({\bf r}) e^{-i {\bf k} \cdot {\bf r}} 
1053: = \frac{\la |\delta\hat\rho_{\bf k}|^2 \ra}{V} - \bar\rho ~,
1054: \label{Db_k}
1055: \end{equation}
1056: where 
1057: $\delta\hat\rho_{\bf k} =
1058: \sum_j \exp[-i {\bf k} \cdot {\bf r}_j] - N\delta_{\bf k,0}$ and 
1059: $V$ is the system volume.
1060: 
1061: Microscopically, since $b^\dagger = d_1^\dagger d_2^\dagger$, for each 
1062: boson added to the system both a $d_1$ and a $d_2$ fermions are added.
1063: As such, the boson and fermion densities are equal: 
1064: $\hat\rho_b = \hat\rho_{d_1} = \hat\rho_{d_2}$.  
1065: However, in the mean field treatment, the two fermion flavors propagate 
1066: independently with different Fermi surfaces, so the corresponding 
1067: fermion density-density correlation functions, which we denote
1068: as $D_{d_\alpha}^{MF}$, do not coincide.  This ambiguity in the
1069: density correlation is an intrinsic deficiency of the 
1070: mean field theory.  As a crude measure, we approximate the boson density 
1071: correlation function as an average of that of the $d_1$ and $d_2$ 
1072: fermions, 
1073: \begin{equation}
1074: D_b^{MF}({\bf r}) \approx \frac{1}{2} 
1075: [ D^{MF}_{d_1}({\bf r}) + D^{MF}_{d_2}({\bf r}) ] ~.
1076: \label{Boson_MF_density_corr}
1077: \end{equation}
1078: The individual density correlation functions in the mean field theory 
1079: are simply,
1080: \begin{equation}
1081: D_{d_\alpha}^{MF}({\bf r}) = -|G^{MF}_{d_\alpha}({\bf r})|^2 \sim 
1082: - \frac{1 + \cos[2 {\bf k}_{F_\alpha} \cdot {\bf r} - 3\pi/2]}
1083:        {c_\alpha |{\bf r}|^3} ~,
1084: \label{DD_MF}
1085: \end{equation}
1086: where again ${\bf k}_{F_\alpha}$ is the place on the $d_\alpha$ Fermi 
1087: surface where the normal is parallel to the observation direction 
1088: $\hat{\bf r}$.
1089: We recognize the $2 k_F$ oscillation with power law envelope, which in 
1090: the momentum space translates to a singularity in the structure factor 
1091: $\sim |2k_F - k|^{3/2}$ across the $2k_F$ line 
1092: (such $2k_F$ surfaces are illustrated in Fig.~\ref{fig:loci2kF}), 
1093: while there is also a $|k|$ singularity at zero momentum.   
1094: As obtained from Eq.~(\ref{Boson_MF_density_corr}), the mean field 
1095: boson density correlator $D_b^{MF}$ has singularities at both 
1096: $2k_{F_1}$ and $2 k_{F_2}$, and ``knows" about the presence of both 
1097: Fermi surfaces.  Remarkably, this mean field approximation is 
1098: consistent with our analysis of the $(\det)_x (\det)_y$ boson 
1099: wavefunction in Sec.~\ref{sec:Wavefnc_props}, which also
1100: reveals singular behavior at both $2 k_{F_1}$ and $2 k_{F_2}$.
1101: 
1102: 
1103: 
1104: \begin{figure}
1105: \centerline{\includegraphics[width=\columnwidth]{figures/loci2kF.eps}}
1106: \vskip -2mm
1107: \caption{Singular lines in the momentum space for the boson
1108: Green's function (left) and the density correlation (right).
1109: In each panel, the thin lines show the elliptical $d_1$ and $d_2$
1110: Fermi surfaces for the parameter $w^4 = 3$, cf.~Eq.~(\ref{wparam}).
1111: In the left panel, the ${\bf k}_{F_1} \pm {\bf k}_{F_2}$ loci are 
1112: constructed by considering a pair of $d_1$ and $d_2$ fermions with 
1113: parallel ($+$ sign) or antiparallel ($-$ sign) group velocities.
1114: The axes are scaled by an ``average'' $k_F$ that would obtain if the 
1115: Fermi surfaces were circular for the same particle density.
1116: }
1117: \label{fig:loci2kF}
1118: \end{figure}
1119: 
1120: 
1121: Let us now consider the two-boson correlator,
1122: \begin{equation}
1123: G_{2b}({\bf r}_1, {\bf r}_2; {\bf r}_1^\prime, {\bf r}_2^\prime) \equiv 
1124: \la b^\dagger({\bf r}_1) b^\dagger({\bf r}_2) 
1125:     b({\bf r}_1^\prime) b({\bf r}_2^\prime ) \ra ~,
1126: \end{equation}
1127: which injects a pair of bosons at ${\bf r}_1, {\bf r}_2$ and removes a 
1128: pair at ${\bf r}_1^\prime, {\bf r}_2^\prime$.
1129: Consider first the limit where the separation between the two 
1130: injected bosons, ${\bf r} = {\bf r}_1 - {\bf r}_2$, and the two 
1131: removed bosons, ${\bf r}^\prime = {\bf r}_1^\prime - {\bf r}_2^\prime$, 
1132: are both small compared to the mean inter-particle spacing, 
1133: $|{\bf r}|, |{\bf r}^\prime| \ll k_F^{-1}$.
1134: Moreover, take the separation
1135: ${\bf R} = \frac{{\bf r}_1 + {\bf r}_2}{2} - 
1136:            \frac{{\bf r}_1^\prime + {\bf r}_2^\prime}{2}$ 
1137: between the injected and removed pair to be much larger than the 
1138: inter-particle spacing, $|{\bf R}| \gg k_F^{-1}$.  
1139: In this limit the mean field 2-boson correlator can be expressed as,
1140: \begin{equation}
1141: G_{2b}^{MF} \sim
1142: \frac{k_F^8}{ [1 + \frac{W}{2} \sin^2 (2 \phi) ]^{5/2} }
1143: \frac{\Psi_{xy}({\bf r}; \hat{\bf R}) 
1144:       \Psi_{xy}^*({\bf r}^\prime; \hat{\bf R})} 
1145:      {(k_F R)^6} ~,
1146: \end{equation}
1147: with ``pair wavefunction",
1148: \begin{equation}
1149: \Psi_{xy}({\bf r}; \hat{\bf R}) = ({\bf r} \cdot \hat{\bf R})^2 
1150: + W  \sin(2\phi) x y ~.
1151: \end{equation}
1152: Here for concreteness we specialized to the elliptical Fermi surfaces,
1153: Eq.~(\ref{wparam}), and $W \ge 0$ characterizes the D-eccentricity,
1154: \begin{equation}
1155: W = \frac{ w^4 + (1/w)^4 }{2} - 1 ~;
1156: \end{equation}
1157: $\hat{\bf R} = {\bf R}/R = \cos(\phi) \hat{\bf x} 
1158: + \sin(\phi) \hat{\bf y}$, is the unit vector pointing from one pair to 
1159: the other.  
1160: With vanishing D-eccentricity, $W=0$, the pair wavefunction is of an
1161: ``extended-$s$" form, for example $\Psi_{xy} \sim x^2$ for $\phi = 0$.
1162: This corresponds to a quadratic nodal line in the pair wavefunction.
1163: With $W=0$ the directions of the nodal lines for each pair are aligned 
1164: perpendicular to the vector connecting one pair to the other.
1165: On the other hand, in the limit of very large D-eccentricity, 
1166: $W \gg 1$, the pair wavefunction takes the $d_{xy}$ form, 
1167: $\Psi_{xy} \sim xy$, corresponding to two nodal lines along the 
1168: $x$ and $y$ axes.  Upon rotating the internal coordinate, the sign
1169: of the pair wavefunction takes the usual $d_{xy}$ form, precisely as for
1170: the Cooper pair wavefunction in a $d_{xy}$ superconductor.  
1171: Thus, the DBL has quasi--long-range order in the $d_{xy}$ boson 
1172: pair channel, although the power law decay exponent within mean field 
1173: theory is very large.
1174: 
1175: It is also interesting to examine the time decay of a ``local" two-boson 
1176: Green's function,
1177: \begin{equation}
1178: G_{2b}({\bf r}, {\bf r}^\prime; \tau-\tau^\prime) \equiv
1179: \la b^\dagger({\bf r}, \tau) b^\dagger({\bf 0}, \tau)
1180:     b({\bf r}^\prime, \tau^\prime) b({\bf 0}, \tau^\prime) \ra ~,
1181: \end{equation}
1182: with $|{\bf r}|,|{\bf r}^\prime| \ll k_F^{-1}$.  
1183: A pair of bosons is injected at positions ${\bf r}$ and ${\bf 0}$
1184: and is removed at later times at positions ${\bf r}^\prime$ and 
1185: ${\bf 0}$.
1186: Within mean field theory,
1187: \begin{equation}
1188: G^{MF}_{2b}({\bf r}, {\bf r}^\prime; \tau) \sim \frac{1}{\tau^4} 
1189: [ ({\bf r} \cdot {\bf r}^\prime)^2 + 2 W (xy) (x^\prime y^\prime) ] ~.
1190: \end{equation}
1191: For large D-eccentricity, $W \gg 1$, this factorizes into a product
1192: of two ``pair wavefunctions" of the $d_{xy}$ form.
1193: This correlator can be used to extract a local pair tunneling 
1194: density of states,
1195: \begin{equation}
1196: \rho_{xy}(E) = \int
1197: \la \hat{\cal D}_{xy}^\dagger(\tau) \hat{\cal D}_{xy}(0) \ra 
1198: e^{-E |\tau|} d\tau ~,
1199: \end{equation}
1200: where $\hat{\cal D}_{xy}^\dagger$ injects a $d_{xy}$ pair centered 
1201: at the origin,
1202: \begin{equation}
1203: \hat{\cal D}_{xy}^\dagger \equiv \int_{\bf r} \,
1204: xy \,\exp(-|{\bf r}|/\xi) b^\dagger({\bf r}) b^\dagger({\bf 0}) ~,
1205: \end{equation}
1206: with pair ``size" $\xi$.
1207: Within mean field theory one obtains a power law tunneling density of 
1208: states $\rho^{MF}_{xy}(E) \sim W E^3$,
1209: with an amplitude that grows with the D-eccentricity parameter $W$.  
1210: The tunneling density of states for an $s$-wave pair also vanishes as 
1211: $E^3$ but the amplitude is independent of $W$.  
1212: This pair tunneling density of states is perhaps the best diagnostic for 
1213: measuring the degree of local $d$-wave two-boson correlations in the DBL.
1214: 
1215: 
1216: Finally, we consider a box correlator for bosons which is defined 
1217: in the Introduction, Eq.~(\ref{Bb_def}).
1218: Within mean field this factorizes as,
1219: ${\cal B}^{MF}_b(x) = {\cal B}_{d_1}(x) {\cal B}_{d_2}(x)$, and from 
1220: Wick's theorem and 90-degree rotational invariance,
1221: \begin{equation}
1222: {\cal B}_{d_1}(x) = G_{d_1}^2(0,x) - G_{d_1}^2(x,0) 
1223: = -{\cal B}_{d_2}(x) ~.
1224: \end{equation}
1225: Thus,
1226: \begin{equation}
1227: \label{Bb_MF}
1228: {\cal B}^{MF}_b(x) = -[ G_{d_1}^2(0,x) - G_{d_1}^2(x,0) ]^2 \le 0 ~.
1229: \end{equation}
1230: Notice that in the case with zero eccentricity the mean field box 
1231: correlator vanishes, but since the $(\det)^2$ wavefunction is 
1232: non-negative, one expects ${\cal B}_b$ to be positive upon inclusion of 
1233: gauge fluctuations.
1234: With non-zero D-eccentricity, the mean field box correlator is 
1235: negative for all spatial separations, which reflects the underlying 
1236: $d_{xy}$ nodal structure, and decays as $x^{-6}$.
1237: It is possible that the inclusion of gauge fluctuations will again
1238: modify the mean field result in the case with closed Fermi surfaces
1239: because of the pairing tendencies of the $d_1$ and $d_2$ fermions.
1240: In the case with open Fermi surfaces to be discussed next,
1241: we conjecture that the exact box correlator will be negative 
1242: at large spatial separations while decaying to zero as the box size is 
1243: taken to infinity.
1244: This conjecture appears to be consistent with the box correlator 
1245: extracted from the $(\det)_x (\det)_y$ wavefunction with open Fermi
1246: surfaces in Sec.~\ref{subsec:lattice_wavefnc}.
1247: 
1248: 
1249: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1250: \subsection{Case with open Fermi surfaces}
1251: \label{subsec:openFS}
1252: 
1253: The preceding analysis holds for arbitrary Fermi surfaces,
1254: in particular, for the lattice bands obtained from Eq.~(\ref{Ht}).
1255: We only need to remember that the fermion Green's function 
1256: Eq.~(\ref{Gd_MF}) is determined by the Fermi surface patches with the 
1257: group velocities that are parallel or antiparallel to $\hat{\bf r}$. 
1258: Once $G_{d_1}$ and $G_{d_2}$ are known, the other correlation functions
1259: follow.
1260: 
1261: An interesting situation occurs on the lattice when the ratio
1262: $t_\parallel / t_\perp$ is such that the $d_1$ and $d_2$ Fermi surfaces 
1263: are open, which is illustrated in the right panel of 
1264: Fig.~\ref{fig:introFS}.
1265: In this case, for an observation direction close to, say, the $y$-axis, 
1266: there are no $d_1$ Fermi surface patches with normals in this direction, 
1267: so the $d_1$ Green's function has an exponential decay in this direction 
1268: instead of the power law Eq.~(\ref{Gd_MF}).  
1269: Since the real-space boson Green's function is the product of the two 
1270: fermion Green's functions, Eq.~(\ref{Gb_MFdef}), we conclude that it 
1271: decays exponentially in the directions near the $x$- and $y$-axes.  
1272: There may still still be directions near $\hat{\bf x} \pm \hat{\bf y}$ 
1273: in which both $G_{d_1}$ and $G_{d_2}$ show power law behavior, 
1274: and so will the boson Green's function.
1275: 
1276: Eventually, for large enough $t_\parallel / t_\perp$, the two Fermi
1277: surfaces will have no parallel patches, and the boson Green's
1278: function will decay exponentially in all directions.
1279: We will call this phase DLBL for D-wave correlated Local Boson Liquid.
1280: Note, however, that the system is still gapless and critical,
1281: as can be measured, e.g., from the local spectral function 
1282: Eq.~(\ref{Aloc}).
1283: Also, the boson density correlations still have the power law envelope, 
1284: Eq.~(\ref{DD_MF}), if one fermion field can propagate in the 
1285: observation direction.
1286: Furthermore, the boson box correlator, Eq.~(\ref{Bb_MF}), exhibits 
1287: power law behavior even though the single and pair boson Green's 
1288: functions are exponentially decaying.
1289: From the boson field perspective, the system is local and bosons
1290: have hard time to propagate; 
1291: nevertheless, the system has power law correlations in other properties 
1292: and, in particular, charges can propagate.
1293: 
1294: Based on an analysis of gauge fluctuations in Sec.~\ref{sec:Gauge_fluct},
1295: we conjecture that the $d_1$ and $d_2$ systems effectively decouple at 
1296: low energies in the DLBL, and at large distances the exact box correlator
1297: is negative and decays to zero as a non-oscillatory power law with 
1298: an integer exponent which is independent of non-universal lattice scale 
1299: physics, ${\cal B}_b(x) \sim - x^{-8}$.  This is the same as in the 
1300: mean field theory except with a larger power.
1301: 
1302: It is also worth mentioning the limiting case $t_\perp=0$ that gives 
1303: completely flat Fermi surfaces, which we will call ``extremal DLBL.''
1304: The $d_1$ fermions can move only along the $x$-axis,
1305: \begin{equation}
1306: G_{d_1}(x,y) = \delta_{y,0} \frac{\sin(k_F x)}{\pi x}~, \quad 
1307: \text{(extremal DLBL)} ~,
1308: \end{equation}
1309: while the $d_2$ fermions can move only along the $y$-axis.
1310: As a result, the boson field cannot propagate at all, not even one 
1311: lattice spacing.  However, this special system still has power law 
1312: density-density correlations, e.g.,
1313: \begin{equation}
1314: D^{MF}_{d_1}(x,y) \sim -\delta_{y,0} \frac{\sin^2(k_F x)}{x^2}, \quad 
1315: \text{(extremal DLBL)} ~,
1316: \label{Dd_MF_XTRM}
1317: \end{equation}
1318: as well as power law box correlation,
1319: \begin{equation}
1320: {\cal B}^{MF}_b(x) \sim - \frac{\sin^4(k_F x)}{x^4}, \quad
1321: \text{(extremal DLBL)} ~.
1322: \label{Bb_MF_XTRM}
1323: \end{equation}
1324: From the numerical study of the extremal DLBL wavefunction, 
1325: Sec.~\ref{subsubsec:xtrmDLBL}, we conjecture that these mean field 
1326: power laws also hold upon the Gutzwiller projection.
1327: In the gauge theory context, we would say that the fermions
1328: remain unaffected by the gauge field fluctuations.  
1329: This extremal DLBL wavefunction is of interest because of its 
1330: similarity to the so-called Excitonic Bose Liquid ground state of a pure 
1331: ring exchange model -- we discuss this in Sec.~\ref{sec:Energetics}.
1332: 
1333: 
1334: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1335: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% 
1336: \section{Properties of the wavefunctions}
1337: \label{sec:Wavefnc_props}
1338: 
1339: In this section we study the $\Psi = (\det)_x (\det)_y$ wavefunctions 
1340: directly and measure numerically their properties such as the boson 
1341: Green's function and the density correlation defined in 
1342: Sec.~\ref{subsec:MF_props}.
1343: A detailed comparison is made with the mean field, which provides an 
1344: initial guide.  This is followed by interpretations of the observed
1345: deviations from the mean field using simple ``Amperean interaction'' 
1346: rules of thumb;
1347: the actual calculations behind these rules within the gauge 
1348: fluctuations theory are given in Sec.~\ref{sec:Gauge_fluct}.
1349: 
1350: We first consider wavefunctions in the continuum so as to avoid 
1351: lattice effects and focus on the consequences of the underlying Fermi 
1352: surfaces.
1353: It appears that the S-type wavefunction, $\Psi = (\det)^2$,
1354: has off-diagonal long-range order, while a generic DBL wavefunction
1355: with non-zero D-eccentricity has only power law boson correlations.
1356: We then consider wavefunctions on the lattice where we can access 
1357: the DLBL phase with open Fermi surfaces described in 
1358: Sec.~\ref{subsec:openFS}; this features boson Green's function that 
1359: decays exponentially in real space.
1360: Finally, we discuss in some detail the extremal wavefunction 
1361: that is obtained when the Fermi surfaces are completely flat.
1362: 
1363: The calculations with the wavefunctions are performed numerically 
1364: using standard determinantal VMC (variational Monte Carlo) 
1365: techniques.\cite{vmc}  The system used in all calculations is a 
1366: square box with periodic boundary conditions.
1367: In each case, the particle number is chosen so as to fill complete 
1368: momentum shells under the Fermi surfaces.
1369: To facilitate the comparison, the presented mean field is calculated 
1370: for the same finite systems.
1371: 
1372: 
1373: 
1374: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1375: \subsection{Wavefunctions in the continuum}
1376: 
1377: Before proceeding with the numerics, the DBL structure factors 
1378: can be found analytically in some ranges in the momentum space:
1379: Thus, by expanding the boson wavefunction in terms of the orbitals that
1380: form each determinant, it is easy to see that
1381: $n_{\bf k}$ vanishes outside the ${\bf k}_{F1} + {\bf k}_{F2}$ 
1382: surface of Fig.~\ref{fig:loci2kF}, while $D({\bf k})$ vanishes
1383: outside the $2{\bf k}_{F1} + 2{\bf k}_{F2}$ surface.
1384: 
1385: 
1386: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1387: \subsubsection{S-type state}
1388: We begin by considering the limit with zero D-eccentricity, 
1389: in which case the boson wavefunction is positive everywhere 
1390: except for the nodes.
1391: Figure~\ref{fig:S-type} shows boson Green's function measured for 
1392: two systems with $N=161$ and $325$ particles.
1393: It appears that at large separations $G_b(r)$ approaches a finite 
1394: positive value which is roughly the same for both system sizes.
1395: To be more quantitative, the ${\bf k}=0$ mode contains
1396: $n_{{\bf k}=0}/N = 0.10$ and $0.083$ fractions of bosons 
1397: in the two systems.  The present data extracted from the continuum 
1398: wavefunction cannot rule out the possibility that the Green's function 
1399: vanishes in the large distance limit.
1400: But at the very least, the result of the projection is clearly dramatic 
1401: in this case with matched Fermi surfaces, since the fall-off of the 
1402: boson correlation is very slow (if any).
1403: We also remark that we unambiguously find off-diagonal long-range order
1404: when this wavefunction is studied on a lattice at fixed boson density
1405: 
1406: The extremely strong enhancement of the boson correlation over the
1407: mean field prediction seen in Fig.~\ref{fig:S-type}
1408: can be understood qualitatively as a result of the pairing of the 
1409: $d_1$ and $d_2$ fermions mediated by the gauge field.
1410: Indeed, as we discuss in Sec.~\ref{sec:Gauge_fluct},
1411: the constituents of such a (zero momentum) ``Cooper pair'' with 
1412: oppositely directed group velocities and opposite gauge charges 
1413: produce parallel gauge currents and therefore experience Amperean 
1414: attraction mediated by the gauge field.  
1415: This ``Amperean attraction'' rule of thumb for the enhancements in
1416: correlations appears to be taken to the extreme in the $(\det)^2$ 
1417: wavefunction, where we can crudely picture the fermions paired 
1418: back into $b$ and condensed, giving rise to long-range order in $G_b$.
1419: We also point out that this wavefunction has rather unusual 
1420: density-density correlations which are singular around the 
1421: $2 k_F$ circle.
1422: Since the presence of the off-diagonal order makes this state 
1423: less interesting to us, we do not consider such details any further.
1424: The observation of the order suggests that the S-type state is unstable; 
1425: the ultimate phase in this case is likely a conventional superfluid, 
1426: and a good superfluid wavefunction needs to be constructed differently, 
1427: e.g., using Jastrow approach that builds in proper density correlations.
1428: 
1429: 
1430: \begin{figure}
1431: \centerline{\includegraphics[width=\columnwidth]{figures/BdagB_CY100_PWR200_NBOS325_10dir.eps}}
1432: \vskip -2mm
1433: \caption{Boson correlation in the $(\det)^2$ wavefunction for two 
1434: systems with $N=161$ and $325$ particles.
1435: The distance is measured in units of $a = \rho^{-1/2}$, 
1436: and the plots are cut when the separation exceeds half of the box size.
1437: Since $G_b(r)$ maintains a sizable limiting value that depends only 
1438: weakly on $N$, we suggest that there is off-diagonal long-range order
1439: in this wavefunction.  
1440: The mean field result is also plotted, and the vertical scale is chosen 
1441: to accommodate its fast decay, $G_b^{MF} \sim 1/r^3$.
