1: \tolerance = 10000
2: \documentclass[prb,twocolumn,superscriptaddress]{revtex4}
3: \usepackage[dvips]{graphicx}
4: \usepackage{latexsym}
5: \begin{document}
6:
7: \newcommand{\fig}[2]{\includegraphics[width=#1]{#2}}
8:
9: \title{The Roton Fermi Liquid}
10:
11:
12: \author{Leon Balents}
13: \affiliation{Department of Physics, University of California,
14: Santa Barbara, CA 93106--4030}
15: \author{Matthew P. A. Fisher}
16: \affiliation{Institute for Theoretical Physics,
17: University of California, Santa Barbara, CA 93106--4030}
18:
19: \date{\today}
20:
21: \begin{abstract}
22: We introduce and analyze a novel metallic phase of two-dimensional
23: (2d) electrons, the Roton Fermi Liquid (RFL), which, in contrast to
24: the Landau Fermi liquid, supports both gapless fermionic and bosonic
25: quasiparticle excitations. The RFL is accessed using a re-formulation
26: of 2d electrons consisting of fermionic quasiparticles and $hc/2e$
27: vortices interacting with a mutual long-ranged statistical
28: interaction. In the presence of a strong vortex-antivortex (i.e.
29: roton) hopping term, the RFL phase emerges as an exotic yet eminently
30: tractable new quantum ground state. The RFL phase exhibits a ``Bose
31: surface'' of gapless roton excitations describing transverse current
32: fluctuations, has off-diagonal quasi-long-ranged order (ODQLRO) at
33: zero temperature ($T=0$), but is not superconducting, having zero
34: superfluid density and no Meissner effect. The electrical resistance
35: {\it vanishes} as $T \rightarrow 0$ with a power of temperature (and
36: frequency), $R(T) \sim T^\gamma$ (with $\gamma >1$), independent of
37: the impurity concentration. The RFL phase also has a full Fermi
38: surface of quasiparticle excitations just as in a Landau Fermi liquid.
39: Electrons can, however, scatter anomalously from rotonic ``current
40: fluctuations'' and ``superconducting fluctuations''. Current
41: fluctuations induced by the gapless rotons scatter anomalously only at
42: ``hot spots'' on the Fermi surface (with tangents parallel to the
43: crystalline axes), while superconducting fluctuations give rise to an
44: anomalous lifetime over the entire Fermi surface {\sl except} near the
45: (incipient) nodal points (``cold spots'').
46: Fermionic quasiparticles dominate the Hall electrical transport.
47: We also find three dominant instabilities of the RFL phase: an
48: instability to a conventional Fermi liquid phase driven by vortex
49: condensation, a BCS-type instability towards fermion pairing and a
50: (non-pairing) superconducting instability. Precisely {\it at} the
51: instability into the Fermi liquid state, the exponent $\gamma$
52: saturates the bound, $\gamma =1$, so that $R(T) \sim T$. Upon
53: entering the superconducting state the rotons are gapped out, and the
54: anomalous quasiparticle scattering is strongly suppressed. We discuss
55: how the RFL phase might underlie the strange metallic state of the
56: cuprates near optimal doping, and outline a phenomenological picture
57: to accomodate the underdoped pseudo-gap regime and the overdoped
58: Landau Fermi liquid phase.
59: \end{abstract}
60:
61: \maketitle
62:
63:
64: \vspace{0.15cm}
65:
66:
67:
68:
69: \section{Introduction}
70: \label{sec:introduction}
71:
72: Despite the appeal of spin-charge separation as an underpinning to
73: superconductivity in the cuprates, there are seemingly fatal obstacles
74: with this approach. Ever since Anderson's initial
75: suggestion\cite{PWA1} of a spinon Fermi surface in the normal state at
76: optimal doping, there have been nagging questions about the {\it
77: charge} sector of the theory. The concept of the ``holon'', a
78: charge $e$ spinless boson, was introduced in the context of the doped
79: spin liquid state\cite{KRS}, and was presumed to be responsible for
80: the electrical conduction. In the simplest theoretical scenario,
81: spin-charge separation occurs on electronic energy scales, thereby
82: liberating the electron's charge from its Fermi statistics. This
83: idea has been actively investigated over the past 15 years -- see
84: Ref.~\onlinecite{SubirReview} for a review. A
85: pervasive challenge to this perspective, however, is the difficulty of avoiding
86: holon condensation and superconductivity at {\it inappropriately} high
87: temperatures. In addition, recent theoretical work, which has
88: elucidated the phenomenology of putative spin-charge separated states,
89: has led to further conflicts with observations. One class of theories
90: have shown how spin-charge separation can emerge from a
91: superconducting phase by pairing and condensing vortices.\cite{NL}\
92: Following this work, a $Z_2$ gauge theory formulation greatly
93: clarified the nature of fractionalization of electronic (and other)
94: quantum numbers.\cite{Z2}\ It has become clear that a necessary
95: requirement for true spin charge separation in two dimensions is the
96: existence of a ``vison'' excitation with a
97: gap.\cite{earlyvison1,earlyvison2,vison1,vison2}\ The vison is perhaps
98: most simplest thought of within the vortex pairing picture, as the
99: remnant of an unbound single vortex. If these ideas were to apply to
100: the cuprates, one would expect this gap to be of order the pseudogap
101: scale $k_B T^*$. Unfortunately for spin-charge separation advocates,
102: experiments designed to detect the vison and measure its
103: gap\cite{visexp1,visexp2}\ have determined an unnaturally low upper
104: bound of approximately $150K$ for the vison gap in underdoped YBCO. Is
105: this the death knell for spin-charge separation?
106:
107: In a very recent paper focusing on the effects of ring-exchange in
108: simple models of bosons hopping on a 2d square lattice\cite{EBL}, we
109: have identified a novel zero temperature {\it normal fluid} phase -
110: (re-)named the ``Exciton Bose Liquid'' (EBL). In the EBL phase
111: boson-antiboson pairs (ie an exciton) are mobile, being carried by a
112: set of gapless collective excitations, while single bosons cannot
113: propagate. The resulting quantum state is ``almost an insulator'',
114: with the d.c. conductivity vanishing as a power of temperature,
115: $\sigma(T) \sim T^\alpha$ with $\alpha \ge 1$. This is in contrast to
116: the ``strange metallic phase'' in the optimally doped cuprates, which
117: is ``almost superconducting''\cite{PWAbook} with an extrapolated zero
118: resistance at $T=0$, as if it were a superconductor with $T_c=0$.
119: This phenomenology suggests the need for a non-superconducting quantum
120: phase in which the vortices are strongly immobilized at low
121: temperatures.
122:
123: Motivated by this, we revisit the $Z_2$ vortex-spinon field theory of
124: interacting electrons\cite{NL,Z2}, in which the $hc/2e$ vortices and
125: the spinons have a long-ranged statistical interaction mediated by two
126: $Z_2$ gauge fields. Rather than gapping out single vortices while
127: condensing pairs (which leads to a spin-charge separated
128: insulator)\cite{Z2}, we would like to find a quantum phase in which
129: the vortices are gapless but nevertheless immobile. To this end, we
130: add an additional ``vortex-ring'' term to the earlier vortex field
131: theory. This term is effectively a kinetic energy for
132: vortex-antivortex pairs, that is for rotons. To access the limit of
133: strong roton hopping requires a further reformulation of the $Z_2$
134: vortex-spinon field theory, replacing the $Z_2$ gauge fields by $U(1)$
135: gauge fields. Similar $U(1)$ vortex-fermion field theories have been
136: explored in Refs.~\onlinecite{tesanovic}. The resulting $U(1)$ vortex-spinon
137: formulation is
138: tractable in the limit of a very strong roton hopping, and describes
139: the ``Roton Fermi Liquid'' (RFL) phase, a novel metallic ground
140: state qualitatively different than Landau's Fermi liquid phase.
141:
142:
143: Both the $Z_2$ vortex-spinon field theory developed in
144: Refs.~[\onlinecite{NL,Z2}] and our $U(1)$ vortex-fermion field theory,
145: described in Section II below, are constructed in terms of operators
146: which create the excitations of a conventional BCS superconductor: the
147: Bogoliubov quasiparticles, the $hc/2e$ vortices and the collective
148: plasmon mode. This choice of ``basis'', however, does not presume
149: that the system is necessarily superconducting at low temperatures,
150: and indeed we intend to employ such a formulation to describe
151: non-superconducting states. When superconductivity {\it is} present
152: at low temperatures, the formulation will also be employed to describe
153: the ``normal'' state above $T_c$. In these approaches, the Bogoliubov
154: quasiparticle is electrically neutralized\cite{NL}, and the resulting
155: ``spinon'' excitation transported around an $hc/2e$ vortex within an
156: ordinary 2d BCS superconductor, acquire a Berry's phase of $\pi$.
157: Within the $Z_2$ and $U(1)$ vortex-spinon formulations, one introduces
158: spinon creation operators, $f^\dagger_{{\bf r}\sigma}$, where we let
159: ${\bf r}$ denote the sites of a 2d square lattice and with $\sigma$
160: the spin. Vortex creation operators are also introduced, conveniently
161: represented in a ``rotor'' representation as $e^{i\theta}$, which live
162: on the plaquettes of the 2d lattice. The vortices are minimally
163: coupled to a gauge field, living on the links of the dual lattice.
164: The ``flux'' in the gauge field describes charge density fluctuations
165: on the original lattice sites, and for example encapsulates the
166: plasmon mode inside the superconducting phase\cite{Duality}. Finally,
167: the long-ranged statistical interaction between the spinons and
168: vortices is incorporated by introducing two Chern-Simons $Z_2$ or
169: $U(1)$ gauge fields. The important new element in the present paper
170: is the inclusion of a roton hopping term. As we shall see, the RFL
171: phase is readily accessed within the $U(1)$ formulation when the
172: magnitude of this roton hopping term is taken signifigantly larger
173: than the single vortex hopping strength.
174:
175: Here, we briefly summarize the main results established in the following
176: Sections. In Sec. II the $Z_2$ vortex-spinon field theory formulation
177: of Ref.~[\onlinecite{Z2}] is recast in terms of a lattice Hamiltonian.
178: Via a sequence of exact unitary transformations on the Hamiltonian, we
179: demonstrate that it is possible to exchange the $Z_2$ Chern-Simons gauge
180: fields for their $U(1)$ counterparts. Within a Lagrangian
181: representation of the resulting $U(1)$ Hamiltonian which we employ
182: throughout the paper, it is possible to choose a gauge for one of the
183: $U(1)$ Chern-Simons fields so that the ``spinon'' is recharged, and has
184: finite overlap with a bare electron. (In App.~\ref{ap:elecform} we show
185: that this gauge choice can effectively be made at the Hamiltonian level,
186: and construct a Hamiltonian theory in terms of vortices and fermionic
187: operators which carry the electron charge and have a finite overlap with
188: a bare electron.)
189:
190: Initially, in Sec. III, we ignore the fermions entirely, and focus on
191: the bosonic charge (or vortex) sector of the theory. A ``spin-wave''
192: expansion valid in the presence of a large roton hopping term, leads
193: to a simple theory which is quadratic - {\it except} for a single
194: vortex hopping term. Dropping this vortex hopping term then leads to
195: a soluble harmonic theory of the ``Roton-Liquid'' (RL) phase. In
196: addition to the gapless 2d plasmon, the RL phase is shown to support a
197: ``Bose surface'' of gapless roton excitations. We compute the Cooper
198: pair propagator in the RL phase, and show that it exhibits
199: off-diagonal quasi-long-ranged order (ODQLRO) at zero temperature, but
200: {\it not} a Meissner effect. The RL phase exhibits a high degree of
201: ``emergent'' symmetry - the number of vortices on every row and column
202: of the 2d dual lattice is asymptotically conserved at low energies.
203: This symmetry implies that the harmonic ``fixed point'' theory of the
204: RL phase has an infinite conductivity at any temperature.
205:
206: In Sec. IV we study the legitimacy of the approximations used to
207: arrive at the harmonic RL theory, focussing first on the neglected
208: vortex hopping term. We show that for a range of parameters the
209: vortex hopping term is ``irrelevant'', scaling to zero at low energies
210: whenever its associated scaling dimension satisfies $\Delta_v \ge 2$.
211: Nevertheless, at finite temperatures vortex hopping leads to
212: dissipation, giving a resistance which vanishing as a power law in
213: temperature, $R(T) \sim T^\gamma$ with $\gamma = 2 \Delta_v -3 \ge 1$.
214: A ``plaquette duality'' transformation\cite{EBL} allows us to next
215: address the legitimacy of the initial ``spin-wave'' expansion, used to
216: obtain the harmonic RL theory. Of paramount importance is the presence
217: of a term in the dual theory which hops a ``charged'' quasiparticle
218: excitation, a term {\it not} present in the harmonic fixed-point
219: theory. We find that the ``charge'' hopping process is irrelevant over
220: a range of parameters - approximately the complement of the range
221: where vortex hopping was irrelevant - implying stability of the RL
222: phase. When relevant, on the other hand, the ``charge'' quasiparticle
223: condenses, leading to a superconducting ground state.
224:
225: The fermions are re-introduced back into the theory in Section V, where
226: we argue for the stability of the Bose surface of rotons and the 2d
227: plasmon in the presence of a {\sl gapless} Fermi sea of fermionic
228: quasiparticles. We denote the corresponding phase by the Roton Fermi
229: Liquid (RFL). The gaplessness of the fermions is somewhat surprising,
230: and deserves some comment. Indeed, it is in sharp contrast to the
231: gapped nature of the quasiparticles in both the superconducting phase
232: and $Z_2$ fractionalized insulator, in which the spinons experience a
233: BCS-like ``pair field''. The cause of this difference is the existence
234: of gapless single vortex excitations (and fluctuations) in the RFL,
235: which according to our analysis leads to the ``irrelevance'' of the
236: fermion pairing term. Crudely, because the bosonic pair field exhibits
237: only OD{\sl Quasi-}LRO rather than ODLRO, there is no average pair field
238: felt by the quasiparticles, and hence no gap. In this sense, the RFL is
239: in fact closer to a Fermi-liquid state than it is a superconductor.
240:
241: Section VI is devoted to an analysis of the properties of the RFL phase.
242: We study both the longitudinal and Hall conductivities, and find that
243: the dissipative electrical resistance vanishes with a power of
244: temperature, $R(T) \sim T^\gamma$ with $\gamma \ge1$, similar to the
245: behavior without fermions present. But the fermions are found to
246: dominate the Hall response, leading within a na\"ive Drude treatment to
247: an inverse Hall angle varying as $\Theta_H^{-1} \sim 1/(\tau_f^2
248: T^\gamma)$, with $\tau_f$ the fermionic quasiparticle transport
249: lifetime. Due to the presence of the ``Bose surface'' of gapless
250: rotons, electrons at finite energy $\omega$ experience anomalous
251: scattering, {\it not} present in a Landau Fermi liquid. Specifically,
252: quasiparticles scatter due to rotonic ``current fluctuations'' and
253: ``superconducting fluctuations'', which contribute additively to the
254: electron decay rate. The former gives rise to especially strong
255: electron scattering at ``hot spots'' -- points on the Fermi surface with
256: tangents parallel to the axes of the square lattice. At such hot spots
257: the associated electron decay rate varies with an anomalous power of
258: energy, $\omega^{(\gamma + 2)/2}$, for $1 < \gamma < 2$. The decay rate
259: from superconducting fluctuations is present everywhere along the Fermi
260: surface {\sl except} near ``cold spots'' at the incipient $d$-wave nodal
261: points. This contribution grows strongly with decreasing
262: energy/temperature, although it has a smaller overall amplitude than the
263: current fluctuation contribution. Upon entering a superconducting state, the
264: rotons -- gapless within the RFL phase -- become gapped out, and all
265: anomalous scattering is strongly suppressed.
266:
267: Finally, in Section VII we briefly discuss possible connections of the
268: present work to the cuprates. We suggest that the RFL phase might
269: underlie the unusual behavior observed near optimal doping in the
270: cuprates, in particular the ``strange metal'' normal state above $T_c$.
271: A scenario is outlined which also incorporates the pseudo-gap regime and
272: the conventional Fermi liquid behavior in the strongly overdoped limit.
273:
274:
275:
276: \section{The Model}
277: \label{sec:model}
278:
279: We are interested in electrons hopping on a 2d square lattice, with
280: electron creation operators $c^\dagger_{{\bf r}\sigma}$. Here, the
281: sites are denoted (in bold roman characters) as, ${\bf r}=x_1 {\bf \hat{x}}
282: + x_2 {\bf \hat{y}}$, where $x_1,x_2$ are integers, $\hat{\bf
283: x}_1=\hat{\bf x}$, $\hat{\bf x}_2=\hat{\bf y}$ are unit vectors
284: along the $x$ and $y$ axes, and $\sigma = \uparrow,\downarrow$
285: denotes the two spin polarizations of the electron (such a spin
286: index will be distinguished from Pauli matrices $\sigma^\mu$ by
287: the lack of any superscript).
288:
289:
290: \subsection{$Z_2$ Chargon-spinon formulation}
291: \label{sec:z_2-chargon-spinon}
292:
293: We begin by formulating the electron problem in spin-charge
294: separated variables using the $Z_2$ gauge theory Hamiltonian. We
295: emphasize that this formulation does not imply that spin-charge
296: separated excitations are deconfined, and indeed this formulation
297: correctly describes the low-energy physics of conventional
298: confined phases as well.
299:
300:
301:
302: The $Z_2$ gauge theory is most readily formulated in terms of a
303: charge $e$ singlet bosonic chargon $b_{\bf r},b_{\bf r}^\dagger$,
304: a neutral spin-$1/2$ fermionic spinon $f_{{\bf r}\sigma}, f_{{\bf
305: r}\sigma}^\dagger$, and an Ising gauge field (Pauli matrix)
306: $\sigma_j^\mu({\bf r})$ residing on the link between sites ${\bf
307: r}$ and ${\bf r}+\hat{\bf x}_j$. It is convenient to use a rotor
308: representation for the chargons, $b_{\bf r} = e^{-i\phi_{\bf r}}$,
309: $b^\dagger_{\bf r} b^{\vphantom\dagger}_{\bf r} = n_{\bf r}$, with
310: $[\phi_{\bf r},n_{\bf
311: r'}]=i\delta_{\bf r,r'}$.
312:
313: The $Z_2$ Hamiltonian is conveniently expressed in terms of the
314: Hamiltonian density, $H_{Z_2}=\sum_{\bf r} {\cal H}_{Z_2}$, which
315: in turn is decomposed into a bosonic charge sector, a fermionic
316: sector and a gauge field contribution, ${\cal H}_{Z_2}={\cal H}_c
317: + {\cal H}^{Z_2}_f + {\cal H}_g$:
318: \begin{eqnarray}
319: \label{eq:Z2chargeham}
320: {\cal H}_c & = & - t_c \sum_j \sigma_j^z({\bf r}) \cos(\partial_j\phi_{\bf
321: r}-A_j({\bf r})) + u_c (n_{\bf r}-\rho_0)^2 , \\
322: {\cal H}_g & = & - t_v \sum_j \sigma^x_j({\bf r}) \!-\!
323: K\!\!\!\!\prod_{\Box({\bf r}+{\bf w})} \!\!\!\!\sigma^z \!\!\!
324: - \kappa_r\! \prod_j \sigma_j^x({\bf r}) \sigma_j^x({\bf
325: r}\!+\!\hat{\bf x}_j), \\
326: {\cal H}^{Z_2}_f & = & - \sum_j \sigma_j^z({\bf r})
327: [ t_s f^\dagger_{{\bf r} + {\bf \hat{x}}_j \sigma}
328: f^{\vphantom\dagger}_{{\bf r} \sigma}+\Delta_j f_{{\bf r}+\hat{\bf
329: x}_j \sigma} \epsilon_{\sigma\sigma'}f_{{\bf r}\sigma'}+ {\rm
330: h.c.}] \nonumber \\
331: & & - t_e \sum_j f^\dagger_{{\bf r} + {\bf \hat{x}}_j
332: \sigma} e^{i (\partial_j \phi_{{\bf r}} -A_j({\bf r}))}
333: f^{\vphantom\dagger}_{{\bf r} \sigma}+ h.c. . \label{HfZ2}
334: \end{eqnarray}
335: Here, $\rho_0 \equiv 1 - x$ is the electron (charge) density with $x$
336: measuring deviations from half-filling. Throughout the paper,
337: $\partial_j$ with $j=1,2$ denotes a {\it discrete} (forward) spatial
338: lattice derivatives in the $x_1$ and $x_2$ directions, for example,
339: $\partial_1 \phi_{\bf r} = \partial_x \phi_{\bf r} = \phi_{{\bf r} +
340: {\bf \hat{x}}} - \phi_{\bf r}$. We have included an external
341: (physical) vector potential $A_j({\bf r})$ in order to calculate
342: electromagnetic response and to include applied fields. The
343: Hamiltonian ${\cal H}_c$ describes the dynamics of the chargons
344: hopping with strength, $t_c$, which are minimally coupled to the $Z_2$
345: gauge field. The dynamics of the gauge fields is primarily determined
346: from ${\cal H}_g$, the first two terms of which constitute the
347: standard pure $Z_2$ gauge theory Hamiltonian. The ``magnetic''
348: contribution involves the plaquette product,
349: \begin{equation}
350: \prod_{\Box({\bf r}+{\bf w})} \sigma^z \equiv \sigma^z_1({\bf r})
351: \sigma^z_1({\bf
352: r}+ \hat{\bf y}) \sigma^z_2({\bf r}) \sigma^z_2({\bf r}+\hat{\bf
353: x}),
354: \end{equation}
355: which is the $Z_2$ analog of the lattice curl. Here we have
356: defined, ${\bf w} = (1/2)(\hat{x}+\hat{y})$, and ${\bf r} +{\bf
357: w}$ denotes the center of the plaquette. We have also included an
358: additional contribution bi-linear in $\sigma^x$, which in the dual
359: vortex representation below will become a ``roton'' hopping term.
360: In the fermion Hamiltonian, ${\cal H}^{Z_2}_f$, we have defined
361: the antisymmetric matrix
362: $\epsilon_{\sigma\sigma'}=i\sigma^y_{\sigma\sigma'}$, and take
363: $\Delta_j=(-1)^j\Delta$, which describes a nearest-neighbor pair
364: field with d-wave symmetry. Apart from the first two terms
365: familiar to aficionados of the $Z_2$ gauge theory\cite{Z2}, we
366: have included a less exotic bare electron hopping amplitude $t_e$.
367: We will primarily be interested in the limit that the spinon
368: hopping strength is significantly larger than the electron hopping
369: strength, $t_s \gg t_e$.
370:
371: In most of the analysis of this paper, we will consider the limit of
372: small spinon pairing $\Delta_j \rightarrow 0$. This can be justified
373: either by the assumption $\Delta_j \ll t_s$, or by the irrelevance in
374: the renormalization group (RG) sense, which will occur occur in some
375: regimes. If strictly $\Delta_j=0$, both fermion (spinon) number and
376: boson charge are conserved, and in principle may be separately fixed.
377: However, for $\Delta_j \rightarrow 0$, even infinitesimal, this is not
378: the case. Instead the spinons will {\sl equilibrate} in some time that
379: diverges as $\Delta_j\rightarrow 0$ but is otherwise finite, and the
380: system will choose a unique fermion density to minimize its (free)
381: energy. We will return to this point in the $U(1)$ formulation in
382: Sec.~\ref{sec:u1-vortex-spinon-1}.
383:
384:
385: The full Hamiltonian above has a set of local $Z_2$ gauge
386: symmetries, commuting with each of the local operators,
387: \begin{equation}
388: \label{eq:Z2constraint}
389: {\cal C}_{\bf r}^1 = (-1)^{n_{\bf r}+ n^f_{{\bf r}}} \prod_{+({\bf r})} \sigma^x ,
390: \end{equation}
391: where $n^f_{{\bf r}} = f_{{\bf r}\sigma}^\dagger f_{{\bf
392: r}\sigma}^{\vphantom\dagger}$ is the fermion density and the $Z_2$
393: lattice divergence is defined as:
394: \begin{equation} \prod_{+({\bf r})}
395: \sigma^x \equiv \prod_j \sigma^x_j({\bf
396: r}) \sigma^x_j({\bf r}-\hat{\bf x}_j) .
397: \end{equation}
398: Physical states are required to be gauge invariant, which is
399: specified by the set of local {\it constraints}: ${\cal C}^1_{\bf
400: r} \equiv 1$. This is the $Z_2$ analog of Gauss law for
401: conventional electromagnetism.
402:
403: The connection between the $Z_2$ gauge theory and a theory of
404: interacting electrons, is most apparent in the limit that $t_v$ is
405: taken to be much larger than all other couplings. In this limit
406: the electron creation operator is equivalent to the product of the
407: chargon and spinon creation operators, $c^\dagger_{{\bf r \sigma}}
408: = b^\dagger_{\bf r} f^\dagger_{{\bf r} \sigma}$. Indeed, when
409: $t_v \rightarrow \infty$ the $Z_2$ ``electric'' field becomes
410: frozen, $\sigma^x_j \approx 1$, and the gauge constraints then
411: imply that on each site of the lattice the sum of the chargon and
412: spinon numbers, $n_{\bf r} + n^f_{\bf r}$, is even. Moreover, for
413: large $t_v$, the chargon and spinon hopping terms are strongly
414: suppressed, and can be considered perturbatively. Upon integrating
415: out the gauge fields, one will thereby generate an additional electron
416: kinetic energy term with amplitude of order $t_c t_s/t_v$. A
417: brief discussion is given in Appendix~\ref{sec:fermiliquid}.