1442: }
1443: \label{fig:S-type}
1444: \end{figure}
1445: 
1446: 
1447: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1448: \subsubsection{DBL state}
1449: 
1450: We now consider in detail a representative case with non-zero
1451: D-eccentricity -- specifically, with $w^4 = 3$ in Eq.~(\ref{wparam}).
1452: The corresponding $d_1$ and $d_2$ Fermi surfaces are shown to scale in 
1453: Fig.~\ref{fig:loci2kF} together with the singular lines for the 
1454: order parameter and density correlations identified in the mean field,
1455: Sec.~\ref{sec:MFT}.
1456: 
1457: Figure~\ref{fig:allq} gives an overall view of $n_{\bf k}$ and 
1458: $D({\bf k})$ in the two-dimensional momentum space, and also shows a 
1459: one-dimensional cut in the $(1,0)$ k-space direction together
1460: with the mean field predictions.
1461: 
1462: \begin{figure}
1463: \centerline{\includegraphics[width=\columnwidth]{figures/allq_cy33_PWR200_NBOS325.eps}}
1464: \vskip -10mm
1465: \centerline{\includegraphics[width=\columnwidth]{figures/q10dir_cy33_PWR200_NBOS325.eps}}
1466: \vskip -2mm
1467: \caption{
1468: Top:  Mode occupation $n_{\bf k}$ and density structure factor
1469: $D({\bf k})$ measured in the continuum DBL wavefunction with $w^4=3$
1470: (note that our $D({\bf k})$ defined in Eq.~\ref{Db_k} vanishes 
1471: for large $k$).
1472: The boson number is $N=325$; the underlying Fermi surfaces and all 
1473: singular lines are shown to scale in Fig.~\ref{fig:loci2kF}.
1474: Bottom:  Cut through the data in the (1,0) direction in the k-space 
1475: and comparison with the mean field results.
1476: In the left panel, the arrows point the appropriate $k_{F1} \pm k_{F2}$ 
1477: momenta (see Fig.~\ref{fig:loci2kF}), while in the right panel the 
1478: arrows point the $2k_{F1}$ and $2k_{F2}$ momenta.
1479: }
1480: \label{fig:allq}
1481: \end{figure}
1482: 
1483: 
1484: Consider first the mode occupation function $n_{\bf k}$.
1485: Upon projection, this develops a squarish top with somewhat sharper 
1486: edges as compared with the mean field.  The latter is not shown in the 
1487: full k-space but looks more smooth, while a (1,0) cut can be seen in the 
1488: bottom panel of Fig.~\ref{fig:allq}.
1489: Coming from large momenta, significant deviations from the mean field 
1490: set in near the ${\bf k}_{F1} - {\bf k}_{F2}$ line of 
1491: Fig.~\ref{fig:loci2kF}
1492: (the corresponding location in the 1D cut is indicated with an arrow).
1493: This surface is singular in the mean field, but the singularities are 
1494: weak and are almost not visible, while they become more pronounced upon 
1495: the projection.  On the other hand, we observe no such enhancement near
1496: the ${\bf k}_{F1} + {\bf k}_{F2}$ line.
1497: 
1498: This difference between the ${\bf k}_{F1} - {\bf k}_{F2}$ and
1499: ${\bf k}_{F1} + {\bf k}_{F2}$ is also visible when examining the boson 
1500: Green's function in real space as shown in the left panels of 
1501: figures~\ref{fig:real-space-10}~and~\ref{fig:real-space-11}.
1502: The mean field $G_b^{MF}({\bf r})$, Eq.~(\ref{Gb_MF}), has both
1503: ${\bf k}_{F1} - {\bf k}_{F2}$ and ${\bf k}_{F1} + {\bf k}_{F2}$ 
1504: components equally present, while after the Gutzwiller projection one
1505: finds that the ${\bf k}_{F1} - {\bf k}_{F2}$ component dominates.
1506: Indeed, for the boson Green's function measured along the $x$-axis,
1507: Fig.~\ref{fig:real-space-10}, the relevant Fermi surface patches have 
1508: normals in the $(x, 0)$ direction and are easily located 
1509: (the relevant momentum space cut of $n_{\bf k}$ is also shown in the 
1510: left panel of Fig.~\ref{fig:allq}).
1511: By comparing with the mean field Eq.~(\ref{Gb_MF}), one finds that after 
1512: the projection the amplitude of the oscillation with the smaller 
1513: wave vector $|k_{F1} - k_{F2}|$ wins over that with the larger wave 
1514: vector $k_{F1} + k_{F2}$. 
1515: Slightly more care is needed when interpreting $G_b({\bf r})$ in the 
1516: $x=y$ diagonal direction, Fig.~\ref{fig:real-space-11}.
1517: In this case, the appropriate locations on the 
1518: ${\bf k}_{F1} - {\bf k}_{F2}$ surface are at the ``cusp'' points with 
1519: $k_x = -k_y$ in Fig.~\ref{fig:loci2kF}, since this is where the normal
1520: to the singular surface is parallel to the observation direction 
1521: $\hat{\bf r}$.  This component does not oscillate when moving along the 
1522: $x=y$ diagonal in real space, and when it is enhanced, the oscillations 
1523: due to the other ${\bf k}_{F1} + {\bf k}_{F2}$ component become less 
1524: visible, which is what what we see in Fig.~\ref{fig:real-space-11}.
1525: 
1526: 
1527: \begin{figure}
1528: \centerline{\includegraphics[width=\columnwidth]{figures/10dir_CY33_PWR200_NBOS325.eps}}
1529: \vskip -2mm
1530: \caption{Boson Green's function $G_b(r)$ and density correlation 
1531: $D_b(r)$ measured along the $r=(x,0)$ direction for the DBL wavefunction 
1532: with $N=325$ bosons and D-eccentricity characterized by $w^4=3$
1533: (this is the same system as in Fig.~\ref{fig:allq}).   
1534: The distance is measured in units of $a = \rho^{-1/2}$,
1535: and the plots are cut roughly where the signal in $G_b(r)$ approaches
1536: the noise level.  Note the logarithmic scales used, so it is the 
1537: absolute values that are plotted, while the positive/negative signs
1538: of the data are indicated with filled/open symbols respectively.
1539: In the case of the mean field data, the signs are indicated using
1540: solid/broken lines; $D^{MF}$ is always negative, see Eq.~(\ref{DD_MF}).
1541: }
1542: \label{fig:real-space-10}
1543: \end{figure}
1544: 
1545: 
1546: \begin{figure}
1547: \centerline{\includegraphics[width=\columnwidth]{figures/11dir_CY33_PWR200_NBOS325.eps}}
1548: \vskip -2mm
1549: \caption{The same as in Fig.~\ref{fig:real-space-10} but in the
1550: $x = y$ diagonal direction in real space.
1551: }
1552: \label{fig:real-space-11}
1553: \end{figure}
1554: 
1555: 
1556: Unfortunately, from this data we cannot attest whether the enhanced
1557: singularities are characterized by new exponents or whether we see
1558: only an amplitude effect.
1559: The locations where the enhancements occur are consistent with the 
1560: Amperean rules of thumb, Sec.~\ref{sec:Gauge_fluct}.
1561: Indeed, the $d_1$ and $d_2$ constituents of the boson operator 
1562: $b^\dagger = d_1^\dagger d_2^\dagger$ move in the opposite directions 
1563: when contributing to the correlation at ${\bf k}_{F1} - {\bf k}_{F2}$ 
1564: and therefore experience Amperean attraction and enhancement 
1565: (remembering that $d_1$ and $d_2$ carry opposite gauge charges), 
1566: while they move in the same direction when contributing at 
1567: ${\bf k}_{F1} + {\bf k}_{F2}$ and therefore experience Amperean 
1568: suppression.
1569: These pictures are made more precise in Sec.~\ref{sec:Gauge_fluct},
1570: where we find that within the gauge fluctuations theory, 
1571: the enhanced correlations are characterized by new exponents that 
1572: depend on the observation direction because of the varying degree 
1573: of the Fermi surface curvature matching in the $d_1$-$d_2$ pairing 
1574: channel.  The strongest such enhancement is expected along the 
1575: diagonals $x = \pm y$, which is roughly what we find in the projected 
1576: wavefunctions
1577: (we repeat again that we cannot make any statement about the exponents 
1578: from the data other than that we see increased numerical correlations 
1579: compared with the mean field).
1580: 
1581: 
1582: Consider now the density correlation $D_b(r)$ and the corresponding 
1583: structure factor $D_b({\bf k})$. 
1584: In Fig.~\ref{fig:allq}, we can see the singular $2{\bf k}_{F1}$ 
1585: and $2{\bf k}_{F2}$ lines, cf.~Fig.~\ref{fig:loci2kF};
1586: the enhancements are peaked where the two curves cross.
1587: The singular $2k_F$ points are also visible in the (1,0) k-space cut,
1588: bottom right panel of Fig.~\ref{fig:allq}.
1589: On the other hand, the $|k|$ singularity near zero momentum is 
1590: not enhanced.
1591: Examination of the density correlation in real space, right panels of
1592: figures~\ref{fig:real-space-10}~and~\ref{fig:real-space-11},
1593: shows an overall increase and dominance of the oscillatory 
1594: components over the zero-momentum component as compared with the 
1595: mean field Eq.~(\ref{DD_MF}).
1596: Again, we cannot tell whether there is a change in the exponents or 
1597: just an amplitude effect.  The $2k_F$ enhancements of the density 
1598: correlation agree qualitatively with the gauge theory expectations.  
1599: Indeed, consider the density operator $d_1^\dagger d_1$.
1600: The particle and hole constituents (which are oppositely gauge charged) 
1601: have antiparallel group velocities when contributing to the $2 k_F$ 
1602: component and therefore experience Amperean attraction, while they 
1603: move in the same direction when contributing at zero momentum and 
1604: therefore repel each other.  
1605: Again, these rules of thumb are made more precise in 
1606: Sec.~\ref{sec:Gauge_fluct}, where such enhancements in the particle-hole 
1607: channel are characterized by new direction-dependent exponents
1608: (the density correlations are found to decay more slowly when less 
1609: curved patches are involved).
1610: Note that the projection imposes 
1611: $\hat{\rho}_b = \hat{\rho}_{d_1} = \hat{\rho}_{d_2}$, and
1612: we expect $D_b$ to acquire both the $2k_{F1}$ and $2k_{F2}$ signatures 
1613: (in the gauge theory, $\rho_{d_1}$ and $\rho_{d_2}$ imprint on each other
1614: via non-singular high-energy connections that are not manifest in our 
1615: low-energy effective description of Sec.~\ref{sec:Gauge_fluct}).
1616: 
1617: 
1618: To summarize, the DBL wavefunction clearly knows about the underlying 
1619: Fermi surfaces; it also contains germs of the gauge fluctuations theory, 
1620: since the enhancements/suppressions of the various ``$2 k_F$'' lines 
1621: appear to agree with the Amperean rules.
1622: However, there are no reasons to believe that the wavefunction 
1623: and the gauge theory will have the same long-distance properties.
1624: In particular, our measurements can not tell whether the long-distance
1625: power laws are changed upon the projection compared with the mean field.
1626: Still, it is gratifying to see that the Amperean rules work for the
1627: DBL wavefunction.
1628: 
1629: 
1630: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1631: \subsection{Wavefunctions on the lattice}
1632: \label{subsec:lattice_wavefnc}
1633: 
1634: We now describe our results for various states on the lattice.
1635: As we have already mentioned, S-type wavefunction, $\Psi = (\det)^2$,
1636: with finite boson density per site, have off-diagonal long-range 
1637: order, which we confirm unambiguously using finite-size scaling.
1638: On the other hand, wavefunctions with non-zero D-eccentricity
1639: do not show such order.  
1640: While we have not performed detailed studies, we expect that the case 
1641: with closed Fermi surfaces is similar to the already described continuum 
1642: DBL wavefunction, with a slight complication that one needs to do
1643: proper Brillouin zone folding when considering singular surfaces in the 
1644: momentum space.
1645: 
1646: \subsubsection{Open Fermi surfaces -- DLBL state}
1647: Of particular interest is the case with open Fermi surfaces,
1648: which can only be realized on the lattice.
1649: Specifically, we studied a $24 \times 24$ system with $N=216$
1650: bosons ($\rho = 0.375$ per site) and the fermion hopping parameters 
1651: $t_\parallel = 1.35$, $t_\perp = 0.65$.  
1652: In this case, the $d_1$ and $d_2$ Fermi surfaces have no parallel 
1653: patches; we therefore expect that the boson Green's function decays 
1654: exponentially.
1655: When we measure $G_b(r)$, we find that it drops below our noise level 
1656: already at 3 lattice spacings, so the corresponding plots are not 
1657: particularly informative and are not shown here.  
1658: Where we can measure reliably, the values after the projection are 
1659: of the same order as the mean field values in the same system.
1660: Since the latter decay exponentially at large distances,
1661: we conjecture the same behavior in the DLBL wavefunction.
1662: 
1663: On the other hand, the boson density correlation along the $x$
1664: and $y$ axes shows oscillations with a clear power law envelope, 
1665: while the correlations are much smaller in the diagonal $x = \pm y$ 
1666: directions; this behavior is again qualitatively consistent with the
1667: mean field, since neither $d_1$ nor $d_2$ Fermi surfaces have normals 
1668: in the diagonal directions, while one or the other has normals 
1669: along the $x$ or $y$ axis.
1670: 
1671: 
1672: 
1673: \subsubsection{Flat Fermi surfaces -- extremal DLBL state}
1674: \label{subsubsec:xtrmDLBL}
1675: Finally, let us discuss the extremal case when the Fermi surfaces
1676: are completely flat.  The $d_1$ fermions can move only along the
1677: $x$-axis and $d_2$ fermions only along the $y$-axis.  
1678: Using fermion orbitals that are localized on individual rows for $d_1$
1679: (or columns for $d_2$), one can see that the boson wavefunction is 
1680: nonzero only when the number of particles on each row (or on each column)
1681: is the same.
1682: Thus, the bosons can not propagate and their inter-site 
1683: correlation is identically zero.
1684: As we discuss in Sec.~\ref{sec:Energetics}, pure ring Hamiltonian
1685: conserves boson number on each row/column, and the extremal DLBL
1686: wavefunction may be useful in this context.
1687: 
1688: We can still use the box correlator, Eq.~(\ref{Bb_def}), to
1689: characterize the state and see some ``gaplessness'' in the system.  
1690: The measurement is shown in Fig.~\ref{fig:corrBOX}, where we also plot 
1691: renormalized mean field result.  
1692: The latter is obtained by dividing Eq.~(\ref{Bb_MF}) by 
1693: $\rho^2 (1-\rho)^2$, and a crude justification for such 
1694: procedure\cite{Zhang} is as follows:  
1695: Each $d_1$ or $d_2$ mean field box calculation contains 
1696: implicitly a weight of order $\rho^2 (1-\rho)^2$, since for the ring
1697: operator to be nonzero, two specific sites need to be occupied
1698: and two need to be empty.  However, after the projection it is enough 
1699: to require that only the $d_1$ configuration is ``correct'' since the 
1700: $d_2$ fermions are tied to $d_1$.
1701: From Fig.~\ref{fig:corrBOX}, we see that the box correlator is negative 
1702: and tracks closely the renormalized mean field values.
1703: 
1704: 
1705: \begin{figure}
1706: \centerline{\includegraphics[width=\columnwidth]{figures/corrBOX_EXTRMDxDy_NVOL576_NBOS216.eps}}
1707: \vskip -2mm
1708: \caption{Box correlator ${\cal B}_b(x)$ in the extremal DLBL wavefunction
1709: with flat Fermi surfaces on a $24 \times 24$ lattice with $N=216$ bosons
1710: (the lines are guide to the eye).
1711: Note that the box correlator is negative in the mean field
1712: and is found to be negative in the DLBL state; 
1713: we plot $-{\cal B}_b(x)$ to be able to use the logarithmic scale.
1714: The mean field Eq.~(\ref{Bb_MF}) is renormalized as explained in the
1715: text using a numerical factor that removes double weighting of 
1716: configurations.  
1717: The measured values below $10^{-5}$ are at the noise threshold.
1718: }
1719: \label{fig:corrBOX}
1720: \end{figure}
1721: 
1722: 
1723: \begin{figure}
1724: \centerline{\includegraphics[width=\columnwidth]{figures/DENScorr_EXTRMDxDy_NVOL576_NBOS216.eps}}
1725: \vskip -2mm
1726: \caption{Boson density correlation for the same extremal DLBL system
1727: as in Fig.~\ref{fig:corrBOX}.  
1728: Left panel:  Real space dependence along the $x$-axis, $D_b(x,0)$, 
1729: together with the mean field expectation.
1730: Right panel:  Density structure factor in the momentum space.  
1731: }
1732: \label{fig:DENScorr}
1733: \end{figure}
1734: 
1735: We can also characterize the extremal DLBL state by measuring the 
1736: density correlations and comparing with the mean field prediction, 
1737: Eq.~(\ref{Dd_MF_XTRM}).  
1738: The results are in Fig.~\ref{fig:DENScorr}.
1739: The left panel shows $D_b(x,0)$ as a function of the distance
1740: along the $x$-axis; it has stronger oscillatory component than the
1741: mean field and swings back and forth across the zero line
1742: while the mean field only touches it, but the overall magnitudes
1743: are comparable and decay as $1/x^2$.
1744: We also find that the density correlation in the $x = \pm y$ diagonal 
1745: directions decays exponentially (not shown);
1746: the mean field predicts zero correlation unless strictly along the axes, 
1747: and we expect that after the projection this corresponds to exponential 
1748: decay.
1749: 
1750: In the right panel of Fig.~\ref{fig:DENScorr}, we show the density
1751: structure factor $D({\bf k})$.  Several features are clearly visible:
1752: rising ``towers'' draw attention to the lines $(\pm 2 k_F, k_y)$ and 
1753: $(k_x, \pm 2 k_F)$;
1754: one can also see a ``cross'' formed by the lines $(0, k_y)$ and 
1755: $(k_x, 0)$ that run along the axes.
1756: As far as we can say, the character of the singularities across these
1757: lines remains the same as in the mean field; there is an amplitude 
1758: enhancement of the $2k_F$, but no qualitative difference otherwise. 
1759: In particular, near the line $(0, k_y)$, we observe
1760: $D_{b, {\rm sing}}(k_x \to 0, k_y) = A |k_x|$ with $A$ which is 
1761: independent of $k_y$ as long as $|k_y| \gg |k_x|$.
1762: 
1763: We thus conjecture that the extremal DLBL wavefunction is adequately 
1764: described by the free fermion mean field. 
1765: Some of our findings, e.g., the cross singularity that has a long 
1766: wavelength character, can be understood semi-analytically, 
1767: since the absolute value of the wavefunction has a Jastrow form with a 
1768: peculiar pseudo-potential,
1769: $u(x, y) \sim -\ln|\sin(\pi x /L)| \delta_{y,0} 
1770:               -\ln|\sin(\pi y /L)| \delta_{x,0}$,
1771: it seems plausible that one can calculate other properties of this 
1772: wavefunction analytically.
1773: It is interesting to note that the above suggests a free fermion
1774: description of this boson state, perhaps with some constraints that become
1775: irrelevant in an infinite system.  
1776: We also note that the extremal DLBL appears to be a relative of the 
1777: so-called Excitonic Bose Liquid (EBL) phase predicted in a pure ring 
1778: model on the square lattice\cite{Paramekanti} that we discuss in 
1779: Sec.~\ref{sec:Energetics};
1780: the above therefore suggests that there may be some such description
1781: of the EBL in terms of fermions carrying fractional $1/2$ charge.
1782: 
1783: 
1784: 
1785: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1786: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1787: \section{Gauge Fluctuations}
1788: \label{sec:Gauge_fluct}
1789: 
1790: We now study the gauge theory using analytic techniques,
1791: focusing on the large $K$ limit where it is reasonable to
1792: ignore the presence of magnetic monopoles in space-time and to 
1793: expand the cosine term in Eq.~(\ref{Hgauge}) treating the 
1794: gauge field as non-compact.  
1795: Moreover, in the $K \to \infty$ limit, the fermions become free and 
1796: one can put them into a Fermi sea state.
1797: A Fermi surface of fermions minimally coupled to a non-compact $U(1)$ 
1798: gauge field has a considerable history.\cite{Holstein, Reizer, PALee, IoffeKotliar, LeeNagaosa, Polchinski, Altshuler, Nayak, YBKim, LeeNagaosaWen, Senthil, Galitski_gauge}
1799: The $2+1$D such system has been studied most notably as a theory for 
1800: the spin sector in the uniform RVB phase in the slave boson approach
1801: to the high-$T_c$ superconductors.
1802: It has been argued that such fermion systems have a stable phase that 
1803: in some crude aspects is similar to a Fermi liquid -- for example, 
1804: it has a finite long-wavelength compressibility and spin susceptibility.
1805: However, the system is strikingly different in other aspects and is 
1806: described by a new non-Fermi liquid fixed point.
1807: A scaling description of this fixed point was developed in 
1808: Refs.~\onlinecite{Polchinski, Altshuler}.
1809: 
1810: While we largely follow the earlier work, we find that it is 
1811: convenient to consider a slight reformulation in which the only 
1812: uncontrolled approximation is to assume that the gauge field dynamics 
1813: can be described correctly by the RPA approximation, retaining only 
1814: terms quadratic in the gauge field with a singular quadratic kernel.
1815: A virtue of this approach is that one thereby obtains a theory which has 
1816: an $N$ flavor extension which is soluble at $N=\infty$.  Moreover, 
1817: a controlled and systematic perturbation expansion in powers of $1/N$ 
1818: can be implemented, which allows one to compute  physical properties in 
1819: terms of non-universal ``bare" parameters such as the shapes of the 
1820: Fermi surfaces.  We employ this approach to calculate the leading
1821: $1/N$ behavior for both the boson correlator, $G_b({\bf r})$, 
1822: and the density-density correlator, $D_b({\bf r})$, in the D-wave Bose 
1823: Liquid.
1824: 
1825: In the next Sec.~\ref{sec:eff_FT}, we will consider an effective field 
1826: theory approach, which allows us to check for the stability of the 
1827: DBL phase that is present at large $N$.  Specifically, we study the 
1828: effects of residual short range attractive interactions between the 
1829: two fermion flavors.  
1830: With vanishing D-eccentricity when the Fermi surfaces for both fermion
1831: species become the same, the $s$-wave Cooper channel is ``nested", and a 
1832: possible instability towards a paired BCS state seems likely. 
1833: The resulting state is a bosonic condensate, since 
1834: $\la b^\dagger \ra = \la d_1^\dagger d_2^\dagger \ra \ne 0$.
1835: However, a non-vanishing D-eccentricity of the Fermi surfaces destroys 
1836: the BCS nesting and the possible pairing instability.
1837: Indeed, we will find that the DBL with a large D-eccentricity can exist 
1838: as a stable phase.
1839: (For smaller D-eccentricity, an instability towards a finite momentum 
1840: Bose condensate or an incommensurate charge density wave co-existing 
1841: with the DBL is a possibility.)
1842: As is discussed in Sec.~\ref{sec:Wavefnc_props}, this is nicely in line 
1843: with the properties of the associated Gutzwiller wavefunctions:
1844: The $(\det)^2$ wavefunction that obtains in the limit of zero 
1845: D-eccentricity appears to have off-diagonal long range order,
1846: whereas the general $(\det)_x (\det)_y$ wavefunction exhibits power law 
1847: correlations consistent with the DBL, as extracted from the gauge theory.
1848: 
1849: 
1850: 
1851: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1852: \subsection{Formulation}
1853: \label{subsec:Fluct_formulation}
1854: 
1855: Being interested in the low energy properties, it is legitimate to 
1856: focus on the fermions living near the Fermi surfaces, just as in 
1857: Fermi liquid theory.  Moreover, the longitudinal Coulomb interactions
1858: mediated by the gauge field are readily screened out within an RPA 
1859: treatment, generating short-ranged screened density-density interactions.
1860: But the transverse fluctuations of the gauge field -- the photon -- are 
1861: incompletely screened by the fermion particle-hole excitations.  
1862: It is thus adequate to just retain the transverse components of the 
1863: vector potential.  Working in the Coulomb gauge with 
1864: $\bm{\nabla} \cdot {\bf a} = 0$, the vector potential reduces to a 
1865: scalar, e.g.,
1866: ${\bf a}({\bf k}) = a({\bf k}) \; {\rm sign}(k_y) 
1867: (k_y \hat{\bf x} - k_x \hat{\bf y}) / |{\bf k}|$. 
1868: 
1869: Moreover, for each fermion species, it will be sufficient to focus on a 
1870: pair of Fermi surface patches with normals parallel to some axis, 
1871: say $\hat{\bf x}$, as shown in Fig.~\ref{fig:patches}.   
1872: It is the $a_x$ component of the vector potential which is minimally 
1873: coupled to these fermions.  The important wave vectors of the gauge field
1874: $a_x({\bf k})$ that are strongly Landau damped by the fermions in these 
1875: patches satisfy $k_x \ll k_y$ (see Fig.~\ref{fig:patches}), 
1876: and in this region of momentum space $a_x({\bf k}) \approx a({\bf k})$.  
1877: Conversely, the modes of the gauge field $a({\bf k})$ with $k_x \ll k_y$
1878: feed back and scatter the patch fermions.
1879: Thus, we can focus on fermion fields $d_1({\bf k}), d_2({\bf k})$ and 
1880: the gauge field $a({\bf k})$ that are confined to their respective 
1881: patches in momentum space, as shown schematically in 
1882: Fig.~\ref{fig:patches}.
1883: 
1884: For concreteness, we assume that for each fermion species there are two 
1885: relevant patches on the opposite sides of the Fermi surface that are 
1886: labelled $s = R/L = +/-$ for the right/left patch as indicated in 
1887: Fig.~\ref{fig:patches}.  We then decompose the Fermion fields into
1888: right and left movers by writing,
1889: \begin{equation}
1890: d_{\alpha} ({\bf r}) \sim \sum_s 
1891: e^{i s {\bf k}_{F_\alpha} \cdot {\bf r}} d_{\alpha s}({\bf r}) ~.
1892: \end{equation}
1893: Here, ${\bf k}_{F_\alpha}$ denote the locations on the two Fermi surfaces
1894: which have normals aligned along the $x$-axis, and the fields 
1895: $d_{\alpha s}({\bf r})$ are assumed to be slowly varying on the scale
1896: of $k_F^{-1}$.  
1897: 
1898: 
1899: The full low energy action consists of three terms describing the 
1900: dynamics of the gauge field, of the fermions, and of their interactions:
1901: \begin{equation}
1902: S^{(0)} = \int d^2{\bf r} \, d\tau \, 
1903: [ {\cal L}_d + {\cal L}^{(0)}_a + {\cal L}_{int} ] ~.
1904: \end{equation}
1905: The ``patch" Lagrangian density for the fermions is simply,
1906: \begin{equation}
1907: {\cal L}_d =  \sum_{\alpha} \sum_{s=\pm} 
1908: d^\dagger_{\alpha s} [ \partial_\tau + s v_\alpha (-i \partial_x) 
1909:                        + \frac{v_\alpha c_\alpha }{2} (-i \partial_y)^2
1910:                       ] d_{\alpha s} ~,
1911: \end{equation}
1912: which is characterized by the local Fermi velocities, $v_\alpha$, 
1913: ($\alpha=1,2$), and the local Fermi surface curvatures, $c_\alpha$.
1914: The curvature is crucial in the fermion-gauge problem since there is an 
1915: effective decoupling of the fermions outside the patch from the 
1916: gauge field, $a = a_x$, which is parallel to the patch normals.
1917: Here $s=\pm 1$ specifies the sign of the group velocity while $v$ will 
1918: always denote the absolute value.  
1919: 
1920: The dynamics for the transverse gauge field describes the free photon,
1921: \begin{equation}
1922: {\cal L}_a^{(0)} = \frac{1}{2} [ (\partial_\tau a)^2 
1923:                                  + \kappa (\bm{\nabla} a )^2 ] ~,
1924: \end{equation}
1925: whereas the gauge-fermion interactions are of the usual minimal coupling 
1926: form:
1927: \begin{equation}
1928: \label{Lint}
1929: {\cal L}_{int} = \sum_{\alpha, s} g_\alpha s v_\alpha 
1930: d^\dagger_{\alpha s} \,a\, d_{\alpha s} ~.
1931: \end{equation} 
1932: Here, $g_1 = -g_2 \equiv g$ with $g=1$, but it will be convenient for 
1933: bookkeeping purposes to retain $g$ as a dimensionless parameter in the 
1934: theory.
1935: We emphasize that both the fermion fields and the gauge field, 
1936: which enter the above action in real space, are slowly varying fields
1937: corresponding to momentum modes inside their respective patches;
1938: in what follows, we always specify the wave vectors of the slow fields 
1939: relative to their respective patch centers.
1940: 
1941: 
1942: \begin{figure}
1943: \centerline{\includegraphics[width=\columnwidth]{figures/patches.eps}}
1944: \vskip -2mm
1945: \caption{Relevant momentum patches of the strongly coupled fermion -
1946: transverse gauge field system.
1947: For simplicity, here and throughout the text, we take the Fermi surface 
1948: patches with normals in the $\hat{\bf x}$ direction; 
1949: the required properties are the Fermi velocities and surface curvatures;
1950: the latter are crucial since differently oriented patch systems
1951: effectively decouple at low energies.
1952: }
1953: \label{fig:patches}
1954: \end{figure}
1955: 
1956: 
1957: 
1958: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1959: \subsection{RPA approximation for the gauge field dynamics}
1960: \label{subsec:RPA}
1961: 
1962: The above theory is strongly coupled and cannot be treated perturbatively
1963: in $g$.  As such, it is necessary to resort to approximations to make any
1964: meaningful progress.  Here we make our central approximation, namely 
1965: replacing the dynamics of the gauge field by a Landau damped form which 
1966: results if one sums the RPA bubble of particle-hole excitations:
1967: $S_a^{(0)} \to S_a$ with
1968: \begin{equation}
1969: \label{Sa}
1970: S_a = \frac{1}{2} \int_{{\bf q}\omega} 
1971: \left(\Gamma \frac{|\omega|}{|q_y|} + \chi |q_y|^2  \right) 
1972: |a ({\bf q}, \omega)|^2 ~.
1973: \end{equation}
1974: The momentum/frequency integral is understood as 
1975: $\int_{{\bf q} \omega} \equiv \int d^2{\bf q} d\omega / (2\pi)^3$,
1976: with the momentum restricted to a region satisfying
1977: $|q_x| < \Lambda_x, |q_y| < \Lambda_y$ with $\Lambda_x \ll  \Lambda_y$.
1978: 
1979: The magnitude of the Landau damping coefficient, $\Gamma$, will be 
1980: determined by the low-energy fermions near the specified Fermi surface 
1981: patches.  Within the RPA approximation one obtains,
1982: \begin{equation}
1983: \Gamma_{RPA} = \frac{g^2}{2\pi} \left( \frac{1}{c_1} + \frac{1}{c_2} \right) ~,
1984: \label{Gamma_RPA}
1985: \end{equation}
1986: but we will retain $\Gamma$ as an independent parameter since beyond 
1987: this point we will not be using RPA in any event.  
1988: It will sometimes prove convenient to consider the limit in which the 
1989: dimensionless gauge charge,
1990: \begin{equation}
1991: e^2 \approx \frac{g^2}{\Gamma c},
1992: \end{equation}
1993: is small.  Here $c$ is a characteristic Fermi surface curvature.
1994: Indeed, within the $N$-flavor extension discussed below, while $1/N$ will
1995: be the small parameter that leads to a controlled analysis, 
1996: already at order $1/N$ there are some physical properties which we can 
1997: only compute analytically when $e^2$ is simultaneously taken as a 
1998: small parameter.
1999: 
2000: The parameter $\chi$ that enters in $S_a$ is a ``stiffness'' which
2001: has to be positive for stability.  
2002: Starting from the lattice gauge theory formulation, $\chi$ gives a 
2003: measure of the energetic cost of setting up static internal fluxes.  
2004: One can very crudely estimate $\chi$ using the expression for the 
2005: diamagnetic response of free fermions,
2006: \begin{eqnarray}
2007: \chi &\approx & \frac{g^2}{12 \pi m_d} ~,
2008: \label{chi}
2009: \end{eqnarray}
2010: where $m_d$ is some effective mass.
2011: We reiterate that this stiffness comes from some energetics that is 
2012: assumed to stabilize the Fermi surfaces of the $d$ fermions 
2013: in the first place.  Therefore, from the low-energy perspective,
2014: $\chi$ should be viewed as a phenomenological parameter that encodes 
2015: some high-energy physics.  Furthermore, the stiffness $\chi$ is 
2016: independent of the patch orientation in momentum space for the 
2017: transverse gauge field $a({\bf k})$, since gauge invariance requires 
2018: that the field energy be proportional to 
2019: $\chi B^2 = \chi (\partial_x a_y - \partial_y a_x)^2$.  
2020: In any event, the magnitude of $\chi$ will play a minor role below.
2021: Indeed, for correlation functions with power law decay as we find below,
2022: changing $\chi$ will only modify the amplitude but not the form of the 
2023: correlators.
2024: 
2025: Before proceeding with an analysis of the full theory, 
2026: $S_d + S_a + S_{int}$, it is important to note that within the above 
2027: approximation the Landau damped dynamics of the gauge field described by 
2028: the action $S_a$ is both singular and harmonic.  We have ignored possible
2029: terms in the action which involve higher powers of the gauge field.
2030: As such, at this stage we could integrate out the gauge field exactly 
2031: generating a non-local and retarded four-fermion interaction.  
2032: It is the absence of non-linearities for the gauge field dynamics which 
2033: allows us to introduce a large-$N$ generalization which is soluble at 
2034: $N=\infty$ and which facilitates a formal and systematic $1/N$ expansion,
2035: as we now describe.
2036: 
2037: 
2038: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2039: \subsection{$N$ flavor extension}
2040: \label{subsec:N_flavor}
2041: 
2042: Purely for technical reasons, then, we generalize the theory to $N$ 
2043: flavors of fermions at the two patches on each Fermi surface, 
2044: $d_{\alpha s} \to d_{i \alpha s}$.
2045: Here $i = 1, \dots, N$ labels the flavor index, while $s = \pm$ labels 
2046: the patch location and $\alpha = 1, 2$ the two Fermion species.  
2047: The action for the fermions is simply generalized as,
2048: \begin{equation}
2049: S_d^{(N)} = \sum_{i=1}^N \sum_{\alpha,s}  \int_{{\bf x} \tau} 
2050: d^\dagger_{i\alpha s} ( \partial_\tau 
2051: - i s v_\alpha \partial_x - \frac{v_\alpha c_\alpha}{2} \partial_y^2 ) 
2052: d_{i\alpha s} ~.
2053: \end{equation}
2054: The real scalar gauge field $a({\bf r})$ is likewise generalized, 
2055: but now as an $N \times N$ Hermitian matrix of complex fields, 
2056: $a_{ij}({\bf r}) = a_{ji}^*({\bf r})$.  When expressed in momentum space,
2057: each component of the matrix, $a_{ij}({\bf q})$, has dynamics
2058: with a Landau damped form,
2059: \begin{equation}
2060: S_a^{(N)} = \frac{1}{2} \sum_{i,j=1}^N \int_{{\bf q} \omega} 
2061: \left(\Gamma \frac{|\omega|}{|q_y|} + \chi |q_y|^2  \right) 
2062: |a_{ij} ({\bf q}, \omega)|^2 ~.
2063: \end{equation}
2064: Since $a_{ij}({\bf q}, \omega) = a_{ji}^*(-{\bf q}, -\omega)$, 
2065: this can also be re-written as a trace over the flavor indices of the 
2066: matrix $\overleftrightarrow{a}({\bf q}, \omega)$:
2067: \begin{equation}
2068: S_a^{(N)} = \frac{1}{2} \int_{{\bf q}\omega} 
2069: \left(\Gamma \frac{|\omega|}{|q_y|} + \chi |q_y|^2  \right) 
2070: {\rm Tr} [ \overleftrightarrow{a}({\bf q}, \omega) 
2071:            \overleftrightarrow{a}({\bf -q}, -\omega) ] ~.
2072: \end{equation}
2073: Finally, the fermion-gauge field interaction is taken as,
2074: \begin{equation}
2075: S^{(N)}_{int} = \sum_{\alpha,s} \frac{g_\alpha}{\sqrt{N}} 
2076: s v_\alpha \sum_{i,j=1}^N \int_{{\bf x} \tau} 
2077: d^\dagger_{i\alpha s} \, a_{ij} \, d_{j \alpha s} ~.
2078: \end{equation}
2079: Notice that the magnitude of the interaction strength (the charge) has
2080: been taken to scale as $1/\sqrt{N}$.
2081: At $N=1$ the theory reduces to that discussed in the previous subsection,
2082: with the assumption of a harmonic Landau damped form for the gauge field 
2083: dynamics as the only approximation.
2084: 
2085: It is worth pointing out that the full action has an enlarged global 
2086: symmetry being invariant under, 
2087: \begin{equation}
2088: d_{i \alpha s} \to e^{i \theta_i} e^{i \phi_\alpha} d_{i \alpha s} ~;
2089: \hskip0.5cm 
2090: a_{ij} \to e^{i(\theta_i - \theta_j)} a_{ij} ~,
2091: \end{equation}
2092: for arbitrary phases $\theta_i$, $i=1, \dots, N$, and $\phi_{\alpha}$, 
2093: $\alpha = 1,2$.  Thus, the densities 
2094: $\sum_\alpha d_{i\alpha}^\dagger d_{i\alpha}$ are independently conserved
2095: for any $i$, and also $\sum_i d_{i\alpha}^\dagger d_{i\alpha}$ are
2096: independently conserved for any $\alpha$.
2097: Note also that the dynamics of each of the $N$ flavor fermions $d_{i1}$ 
2098: is identical to one another, and similarly for all the $d_{i2}$ fermions.
2099: 
2100: 
2101: Before solving the full $N$-flavor theory, 
2102: $S_d^{(N)} + S_a^{(N)} + S_{int}^{(N)}$, exactly at $N=\infty$,  
2103: we define large $N$ generalizations of the various correlation functions.
2104: As in the gauge mean field theory, we anticipate that the exact fermion 
2105: Green's function for (any) one of the $N$ fermion flavors, which we 
2106: denote $G^{(N)}_{d_\alpha}({\bf r})$, will be dominated at large 
2107: distances by the patch fermions with normals parallel or antiparallel
2108: to the observation direction $\hat{\bf r}$.
2109: We can thus expand $G^{(N)}_{d_\alpha}({\bf r})$ for large $|{\bf r}|$ 
2110: in terms of the ``patch" fermion Green's functions as,
2111: \begin{equation}
2112: G^{(N)}_{d_\alpha}({\bf r}, \tau) \approx \sum_s 
2113: e^{i s {\bf k}_{F_\alpha} \cdot {\bf r}} G^{(N)}_{\alpha s}({\bf r},\tau)
2114: ~,
2115: \end{equation}
2116: with the definition,
2117: \begin{equation}
2118: G^{(N)}_{\alpha s} ({\bf r}-{\bf r}^\prime, \tau-\tau^\prime) = 
2119: \la d^\dagger_{i\alpha s}({\bf r}, \tau) 
2120:     d_{i\alpha s}({\bf r}^\prime, \tau^\prime) \ra ~.
2121: \end{equation}
2122: Note that there is no summation over $i$; throughout, any summation over
2123: indices will be shown explicitly.
2124: 
2125: There are now $N$ flavors of bosons with creation operators,
2126: \begin{equation}
2127: b_i^\dagger({\bf r}) = d_{i1}^\dagger({\bf r}) d_{i2}^\dagger({\bf r}) ~,
2128: \end{equation}
2129: with all $N$ boson flavors having identical dynamics.
2130: The large $N$ generalization of the boson correlator is then simply,
2131: \begin{equation}
2132: G_{b;ij}^{(N)} ({\bf r}-{\bf r}^\prime, \tau-\tau^\prime) = 
2133: \la b_i^\dagger({\bf r}, \tau) b_j({\bf r}^\prime, \tau^\prime) \ra ~.
2134: \end{equation}
2135: Due to the global symmetries, the correlators which are off-diagonal in 
2136: the flavor indices vanish, leaving a uniquely defined boson Green's 
2137: function, $G_{b;ij}^{(N)} = \delta_{ij} G_{b}^{(N)}$.  
2138: As desired, $G_b^{(N)}$ reduces at $N=1$ to the boson correlator 
2139: defined in Sec.~\ref{subsec:MF_props}.
2140: 
2141: The boson density-density correlator, $D_b^{(N)}$, is readily defined in 
2142: terms of (any) of the boson densities $\hat\rho_{ib}$ as in 
2143: Eq.~(\ref{densitycorr:def}).  
2144: Similarly, we can define the fermion density-density correlators for 
2145: each species, $D^{(N)}_{d_\alpha}$, in terms of any one of the fermion 
2146: densities, $\hat\rho_{i\alpha}$.
2147: As before, we assume that the boson density-density correlator can be 
2148: approximated as an average over $\alpha$ of the fermion density-density 
2149: correlators, 
2150: $D_b^{(N)} \approx \frac{1}{2} \sum_\alpha D^{(N)}_{d_\alpha}$.
2151: 
2152: 
2153: In order to extract the above correlation functions, it is convenient to
2154: define patch fermion particle-hole and particle-particle bubbles, 
2155: which we denote as $\Pi^{(N)}_{ph}$ and $\Pi^{(N)}_{pp}$, respectively.
2156: The exact bubble correlators can be formally expressed in terms of the 
2157: exact patch fermion Green's functions and the exact particle-hole and 
2158: particle-particle vertices,
2159: \begin{widetext}
2160: \begin{eqnarray}
2161: \Pi^{(N)}_{ph;\, \alpha s; \alpha^\prime s^\prime}({\bf q}, \omega) &=& 
2162: \int_{{\bf q}^\prime \omega^\prime} 
2163: V^{(N)}_{ph;\, \alpha s; \alpha^\prime s^\prime}
2164: ({\bf q} + {\bf q}^\prime, \omega + \omega^\prime; 
2165:  {\bf q}^\prime, \omega^\prime) \,
2166: G^{(N)}_{\alpha s}({\bf q} + {\bf q}^\prime, \omega + \omega^\prime) \,
2167: G^{(N)}_{\alpha^\prime s^\prime}({\bf q}^\prime, \omega^\prime) ~, \\
2168: \Pi^{(N)}_{pp;\, \alpha s; \alpha^\prime s^\prime}({\bf q}, \omega) &=& 
2169: \int_{{\bf q}^\prime \omega^\prime} 
2170: V^{(N)}_{pp;\, \alpha s; \alpha^\prime s^\prime}
2171: ({\bf q} + {\bf q}^\prime, \omega + \omega^\prime; 
2172:  {\bf q}^\prime, \omega^\prime) \,
2173: G^{(N)}_{\alpha s}({\bf q} + {\bf q}^\prime, \omega + \omega^\prime) \,
2174: G^{(N)}_{\alpha^\prime s^\prime}(-{\bf q}^\prime, -\omega^\prime) ~.
2175: \label{bubble;vertex;Greens}
2176: \end{eqnarray}
2177: \end{widetext}
2178: Here $V^{(N)}_{ph}$ and $V^{(N)}_{pp}$ denote the fully renormalized 
2179: vertices, and $\alpha,s$ and $\alpha^\prime,s^\prime$ label the patch 
2180: locations of the two fermion lines that come out of the vertex.  
2181: The total momentum and frequency running through the bubble is 
2182: ${\bf q},\omega$, and is divided between the two fermions as shown above.
2183: At $N=\infty$ there are no vertex corrections, 
2184: $V^{(\infty)}_{ph} = 1$ and $V^{(\infty)}_{pp} = 1$, and each bubble
2185: is just a convolution of the two Green's functions.
2186: 
2187: The boson correlator at large distances and long times can be readily 
2188: expressed in terms of the bubbles,
2189: \begin{equation}
2190: G_b^{(N)}(s {\bf k}_{F_1} + s^\prime {\bf k}_{F_2} + {\bf q}, \omega) 
2191: \approx \Pi^{(N)}_{pp;\, 1s; 2s^\prime}({\bf q}, \omega) ~,
2192: \end{equation}
2193: and similarly for the fermion density-density correlator,
2194: \begin{equation}
2195: D_{d_\alpha}^{(N)}(s {\bf k}_{F_\alpha} - s^\prime {\bf k}_{F_\alpha}
2196:                    + {\bf q}, \omega) 
2197: \approx \Pi^{(N)}_{ph;\, \alpha s; \alpha s^\prime}({\bf q}, \omega) ~.
2198: \end{equation}
2199: 
2200: 
2201: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2202: \subsection{Solution at $N=\infty$ }
2203: \label{subsec:N_infinity}
2204: 
2205: At $N=\infty$ the fermion Green's function 
2206: $G^{(N)}_{\alpha s}({\bf q}, \omega)$ can be readily obtained by 
2207: summing the nested rainbow diagrams to give,
2208: \begin{equation}
2209: G^{(\infty)}_{\alpha s}({\bf q}, \omega)
2210: = \frac{1} {i\omega \left(1 + |\Omega_\alpha /\omega|^{1/3} \right) 
2211:             - s v_\alpha q_x - \frac{v_\alpha c_\alpha}{2} q_y^2 }
2212: \label{N=infinity_Greens}
2213: \end{equation}
2214: with
2215: \begin{equation}
2216: \Omega_\alpha^{1/3} = \frac{v_\alpha}{2 \pi \sqrt{3}} 
2217: \frac{g^2}{\chi^{2/3} \Gamma^{1/3}} ~.
2218: \label{Omega}
2219: \end{equation}
2220: Because of the gauge interactions, each fermion species obtains an 
2221: anomalous self-energy of the form $-i\omega |\Omega/\omega|^{1/3}$ 
2222: which dominates over the bare term $-i\omega$ on energy scales below 
2223: $\Omega$.  Physically, the fermions cease to propagate as free particles 
2224: due to the strong scattering from the dynamical gauge field, 
2225: becoming ``incoherent".