418:
419:
420: In what follows, we will study the $Z_2$ gauge theory more
421: generally, away from the large $t_v$ limit. Of interest is the
422: electron Greens function,
423: \begin{equation}
424: {\cal G}_e({\bf r}_1, \tau_1; {\bf r}_2, \tau_2) = - \langle
425: T_\tau \hat{c}_{{\bf r}_1\sigma}(\tau_1) \hat{c}^\dagger_{{\bf
426: r}_2 \sigma}(\tau_2) \rangle .
427: \end{equation}
428: We will express the electron operators as,
429: \begin{equation}
430: c_{{\bf r}\sigma} \equiv b_{\bf r} f_{{\bf r}\sigma} ,
431: \end{equation}
432: which is exact as $t_v \rightarrow \infty$, but more generally
433: should be sufficient to extract the universal low energy and long
434: length scale behavior of the electron Green's function. We will
435: also be interested in correlation functions involving the Cooper
436: pair creation and destruction operators, $B^\dagger_{\bf r},B_{\bf
437: r}$, with $B_{\bf r} = (b_{\bf r})^2 = e^{-2i\phi_{\bf r}}$.
438:
439:
440:
441: \subsection{$Z_2$ Vortex-spinon formulation}
442: \label{sec:z_2-vortex-spinon}
443:
444: In what follows, it will prove particularly convenient to work
445: with vortex degrees of freedom, rather than the chargon fields. To
446: arrive at such a description, we use the U(1) duality
447: transformation\cite{Duality}, in which the dual variables sit
448: naturally on the 2d dual lattice. We denote the sites of the 2d
449: dual lattice by sans serif characters as ${\sf r}=x_1 {\bf
450: \hat{x}} + x_2 {\bf \hat{y}} + {\bf
451: w}$, with ${\bf w} = (1/2)({\bf \hat{x}} + {\bf \hat{y}})$ and
452: integer $x_1,x_2$. The duality transformation itself defines two
453: conjugate gauge fields $[a_i({\sf r}), e_j({\sf r}')] = i\delta_{ij}\delta_{\sf
454: r,r'}$, where
455: \begin{eqnarray}
456: \label{eq:duality}
457: n_{\bf r} & = & \frac{1}{\pi} \epsilon_{ij} \partial_i a_j({\bf
458: r-w}), \\
459: \partial_i \phi_{\bf r} & = & \pi \epsilon_{ij} e_j({\bf
460: r+w}-\hat{\bf x}_j).
461: \end{eqnarray}
462: Here, as defined, due to the discreteness of the $n_{\bf r}$
463: variables, $a_j({\sf r})$ takes on values that are integer multiples
464: of $\pi$, while $e_j({\sf r})$ is a periodic variable with period
465: $2$. This transformation is faithful provided the constraint
466: \begin{equation}
467: \label{eq:U1constraint}
468: (\vec{\nabla}\cdot \vec{e})({\sf r}) \equiv \sum_j \partial_j
469: e_j({\sf r}-{\bf x}_j) = 0 \;\; ({\rm mod}\, 2),
470: \end{equation}
471: or equivalently
472: \begin{equation}
473: \label{eq:U1constrainta}
474: {\cal C}^2_{\sf r} = e^{i\pi (\vec{\nabla}\cdot \vec{e})({\sf r})} = 1
475: \end{equation}
476: is imposed at every site ${\sf r}$ of the dual lattice.
477: Rewriting the charge Hamiltonian, one has
478: \begin{eqnarray}
479: \label{eq:DualChargeHam}
480: {\cal H}_c & = & - t_c\sum_i \sigma_i^z({\bf r}) \cos(\pi
481: \epsilon_{ij} e_j({\bf r+w}-\hat{\bf x}_j)-A_i({\bf r}) )\nonumber \\
482: && + \frac{u_c}{\pi^2}(\epsilon_{ij}\partial_i
483: a_j- \pi\rho_0)^2.
484: \end{eqnarray}
485:
486: Conventional $hc/2e$ superconducting vortices are composites of a
487: ``vison'' (topological excitation in $\sigma_j^z$) and a
488: half-vortex in $\phi$. To describe them, we perform a unitary
489: transformation to a new Hamiltonian $\tilde{H}_{Z_2}$ with new
490: constraints $\tilde{\cal C}^a$ ,
491: \begin{equation}
492: \label{eq:Hdual1}
493: \tilde{H}_{Z_2} = U^\dagger H_{Z_2} U^{\vphantom\dagger} , \qquad \tilde{\cal
494: C}^a = U^\dagger {\cal
495: C}^a U^{\vphantom\dagger},
496: \end{equation}
497: with the unitary operator,
498: \begin{eqnarray}
499: \label{eq:unitary}
500: U & = & \exp[\frac{i}{2} \sum_{{\bf r},i,j} \epsilon_{ij}a_i({\bf r+w}-\hat{\bf
501: x}_i) (\sigma_j^z({\bf r})-1)] \nonumber\\
502: & = & \prod_{\bf r} (\sigma^z_1({\bf r}))^{\frac{a_2({\bf r}+\overline{\bf w})}{\pi}}
503: (\sigma^z_2({\bf r}))^{\frac{a_1({\bf r}-\overline{\bf w})}{\pi}},
504: \end{eqnarray}
505: with $\overline{\bf w} = (1/2)(\hat{\bf x} - \hat{\bf y}) = {\bf
506: w}-\hat{\bf y}$. The transformed constraints are,
507: \begin{eqnarray}
508: \label{eq:Z2constraint2a}
509: \tilde{\cal C}_{\bf r}^1 & = & (-1)^{n^f_{{\bf r}}} \prod_{+({\bf
510: r})} \sigma^x = 1, \\
511: \tilde{\cal C}^2_{\sf r} & = & (-1)^{(\vec{\nabla}\cdot
512: \vec{e})({\sf r})} \prod_{\Box({\sf r})} \sigma^z = 1.
513: \label{eq:Z2constraint2b}
514: \end{eqnarray}
515: Under this unitary transformation,
516: \begin{eqnarray}
517: \label{eq:Z2ham2}
518: \tilde{\cal H}_c & = & - t_c \sum_j \cos( \pi e_j({\sf
519: r})+\epsilon_{jk}A_k({\sf r}+{\bf\overline{w}})) \nonumber \\ && +
520: \frac{u_c}{\pi^2}(\epsilon_{ij} \partial_i a_j -
521: \pi\rho_0)^2 , \\
522: \tilde{\cal H}_g & = & \!\!\!\!- t_v \!\sum_j
523: \!\overline{\sigma}_j^x({\sf r})
524: \cos(a_j({\sf r})) \!-\! K
525: (-1)^{(\vec{\nabla}\cdot \vec{e})({\sf r})} + {\cal H}^{Z_2}_r.
526: \end{eqnarray}
527: Here we have defined, $\overline{\sigma}^x_1({\sf r}) =
528: \sigma^x_2({\sf
529: r}+\overline{\bf w}), \overline{\sigma}^x_2({\sf r}) =
530: \sigma^x_1({\sf r}-\overline{\bf w})$, and have used
531: Eq.~(\ref{eq:Z2constraint2b}). The transformed ``roton'' hopping
532: term becomes,
533: \begin{equation}
534: {\cal H}^{Z_2}_r = -\kappa_r \overline{\sigma}_1^x({\sf
535: r}) \overline{\sigma}_1^x({\sf
536: r}+\hat{\bf x}_2)\cos(\partial_y a_x({\sf r})) + (x\leftrightarrow
537: y) .
538: \end{equation}
539: The fermion Hamiltonian is almost unchanged in the dual
540: vortex-spinon representation,
541: \begin{eqnarray}
542: \label{eq:fdv}
543: \tilde{\cal H}^{Z_2}_f & = & - \sum_j \sigma_j^z({\bf r})
544: [ t_s f^\dagger_{{\bf r} + {\bf \hat{x}}_j \sigma}
545: f^{\vphantom\dagger}_{{\bf r} \sigma}+\Delta_j f_{{\bf r}+\hat{\bf
546: x}_j \sigma} \epsilon_{\sigma\sigma'}f_{{\bf r}\sigma'}+ {\rm
547: h.c.}] \nonumber \\
548: & & - t_e \sum_j \sigma_j^z({\bf r}) e^{i(\pi \overline{e}_j({\bf
549: r})-A_j({\bf r}))}
550: f^\dagger_{{\bf r} + {\bf \hat{x}}_j
551: \sigma} f^{\vphantom\dagger}_{{\bf r} \sigma}+ h.c. ,
552: \label{HtildefermZ2}
553: \end{eqnarray}
554: the changes appearing only in the electron hopping (the last term).
555: Here we have defined,
556: \begin{equation}
557: \overline{e}_i({\bf r}) = \epsilon_{ij} e_j({\bf r} + {\bf w} - \hat{x}_j) .
558: \label{eq:defineebar}
559: \end{equation}
560: Notice that $\pi \overline{e}_j$ couples ``like a gauge field'' to
561: the spinons in the final electron term.
562:
563: To arrive at the final $Z_2$ vortex-spinon theory, we split the
564: electric and magnetic fields into longitudinal and transverse parts,
565: $a_j = a_j^l + a_j^t$, $e_j=e_j^l+ e_j^t$, with $\vec{\nabla}\cdot
566: \vec{a}^t=\vec{\nabla}\cdot\vec{e}^t=0$, $\epsilon_{ij}\partial_i
567: a_j^l=\epsilon_{ij}\partial_i e_j^l= 0$. For future purposes, we note
568: that the longitudinal part of $e_j$ is related to the transverse part
569: of $\overline{e}_j$ and vice versa, i.e. $\overline{e}_i^{l/t}({\bf
570: r}) = \epsilon_{ij} e_j^{t/l}({\bf r} + {\bf w} - \hat{x}_j)$. It
571: is convenient then to solve the constraint for the longitudinal
572: fields,
573: \begin{eqnarray}
574: \label{eq:vortexvarstheta}
575: a_j^l({\sf r}) & = & -\partial_j \theta_{\sf r}, \\
576: (\vec{\nabla}\cdot\vec{e}^l)({\sf r}) & = & N_{\sf r}.\label{eq:vortexvarsN}
577: \end{eqnarray}
578: The fields $e^{i\theta_{\sf r}}$ and $N_{\sf r}$ have the
579: interpretation of vortex creation and number operators, as can be seen
580: by tracking the circulation defined from the original chargon phase
581: variable $\phi$. One finds canonical commutation relations,
582: \begin{equation}
583: [\hat{\theta}_{\sf r}, \hat{N}_{\sf r^\prime}] = i \delta_{{\sf r}{\sf
584: r^\prime}} .\label{eq:Nthetacom}
585: \end{equation}
586:
587: At this stage, $N_{\sf r}$ is a periodic variable with period $2$,
588: and $-\partial_j \theta_{\sf r} + a^t_j({\sf r})$ is constrained
589: to be an integer multiple of $\pi$. It is convenient to soften
590: the latter constraint, and in order to respect the uncertainty
591: relation implied by Eq.~(\ref{eq:Nthetacom}), at the same time
592: relax the periodicity of $N_{\sf r}$. Formally, this is
593: accomplished by replacing,
594: \begin{equation}
595: \label{eq:replacecosine}
596: -t_c \sum_j \cos (\pi e_j+\epsilon_{jk}A_k) \rightarrow {\rm Const.} +
597: \frac{u_v}{2}
598: \sum_j (e_j+\epsilon_{jk} A_k/\pi)^2,
599: \end{equation}
600: with $u_v \approx t_c \pi^2$, and adding a term to the
601: Hamiltonian, $\tilde{H}_{Z_2} \rightarrow \tilde{H}_{Z_2} +
602: \sum_{\sf r} {\cal H}^{Z_2}_{2v}$, with
603: \begin{equation}
604: \label{eq:pairvortex}
605: {\cal H}^{Z_2}_{2v} = - t_{2v} \sum_j \cos (2\partial_j \theta-
606: 2a_j^t).
607: \label{pairvhop}
608: \end{equation}
609: The constraint is recovered for large $t_{2v}$, but we will
610: consider a renormalized theory in which $t_{2v}$ may be considered
611: a small perturbation.
612:
613: It is convenient to regroup the longitudinal contribution to
614: $\tilde{H}_c$ along with the terms in $\tilde{H}_g$ and ${\cal
615: H}^{Z_2}_{2v}$ into vortex ``potential'' and ``kinetic'' terms,
616: $H_N$ and $H^{Z_2}_{kin}$. We thereby arrive at the final form for
617: the $Z_2$ vortex-spinon Hamiltonian,
618: \begin{equation}
619: \tilde{H}_{Z_2} = H_{pl} + H_N + H^{Z_2}_{kin} +
620: \tilde{H}_f^{Z_2}, \label{HamZ2vs}
621: \end{equation}
622: with the fermion Hamiltonian $\tilde{\cal H}_s^{Z_2}$ given in
623: Eq.~(\ref{HtildefermZ2}) and with,
624: \begin{eqnarray}
625: \label{eq:z2vortexspinon}
626: {\cal H}_{pl} & = & {u_v \over 2} [e^t_j({\sf
627: r})\!+\!\epsilon_{jk}A^l_k({\sf r}+\overline{\bf w})]^2 \nonumber
628: \\ && +
629: \frac{v_0^2}{2u_v}
630: [\epsilon_{ij} \partial_i a^t_j({\sf r})\! - \pi \rho_0]^2 , \\
631: H_{N} & = & {u_v \over 2} \sum_{{\sf r},{\sf r}^\prime}
632: (\hat{N}_{\sf r}-\frac{B_{\sf r}}{\pi})
633: (\hat{N}_{{\sf r}^\prime}-\frac{B_{\sf r}}{\pi}) V({\sf r} - {\sf
634: r}^\prime) , \\
635: {\cal H}^{Z_2}_{kin} & = & {\cal H}^{Z_2}_v + {\cal H}^{Z_2}_{2v}
636: + {\cal H}^{Z_2}_r.
637: \end{eqnarray}
638: Here $B_{\sf r} = \epsilon_{ij}\partial_i A_j({\sf r}-{\bf w})$ is the
639: physical flux through the plaquette of the original lattice located at
640: the site ${\sf r}$ of the dual lattice.
641: In the ``plasmon'' Hamiltonian, ${\cal H}_{pl}$, an implicit sum
642: over $j=1,2$ is understood and we have defined a (bare) plasmon
643: velocity, $v_0$, as $v_0^2 = 2 u_ct_c = 2u_c t_v /\pi^2$. Inside
644: the superconducting phase, this Hamiltonian describes the
645: Goldstone mode - or sound mode - and can be readily diagonalized
646: to give the dispersion, $\omega^2_{pl}({\bf k}) = v_0^2 \sum_j [2
647: \sin(k_j/2)]^2$ with $|k_j| < \pi$ in the first Brillouin zone.
648: Since the electron charge density is given by, ${1 \over \pi}
649: \epsilon_{ij}
650: \partial_i a_j$, one can readily include long-ranged Coulomb interactions
651: which will modify the plasmon at small ${\bf k}$.
652:
653:
654: In $H_N$ above we have set $K=0$, dropping henceforth
655: the term $\delta H_N(K) = - K\sum_{\sf r}
656: (-1)^{N_{\sf r}}$, since it will not play an important role in the
657: phases of interest.
658: The vortex-vortex interaction energy $V({\sf r})$ is the Fourier
659: transformation of the discrete inverse Laplacian operator,
660: $V^{-1}({\bf k}) \equiv {\cal K}^2({\bf k})$, with ${\cal
661: K}^2({\bf k}) = \sum_j 2(1-\cos k_j)$, and has the expected
662: logarithmic behavior at large distances, $V({\sf r}) \sim {1 \over
663: 2 \pi} \ln(|{\sf r}|)$. The vortex kinetic energy, ${\cal
664: H}^{Z_2}_{kin}$, is a sum of three contributions - a single vortex
665: hopping term,
666: \begin{equation}
667: {\cal H}^{Z_2}_v = - t_{v} \sum_j \overline{\sigma}_j^x({\sf
668: r})\cos (\partial_j \theta_{\sf r}- a_j^t) ,
669: \end{equation}
670: a pair-vortex hopping term ${\cal H}^{Z_2}_{2v}$ given in
671: Eq.~(\ref{pairvhop}), and a ``roton'' hopping term,
672: \begin{equation}
673: {\cal H}^{Z_2}_r = -\kappa_r \overline{\sigma}_2^x({\sf
674: r}) \overline{\sigma}_2^x({\sf
675: r}+\hat{\bf x}_1)\cos(\Delta_{xy} \theta_{\sf r} -
676: \partial_x a^t_y({\sf r})) + (x\leftrightarrow y) ,
677: \end{equation}
678: where we have defined,
679: \begin{equation}
680: \Delta_{xy} \theta_{{\sf r}} = \Delta_{yx} \theta_{{\sf r}} \equiv
681: \sum_{e_1,e_2 = 0,1} (-1)^{e_1 + e_2} \theta_{{\sf r}+ e_1 {\bf
682: \hat{x}}+ e_2 {\bf \hat{y}} } . \label{deltaxy}
683: \end{equation}
684: Notice that ${\cal H}^{Z_2}_r$ hops two vortices, originally at
685: sites of the dual lattice on opposite corners of an elementary
686: square, to the other two sites. Equivalently, this term can be
687: interpreted as hopping a vortex-antivortex pair on neighboring
688: sites (ie. a vortex ``dipole'' or more simply a {\it roton}) one
689: lattice spacing in a direction perpendicular to the dipole. Such
690: a roton is a 2d analog of a 3d vortex ring, and in a Galilean
691: invariant superfluid (such as $4-He$) vortex rings propagate in
692: precisely this manner. Henceforth we shall refer to this process
693: as a ``roton hopping'' process.
694:
695:
696:
697: The above $Z_2$ vortex-spinon Hamiltonian must be supplemented by
698: the two gauge constraints, which from
699: Eqs.~(\ref{eq:Z2constraint2a},\ref{eq:Z2constraint2b}) and
700: (\ref{eq:vortexvarsN}) can be cast into an appealingly simple and
701: symmetrical form:
702: \begin{equation}
703: \tilde{\cal C}_{\bf r}^1 = (-1)^{n^f_{\bf r}} \prod_{\Box({\bf
704: r})} \overline{\sigma}^x = 1,
705: \label{Z2con1}
706: \end{equation}
707: \begin{equation}
708: \tilde{\cal C}^2_{\sf r} = (-1)^{N_{\sf r}} \prod_{\Box({\sf
709: r})} \sigma^z = 1.
710: \label{Z2con2}
711: \end{equation}
712:
713:
714: These constraints correspond to an attachment of a $Z_2$ flux in
715: $\sigma^z$ and $\overline{\sigma}^x$ to the vortex number {\it
716: parity} and the spinon number parity, respectively. Since the
717: spinons are minimally coupled to $\sigma^z$ and the vortices to
718: $\overline{\sigma}^x$, this implies a sign change upon hopping a
719: spinon around a vortex -- or vice versa. Indeed, if the partition
720: function for the vortex-spinon Hamiltonian (without ${\cal
721: H}^{Z_2}_r$ and with $t_e=0$) is expressed as an imaginary time
722: path integral with the $Z_2$ constraints in
723: Eqs.~(\ref{Z2con1},~\ref{Z2con2}) imposed, the resulting Euclidean
724: action becomes identical to Eqs (109)-(113) in
725: Ref.~[\onlinecite{Z2}].
726:
727: It is instructive to obtain explicit expressions for the electron
728: and Cooper pair creation operators in terms of the dualized vortex
729: degrees of freedom. From Eq.~(\ref{eq:duality}) we can directly
730: obtain an expression for the Cooper pair destruction operator,
731: $B_{\bf r} = e^{-2i \phi_{\bf r}}$, as
732: \begin{equation}
733: B_{\bf r} = \prod_{\bf r}^{\infty} e^{2\pi i \overline{e}_j({\bf r}')
734: d{\bf r}'_j} ,
735: \label{pairop}
736: \end{equation}
737: where the $\prod$ symbol here denotes a product along a semi-infinite
738: ``directed'' string running on the links of the original lattice,
739: originating at ${\bf r}$ and terminating at spatial infinity, with
740: $d{\bf r}'$ the unit vector from the point ${\bf r}'$ along the string
741: to the next point. In terms of $e_j$ (rather than $\overline{e}_j$),
742: the product contains a factor of $\exp(\pm 2\pi i e_j({\sf r}))$ for
743: each link of the dual lattice that crosses the string, taking the
744: positive/negative sign for directed links crossing the string from
745: right/left to left/right proceeding from ${\bf r}$ to $\infty$. We
746: will use the above notation when possible to present precise analytic
747: expressions for such strings.
748: The path independence of the string is assured by the second gauge
749: constraint, ${\cal C}^2_{\sf r} = 1$. Since the unitary
750: transformation, $U$ in Eq.~(\ref{eq:unitary}), commutes with
751: $e^{2\pi i e_j}$, this is the correct expression for the Cooper
752: pair operator within the $Z_2$ vortex-spinon theory.
753:
754: An expression for the electron operator, $c_{{\bf r}\sigma} =
755: b_{\bf r} f_{{\bf r}\sigma}$, in the dual vortex-spinon theory can
756: be extracted by moreover re-expressing the ``chargon'' operator
757: $b_{\bf r} = e^{-i\phi_{\bf r}}$ as a string,
758: \begin{equation}
759: b_{\bf r} = \prod_{\bf r}^\infty e^{i\pi \overline{e}_j d{\bf r}'_j} = {\cal
760: S}_{vort}({\bf r}) {\cal S}_\phi({\bf r}) . \label{bstring}
761: \end{equation}
762: For later convenience, we have here decomposed this expression
763: into a piece depending on the vortex configurations through the
764: longitudinal ``electric'' field, $e_j^{\ell}$, and a contribution
765: depending on the smooth part of the phase, $\phi$, through the
766: transverse field, $e_j^t$:
767: \begin{equation}
768: {\cal S}_{vort}({\bf r}) = \prod_{\bf r}^\infty
769: e^{i \pi \overline{e}_j^{t}d{\bf r}'_j} ; \hskip1cm {\cal
770: S}_\phi({\bf r}) = \prod_{\bf
771: r}^\infty e^{i\pi \overline{e}_j^l d{\bf r}'_j} . \label{SvortSphi}
772: \end{equation}
773: But unlike the Cooper pair operator, the ``chargon'' operator
774: transforms non-trivially under the unitary transformation in
775: Eq.~(\ref{eq:unitary}):
776: \begin{equation}
777: \tilde{b}_{\bf r} = U^\dagger b_{\bf r} U = \prod_{\bf r}^\infty
778: [\sigma^z_i e^{i\pi \overline{e}_j d{\bf r}'_j} ] , \label{btilde}
779: \end{equation}
780: now including a factor of $\sigma^z_j({\bf r})$ for each link of
781: the string. Again, the path independence is guaranteed by the
782: gauge constraint, $\tilde{\cal C}^2_{\sf r} = 1$.
783: The final expression for the electron operator within the dualized
784: $Z_2$ vortex-spinon field theory follows simply as,
785: \begin{equation}
786: \tilde{c}_{{\bf r}\sigma} = U^\dagger c_{{\bf r}\sigma} U =
787: \tilde{b}_{\bf r} f_{{\bf r}\sigma} .
788: \label{electronZ2}
789: \end{equation}
790:
791: As discussed at length in Ref.~[\onlinecite{Z2}], the $Z_2$
792: vortex-spinon formulation is particularly well suited for
793: accessing spin-charge separated insulating states. Specifically,
794: when pairs of vortices hop around they ``see'' an average gauge
795: flux of, $2 \epsilon_{ij}
796: \partial_i a_j = 2\pi \rho_0$. Thus at half-filling with
797: $\rho_0=1$, vortex-pairs are effectively moving in zero flux and
798: at large pair-hopping strength, $t_{2v} \rightarrow \infty$, will
799: readily condense - driving the system into an insulating state
800: with a charge gap. In the simplifying limit with vanishing {\it
801: single} vortex hopping strength, $t_v=0$, all of the vortices are
802: paired, $(-1)^N =1$, and the $Z_2$ gauge constraint in
803: Eq.~(\ref{Z2con2}) reduces to $\prod_{\Box} \sigma^z
804: =1$. The ``vison'' excitations (plaquettes with $\prod_ {\Box}
805: \sigma^z = -1$) are gapped out of the ground state, and the
806: spinons, being minimally coupled to $\sigma^z$, can propagate as
807: deconfined excitations. The charge sector supports gapped but
808: deconfined chargon excitations, which can be viewed as topological
809: defects in the pair-vortex condensate.
810:
811: But as we shall see below, to access the new Roton Fermi Liquid phase
812: requires taking the strength of the roton hopping strength large, and
813: the $Z_2$ formulation proves inadequate. To remedy this, we introduce
814: in subsection C below, a new $U(1)$ formulation of the vortex-spinon
815: field theory. As we shall demonstrate, the $U(1)$ and $Z_2$
816: vortex-spinon formulations are formally equivalent, and by a sequence
817: of unitary transformations it is possible to pass from one
818: representation to the other. Care should be taken when considering
819: operators which transform non-trivially under the unitary operations
820: related different representations, however. A third dual vortex
821: formulation involving electron (rather than spinon) operators is
822: briefly discussed below in Appendix \ref{ap:elecform}. The Hamiltonian
823: in this ``vortex-electron'' formulation is equivalent under a sequence
824: of unitary transformations to both the $Z_2$ and $U(1)$ vortex-spinon
825: Hamiltonians.