2226: 
2227: We first consider the long distance and time behavior of 
2228: $G^{(\infty)}_{\alpha s}({\bf r},\tau)$.
2229: As in the mean field Sec.~\ref{sec:MFT}, the Green's function is 
2230: dominated at large distances by the patches of the Fermi surface with 
2231: normals along $\pm \hat{\bf r}$.  
2232: For the equal-time Green's function, we obtain,
2233: \begin{equation}
2234: G^{(\infty)}_{d_\alpha} ({\bf r}, 0) \approx \frac{3 \sqrt{\pi}}{4} 
2235: \left( \frac{v_\alpha}{\Omega_\alpha} \right)^{1/2} 
2236: \frac{1}{\sqrt{|{\bf r}|}} G_{d_\alpha}^{MF}({\bf r}) ~,
2237: \end{equation}
2238: where the velocity $v_\alpha$ and curvature $c_\alpha$ characterize the 
2239: $d_\alpha$ Fermi surface patches with normals along $\pm \hat{\bf r}$, 
2240: and $G_{d_\alpha}^{MF}({\bf r})$ is given in Eq.~(\ref{Gd_MF}).
2241: On the other hand, the long time dependence of the local fermion Green's 
2242: function is not affected by the anomalous self-energy and is determined 
2243: solely by the density of states $\nu_0$ at the Fermi energy,
2244: \begin{equation}
2245: G^{(\infty)}_{d_\alpha} ({\bf 0}, \tau) = -\frac{\nu_0}{\tau} ~.
2246: \end{equation}
2247: 
2248: 
2249: Turning to the physical correlators,
2250: at $N=\infty$ there are no vertex corrections to the fermion bubble, 
2251: so we have simply,
2252: \begin{equation}
2253: D^{(\infty)}_{d_\alpha}({\bf r}) = - |G^{(\infty)}_{d_\alpha}({\bf r})|^2
2254: ~,
2255: \end{equation}
2256: which gives for the $N=\infty$ equal-time density correlator,
2257: \begin{equation}
2258: D^{(\infty)}_{b}({\bf r}) \approx \frac{1}{2} \sum_\alpha \frac{9\pi}{16}
2259: \frac{v_\alpha}{\Omega_\alpha}
2260: \frac{1}{|{\bf r}|} D^{MF}_{d_\alpha}({\bf r}) ~,
2261: \end{equation}
2262: where $D^{MF}_{d_\alpha}({\bf r})$ is given in Eq.~(\ref{DD_MF}).
2263: Similarly, in the absence of vertex corrections at $N=\infty$ in the 
2264: particle-particle bubble, one has for the equal-time boson Green's 
2265: function, 
2266: \begin{eqnarray}
2267: G_b^{(\infty)} ({\bf r}) &=& G_{d_1}^{(\infty)} ({\bf r})  
2268: G_{d_2}^{(\infty)} ({\bf r}) \\
2269: &\approx& 
2270: \frac{9\pi}{16}
2271: \left( \frac{v_1 v_2}{\Omega_1 \Omega_2} \right)^{1/2} 
2272: \frac{1}{|{\bf r}|} G_b^{MF}({\bf r}) ~,
2273: \end{eqnarray}
2274: where $G_b^{MF}({\bf r})$ is given in Eq.~(\ref{Gb_MF}).
2275: However, the time dependence of the local boson Green's function 
2276: $G_b({\bf 0}, \tau)$ is unchanged from the mean field Eq.~(\ref{Gtau}).
2277: Finally, the box correlator is given by Eq.(\ref{Bb_MF}) with 
2278: appropriate $N=\infty$ fermion Green's functions and decays
2279: with a power law envelope of $-x^{-8}$.
2280: 
2281: 
2282: Notice that both the boson and the boson density-density correlators
2283: at $N=\infty$ fall off in space as $|{\bf r}|^{-4}$, which is faster 
2284: than their mean field counterparts.  Injecting a boson corresponds to 
2285: creating two fermions.  In the mean field, these fermions propagate 
2286: independently and as free particles.  On the other hand, in the 
2287: present $N=\infty$ theory that sums the nested rainbow diagrams, 
2288: both fermions propagate ``incoherently" due to their scattering from the 
2289: gauge field, which effectively reduces the ability of the created boson 
2290: to propagate, thereby leading to a faster decay of the boson correlator.
2291: At this level of approximation, the two injected fermions, while 
2292: scattered by the gauge fluctuations, do not scatter off one another.  
2293: The effects of interactions between the two injected fermions appear at 
2294: order $1/N$, entering as vertex corrections.  Since the two fermion 
2295: species have opposite gauge charge, one expects that they will attract 
2296: one another.  As we shall see, though, they do not form a composite 
2297: boson (a bound state, for example).
2298: But their motion becomes strongly correlated allowing them to propagate 
2299: more effectively when close by spatially.  The net result, as we find 
2300: below, is that the boson can propagate more effectively when the 
2301: motion of the fermion pair is correlated, leading to a slower decay 
2302: of the correlator compared to the $N=\infty$ result.
2303: Thus, the effect on the boson dynamics due to the decoherence 
2304: experienced by the fermions from scattering off the gauge fluctuations
2305: is compensated, perhaps only partially, by the attractive interaction 
2306: between the pair of fermions.
2307: 
2308: 
2309: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2310: \subsection{Fermion Green's function at order $1/N$}
2311: \label{subsec:1_N_fermion_Greens}
2312: 
2313: The $1/N$ contributions to the fermion Green's function are obtained
2314: by considering the nested rainbow diagrams and replacing one gauge field 
2315: propagator 
2316: $G_a({\bf q}, \omega) = (\Gamma |\omega|/|q_y| + \chi |q_y|^2)^{-1}$ 
2317: with the bubble correction
2318: $\delta G_a \sim \frac{1}{N} G_a^2 \Pi_{ph}^{(\infty)}$.
2319: Formally, we can first calculate the rainbows using $G_a + \delta G_a$
2320: everywhere and then extract the $1/N$ piece.
2321: As we now argue, the functional form of the fermion Green's function 
2322: in Eq.~(\ref{N=infinity_Greens}) remains unchanged except for finite 
2323: $1/N$ shifts in the parameters.
2324: Indeed, evaluating the $N=\infty$ particle/hole bubble at small 
2325: wave vector and frequency gives 
2326: $\Pi^{(\infty)}_{ph}({\bf q}, \omega) \sim |\omega|/|q_y|$, 
2327: which only leads to a finite $1/N$ shift in $\Gamma$.
2328: This is not surprising, since the bare singular gauge propagator
2329: postulated at the outset in our theory is motivated by the free fermion 
2330: particle-hole bubble, and one can verify that the bubble remains unchanged
2331: also in the presence of an arbitrary fermion self-energy that depends 
2332: on the frequency only.
2333: Then, using this renormalized gauge propagator to evaluate the rainbow 
2334: diagram gives a contribution to the fermion Green's function of the form,
2335: $\delta G_{d_\alpha}^{(N)} \sim \frac{1}{N} \omega^{2/3} 
2336: [G^{(\infty)}_{d_\alpha}]^2$.
2337: This will at most give finite $1/N$ shifts to 
2338: $\Omega_\alpha, v_\alpha, c_\alpha$. 
2339: Next we consider the vertices at $1/N$, which undergo more dramatic 
2340: modifications.
2341: 
2342: 
2343: 
2344: 
2345: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2346: \subsection{Vertices at order $1/N$}
2347: \label{subsec:1_N_vertex}
2348: 
2349: \subsubsection{Single photon contribution}
2350: \label{subsubsec:V1a}
2351: 
2352: Consider first dressing the bare vertices with a single gauge propagator.
2353: We denote the value of the corresponding Feynman diagrams as
2354: $V^{(1a)}_{ph}$ and $V^{(1a)}_{pp}$, where $(1a)$ stands for one 
2355: (Landau damped) photon exchanged.  These diagrams are the
2356: first ones in Figs.~\ref{fig:Vph}~and~\ref{fig:Vpp}.
2357: As we will discuss below, the full structure at $1/N$ is more involved, 
2358: but the calculation of $V^{(1a)}$ already captures the main qualitative
2359: physics.
2360: 
2361: For simplicity, we first focus on the situation with zero external $y$ 
2362: momentum and frequency, defining,
2363: \begin{equation}
2364: V^{(1a)}_{q_x^-} = 
2365: V^{(1a)}({\bf q}, \omega; {\bf q}^\prime, \omega^\prime)|_{q_y=q_y^\prime=\omega=\omega^\prime=0} ~,
2366: \end{equation}
2367: where ${\bf q},{\bf q}^\prime$, $\omega,\omega^\prime$ denote 
2368: the momentum/frequencies in and out of the two legs and 
2369: $q_x^- = q_x - q_x^\prime$ is the momentum running through the vertex.  
2370: For notational simplicity we have suppressed the corresponding flavor 
2371: labels, $\alpha,s; \alpha^\prime,s^\prime$.
2372: Evaluating the Feynman diagrams gives,
2373: \begin{eqnarray}
2374: V^{(1a)}_{ph;\, q_x^-} &=& 
2375: \frac{1}{N} \,{\rm sign}(g_\alpha g_{\alpha^\prime})\, \delta_{s, -s^\prime}
2376: \lambda_{ph} \ln(\Lambda_x/|q_x^-|) ~, 
2377: \label{Vph1a}
2378: \\
2379: V^{(1a)}_{pp;\, q_x^-} &=& 
2380: -\frac{1}{N} \,{\rm sign}(g_\alpha g_{\alpha^\prime})\, \delta_{s, -s^\prime}
2381: \lambda_{pp} \ln(\Lambda_x/|q_x^-|) ~,
2382: \label{Vpp1a}
2383: \end{eqnarray}
2384: revealing a logarithmic divergence cut off by the total $x$-momentum 
2385: running through the vertex.
2386: Explicit expressions for the dimensionless non-negative coefficients 
2387: $\lambda_{ph}$ and $\lambda_{pp}$ are given below.
2388: 
2389: Notice that both the particle-hole and the particle-particle 
2390: vertex corrections vanish unless the two patches have opposite group 
2391: velocities.  Moreover, the sign of the correction is given by the 
2392: sign of the product of the two charges and differs for the particle-hole 
2393: and particle-particle channels.  The origin of these signs is essentially
2394: Ampere's law familiar from electrodynamics. 
2395: Two moving charges producing parallel charge currents experience an 
2396: attractive interaction, bringing them closer together and enhancing their
2397: interaction strength, which is encapsulated by the magnitude of the 
2398: interaction vertex. 
2399: For example, a particle and a hole of the same fermion species 
2400: ($\alpha = \alpha^\prime$) residing on the opposite Fermi surface patches
2401: ($s = -s^\prime$) have an enhanced interaction strength, which is
2402: encoded in $V_{ph}^{(1a)} > 0$.
2403: On the other hand, the interaction vertex is suppressed when the
2404: charge currents are anti-parallel.
2405: For example, a pair of particles with the same charge ($g = g^\prime$) 
2406: experience a suppressed vertex interaction, $V_{pp}^{(1a)} < 0$, 
2407: when their group velocities are antiparallel.
2408: 
2409: The absence of singular vertex corrections when the two patch fermions
2410: have parallel group velocities, $s = s^\prime$, is, we believe, not just 
2411: a peculiarity of this lowest order contribution, but will be valid 
2412: generally for the exact vertices at arbitrary $N$.  This expectation is 
2413: based on the specific form of the propagators: the fermion propagator 
2414: has simple poles in the complex $q_x$ plane while the gauge field
2415: propagator is independent of $q_x$, the latter a reflection of the 
2416: transverse nature of the gauge field.
2417: For any given Feynman diagram, it is the sign of the internal frequency
2418: variables running around the loops together with $s, s^\prime$ 
2419: which will determine whether the poles in the integrand will be in the 
2420: upper or the lower complex $q_x$ planes.
2421: For vertices with $s = s^\prime$ all of the fermion lines internal to 
2422: any diagram will have the same group velocity, and with energy 
2423: (frequency) conservation one expects that at least one of the 
2424: $q_x$ variables will only have poles in the upper (or lower) half-plane, 
2425: so that the contour can be appropriately deformed to show that the 
2426: diagram vanishes.  
2427: Thus, we henceforth restrict attention exclusively to vertices with
2428: $s = -s^\prime$, and to ease the notation will drop the explicit
2429: $s, s^\prime$ indices in the following.
2430: 
2431: The dimensionless magnitudes of the vertex enhancements in the 
2432: particle-hole and the particle-particle channel, 
2433: $\lambda_{ph}$ and $\lambda_{pp}$ respectively, are given by,
2434: \begin{eqnarray}
2435: \lambda_{ph} &=& 
2436: \frac{g^2}{\sqrt{3}\pi \Gamma^{1/3} \chi^{2/3}
2437:            \left( \frac{\Omega^{1/3}}{v} 
2438:                  + \frac{\Omega^{\prime 1/3}}{v^\prime} \right)} 
2439: {\cal E}[\zeta_{ph}] ~, 
2440: \label{lambda_ph}
2441: \\
2442: \lambda_{pp} &=& 
2443: \frac{g^2}{\sqrt{3}\pi \Gamma^{1/3} \chi^{2/3}
2444:            \left( \frac{\Omega^{1/3}}{v} 
2445:                  + \frac{\Omega^{\prime 1/3}}{v^\prime} \right)} 
2446: {\cal E}[\zeta_{pp}] ~;
2447: \label{lambda_pp}
2448: \\
2449: \lambda_{ph} &=& {\cal E}[\zeta_{ph}] ~,  \quad\quad
2450: \lambda_{pp} = {\cal E}[\zeta_{pp}] ~.
2451: \label{lambda_Ninfty}
2452: \end{eqnarray}
2453: In the last line, we specialized to the $N=\infty$ expressions for 
2454: $\Omega$ and $\Omega^\prime$ (Eq.~\ref{Omega}),
2455: which is valid for extracting the order $1/N$ vertex corrections.
2456: Here, ${\cal E}[\zeta]$ is a dimensionless function of a dimensionless 
2457: argument given by
2458: \begin{widetext}
2459: \begin{equation}
2460: {\cal E}[\zeta] = \frac{3\sqrt{3}}{2\pi} \int_0^\infty \frac{t dt}{(1+t^3) (1 + \zeta^2 t^4)} 
2461:  =  \frac{3\pi\zeta^5 + 8\pi(1-\zeta^4)/\sqrt{3}
2462: - 3\sqrt{2}\pi(\zeta^{1/2} - \zeta^{7/2}) - 6\zeta^2 \ln\zeta}
2463: {8\pi (1+\zeta^6) / \sqrt{3}} ~,
2464: \label{Ffunc}
2465: \end{equation}
2466: \end{widetext}
2467: and has been normalized so that ${\cal E}[0]=1$.  
2468: This function is monotonically decreasing with increasing argument
2469: varying as ${\cal E}[\zeta] \approx \frac{3 \sqrt{3}}{8 \zeta}$ for 
2470: $\zeta \to \infty$.
2471: The parameters $\zeta_{ph}$ and $\zeta_{pp}$ are given by,
2472: \begin{eqnarray}
2473: \zeta_{ph} &=& 
2474: \frac{ \Gamma^{2/3} (c + c^\prime)}
2475:      { 2 \chi^{2/3}
2476:        \left( \frac{\Omega^{1/3}}{v} 
2477:              + \frac{\Omega^{\prime 1/3}}{v^\prime} \right) } ~, 
2478: \label{zeta_ph} \\
2479: \zeta_{pp} &=& 
2480: \frac{ \Gamma^{2/3} |c - c^\prime|}
2481:      { 2 \chi^{2/3}
2482:        \left( \frac{\Omega^{1/3}}{v} 
2483:              + \frac{\Omega^{\prime 1/3}}{v^\prime} \right) } ~;
2484: \label{zeta_pp} \\
2485: \zeta_{ph} &=& \frac{\sqrt{3}\pi}{2 g^2} \Gamma (c + c^\prime) ~, 
2486: \quad\;
2487: \zeta_{pp} = \frac{\sqrt{3}\pi}{2 g^2} \Gamma |c - c^\prime| ~.
2488: \label{zeta_Ninfty}
2489: \end{eqnarray}
2490: In the last line, we again used the $N=\infty$ values for 
2491: $\Omega,\Omega^\prime$.
2492: Notice that the vertex interaction strength between two fermions on 
2493: opposing patches depends on their particular Fermi surface curvatures 
2494: $c, c^\prime$.  Moreover, in the particle-particle channel the 
2495: interaction strength is maximal when the curvatures are equal.  
2496: This reflects a particle-particle nesting, which is responsible for the 
2497: BCS pairing instability in the Fermi liquid context.
2498: 
2499: 
2500: \begin{figure}
2501: \centerline{\includegraphics[width=2.0in]{figures/Vph.eps}}
2502: \vskip -2mm
2503: \caption{
2504: Ladder diagrams that contribute to $V_{ph}$ at $1/N$.
2505: }
2506: \label{fig:Vph}
2507: \end{figure}
2508: 
2509: 
2510: \begin{figure}
2511: \centerline{\includegraphics[width=2.0in]{figures/Vpp.eps}}
2512: \vskip -2mm
2513: \caption{
2514: Crossed diagrams that contribute to $V_{pp}$ at $1/N$.
2515: }
2516: \label{fig:Vpp}
2517: \end{figure}
2518: 
2519: 
2520: \vskip 2mm
2521: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2522: \subsubsection{Photon ladders}
2523: 
2524: Having evaluated the one-photon corrections $V^{(1a)}$ to the bare 
2525: vertices, we now discuss the full set of diagrams that appear at 
2526: order $1/N$ in each case.
2527: First consider the fermion particle-hole vertex which enters into the
2528: density-density correlator: 
2529: $V^{(N)}_{\rho_\alpha} = V_{ph;\, \alpha, \alpha}^{(N)}$.
2530: The $1/N$ diagrams are shown in Fig.~\ref{fig:Vph} and contain an 
2531: arbitrary number of non-crossing photon lines,\cite{footnote:N_flavor} 
2532: beginning with the one-photon diagram $V_{ph}^{(1a)}$, Eq.~(\ref{Vph1a}).  
2533: We shall denote the contribution from the diagram with $n$ photon lines 
2534: as $V_{ph}^{(na)}$.  The magnitude of the singular contributions from 
2535: each diagram in the ladder sum will depend on a dimensionless gauge 
2536: charge,
2537: \begin{equation}
2538: e_\alpha^2 \equiv g^2/\Gamma c_\alpha ~.
2539: \end{equation}
2540: Above we found
2541: $V_{ph}^{(1a)} = (1/N) {\cal E}[\sqrt{3}\pi/e_\alpha^2] \ln(1/q_x^-)
2542: \equiv (1/N) A^{(1)}_1 \ln(1/q_x^-)$.
2543: The singular contributions from the diagram with two photon lines 
2544: can also be extracted from the Feynman diagram,
2545: $V_{ph}^{(2a)} = (1/N) 
2546: [ A^{(2)}_1 \ln(1/q_x^-) + A^{(2)}_2 \ln^2(1/q_x^-) ]$, 
2547: with $A^{(2)}_1, A^{(2)}_2$ universal dimensionless functions
2548: of $e_\alpha^2$.
2549: From this we infer the general structure for the diagram with $n$ 
2550: photon lines,
2551: \begin{equation}
2552: V_{ph}^{(na)} \sim \frac{1}{N} \sum_{m=1}^n A^{(n)}_m \ln^m(1/q_x^-) ~.
2553: \end{equation}
2554: This form suggests that the full set of $1/N$ ladder diagrams can
2555: be exponentiated,
2556: \begin{equation}
2557: V_{\rho_\alpha}^{(N)} 
2558: \sim \frac{1}{N} \exp[\gamma_\alpha \ln(1/q_x^-)] 
2559: \sim \frac{1}{N} |q_x - q_x^\prime|^{-\gamma_\alpha} ~,
2560: \label{ph_ladder_sum}
2561: \end{equation}
2562: with $\gamma_\alpha = \sum_{n=1}^\infty A^{(n)}_1$.
2563: The validity of this exponentiation can be justified by formally 
2564: summing the ladder series to obtain an integral equation for the full 
2565: vertex.  The integral equation has a singular kernel and its solution is 
2566: expected to have the power-law form with some exponent $\gamma_\alpha$ 
2567: which will depend on the dimensionless charge $e_\alpha^2$.
2568: 
2569: Unfortunately, we have been unable to solve the integral equation
2570: to obtain $\gamma_\alpha$.
2571: However, we observe that for small charge $e_\alpha^2$,
2572: $A^{(1)}_1 = 3 e_\alpha^2 / (8\pi) + O(e_\alpha^4)$.
2573: This gives the leading behavior even though $A^{(1)}_1$ is not analytic 
2574: in $e_\alpha^2$, as can be seen from Eq.~(\ref{Ffunc});
2575: more generally, one finds $A^{(n)}_m = O(e_\alpha^{2n})$.
2576: Thus, assuming small parameter $e_\alpha^2$, the leading contribution to 
2577: $\gamma_\alpha$ comes just from the single photon diagram,
2578: \begin{equation}
2579: \gamma_\alpha = \frac{3}{8\pi} e_\alpha^2 + O(e_\alpha^4, 1/N) ~.
2580: \label{anomalous;density;exponent}
2581: \end{equation}
2582: 
2583: For $e_\alpha^2$ of order one, the exponent $\gamma_\alpha$, while 
2584: obtained by summing diagrams at order $1/N$, is itself of order one.
2585: It is the amplitude of the vertex in Eq.~(\ref{ph_ladder_sum})
2586: which is of order $1/N$.
2587: This reflects the particular structure of the large $N$ theory,
2588: which, constructed to sum the nested rainbows for the fermion self energy
2589: at leading order, also exponentiates the ladder sum for the 
2590: particle-hole vertex at order $1/N$.
2591: Notice that $\gamma_\alpha$ monotonically decreases as the curvature
2592: of the Fermi surface is taken large, due to the increasingly restrictive 
2593: phase space requirements on the particle-hole vertex in this limit.  
2594: We expect that this trend is valid more generally, for larger 
2595: $e_\alpha^2$ and for arbitrary $N$.
2596: 
2597: 
2598: Summarizing the discussion of the particle-hole vertex, we can also 
2599: write a scaling form for more general external momenta and frequencies,
2600: \begin{equation}
2601: V^{(N)}_{\rho_\alpha} \sim |q_x-q_x^\prime|^{-\gamma_\alpha}
2602: \tilde{V}_{\rho_\alpha} 
2603: \left(\frac{q_y^2}{q_x^-}, \frac{q_y^{\prime 2}}{q_x^-},
2604:       \frac{\omega^{2/3}}{q_x^-}, \frac{\omega^{\prime 2/3}}{q_x^-} 
2605: \right) ~.
2606: \label{Vrho;scaling}
2607: \end{equation}
2608: 
2609: 
2610: \vskip 2mm
2611: 
2612: We next consider the vertex which enters in the boson-boson correlator, 
2613: $V_b^{(N)} = V^{(N)}_{pp;\, 1,2}$.
2614: The full set of diagrams that contribute to the particle-particle vertex 
2615: are shown in Fig.~\ref{fig:Vpp} and contain an arbitrary number of 
2616: maximally-crossed photon lines.\cite{footnote:N_flavor}
2617: The first term is the one-photon diagram that we evaluated in 
2618: Eq.~(\ref{Vpp1a}), 
2619: $V_{pp;\, 1,2}^{(1a)} = (1/N) {\cal E}[\zeta_{pp}] \ln(1/q_x^-)
2620: \equiv (1/N) B_1 \ln(1/q_x^-)$
2621: (plus non-singular contributions).
2622: An evaluation of the two-photon diagram gives the same form,
2623: $V_{pp;\, 1,2}^{(2a)} = (1/N) B_2 \ln(q_x^-)$.
2624: In contrast to the particle-hole case, here the two-photon diagram 
2625: does not contain a log-squared term, which indicates that the crossed 
2626: ladder series cannot be exponentiated.   While we cannot evaluate the 
2627: higher order diagrams, we expect that each term will have a log 
2628: contribution with some dimensionless amplitudes $B_n$.
2629: If this is the case, the full structure of the particle-particle vertex 
2630: up to order $1/N$ is,
2631: \begin{equation}
2632: V_b^{(N)} = 1 + \frac{\eta_1 \ln(\Lambda_x/q_x^-)}{N} + O(1/N^2) ~,
2633: \label{bosonvertex_1/N}
2634: \end{equation}
2635: with $\eta_1 = \sum_{n=1}^\infty B_n$.
2636: To check whether the latter series can be defined in principle,
2637: we again take the limit of small dimensionless charges $e_\alpha^2$.
2638: Consider the least favorable case with matched Fermi surface curvatures,
2639: $c_1 = c_2$, so $e_1^2 = e_2^2 \equiv e^2 = g^2/(\Gamma c)$.
2640: We find that the first term is independent of $e^2$,
2641: $B_1 = 1$.  However, evaluating the two-photon diagram gives
2642: $B_2 \sim (e^2)^2 \ln^3(1/e^2)$ for small $e^2$.
2643: The diagrams are expected to be smaller when $c_1 \neq c_2$
2644: (for example, even the first term $B_1 = {\cal E}[\zeta_{pp}]$
2645: can be made small when $g^2 \ll \Gamma |c_1 - c_2|$).
2646: We thus conjecture that the series converges for small enough
2647: $e_\alpha^2$.
2648: 
2649: Provided it is legitimate to exchange the order of limits $N \to \infty$ 
2650: and $q_x \to 0$ in Eq.~(\ref{bosonvertex_1/N}),
2651: the  boson vertex can be written in a power law form,
2652: \begin{equation}
2653: V_b^{(N)}({\bf q}, \omega; {\bf q}^\prime, \omega^\prime)
2654: |_{q_y=q_y^\prime=\omega=\omega^\prime=0} 
2655: \sim |q_x - q_x^\prime|^{-\eta} ~,
2656: \end{equation}
2657: with
2658: \begin{equation}
2659: \eta = \eta_1/N + O(1/N^2) ~.
2660: \label{eta_1/N}
2661: \end{equation}  
2662: We will discuss the legitimacy of this exponentiation procedure in 
2663: Sec.~\ref{sec:eff_FT}.
2664: For more general external momenta and frequencies the full boson 
2665: vertex will satisfy a scaling form,
2666: \begin{equation}
2667: V_b^{(N)}({\bf q}, \omega; {\bf q}^\prime, \omega^\prime) \sim  
2668: |q_x - q_x^\prime|^{-\eta} \,
2669: \tilde{V}_b \left(\frac{q_y^2}{q_x^{-}}, \frac{q_y^{\prime 2}}{q_x^-},
2670:                   \frac{\omega^{2/3}}{q_x^-}, 
2671:                   \frac{\omega^{\prime 2/3}}{q_x^-} \right) ~,
2672: \label{Vb;scaling}
2673: \end{equation}
2674: where we have used the fact that the vertex depends only on the
2675: total $x$-momentum running through it, $q_x^- = q_x - q_x^\prime$.
2676: The scaling function will vary as, $\tilde{V}_b(X,0,0,0) \sim X^{-\eta}$
2677: when any of its arguments $X \to \infty$.  For example, one has,
2678: $V_b \sim |q_y|^{-2\eta} \sim |\omega|^{-2\eta/3}$.
2679: 
2680: As will be detailed shortly, the singular vertices
2681: Eq.~(\ref{Vrho;scaling}) and Eq.~(\ref{Vb;scaling}) will lead to 
2682: anomalous decay exponents for the boson density-density correlator
2683: and the boson Green's function.
2684:  
2685: 
2686: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2687: \subsection{ Boson and density correlators at finite $N$}
2688: \label{subsec:1_N_corr}
2689: 
2690: Since there are no singular $1/N$ corrections to the fermion 
2691: Green's functions, to extract the boson and density correlators up to 
2692: order $1/N$ one can use the $N=\infty$ fermion Green's functions when 
2693: evaluating the bubble diagram.
2694: Using Eq.~(\ref{N=infinity_Greens}) and the scaling form 
2695: Eq.~(\ref{Vb;scaling}) for the $V_b$ vertex 
2696: enables the particle-particle bubble to be expressed as,
2697: \begin{eqnarray*}
2698: \Pi^{(N)}_{pp} ({\bf q},\omega) \sim |q_x|^{-(2+\eta)}
2699: \int_{{\bf q}^\prime \omega^\prime}
2700: P \left( \frac{q_y^2}{q_x}, \frac{\omega^{2/3}}{q_x},
2701:          \frac{q_x^\prime}{q_x}, 
2702:          \frac{q_y^{\prime 2}}{q_x},\frac{\omega^{\prime 2/3}}{q_x} 
2703: \right) ~,
2704: \end{eqnarray*} 
2705: where $P$ is an appropriate scaling function.
2706: One can then scale out the $q_x$ dependence under the integrals 
2707: to obtain,
2708: \begin{equation}
2709: \Pi^{(N)}_{pp}({\bf q},\omega) \sim |q_x|^{1-\eta} \,
2710: \tilde{\Pi}_{pp}\left(\frac{q_y^2}{q_x}, \frac{\omega^{2/3}}{q_x} \right)
2711: ~.
2712: \end{equation}
2713: We thereby arrive at the desired expression for the boson correlator,
2714: \begin{equation}
2715: G_b^{(N)}({\bf k}_{F_1} - {\bf k}_{F_2} + {\bf q}, \omega) 
2716: \sim |q_x|^{1-\eta} 
2717: \tilde{G} \left( \frac{q_y^2}{q_x}, \frac{\omega^{2/3}}{q_x} \right) ~,
2718: \label{Gb_scaling}
2719: \end{equation}
2720: with exponent $\eta$ given in Eq.~(\ref{eta_1/N}).
2721: While this result was obtained systematically to order $1/N$, 
2722: we expect it will remain valid more generally, and in particular will 
2723: involve the same scaling combinations $q_y^2/q_x$, $\omega^{2/3}/q_x$.
2724: The scaling function $\tilde{G}$ and the dependence of the exponent 
2725: $\eta$ on the bare parameters such as the Fermi surface curvatures will 
2726: presumably vary with $N$.
2727: Near the other momentum ${\bf k}_{F_1} + {\bf k}_{F_2}$, 
2728: the boson correlator will satisfy a similar scaling form except with 
2729: $\eta=0$.
2730: As argued before, we expect that the anomalous exponent at momentum
2731: ${\bf k}_{F_1} +{\bf k}_{F_2}$ will vanish in general,
2732: not just to leading order in $1/N$.
2733: 
2734: The large distance behavior of the equal-time boson Green's function in 
2735: the DBL follows directly from Eq.~(\ref{Gb_scaling}) and was already 
2736: listed in the Introduction, Eq.~(\ref{Gb_final}).
2737: We remark that simple ``power counting'' using Eq.~(\ref{Gb_scaling}) 
2738: works here because the relevant frequencies and momenta are indeed in 
2739: the ``scaling regime''; however, it should be kept in mind that we are 
2740: dealing with line singularities and should exercise more caution with 
2741: such arguments in general.
2742: Once again, both the Fermi wave vectors ${\bf k}_{F_1}, {\bf k}_{F_2}$ 
2743: and the scaling exponent $\eta$ will depend on the location and shape of 
2744: the Fermi surface patches with normals along the observation direction 
2745: $\hat{\bf r}$.   At wave vector ${\bf k}_{F_1} - {\bf k}_{F_2}$, the 
2746: Green's function decay becomes more slow due to the Amperean attraction 
2747: between the $d_1$ and $d_2$ fermions.
2748: The fermions tend to move as pairs, but unlike a Cooper pair their 
2749: motion is not phase coherent -- the pairs are uncondensed.
2750: The boson correlator, while decaying more slowly with positive $\eta$,
2751: does not exhibit ODLRO in the DBL phase.
2752: 
2753: As we found to leading order in $1/N$, Eq.~(\ref{eta_1/N}), $\eta$ is 
2754: expected to be largest when the two Fermi surface curvatures are equal.
2755: In the DBL phase which has closed Fermi surfaces, the characteristics
2756: of the $d_1$ and $d_2$ patches with normals along either diagonal 
2757: $\hat{\bf x} \pm \hat{\bf y}$ are always equal due to the symmetry of 
2758: the square lattice.   Moreover, the decay of the boson Green's function 
2759: $G_b({\bf r})$ will be non-oscillatory along the diagonals,
2760: \begin{equation}
2761: G_b({\bf r}) \sim \frac{1}{|{\bf r}|^{4-\eta}} ~, 
2762: \hskip0.5cm \hat{\bf r} \parallel (\hat{\bf x} \pm \hat{\bf y}) ~,
2763: \end{equation}
2764: a prediction that we find consistent with the properties of the 
2765: $(\det)_x (\det)_y$ wavefunctions studied in Sec.~\ref{sec:Wavefnc_props}
2766: (see Fig.~\ref{fig:real-space-11}).
2767: For $c_1=c_2$, a very rough estimate can be extracted from the leading 
2768: behavior, $\eta = (1/N) + O(e^2/N)$ by setting $N=e^2=1$ giving 
2769: $\eta \approx 1$.
2770: This would shift the $N=\infty$ decay exponent back to its mean field 
2771: value of $1/|{\bf r}|^3$.   
2772: As noted earlier, the gauge fluctuations encapsulated already at 
2773: $N=\infty$ strongly modify the motion of each fermion leading to a 
2774: boson that moves less coherently decaying with a larger exponent than 
2775: in the mean field theory.  The Amperean attraction between the pair of 
2776: fermions which enters at order $1/N$ compensates this effect, 
2777: leading to a slower decay of the boson correlator.  
2778: For $\eta=1$ the two competing effects exactly compensate one another.
2779: 
2780: 
2781: Next we consider the density-density correlator.
2782: Just as for the boson correlator, the density correlator will
2783: satisfy a scaling form:
2784: \begin{equation}
2785: D_{d_\alpha}^{(N)}(2{\bf k}_{F_\alpha} + {\bf q}, \omega) 
2786: \sim |q_x|^{1-\gamma_\alpha} 
2787: \tilde{D}_{2{\bf k}_{F_\alpha}} 
2788: \left( \frac{q_y^2}{q_x}, \frac{\omega^{2/3}}{q_x} \right) ~,
2789: \end{equation}
2790: with an anomalous exponent that depends on the Fermi surface patches,
2791: see Eq.~(\ref{anomalous;density;exponent}).
2792: Back in real space, the dominant density correlator oscillates with 
2793: wave vector $2{\bf k}_{F_\alpha}$, with an envelope decaying as a power 
2794: law with the anomalous exponent,
2795: \begin{equation}
2796: D_{d_\alpha}^{(N)}({\bf r}) \sim 
2797: - \frac{\cos[2{\bf k}_{F_\alpha} \cdot {\bf r} - 3\pi/2]}
2798:        {|{\bf r}|^{4-\gamma_\alpha}}
2799: - \frac{1}{|{\bf r}|^4} ~.
2800: \end{equation}
2801: We have also indicated that the zero momentum component is
2802: not modified relative to the $N=\infty$ behavior.
2803: Again, both ${\bf k}_{F_\alpha}$ and $\gamma_\alpha$ are particular to 
2804: the Fermi surface patches with normals parallel to $\pm \hat{\bf r}$.
2805: In contrast to the particle-particle channel, the particle-hole channel 
2806: is not nested so that we suspect the exponents $\gamma_\alpha$ 
2807: will be smaller than $\eta$ for the physically relevant case
2808: with $N=1$ . 
2809: 
2810: In sum, at leading order in our systematic $1/N$ expansion, the boson 
2811: correlator was found to exhibit power law decay with an exponent that 
2812: depends on the bare Fermi surface curvatures, varying continuously 
2813: around the Fermi surface.
2814: Similar behavior was found for the density-density correlator.
2815: Provided this qualitative behavior persists down to the physically 
2816: relevant case of $N=1$, we conclude that the DBL phase is described 
2817: by a manifold of scale invariant theories rather than an isolated 
2818: fixed point.  Within the $1/N$ expansion, the power law form obtained 
2819: for the boson correlator relied on the exponentiation of the logarithmic 
2820: behavior on momentum and frequency of the leading $1/N$ correction.  
2821: In the next section, we check the legitimacy of this procedure 
2822: by revisiting the renormalization group approach developed earlier
2823: to describe the low energy physics of a sea of fermions coupled
2824: to a $U(1)$ gauge field.  
2825: 
2826: We conclude this section by pointing out that throughout we worked
2827: on the assumption of having both $d_1$ and $d_2$ parallel patches
2828: present as in Fig.~\ref{fig:patches}.  In the case with open Fermi 
2829: surfaces shown in the right panel of Fig.~\ref{fig:introFS}, 
2830: we encounter situations when for a given observation direction 
2831: $\hat{\bf r}$ only one Fermi surface (or even none at all) has patches 
2832: with normals in this direction.
2833: In such cases, the boson correlator will decay exponentially; 
2834: there is no inter-species gauge interactions (one species is simply
2835: absent), but the fermions that are present are still strongly 
2836: affected by the gauge field, and in particular the preceding analysis of 
2837: their density correlations remains the same.
2838: In the DLBL phase, there are no parallel patches on the entire two Fermi 
2839: surfaces, so the $d_1$ and $d_2$ fermions effectively decouple from 
2840: each other.  It is also interesting to remark that
2841: in the limit of extreme D-eccentricity when the Fermi surfaces
2842: are completely flat, we expect that the relevant gauge field
2843: is very strongly damped.  This can be seen, e.g., from the naive
2844: RPA approximation for the Landau damping coefficient $\Gamma$, 
2845: Eq.~(\ref{Gamma_RPA}), in the limit of vanishing curvatures.
2846: In this limit, the energy scale $\Omega$, Eq.~(\ref{Omega}), 
2847: below which fermions become incoherent, goes to zero, 
2848: which leads us to speculate that perhaps in this case the 
2849: $d_1$ and $d_2$ fermions behave as essentially free.
2850: 
2851: 
2852: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2853: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2854: \section{Effective field theory for the DBL: Criticality and Stability}
2855: \label{sec:eff_FT}
2856: 
2857: Polchinski\cite{Polchinski} and others\cite{Altshuler, Nayak} 
2858: argued that 2d fermions interacting with a $U(1)$ gauge field can be 
2859: fruitfully studied within a renormalization group analysis of a 
2860: particular effective field theory.  
2861: In the present case, the fixed point theory has an action given by,
2862: \begin{equation}
2863: S_{\rm fixed\, point} = S_d + S_a + S_{int} ~,
2864: \end{equation}
2865: with
2866: \begin{widetext}
2867: \begin{equation}
2868: S_d = \sum_{\alpha s} \int_{{\bf q}\omega} 
2869: d^\dagger_{\alpha s}({\bf q},\omega) 
2870: \left( -i \omega \left|\frac{\Omega_\alpha}{\omega}\right|^{1/3} 
2871:        + s v_\alpha q_x + \frac{v_\alpha c_\alpha}{2} q_y^2 
2872: \right) d_{\alpha s}({\bf q},\omega) ~,
2873: \end{equation}
2874: \end{widetext}
2875: while the gauge field part $S_a$ is given in Eq.~(\ref{Sa}) and
2876: the fermion-gauge coupling is specified in Eq.~(\ref{Lint}).
2877: Within this fixed point ansatz, both the gauge field and the fermions 
2878: have singular propagators, but their interaction is local.
2879: The advantage of this effective field theory is that it is possible to 
2880: define a simple renormalization group transformation which leaves the 
2881: Gaussian part of the theory invariant, and under which the interaction 
2882: strength $g$ is a marginal perturbation.  If $g$ is assumed to be small,
2883: it can be treated via a conventional perturbative RG approach as we 
2884: describe below.  One finds that the full effective field theory is 
2885: invariant under the RG, i.e., it is at a fixed point.  This strongly 
2886: suggests that exponentiating the logarithmic singularities present in the
2887: large $N$ approach, as we did to order $1/N$ above, is a correct 
2888: procedure at all orders in $1/N$.  If it is, then at order $1/N^2$ 
2889: one must find $(\log)^2$ singularities with particular coefficients
2890: such that the exponentiation procedure to obtain a power law behavior 
2891: is consistent.
2892: 
2893: An unsettling drawback with this effective field theory approach is that 
2894: it leaves unclear what constraints must be placed on additional 
2895: interactions that can be added to the theory.
2896: At the very least these interactions must be gauge invariant and 
2897: consistent with momentum conservation, but is it legitimate to require 
2898: that the interactions be local in the fermion fields?  
2899: For example, can one require that four-fermion interactions be local 
2900: with non-singular coefficients, or will singular interactions necessarily
2901: be generated?  Indeed, if one were to formally integrate out the gauge 
2902: field with its singular propagator, one would generate four-fermion 
2903: interactions with a particular singular form.
2904: In what follows we will ignore these subtleties, exploring the possible 
2905: perturbative instabilities driven by non-singular fermion interactions.  
2906: Specifically, we will consider all non-singular quartic interactions 
2907: involving four fermions living near the Fermi surfaces that are 
2908: consistent with the relevant symmetries, most importantly momentum 
2909: conservation.
2910: 
2911: The fixed point action, as it stands, has 9 parameters,
2912: \begin{eqnarray}
2913: \Gamma, \chi, v_\alpha, c_\alpha, \Omega_\alpha, g ~.
2914: \label{all_params}
2915: \end{eqnarray}
2916: But with an appropriate rescaling of the fermion fields and the gauge 
2917: field together with the momenta and frequency, it is possible to set 
2918: 5 of these parameters to unity, 
2919: $\tilde{\Gamma} = \tilde{\chi} = \tilde{g} = \tilde{v}_\alpha = 1$.  
2920: The effective field theory is then specified by 4 dimensionless 
2921: parameters, $\tilde{c}_\alpha = c_\alpha \Gamma / g^2$ and 
2922: $\tilde{\Omega}_\alpha = \Omega_\alpha \chi^2 \Gamma / (g^2 v_\alpha^3)$.
2923: Evidently, this theory is not describing a fixed ``point" per se,
2924: but constitutes a four-dimensional manifold of theories which are 
2925: invariant under the RG.  Establishing definitively that a particular 
2926: bare (lattice) gauge theory Hamiltonian is attracted to this manifold is 
2927: exceedingly difficult.  Arguably, it is even harder to deduce where on 
2928: this manifold the theory flows.  
2929: The values of the dimensionless parameters which are obtained from the 
2930: leading large $N$ analysis can, nevertheless, be used as a rough guide
2931: in addressing both questions.
2932: 
2933: The RG analysis of the fermion-gauge action, $S_{\rm fixed\, point}$, 
2934: proceeds as follows.  At each stage, the fermion fields reside in the 
2935: appropriate momentum space patches 
2936: $|q_x| < \Lambda$, $|q_y| < (\Lambda/c)^{1/2}$,
2937: where $\Lambda$ is the shell width in the direction normal to the Fermi 
2938: surface.  The corresponding restriction on the frequency is
2939: $|\omega| < (v \Lambda)^{3/2} / \Omega^{1/2}$.
2940: The gauge fields reside in similar momentum-frequency regions but 
2941: centered around zero momentum, and the overall setup is illustrated in 
2942: Fig.~\ref{fig:patches}.
2943: If the dimensionless parameters exhibited earlier are of order one
2944: (which is the case if the parameters are taken from the large-$N$ 
2945: analysis), the corresponding regions are roughly similar for all fermion
2946: and gauge fields.  
2947: Also, in practical RG calculations beyond the tree level it is 
2948: convenient to keep the cutoff only on the frequencies and perform 
2949: unrestricted integrations over the momenta.
2950: 
2951: We integrate out the high-energy fields from the shell between 
2952: $\Lambda$ and $\Lambda/b$ and then rescale the momenta and
2953: frequencies in order to recover the initial cutoff:
2954: $q_x = q_x'/b$, $q_y = q_y'/b^{1/2}$, and $\omega = \omega'/b^{3/2}$.
2955: We also perform appropriate rescaling of the fields:
2956: $d = b^2 d'$, $a = b^2 a'$.
2957: Upon such tree-level scaling, the fixed-point action remains 
2958: exactly as before, i.e., all couplings remain unchanged.
2959: It is also useful to bear in mind that even though we nominally
2960: restore all cutoffs as we proceed, in terms of the original momenta 
2961: the patches become more and more elongated in the $\hat{\bf y}$ 
2962: direction in Fig.~\ref{fig:patches}.  
2963: In particular, any overlap between two patches that correspond to nearby 
2964: but non-parallel tangents $\hat{\bf y}$ and $\hat{\bf y}'$ 
2965: goes to zero in the low-energy limit.
2966: 
2967: To proceed with the analysis beyond the tree level, we will work
2968: perturbatively in the (dimensionless) fermion-gauge coupling, 
2969: assuming that it is small.
2970: Quite generally, since the momentum shell RG cannot produce terms that 
2971: are singular at small frequencies and momenta, the couplings 
2972: $\Omega_\alpha$ and $\Gamma$ will not renormalize at any order in the 
2973: perturbation expansion.  On the other hand, the parameters 
2974: $v_\alpha$, $c_\alpha$, $\chi$, and $g$, can potentially flow.
2975: However, a direct examination shows that they do not renormalize at the 
2976: lowest (one-loop) order, and we strongly suspect that this remains true 
2977: at all orders.  The implication is that the fermion-gauge interaction is 
2978: exactly marginal and that $S_{\rm fixed\, point}$ describes a 
2979: manifold of RG fixed points parameterized by 4 dimensionless couplings.
2980: Although the fermion-gauge coupling need not be small, we expect that 
2981: this RG-invariant manifold will extend outside the perturbatively 
2982: accessible regime.
2983: This manifold describes the putative DBL phase.
2984: 
2985: In order to establish the stability of the DBL, however, we need to 
2986: consider the effects of all symmetry allowed perturbations that can be 
2987: added to $S_{\rm fixed\, point}$.
2988: Stability requires that all such perturbations are irrelevant 
2989: under the RG.  Restricting ourselves to local terms, we focus now on the 
2990: four-fermion interactions which are most likely to destabilize the 
2991: fixed point manifold.