826:
827: To establish these equivalences, it is convenient to ``choose a
828: gauge'' in the $Z_2$ theory. As detailed in Appendix
829: \ref{ap:enslaveZ2}, it is possible to unitarily transform to a
830: basis in which the $Z_2$ gauge fields are completely ``slaved'' to
831: the vortex and spinon operators, and can be eliminated completely
832: from the theory. Specifically, in the chosen gauge the
833: $x-$components of both $\sigma^z$ and $\overline{\sigma}^x$ are
834: set to unity on every link of the lattice:
835: $\overline{\sigma}_1^{x}({\sf r})= \sigma_1^{z}({\bf r}) = 1$. As
836: we shall see in subsection C below, the $U(1)$ gauge fields in the
837: $U(1)$ vortex-spinon formulation can be similarly enslaved.
838: Remarkably one arrives at the {\it identical} ``enslaved''
839: Hamiltonian in both cases, thereby establishing the formal
840: equivalence between the two formulations.
841:
842:
843:
844:
845: \subsection{U(1) Vortex-spinon formulation}
846: \label{sec:u1-vortex-spinon}
847:
848:
849: In the $U(1)$ formulation of the vortex-spinon field theory, the
850: Pauli matrices $\sigma^z, \overline{\sigma}^x$ which live as $Z_2$
851: gauge fields on the links of the original and dual lattice,
852: respectively, are
853: effectively replaced by exponentials of two
854: $U(1)$ gauge fields: $\overline{\sigma}^x_j({\sf r}) \rightarrow
855: exp[i \alpha_j({\sf r})]$ and $\sigma^z_j({\bf r})
856: \rightarrow
857: exp[i\beta_j({\bf r})]$. These two $U(1)$ gauge fields are
858: canonically conjugate variables taken to satisfy,
859: \begin{equation}
860: [\alpha_i({\sf r} - {\bf \hat{x}}_i),\beta_j({\bf r}^\prime)] = i
861: \pi \epsilon_{ij} \delta^2({\bf r} - {\bf r}^\prime) ,
862: \end{equation}
863: with ${\sf r} = {\bf r} + {\bf w}$. For two ``crossing'' links
864: these commutation relations imply that the exponentials,
865: $e^{i\alpha}, e^{i\beta}$, anticommute with one another, $[e^{i
866: \alpha}, e^{i \beta}]_- =0$.
867:
868: \subsubsection{$U(1)$ Vortex-spinon Hamiltonian}
869: \label{sec:u1-vortex-spinon-1}
870:
871: The full Hamiltonian for the $U(1)$ vortex-spinon field theory
872: takes the same form as the $Z_2$ vortex-spinon Hamiltonian in
873: Eq.~(\ref{HamZ2vs}),
874: \begin{equation}
875: H = H_{pl} + H_N + H_{kin} + H_f,
876: \label{HamU1vs}
877: \end{equation}
878: with $H_{pl}$ and $H_N$ given as before in
879: Eq.~(\ref{eq:z2vortexspinon}). Only the vortex kinetic terms and
880: the fermion Hamiltonian are modified. Once again the vortex
881: kinetic energy terms are decomposed into single vortex,
882: pair-vortex and roton hopping processes:
883: \begin{equation}
884: {\cal H}_{kin} = {\cal H}_v + {\cal H}_{2v}
885: + {\cal H}_r .
886: \end{equation}
887: Since the vortices in the $U(1)$ formulation are minimally coupled
888: to the $U(1)$ gauge field, $\alpha_j({\sf r})$, each of these
889: three terms will be modified from their $Z_2$ forms. Specifically,
890: in terms of the associated Hamiltonian densities we have,
891: \begin{equation}
892: {\cal H}_v = -t_v \sum_{j=1,2} \cos(\partial_j \theta - a^t_j + \alpha_j) ,
893: \label{vorthop}
894: \end{equation}
895: \begin{equation}
896: {\cal H}_{2v} = -t_{2v} \sum_{j=1,2}
897: \cos(2\partial_j \theta - 2a^t_j + 2\alpha_j) ,
898: \label{vortpairhop}
899: \end{equation}
900: \begin{equation}
901: {\cal H}_{r} = -{\kappa_r \over 2} \cos[\Delta_{xy} \theta -
902: \partial_x(a^t_y - \alpha_y)] + (x \leftrightarrow y) ,
903: \label{rothop}
904: \end{equation}
905: with $\Delta_{xy} \theta_{{\sf r}}$ defined in
906: Eq.~({\ref{deltaxy}).
907:
908:
909:
910:
911:
912: The Hamiltonian density for the fermions in the $U(1)$ formulation
913: is given by
914: \begin{eqnarray}
915: {\cal H}_f & = & - \sum_j e^{i\beta_j({\bf r})} [ (t_s + t_e e^{i(\pi
916: \overline{e}_j({\bf r})-A_j({\bf r}))})
917: f^\dagger_{{\bf r} + {\bf \hat{x}}_j \sigma}
918: f^{\vphantom\dagger}_{{\bf r} \sigma} \nonumber \\
919: & & + \Delta_j
920: [{\cal S}_{\bf r}]^2 f_{{\bf r}+\hat{\bf
921: x}_j \sigma} \epsilon_{\sigma\sigma'}f_{{\bf r}\sigma'}+ {\rm
922: h.c.}]
923: \label{eq:Hf}
924: \end{eqnarray}
925: In the $U(1)$ formulation, the (average) density of spinons,
926: $\langle f^\dagger_{\sigma} f^{\vphantom\dagger}_{\sigma} \rangle
927: = \langle {1 \over \pi} \epsilon_{ij} \partial_i \alpha_j \rangle
928: $ (as follows from Eq.~(\ref{U1constraint})) is taken to be {\it
929: equal} to the (average) charge density, $\langle {1 \over \pi}
930: \epsilon_{ij}
931: \partial_i a_j \rangle$.
932: A new element, not present in the $Z_2$ fermion Hamiltonian,
933: ${\tilde{\cal H}}^{Z_2}_s$ in Eq.~(\ref{HtildefermZ2}), is the
934: ``string operator'', ${\cal S}_{\bf r}$. The string operator is
935: given as a product of $e^{i\beta}$ running along directed links of
936: the original lattice from the site ${\bf r}$ to spatial infinity:
937: \begin{equation}
938: {\cal S}_{\bf r} = \prod_{\bf r}^\infty e^{i\beta_j d{\bf r}'_j} .
939: \end{equation}
940: As we shall discuss below, once we restrict the Hilbert space to
941: gauge invariant states other choices for the ``path'' of the
942: string are formally equivalent.
943:
944:
945:
946: In addition to global spin and charge conservation, the full
947: $U(1)$ vortex-spinon Hamiltonian, ${\cal H}$ above, has a number
948: of {\it local} gauge symmetries. To fully define the model we need
949: to specify the set of gauge invariant states that are allowed.
950: Associated with each of the ``Chern-Simons'' fields, $\alpha$ and
951: $\beta$, is a $U(1)$ gauge symmetry. Specifically, the full
952: Hamiltonian is invariant under {\it independent} gauge
953: transformations:
954: \begin{eqnarray}
955: e^{-i \theta_{\sf r}} & \rightarrow & e^{-i \theta_{\sf r}} e^{i \chi_r}, \nonumber \\
956: \alpha_j({\sf r}) & \rightarrow & \alpha_j({\sf r}) + \partial_j \chi_{\sf r} ,
957: \label{Hgauge1}
958: \end{eqnarray}
959: and
960: \begin{eqnarray}
961: f_{{\bf r}\sigma} & \rightarrow & f_{{\bf r}\sigma}e^{i\Lambda_{\bf
962: r}}, \nonumber \\
963: \beta_j({\bf r}) & \rightarrow & \beta_j({\bf r}) +
964: \partial_j \Lambda_{\bf r},
965: \label{Hgauge2}
966: \end{eqnarray}
967: for arbitrary real functions, $\Lambda_{\bf r}$ and $\chi_{\sf r}$,
968: living on the original and dual lattices, respectively.
969: The corresponding operators which
970: transform the fields in this way are,
971: \begin{equation}
972: {\cal G}_v(\chi_{\sf r}) = e^{-i \sum_{\sf r} \chi_{\sf r} [
973: N_{\sf r} - {1 \over \pi} \epsilon_{ij} \partial_i \beta_j({\sf r}
974: - {\bf w}) ]} , \label{U1opvortex}
975: \end{equation}
976: and
977: \begin{equation}
978: {\cal G}_f(\Lambda_{\sf r}) = e^{i \sum_{\bf r} \Lambda_{\bf r}
979: [ f^\dagger_{{\bf r} \sigma}
980: f_{{\bf r} \sigma} - {1 \over \pi}
981: \epsilon_{ij} \partial_i \alpha_j({\bf r} - {\bf w}) ]} .
982: \label{U1opspinon}
983: \end{equation}
984: Both of these operators commute with the full Hamiltonian
985: $H$. The $U(1)$ sector is specified by simply setting ${\cal
986: G}_v = {\cal G}_f =1$ for arbitrary $\chi_{\sf
987: r}$ and $\Lambda_{\bf r}$. From
988: Eqs.~(\ref{U1opvortex},\ref{U1opspinon}) this is equivalent to
989: attaching $\pi$ flux in the statistical gauge fields $\alpha$ and
990: $\beta$ to the spinons and vortices:
991: \begin{equation}
992: \epsilon_{ij} \partial_i \alpha_j({\bf r} - {\bf w}) = \pi f^\dagger_{{\bf r} \sigma}
993: f_{{\bf r} \sigma} ; \hskip0.5cm \epsilon_{ij} \partial_i \beta_j({\sf r} - {\bf w}) = \pi N_{\sf r} .
994: \label{U1constraint}
995: \end{equation}
996: Notice that this is simply the $U(1)$ analog of the $Z_2$ flux
997: attachment in Eqs.~(\ref{Z2con1},\ref{Z2con2}),
998: and implies the same sign change
999: when a spinon is transported around a vortex or vice versa. The
1000: only difference is that in the $U(1)$ formulation the phase of
1001: $\pi$ is accumulated gradually when the spinon is taken around the
1002: vortex, whereas the sign change in the $Z_2$ theory can occur when
1003: the spinon hops across a single link. The formal equivalence of
1004: the $Z_2$ and $U(1)$ formulations will be established below.
1005:
1006:
1007:
1008: As we shall see, the ``smearing'' of the accumulated $\pi$ phase
1009: change, makes the theory in the $U(1)$ formulation eminently more
1010: tractable. The one notable complication is the square of the ``string operator'',
1011: which in the $U(1)$ sector is a non-trivial function of
1012: $e^{2i\beta}$ along the string, rather than equaling unity as in
1013: the $Z_2$ sector. However, it is worth emphasizing that within the
1014: $U(1)$ sector of the theory, the value of the operator ${\cal O}_{\bf r}
1015: \equiv {\cal S}_{\bf r}^2$ is
1016: {\it independent} of the chosen path. Specifically, consider two
1017: (unitary) string operators, denoted ${\cal O}_1, {\cal O}_2$, with
1018: different paths running from the same site ${\bf r}$ off to
1019: spatial infinity. The ``difference'' between the two string
1020: operators, ${\cal O}_1^{-1}{\cal O}_2$, is a product of
1021: $e^{2i\beta}$ around {\it closed loops}. But due to the $U(1)$
1022: gauge constraint in Eq.~(\ref{U1constraint}), $\epsilon_{ij}
1023: \partial_i \beta_j = \pi N$, this product is an exponential
1024: of the total vorticity $N_{tot}$ inside the closed loops, ${\cal
1025: O}_1^{-1}{\cal O}_2 = exp(2\pi i N_{tot})$. Since the vorticity is
1026: integer, one deduces that the string operator is indeed path
1027: independent, ${\cal O}_1 = {\cal O}_2$.
1028:
1029: It will prove useful to obtain expressions for the electron and
1030: Cooper pair creation operators within the $U(1)$ vortex-spinon
1031: formulation. The Cooper pair operator has the same form as in the
1032: $Z_2$ vortex-spinon formulation, given explicitly in
1033: Eq.~(\ref{pairop}), but the electron operator is modified in a
1034: non-trivial way. As we shall check explicitly below, the electron
1035: operator within the $U(1)$ formulation involves a string depending
1036: on both the dual ``electric field'', $e_j$, as well as the
1037: statistical gauge field, $\beta_j$:
1038: \begin{equation}
1039: c_{{\bf r}\sigma} = f_{{\bf r}\sigma} \prod_{\bf r}^\infty
1040: e^{i[\pi \overline{e}_i+ \beta_i]d{\bf r}'_i} , \label{electronU1}
1041: \end{equation}
1042: with $\overline{e}_i({\bf r})$ defined in Eq.~(\ref{eq:defineebar}).
1043: The path independence follows from the condition in
1044: Eq.~(\ref{eq:vortexvarsN}), together with the second $U(1)$ gauge
1045: constraint in Eq.~(\ref{U1constraint}) above;
1046: \begin{equation}
1047: \pi \partial_j e_j({\sf r} - {\bf x}_j) = \epsilon_{ij} \partial_i
1048: \beta_j({\sf r} - {\bf w}).
1049: \end{equation}
1050: This implies an equality between the longitudinal ``electric
1051: field'' and the transverse statistical field:
1052: \begin{equation}
1053: \beta^t_1({\bf r}) = \pi e^l_2({\bf r} + \overline{\bf w}) ;
1054: \hskip1cm \beta^t_2({\bf r}) = - e^l_1({\bf r} -
1055: \overline{\bf w}) ,
1056: \end{equation}
1057: or
1058: \begin{equation}
1059: \label{eq:bteqebt}
1060: \beta_i^t = -\pi \overline{e}_i^t,
1061: \end{equation}
1062: and enables the electron destruction operator to be re-expressed
1063: as,
1064: \begin{equation}
1065: c_{{\bf r}\sigma} = f_{{\bf r}\sigma} {\cal S}_\phi({\bf r})
1066: \prod_{\bf r}^\infty e^{i\beta_j^l d{\bf r}'_j} \label{elecspinon}
1067: \end{equation}
1068: with ${\cal S}_\phi({\bf r})$ defined in Eq.~(\ref{SvortSphi}).
1069: Similarly, the string operator that enters into the $U(1)$ fermion
1070: Hamiltonian, $H_f$ in Eq.~(\ref{eq:Hf}), can be written as,
1071: \begin{equation}
1072: {\cal S}_{\bf r} = {\cal S}_{vort}({\bf r})
1073: \prod_{\bf r}^\infty e^{i\beta^l_j d{\bf r}'_j} .
1074: \end{equation}
1075: Upon combining the above two equations, and recalling the identity for
1076: the ``chargon'' operator in the original $Z_2$ theory, $b_{\bf r} =
1077: {\cal S}_{vort}({\bf r}) {\cal S}_\phi({\bf r})$ in
1078: Eq.~({\ref{bstring}), implies that, ${\cal S}^\dagger_{\bf r} f_{{\bf
1079: r}\sigma} = b^\dagger_{\bf r} c_{{\bf r}\sigma}$. Consequently,
1080: ``spinon pairing'' terms in the $Z_2$ gauge theory are seen to be
1081: equivalent to the usual Bogoliubov-deGennes form:
1082: \begin{equation}
1083: [{\cal S}_{\bf r} ]^2 \epsilon_{\sigma \sigma^\prime}
1084: f_{{\bf r}\sigma} f_{{\bf r}\sigma^\prime} = B^\dagger_{\bf r}
1085: \epsilon_{\sigma \sigma^\prime} c_{{\bf r}\sigma} c_{{\bf
1086: r}\sigma^\prime} ,
1087: \label{eq:bdgform}
1088: \end{equation}
1089: with $B^\dagger_{\bf r}$ the Cooper pair creation operator. For
1090: $d$-wave pairing, the pair field lives on links, and a similar identity
1091: obtains:
1092: \begin{eqnarray}
1093: \label{eq:dwaveid}
1094: \!\!\! e^{i\beta_j({\bf r})}
1095: [{\cal S}_{\bf r}]^2 f_{{\bf r}+\hat{\bf
1096: x}_j \sigma} \epsilon_{\sigma\sigma'}f_{{\bf r}\sigma'} & = &
1097: \!\!\!B_{{\bf r},{\bf r}+\hat{\bf x}_j}^\dagger
1098: c_{{\bf r}+\hat{\bf
1099: x}_j \sigma} \epsilon_{\sigma\sigma'}c_{{\bf
1100: r}\sigma'}. \end{eqnarray}
1101: Here we have introduced a bond-Cooper pair operator,
1102: \begin{eqnarray}
1103: \label{eq:bondcp}
1104: B_{{\bf r},{\bf r}+\hat{\bf x}_j}^\dagger & = & \tilde{b}^\dagger_{\bf
1105: r} \tilde{b}^\dagger_{{\bf r}+\hat{\bf x}_j}, \\
1106: \tilde{b}^\dagger_{\bf r} & = & \prod_{\bf r}^\infty e^{-i\pi
1107: \overline{e}_i d{\bf r}'_i}.
1108: \end{eqnarray}
1109: Note that in the un-transformed chargon variables of the $Z_2$
1110: gauge theory formulation, $B_{{\bf r},{\bf r}+\hat{\bf
1111: x}_j}^\dagger=\sigma^z_j({\bf r})b^\dagger_{\bf r} b^\dagger_{{\bf
1112: r}+\hat{\bf x}_j}$.
1113:
1114: Recalling the discussion in Sec.~\ref{sec:z_2-chargon-spinon}, it is
1115: still the case that for $\Delta_j=0$, when this pairing term is absent,
1116: the fermion number is conserved, and naively may be chosen arbitrarily.
1117: However, in the limit $\Delta_j\rightarrow 0$, which we consider here,
1118: this conservation is weakly violated, and only the total charge is
1119: conserved. The physics at work is clear from Eq.~(\ref{eq:bdgform}):
1120: for non-vanishing $\Delta_j$, quasiparticle pairs and boson pairs are
1121: interchanged, and the two charged fluids come to equilibrium. Thus in
1122: what follows, we should choose to divide the charge density amongst the
1123: fermions and bosons in such a way as to minimize the total (free)
1124: energy. This division will therefore shift as parameters of the
1125: Hamiltonian are changed. How it does so is crucial to the ultimate
1126: low energy physical properties of the system, as is clear from
1127: Eqs.~(\ref{vorthop})-~(\ref{rothop}), which show that the vortices
1128: experience an effective ``flux'' proportional to the difference of the
1129: total charge density ($\epsilon_{ij}\partial_i a_j/\pi$) and the
1130: fermionic density $n_f=\epsilon_{ij}\partial_i\alpha_j/\pi$. As the
1131: fermion density is varied, the effective flux seen by the vortices
1132: changes. Significantly, in the limit $t_v\rightarrow \infty$, vortex
1133: hopping dominates the energetics, and is minimized when the fermion
1134: density equals the total charge density, $n_f =
1135: \epsilon_{ij}\partial_i a_j/\pi$. This naturally
1136: recovers the Fermi liquid phase (App.~\ref{sec:fermiliquid}) by binding
1137: charge $e$ firmly to each fermion, fully accommodating all the electrical
1138: charge.
1139:
1140: In the rest of the paper we work exclusively within the $U(1)$
1141: vortex-spinon formulation, which is particularly suitable for
1142: extracting the properties of the Roton Fermi liquid. Before
1143: embarking on that, we first establish the formal equivalence
1144: between the two formulations by enslaving the $U(1)$ gauge fields.
1145: As detailed in Appendix ~\ref{ap:enslaveU1}, it is possible to
1146: unitarily transform to a gauge with $\vec{\nabla}\cdot\vec{\alpha}
1147: = \vec{\nabla}\cdot\vec{\beta}^{slave} = 0$. In this gauge, both
1148: $\alpha_j$ and $\beta_j$ are ``enslaved", being fully expressible
1149: in terms of the spinon and vortex densities, $n^f_{\bf r}$ and
1150: $N_{\sf r}$, respectively. Moreover, the enslaved $U(1)$
1151: Hamiltonian is found to be {\it identical} to the enslaved $Z_2$
1152: Hamiltonian obtained in Appendix ~\ref{ap:enslaveZ2}, and the
1153: enslaved expressions for the electron operators also
1154: coincide.
1155:
1156: Having thereby established the equivalence between the $Z_2$ and
1157: $U(1)$ vortex-spinon field theories, in the remainder of the paper
1158: we choose to work exclusively within the $U(1)$ formulation,
1159: employing the Hamiltonian $H$ defined in Eq.~(\ref{HamU1vs}),
1160: together with the gauge constraints in Eq.~(\ref{U1constraint}).
1161: In practice, it is far simpler to work within a Lagrangian
1162: formulation, where the gauge constraints can be imposed explicitly
1163: within a path integral, as detailed in the next subsection.
1164:
1165:
1166:
1167:
1168:
1169: \subsubsection{Lagrangian for $U(1)$ Vortex-spinon theory }
1170: \label{sec:lagr-form-u1}
1171:
1172: In order to impose the $U(1)$ gauge constraints in
1173: Eq.~(\ref{U1constraint}) on the Hilbert space of the full
1174: vortex-spinon Hamiltonian, $H$, we will pass to a Euclidean
1175: path integral representation of the partition function. The
1176: associated Euclidean Lagrangian is readily obtained as a sum of
1177: three contributions,
1178: \begin{equation}
1179: S = \int d \tau [ H + L_{B} + L_{con}],
1180: \label{Sfulla}
1181: \end{equation}
1182: with $L_{B}$ involving the generalized coordinates and conjugate momenta,
1183: \begin{eqnarray}
1184: L_{B} & = &
1185: \sum_{{\bf r} = {\sf r} + {\bf w}}
1186: [i N_{\sf r} \partial_0 \theta_{\sf r}
1187: - {i \over \pi} \beta_i({\bf r}) \epsilon_{ij} \partial_0 \alpha_j ({\sf r} + \hat{x}_i )] \nonumber \\
1188: & + & \sum_{{\bf r} = {\sf r} + {\bf w}}
1189: [i e^t_j({\sf r}) \partial_0 a^t_j({\sf r})
1190: + f_{{\bf r} \sigma}^\dagger \partial_0
1191: f^{\vphantom\dagger}_{{\bf r} \sigma} ] ,
1192: \end{eqnarray}
1193: with $\partial_0 \equiv \partial/\partial \tau$ denoting an
1194: imaginary time derivative and $L_{con}$ is a Lagrange multiplier
1195: term imposing the two independent $U(1)$ gauge constraints,
1196: \begin{eqnarray}
1197: L_{con} & = & {i \over \pi} \sum_{{\bf r} = {\sf r} - {\bf w}}
1198: \alpha_0({\sf r}) [\epsilon_{ij} \partial_i \beta_j({\bf r}) - \pi N_{\sf r} ] \nonumber \\
1199: & +& {i \over \pi} \sum_{{\bf r} = {\sf r} + {\bf w}}
1200: \beta_0({\bf r})
1201: [ \epsilon_{ij} \partial_i \alpha_j({\sf r}) - \pi f_{{\bf r}
1202: \sigma}^\dagger f^{\vphantom\dagger}_{{\bf r} \sigma} ] .
1203: \end{eqnarray}
1204: Here we have introduced two Chern-Simons scalar potentials as
1205: Lagrange multipliers, denoted $\alpha_0({\sf r})$ and
1206: $\beta_0({\bf r})$, which live on the dual and original lattice
1207: sites respectively.
1208:
1209: Upon introducing another scalar potential, $a_0({\sf r})$, living on the sites
1210: of the dual lattice, and collecting together the longitudinal and transverse parts
1211: of $a_j$ and $e_j$, the full Euclidean action
1212: can be compactly cast into a simple form.
1213: In order to make the vortex physics more explicit we choose to re-introduce the
1214: vortex phase-field $\theta$ within the Lagrangian formulation.
1215: Specifically, we shift
1216: $a_\mu \rightarrow a_\mu + \partial_\mu \theta$
1217: with $\mu = 0,x,y$,
1218: and then integrate over {\it both} $a_\mu$ and $\theta$.
1219: In this way we arrive at the final form for the full Euclidean Lagrangian:
1220: \begin{equation}
1221: S = S_c + S_f + S_{cs} + S_A .
1222: \label{Sfullb}
1223: \end{equation}
1224: The charge sector action $S_c =\int_\tau \sum_{\sf r} {\cal
1225: L}_c$ can be expressed in terms of the Lagrangian density,
1226: \begin{equation}
1227: {\cal L}_c= {\cal L}_{a} + {1 \over 2u_0} (\partial_0 \theta - a_0 + \alpha_0)^2 + {\cal L}_{kin},
1228: \label{Lc}
1229: \end{equation}
1230: with the $u_0 \rightarrow 0$ limit enslaving $a_0 = \partial_0 \theta +
1231: \alpha_0$ and
1232: \begin{equation}
1233: {\cal L}_{a} = {u_v \over 2} (\overline{e}_j-\frac{A_j}{\pi})^2 - i e_j
1234: (\partial_0 a_j -
1235: \partial_j a_0) + {v_0^2 \over 2 u_v} (\epsilon_{ij} \partial_i a_j - \pi
1236: \rho_0)^2 .
1237: \end{equation}
1238: The vortex kinetic
1239: energy terms ${\cal L}_{kin}$ are given explicitly by ${\cal H}_{kin}$ in Eqs.~(\ref{vorthop}-\ref{rothop})
1240: except with $a_j^t \rightarrow a_j$.
1241:
1242:
1243:
1244: The fermion action $S_f = \int_\tau \sum_{\bf r} {\cal L}_f$ is
1245: given by,
1246: \begin{equation}
1247: \label{eq:Lf}
1248: {\cal L}_f = f_{{\bf r} \sigma}^\dagger (\partial_0 - i \beta_0)
1249: f^{\vphantom\dagger}_{{\bf r} \sigma} + {\cal H}_f ,
1250: \end{equation}
1251: with ${\cal H}_f$ given in Eq.~(\ref{eq:Hf}). The two sectors are
1252: coupled together by the electric field in the electron hopping term, and
1253: by the Chern-Simons action $S_{cs} = \int_\tau \sum_{{\bf r} = {\sf r} +
1254: {\bf w}} {\cal L}_{cs}$ with
1255: \begin{eqnarray}
1256: {\cal L}_{cs} & = & {i \over \pi} \beta_0({\bf r}) \epsilon_{ij} \partial_i \alpha_j({\sf r}) \nonumber \\
1257: && + {i \over \pi} \beta_i({\bf r}) \epsilon_{ij} \big[
1258: \partial_j \alpha_0({\sf r} + \hat{x}_i ) - \partial_0 \alpha_j
1259: ({\sf r} + \hat{x}_i ) \big] .