2992: 
2993: 
2994: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2995: \subsection{Short-range fermion interactions}
2996: \label{subsec:4_fermion}
2997: Let us consider a general quartic term,
2998: \begin{equation}
2999: W\; d^\dagger_{\alpha_1 s_1} d^\dagger_{\alpha_2 s_2}
3000:     d_{\alpha_3 s_3} d_{\alpha_4 s_4} ~,
3001: \end{equation}
3002: which can contain different fermion species labelled by $\alpha$ and 
3003: involve different patches labelled by $s$.
3004: The fermion/patch labels are assumed to be such as to satisfy the 
3005: momentum conservation.  Also, we do not show explicitly the (conserved) 
3006: frequency and momentum of the fermion fields.
3007: The amplitude for this four-fermion interaction, $W$, measures the 
3008: strength of a direct scattering between two fermions. 
3009: 
3010: At the tree level in the RG, all such short-range interactions
3011: are irrelevant,
3012: \begin{equation}
3013: W \to W' = W/b ~,
3014: \label{Wtree}
3015: \end{equation}
3016: scaling towards zero at low energies.  Physically, the incoherent motion 
3017: of the fermions, manifest in their $\omega^{2/3}$ self energy, leads to 
3018: a reduced phase space for direct two-body interactions relative to the 
3019: case for free fermions. 
3020: 
3021: 
3022: \begin{figure}
3023: \centerline{\includegraphics[width=2.5in]{figures/4ferm.eps}}
3024: \vskip -2mm
3025: \caption{One-loop diagrams needed to calculate renormalization
3026: of four-fermion interactions.
3027: }
3028: \label{fig:4ferm}
3029: \end{figure}
3030: 
3031: 
3032: At one loop, a gauge propagator can connect any two legs of the quartic 
3033: vertex which involve fermions located on patches with parallel or 
3034: antiparallel Fermi surface normals.
3035: Two such physically distinct diagrams are shown in Fig.~\ref{fig:4ferm}.
3036: Both processes correspond to an interaction between the two fermions
3037: mediated by the Landau damped gauge field.  The one-loop integrals
3038: needed to evaluate these diagrams are identical to those for the 
3039: vertex corrections in subsection~\ref{subsubsec:V1a}, and give the 
3040: following contributions to the four-fermion interaction before any 
3041: rescaling:
3042: \begin{eqnarray}
3043: \text{a):} \quad
3044: \delta W &=& \text{sign}(g_\alpha g_{\alpha^\prime})
3045: \delta_{s, -s^\prime} \lambda_{ph} \, W \, \ln(b)  ~,
3046: \label{Wph}
3047: \\
3048: \text{b):} \quad
3049: \delta W &=& - \text{sign}(g_\alpha g_{\alpha^\prime})
3050: \delta_{s, -s^\prime} \lambda_{pp} \, W \, \ln(b)  ~.
3051: \label{Wpp}
3052: \end{eqnarray}
3053: Here, the primed and unprimed parameters refer to the possibly distinct 
3054: species of the two fermions connected by the gauge propagator.
3055: The anomalous exponents $\lambda_{ph}$ and $\lambda_{pp}$ are given in 
3056: Eqs.~(\ref{lambda_ph})~and~(\ref{lambda_pp})
3057: -- these formulas are general and do not assume any relations among the 
3058: parameters, as appropriate in the present RG setup; 
3059: sometimes, which will be indicated explicitly, we will use the 
3060: $N=\infty$ expressions Eq.~(\ref{lambda_Ninfty}) to get a crude 
3061: estimate of the numerical values.
3062: 
3063: 
3064: Similarly to the analysis in Sec.~\ref{subsubsec:V1a},
3065: either diagram gives zero if the fermions involved have 
3066: parallel group velocities, whereas a nontrivial contribution is found 
3067: when the two fermion fields reside on the opposite patches with 
3068: antiparallel group velocities.
3069: The sign is again determined by Ampere's law:  Two fermions emerging 
3070: from a scattering process with parallel charge currents attract one 
3071: another, spend more time close together, and consequently enhance the 
3072: amplitude for the scattering strength $W$.  Conversely, the interaction 
3073: strength is suppressed when the pair of scattered fermions have 
3074: anti-parallel charge currents.  
3075: The difference in the absolute values of the contributions from the 
3076: diagrams a) and b) is similar to the distinction between the 
3077: particle-hole (CDW) and particle-particle (BCS) processes in a Fermi 
3078: liquid.  Thus, in the situation when $c$ and $c^\prime$ have the same 
3079: sign, the particle-hole process is not ``nested'' because of the 
3080: curvature of the Fermi surface.  On the other hand, the 
3081: particle-particle process is somewhat better nested and becomes 
3082: perfectly nested when $c=c^\prime$, which is the familiar phase space 
3083: reasoning in the theory of the BCS instability.
3084: This distinction between the processes can be clearly seen from the 
3085: arguments $\zeta_{ph}$ and $\zeta_{pp}$, 
3086: Eqs.~(\ref{zeta_ph}) and (\ref{zeta_pp}),
3087: of the function ${\cal E}[\zeta]$ in the two cases.
3088: 
3089: 
3090: With the above general expressions in hand, we now consider specific 
3091: four-fermion interactions.  Of interest is the one-loop contribution
3092: $\Upsilon \propto g^2$ to the eigenvalue of such perturbations at the 
3093: fixed point,
3094: \begin{equation}
3095: \frac{dW}{dl} = (-1 + \Upsilon) W + O(W^2) ~.
3096: \end{equation}
3097: The $O(W^2)$ terms, which in general mix different four-fermion terms,
3098: can be deduced similarly to the RG treatment of quartic interactions 
3099: around the Fermi liquid fixed point\cite{Shankar, Polchinski_FL}
3100: and in fact have very similar structure.  In the Fermi liquid case, 
3101: the interactions are marginal and there is no term linear in $W$, 
3102: so the $O(W^2)$ terms determine the physics.
3103: In the fermion-gauge case, on the other hand, the interactions can
3104: be either relevant when $\Upsilon > 1$ or irrelevant when $\Upsilon < 1$.
3105: For $\Upsilon < 1$, the DBL phase is stable to weak perturbations,
3106: but sufficiently strong bare interactions might still drive the 
3107: DBL through a quantum phase transition into another phase.  
3108: In this case the $O(W^2)$ terms could be helpful in discerning the 
3109: nature of the new phase.
3110: 
3111: As in Fermi liquid theory, the strongest constraint on the allowed
3112: quartic interactions comes from momentum conservation, since all four 
3113: fermions have to live near the Fermi surfaces.
3114: In Fermi liquid theory there are only two allowed types of vertices,
3115: the forward scattering interactions which contribute to the Landau 
3116: parameters and the Cooper vertices.  
3117: We first consider interactions in which all four fermions are of the 
3118: same species.  A general forward scattering interaction then takes the 
3119: form,
3120: \begin{equation}
3121: W^{f}_{{\bf k}, {\bf k}^\prime} 
3122: d^\dagger_{{\bf k}\alpha} d_{{\bf k}\alpha} 
3123: d^\dagger_{{\bf k}^\prime \alpha} d_{{\bf k}^\prime \alpha} ~,
3124: \end{equation}
3125: where ${\bf k}, {\bf k}^\prime$ are two wave vectors on the species 
3126: $\alpha$ Fermi surface.
3127: Consider first the generic case with ${\bf k} \ne \pm {\bf k}^\prime$.
3128: Since this interaction does not involve opposite patches on the 
3129: Fermi surface, the one-loop contribution will vanish, $\Upsilon^f = 0$.
3130: This is also the case when ${\bf k} = {\bf k}^\prime$.
3131: The forward scattering vertex with ${\bf k} = -{\bf k}^\prime$
3132: is part of the Cooper channel to which we now turn.
3133: 
3134: For a single species of fermion, there exists only the odd angular 
3135: momentum Cooper pairing channel (``triplet pairing''),
3136: \begin{equation}
3137: W^t_{{\bf k}, {\bf k}^\prime} 
3138: d^\dagger_{{\bf k} \alpha} d^\dagger_{-{\bf k} \alpha}
3139: d_{-{\bf k}^\prime \alpha} d_{{\bf k}^\prime \alpha} ,
3140: \end{equation}
3141: with 
3142: $W^{t}_{{\bf k}, {\bf k}^\prime} = -W^{t}_{-{\bf k}, {\bf k}^\prime} 
3143: = -W^{t}_{{\bf k}, -{\bf k}^\prime}$.  
3144: The one loop contribution is negative,
3145: $\Upsilon^t_{{\bf k}, {\bf k}^\prime} = 
3146: - (\lambda_{pp}^{\bf k} + \lambda_{pp}^{{\bf k}^\prime})$
3147: for ${\bf k} \ne \pm {\bf k}^\prime$, due to the Amperean repulsion
3148: between the two particles (and between the two holes) with antiparallel 
3149: group velocities.  
3150: For the special case ${\bf k} = \pm {\bf k}^\prime$ one has 
3151: $\Upsilon^t_{{\bf k}, \pm {\bf k}} = 
3152: - 2 (\lambda_{pp}^{\bf k} - \lambda_{ph}^{\bf k})$ 
3153: since there is now also an Amperean attraction between particles 
3154: and holes on opposite patches of the Fermi surface.  
3155: But since we expect $\lambda_{pp}^{\bf k} \ge \lambda_{ph}^{\bf k}$ 
3156: due to the nesting in the particle-particle channel, one has 
3157: $\Upsilon^t \le 0$ in this case as well.
3158: Thus, the one-loop contribution makes this triplet pairing vertex 
3159: more irrelevant.  Physically the repulsive gauge interaction between 
3160: two $d_\alpha$ fermions is unfavorable for their pairing.  
3161: 
3162: 
3163: We now turn to vertices which involve two $d_1$ fermions and two $d_2$ 
3164: fermions.  For simplicity, we first focus on the situation with 
3165: vanishing D-eccentricity, so that the two Fermi surfaces coincide.
3166: Consider first the forward scattering interactions coupling the 
3167: densities of the two fermion species,
3168: \begin{equation}
3169: W^{f;1,2}_{{\bf k}, {\bf k}^\prime} 
3170: d^\dagger_{{\bf k} 1} d_{{\bf k} 1} 
3171: d^\dagger_{{\bf k}^\prime 2} d_{{\bf k}^\prime 2} ~.
3172: \label{forward_scattering}
3173: \end{equation}
3174: The only non-vanishing one loop correction is when
3175: ${\bf k} = -{\bf k}^\prime$, and gives 
3176: $\Upsilon^{f;1,2}_{{\bf k}, -{\bf k}} = 
3177: 2 (\lambda^{\bf k}_{pp} - \lambda^{\bf k}_{ph}) \geq 0$.
3178: If this is greater than one, it would signal a possible instability, 
3179: but in any event we will see that this is smaller than the eigenvalue 
3180: in the conventional BCS ``singlet" pairing channel that we consider next.
3181: 
3182: The singlet BCS interaction is of the form,
3183: \begin{equation}
3184: W^{s}_{{\bf k}, {\bf k}^\prime} 
3185: d^\dagger_{{\bf k} 1} d^\dagger_{-{\bf k} 2}
3186: d_{-{\bf k}^\prime 2} d_{{\bf k}^\prime 1} ~.
3187: \end{equation}
3188: With zero D-eccentricity as assumed here, the BCS interaction is nested.
3189: Moreover, since the two particles (or the two holes) here carry opposite 
3190: gauge charge and therefore have parallel gauge currents, there will be 
3191: an Amperean attraction giving a positive contribution to the Cooper 
3192: vertex: 
3193: $\Upsilon^{s}_{{\bf k}, {\bf k}^\prime} = 
3194: \lambda_{pp}^{\bf k} + \lambda_{pp}^{{\bf k}^\prime}$,
3195: for ${\bf k} \ne \pm {\bf k}^\prime$.
3196: For a circular Fermi surface, symmetry dictates that 
3197: $\lambda_{pp}^{\bf k}$ will be ${\bf k}$-independent,
3198: but this is actually the case generally for such one-loop contribution
3199: in the case with matched $d_1$ and $d_2$ Fermi surfaces since the 
3200: curvatures on the opposing patches are equal.  
3201: Thus, as claimed earlier, the eigenvalue of the singlet BCS pairing 
3202: interaction, $\Upsilon^s = 2\lambda_{pp}$, is larger than the 
3203: corresponding eigenvalue of the forward scattering interaction coupling 
3204: the densities of the two species on opposite sides of the Fermi surface, 
3205: $\Upsilon^s > \Upsilon^{f;1,2}_{{\bf k}, -{\bf k}}$.
3206: For the special case ${\bf k} = - {\bf k}^\prime$, there will be an 
3207: additional positive contribution to the eigenvalue of the singlet BCS 
3208: pairing interaction due to the Amperean attraction between particles and 
3209: holes,
3210: $\Upsilon^s_{{\bf k}, -{\bf k}} =  
3211: 2 (\lambda_{pp}^{\bf k} + \lambda_{ph}^{\bf k})$.  
3212: But being a set of measure zero, this will not contribute to a BCS-type 
3213: pairing instability.
3214: At the present it is not clear what state would be preferred by a 
3215: large interaction of this form and what significance this might have.
3216: 
3217: It is interesting to use the $N=\infty$ parameters to deduce an 
3218: approximate value for the one-loop contribution to $\Upsilon^s$.  
3219: With zero D-eccentricity, the two Fermi surfaces coincide, 
3220: so the curvatures are equal, $c_1=c_2$, and 
3221: $\lambda_{pp}^{\bf k} = {\cal E}[0] = 1$ for all ${\bf k}$.
3222: This gives  $\Upsilon^s = 2$ and implies that the BCS interaction is 
3223: strongly relevant,
3224: \begin{equation}
3225: \frac{d W^s} {d l} = (-1 + \Upsilon^s) W^s = W^s ~,
3226: \end{equation}
3227: within this approximation.  Since we expect some bare short-range 
3228: attraction between the $d_1$ and $d_2$ particles, the runaway flows will 
3229: lead to BCS pairing.  The composite boson 
3230: $b^\dagger = d_1^\dagger d_2^\dagger$ will condense, resulting in a 
3231: superfluid phase.  Thus, based on this analysis, we suspect that the 
3232: putative S-type Bose liquid phase accessed from the gauge theory with 
3233: identical Fermi surfaces for $d_1$ and the $d_2$ will be generically 
3234: unstable towards superfluidity.
3235: 
3236: 
3237: We now consider the situation with non-zero D-eccentricity.
3238: In this case, $d_1$ and $d_2$ have different Fermi surfaces and the 
3239: Cooper vertex will no longer be nested.  This precludes a conventional 
3240: weak coupling BCS instability for the D-wave Bose liquid phase.
3241: The remaining channels that could potentially drive a weak coupling 
3242: instability are the forward scattering interactions given in 
3243: Eq.~(\ref{forward_scattering}) with the locations ${\bf k}$ 
3244: and ${\bf k}^\prime$ on the two Fermi surfaces tuned to have 
3245: anti-parallel patch normals.
3246: 
3247: For bosons on the square lattice, with increasing D-eccentricity the two 
3248: closed Fermi surfaces will eventually open up, and above some critical 
3249: eccentricity the requirement of anti-parallel $d_1$ and $d_2$ Fermi 
3250: surface normals will be no longer possible.
3251: In this large D-eccentricity regime, when the two Fermi surfaces are so
3252: dissimilar, there are no such forward scattering interactions coupling
3253: the two species which can be enhanced by the gauge fluctuations.
3254: Sometimes one can find interactions that can be enhanced via the 
3255: particle-hole attraction within the same species, e.g., 
3256: $d_{{\bf k} 1}^\dagger d_{{\bf -k} 1} 
3257:  d_{{\bf k}^\prime 2}^\dagger d_{{\bf k}^{\prime\prime} 2}$,
3258: which requires tuning both ${\bf k}^\prime$ and ${\bf k}^{\prime\prime}$
3259: to satisfy momentum conservation but this is not always possible.
3260: In any event, the corresponding one-loop $\Upsilon = \lambda_{ph}^{\bf k}$ 
3261: is not likely to change the irrelevance of this term from the tree level;
3262: e.g. the $N=\infty$ estimate of $\lambda_{ph}^{\bf k}$, 
3263: Eq.~(\ref{lambda_Ninfty}), is always smaller than 1.
3264: Regarding other interactions, none of the intra-species four fermion 
3265: terms have positive eigenvalues.  
3266: We conclude that in this large D-eccentricity regime the D-wave Local 
3267: Bose liquid (as specified in Sec.~\ref{subsec:openFS}) will exist 
3268: as a stable phase.
3269: 
3270: Returning to the case with smaller D-eccentricity, the stability of the 
3271: DBL phase will depend on the eigenvalues $\Upsilon^f(\hat{\bf n})$ of 
3272: the forward scattering interaction that couples a patch on one Fermi 
3273: surface with normal $\hat{\bf n}$ to a patch on the other Fermi surface 
3274: with anti-parallel normal.  The particles on opposing patches 
3275: experience a strong Amperean attraction, whereas a particle and hole 
3276: experience a weaker repulsion, so that the eigenvalues 
3277: $\Upsilon^f(\hat{\bf n}) = 2(\lambda_{pp} -\lambda_{ph})$ 
3278: will be positive for all orientations of $\hat{\bf n}$.  
3279: However, the magnitude of $\Upsilon^f(\hat{\bf n})$ depends on the 
3280: Fermi surface curvatures and will vary with orientation.
3281: This can be made explicit by evaluating the one-loop contribution using 
3282: the $N=\infty$ values of the various parameters, thereby we obtain an 
3283: approximate expression for the eigenvalue,
3284: \begin{eqnarray*}
3285: \Upsilon^f(\hat{\bf n}) \approx 
3286: 2{\cal E} \left[ \frac{\sqrt{3}\pi}{2 g^2} \Gamma |c_1 - c_2| \right]
3287: - 2{\cal E} \left[ \frac{\sqrt{3}\pi}{2 g^2} \Gamma (c_1 + c_2) \right]
3288: ~.
3289: \end{eqnarray*}
3290: The first contribution is maximal and equal to $2$ when the curvatures 
3291: on the two Fermi surfaces coincide, which they will when $\hat{\bf n}$ 
3292: is along a diagonal of the square lattice.
3293: We have no such estimate of the second contribution since 
3294: $g, \Gamma, c_\alpha$ are independent in our $N=\infty$ theory; 
3295: if we use in addition a crude RPA approximation Eq.~(\ref{Gamma_RPA}), 
3296: we get the total
3297: $\Upsilon^f = 2 - 2{\cal E}[\sqrt{3}] = 1.53 > 1$.
3298: In this case, the D-wave Bose liquid in this regime of D-eccentricity 
3299: will presumably be unstable, driven by the forward scattering 
3300: interactions with $\hat{\bf n} = \hat{\bf x} \pm \hat{\bf y}$ 
3301: along the diagonals of the square lattice.
3302: 
3303: 
3304: The above numerical estimate is very uncontrolled and is given only 
3305: to see possible trends.  Still, let us speculate what the resulting
3306: phase might if there is indeed such instability.
3307: Most naively, there are two guesses, obtained by a mean field 
3308: decoupling of the four fermion interaction
3309: $d^\dagger_{{\bf k} 1} d_{{\bf k} 1} 
3310:  d^\dagger_{{\bf k}^\prime 2} d_{{\bf k}^\prime 2}$,
3311: where ${\bf k}$ and ${\bf k}^\prime$ are the locations of the two
3312: patches with normals $\hat{\bf n}$ and $-\hat{\bf n}$ that
3313: give the largest $\Upsilon^f$.
3314: We can assume a non-vanishing particle-particle or particle-hole 
3315: condensate, 
3316: $\la d^\dagger_{{\bf k} 1} d^\dagger_{{\bf k}^\prime 2} \ra $ or
3317: $\la d^\dagger_{{\bf k} 1} d_{{\bf k}^\prime 2} \ra$, respectively.
3318: Both condensates carry non-zero momentum,
3319: ${\bf k} + {\bf k}^\prime$ for the particle-particle condensate, and 
3320: ${\bf k} - {\bf k}^\prime$ for the particle-hole.  
3321: But since the interaction is repulsive in the particle-hole channel,
3322: such a condensate seems rather unlikely.
3323: Also, the combination $d_1^\dagger d_2$ carries a non-zero gauge charge, 
3324: so is unphysical and cannot serve as a legitimate order parameter
3325: [but the product 
3326: $\la d^\dagger_{{\bf k} 1} d_{{\bf k}^\prime 2} 
3327:      d^\dagger_{-{\bf k}^\prime 2} d_{-{\bf k} 1} \ra$ 
3328: is gauge invariant, and if condensed would correspond to
3329: an energy density wave at momentum $2 ({\bf k} - {\bf k}^\prime)$].
3330: 
3331: If there is an instability, we think it is more likely to occur in the 
3332: particle-particle channel.  The order parameter 
3333: $\la d^\dagger_{{\bf k} 1} d^\dagger_{{\bf k}^\prime 2} \ra $ 
3334: is gauge invariant and correspond physically to a finite momentum 
3335: Bose condensate.  
3336: The situation is analogous to the FFLO problem, and as there, provided 
3337: the condensate is not too strong, it will only gap out parts of the 
3338: Fermi surfaces.  As such, this would still be a very unusual type of 
3339: superfluid, which in a mean field description would have residual 
3340: gapless fermionic excitations.  Beyond the mean field, the gauge 
3341: fluctuations would still scatter the gapless fermions, rendering them 
3342: incoherent.  This state would thus correspond to finite momentum 
3343: Bose condensation co-existing with the DBL.  
3344: 
3345: 
3346: In sum, whether or not the DBL for small D-eccentricity is unstable 
3347: to such finite momentum pairing is a quantitative issue which will 
3348: depend on the values of the anomalous dimensions.  
3349: An analysis of the actual wavefunctions in Sec.~\ref{sec:Wavefnc_props}
3350: reveals no such tendencies, so might be taken as an indication against 
3351: the instability.  
3352: Indeed, as we saw in Sec.~\ref{sec:Wavefnc_props}, the S-wave Bose 
3353: liquid wavefunction does appear generically unstable towards a 
3354: conventional superfluid, consistent with the expectation of a 
3355: zero momentum BCS instability in the gauge theory.   
3356: In the same spirit, we will see in the next section that for hard-core 
3357: bosons at half-filling the DBL wavefunction in the extreme 
3358: D-eccentricity limit (corresponding to fermions which can only hop in 
3359: one of the two directions on the square lattice) 
3360: reveals a co-existence of a commensurate $(\pi, \pi)$ CDW with a gapless 
3361: DBL Fermi surface.
3362: This indicates that the Gutzwiller wavefunctions for the DBL, 
3363: at least in some special instances, can also reveal translational 
3364: symmetry breaking instabilities,
3365: while we repeat that no such instabilities are observed for generic DBL
3366: wavefunctions.
3367: 
3368: This concludes our discussion of the effective field theory description
3369: and possible instabilities of the DBL phase.
3370: As we have seen, the main potential instability involves pairing 
3371: $d_1$ and $d_2$ fermions moving with opposite group velocities,
3372: but such BCS channel is suppressed for mismatched Fermi surfaces
3373: and is completely eliminated in the DLBL regime, which appears to
3374: be particularly stable.
3375: 
3376: 
3377: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3378: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3379: \section{Ring Hamiltonian Energetics for the DBL phase}
3380: \label{sec:Energetics}
3381: 
3382: In this section we ask what energetics may stabilize the DBL or DLBL
3383: phases.  Specifically, we motivate and study the $J - K_4$ Hamiltonian 
3384: Eq.~(\ref{Hring}) with competing boson hopping and ring exchange terms.
3385: When $J>0$ and $K_4>0$, this Hamiltonian does not satisfy the Marshall 
3386: sign conditions, so one expects the ground state wavefunction to take 
3387: both positive and negative values.
3388: 
3389: The Hamiltonian~(\ref{Hring}) is motivated by considering the 
3390: gauge theory description Eq.~(\ref{HU1}) of the DBL phase in the 
3391: strong coupling limit of the gauge theory, 
3392: $h \gg K, t_\parallel, t_\perp$.  
3393: In this limit, one can perturbatively eliminate the gauge field;
3394: the resulting boson Hamiltonian contains, among other terms, both $H_J$ 
3395: and $H_4$ with 
3396: $J = A\, t_\parallel t_\perp / h$ and
3397: $K_4 = A'\, K t_\parallel^4 / h^4 
3398: - A''\, t_\parallel^2 t_\perp^2 / h^3$ 
3399: with positive numerical coefficients $A, A', A''$.
3400: Only the signs of the contributions to $J$ and $K_4$ are of 
3401: interest here, since in what follows both $J$ and $K_4$ are taken as 
3402: free parameters of the Hamiltonian Eq.~(\ref{Hring}).
3403: Coming from the gauge theory, we have $J>0$ and $K_4>0$ in the regime of 
3404: primary interest when $K \sim h$ and $t_\parallel \gg t_\perp$.
3405: To define the system, we also need to specify the boson density per site 
3406: $\rho$.  The hard-core boson model has particle-hole symmetry,
3407: so it is enough to consider $\rho \leq 1/2$.
3408: 
3409: Before proceeding, we note that the above Hamiltonian with $J>0$ and 
3410: $K_4<0$ was introduced and analyzed in Ref.~\onlinecite{Paramekanti}.
3411: Since there is no sign problem in this case, extensive quantum Monte 
3412: Carlo studies of Refs.~\onlinecite{Sandvik, Melko, Rousseau} were
3413: able to map out the phase diagram of the model.
3414: In addition to a bosonic superfluid phase, the numerical studies 
3415: found two ordered phases at half-filling, a bond-ordered stripe phase 
3416: and a staggered charge density wave (CDW).  When the Monte Carlo was 
3417: performed at fixed chemical potential,\cite{Melko} varying the chemical 
3418: potential drove first-order transitions out of the commensurate solid 
3419: phases into the superfluid.
3420: On the other hand, when the Monte Carlo was performed at fixed 
3421: density,\cite{Rousseau} the superfluid phase was observed as soon as the 
3422: density dropped below half-filling, while a phase separation occurred
3423: when the density was decreased further.
3424: 
3425: The case $J>0$ and $K_4>0$ was not studied in Monte Carlo since there is 
3426: a sign problem.  Below, we present a rudimentary variational energetics 
3427: study of this regime.  
3428: The results are summarized in Fig.~\ref{fig:tKphased} and lend some 
3429: support that the DLBL phase of Sec.~\ref{subsec:openFS} may be 
3430: stabilized by such competing hopping and ring exchange terms.
3431: 
3432: 
3433: \subsection{Model with ring exchanges only}
3434: Consider first the Hamiltonian with the ring term only.
3435: In this case, we can change the sign of $K_4$ by dividing the 
3436: square lattice into four sublattices and performing a transformation 
3437: $b \to -b$ on one sublattice.  
3438: The cited quantum Monte Carlo studies did not consider the pure ring 
3439: model, but extrapolating these to $J=0$, it seems likely that at 
3440: half-filling the ground state is the $(\pi, \pi)$~CDW.
3441: However, we also expect that there is a regime away from half-filling
3442: that realizes a novel ``Excitonic Bose Liquid'' (EBL) phase
3443: discovered and studied by Paramekanti\etal,\cite{Paramekanti}
3444: which we now discuss.
3445: 
3446: The EBL was accessed by
3447: considering a rotor version of the ring Hamiltonian,
3448: \begin{eqnarray*}
3449: H_{\rm rotor} &=& \frac{\cal U}{2} \sum_{\bf r} (n_{\bf r} - \bar{n})^2
3450: \\
3451: &-& |{\cal K}|
3452: \sum_{\bf r} \cos(\phi_{\bf r}
3453:                   - \phi_{{\bf r} + \hat{\bf x}}
3454:                   + \phi_{{\bf r} + \hat{\bf x} + \hat{\bf y}} 
3455:                   - \phi_{{\bf r} + \hat{\bf y}}) ~,
3456: \end{eqnarray*}
3457: where the phase $\phi_{\bf r}$ and the boson density $n_{\bf r}$ are 
3458: canonically conjugate.
3459: When the cosine in the ring term is expanded, i.e., in the phase with 
3460: no topological defects, one finds a quasi--one-dimensional dispersion 
3461: relation for the excitons (density waves),
3462: $\omega_{\bf k}^2 \sim \sin^2(k_x/2) \sin^2(k_y/2)$.
3463: This gapless normal Bose state is expected to be stable 
3464: against becoming an insulator when $|{\cal K}|$ dominates over 
3465: ${\cal U}$.
3466: 
3467: The ring only model on the square lattice has special conservation laws: 
3468: particle numbers on each column and each row are individually conserved.
3469: It is this property that is responsible for the vanishing of 
3470: $\omega_{\bf k}$ along the lines $k_x=0$ and $k_y=0$ in the EBL phase.  
3471: Paramekanti\etal\cite{Paramekanti} showed that due to the gapless lines,
3472: EBL is a critical (power-law) 2d quantum phase with continuously
3473: varying exponents.
3474: 
3475: For our energetics study, it is useful to have a good trial wavefunction 
3476: for the EBL phase of hard-core bosons.  Motivated by the described 
3477: ``spin wave'' theory, we write the following wavefunction
3478: \begin{equation}
3479: \Psi_{\rm EBL}^{(K_4 < 0)}({\bf r}_1, {\bf r}_2, \dots)
3480: \;\;\propto\;\; \exp[-\sum_{i<j} u_{\rm EBL}({\bf r}_i - {\bf r}_j)] ~,
3481: \end{equation}
3482: with
3483: \begin{equation}
3484: u_{\rm EBL}({\bf r}-{\bf r}') = 
3485: \frac{1}{V} \sum_{\bf k} \frac{\cal W}{4 |\sin(k_x/2)| |\sin(k_y/2)|} 
3486: e^{i {\bf k} \cdot ({\bf r} - {\bf r}')} ~. 
3487: \end{equation}
3488: In the spin wave theory, we have ${\cal W} = \sqrt{\cal U/|K|}$, 
3489: but more generally ${\cal W}$ is treated as a variational parameter.
3490: The summation over ${\bf k}$ excludes the lines where either
3491: $k_x=0$ or $k_y = 0$.
3492: Such wavefunction can be defined in any sector with fixed number
3493: of particles on each row and column as is appropriate in the study
3494: of the ring-only model.
3495: We find that for moderate ${\cal W}$ and away from half-filling,
3496: this wavefunction has a density structure factor of the form,
3497: $\la \hat\rho(-{\bf k}) \hat\rho({\bf k}) \ra \sim |k_x| |k_y|$
3498: at long wavelengths as expected in the EBL.
3499: Furthermore, the EBL theory predicts that the box correlator defined in 
3500: Eq.~(\ref{Bb_def}) decays as a power law with a continuously varying
3501: exponent, and this is what we observe for the wavefunction.
3502: In the energetics study, the parameter ${\cal W}$ is varied to minimize
3503: the ring energy $H_4$.
3504: 
3505: Returning to the model with $K_4 > 0$, the trial wavefunction is
3506: \begin{equation}
3507: \Psi_{\rm EBL}^{(K_4 > 0)}({\bf r}_1, {\bf r}_2, \dots)
3508: \propto (-1)^{N_{\rm I}} 
3509: \exp[-\sum_{i<j} u_{\rm EBL}({\bf r}_i - {\bf r}_j)] ~,
3510: \label{Psi_EBL}
3511: \end{equation}
3512: where $N_{\rm I}$ is the number of bosons on the sublattice ${\rm I}$ of 
3513: the four sublattices.  One can readily translate the preceding results
3514: to this wavefunction, which has the sign structure appropriate 
3515: for the pure ring exchange model with $K_4 > 0$.
3516: 
3517: 
3518: Interestingly, the DLBL wavefunction in the extreme D-eccentricity
3519: limit, which was introduced towards the end of Sec.~\ref{subsec:openFS}
3520: and discussed further in Sec.~\ref{subsubsec:xtrmDLBL},
3521: also has the correct sign structure for the ring Hamiltonian.
3522: This follows from an observation that this wavefunction takes 
3523: opposite signs for any two boson configurations that are connected by 
3524: $H_4$.  It is important here that $H_4$ contains only unit square 
3525: plackets, and one subtlety is that such unit ring moves are not ergodic 
3526: in the sectors mentioned earlier with fixed boson numbers on each row 
3527: and column.  The signs of the EBL and extremal DLBL wavefunctions 
3528: as written will not agree everywhere.
3529: However, in the study of the ring model, it is natural to consider 
3530: smaller sectors comprising states that are connected by repeated 
3531: applications of $H_4$, and in such sectors the signs of the two 
3532: wavefunctions agree.  
3533: Thus, one needs to carefully specify what sector restriction is being 
3534: made when working with these wavefunctions.  As far as the $K_4$ 
3535: energetics is concerned, we find essentially no difference between the 
3536: wavefunctions obtained by restricting to the smaller (ergodic) sectors 
3537: or the larger sectors which fix only the boson numbers on each row and 
3538: column.
3539: The presented results are all for the latter choice of sectors;
3540: the Monte Carlo sampling is performed using arbitrary rectangular 
3541: ring moves, which are ergodic in this sector.
3542: 
3543: 
3544: Figure~\ref{fig:etrial_ring} shows the ring exchange energy per site 
3545: for the optimized EBL and extremal DLBL wavefunctions plotted vs boson 
3546: density.  The two trial energies are fairly close; the EBL state has 
3547: somewhat lower energy for $\rho \gtrsim 1/4$, and the difference is 
3548: largest near $\rho = 1/2$.
3549: Note that the extremal DLBL wavefunction has no variational 
3550: parameters, while the EBL state has one parameter.
3551: The trial DLBL energy can be somewhat improved by maintaining the 
3552: same sign structure but taking a variable power of the determinants; 
3553: such trial energies (which we do not show) approach closer to the 
3554: optimal EBL energies.
3555: Figure~\ref{fig:etrial_ring} suggests that the DLBL wavefunction 
3556: not only has the right sign structure, but also has similar short range 
3557: correlations to the EBL state.
3558: Of course, additional Jastrow-type factors can be used to further
3559: improve the DLBL energy, but here and below we emphasize the good sign 
3560: structure and correlations that are already present in the 
3561: determinantal construction itself.
3562: 
3563: 
3564: 
3565: \begin{figure}
3566: \centerline{\includegraphics[width=\columnwidth]{figures/xtrmDBL_EBL.eps}}
3567: \vskip -2mm
3568: \caption{
3569: Trial ring energy in the optimized EBL wavefunction (filled circles)
3570: and extremal DLBL wavefunction (open circles).
3571: Measurements are performed on a $24 \times 24$ system;
3572: the boson number is always a multiple of 24 since the extremal DLBL 
3573: requires that the number of particles is exactly the same in each 
3574: row/column.
3575: Away from half-filling the optimized EBL and DLBL wavefunctions
3576: produce uniform liquids, but exactly at half-filling we find a long-range
3577: $(\pi,\pi)$ CDW order in both states, which is emphasized with 
3578: large box symbols.
3579: The thin lines are guides to eye, while the thick lines are the result 
3580: of the Maxwell construction for the two trial energies that addresses
3581: the possibility of the phase separation at low densities.  
3582: From this construction, the uniform liquids appear to be stable in the 
3583: the density windows  $0.42 < \rho < 0.50$.
3584: }
3585: \label{fig:etrial_ring}
3586: \end{figure}
3587: 
3588: 
3589: The similarity between the EBL and extremal DLBL wavefunctions is also 
3590: revealed in their density structure factors, 
3591: see also Sec.~\ref{subsubsec:xtrmDLBL}.
3592: In both states away from half-filling, for small $|k_x| \ll |k_y|$, 
3593: we can write $D_b(k_x, k_y) = A(k_y) |k_x|$.  
3594: In the DLBL state, $A(k_y)$ appears to be independent of $k_y$.  
3595: On the other hand, in the EBL state, $A(k_y) \sim |k_y|$.  
3596: Thus, the character of the $k_x \to 0$ singularity is the same in both 
3597: states for any non-zero $k_y$, and it is only in the limit 
3598: $k_x, k_y \to 0$ that the two differ.
3599: Interestingly, both wavefunctions also exhibit ``$2 k_F$'' singularities
3600: in the density structure factor.
3601: Such singularities are not apparent in the naive continuum ``spin wave'' 
3602: EBL theory, but manifest themselves in the lattice hard-core boson 
3603: EBL wavefunction.
3604: 
3605: The study of density correlations also reveals that these optimized 
3606: states at half-filling in fact have spontaneous $(\pi, \pi)$ 
3607: charge order, while away from half-filling no order is observed.
3608: Thus, even though in the present crude energetics work we do not consider
3609: directly conventional ordered states of bosons such as charge or bond 
3610: density waves, the EBL and DLBL wavefunctions provide some access to 
3611: the energetics of such ordered states.  
3612: (Note, however, that these trial states likely have some degree of 
3613: EBL-ness, e.g., finite compressibility; this is analogous to the 
3614: well-known observation that when a conventional Jastrow wavefunction 
3615: is driven into a regime of charge order, it in fact represents a 
3616: supersolid that also supports off-diagonal long range order.)
3617: As we have already mentioned, Monte Carlo studies\cite{Melko, Rousseau}
3618: suggest staggered CDW at $\rho = 1/2$, and our finding of
3619: the same tendency in the considered wavefunctions lends some support
3620: to the goodness of the variational approach.
3621: Conventional boson orders are potentially important near other 
3622: commensurate boson densities but are less important at generic 
3623: $\rho$, and we will not consider charge ordered states further.
3624: 
3625: Leaving the charge orders aside, an important feature is prominent 
3626: in the energy plots in Fig.~\ref{fig:etrial_ring} for the considered 
3627: uniform liquids.  
3628: We see regions where $\epsilon(\rho)$ is concave, which signals an 
3629: instability towards phase separation.  Using the Maxwell construction 
3630: for the available data, we conclude that the system phase separates
3631: for boson densities $\rho \lesssim 0.42$.  
3632: At low densities, the system will spontaneously separate into an empty 
3633: phase and the EBL phase at density $\rho = 0.42$.
3634: In the region near $\rho = 1/2$, the considered uniform liquids
3635: appear to be stable, except perhaps very close to half-filling.
3636: 
3637: The tendency in the ring model towards phase separation at low densities
3638: was noted by Rousseau\etal\cite{Rousseau} 
3639: Indeed, if we consider two bosons on an otherwise empty lattice, 
3640: the lowest ring energy is achieved when the two reside on the same 
3641: square placket, since only in this case the ring term is nonzero.
3642: Thus, two bosons can be bound together by the ring energy.
3643: By performing exact diagonalization of small clusters up to $4 \times 4$ 
3644: with open boundary conditions and for different boson numbers, 
3645: we find that the energy per particle is the lowest when the density is 
3646: near half-filling; also, larger open clusters achieve lower energies per 
3647: boson.  
3648: One can construct eigenstates of the ring Hamiltonian at low densities 
3649: composed of such isolated clusters, and the discussed tendencies suggest 
3650: that not only two but many bosons tend to clump together in the 
3651: presence of the ring exchanges.
3652: 
3653: One more notable feature in the energy plots occurs near half-filling.  
3654: We show only the $\rho \leq 1/2$ part, while the $\rho \geq 1/2$ part 
3655: would be the mirror image due to particle-hole symmetry.
3656: Therefore, $\epsilon(\rho)$ shows a cusp precisely at half-filling --
3657: at least in the EBL case and for the system sizes studied.
3658: Since $\mu = d\epsilon/d\rho$, the cusp may be interpreted as an 
3659: evidence for a charge gap at this density, which appears to be in line 
3660: with the proposal of the CDW order by the Monte Carlo studies.
3661: 
3662: We do not know whether the above observations persist in the limit
3663: of large systems.  Thus, it may also happen that there will be
3664: no regime of the EBL phase in the specific model if the system chooses to
3665: phase separate into empty space and a half-filled charge insulator.
3666: If there is a stable EBL regime over some density window near 
3667: half-filling, then the phase separation at lower densities is likely to 
3668: be into empty space and an EBL phase of an appropriate density 
3669: determined from the Maxwell construction.
3670: The existence of an EBL phase near half-filling could presumably be 
3671: checked by careful Monte Carlo simulations for the pure ring model.
3672: It is also plausible that the regime of the stable uniform EBL phase 
3673: could be extended by adding interactions that disfavor phase separation, 
3674: but we have not explored this in the present work.
3675: 
3676: 
3677: \subsection{Full $J$-$K_4$ Hamiltonian}
3678: We now turn to the case with non-zero boson hopping $J > 0$.
3679: Consider first small $J$, and suppose we start in the uniform
3680: EBL or DLBL regime assuming such exists.  As we have discussed, 
3681: the hopping and ring terms frustrate one another.  
3682: At present, we do not know how to include the effects of
3683: hopping if we begin with the EBL.
3684: Thus, the EBL wavefunction is defined in sectors with fixed 
3685: boson numbers on each row/column, but what relative signs should 
3686: we take for the different sectors that are now mixed by the boson
3687: hopping?  If, for example, we naively extend Eq.~(\ref{Psi_EBL}) to all 
3688: boson configurations, with perhaps modified Jastrow factors, the
3689: expectation value of $H_J$ remains precisely zero.
3690: It seems that one needs to change the sign structure of the wavefunction
3691: in order to treat the frustration accurately.
3692:  
3693: On the other hand, the DLBL wavefunction provides a natural 
3694: continuation from the extremal case of the ring only model, 
3695: and we find that it is indeed capable of utilizing both the ring energy 
3696: and the hopping energy.  More precisely, for non-zero $J$ the optimized 
3697: D-eccentricity is finite and varies as $J$ is increased,
3698: moving towards zero D-eccentricity at large $J$.
3699: Within our restricted variational energetics study this is the strongest 
3700: evidence we can offer that the ground state of the ring Hamiltonian in 
3701: this regime is in the DLBL phase.
3702: 
3703: 
3704: \begin{figure}
3705: \centerline{\includegraphics[width=\columnwidth]{figures/tKphased.eps}}
3706: \vskip -2mm
3707: \caption{
3708: Schematic ``variational phase diagram'' of the $J - K_4$ model.
3709: Numerics is done for the same sizes as in Fig.~\ref{fig:etrial_ring}.
3710: The phase boundaries are obtained by examining the DBL state, allowing 
3711: for the possibility of phase separation, and the superfluid state
3712: (the latter appears to be always uniform in our study).
3713: The optimal values of $t_\perp / t_\parallel$ throughout the whole 
3714: DBL phase are such that the fermions have open Fermi surfaces,
3715: i.e., this is the DLBL phase of Sec.~\ref{subsec:openFS}.
3716: }
3717: \label{fig:tKphased}
3718: \end{figure}
3719: 
3720: 
3721: Some details of the variational study are summarized in the 
3722: ``phase diagram'' in Fig.~\ref{fig:tKphased}.
3723: At a given density, the trial energy of the DBL state is optimized
3724: using the ratio $t_\perp / t_\parallel$ as a variational parameter.  
3725: Similarly to the pure ring model, we also analyze the possibility of 
3726: phase separation, which is done in the spirit of 
3727: Fig.~\ref{fig:etrial_ring}.
3728: We find a stable uniform DBL phase in the density range roughly 
3729: similar to the ring only model.  
3730: Physically, we expect the phase separation to be suppressed when the
3731: hopping becomes non-zero, but our study is too limited to explore this 
3732: systematically.
3733: 
3734: The phase diagram Fig.~\ref{fig:tKphased} also includes a conventional 
3735: superfluid state of bosons, which is expected in the regime of large $J$.
3736: We take the Jastrow-type form for the trial wavefunction,
3737: Eq.~(\ref{Psi_Jastrow}), and use the pseudopotential 
3738: $u({\bf r} - {\bf r}') = W/|{\bf r} - {\bf r}'|^p$ 
3739: with two variational parameters $W$ and $p$.
3740: Allowing more parameters such as an independent nearest-neighbor 
3741: pseudopotential does not visibly change the optimized energies.
3742: Interestingly, as deduced from the energetics, the superfluid-DBL phase 
3743: boundary roughly coincides with the limit of vanishing D-eccentricity
3744: $t_\perp = t_\parallel$.
3745: The resulting $(\det)^2$ wavefunction is positive, and, as we discussed
3746: at length in Sec.~\ref{sec:Wavefnc_props}, this state has off-diagonal 
3747: long range order.  
3748: Of course, it does not have the same long-wavelength correlations 
3749: as the superfluid (e.g., it has unusual $2k_F$ density correlations),
3750: but it appears to have reasonable short-range correlations.
3751: 
3752: Finally, we also remark that in the present $J - K_4$ study, the 
3753: optimal DBL phase when it wins over the superfluid is in fact the DLBL 
3754: phase with open Fermi surfaces discussed in Sec.~\ref{subsec:openFS} 
3755: and is in the regime where the boson Green's function decays 
3756: exponentially in all directions.
3757: As we discussed in Secs.~\ref{sec:MFT} and \ref{sec:Wavefnc_props},
3758: despite such ``locality'' in the boson correlator, this unusual phase 
3759: is still critical with many low energy excitations.
3760: 
3761: Summarizing, the DBL wavefunction is able to interpolate 
3762: between the EBL and the superfluid regimes in the frustrated 
3763: $J - K_4$ model.  This is the main argument supporting our proposal 
3764: that the ground state of the ring Hamiltonian in the intermediate regime 
3765: is in the DLBL phase.
3766: 
3767: 
3768: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3769: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3770: \section{Summary and Discussion}
3771: \label{sec:concl}
3772: 
3773: Our main results were summarized in the Introduction,
3774: so here we only highlight the most interesting points and discuss
3775: possible future directions.
3776: We are searching for examples of uncondensed quantum boson liquids that respect 
3777: time reversal and occur at generic continuously varying densities.
3778: In this paper, we proposed several wavefunctions that produce such
3779: liquids and studied in detail the specific states dubbed DBL and DLBL.
3780: While the initial wavefunction construction appears somewhat ad-hoc,
3781: we argued that it is no more so than the slave particle construction
3782: of spin liquid states in frustrated antiferromagnets, which has been
3783: developed to maturity over the last two decades.  