1260: \label{Lcs}
1261: \end{eqnarray}
1262: Notice that in the absence of the electron hopping term ($t_e =
1263: 0$), the ``electric field'' $e_j$ enters quadratically in the
1264: action, and can then be integrated out to give ${\cal L}_a
1265: \rightarrow \tilde{\cal L}_a$, with
1266: \begin{equation}
1267: \tilde{\cal L}_{a} = {1 \over 2u_v }(\partial_0 a_j - \partial_j a_0)^2 +
1268: v_0^2 (\epsilon_{ij} \partial_i a_j - \pi \rho_0)^2 .
1269: \end{equation}
1270:
1271:
1272: Finally, it is useful in some circumstances to treat the external gauge
1273: field by making the shift
1274: $\overline{e}_i \rightarrow \overline{e}_i+A_i/\pi$, which removes all
1275: coupling of $A_i$ to the fermions, and furthermore leaves $A_i$ {\sl linearly
1276: coupled} to a charge ``3-current'' of the usual form,
1277: $S_A = \int_\tau \sum_{r} {\cal L}_A$,
1278: with
1279: \begin{equation}
1280: \label{LA}
1281: {\cal L}_A = i A_\mu({\bf r}) J_\mu ({\bf r}) .
1282: \end{equation}
1283: Here $J_\mu$ is the charge 3-current given explicitly by,
1284: \begin{equation}
1285: \label{eq:Jzero}
1286: J_0({\bf r}) = {1 \over \pi} \epsilon_{ij} \partial_i a_j({\sf r}) ,
1287: \end{equation}
1288: \begin{equation}
1289: \label{eq:Jj}
1290: J_i({\bf r}) = {1 \over \pi} \epsilon_{ij} [ \partial_j a_0({\sf r} +
1291: \hat{x}_i ) - \partial_0 a_j({\sf r} +
1292: \hat{x}_i ) ] ,
1293: \end{equation}
1294: with ${\bf r} = {\sf r} + {\bf w}$. Notice that the 3-current is
1295: conserved as required: $\partial_0 J_0({\bf r}) + \partial_i J_i({\bf r}
1296: - \hat{x}_i ) =0$. This form is useful for a variety of calculations,
1297: particularly within the purely bosonic RL model discussed in
1298: Secs.~\ref{sec:roton-liquid}-\ref{sec:inst-roton-liqu}, but less so in
1299: some RFL calculations best done in the ``electron gauge'' (see below),
1300: which is incompatible with the above shift of $\overline{e}_i$.
1301:
1302: As can be seen
1303: from the equations of motion obtained from, $\delta {\cal
1304: L}/\delta \alpha =0$ and $\delta {\cal L} / \delta \beta =0$, the
1305: effect of the Chern-Simons term is to attach $\pi$ flux in
1306: $\alpha$ ($\beta$) to the spinon (vortex) world-lines;
1307: \begin{equation}
1308: \epsilon_{\mu \nu \lambda} \partial_\nu \alpha_\lambda = \pi
1309: J^s_\mu ; \hskip0.5cm \epsilon_{\mu \nu \lambda} \partial_\nu
1310: \beta_\lambda = \pi J^v_\mu ,
1311: \end{equation}
1312: where $J^s_\mu$ and $J^v_\mu$ are the spinon and vortex
1313: three-currents. Here $\mu,\nu,\lambda = 0,x,y$ run over the three
1314: space-time coordinates.
1315:
1316:
1317:
1318:
1319: Finally we comment on the nature of the gauge symmetries of the
1320: full action $S$ in the Lagrangian representation. In particular,
1321: associated with the three gauge fields $a_\mu,\alpha_\mu$ and
1322: $\beta_\mu$ are three {\it independent} space-time gauge
1323: symmetries. Specifically, these are:
1324: \begin{enumerate}
1325: \item
1326: \begin{eqnarray}
1327: \label{eq:gauge1}
1328: \theta_{\sf r} & \rightarrow & \theta_{\sf r} + \Theta_{\sf r}, \nonumber \\
1329: a_\mu({\sf r}) & \rightarrow & a_\mu({\sf r}) + \partial_\mu \Theta_{\sf
1330: r},
1331: \end{eqnarray}
1332: \item
1333: \begin{eqnarray}
1334: \label{eq:gauge2}
1335: \theta_{\sf r} & \rightarrow & \theta_{\sf r} + \chi_{\sf r}, \nonumber \\
1336: \alpha_\mu({\sf r}) & \rightarrow & \alpha_\mu({\sf r}) - \partial_\mu \chi_{\sf
1337: r},
1338: \end{eqnarray}
1339: \item
1340: \begin{eqnarray}
1341: \label{eq:gauge3}
1342: f_{{\bf r}\sigma} & \rightarrow & f_{{\bf r}\sigma}e^{i\Lambda_{\bf
1343: r}}, \nonumber \\
1344: \beta_\mu({\bf r}) & \rightarrow & \beta_\mu ({\bf r}) +
1345: \partial_\mu \Lambda_{\bf r} ,
1346: \end{eqnarray}
1347: \end{enumerate}
1348: with $\Theta_{\sf r}, \chi_{\sf r}$ and $\Lambda_{\bf r}$
1349: arbitrary functions of space and imaginary time. Because of the
1350: gauge invariance of $S$ under these three distinct
1351: transformations, we are free to fix gauges {\it independently} for
1352: the three gauge fields.
1353:
1354: In addition to these three gauge symmetries, the full Lagrangian
1355: has two {\it global} symmetries. By construction,
1356: the spinon Lagrangian ${\cal L}_s$ conserves the total
1357: spin, and since the electrical 3-current in Eqs.~(\ref{eq:Jzero}-\ref{eq:Jj}) satisfies
1358: a continuity equation, $\partial_\mu J_\mu =0$,
1359: the total electrical charge $Q = \sum_{\bf r} J_0({\bf r})$ is
1360: also conserved.
1361:
1362:
1363:
1364: A particularly convenient gauge choice for the gauge field $\beta_\mu$
1365: is,
1366: \begin{equation}
1367: \beta^l_i({\bf r}) = - \pi \overline{e}^l_i({\bf r}) ,
1368: \end{equation}
1369: with $\overline{e}_i$ defined in terms of the ``electric field''
1370: $e_j({\sf r})$ in Eq.~(\ref{eq:defineebar}). Remarkably, in this
1371: gauge the electron creation operator equals the spinon creation
1372: operator. To see this, it is convenient to shift $\alpha_0
1373: \rightarrow \alpha_0 + a_0$ and then integrate out $a_0$, which
1374: constrains $\vec\nabla\times\vec\beta = \pi\vec\nabla\cdot
1375: \vec{e}$, or equivalently $\beta_i^t = -\pi \overline{e}^t_i$.
1376: Together with the above gauge choice this implies that $\beta_i
1377: \equiv - \pi \overline{e}_i$, so that from Eq.~(\ref{electronU1})
1378: one has $c_{{\bf r}\sigma} = f_{{\bf r}\sigma}$. We refer to this
1379: as the ``electron gauge''. The possibility to choose a gauge
1380: within the Lagrangian formulation of the $U(1)$ vortex-spinon
1381: theory which makes $f_{{\bf r}\sigma}$ an electron operator,
1382: suggest that it should be possible to re-formulate the
1383: vortex-spinon {\it Hamiltonian} entirely in terms of vortices and
1384: electrons. This is indeed the case, as we demonstrate briefly in
1385: Appendix \ref{ap:elecform}.
1386:
1387:
1388:
1389:
1390:
1391:
1392: \section{The Roton Liquid}
1393: \label{sec:roton-liquid}
1394:
1395: We first focus on the bosonic charge sector of the theory, ignoring
1396: entirely the fermions. Specifically, in the full Euclidean action in
1397: Eq.~(\ref{Sfullb}) we retain {\it only} the charge action, $S_c$, and
1398: the coupling to the external electromagnetic field, $S_A$. We take the
1399: fermionic density as a non-fluctuating constant. As discussed in
1400: Sec.~\ref{sec:u1-vortex-spinon-1}, this constant should be determined by
1401: energetics. We assume here that the largest energy scales in the
1402: problem are those of the vortices, i.e. $\kappa_r,u_v$ etc. In this
1403: case, one expects that vortex kinetic energy (rotonic or otherwise) is
1404: minimized when the vortices experience zero average magnetic flux. We
1405: therefore choose the fermionic density {\sl equal} to the total charge
1406: density, setting ${1 \over \pi} \epsilon_{ij} \partial_i \alpha_j =
1407: \rho_0$ and also putting $\alpha_0=0$. Note that this choice is
1408: essential to recovering an ordinary Fermi liquid state (see
1409: Sec.~\ref{sec:vortex-hopping},App.~\ref{sec:fermiliquid}) and hence is also
1410: natural in this sense. We remark that, while we will continue to assume
1411: the average fermion density is equal to $\rho_0$ in the bulk of this
1412: paper, we will return to another possibility -- and its physical regime
1413: of relevance -- in the discussion section.
1414:
1415: It is also convenient to isolate the
1416: fluctuations in the charge density by defining,
1417: \begin{equation}
1418: a_j = a_j^b + \tilde{a}_j ,
1419: \end{equation}
1420: with ``background'' density ${1 \over \pi} \epsilon_{ij} \partial_i a^b_j
1421: = \rho_0 $.
1422: We can then take $\alpha_j = a^b_j$,
1423: so that $a_j - \alpha_j = \tilde{a}_j$.
1424: Of the three vortex kinetic energy processes which enter in ${\cal L}_{kin}$,
1425: we hereafter and in the rest of the paper drop entirely the vortex pair hopping process
1426: in Eq.~(\ref{vortpairhop})
1427: putting $t_{2v}=0$. For now we also set the single vortex hopping processes
1428: to zero, putting $t_v=0$ in Eq.~(\ref{vorthop}), but will return to their effects
1429: in Section IV. Of interest here is the new roton hopping process,
1430: ${\cal L}_{r} \equiv {\cal H}_{r}$ in Eq.~(\ref{rothop}), which can be conveniently recast in
1431: the form:
1432: \begin{equation}
1433: {\cal L}_{r} = - \kappa_r {\cal C}[\tilde{a}] \cos[\Delta_{xy} \theta -
1434: {1 \over 2}(\partial_x \tilde{a}_y +\partial_y \tilde{a}_x)] ,
1435: \end{equation}
1436: with
1437: \begin{equation}
1438: {\cal C}[\tilde{a}] = \cos(\epsilon_{ij}
1439: \partial_i \tilde{a}_j/2) .
1440: \label{rotamp}
1441: \end{equation}
1442: More generally, with spinon fluctuations included
1443: one has ${\cal C} = \cos[\epsilon_{ij} \partial_i (a_j - \alpha_j)/2]$.
1444: Clearly, ${\cal C}$ is maximal -- and hence the rotonic kinetic energy
1445: most negative -- for
1446: $\epsilon_{ij}\partial_i\tilde{a}_j=0$, which is true on average for
1447: this choice of fermion density.
1448:
1449: We next choose the gauge $\vec{\nabla}\cdot\vec{\tilde{a}} =0$, and
1450: integrate over $a_0$ (with $u_0 \rightarrow 0$).
1451: Having dropped the vortex hopping processes, the remaining charge Lagrangian is then given by,
1452: \begin{equation}
1453: {\cal L}_c = {\cal L}_{pl} + {\cal L}_{\theta} ,
1454: \end{equation}
1455: with
1456: \begin{equation}
1457: {\cal L}_{pl} = {1 \over 2u_v} [ (\partial_0 \tilde{a}_j)^2 + v_0^2 (\epsilon_{ij} \partial_i \tilde{a}_j)^2 ] ,
1458: \label{Lplas}
1459: \end{equation}
1460: \begin{equation}
1461: {\cal L}_{\theta} = {1 \over 2u_v} (\partial_j \partial_0 \theta)^2 -
1462: \kappa_r{\cal C}[\tilde{a}] \cos[\Delta_{xy}\theta -
1463: {1 \over 2}(\partial_1 \tilde{a}_2+\partial_2 \tilde{a}_1)] .
1464: \label{Ltheta}
1465: \end{equation}
1466: To analyze the phases of this model
1467: it is instructive to represent the Lagrangian ${\cal L}_\theta$
1468: in Hamiltonian
1469: form by re-introducing the vortex number operator, $\hat{N}_{\sf
1470: r}$:
1471: \begin{eqnarray}
1472: \hat{H}_\theta & = & {u_v \over 2} \sum_{{\sf r},{\sf r}^\prime}
1473: \hat{N}_{\sf r}
1474: \hat{N}_{{\sf r}^\prime} V({\sf r} - {\sf r}^\prime) \nonumber \\ &&
1475: \hspace{-0.3in}- \kappa_r \sum_{\sf r} {\cal C}[\tilde{a}]
1476: \cos(\Delta_{xy}\hat{\theta}_{\sf
1477: r}-\frac{1}{2}(\partial_x \tilde{a}_1+\partial_y\tilde{a}_x)) .
1478: \label{Htheta}
1479: \end{eqnarray}
1480: Again $V({\sf r})$ is Fourier
1481: transform of the discrete inverse Laplacian operator, $V({\bf k})
1482: \equiv 1/{\cal K}^2({\bf k})$, with ${\cal K}^2({\bf k}) = \sum_j
1483: 2(1-\cos k_j)$.
1484:
1485: The first term in Eq.~(\ref{Htheta}) describes a logarithmically
1486: interacting gas of (integer strength) vortices moving on
1487: the dual 2d square lattice. When $\kappa_r =0$, this model
1488: will undergo a finite temperature Kosterlitz-Thouless transition\cite{2dXY} from
1489: a high temperature ``vortex plasma'' into the low temperature ``vortex
1490: dielectric''. This corresponds, of course, to a transition into
1491: a superconducting phase. With $\kappa_r=0$ the Kosterlitz-Thouless
1492: transition temperature will be set by the vortex interaction strength,
1493: $u_v$. But upon increasing the strength of the roton hopping,
1494: one expects the transition temperature to be suppressed, and for
1495: $\kappa_r \gg u_v$ to be driven all the way to zero.
1496: Thus, at zero temperature, upon increasing the single dimensionless
1497: ratio, $\kappa_r/u_v$, one expects a quantum phase transition
1498: out of the superconducting phase and into a new phase (see Fig. 1) -
1499: the ``Roton Liquid''.
1500:
1501:
1502:
1503: \begin{figure}
1504: \begin{center}
1505: \vskip-2mm
1506: \hspace*{0mm}
1507: \centerline{\fig{2.8in}{fig1.eps}}
1508: \vskip-2mm
1509: \caption{Schematic phase diagram of the Hamiltonian in the charge sector, $H_c = H_\theta + H_{pl}$,
1510: with zero vortex hopping strength, $t_v=0$. When the roton hopping vanishes,
1511: $\kappa_r=0$, $H_c$ describes a classical
1512: logarithmically interacting vortex gas, and
1513: has a superconductor-to-normal transition
1514: at a Kosterlitz-Thouless temperature, $T_{KT} \approx u_v$.
1515: With increasing roton hopping, $T_{KT}$ decreases being driven to zero
1516: at $\kappa_r^* \approx u_v$ where there is a quantum phase transition
1517: from the superconductor into the Roton liquid phase.
1518: }
1519: \label{fig:RLphasediag}
1520: \end{center}
1521: \end{figure}
1522:
1523:
1524: \subsection{Harmonic theory and Excitations}
1525: \label{sec:harm-theory-excit}
1526:
1527: To access the properties of the Roton Liquid (RL), we consider
1528: $\kappa_r \gg u_v$, where it is presumably valid to expand
1529: the cosine terms in Eq.~(\ref{Htheta}) for small argument,
1530: giving $\hat{H}_\theta = \hat{H}_{rot} + ...$, with
1531: \begin{eqnarray}
1532: \label{eq:Hrot}
1533: \hat{H}_{rot} & = & {u_v \over 2} \sum_{{\sf r},{\sf r}^\prime} \hat{N}_{\sf r}
1534: \hat{N}_{{\sf r}^\prime} V({\sf r} - {\sf r}^\prime) +
1535: \frac{\kappa_r}{8} \sum_{\sf r} (\epsilon_{ij}\partial_i
1536: \tilde{a}_j)^2 \nonumber \\
1537: & & + \frac{\kappa_r}{2}
1538: \sum_{\sf r} [\Delta_{xy}\hat{\theta}_{\sf
1539: r}- \frac{1}{2} (\partial_x \tilde{a}_y+\partial_y \tilde{a}_x)]^2 .
1540: \end{eqnarray}
1541: With this expansion, it is no longer legitimate to
1542: restrict $\theta$ to the interval $[0,2\pi]$. Consistency then dictates that
1543: the eigenvalues of the vortex number operator no longer be restricted
1544: to integers, but allowed to take on any real value from $[- \infty, \infty]$.
1545:
1546: The full Roton Liquid Hamiltonian, $\hat{H}_{RL} =\hat{H}_{pl}+\hat{H}_{\rm rot}$ is
1547: quadratic and can be readily diagonalized. This is most conveniently
1548: done by returning to the Lagrangian framework, described now by
1549: \begin{eqnarray}
1550: {\cal L}_{RL} & = & {1 \over 2u_v} [ (\partial_0 \tilde{a}_j)^2 +
1551: \tilde{v}_0^2 (\epsilon_{ij} \partial_i \tilde{a}_j)^2
1552: ] \nonumber \\
1553: & & \hspace{-0.5in}+
1554: {1 \over 2u_v} (\partial_j \partial_0 \theta)^2 +{ \kappa_r \over 2}
1555: [\Delta_{xy} \theta -
1556: {1 \over 2}(\partial_x \tilde{a}_y+\partial_y \tilde{a}_x)]^2 ,
1557: \label{LRL}
1558: \end{eqnarray}
1559: with $\tilde{v}_0=\sqrt{v_0^2 + \kappa_r u_v/4}$.
1560: To proceed to describe the normal modes of this quadratic Lagrangian,
1561: we define Fourier transforms
1562: \begin{eqnarray}
1563: O({\bf r},\tau) & = & \int_{{\bf k},\omega_n} e^{i {\bf k}\cdot{\bf r} -i\omega_n\tau} O({\bf k},\omega_n), \\
1564: {\sf O}({\sf r},\tau) & = & \int_{{\bf k},\omega_n} e^{i {\bf k}\cdot{\sf r}-i\omega_n\tau} {\sf O}({\bf k},\omega_n),
1565: \end{eqnarray}
1566: for fields $O,{\sf O}$ on the original and dual lattices,
1567: respectively. Here integration $\int_{\bf k} \equiv \int d^2{\bf
1568: k}/(2\pi)^2$ is taken over the Brillouin zone $|k_1|,|k_2| < \pi$ and
1569: $\int_{\omega_n} \equiv \int_{- \infty}^\infty d\omega_n/(2\pi)$
1570: defines the integration measure (at zero temperature) for the
1571: Matsubara frequency $\omega_n$. At non-zero temperature, one simply
1572: replaces $\int_{\omega_n} \rightarrow \beta^{-1}\sum_{\omega_n}$, with
1573: $\omega_n=2\pi n/\beta$. It is moreover convenient to define
1574: \begin{equation}
1575: {\cal K}_j({\bf k}) = -i(e^{ik_j}-1),
1576: \end{equation}
1577: so that upon Fourier transformation, the discrete derivatives behave
1578: intuitively,
1579: \begin{equation}
1580: \partial_j \rightarrow_{FT} i {\cal K}_j({\bf k}),
1581: \end{equation}
1582: and of course $\partial_0 \rightarrow -i\omega_n$ as usual. We also
1583: introduce the transverse gauge field:
1584: \begin{equation}
1585: \tilde{a}_i({\bf k}) =\epsilon_{ij} { i{\cal K}^*_j({\bf k}) \over {\cal
1586: K}({\bf k}) } a({\bf k}) ,
1587: \end{equation}
1588: with ${\cal K}^2({\bf k}) = \sum_j |{\cal K}_j({\bf k})|^2$
1589: and $|{\cal K}_j({\bf k})| = 2 |\sin(k_j/2)|$.
1590:
1591:
1592: To now diagonalize ${\cal L}_{RL}$, it is convenient to define
1593: a real two-component field
1594: $\Upsilon_a$ via
1595: \begin{eqnarray}
1596: a({\bf k},\omega_n) & = & \sqrt{u_v} e^{i{\bf k}\cdot{\bf w}} \Upsilon_1({\bf
1597: k},\omega_n), \\
1598: \theta({\bf k},\omega_n) & = & \frac{\sqrt{u_v}}{{\cal K}({\bf
1599: k}) } \Upsilon_2({\bf k},\omega_n).
1600: \end{eqnarray}
1601: Then the action, $S_{RL}=\sum_{\sf r}\int_\tau {\cal L}_{RL}$ is
1602: \begin{eqnarray}
1603: S_{RL} & = & \frac{1}{2}
1604: \int_{{\bf k},\omega_n} \hspace{-0.2in} \Upsilon_\alpha({\bf k},\omega_n) G^{-1}_{\alpha\beta}({\bf k},\omega_n)
1605: \Upsilon_\beta({\bf -k},-\omega_n),
1606: \label{eq:upsilons}
1607: \end{eqnarray}
1608: with
1609: \begin{equation}
1610: \label{eq:gf}
1611: G_{\alpha\beta} =
1612: \frac{G^0 \delta_{\alpha\beta} + G^z \sigma^z_{\alpha\beta} + G^x
1613: \sigma^x_{\alpha\beta}}{(\omega_n^2+\omega_{pl}^2)(\omega_n^2+\omega_{rot}^2)} , \end{equation}
1614: where $\vec{\sigma}$ is the usual vector of Pauli matrices.
1615: Here we have defined,
1616: \begin{eqnarray}
1617: G^0 & = & \omega_n^2 + \frac{1}{2} v_+^2 {\cal K}^2, \\
1618: G^z & = & -\frac{1}{2} v_+^2 {\cal K}^2 + v_1^2 \frac{|{\cal
1619: K}_x{\cal K}_y|^2}{{\cal K}^2}, \\
1620: G^x & = & \frac{v_1^2}{2} (|{\cal K}_x|^2 -
1621: |{\cal K}_y|^2)\frac{\tilde{\cal K}_x\tilde{\cal K}_y}{{\cal K}^2},
1622: \label{eq:gf1}
1623: \end{eqnarray}
1624: with $v_1=\sqrt{\kappa_r u_v}$ and $\tilde{\cal K}_j = 2\sin (k_j/2)$. The poles in
1625: $G_{\alpha\beta}$ at $\omega=i\omega_n=\pm \omega_{pl}, \pm \omega_{rot}$
1626: describe two types of collective modes.
1627:
1628: The first excitation is a plasmon with a
1629: renormalized dispersion,
1630: \begin{equation}
1631: \omega_{pl}^2({\bf k}) = \frac{1}{2} \left[ v_{+}^2 {\cal K}^2 +
1632: \sqrt{ v_{-}^4 {\cal K}^4 + \tilde{v}_0^2 v_1^2 (|{\cal K}_x|^2
1633: - |{\cal K}_y|^2 )^2 }\right],
1634: \end{equation}
1635: with velocities,
1636: \begin{equation}
1637: v_{\pm}=\sqrt{\tilde{v}_0^2 \pm v_1^2/4} .
1638: \end{equation}
1639: The plasmon frequency vanishes at the center of the Brillouin zone,
1640: ${\bf k}=0$, and in the absence of long-ranged Coulomb interactions
1641: disperses linearly $\omega_{pl} = v_{pl} |{\bf k}|$ at
1642: small wavevectors, $k_j \rightarrow 0$.
1643: But the associated plasmon velocity, $v_{pl}(\phi)$,
1644: depends upon the ratio, $k_y/k_x = \tan(\phi)$.
1645: In particular, along the zone diagonals with $k_y= \pm k_x$
1646: the velocity is minimal and unaffected by the vortices with $v_{pl} = \tilde{v}_0$, whereas
1647: it takes it's maximum
1648: value, $v_+$, along the $k_x$ or $k_y$ axes.
1649:
1650:
1651: This upward shift in the plasmon frequency is due to a
1652: ``level repulsion'' with the second collective mode - the gapless roton,
1653: which disperses as,
1654: \begin{equation}
1655: \omega_{rot}^2({\bf k}) = \frac{1}{2} \left[ v_+^2 {\cal K}^2 -
1656: \sqrt{ v_{-}^4 {\cal K}^4 + \tilde{v}_0^2 v_1^2 (|{\cal K}_x|^2
1657: - |{\cal K}_y|^2 )^2 }\right] .
1658: \end{equation}
1659: For $|k_x| \ll 1$ {\sl and fixed $k_y$} the roton dispersion {\it vanishes},
1660: $\omega_{rot} \sim v_{rot} |k_x|$, with
1661: \begin{equation}
1662: v_{rot} = \frac{\tilde{v}_0}{v_{+}} v_1 .
1663: \end{equation}
1664: Remarkably, the Roton Liquid phase supports
1665: a gapless ``Bose surface'' of roton excitations,
1666: along the $k_x=0$ and $k_y=0$ axes. These roton excitations
1667: describe {\it gapless and transverse} current fluctuations,
1668: which are obviously not present in a conventional bosonic superfluid.
1669:
1670:
1671: With long-range Coulomb interactions present one would have simply,
1672: $v_0^2 \rightarrow v_0^2({\bf k}) \sim {1 \over |{\bf k}|}$, giving
1673: the familiar 2d plasmon dispersion, $\omega_{pl} \sim \sqrt{|{\bf
1674: k}|}$. In addition, the roton velocity $v_{rot}$ becomes
1675: dependent upon $k_y$. We note in passing that the roton velocity is
1676: in either case determined not only from the dynamics of $\theta$ but
1677: also that of $\tilde{a}$, as is evident from its dependence upon
1678: $\tilde{v}_0/v_{+}$. It will be sometimes instructive in the
1679: following to consider the simple limit $v_0\sim\tilde{v}_0 \rightarrow
1680: \infty$, in which the spatial fluctuations of $\tilde{a}$ vanish and
1681: the roton mode is entirely decoupled from $\tilde{a}$.
1682:
1683:
1684:
1685:
1686:
1687:
1688:
1689:
1690: \subsection{No Meissner effect in roton liquid}
1691: \label{sec:no-meissner-effect}
1692:
1693: We now employ the Gaussian theory to examine some of the electrical
1694: properties of the Roton liquid phase.
1695: Consider first the response of the RL phase to
1696: an applied magnetic field.
1697: In the presence of a magnetic field, $B = \epsilon_{ij} \partial_i A_j$, there is an additional term that
1698: one must add to the Lagrangian, which
1699: from Eq.~(\ref{LA}) takes the form:
1700: \begin{equation}
1701: {\cal L}_A = {i \over \pi} a_0 B .
1702: \end{equation}
1703: If $a_0$ is integrated out from ${\cal L}_c$ in Eq.~(\ref{Lc}) with this
1704: additional term present, the Hamiltonian, $H_\theta(\hat{N}_{\sf r},
1705: \hat{\theta}_{\sf r})$ in Eq.~(\ref{Htheta}) becomes simply,
1706: $H_\theta(\hat{N}_{\sf r}- {1 \over \pi} B, \hat{\theta}_{\sf r})$. As
1707: expected, the vortex density will become non-zero in the presence of the
1708: magnetic field. Since the vortex number operator in $H_\theta$ has
1709: integer eigenvalues, it is not generally possible to shift away the
1710: applied B-field. But in the Roton Liquid phase (at zero temperature)
1711: where the cosine term can be expanded to quadratic order (as in
1712: $H_{rot}$), the vortex number operator has a continuous spectra, and one
1713: can formally eliminate the B-field by shifting, $\hat{N}_{\sf r}
1714: \rightarrow \hat{N}_{\sf r} + {1 \over \pi} B$, for all ${\sf r}$.
1715: Since the ground state energy of the RL phase is thus independent of the
1716: applied B-field, both the magnetization and the magnetic susceptibility,
1717: $\chi = \partial M/\partial B$ vanish. Unlike in a superconductor,
1718: where $\chi = -{1 \over 4\pi}$, there is {\it no} Meissner effect in the
1719: Roton Liquid (strictly speaking, there is never a Meissner effect in a
1720: single two dimensional layer, but one can consider an infinite stack of
1721: electrically decoupled but magnetically coupled layers, which would
1722: exhibit a Meissner effect when the layers are true 2d superconductors,
1723: but not when they are RLs). Physically, since the RL phase supports
1724: gapless roton excitations, the state cannot screen out an applied
1725: magnetic field.
1726:
1727: \subsection{Off-Diagonal Quasi-Long-Range Order}
1728: \label{sec:quasi-diagonal-long}
1729:
1730:
1731:
1732: We next consider the Cooper pair propagator in the Roton Liquid,
1733: \begin{equation}
1734: G^{cp}({\bf r}_1 - {\bf r}_2,\tau_1-\tau_2) = \langle
1735: B_{{\bf r}_1}(\tau_1)
1736: B^\dagger_{{\bf r}_2}(\tau_2) \rangle ,
1737: \end{equation}
1738: with the Cooper pair destruction operator $B_{\bf r}$ given in
1739: Eq.~(\ref{pairop}) as an infinite product of exponentials, $e^{2\pi i
1740: e_j}$, running along the string. The propagator for the $d$-wave pair
1741: field, $B_{{\bf r},{\bf r}+\hat{\bf x}_j}$,
1742: \begin{eqnarray}
1743: \label{eq:Gcpij}
1744: G^{cp}_{ij}({\bf r},\tau) & = &
1745: \left\langle B^{\vphantom\dagger}_{{\bf 0},\hat{\bf
1746: x}_i}(0)B^\dagger_{{\bf r},{\bf
1747: r}+\hat{\bf x}_j}(\tau)\right\rangle,
1748: \end{eqnarray}
1749: behaves similarly, and will
1750: be discussed at the end of this sub-section.
1751:
1752: \subsubsection{Equal time correlator}
1753:
1754: We consider at first the equal time correlator, with $\tau_1=\tau_2$.
1755: The path independence of the string rests on the condition,
1756: $(\vec{\nabla}\cdot\vec{e})({\sf r}) = N_{\sf r}$, with {\it integer}
1757: vortex number, $N_{\sf r}$. Unfortunately, within the tractable
1758: harmonic approximation valid for most quantities in the roton liquid
1759: phase (with cosine terms in the roton hopping expanded to quadratic
1760: order), the condition of integer vortex number is {\it not} satisfied,
1761: and the results for $G^{cp}({\bf r},0)$ depend upon the choice of
1762: string. We believe that the correct behavior can be extracted by taking
1763: the string running along the straightest and ``shortest'' (using the
1764: ``city block'' metric $|x_1-x_2|+|y_1-y_2|$) path between the two
1765: points, ${\bf r}_1$ and ${\bf r}_2$. As we shall see, the Cooper pair
1766: propagator calculated in this way has an anisotropic spatial power-law
1767: decay. Preliminary calculations suggest that, once perturbative
1768: corrections to the harmonic approximation (using the formalism
1769: established in Sec.~\ref{sec:superc-inst}) are taken into account (even
1770: if they are irrelevant in the renormalization group sense), simple
1771: variations in the string do not modify the power-law decay of
1772: $G^{cp}({\bf r},0)$, but only change the (non-universal)
1773: prefactor.\cite{BFunpub}\
1774:
1775: We take ${\bf r}_1 - {\bf r}_2= X\hat{x}+Y\hat{y}$, and ${\bf r}_2
1776: = {\bf w}$, with integer $X,Y\geq 0$. Then, upon expressing the
1777: correlator $G^{cp}(X,Y,0)$ as an imaginary time path integral, one
1778: obtains an extra term in the Euclidean action, $S=\int_\tau
1779: \sum_{\sf r} {\cal
1780: L}$ in Eq.~(\ref{Sfullb}), with
1781: \begin{equation}
1782: {\cal L} \rightarrow {\cal L} + i e_j({\sf r}
1783: ,\tau) {\cal J}_j({\sf r},\tau) .
1784: \end{equation}
1785: The c-number ``source'' field is given as,
1786: \begin{equation}
1787: {\cal J}_j({\sf r},\tau) = 2 \pi \delta(\tau)\big[
1788: \delta_{j2}\sum_{x'=0}^{X-1} \delta_{x,x'+1}\delta_{y,0}-\delta_{j1}
1789: \sum_{y'=0}^{Y-1} \delta_{x,X}\delta_{y,y'} \big] .
1790: \end{equation}
1791: The Fourier transform is simply,
1792: \begin{equation}
1793: {\cal J}_j({\bf k},\omega_n) = 2\pi \big[\delta_{j2}
1794: \frac{1-e^{-ik_x X}}{i{\cal K}_x} - \delta_{j1} \frac{e^{-ik_x
1795: X}(1-e^{-ik_y Y})}{i{\cal K}_y}\big].
1796: \end{equation}
1797: Integrating out the electric field $e_j$ in the gauge
1798: $\vec{\nabla}\cdot \vec{a}=0$ and decomposing the source
1799: fields into transverse and longitudinal parts, ${\cal J}_t = i
1800: \epsilon_{ij} {\cal K}_i {\cal J}_j/{\cal K}$, ${\cal J}_l=i {\cal
1801: K}_i^* {\cal J}_i/{\cal K}$, one obtains the result
1802: \begin{eqnarray}
1803: \label{eq:stringups}
1804: G^{cp}(X,Y,0) & = & \Big\langle \exp\big\{ -\int_{{\bf
1805: k},\omega_n}\big[
1806: \frac{i\omega_n}{\sqrt{u_v}}( {\cal J}_t \Upsilon_1-{\cal J}_l
1807: \Upsilon_2 ) \nonumber \\ && + \frac{1}{2u_v} ({\cal J}_t^2+{\cal
1808: J}_l^2) \big] \big\}
1809: \Big\rangle_{\Upsilon} ,
1810: \end{eqnarray}
1811: where the Gaussian average over $\Upsilon$ is to be taken with respect
1812: to $S_{RL}$ in Eq.~(\ref{eq:upsilons}). Performing this Gaussian
1813: integral, one obtains
1814: \begin{eqnarray}
1815: \label{eq:Gcpthree}
1816: G^{cp}(X,Y,0) & = & \exp[ -\Gamma_l -\Gamma_t
1817: -\Gamma_{lt}],
1818: \end{eqnarray}
1819: where
1820: \begin{eqnarray}
1821: \label{eq:Gammacps}
1822: \Gamma_l & = & \int_{{\bf k}\omega_n} \frac{|{\cal J}_l|^2
1823: (1-\omega_n^2 G_{22})}{2u_v}, \\
1824: \Gamma_t & = & \int_{{\bf k}\omega_n} \frac{|{\cal J}_t|^2
1825: (1-\omega_n^2 G_{11})}{2u_v}, \\
1826: \Gamma_{lt} & = & -\int_{{\bf k}\omega_n} \frac{{\cal J}_l{\cal J}_t^*
1827: \omega_n^2 G_{12}}{u_v},
1828: \end{eqnarray}
1829: with $G_{ij}$ given in Eq.~(\ref{eq:gf}).
1830:
1831: Investigation of $\Gamma_t$ and $\Gamma_{lt}$ shows that the
1832: corresponding integrands are non-singular at small $k_x$ or $k_y$, and
1833: hence go to finite limits for large $|X|$ and/or large $|Y|$.
1834: They will thus affect only the amplitude of the Cooper pair propagator at
1835: large distances, and we henceforth neglect them. Singular behavior at
1836: long distances {\sl does} arise from $\Gamma_l$, in line with the
1837: intuition that it is vortex fluctuations which disrupt the
1838: superconducting phase, since ${\cal J}_l$ couples to the longitudinal
1839: electric field, which through $\vec\nabla\cdot\vec{e} = N$ describes
1840: the vorticity. To evaluate $\Gamma_l$, we first perform the frequency
1841: integration to obtain
1842: \begin{eqnarray}
1843: \label{eq:gammal1}
1844: \Gamma_l & = & \int_{\bf k} |{\cal J}_l|^2 \frac{\tilde{v}_0 v_1
1845: |{\cal K}_x {\cal K}_y| + v_1^2|{\cal K}_x {\cal K}_y|^2/{\cal
1846: K}^2}{4u_v (\omega_{pl}+\omega_{rot})} .
1847: \end{eqnarray}
1848: Next, we explicitly express the square of the longitudinal string,
1849: \begin{eqnarray}
1850: \label{eq:stringl}
1851: |{\cal J}_l|^2 & = & \frac{(2\pi)^2}{{\cal K}^2} \Big[
1852: \left|\frac{{\cal K}_x}{{\cal K}_y}\right|^2 |1-e^{-ik_y Y}|^2 +
1853: \left|\frac{{\cal K}_y}{{\cal K}_x}\right|^2 |1-e^{-ik_x X}|^2
1854: \nonumber \\
1855: & & + 2{\rm Re} \{ (1-e^{-ik_x X})(1-e^{-ik_y Y}) \} \Big].
1856: \end{eqnarray}
1857: The first two terms in Eq.~(\ref{eq:stringl}) are singular for small
1858: $k_y$, $k_x$ respectively for very large $Y$,$X$, leading to
1859: a logarithmic dependence when inserted in Eq.~(\ref{eq:gammal1}). The
1860: final term in Eq.~(\ref{eq:stringl}), by contrast, is singular only
1861: for {\sl both} $k_x$,$k_y$ small, and this singularity, inserted into
1862: Eq.~(\ref{eq:gammal1}), is weak and integrable. Extracting the
1863: logarithmic parts, one finds
1864: \begin{equation}
1865: \label{eq:gammal2}
1866: \Gamma_l \sim \Delta_c (\ln |X| + \ln|Y|),
1867: \end{equation}
1868: for $|X|,|Y| \gg 1$, with
1869: \begin{equation}
1870: \Delta_c = 2\pi {\tilde{v}_0 \over v_+ } \sqrt{ {\kappa}_r \over
1871: u_v } .
1872: \label{Cooperexp}
1873: \end{equation}
1874: Hence we have
1875: \begin{equation}
1876: \label{eq:ODQLRO}
1877: G^{cp}(X,Y,0) \sim \frac{{\rm const}}{|X|^{\Delta_c}|Y|^{\Delta_c}}.
1878: \end{equation}
1879: This establishes
1880: that the Roton liquid phase has off-diagonal quasi-long-ranged
1881: order (ODQLRO) at zero temperature.
1882:
1883: \subsubsection{Unequal time correlator}
1884:
1885: We now consider the Cooper pair propagator at unequal times.
1886: Unfortunately, it is difficult to produce a simple and general
1887: calculation for arbitrary spatial and time separations. In
1888: particular, clearly, by square symmetry, we expect $G^{cp}(X,Y,\tau) =
1889: G^{cp}(Y,X,\tau)$. Any choice of strings, however, necessarily
1890: creates an asymmetry between the two spatial directions. As we have
1891: been unable to resolve this dilemma, we instead focus on the simple
1892: case in which the pair is created and annihilated on a single row of
1893: the lattice, i.e. $G^{cp}(X,0,\tau)$. We will see that this
1894: correlator decays as a power law both in space and time.
1895:
1896: To proceed, we take ${\bf r}_1 - {\bf r}_2= X\hat{x}$, ${\bf r}_2 =
1897: {\bf w}$, $\tau_1=\tau$, $\tau_2=0$, with $X,Y\geq 0$. With this
1898: choice, the string in Fourier space becomes
1899: \begin{equation}
1900: \label{eq:stringtau}
1901: {\cal J}_j({\bf k},\omega_n) = 2\pi \delta_{j2}
1902: \frac{1-e^{-ik_x X+i\omega_n\tau}}{i{\cal K}_x}.
1903: \end{equation}
1904: Repeating the same manipulations as above, one again obtains (with
1905: negligible contributions from the transverse part of the string)
1906: \begin{equation}
1907: \label{eq:cptau}
1908: G^{cp}(X,0,\tau) \sim \exp[ -\tilde\Gamma_l ],
1909: \end{equation}
1910: with
1911: \begin{equation}
1912: \label{eq:gammalt}
1913: \tilde\Gamma_l \sim \frac{\Delta_c}{2} \ln ( X^2 + v_{rot}^2 \tau^2 ),
1914: \end{equation}
1915: where for simplicity we have taken $v_{rot}$ independent of $k_y$ (for
1916: a non-trivial $v(k_y)$, the logarithm is simply averaged uniformly
1917: over the $k_y$ axis). This gives
1918: \begin{equation}
1919: \label{eq:Gctau}
1920: G_c(X,0,\tau) \sim \frac{\rm const.}{(X^2+v_{rot}^2 \tau^2)^{\Delta_c/2}}.
1921: \end{equation}
1922: Note that this power-law form implies a power-law local tunneling
1923: density of states for Cooper pairs, $\rho_{cp}(\epsilon) \sim
1924: \epsilon^{\Delta_c-1}$.
1925:
1926: We conjecture that the full
1927: correlator satisfies a simple scaling form with ``z=1'' scaling:
1928: \begin{equation}
1929: \label{eq:scaling}
1930: G^{cp}(X,Y,\tau) \sim \frac{{\rm
1931: const}}{|X|^{\Delta_c}|Y|^{\Delta_c}} {\cal
1932: G}(\frac{X}{v_{rot}\tau},\frac{Y}{v_{rot}\tau}).
1933: \end{equation}
1934: Combining our two calculations above implies ${\cal G}(\chi,0) =
1935: (\chi^2/(\chi^2+1))^{\Delta_c/2}$.
1936:
1937: \subsubsection{ODQLRO of the $d$-wave pair field}
1938: \label{sec:odqlro-d-wave}
1939:
1940: We now briefly discuss the analogous ODQLRO of the $d$-wave pair field
1941: described by $G^{cp}_{ij}({\bf r},\tau)$ in Eq.~(\ref{eq:Gcpij}). This
1942: quantity is plagued by the same string ambiguities as the local Cooper
1943: pair propagator, but to a larger degree, since it involves {\sl two}
1944: separate strings emanating from the two sites shared by the initial bond
1945: and ending at the two sites shared by the final bond. Although we are
1946: confident $G^{cp}_{ij}({\bf r},\tau)$ has a power-law form consistent
1947: with ODQLRO, we are unable to determine the precise nature of these
1948: correlations with reliability. For instance, consider the equal time
1949: pair field correlator for two bonds along the $x$-axis,
1950: $G^{cp}_{11}({\bf r},0)$. For two bonds in the same row, ${\bf
1951: r}=X\hat{\bf x}_1$, the strings can be chosen to line all in the same
1952: row, and since the logarithmic divergence controlling the ODQLRO arises
1953: from small $k_x$ in this case, the power law exponent is unchanged, i.e.
1954: $G^{cp}_{11}(X,0,0) \sim {\rm Const}/|X|^{\Delta_c}$, with $\Delta_c$
1955: given above. We believe that, since this choice of string is by far the
1956: most natural, this is probably the correct result. If, instead, we
1957: choose to separate the two pair fields along a single column, ${\bf
1958: r}=Y\hat{\bf x}_2$, then the two strings involved cannot be taken
1959: entirely atop one another. Different choices for the strings then give
1960: different results. For instance, making the symmetric choice of two
1961: parallel strings (each of strength $\pi$ rather than $2\pi$) gives a
1962: decay exponent reduced from $\Delta_c$ to $\Delta_c/2$ in the $Y$
1963: direction, while choosing the strings to overlap everywhere except the
1964: two ends reproduces the previous exponent $\Delta_c$ without any
1965: reduction. Since we are unable to reliably resolve this ambiguity, we
1966: are unable to determine the exact form of the $d$-wave pair field
1967: correlator. Instead, we will take the pragmatic approach of
1968: approximating the correlations by those of the local pair field,
1969: $G^{cp}_{ij}({\bf r},\tau) \approx G^{cp}({\bf r},\tau)$. It should be
1970: understood that the decay exponent $\Delta_c$ may need to be
1971: renormalized and/or the correlator corrected slightly to obtain detailed
1972: results for a specific model.
1973:
1974: \subsection{Conductivity in the harmonic theory}
1975: \label{sec:cond-harm-theory}
1976:
1977: Given the above result of ODQLRO, it is natural to expect a very
1978: large and perhaps infinite conductivity in the Roton liquid.
1979: Indeed, neglecting the effects of vortex hopping, ${\cal L}_v$ in
1980: Eq.~(\ref{vorthop}), one can readily see (e.g. from
1981: Eq.~(\ref{Htheta})) that the total vortex number on each row and
1982: each column of the 2d lattice is separately conserved. Thus it is
1983: impossible to set up a vortex flow, and hence by the Josephson
1984: relation to generate an electric field. Using the quadratic roton
1985: liquid Lagrangian ${\cal L}_{RL}$, this expectation can be
1986: directly confirmed. In particular, we can integrate out {\sl all}
1987: dynamical fields ($\theta,a$) from the Lagrangian leaving only the
1988: external gauge field $A^\mu$, and thereby extract the polarization
1989: tensor and conductivity. This is most conveniently carried out in
1990: the gauge $A_0=0$, and assuming no magnetic field $\partial_x
1991: A_y-\partial_y A_x=0$. In this case, one may write
1992: \begin{equation}
1993: S_A = \int_{{\bf k},\omega_n} \frac{-i\omega_n {\cal K}_j({\bf
1994: k})}{\pi{\cal K}({\bf k})}e^{-i{\bf k}\cdot{\bf w}} a({\bf k},\omega_n)
1995: A_j(-{\bf k},-\omega_n).
1996: \end{equation}
1997: Integrating out $a$ and $\theta$ using the Green's function $G_{11}$
1998: in Eq.~(\ref{eq:gf}) , one obtains, in the limit $|{\bf k}| \rightarrow
1999: 0$, the effective action
2000: \begin{equation}
2001: S^{0}_A = \frac{1}{2} \int_{{\bf k},\omega_n} \Pi^0_{ij}({\bf
2002: k},\omega_n) A_i({\bf k},\omega_n) A_j(-{\bf k},-\omega_n),
2003: \end{equation}
2004: with
2005: \begin{equation}
2006: \Pi^0_{ij}({\bf k},\omega_n) \sim \frac{u_v}{\pi^2}\frac{k_i k_j}{k^2} ,
2007: \end{equation}
2008: as $|{\bf k}| \rightarrow 0$. The dependence of $\Pi^0_{ij}$ on the
2009: orientation of ${\bf k}$ is due to the fact that we have assumed
2010: $B=0$, which according to Faraday's law requires ${\bf \nabla}\times
2011: {\bf E} = -\partial_t {\bf B}=0$. Hence we must choose ${\bf k}$
2012: parallel to the electric field ${\bf A}={\bf E}/(-i\omega)$. Thus we
2013: extract an {\sl isotropic} conductivity tensor $\sigma^0_{ij} =
2014: \delta_{ij} \sigma^0$, with
2015: \begin{equation}
2016: \sigma^0(\omega) = \frac{u_v}{\pi^2} \frac{1}{-i\omega},
2017: \end{equation}
2018: characteristic of a system with no dissipation.
2019:
2020: The above conclusion for the quadratic RL Lagrangian is, however,
2021: modified by the vortex hopping terms. As we detail in the next section,
2022: the effects of a small vortex hopping term depend sensitively on the
2023: parameters that enter in the harmonic theory of the roton liquid - in
2024: particular the dimensionless ratio $u_v/\kappa_r$. There are two
2025: regimes. When this ratio is larger than a critical value, vortex
2026: hopping is ``relevant'' and grows at low energies destabilizing the
2027: roton liquid phase. On the other hand, for small enough $u_v/\kappa_r$
2028: the vortex hopping strength scales to zero and the roton liquid phase is
2029: stable. In this latter case, the effects of vortex hopping on physical
2030: quantities can be treated perturbatively. In particular, we find that
2031: the conductivity in the roton liquid diverges as a power law in the low
2032: frequency and low temperature limit.
2033:
2034:
2035: \section{Instabilities of the Roton Liquid}
2036: \label{sec:inst-roton-liqu}
2037:
2038: We first consider the instabilities of the roton liquid due to the
2039: presence of a vortex hopping term, and examine the effects of such
2040: processes on the electrical transport. In the subsequent subsection
2041: we consider the legitimacy of the harmonic expansion required to
2042: obtain the quadratic RL Lagrangian. This is achieved by performing a
2043: ''plaquette duality'' transformation\cite{EBL}, where it is possible to address
2044: this issue perturbatively. Again we find two regimes depending on the
2045: parameters in the quadratic roton liquid Lagrangian. A stable regime
2046: for small $u_v/\kappa_r$ wherein the harmonic roton liquid description
2047: is valid, and an instability towards a superconducting phase when this
2048: ratio is large.
2049:
2050:
2051: \subsection{Vortex hopping}
2052: \label{sec:vortex-hopping}
2053:
2054: The analysis of the stability of the RL to vortex hopping is
2055: mathematically nearly identical to the stability analysis of the
2056: Exciton Bose Liquid (EBL) of Ref.~[\onlinecite{EBL}] with respect to
2057: boson hopping. Taking over those methods, we note that the vortex
2058: hopping operator exhibits {\sl one-dimensional power-law
2059: correlations},
2060: \begin{equation}
2061: \label{eq:hoppower}
2062: \langle e^{i\partial_y\theta_{\sf r}(0)} e^{-i\partial_y\theta_{{\sf
2063: r}+{\bf r}}(\tau)} \rangle_{RL} = \delta_{y,0} {\cal R}(x,\tau) ,
2064: \end{equation}
2065: with a power law form at large space-time separations,
2066: \begin{equation}
2067: \label{rotgreen}
2068: {\cal R}(x,\tau) =
2069: \frac{c}{(x^2+v_{rot}^2 \tau^2)^{\Delta_v}},
2070: \end{equation}
2071: where ${\bf r}=x{\bf\hat{x}}+y{\bf\hat{y}}$
2072: and $c$ is a dimensionless constant.
2073: Here, ${\cal R}(x,\tau)$ is essentially the single roton
2074: Greens function,
2075: describing the space-time propagation of a roton with a dipole
2076: oriented
2077: along the $\hat{y}$ axis.
2078: When calculated using the RL Lagrangian, one finds
2079: \begin{equation}
2080: \label{DeltavRL}
2081: \Delta_v = \frac{1}{4\pi} \sqrt{\frac{u_v}{\kappa_r}} \frac{v_+}{\tilde{v}_0}.
2082: \end{equation}
2083: Simple calculations show that this power-law behavior is not modified
2084: by including the fluctuating $a_y$ field in the vortex hopping
2085: operator, $\hat{T}_j({\sf r},\tau) = e^{i(\partial_y\theta_{\sf r} +
2086: a_y({\sf r}))}$, i.e.
2087: \begin{equation}
2088: \label{eq:hoppower2}
2089: \langle \hat{T}_y({\sf r},0) \hat{T}^*_y({\sf r}+{\bf r},\tau)
2090: \rangle_{RL} \sim
2091: \frac{\delta_{y,0}}{(x^2+v_{rot}^2 \tau^2)^{\Delta_v}},
2092: \end{equation}
2093: with only a change in the prefactor. Notice that the exponent
2094: $\Delta_v$ characterizing the power law decay of the roton
2095: propagator is inversely proportional to the analogous exponent
2096: $\Delta_c$ in Eq.~(\ref{Cooperexp}) which gives the power law
2097: decay of the Cooper pair propagator. Indeed, for the Roton liquid
2098: Lagrangian studied here we find the simple identity, $\Delta_v
2099: \Delta_c = 1/2$. But with inclusion of other terms in the
2100: original Hamiltonian such as the spinons or further neighbor roton
2101: hopping terms, this equality will be modified.
2102:
2103: The arguments of Ref.~[\onlinecite{EBL}] imply that the vortex hopping
2104: term is then relevant for $\Delta_v <2$. In this regime, the vortex
2105: hopping strength grows large when scaling to low energies, and one
2106: expects the vortices to condense at zero temperature. In this case it is legitimate to
2107: expand the cosine term in Eq.~(\ref{vorthop})
2108: and one generates a ``dual Meissner
2109: effect'' where the gauge fields that are minimally coupled to the
2110: vortices become massive. In the presence of spinons this leads to a
2111: mass term of the form, ${\cal L}_v \sim {t_v \over 2} (a_j -
2112: \alpha_j)^2$, which confines one unit of electrical charge to each
2113: spinon presumably driving one into a Fermi liquid phase. If we ignore
2114: fluctuations in the spinon density, or drop the spinons entirely
2115: retaining a theory of Cooper pairs, the resulting phase is a charge
2116: ordered bosonic insulator. The nature of the charge ordering will in
2117: this case depend sensitively on the commensurability of the Cooper
2118: pair density ($\rho_0/2$) with the underlying square lattice. In the
2119: simplest commensurate case with one Cooper pair per site ($\rho_0 =
2120: 2$), a featureless Mott insulating state obtains.
2121:
2122: \subsubsection{Electrical Resistance in the
2123: roton liquid}
2124: \label{sec:electr-resist-roton}
2125:
2126: When $\Delta_v>2$, on the other hand, vortex hopping is ''irrelevant'',
2127: and the effects of the hopping on physical quantities can be treated
2128: perturbatively. Despite its irrelevance, we expect the vortex hopping
2129: to strongly modify the Gaussian result for the conductivity by
2130: introducing dissipation. To understand how this occurs, it is
2131: instructive to first consider the simple limit alluded to earlier in
2132: which $v_0 \rightarrow \infty$. In this limit, the longitudinal density
2133: fluctuations described by $\tilde{a}$ at non-zero wavevector are
2134: suppressed, and the roton mode is purely captured by the $\theta$ field.
2135: The zero wavevector (but non-zero frequency) piece of $a$, however,
2136: remains non-zero in this limit and can be used to calculate the
2137: conductivity in an RPA fashion. In particular, we take into account the
2138: roton fluctuations and their associated dissipation induced by vortex
2139: hopping by calculating the effective action for $\tilde{a},A$ upon
2140: integrating out $\theta$ to second order (the lowest non-trivial
2141: contribution) in $t_v$. Starting then with the Lagrangian, ${\cal
2142: L}_{RL} + {\cal L}_A + {\cal L}_v$, expanding to second order in the
2143: vortex hopping action $S_v$, and integrating over the $\theta$ field and
2144: the gauge field $a({\bf k} \ne 0)$ with $v_0 \rightarrow \infty$ gives,
2145: \begin{eqnarray}
2146: S_{a,A}^{\rm eff} & = & \sum_{{\bf r}={\sf r}+{\bf w}}\int_\tau
2147: [\frac{1}{2u_v}
2148: (\partial_0\tilde{a}_j)^2 -\frac{i}{\pi}\epsilon_{ij} \tilde{a}_i({\sf r})
2149: \partial_0 A_j({\bf r}-{\bf\hat{x}_j})] \nonumber \\
2150: & & + S^{(2)}_{a,A},
2151: \end{eqnarray}
2152: where
2153: \begin{equation}
2154: S^{(2)}_{a,A} = -\frac{t_v^2}{2} \sum_{{\sf r,r'}}\int_{\tau,\tau'}
2155: \left\langle \cos(\partial_i \theta - \tilde{a}_i)_{{\sf r}\tau}
2156: \cos(\partial_j \theta-\tilde{a}_j)_{{\sf r'}\tau'} \right\rangle_\theta,
2157: \end{equation}
2158: where $\langle\cdot\rangle_\theta$ indicates the average with respect
2159: to the Gaussian action for $\theta$. From Eq.~(\ref{eq:hoppower}),
2160: one can carry out this average to obtain,
2161: \begin{eqnarray}
2162: S^{(2)}_{a,A} & \sim & -\frac{t_v^2}{4} \bigg\{
2163: \sum_{{\sf r},x}\int_{\tau,\tau'}
2164: {\cal R}(x,\tau - \tau^\prime) \nonumber \\
2165: & \times & \cos[\tilde{a}_y({\sf r},\tau)-\tilde{a}_y({\sf
2166: r}+x{\bf\hat{x}},\tau')] + (x \leftrightarrow y)\bigg\} ,
2167: \end{eqnarray}
2168: with the roton propagator ${\cal R}(x,\tau)$ given as in Eq.~(\ref{rotgreen}),
2169: except with $\Delta_v \rightarrow \sqrt{u_v/\kappa_r}/(4\pi)$ in this limit.
2170: Following the usual RPA strategy, we expand $S^{(2)}$ to quadratic
2171: order in $\tilde{a}$ to obtain,
2172: \begin{equation}
2173: \label{RPArot}
2174: S_{a,A}^{(2)} = \frac{t_v^2}{4} \int_{{\bf
2175: k},\omega_n} \tilde{{\cal R}}(\omega_n)
2176: |\tilde{a}_j(\omega_n)|^2,
2177: \end{equation}
2178: with the definition, $\tilde{{\cal R}}(\omega_n) = {\cal R}(0)
2179: - {\cal R}(\omega_n)$ and,
2180: \begin{equation}
2181: \label{Protgreen}
2182: {\cal R}(\omega_n) \equiv {\cal R}(k_x=0,\omega_n) .
2183: \end{equation}
2184: The ${\bf k} = 0$ limit
2185: is valid when
2186: $v_0\rightarrow\infty$.
2187: At low frequencies one has,
2188: \begin{equation}
2189: {\cal R}(\omega_n) = - {\cal C}_\gamma \frac{| \omega_n /v_{rot} |^{1 + \gamma} }{v_{rot} \sin \frac{\pi}{2}
2190: (1 + \gamma) } + ... ,
2191: \label{rotpol}
2192: \end{equation}
2193: with ${\cal C}_\gamma > 0$ a dimensionless constant. Here we have retained
2194: explicitly the leading singular frequency dependence and
2195: dropped analytic terms consisting of even powers of $\omega_n$.
2196: The exponent $\gamma$ is defined as,
2197: \begin{equation}
2198: \gamma = 2 \Delta_v - 3 .
2199: \end{equation}
2200: and in the stable regime of the RL phase, $\gamma >1$.
2201:
2202: Finally, upon
2203: integrating out $\tilde{a}_j$, one obtains a renormalized
2204: electromagnetic response tensor
2205: \begin{equation}
2206: \Pi_{ij} = \frac{\omega_n^2}{u_v^{-1} (\pi \omega_n)^2 +
2207: \frac{\pi^2 t_v^2}{2} \tilde{{\cal R}}(\omega_n) } \frac{k_i
2208: k_j}{k^2}.
2209: \label{eq:piRPA}
2210: \end{equation}
2211:
2212: It is now straightforward to extract the conductivity by analytic
2213: continuation,
2214: \begin{equation}
2215: \sigma(\omega) = \left.\frac{\Pi_{xx}(k_x\rightarrow
2216: 0,k_y=0,\omega_n)}{\omega_n} \right|_{i\omega_n\rightarrow
2217: \omega+i\delta}.
2218: \end{equation}
2219: One obtains an appealing Drude form:
2220: \begin{equation}
2221: \label{eq:drude}
2222: \sigma(\omega) = \frac{1}{-i\omega ( \pi^2/u_v) +
2223: i \frac{\pi^2 t_v^2}{2} \tilde{{\cal R}}_{ret}(\omega) /\omega } ,
2224: \end{equation}
2225: with the retarded propagator obtained from analytic continuation:
2226: \begin{equation}
2227: {\cal R}_{ret}(\omega)= {\cal R}(\omega_n) |_{i\omega_n\rightarrow
2228: \omega+i\delta}.
2229: \end{equation}
2230: The non-analytic frequency dependence of ${\cal R}_{ret} (\omega)$
2231: contributes to the dissipative (real) part of the resistance
2232: (per square), $R(\omega) \equiv {\rm Re} \sigma^{-1}(\omega)$,
2233: which is quadratic in the vortex hopping amplitude, $t_v^2$:
2234: \begin{equation}
2235: \label{eq:resist}
2236: R(\omega) = \frac{\pi^2 t_v^2} {2\omega} Im {\cal R}_{ret}(\omega) =
2237: {\cal C}_\gamma \frac{ \pi^2 t_v^2}{2v_{rot}^2} \left[ \frac{|\omega|}{v_{rot}}\right]^\gamma . \end{equation}
2238: Note that at the point for which vortex hopping is just marginal,
2239: $\gamma =1$, the resistance becomes linear in frequency ${\rm Re}
2240: \sigma(\omega) \sim 1/\omega$.
2241:
2242: We can readily extend this result to finite temperatures,
2243: by using the finite temperature roton propagator:
2244: \begin{equation}
2245: {\cal R}(\tau) = c_\gamma \left[
2246: \frac{ \pi/v_{rot}\beta } { sin(\pi \tau/\beta) } \right]^{\gamma + 2} .
2247: \end{equation}
2248: The retarded roton propagator follows upon analytic continuation,
2249: and can be most readily extracted by using the identity:
2250: \begin{equation}
2251: \label{eq:identity}
2252: Im {\cal R}_{ret}(\omega) = \sinh(\beta \omega/2) \int^\infty_{-\infty}
2253: dt e^{i \omega t} {\cal R}(\tau \rightarrow \frac{\beta}{2} + it ) .
2254: \end{equation}
2255: Upon combining Eq.~(\ref{eq:resist}-\ref{eq:identity}) we thereby obtain
2256: a general expression for the finite temperature and frequency
2257: (dissipative) resistance in the roton liquid phase:
2258: \begin{equation}
2259: R(\omega,T) = c_\gamma \frac{(\pi t_v)^2}{v_{rot}^2}
2260: \left[ \frac{\pi T}{v_{rot}}\right]^\gamma
2261: \tilde{R}_\gamma(\omega/\pi T) ,\label{eq:RLresist}
2262: \end{equation}
2263: with a universal crossover scaling function,
2264: \begin{equation}
2265: \tilde{R}_\gamma(X) = \frac{2^\gamma}{\Gamma(2+\gamma)} |\Gamma(1 + \frac{\gamma
2266: +iX}{2})|^2 \frac{\sinh(\pi X/2)}{X} ,
2267: \end{equation}
2268: interpolating between the d.c. resistance at finite temperature
2269: and
2270: the $T=0$ a.c. behavior.
2271: Since $\tilde{R}_\gamma(X \rightarrow 0)$ is finite, the d.c. resistance varies as a power law in temperature: $R(T) \sim T^\gamma$.
2272: At the boundary of the
2273: RL phase with $\gamma =1$, a linear temperature dependence is predicted.
2274: At large argument,
2275: \begin{equation}
2276: \tilde{R}_\gamma(X \rightarrow \infty) = \frac{\pi}{ \Gamma(2 + \gamma)}
2277: X^\gamma ,
2278: \end{equation}
2279: so that the resistance crosses over smoothly to
2280: the zero temperature form, $R(\omega, T=0) \sim |\omega|^\gamma$.
2281:
2282: This RPA treatment has the appeal that it produces the natural physical
2283: result that the effect of the weak (irrelevant) vortex hopping is to
2284: generate a small {\sl resistivity} $\sim t_v^2$. Formally, it is
2285: correct for $v_0=\infty$ because the RPA reproduces the exact
2286: perturbative result for the electromagnetic response tensor to
2287: $O(t_v^2)$ in this case. Unfortunately, when the spatial fluctuations
2288: of $\tilde{a}_j$ are not negligible, i.e. for $v_0 < \infty$, even the
2289: $O(t_v^2)$ term is not obtained correctly. More generally, the
2290: fluctuations of $\tilde{a}_j$ and $\theta$ must be treated on the same
2291: footing. Therefore in the general case we instead integrate out {\sl
2292: both} fields and obtain more directly the correction to $\Pi_{ij}$ to
2293: $O(t_v^2)$. The calculations are described in
2294: appendix~\ref{sec:appendix} . This does not yield the appealing
2295: ``Drude'' form in Eq.~(\ref{eq:drude}) but instead the Taylor expansion
2296: of Eq.~(\ref{eq:piRPA}) to $O(t_v^2)$,
2297: \begin{equation}
2298: \label{eq:polar3}
2299: \Pi_{ij}^{(2)} \sim - \frac{t_v^2 u_v^2}{2v_{rot}^{2\Delta_v -1}} |\omega_n|^{2\Delta_v-4} \frac{k_i
2300: k_j}{k^2} ,
2301: \end{equation}
2302: except with the scaling dimension $\Delta_v$, given explicitly in
2303: Eq.~(\ref{DeltavRL}), now fully renormalized by the plasmon
2304: fluctuations. Provided the vortex hopping is irrelevant, this is
2305: sufficient to recover properly the low-frequency behavior of the
2306: resistivity, Eq.~(\ref{eq:resist}) to $O(t_v^2)$. In particular,
2307: formally inverting the perturbative result for $\sigma(\omega,T)$ to
2308: $O(t_v^2)$, we infer the appropriate d.c. dissipative
2309: resistance $R(T) \sim t_v^2 T^{\gamma}$, with $\gamma =
2310: 2\Delta_v -3$.
2311:
2312:
2313: \subsection{``Charge Hopping''}
2314: \label{sec:charge-hopping}
2315:
2316: In this subsection we examine the legitimacy of the ``spin wave''
2317: expansion employed in Sec. IIIA to obtain the Gaussian Roton
2318: liquid Lagrangian. This is most readily achieved by passing to a
2319: dual representation, which exchanges vortex operators for new
2320: ``charge'' operators. This procedure is a quantum analog of the
2321: mapping from the classical 2d $XY$-model to a sine-Gordon
2322: representation\cite{2dXY}, the latter suited to examine
2323: corrections to the low temperature ``spin wave'' expansion. As we
2324: shall find, there are parameter regimes where the ``charge''
2325: hopping terms are irrelevant and the RL phase is stable. But
2326: outside of these regimes, the ``charge'' quasiparticles become
2327: mobile at low energies and condense - driving an instability into
2328: a conventional superconducting phase. Throughout this subsection
2329: we will drop the vortex hopping term, $H_v$, focusing on the
2330: parameter regimes of the Roton Liquid phase where it is irrelevant
2331: (i.e. $\Delta_v > 2$).
2332:
2333: \subsubsection{Plaquette Duality}
2334: \label{sec:plaquette-duality}
2335:
2336: To this end we now employ the ``plaquette duality''
2337: transformation, originally introduced in Ref.~[\onlinecite{EBL}]
2338: in the context of the Exciton-Bose-Liquid phase. Consider
2339: the charge sector
2340: of the theory, with Hamiltonian $H_c = H_{pl} + H_{\theta} + H_v$.
2341: This Hamiltonian
2342: is a function of the vortex phase field and number operators,
2343: $\theta_{\sf r}, N_{\sf r}$ (living on the sites of the dual lattice),
2344: as well as the (transverse) gauge field
2345: $a^t_j$ and it's conjugate transverse ``electric field'', $e^t_j$.
2346: The plaquette duality exchanges $\theta_{\sf r}$ and $N_{\sf r}$
2347: for a new set of canonically conjugate fields which live on the sites of the
2348: original 2d square lattice. The two new fields, denoted $\tilde{\phi}_{\bf r}$ and
2349: $\tilde{n}_{\bf r}$, are defined
2350: via the relations,
2351: \begin{equation}
2352: \pi N_{{\bf r} + {\bf w}} \equiv \Delta_{xy} \tilde{\phi}_{\bf r} ,
2353: \end{equation}
2354: \begin{equation}
2355: \pi \tilde{n}_{\bf r} \equiv \Delta_{xy} \theta_{{\bf r} - {\bf w}} .
2356: \end{equation}
2357: Although $\tilde{\phi}_{\bf r}$ and $\tilde{n}_{\bf r}$ are conjugate fields satisfying,
2358: \begin{equation}
2359: \label{nphicomm}
2360: [\tilde{n}_{\bf r},\tilde{\phi}_{{\bf r}^\prime} ]= i \delta_{{\bf r} {\bf r}^\prime} ,
2361: \end{equation}
2362: they can not strictly be interpreted as phase and number operators
2363: since the eigenvalues of $\tilde{\phi}_{\bf r} = \pi m$ for arbitrary
2364: integer $m$, whereas $\pi \tilde{n}_{\bf r}$ is $2 \pi$ periodic.
2365: It is important that $\tilde{\phi}_{\bf r}$ and $\tilde{n}_{\bf r}$
2366: not be confused with the ``chargon'' phase and number operators introduced in Section II
2367: which were denoted $\phi_{\bf r},n_{\bf r}$ -
2368: {\it without} the tildes. As we discuss below, $e^{i\tilde{\phi}_{\bf r}}$
2369: does in fact create a charge-like excitation, but it is {\it not} the chargon.
2370:
2371: Under the change of variables, $H_\theta(\theta,N) \rightarrow
2372: H_\phi(\tilde{\phi},\tilde{n})$,
2373: \begin{equation}
2374: H_\phi = H_u - \kappa_r \sum_{\sf r} \cos[\pi \tilde{n}_{{\sf r} + {\bf w}} -
2375: {1 \over 2} (\partial_x \tilde{a}_y + \partial_y \tilde{a}_x)] ,
2376: \end{equation}
2377: with the definition,
2378: \begin{equation}
2379: H_u = {u_v \over 2 \pi^2 } \sum_{{\bf r} {\bf r}^\prime}
2380: \Delta_{xy} \tilde{\phi}_{\bf r} \Delta_{xy} \tilde{\phi}_{{\bf r}^\prime} V({\bf r} - {\bf r}^\prime) .
2381: \end{equation}
2382:
2383: In the Roton Liquid phase the cosine term in $H_\phi$ is expanded
2384: to quadratic order, and $H_\phi + H_{pl} \rightarrow H_{RL}$.
2385: To be consistent, both $\tilde{n}$ and $\tilde{\phi}$ must then be allowed to take on
2386: any real value, and it is convenient to pass to a Lagrangian written
2387: just in terms of $\tilde{\phi}$:
2388: \begin{eqnarray}
2389: L_{RL} & = & L_{pl} + H_u + {1 \over 2 } \sum_{\bf r}
2390: { (\partial_0 \tilde{\phi}_{\bf r} )^2 \over \pi^2 \kappa_r}
2391: \nonumber \\
2392: & & + \frac{i}{2\pi}
2393: \sum_{\sf r} \partial_0 \tilde{\phi}_{{\sf r} + {\bf w}}
2394: (\partial_x \tilde{a}_y +
2395: \partial_y \tilde{a}_x ) .
2396: \label{LRLdual}
2397: \end{eqnarray}
2398: In this dual form the Roton Liquid Lagrangian depends
2399: (quadratically) on the field $\tilde{\phi}_{\bf r}$, which
2400: lives on the sites of the direct lattice,
2401: and the (transverse) gauge field, $\tilde{a}_j({\sf r})$,
2402: defined on the links of the dual lattice.
2403:
2404:
2405: \subsubsection{Superconducting Instabilities}
2406: \label{sec:superc-inst}
2407:
2408: This dual formulation is ideal for studying the legitimacy
2409: of the ``spin-wave'' expansion needed to obtain the quadratic
2410: RL Lagrangian. The crucial effect of the spin-wave expansion
2411: was in softening the integer constraint on
2412: the eigenvalues of $\tilde{\phi}/\pi$,
2413: allowing $\tilde{\phi}$ to take on all real values in $L_{RL}$.
2414: It is, however, possible to mimic the effects
2415: of this constraint by adding a potential term to $L_{RL}$ of the form,
2416: \begin{equation}
2417: L_\lambda = - \lambda \sum_{\bf r} \cos(2 \tilde{\phi}_{\bf r}) .
2418: \label{cospert}
2419: \end{equation}
2420: When $\lambda \rightarrow \infty$ the integer constraint is enforced,
2421: whereas the RL phase corresponds to $\lambda =0$.
2422: Stability of the RL phase can be studied by treating $\lambda$ as a {\it small}
2423: perturbation to the quadratic RL Lagrangian.
2424: But as discussed in Ref.~[\onlinecite{EBL}], one should also consider other
2425: perturbations to $L_{RL}$ which might be even more relevant.
2426: Generally, one can add any local operator involving
2427: $\tilde{\phi}_{\bf r}$ at a set of nearby spatial points
2428: which is $2 \pi$ periodic in $2 \tilde{\phi}$, and satisfies all
2429: the discrete lattice symmetries (i.e.. translations, rotations
2430: and parity). For example, terms of the form $\cos(2 \ell \tilde{\phi})$
2431: for arbitrary integer $\ell$ are allowed, although these
2432: will generically become less relevant with increasing $\ell$.
2433: as we shall see,
2434: for our ``minimal'' model of the RL phase, the most relevant
2435: perturbation is of the form,
2436: \begin{equation}
2437: L_t = - t_c \sum_{\bf r} \sum_{j=1,2} \cos(2\partial_j \tilde{\phi}_{\bf r} ) .
2438: \label{chhoppert}
2439: \end{equation}
2440:
2441: Before studying the perturbative stability to such operators, we
2442: try to get some physical intuition for the meaning of the operator
2443: $e^{i\tilde{\phi}}$. From the commutation relations in
2444: Eq.~(\ref{nphicomm}), the operator $e^{i\tilde{\phi}_{\bf r}}$ increase
2445: $\tilde{n}_{\bf r}$ by one, and creates some sort of quasiparticle
2446: excitation on the spatial site ${\bf r}$. Since the perturbation
2447: in Eq.~(\ref{cospert}) changes the number $\tilde{n}_{\bf r}$ by $\pm 2$,
2448: the total number of these quasiparticles, $\tilde{n}_{tot} = \sum_{\bf r}
2449: \tilde{n}_{\bf r}$ is {\it not} conserved, but the ``complex charge'', $Q_c
2450: = e^{i\pi \tilde{n}_{tot}}$ is conserved. The perturbation in
2451: Eq.~(\ref{chhoppert}) can then be interpreted as a ``charge
2452: hopping'' process. To get some feel for the nature of the
2453: quasiparticle, it is instructive to introduce an external magnetic
2454: field, $B = \epsilon_{ij}
2455: \partial_i A_j$, which enters into $H_u$ above via the
2456: substitution, $\Delta_{xy} \phi_{\bf r} \rightarrow \Delta_{xy}
2457: \phi_{\bf r} - B_{{\bf r} + {\bf w}}$. A spatially
2458: uniform field, $B$, can readily be removed from the Gaussian RL
2459: Lagrangian by letting
2460: \begin{equation}
2461: \tilde{\phi}_{\bf r} \rightarrow \tilde{\phi}_{\bf r} + B x y .
2462: \end{equation}
2463: But then $B$ appears in the cosine perturbations, in ``almost'' a
2464: minimal coupling form. For example, one has the combination
2465: $\partial_x \tilde{\phi} - 2A_x$ where we have chosen the gauge $A_x = -B
2466: y/2$ and $A_y =Bx/2$, suggesting that $e^{i\tilde{\phi}}$ carries the
2467: Cooper pair charge. But the $y-$derivative enters as $\partial_y
2468: \tilde{\phi} + 2A_y$ - with the {\it wrong} sign. Thus, this quasiparticle
2469: is {\it not} a conventional electrically charged particle.
2470: Nevertheless, as we show below, condensation of the quasiparticle
2471: with $\langle e^{i\tilde{\phi}} \rangle \ne 0$ does drive the RL phase
2472: into a superconducting state.
2473:
2474: To evaluate the relevance of the various perturbations
2475: in the RL phase, requires diagonalizing the associated action, $S_{RL}$.
2476: In momentum space one has,
2477: \begin{equation}
2478: S_{RL} = \int_{k_\mu} [ \frac{D_{\phi \phi}}{2} |\tilde{\phi}|^2
2479: + \frac{D_{aa}}{2} |a|^2 + D_{a \phi} a(k_\mu) \tilde{\phi}(-k_\mu) ],
2480: \end{equation}
2481: \begin{equation}
2482: D_{\phi \phi} = {1 \over \pi^2 \kappa_r } [ \omega_n^2 + { u_v \kappa_r
2483: |{\cal K}_x {\cal K}_y |^2 \over {\cal K}^2 } ] ,
2484: \end{equation}
2485: \begin{equation}
2486: D_{aa} = {1 \over u_v } [ \omega_n^2 + \tilde{v}_0^2 {\cal K}^2 ] ,
2487: \end{equation}
2488: \begin{equation}
2489: D_{a \phi} = { \omega_n ({\cal K}_x^2 - {\cal K}_y^2) e^{i{\bf k} \cdot {\bf w}} \over 2 \pi {\cal K} } ,
2490: \end{equation}
2491: with $k_\mu = ({\bf k}, \omega_n)$.
2492: Evaluation of the two-point function of $e^{2i\tilde{\phi}}$ then gives,
2493: \begin{equation}
2494: \langle e^{i 2\tilde{\phi}_{\bf r}(\tau)} e^{-i2\tilde{\phi}_{\bf 0}(0)} \rangle_{RL}
2495: \sim { \delta_{{\bf r},{\bf 0}} |\tau|^{-2 \Delta_c } } ,
2496: \label{eq:e2phi}
2497: \end{equation}
2498: with scaling dimension,
2499: \begin{equation}
2500: \Delta_c = 2 \pi \sqrt{ \kappa_r \over u_v } { \tilde{v}_0 \over v_+ } .
2501: \label{Deltac}
2502: \end{equation}
2503: We note in passing that this power-law form of the ``charge''
2504: operator ($e^{2i\tilde{\phi}}$) two-point function in Eq.~(\ref{eq:e2phi})
2505: differs from $\exp(-\ln^2 \tau)$ behavior of the corresponding
2506: object in the Exciton Bose liquid of Ref.~\onlinecite{EBL}, due to
2507: the long-range logarithmic interactions between vortices.
2508: Stability of the RL phase requires $\Delta_c > 1$. Evaluation of
2509: the two-point function for the ``charge'' hopping operator,
2510: $e^{2i\partial_y \tilde{\phi}}$, gives,
2511: \begin{equation}
2512: \langle e^{i 2 \partial_y \tilde{\phi}_{\bf r}(\tau)} e^{-i2
2513: \partial_y \tilde{\phi}_{\bf 0}(0)} \rangle_{RL}
2514: \sim { \delta_{y,0} \over (x^2 + v_{rot}^2 \tau^2 )^{\Delta_c} } ,
2515: \label{eq:chargehopping}
2516: \end{equation}
2517: with ${\bf r} = x{\bf\hat{x}} + y {\bf\hat{y}}$, and with
2518: the {\it same} scaling dimension $\Delta_c$ as above.
2519: However, since this two-point function decays algebraically in
2520: two (rather than one) space-time dimensions, the perturbation
2521: $t_c$ is relevant for $\Delta_c < 2$.
2522:
2523: When $\Delta_c <2$, the ``charge'' hopping process will grow at low
2524: energies, and will destabilize the roton liquid phase. Not
2525: surprisingly, the resulting quantum ground state is
2526: superconducting. Indeed, the exponent $\Delta_c$ in Eq.~(\ref{Deltac}) above
2527: is in fact {\it identical} to the exponent characterizing the
2528: power law decay of the Cooper pair propagator in Eq.~(\ref{Cooperexp}),
2529: so that for small $\Delta_c$ the Roton liquid phase is already ``almost'' superconducting.
2530: Moreover, when the vortex core
2531: energy greatly exceeds the roton hopping strength, $u_v \gg
2532: \kappa_r$, the Hamiltonian $H_\theta$ in Eq.~(\ref{Htheta}) is
2533: deep within it's superconducting phase. This limit precisely
2534: corresponds to $\Delta_c \ll 1$, the limit where the ``charge''
2535: hopping is strongly relevant. More directly, when the ``charge''
2536: hopping strength $t_c$ grows large, the field $\tilde{\phi}$ gets trapped
2537: at the minimum of the cosine potentials in
2538: Eqs.~(\ref{cospert},\ref{chhoppert}), and it is legitimate to
2539: expand the cosine potentials to quadratic order. Once massive,
2540: the expectation value $\langle e^{i\tilde{\phi}} \rangle \ne 0$ - and the
2541: charge quasiparticle has condensed. Moreover, setting $\tilde{\phi}=0$ in
2542: Eq.~(\ref{chhoppert}) in the presence of an applied magnetic field
2543: will generate a term of the form,
2544: \begin{equation}
2545: {\cal L}_t \sim {t_c \over 2} \vec{A}^2 ,
2546: \end{equation}
2547: indicative of a Meissner response.
2548:
2549: In the resulting superconducting phase, the rotons - gapless in the
2550: RL phase - become gapped.
2551: This follows upon expanding the cosine term in Eq.~(\ref{cospert}),
2552: ${\cal L}_\lambda = 2 \lambda \tilde{\phi}^2$,
2553: which leaves the roton liquid Lagrangian quadratic, and allow one to readily extract the
2554: modified roton dispersion. For $k_x \rightarrow 0$ at fixed $k_y$, this gives:
2555: \begin{equation}
2556: \omega_{rot}({\bf k}) = \sqrt { v_{rot}^2 k_x^2 + m_{rot}^2 } ,
2557: \end{equation}
2558: with the ``roton mass gap'' given by, $m_{rot} = 2\pi \sqrt{\kappa_r \lambda}$.
2559:
2560:
2561: Since the product $\Delta_c \Delta_v = {1 \over 2}$, it is not
2562: possible to have both the vortex hopping and the ``charge'' hopping
2563: terms simultaneously irrelevant.
2564:
2565:
2566:
2567:
2568: \section{The Roton Fermi Liquid Phase}
2569: \label{sec:roton-spinon-liquid}
2570:
2571: We now put the fermions back into the description of the Roton liquid.
2572: We first consider setting the explicit pairing term in the fermion
2573: Hamiltonian ${\cal H}_f$ in Eq.~(\ref{eq:Hf}) to zero: $\Delta = 0$. As
2574: in Sec.~\ref{sec:roton-liquid}, we will assume for the most part (with
2575: the exception of Sec.~\ref{sec:phase-diagram} in the discussion) that
2576: the equilibrium fermion density is equal to the charge density $\rho_0$.
2577: This choice naturally minimizes Coulomb energy and vortex kinetic
2578: energy, as discussed therein. As we shall see, in this way we will
2579: arrive at a description of a novel non-Fermi liquid phase - the Roton
2580: Fermi liquid - which supports a gapless Fermi surface of quasiparticles
2581: coexisting with a gapless set of roton modes. We then reintroduce a
2582: non-zero pairing term, and study the perturbative effects of $\Delta$.
2583: We argue that, when the scaling exponent that describes the decay of the
2584: off-diagonal order in the Roton liquid is large enough, $\Delta_c>
2585: \Delta_c^* > 3/2$, the explicit pairing term is perturbatively
2586: irrelevant, and the RFL phase with a full gapless Fermi surface is
2587: stable.
2588:
2589: Even in the absence of the explicit pairing term -- which couples
2590: the fermionic and vortex degrees of freedom in a highly non-linear
2591: manner -- the rotons and quasiparticles interact through
2592: (three-)current-current interactions ``mediated'' by the
2593: $\beta_\mu$ and $e_i$ fields. Although these interactions are
2594: long-ranged for individual vortices, they are not for rotons,
2595: which carry no net vorticity. Moreover, due to phase space
2596: restrictions we find that the residual short-ranged interactions
2597: asymptotically decouple at low energies. The resulting RFL phase
2598: supports both gapless charge and spin excitations with no broken
2599: spatial or internal symmetries, just as in a conventional Fermi
2600: liquid. But, due to the vortex sector of the theory, the RFL
2601: phase is demonstrably a non-Fermi liquid, with a gapless ``Bose
2602: surface'' of Rotons and with ODQLRO in the Cooper pair field but
2603: no Meissner effect (see below). Moreover, the quasiparticles at
2604: the RFL Fermi surface are sharp (in the sense of the electron
2605: spectral function), but electrical currents are carried by the
2606: (quasi-)condensate. Even with impurities present the resistivity
2607: vanishes as a power law of temperature in the RFL. The power law
2608: exponent, $\gamma$, varies continuously, but is greater than or
2609: equal to one. For $\gamma <1$, the zero temperature RFL phase is
2610: unstable to a quantum confinement transition, which presumably
2611: drives the system into a conventional Fermi liquid phase.
2612:
2613: To describe the RFL, we start with the general Lagrangian,
2614: Eqs.~(\ref{Sfullb}-\ref{eq:Lf}), and first make the same approximation
2615: as in the RL of expanding the roton hopping term to quadratic order.
2616: That is, to leading order, ${\cal L}_{kin} \approx {\cal L}_r^0$, with
2617: \begin{eqnarray}
2618: {\cal L}_r^0 = {\kappa_r \over 2} [ \partial_{xy} \theta & -
2619: & {1 \over 2} \lbrace \partial_x (a_y - \alpha_y) + \partial_y(a_x - \alpha_x)
2620: \rbrace ]^2 \nonumber \\
2621: & & + {\kappa_r \over 8} [\epsilon_{ij} \partial_i (a_j - \alpha_j)]^2.
2622: \end{eqnarray}
2623: As before for the RL, this approximation will be corrected
2624: perturbatively by ``charge'' and vortex hopping terms, which will {\sl
2625: not} be expanded.
2626:
2627: Turning to the fermionic sector, we assume for the moment that the
2628: power-law decay (ODQLRO) of the Cooper pair field is sufficiently
2629: rapid that the pair field term $\Delta_j$ can be neglected. We
2630: argue later this is correct for $\Delta_c> \Delta_c^* > 3/2$.
2631: This gives a non-anomalous fermionic Lagrange density $L_f$, which
2632: we further presume is well-described by Fermi liquid theory, again
2633: checking the correctness of this assumption perturbatively in the
2634: couplings to $\beta_\mu$ and $e_i$. Hence we replace ${\cal L}_f
2635: \approx {\cal L}_f^0 + {\cal
2636: L}_f^1$, with (working for
2637: simplicity at zero temperature)
2638: \begin{eqnarray}
2639: \label{eq:Lf0}
2640: {\cal L}_f^0 & = & f^\dagger_{{\bf r}\sigma}(\partial_0 - \mu)
2641: f^{\vphantom\dagger}_{{\bf r}\sigma} -t \sum_j f^\dagger_{{\bf r} +
2642: {\bf \hat{x}}_j \sigma} f^{\vphantom\dagger}_{{\bf r} \sigma}, \\
2643: \label{eq:Lf1}
2644: {\cal L}_f^1 & = & -i\beta_0({\bf r}) f_{{\bf r}\sigma}^\dagger
2645: f^{\vphantom\dagger}_{{\bf r}\sigma} \\ && - i\sum_j
2646: [t\beta_j({\bf r}) +
2647: t_e (\pi\overline{e}_j({\bf
2648: r})+A_j({\bf r}))] f^\dagger_{{\bf r} +
2649: {\bf \hat{x}}_j \sigma} f^{\vphantom\dagger}_{{\bf r} \sigma}
2650: \nonumber \\
2651: & & \hspace{-0.4in} + \sum_j [\frac{t_s}{2} \beta_j({\bf r})^2 +
2652: \frac{t_e}{2}(\beta_j({\bf r})+ \pi\overline{e}_j({\bf
2653: r})+A_j({\bf r}))^2]
2654: f^\dagger_{{\bf r} +
2655: {\bf \hat{x}}_j \sigma} f^{\vphantom\dagger}_{{\bf r} \sigma} . \nonumber
2656: \end{eqnarray}
2657: Note that, to leading order, the fermionic dispersion is
2658: controlled by the sum of the two hopping amplitudes, $t=t_s+t_e$.
2659: Here we have included explicitly the physical external vector
2660: potential, $A_j({\bf r})$, in the electron hopping term. Some
2661: care needs to be taken when treating $A_j({\bf r})$. The above
2662: form is correct provided $A_j({\bf r})$ is also coupled into $e_j$
2663: in the quadratic Hamiltonian in the ``canonical'' fashion $\pi
2664: \overline{e}_j \rightarrow \pi \overline{e}_j + A_j$, in the
2665: roton/plasmon portion of the Hamiltonian. It is {\sl not} correct
2666: if this canonical coupling is removed by shifting
2667: $\overline{e}_j$, which is the procedure needed to generate the
2668: $A_\mu J_\mu$ coupling in Eq.~(\ref{LA}). If the latter form of
2669: the Lagrangian is used, the vector potential should be removed
2670: from the electron hopping term. Either choice is correct if used
2671: consistently.
2672:
2673: The full Lagrangian that we then use to access the RFL phase is
2674: given by,
2675: \begin{equation}
2676: {\cal L}_{RFL} = {\cal L}_a + {\cal L}_0 + {\cal L}_r^0 + {\cal
2677: L}_{cs} + {\cal L}_f^0 + {\cal L}_f^1 , \label{LRFL}
2678: \end{equation}
2679: with the definition,
2680: \begin{equation}
2681: {\cal L}_0 = {1 \over 2 u_0} (\partial_0 \theta - a_0 +
2682: \alpha_0)^2 .
2683: \end{equation}
2684: The interaction terms between the Fermions and the fields
2685: $\beta_j$ and $e_j$ in ${\cal L}_f^1$ will be treated in the
2686: random phase approximation. Doing so, one arrives at a tractable,
2687: if horrendously algebraically complicated Lagrangian describing
2688: the RFL, which is quadratic in the fields
2689: $\theta,a_\mu,e_j,\alpha_\mu$ and $\beta_\mu$. This makes
2690: calculations of nearly any physical quantity possible in the RFL.
2691:
2692: Before turning to these properties, we verify (in the remainder of this
2693: section) the above claims that the coupling of fermions and vortices
2694: does not destabilize the RFL -- i.e. it neither modifies the form of the
2695: low energy roton excitations at the ``Bose surface'' nor the fermionic
2696: quasiparticles at the Fermi surface.
2697:
2698: \subsection{Quasiparticle Scattering by Rotons}
2699: \label{sec:selectr-scatt-rotons}
2700:
2701: In this subsection, we show that the coupling of the electronic
2702: quasiparticles to the vortices does not destroy the Fermi surface.
2703: To do so, we will integrate out the vortex degrees of freedom to
2704: arrive at effective interactions amongst the quasiparticles. This
2705: procedure is somewhat gauge dependent. To provide a useful
2706: framework for the calculation of the electron spectral function in
2707: the following section, we will choose the gauge $\vec{\nabla}\cdot
2708: \vec{\beta} = \pi \epsilon_{ij}\partial_i e_j$, in which the
2709: $f,f^\dagger$ operators create fermionic quasiparticles with
2710: non-vanishing overlap with the bare electrons, without the need
2711: for any additional string operators. This is essentially
2712: equivalent to working in the electron formulation of Appendix
2713: \ref{ap:elecform}.
2714:
2715: Since this gauge choice explicitly involves
2716: $\epsilon_{ij}\partial_i e_j$, we must employ a path integral
2717: representation in which the transverse component of the electric
2718: field, $e^t_j$, has not been integrated out. It is further
2719: convenient to fix the two remaining gauge choices according to
2720: $\vec{\nabla}\cdot\vec{a}=\alpha_0=0$, and to integrate out the
2721: field $a_0$ in the $u_0 \rightarrow 0$ limit. The full RFL
2722: Lagrangian density (before imposing the constraint $\beta^l_j =
2723: \pi e^t_j$) then takes the form,
2724: \begin{equation}
2725: \label{eq:laget0}
2726: {\cal L}_{RFL} = {\cal L}_{vort} + {\cal L}^0_f(\beta_\mu) + {\cal
2727: L}_f^1(\beta_\mu) ,
2728: \end{equation}
2729: with
2730: \begin{eqnarray}
2731: \label{eq:laget}
2732: {\cal L}_{vort} & = & \frac{u_v}{2} (e^t_j)^2 + i e^t_j \partial_0 a^t_j +
2733: \frac{v_0^2}{2u_v} (\epsilon_{ij}\partial_i a^t_j - \pi
2734: \overline{\rho})^2 \\
2735: & & + \frac{i}{\pi}\epsilon_{ij} \alpha_i (\partial_j \beta_0 -
2736: \partial_0 \beta_j) + {\cal L}_\theta(\alpha_j-a^t_j) ,
2737: \nonumber
2738: \end{eqnarray}
2739: where
2740: \begin{eqnarray}
2741: \label{eq:Lthetaet}
2742: {\cal L}_\theta(\alpha_j) & = & \frac{1}{2u_v} (\partial_0 \partial_i
2743: \theta)^2 + \frac{\kappa_r}{8}(\epsilon_{ij}\partial_i \alpha_j)^2
2744: \nonumber \\
2745: & + & \frac{\kappa_r}{2}[\Delta_{xy}\theta +
2746: \frac{1}{2}(\partial_x\alpha_y+\partial_y\alpha_x)]^2 .
2747: \end{eqnarray}
2748:
2749: To assess the perturbative effects of the vortices upon the electronic
2750: quasiparticles, we wish to integrate out the vortices perturbatively
2751: in the coupling of the gauge field $\beta_\mu$ to the fermions. For
2752: this we require the correlation functions of the $\beta_\mu$ fields
2753: (neglecting the couplings inside ${\cal L}_f$, and in particular to
2754: lowest order just $\langle \beta_\mu \beta_\nu \rangle$). To obtain
2755: the latter, we add a source term to the Lagrangian,
2756: \begin{equation}
2757: \label{eq:betasource}
2758: {\cal L}_{vort} \rightarrow {\cal L}_{vort} +i( \lambda_0 i\beta_0 +
2759: \vec{\lambda}\cdot\vec{\beta}).
2760: \end{equation}
2761: Here we have included an extra factor of $i$ with $\beta_0$ to
2762: compensate for the factor of $i$ present in the coupling of
2763: $\beta_0$ to the fermion density. Upon fully integrating out the
2764: $e^t, \theta,a^t,\beta_\mu,\alpha_j$ fields, the coefficient of
2765: $\lambda_\mu \lambda_\nu$ in the effective action will give (half)
2766: the desired correlator. We perform the integration in two stages.
2767: First, imposing the constraint $\beta^l = \pi e^t$, we eliminate
2768: $e^t$, shift $\alpha_j \rightarrow \alpha_j - (\partial_j \theta -
2769: a^t_j)$, and integrate out the $\theta$, $a^t_j$, and $\beta_\mu$
2770: fields. One obtains ${\cal L}_{vort} \rightarrow \tilde{\cal
2771: L}_{vort}(\alpha_j,\lambda_\mu)$, with
2772: \begin{eqnarray}
2773: \label{eq:lvorttilde}
2774: \tilde{\cal L}_{vort} & = & \frac{v_0^2}{2u_v}
2775: (\epsilon_{ij}\partial_i \alpha_j +i\pi \tilde{\lambda}_0)^2 +
2776: \frac{\kappa_r}{4} [(\partial_x \alpha_y)^2 + (\partial_y
2777: \alpha_x)^2] \nonumber \\
2778: & & + \frac{1}{2u_v} (\partial_0 \alpha_i - \pi \epsilon_{ij}
2779: \lambda_j)^2 -t_v \cos(\alpha_j ),
2780: \end{eqnarray}
2781: with $\tilde\lambda_0 = \lambda_0 -i\overline{\rho}$.
2782:
2783: In Eq.~(\ref{eq:lvorttilde}) we have added back in the vortex
2784: hopping term, ${\cal L}_v = -t_v \cos(\alpha_j)$, neglected in
2785: the RFL Lagrangian. In the RFL phase, the vortex hopping is
2786: irrelevant, and scales to zero at low energies. If we put
2787: $t_v=0$, the remaining integrations over $\alpha_j$ are Gaussian
2788: and can be readily performed. We will return to the effects of
2789: non-zero vortex hopping upon the fermions in
2790: Sec.~\ref{subsec:resist}. The final effective action then takes
2791: the form,
2792: \begin{equation}
2793: \label{eq:Svorteff}
2794: S_{eff}(\lambda_\mu) = - \frac{1}{2}\int_{{\bf k},\omega_n}\!
2795: U^{(0)}_{\mu\nu}({\bf k},\omega_n) \lambda_\mu({\bf k},\omega_n)
2796: \lambda_\nu(-{\bf k},-\omega_n).
2797: \end{equation}
2798: Here
2799: \begin{equation}
2800: \label{eq:Umunu}
2801: U^{(0)}_{\mu\nu}({\bf k},\omega_n) = \frac{\pi^2}{u_v(\omega_n^2 +
2802: \omega_{pl}^2)(\omega_n^2+\omega_{rot}^2)} u_{\mu\nu}({\bf k},\omega_n)
2803: \end{equation}
2804: specifies the $\beta_\mu$ propagator: $\langle
2805: \beta_0\beta_0\rangle_0=U^{(0)}_{00}$ , $\langle
2806: \beta_i\beta_j\rangle_0=-U^{(0)}_{ij}$,$\langle
2807: i\beta_0\beta_j\rangle_0=-U^{(0)}_{0j}$, where the superscript
2808: zero reminds us that this is the result to zeroth order in
2809: $t_v=0$, and we have the definitions,
2810: \begin{eqnarray}
2811: \label{eq:lambdamunu}
2812: u_{00} & = & v_0^2[\omega_n^4 + \frac{v_1^2 {\cal K}^2}{2} \omega_n^2 +
2813: \frac{v_1^4}{4} |{\cal K}_x {\cal K}_y|^2], \\
2814: u_{xx} & = & -v_1^2[\tilde{v}_0^2 |{\cal K}_x {\cal K}_y|^2 + v_+^2
2815: |{\cal K}_x|^2 \omega_n^2], \\
2816: u_{yy} & = & -v_1^2[\tilde{v}_0^2 |{\cal K}_x {\cal K}_y|^2 + v_+^2
2817: |{\cal K}_y|^2 \omega_n^2], \\
2818: u_{xy} & = & -v_0^2 {\cal K}_x {\cal K}_y \omega_n^2, \\
2819: u_{0j} & = & -v_0^2 {\cal K}_x i\omega_n [2\omega_n^2 + v_1^2 ({\cal
2820: K}^2 - {\cal K}_j^2)].\label{eq:lambda0j}
2821: \end{eqnarray}
2822:
2823: In the d.c. limit, these interactions simplify considerably, and one
2824: obtains the simple results
2825: \begin{eqnarray}
2826: \label{eq:Udc}
2827: U^{(0)}_{00}({\bf k},\omega_n=0) & = & \frac{\pi^2 \kappa_r}{4}
2828: \left(\frac{v_0}{\tilde{v}_0}\right)^2, \\
2829: U^{(0)}_{ij}({\bf k},\omega_n=0) & = & -\frac{\pi^2}{u_v} \delta_{ij},
2830: \end{eqnarray}
2831: and $U^{(0)}_{0i}({\bf k},\omega_n=0) = 0$.
2832:
2833:
2834:
2835:
2836:
2837:
2838:
2839: Since $i\beta_0$ and $\beta$ couple to the fermion density and
2840: current respectively, ${\cal S}_{eff}$ mediates an effective
2841: frequency and wavevector dependent interaction between fermions.
2842: Since, in the d.c. limit, $-U^{(0)}_{00}({\bf k},\omega_n=0)$ and
2843: $U^{(0)}_{ii}({\bf k},\omega_n=0)$ are finite and wavevector
2844: independent, they describe local (on-site) repulsive quasiparticle
2845: density-density and attractive current-current interactions,
2846: respectively.
2847:
2848: Generally, a repulsive density-density interaction between fermions
2849: will lead to Fermi liquid corrections in the quasiparticle dynamics
2850: and thermodynamics, but will not destroy the Fermi surface. On the
2851: other hand, the current-current interaction for near-neighbor
2852: quasiparticle hopping can be rewritten in terms of antiferromagnetic,
2853: near-neighbor repulsive, and pairing interactions:
2854: \begin{equation}
2855: H_J = J \sum_{{\bf r}, {\bf r}^\prime} \left[{\bf S}_{\bf r} \cdot
2856: {\bf S}_{{\bf r}^\prime} + \frac{1}{4} n^f_{\bf r} n^f_{\bf r'}
2857: + 2\overline{\Delta}^f_{\bf r} \Delta^f_{\bf r'} \right] ,
2858: \end{equation}
2859: up to a shift of the fermion chemical potential, with ${\bf S} \equiv
2860: f^\dagger \frac{\vec\sigma}{2} f^{\vphantom\dagger}$, (with $\vec\sigma$
2861: the vector of Pauli matrices), $n^f = f^\dagger f^{\vphantom\dagger}$,
2862: $\Delta^f = f_\uparrow f_\downarrow$ and $\overline{\Delta}^f =
2863: f^\dagger_\downarrow f^\dagger_\uparrow$. Here, $J \sim t_s^2 /u_v$ is
2864: inversely proportional to the vortex core energy. One expects that the
2865: antiferromagnetic interaction can lead to an a Cooper instability in the
2866: $d$-wave (or extended $s-$wave) channel. The repulsive pairing
2867: interaction interaction dis-favors an $s$-wave Cooper instability.
2868: Hence it seems possible that for small $u_v$, for which this $J$ is
2869: large, a spontaneous quasiparticle pairing instability may occur, most
2870: probably of $d$-wave symmetry. Another possibility, which appears very
2871: natural for {\sl extremely} small doping $x\rightarrow 0^+$ (and $u_v
2872: \rightarrow 0^+$), is that the antiferromagnetic interaction drives an
2873: antiferromagnetic spin-density-wave instability. We will discuss both
2874: possibilities in the discussion section.
2875:
2876:
2877:
2878:
2879: \subsection{Roton Scattering by Quasiparticles}
2880: \label{sec:roton-scatt-selectr}
2881:
2882: Having established that the Fermi surface remains intact (apart
2883: from a possible BCS-type pairing instability) in the presence of
2884: gapless rotons, we need to see how the fermions feed back and
2885: effect the roton modes. To this end, we will integrate out all
2886: fields except $\theta$, treating the fermions within the random
2887: phase approximation (RPA), to study the effects upon the roton
2888: dispersion. The RPA is complicated by the two distinct
2889: amplitudes, $t_s$ and $t_e$, describing spinon and electron
2890: hopping processes, the latter coupling to the electric field $e_j$
2891: as well as the $\beta_j$ gauge field. We begin with the RFL
2892: Lagrangian in Eq.~(\ref{LRFL}). It is convenient to first shift
2893: $\alpha_\mu \rightarrow \alpha_\mu + a_\mu$ then integrate out
2894: $a_0$ and take $u_0 \rightarrow 0$, which constrains
2895: $\vec\nabla\times\vec\beta = \pi\vec\nabla\cdot \vec{e}$ and
2896: $\alpha_0 = -\partial_0 \theta$. Choosing in addition
2897: $\vec\nabla\cdot\vec\beta =- \pi \vec\nabla\times\vec{e}$, we
2898: essentially return to the electron formulation, with $\beta_i=-\pi
2899: \overline{e}_i$. This choice is convenient, since the $\beta_i$
2900: and $\pi\overline{e}_i$ fields present in the electron hopping
2901: term precisely cancel, and only the $-\pi \overline{e}_i$ appears
2902: ``like a gauge field'' in the spinon hopping term. The fermionic
2903: quasiparticles can then be integrated out in the random phase
2904: approximation (RPA). One finds $S_f \rightarrow S_{RPA}$, with
2905: \begin{eqnarray}
2906: S_{RPA}(\beta_\mu) & = & {1 \over 2} \int_{{\bf k},\omega_n}\!\!\!
2907: [ \Pi_{00} |\beta_0({\bf k},\omega_n)|^2 + \frac{\pi^2
2908: t_s^2}{t^2}\Pi_{11} |e_l({\bf k},\omega_n)|^2 \nonumber \\
2909: & & + \pi^2
2910: t_s(1\!-\!\frac{t_s}{t}\!) T_{nn} |e({\bf k},\omega_n)|^2 ] ,\label{eq:sRPA}
2911: \end{eqnarray}
2912: where $T_{nn} = \langle f^\dagger_{{\bf r}\sigma}
2913: f^{\vphantom\dagger}_{{\bf r+\hat{x}}\sigma}\rangle$ is the nearest
2914: neighbor kinetic energy, and $t=t_s+t_e$. Here $\Pi_{00}$ and
2915: $\Pi_{11}$ are respectively the density-density and current-current
2916: response kernels (polarization bubbles + diamagnetic contribution in the
2917: case of $\Pi_{11}$) that would obtain for the Fermi sea were a single
2918: gauge field minimally coupled to the fermions. These depend upon the band
2919: structure. We will not require particular expressions for these
2920: quantities, but will use the fact that both $\Pi_{00}({\bf k},\omega_n)$
2921: and $\Pi_{11}({\bf k},\omega_n)$ are finite and generally non-vanishing
2922: at {\sl fixed wavevector} $|{\bf k}|>0$ and $\omega_n=0$. Since we will
2923: focus upon the low energy rotons, which occur near the principle axes in
2924: the Brillouin zone, we have dropped terms in Eq.~(\ref{eq:sRPA}) that
2925: are small for $\omega_n \sim k_x \ll k_y$ and $\omega_n \sim k_y \ll
2926: k_x$ (e.g. due to the parity symmetry, $x \rightarrow -x$, of the 2d
2927: square lattice, $\Pi_{01}(k_x=0,k_y,\omega_n)$ vanishes). We also note
2928: that Eq.~(\ref{eq:sRPA}) does not have the usual RPA form even in this
2929: limit, due to the non-minimal coupling form of the fermion Lagrangian
2930: (minimal coupling is restored only for $t_s\rightarrow t$).
2931:
2932: It may be helpful to keep in mind the forms for a circular Fermi
2933: surface (e.g. valid at small electron densities), where, at small
2934: wavevectors and frequencies, one has
2935: \begin{equation}
2936: \Pi_{00}({\bf k}, \omega_n) \sim m^* ,
2937: \end{equation}
2938: where $m^* \sim t^{-1}$ is the effective mass and
2939: \begin{equation}
2940: \Pi_{11} ({\bf k}, \omega_n) \sim {1 \over m^*} [ {\bf k}^2 +
2941: {|\omega| \over v_F k } ] ,
2942: \end{equation}
2943: which is valid for $|\omega | < v_F k $. We stress, however, that our
2944: results do not depend upon these particular forms.
2945:
2946: Since $S_{RPA}$ is quadratic, one can perform the remaining
2947: integrals straightforwardly. We choose
2948: $\vec{\nabla}\cdot\vec{a}=\vec{\nabla}\cdot\vec{\alpha}=0$, and
2949: integrate out $a$, $\beta_0$ and $e$, to obtain
2950: \begin{eqnarray}
2951: \label{eq:atlag}
2952: S & = & \int_{{\bf k},\omega_n}\!\!\! \big\{
2953: \frac{\omega_n^2}{2\tilde{u}_v} \alpha^2 + \frac{\kappa_r}{8}\frac{4v_0^2 +
2954: \tilde{v}_1^2}{\tilde{v}_1^2}\alpha^2 + \frac{\omega_n^2 {\cal
2955: K}^2}{\overline{u}_v} \theta^2 \nonumber \\
2956: & &
2957: + \frac{\kappa_r}{2}|{\cal K}_x{\cal K}_y \theta + \frac{({\cal
2958: K}_x^2 - {\cal K}_y^2)}{2{\cal K}} \alpha|^2 \big\},
2959: \end{eqnarray}
2960: where $\tilde{u}_v = u_v + \pi^2 t_s (1-t_s/t) T_{nn}$,
2961: $\overline{u}_v = \tilde{u}_v + \pi^2 t_s^2/t^2 \Pi_{11}$, and
2962: $\tilde{v}_1^2 = v_1^2 + \kappa_r \pi^2 \Pi_0 v_0^2$. Without loss of
2963: generality, we focus upon the branch of rotons with $\omega_n \sim k_x
2964: \ll k_y \sim O(1)$, for which the first term in Eq.~(\ref{eq:atlag})
2965: is negligible, and the remaining $\alpha$ integral can be carried out
2966: to obtain finally the effective action
2967: \begin{equation}
2968: S_{rot} = \frac{1}{2}\int_{{\bf k},\omega_n}^\prime \frac{\tilde{\cal
2969: K}_y^2}{\overline{u}_v} \left[ \omega_n^2 + v_{rot}^2(k_y)
2970: k_x^2\right] |\theta|^2, \label{eq:Srotsec5}
2971: \end{equation}
2972: with
2973: \begin{equation}
2974: v_{rot}^2(k_y) = v_1^2 \frac{\overline{u}_v}{u_v}
2975: { v_0^2 + {1 \over 4} \tilde{v}_1^2 \over v_0^2 + {1 \over 2}
2976: \tilde{v}_1^2 } .\label{eq:vrotfull}
2977: \end{equation}
2978:
2979:
2980: We have thereby shown that even a gapless Fermi sea {\it
2981: does not} lead to a qualitative change in the gapless Bose-surface
2982: of rotons. Due to the $k_y$-dependence of $\Pi_{00}$ and $\Pi_{11}$
2983: (implicit in $\tilde{v}_1$ and $\overline{u}_v$), the roton velocity
2984: is now seen to depend upon $k_y$. Additional ``direct''
2985: quasiparticle-quasiparticle interactions (not mediated by the rotons)
2986: would in any case similarly renormalize $v_{rot}$. But the location
2987: of the Bose surface and the qualitative low energy dispersion of the
2988: roton modes are seen to be unaffected by the fermions. Together with
2989: the earlier demonstration of the stability of the Fermi
2990: surface, this result establishes the RFL phase as a stable 2d
2991: non-Fermi liquid with {\it both} gapless charge and spin excitations.
2992:
2993:
2994:
2995: \section{Instabilities of the Roton Fermi Liquid}
2996: \label{sec:inst-roton-spin}
2997:
2998: As for the RL, the RFL has potential instabilities to superconducting
2999: and Fermi Liquid states, driven by effects/terms neglected in the
3000: previous subsection. We address each in turn now.
3001:
3002: \subsection{Landau Fermi Liquid Instability}
3003: \label{sec:fermi-liqu-inst}
3004:
3005: The arguments of Sec.~\ref{sec:vortex-hopping} for determining the
3006: relevance or irrelevance of the neglected vortex hopping term
3007: within the simpler RL continue to hold for the full RFL, provided
3008: the proper renormalized roton liquid parameters are employed. In
3009: particular, the vortex hopping term continues to be described by a
3010: $1+1$-dimensional scaling dimension $\Delta_v$, which is, however,
3011: renormalized by the statistical interactions with the fermions, to
3012: wit:
3013: \begin{equation}
3014: \label{eq:deltavrenorm}
3015: \Delta_v = \frac{1}{4\pi} \int_{-\pi}^{\pi} \! \frac{dk_y}{2\pi}\,
3016: \frac{\overline{u}_v(k_y)}{v_{rot}(k_y)},
3017: \end{equation}
3018: with $v_{rot}(k_y)$ given from Eq.~(\ref{eq:vrotfull}). The same
3019: arguments thus continue to apply, and the vortex hopping term is
3020: irrelevant for $\Delta_v>2$. For $\Delta_v<2$, we expect an
3021: instability to a state with proliferated vortices. The dual
3022: Meissner effect for this vortex condensate confines particles with
3023: non-zero gauge charge, as discussed in
3024: Sec.~\ref{sec:vortex-hopping}. Coming from the RFL, the natural
3025: expectation is then that the system becomes a Fermi liquid. This
3026: hypothesis is fleshed out in Appendix \ref{sec:fermiliquid}.
3027:
3028: \subsection{Superconducting Instabilities}
3029: \label{sec:superc-inst-1}
3030:
3031: As for the RL, the RFL can also be unstable to a superconducting
3032: state. However, the inclusion of the fermionic quasiparticles opens new
3033: routes to superconductivity from the RFL. We explore each of
3034: these in turn.
3035:
3036: \subsubsection{``Charge'' hopping}
3037: \label{sec:charge-hopping-1}
3038:
3039: As for the vortex hopping considered above, the arguments for the
3040: relevance of the ``charge'' hopping in Sec.~\ref{sec:superc-inst} for
3041: the RL continue to apply with only a renormalization of the roton
3042: liquid parameters. In particular, the scaling dimension defined from
3043: the charge hopping, Eq.~(\ref{eq:chargehopping}), is modified to
3044: \begin{equation}
3045: \label{eq:deltacrenorm}
3046: \Delta_c = \pi \int_{-\pi}^\pi \! \frac{dk_y}{2\pi} |{\cal K}_y(k_y)|^2
3047: \frac{v_{rot}(k_y)}{\overline{u}_v(k_y)}.
3048: \end{equation}
3049: With this modification, the charge hopping processes become
3050: relevant for $\Delta_c<2$, as before. It should be noted that,
3051: including the renormalizations due to scattering by fermions,
3052: $\Delta_c \Delta_s \neq 1/2$, owing to the $k_y$ dependence of the
3053: $v_{rot}$ and $\overline{u}_v$.
3054:
3055: As we established in Sec.~\ref{sec:superc-inst} by employing the
3056: plaquette duality transformation, when the charge hopping
3057: processes are relevant the rotons become gapped out and the system
3058: exhibits a Meissner effect. Here, we briefly comment on the
3059: corresponding fate of the fermionic quasiparticles, which are
3060: gapless across the Fermi surface in the RFL phase. The relevance
3061: of the charge hopping in the plaquette dualized representation,
3062: indicates that it is not legitimate to expand the cosine term that
3063: enters into the roton hopping Lagrangian, ${\cal L}_r$. Rather,
3064: in the original vortex representation, the state corresponds to a
3065: ``vortex vacuum", and the properties of the state can be accessed
3066: by taking all of the vortex hopping processes small,
3067: $t_v,t_{2v},\kappa_r \ll u_v$. At zeroth order in the vortex
3068: kinetic terms, the full Hamiltonian of the $U(1)$ vortex-spinon
3069: field theory in Eq.~(\ref{HamU1vs}) is independent of $\alpha_j$,
3070: so that $\beta_j$ commutes with $\hat{H}$ and naively can be taken
3071: as c-numbers. A simple choice consistent with the condition
3072: $N_{\sf r}=0$ is $\beta_j=0$, and in the vortex vacuum one also
3073: has that $e^\ell_j =0$. In this case, the full fermionic
3074: Hamiltonian describing the quasiparticles in Eq.~(\ref{eq:Hf})
3075: reduces to the Bogoliubov-deGennes form, apart from the coupling
3076: to $e^t_j$ which describes the Doppler shift couplings to the
3077: super-flow. In particular, the string operator which enters into
3078: the explicit pairing term equals unity, ${\cal S}^2_{\bf r}=1$, so
3079: that the fermionic quasiparticles are paired. We denote by
3080: $|0\rangle_f$ the ground state of $\hat{H}_f$ with
3081: $\beta_j=e_j=0$, which is easily found by filling the
3082: Bogoliubov-deGennes levels below the Fermi energy. Our naive
3083: ground state is then
3084: \begin{equation}
3085: \label{eq:u1naive}
3086: |0\rangle_0 = |0\rangle_f \otimes |\beta_j=0\rangle \otimes |N_{\sf
3087: r}=0\rangle.
3088: \end{equation}
3089: In the U(1) sector, unfortunately, we must take somewhat more
3090: care, since the condition $\beta_j=0$ is in fact inconsistent with
3091: the gauge constraint ${\cal G}_f(\Lambda_{\bf r})=1$. We can,
3092: however, project into the U(1) subspace using the operator
3093: \begin{equation}
3094: \label{eq:projector}
3095: \hat{P}=\prod_{\bf r} \delta(f^\dagger_{{\bf r} \sigma}
3096: f^{\vphantom\dagger}_{{\bf r} \sigma} - {1 \over \pi}
3097: \epsilon_{ij} \partial_i \alpha_j({\bf r} - {\bf w})).
3098: \end{equation}
3099: Since $[\hat{P},\hat{H}]=[\hat{P},{\cal G}_v(\chi_{\sf r})]=0$,
3100: the state
3101: \begin{equation}
3102: \label{eq:projgs}
3103: |0\rangle = \hat{P}|0\rangle_0
3104: \end{equation}
3105: satisfies all gauge constraints and remains an eigenstate of $H$
3106: (with zero kinetic terms). This establishes that the resulting
3107: superconducting state is an ordinary BCS-type superconductor with
3108: paired electrons.
3109:
3110:
3111:
3112:
3113: \subsubsection{BCS instability}
3114: \label{sec:selectr-bcs-inst}
3115:
3116: As we saw in Section V, the coupling of the spinons to the gapless
3117: rotons in the RFL phase leads to Heisenberg exchange and pair
3118: field interactions between the fermions with strength $J \sim
3119: t_s^2/u_v$. In the physically interesting limit $t \approx t_s
3120: \gg t_e$, the quasiparticles move primarily as spinons, i.e.
3121: without any associated electrical charge, and hence do not
3122: experience a long-ranged Coulomb repulsion. Thus one may imagine
3123: that the above interactions could lead to a BCS pairing
3124: instability at ``high'' energies (still below the quasiparticle
3125: Fermi energy to make the Fermi liquid description appropriate, but
3126: quantitatively large compared to e.g. more conventional BCS
3127: critical temperatures) of order $J$. Here we explore the
3128: properties of the resulting phase which emerges when the fermions
3129: pair spontaneously and condense. To this end, we focus on the
3130: fluctuations of the phase, $\Phi$ of the fermion pair field,
3131: $\langle f_{\uparrow} f_{\downarrow} \rangle = \Delta_f
3132: e^{i\Phi}$. Keeping $\Delta_f$ fixed, and working once more in
3133: the electron gauge $\beta_i = -\pi \overline{e}_i$, we integrate
3134: out the fermions entering in ${\cal L}_{RFL}$ to generate an
3135: effective action for $\Phi, \beta_0$, and $\overline{e}_i$.
3136: Specifically, one obtains ${\cal L}^0_f + {\cal L}_f^1 \rightarrow
3137: {\cal
3138: L}_\Phi$, with
3139: \begin{eqnarray}
3140: {\cal L}_\Phi & = & \frac{g_f}{2} (\partial_0 \Phi - 2 \beta_0)^2 + \frac{
3141: \rho_s^f t_s}{2}(\partial_i\Phi + 2\pi
3142: \overline{e}_i)^2
3143: \\ & & \hspace{-0.5in} + \frac{\rho_s^f t_e}{2}(\partial_i \Phi - 2A_i)^2
3144: -\alpha\frac{ k_B T}{t\Delta_f} (t \partial_i \Phi + 2 \pi
3145: t_s\overline{e}_i - 2 t_e A_i)^2. \nonumber
3146: \label{eq:LPhi}
3147: \end{eqnarray}
3148: Here $g_f$ is of order the density of states at the Fermi surface, $A_i$
3149: is (time-independent for simplicity) external vector potential, and
3150: $\rho_s^f$ is the fermion pair (``superfluid'') density. The final term
3151: represents the low but non-zero temperature corrections to the
3152: superfluid density appropriate to the case of d-wave pairing. Here
3153: $\alpha$ is a non-universal constant of order one representing the
3154: effects of Fermi liquid corrections, or equivalently, high-energy
3155: renormalizations of the ``Doppler shift'' coupling constant.
3156:
3157: The excitations and response functions of the system can then be
3158: obtained from the RFL Lagrangian by shifting $\alpha_\mu
3159: \rightarrow \alpha_\mu + a_\mu$ and then integrating out $a_0,
3160: a_j$ and $\alpha_0$. Having made the replacement, ${\cal L}_f
3161: \rightarrow {\cal L}_\Phi$, one thereby obtains ${\cal L}_{RFL}
3162: \rightarrow {\cal L}_{eff}$, with an effective Lagrangian given by
3163: ${\cal L}_{eff} =\hat{\cal L}_{rot}+{\cal L}_\Phi$, where
3164: \begin{eqnarray}
3165: \label{eq:Lrothat}
3166: \hat{\cal L}_{rot} & = & \frac{u_v}{2} (e_i+ \frac{1}{\pi}
3167: \epsilon_{ij}A_j)^2 + i \partial_0 e_i (\partial_i\theta+\alpha_i)
3168: \nonumber \\
3169: & & +
3170: \frac{u_v}{2\pi^2 v_0^2}\beta_0^2 + \frac{i}{\pi} \beta_0
3171: \epsilon_{ij} \partial_i \alpha_j \\
3172: & & + \frac{\kappa_r}{8}(\epsilon_{ij}\partial_i \alpha_j)^2 +
3173: \frac{\kappa_r}{2}[ \Delta_{xy}\theta + \frac{1}{2}(\partial_x
3174: \alpha_y + \partial_y \alpha_x)]^2. \nonumber
3175: \end{eqnarray}
3176:
3177: First, it is instructive to consider the transverse electromagnetic
3178: response, in particular consider a static, transverse external gauge
3179: field, $\partial_0 A_i = \vec\nabla\cdot\vec{A}=0$. Since the external
3180: gauge field is at zero frequency, $\alpha_j,\beta_0$, $\theta$, and $e_t$ are
3181: decoupled from $A_i$ and $e_l$ in this limit, and we may thus neglect all but
3182: the first term in Eq.~\ref{eq:Lrothat}. In addition, $\Phi$ is
3183: decoupled from $e_l$ ($\overline{e}_t$) as well, so we may simplify the
3184: effective Lagrangian to
3185:
3186: \begin{eqnarray}
3187: \label{eq:efflagmeissner}
3188: {\cal L}_{eff}=
3189: \frac{u_v}{2}({e_l}+ \frac{A_t}{\pi })^2 +
3190: 2 t_e \rho_s^f A_t^2+ 2\pi^2t_s\rho_s^f e_l^2
3191: \nonumber \\ -\alpha\frac{ k_B T}{t\Delta_f} (2 \pi
3192: t_s e_l - 2 t_e A_t)^2.
3193: \end{eqnarray}
3194:
3195: Integrating over $e_l$ then gives the superfluid stiffness $K_s$ as the
3196: coefficient of $2 A_t^2$ (corresponding to the pair-field phase
3197: stiffness) in the effective action:
3198: \begin{eqnarray}
3199: \label{eq:rhos}
3200: K_s & & = \\
3201: & & \frac{\rho_s^f(t u_v + 4\pi^2 t_e t_s \rho_s^f
3202: )}{u_v + 4\pi^2 t_s\rho_s^f} -
3203: \frac{( t u_v + 4\pi^2 t_e t_s \rho_s^f )^2}{(u_v + 4\pi^2
3204: t_s\rho_s^f )^2} \frac{2\alpha k_B T}{t\Delta_f}\nonumber
3205: \end{eqnarray}
3206: to linear order in temperature. Note that, for $\rho_s^f\neq 0$, the
3207: system is a true charge superfluid, displaying a Meissner effect.
3208:
3209: We would next like to establish the fate of the roton excitations.
3210: To do these, we let $A_i=0$, specialize to $T=0$, and consider the
3211: appropriate limit $\omega_n \sim k_x \ll k_y\sim O(1)$,
3212: integrating out fields to obtain an effective action for $\theta$
3213: alone. Integrating out $\Phi$, one obtains in this limit
3214: \begin{eqnarray}
3215: \label{eq:Phiintegral}
3216: {\cal L}_\Phi & \rightarrow & 2 g_f \beta_0^2 + 2\pi^2 \rho_s^f t_s
3217: e_l^2,
3218: \end{eqnarray}
3219: using the above conditions on $\omega_n$ and ${\bf k}$, and $\partial_0
3220: \beta_0 \ll \epsilon_{ij}\partial_i e_j$, which follows in this limit.
3221: Further taking the gauge $\vec\nabla\cdot\vec\alpha=0$, one sees that
3222: $e_t$ couples only to $\partial_0\alpha$ in Eq.~(\ref{eq:Lrothat}), and
3223: thus generates only negligible terms at low frequencies and may be
3224: dropped. We therefore need only the effective action for
3225: $\alpha=\alpha_t$, $\theta$, $\beta_0$ and $e_l$, $S_{eff}=\int_{{\bf
3226: k}\omega_n} s_{eff}$, which in this limit becomes
3227: \begin{eqnarray}
3228: \label{eq:seffatbe}
3229: s_{eff} & & = \frac{u_v + 4\pi^2 \rho_s^f t_s}{2} e_l^2 +
3230: \frac{u_v+4\pi^2 v_0^2 g_f}{2\pi^2 v_0^2}\beta_0^2 - |{\cal
3231: K}_y|\omega_n e_l \theta \nonumber \\ && \hspace{-0.3in} + \frac{i|{\cal
3232: K}_y|}{\pi}\beta_0 \alpha
3233: +
3234: \frac{\kappa_r}{8}{\cal K}_y^2 \alpha^2 + \frac{\kappa_r {\cal
3235: K}_y^2}{2}|k_x \theta + \frac{{\rm sign}(k_y)}{2} \alpha|^2.
3236: \end{eqnarray}
3237: Performing the integration over $\alpha$, $\beta_0$ and $e_l$, one
3238: obtains the final effective action for $\theta$:
3239: \begin{eqnarray}
3240: \label{eq:Seffthetafinal}
3241: s_{eff}(\theta) & = & \frac{1}{2}\frac{{\cal K}_y^2}{u_v+4\pi^2
3242: \rho_s^f t_s}\left(\omega_n^2 + \tilde{v}_{rot}^2
3243: k_x^2\right)|\theta|^2,
3244: \end{eqnarray}
3245: where
3246: \begin{eqnarray}
3247: \label{eq:vrotss}
3248: \tilde{v}_{rot}^2 = v_1^2 \frac{u_v+4\pi^2 \rho_s^f t_s}{u_v}
3249: \frac{u_v(v_0^2+\frac{v_1^2}{4}) + \pi^2 g_f v_0^2 v_1^2}{u_v(v_0^2 +
3250: \frac{v_1^2}{2})
3251: + 2\pi^2 g_f v_0^2 v_1^2}.
3252: \end{eqnarray}
3253: Thus, despite the superconductivity induced by quasiparticle pairing,
3254: the gapless roton excitations survive, with some quantitative
3255: modifications to their velocity and correlations.
3256:
3257: As we shall explore further in the concluding sections, in the limit of
3258: a very small ``bare'' vortex core energy appropriate in the under-doped
3259: limit, $u_v \sim x \rightarrow 0$, there is a large energy scale for
3260: pairing, $J \sim t_s^2 /u_v$. It is then natural to assume $t\rho_s^f
3261: \approx t(1-x)
3262: \gg u_v$. If we presume that the fermionic kinetic energy is
3263: predominantly due to ``spinon hopping'', $t\approx t_s \gg t_e$, then
3264: one has
3265: \begin{equation}
3266: \label{eq:Ksapprox}
3267: K_s \approx \frac{u_v}{4\pi^2} + t_e\rho_s^f - (\frac{u_v}{4\pi^2} +
3268: t_e\rho_s^f)^2 \frac{2\alpha k_{\scriptscriptstyle B}T}{t\Delta_f
3269: (\rho_s^f)^2}.
3270: \end{equation}
3271: Thus, despite the large bare superfluid stiffness coming from the BCS
3272: pairing of the quasiparticles, the renormalized stiffness is small, set
3273: by the bare vortex core energy, $K_s = {u_v \over 4\pi^2} \sim x$ or the
3274: electron hopping $t_e$ (presumed small). It is the renormalized
3275: stiffness which determines the vortex core energy, and sets the scale
3276: for the finite temperature Kosterlitz-Thouless superconducting
3277: transition, $T_{KT} \sim u_v \sim x$. In this way, one can understand
3278: the large separation in energy scales between the pseudo-gap line at
3279: scale $J$ and the superconducting transition temperature, $T_{KT} \sim
3280: x$. Unfortunately (since it is in apparent conflict with the small
3281: amount of experimental data for this quantity), along with this small
3282: superfluid stiffness, one obtains a small linear temperature derivative,
3283: $\left.\partial K_s/\partial T\right|_{T=0}$, due to the same mechanism.
3284: This is similar to results for the U(1) gauge theory for the $t-J$
3285: model.
3286:
3287: To complete the analysis, one should consider the effects of the
3288: heretofore neglected explicit pairing term in this novel