3784: We introduced a particular slave fermion treatment of hard core bosons
3785: on a square lattice that corresponds to the proposed DBL wavefunctions.
3786: This approach allows to go beyond the trial ground state wavefunction,
3787: suggesting natural excitations, and eventually leads to a gauge theory 
3788: description of the low-energy properties of the putative boson liquid 
3789: phase.
3790: Using techniques previously developed for the so-called uniform RVB
3791: spin liquid with a Fermi sea of spinons,
3792: we argued for the possible stability of the new DBL phases to gauge 
3793: fluctuations and found the DLBL regime to be particularly stable.
3794: Within the gauge theory formulation, we also analyzed boson and boson
3795: density correlators at long distances, which reveal special singular 
3796: surfaces in the momentum space.
3797: 
3798: The singular surfaces were confirmed by measuring the properties of
3799: the proposed DBL/DLBL wavefunctions directly; in the same numerics,
3800: no potential orders were observed, indicating stability of the states.
3801: We remark here that the Gutzwiller wavefunctions and the gauge theory 
3802: should not be viewed as identical descriptions, and, in particular,
3803: the detailed long-distance properties may be different.
3804: The Gutzwiller wavefunctions can be used for crude energetics
3805: and qualitative characterization, but in our opinion the gauge theory 
3806: is more complete as far as the actual quantum phase is concerned.
3807: Indeed, the wavefunctions as constructed do not incorporate any
3808: dynamics of the gauge fluxes, which are clearly additional low-energy 
3809: ``variational'' degrees of freedoms that the quantum system is going to 
3810: explore and utilize.
3811: 
3812: From the singular momentum surfaces in the boson properties, 
3813: one can in principle recover the underlying Fermi surfaces of the 
3814: slave particle construction; the latter have correct volumes for 
3815: the boson density, offering a tantalizing possibility of some 
3816: ``Luttinger theorem'' for uncondensed liquids of interacting
3817: bosons.
3818: 
3819: Some striking thermodynamic properties of the proposed boson liquids
3820: follow from their fermion - gauge character.\cite{Holstein, Reizer, PALee, IoffeKotliar, LeeNagaosa, Polchinski, Altshuler, Nayak, YBKim, LeeNagaosaWen, Senthil} 
3821: In both the DBL and DLBL phases, because of the gapless Fermi surfaces
3822: and the gauge interactions, the specific heat at low temperatures is 
3823: expected to behave as $C \sim T^{2/3}$, while the resistivity is 
3824: expected to vanish as $R_b \sim T^{4/3}$.
3825: The last result in the DBL phase with closed Fermi surfaces is simply a 
3826: quote from earlier studies of such Fermi sea - gauge 
3827: systems,\cite{IoffeKotliar, LeeNagaosa}
3828: while the case with open Fermi surfaces requires slightly more care.  
3829: Our argument is based on the application of the Ioffe-Larkin 
3830: rule\cite{IoffeLarkin} which gives the boson resistivity as 
3831: $R_b = R_{d_1} + R_{d_2}$.
3832: For the Fermi surfaces as in the right panel of Fig.~\ref{fig:introFS},
3833: and the conductivity measured, say, in the $\hat{\bf x}$ direction,
3834: the small-angle scattering by the gauge field can effectively degrade 
3835: the $d_2$ fermion current, therefore the standard result 
3836: $R_{d_2} \sim T^{4/3}$.
3837: On the other hand, such scattering cannot completely degrade the $d_1$ 
3838: fermion current, so $R_{d_1} \ll R_{d_2}$, and $R_b \sim T^{4/3}$ 
3839: follows. 
3840: (Parenthetically, the $d_1$ current can be degraded by quartic or more 
3841: particle interactions, but such contributions vanish as $T^2$ or faster.)
3842: 
3843: Intriguingly, in the DLBL phase, the boson correlators are expected 
3844: to be short-ranged despite such manifest thermodynamic signatures of 
3845: gaplessness and criticality.
3846: On the other hand, the boson box correlator in the DLBL decays as a 
3847: power law and is negative, $-x^{-8}$, offering perhaps a more physical 
3848: glimpse of the fermionic partons in this phase.
3849: Indeed, consider inserting bosons at $(0,0)$ and $(x,x)$
3850: and removing at $(x,0)$ and $(0,x)$.  The dominant contribution
3851: to the box correlator comes from the $d_1$ fermions propagating
3852: $(0,0) \to (x,0)$ and $(x,x) \to (0,x)$, while the $d_2$ paths
3853: are $(0,0) \to (0,x)$ and $(x,x) \to (x,0)$,
3854: which is illustrated in Fig.~\ref{fig:partons}.
3855: The minus sign of the box correlator is then due to one fermionic
3856: exchange needed so that the two pairs of injected fermions are removed 
3857: with the same pairing.
3858: 
3859: \begin{figure}
3860: \centerline{\includegraphics[width=3in]{figures/partons.eps}}
3861: \vskip -2mm
3862: \caption{Boson box correlator 
3863: $\la b^\dagger(0,0) b^\dagger(x,x) b(x,0) b(0,x) \ra$
3864: in the DLBL phase reveals the fermionic character of the partons,
3865: decaying as power law and having negative sign from the fermion
3866: exchange.
3867: }
3868: \label{fig:partons}
3869: \end{figure}
3870: 
3871: The proposed DBL/DLBL wavefunctions do not satisfy Marshall signs,
3872: and their interesting sign structure is brought out by the nodal 
3873: pictures in the continuum such as Fig.~\ref{fig:nodes},
3874: which give some caricature of the wavefunction signs also on the 
3875: lattice even though the nodes are not sharply defined in this case.
3876: Clearly, these wavefunctions cannot be ground states of boson models 
3877: without frustration.
3878: To address the question of what Hamiltonians may stabilize such phases,
3879: we considered a particular frustrated hard core boson model with
3880: competing hopping and four-site ring exchange terms, 
3881: and our energetics study suggests that the DLBL state is a good 
3882: candidate in this model near half-filling.  
3883: Unfortunately, the frustrated nature of the boson motion
3884: makes this model not suitable for large system quantum Monte Carlo 
3885: studies.  We still suggest it as an interesting model for numerical
3886: studies such as exact diagonalization of small systems or DMRG; 
3887: the gapless nature of the boson liquid may make results hard to 
3888: interpret, but the short-range character of the correlations and 
3889: particularly the quasi-local nature of the DLBL phase can perhaps 
3890: facilitate DMRG to access larger systems.
3891: 
3892: 
3893: Looking ahead, we now describe one of the main drives behind the
3894: study of uncondensed boson liquids, despite its own intrinsic interest.
3895: We are searching for electronic conducting non-Fermi liquids,
3896: and the ideas of the present work suggest some avenues in this
3897: direction.
3898: As a specific example, consider the slave boson approach that is
3899: popular in the context of the $t-J$ model of high-$T_c$ superconductors.
3900: The electron operator is written as 
3901: $c^\dagger_\sigma = b^\dagger f^\dagger_\sigma$,
3902: leading to a theory of spinons and slave bosons strongly coupled via
3903: an emergent fluctuating gauge field.
3904: Phenomenologically, a Fermi sea of spinons is very
3905: appealing in the strange metal, but what about the slave bosons?
3906: If they condense, we would recover the conventional Fermi liquid.
3907: It has been argued that the fluctuating gauge field frustrates the 
3908: motion of the bosons and suppresses the tendency to 
3909: condense.\cite{IoffeLarkin, IoffeKotliar, LeeNagaosa, Feigelman, LeeNagaosaWen}
3910: If this suppression could persist to zero temperature, we would
3911: obtain a conducting non-Fermi liquid state.
3912: Such a possibility is precisely what we are trying to establish in the 
3913: present work, and our approach would be to write the slave boson in 
3914: terms of new (second generation) slave fermions, 
3915: $b^\dagger = d_1^\dagger d_2^\dagger$.
3916: 
3917: On the level of wavefunctions, we would then write an electronic state 
3918: of the form:
3919: \begin{eqnarray*}
3920: \Psi_{\rm electron}(\up, \dn)
3921: = [(\det)_x \times (\det)_y] (\up, \dn) \,\, \det(\up) \, \det(\dn) ~,
3922: \end{eqnarray*}
3923: where schematically $(\up)$ or $(\dn)$ denotes locations
3924: of all electrons with spin-up or spin-down respectively.
3925: Each determinant in the slave boson wavefunction 
3926: $(\det)_x \times (\det)_y$ evaluates appropriate $d$-particle
3927: orbitals at the locations of all electrons irrespective of their
3928: spin, which assures the no-double occupancy constraint.  
3929: There is significant freedom in specifying these orbitals,
3930: and whether such wavefunctions are useful for any electronic system
3931: requires detailed energetics studies of specific Hamiltonians.
3932: One can nevertheless develop a low-energy theory of such a $T=0$ phase 
3933: in the spirit of the present paper, thereby obtaining an itinerant 
3934: non-Fermi liquid conducting phase.
3935: If the DBL or DLBL bosonic states studied in this work appear as
3936: useful caricatures of the behavior of the electronic charge, 
3937: we may even call the resulting phase a ``D-wave metal''!
3938: Such all-fermion description of electrons on the square lattice with 
3939: no double occupancy has some phenomenological appeal, but detailed 
3940: explorations are left for future work.
3941: 
3942: 
3943: 
3944: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3945: \acknowledgments
3946: We would like to acknowledge discussions with T.~Senthil and
3947: A.~Vishwanath, and thank Jason Alicea for help with some calculations.
3948: The work at KITP was supported by the National Science Foundation 
3949: through grants PHY-9907949 and DMR-0529399.
3950: 
3951: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3952: 
3953: 
3954: 
3955: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3956: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3957: \begin{thebibliography}{10}
3958: 
3959: 
3960: \bibitem{Anderson}
3961: P. W. Anderson, Science {\bf 235}, 1196 (1987).
3962: 
3963: \bibitem{Baskaran}
3964: % The resonating valence bond state and high-T_c superconductivity
3965: % - a mean field theory
3966: G. Baskaran, Z. Zou and P. W. Anderson, 
3967: Solid State Commun. {\bf 63}, 973 (1987).
3968: 
3969: \bibitem{KotliarLiu}
3970: % Superexchange mechanism and d-wave superconductivity
3971: G. Kotliar and J. Liu,
3972: Phys. Rev. B {\bf 38}, 5142 (1988). 
3973: 
3974: \bibitem{IoffeLarkin}
3975: % Gapless fermions and gauge fields in dielectrics
3976: L. B. Ioffe and A. I. Larkin,
3977: Phys. Rev. B {\bf 39}, 8988 (1989).
3978: 
3979: \bibitem{LeeNagaosaWen}
3980: % Doping a Mott insulator: Physics of high-temperature superconductivity
3981: P. A. Lee, N. Nagaosa, and X.-G. Wen,
3982: Rev. Mod. Phys. {\bf 78}, 17 (2006).
3983: 
3984: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3985: 
3986: 
3987: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3988: 
3989: \bibitem{Feigelman}
3990: % Two-dimensional Bose liquid with strong gauge-field interaction
3991: M. V. Feigelman, V. B. Geshkenbein, L. B. Ioffe, and A. I. Larkin
3992: Phys. Rev. B {\bf 48}, 16641 (1993).
3993: 
3994: \bibitem{Dalidovich}
3995: % Interaction-induced Bose metal in two dimensions
3996: D. Dalidovic and P. Phillips, Phys. Rev. B 64, 052507 (2001).
3997: 
3998: \bibitem{Galitski}
3999: % Vortices and Quasiparticles near the Superconductor-Insulator 
4000: % Transition in Thin Films
4001: V. M. Galitski, G. Refael, M. P. A. Fisher, and T. Senthil
4002: Phys. Rev. Lett. {\bf 95}, 077002 (2005).
4003: 
4004: \bibitem{Alicea}
4005: % Algebraic Vortex Liquid in Spin-1/2 Triangular Antiferromagnets: 
4006: % Scenario for Cs2CuCl4
4007: J. Alicea, O. I. Motrunich, and M. P. A. Fisher,
4008: Phys. Rev. Lett. {\bf 95}, 247203 (2005);
4009: % Theory of the algebraic vortex liquid in an anisotropic spin-1/2 
4010: % triangular antiferromagnet
4011: J. Alicea, O. I. Motrunich, and M. P. A. Fisher,
4012: Phys. Rev. B {\bf 73}, 174430 (2006).
4013: 
4014: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4015: 
4016: 
4017: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4018: 
4019: \bibitem{Ceperley91}
4020: % Fermion Nodes
4021: D. M. Ceperley,
4022: J. Stat. Phys. {\bf 63}, 1237 (1991).
4023: 
4024: \bibitem{WenPSG}
4025: % Quantum orders and symmetric spin liquids
4026: X.-G. Wen, 
4027: Phys. Rev. B {\bf 65}, 165113 (2002); cond-mat/0107071.
4028: 
4029: \bibitem{GirvinMacDonald}
4030: % Off-diagonal long-range order, oblique confinement, and the fractional 
4031: % quantum Hall effect
4032: S. M. Girvin and A. H. MacDonald,
4033: Phys. Rev. Lett. {\bf 58}, 1252 (1987).
4034: 
4035: \bibitem{Kane}
4036: % General validity of Jastrow-Laughlin wave functions
4037: C. L. Kane, S. Kivelson, D. H. Lee, and S. C. Zhang,
4038: Phys. Rev. B {\bf 43}, 3255 (1991).
4039: 
4040: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4041: 
4042: 
4043: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4044: 
4045: \bibitem{Paramekanti}
4046: % Ring exchange, the exciton Bose liquid, and bosonization in two dimensions
4047: A. Paramekanti, L. Balents, and M. P. A. Fisher,
4048: Phys. Rev. B {\bf 66}, 054526 (2002).
4049: 
4050: \bibitem{Sandvik}
4051: % Striped Phase in a Quantum XY Model with Ring Exchange
4052: A. W. Sandvik, S. Daul, R. R. P. Singh, and D. J. Scalapino, 
4053: Phys. Rev. Lett. {\bf 89}, 247201 (2002).
4054: 
4055: \bibitem{Melko}
4056: % Two-dimensional quantum XY model with ring exchange and external field
4057: R. G. Melko, A. W. Sandvik, and D. J. Scalapino,
4058: Phys. Rev. B {\bf 69}, 100408 (2004).
4059: 
4060: \bibitem{Rousseau}
4061: % Ring exchange and phase separation in the two-dimensional boson 
4062: % Hubbard model
4063: V. G. Rousseau, R. T. Scalettar, and G. G. Batrouni,
4064: Phys. Rev. B {\bf 72}, 054524 (2005);
4065: % Phase Separation in the Two-Dimensional Bosonic Hubbard Model with 
4066: % Ring Exchange
4067: V. Rousseau, G. G. Batrouni, and R. T. Scalettar,
4068: Phys. Rev. Lett. {\bf 93}, 110404 (2004).
4069: 
4070: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4071: 
4072: 
4073: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4074: 
4075: \bibitem{Zhang_CS}
4076: % Effective-Field-Theory Model for the Fractional Quantum Hall Effect 
4077: S.C. Zhang, T.H. Hansson and S. Kivelson, 
4078: Phys. Rev. Lett. {\bf 62}, 82 (1989).
4079: 
4080: \bibitem{MPAF_Lee}
4081: % Anyon superconductivity and the fractional quantum Hall effect 
4082: D.H. Lee and M.P.A. Fisher,
4083: Phys. Rev. Lett. {\bf  63}, 903 (1989).
4084: 
4085: \bibitem{Wen_edge}
4086: % Gapless boundary excitations in the quantum Hall states and in the 
4087: % chiral spin states
4088: X.G. Wen,
4089: Phys. Rev. B {\bf 43}, 11025 (1991); 
4090: % Electrodynamical properties of gapless edge excitations in the 
4091: % fractional quantum Hall states
4092: Phys. Rev. Lett. {\bf 64}, 2206 (1990). 
4093: 
4094: 
4095: \bibitem{Wen_proj4FQHE}
4096: %Projective Construction of Non-Abelian Quantum Hall Liquids
4097: X.-G. Wen, Phys. Rev. B {\bf 60}, 8827 (1999).
4098: 
4099: \bibitem{KL}
4100: % Theory of the spin liquid state of the Heisenberg antiferromagnet
4101: V. Kalmeyer and R. B. Laughlin, Phys. Rev. B {\bf 39}, 11879 (1989);
4102: % Equivalence of the resonating-valence-bond and fractional 
4103: % quantum Hall states
4104: Phys. Rev. Lett. {\bf 59}, 2095 (1987).
4105: 
4106: \bibitem{WenWilczekZee}
4107: % Chiral spin states and superconductivity
4108: X. G. Wen, F. Wilczek, and A. Zee,
4109: Phys. Rev. B {\bf 39}, 11413 (1989).
4110: 
4111: \bibitem{LaughlinZou}
4112: % Properties of the chiral-spin-liquid state
4113: R. B. Laughlin and Z. Zou, Phys. Rev. B {\bf 41}, 664 (1990).
4114: 
4115: \bibitem{ZhouWen}
4116: % Quantum Orders and Spin Liquids in Cs$_2$CuCl$_4$
4117: Y. Zhou and X.-G. Wen, cond-mat/0210662.
4118: 
4119: \bibitem{Yunoki06}
4120: % Two spin liquid phases in the spatially anisotropic triangular 
4121: % Heisenberg model
4122: S. Yunoki and S. Sorella,
4123: Phys. Rev. B {\bf 74}, 014408 (2006).
4124: 
4125: \bibitem{Rantner}
4126: % Spin correlations in the algebraic spin liquid: 
4127: % Implications for high-Tc superconductors
4128: W. Rantner and X.-G. Wen, 
4129: Phys. Rev. B {\bf 66}, 144501 (2002).
4130: 
4131: \bibitem{Hermele}
4132: % Stability of U(1) spin liquids in two dimensions
4133: M. Hermele, T. Senthil, M. P. A. Fisher, P. A. Lee, N. Nagaosa, 
4134: and X.-G. Wen,
4135: Phys. Rev. B {\bf 70}, 214437 (2004);
4136: % Algebraic spin liquid as the mother of many competing orders
4137: M. Hermele, T. Senthil, and M. P. A. Fisher,
4138: Phys. Rev. B {\bf 72}, 104404 (2005).
4139: 
4140: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4141: 
4142: 
4143: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4144: 
4145: \bibitem{vmc}
4146: % Monte Carlo simulation of a many-fermion study
4147: D. M. Ceperley, G. V. Chester and M. H. Kalos,
4148: Phys. Rev. B {\bf 16}, 3081 (1977).
4149: 
4150: \bibitem{Zhang}
4151: % A renormalised hamiltonian approach for a resonant valence bond
4152: % wavefunction
4153: F. C. Zhang, C. Gros, T. M. Rice, and H. Shiba,
4154: Supercond. Sci. Technol. {\bf 1}, 36 (1988) (cond-mat/0311604).
4155: 
4156: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4157: 
4158: 
4159: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4160: 
4161: \bibitem{Holstein}
4162: % de Haas-van Alphen Effect and the Specific Heat of an Electron Gas
4163: T. Holstein, R. E. Norton, and P. Pincus,
4164: Phys. Rev. B {\bf 8}, 2649 (1973).
4165: 
4166: \bibitem{Reizer}
4167: % Relativistic effects in the electron density of states, 
4168: % specific heat, and the electron spectrum of normal metals
4169: M. Reizer, Phys. Rev. B {\bf 40}, 11571 (1989).
4170: 
4171: \bibitem{PALee}
4172: % Gauge field, Aharonov-Bohm flux, and high-Tc superconductivity
4173: P. A. Lee, Phys. Rev. Lett. {\bf 63}, 680 (1989).
4174: 
4175: \bibitem{IoffeKotliar}
4176: % Transport phenomena near the Mott transition
4177: L. B. Ioffe and G. Kotliar,
4178: Phys. Rev. B {\bf 42}, 10348 (1990).
4179: 
4180: \bibitem{LeeNagaosa}
4181: % Gauge theory of the normal state of high-Tc superconductors
4182: P. A. Lee and N. Nagaosa,
4183: Phys. Rev. B {\bf 46}, 5621 (1992).
4184: 
4185: \bibitem{Polchinski}
4186: % Low-energy dynamics of the spinon-gauge system
4187: J. Polchinski, Nucl. Phys. B {\bf 422}, 617 (1994).
4188: 
4189: \bibitem{Altshuler}
4190: % Low-energy properties of fermions with singular interactions
4191: B. L. Altshuler, L. B. Ioffe, and A. J. Millis
4192: Phys. Rev. B {\bf 50}, 14048 (1994) 
4193: 
4194: \bibitem{Nayak}
4195: % Non-Fermi liquid fixed point in 2 + 1 dimensions
4196: C. Nayak and F. Wilczek, Nucl. Phys. B {\bf 417}, 359 (1994);
4197: % Renormalization group approach to low temperature properties of a non-Fermi liquid metal
4198: {\it ibid.} {\bf 430}, 534 (1994).
4199: 
4200: \bibitem{YBKim}
4201: % Gauge-invariant response functions of fermions coupled to a gauge field
4202: Y. B. Kim, A. Furusaki, X. G. Wen, and P. A. Lee
4203: Phys. Rev. B {\bf 50}, 17917 (1994);
4204: % Quantum Boltzmann equation of composite fermions interacting with a 
4205: % gauge field
4206: Y. B. Kim, P. A. Lee, and X. G. Wen,
4207: Phys. Rev. B {\bf 52}, 17275 (1995).
4208: 
4209: \bibitem{Senthil}
4210: % Weak magnetism and non-Fermi liquids near heavy-fermion critical points
4211: T. Senthil, M. Vojta, and S. Sachdev,
4212: Phys. Rev. B {\bf 69}, 035111 (2004).
4213: 
4214: \bibitem{Galitski_gauge}
4215: % Metallic phase in a two-dimensional disordered Fermi system with 
4216: % singular interactions
4217: V. M. Galitski, 
4218: Phys. Rev. B {\bf 72}, 214201 (2005).
4219: 
4220: 
4221: \bibitem{footnote:N_flavor}
4222: The structure of the $1/N$ diagrams for the particle-hole and 
4223: particle-particle vertices, figures \ref{fig:Vph} and \ref{fig:Vpp}, 
4224: is specific to our $N$-flavor extension defined in Sec.~\ref{subsec:N_flavor}, 
4225: while different extensions can lead to different organization of 
4226: diagrams.
4227: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4228: 
4229: 
4230: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4231: 
4232: \bibitem{Shankar}
4233: % Renormalization-group approach to interacting fermions
4234: R. Shankar,
4235: Rev. Mod. Phys. {\bf 66}, 129 (1994).
4236: 
4237: \bibitem{Polchinski_FL}
4238: J. Polchinski, Effective Field Theory and the Fermi Surface,
4239: {\it in} Recent directions in particle theory,
4240: Proc. 1992 TASI, eds. J. Harvey and J. Polchinski
4241: (World Scientific, Singapore, 1993);
4242: hep-th/9210046
4243: 
4244: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4245: 
4246: \end{thebibliography}
4247: 
4248: \end{document}
4249: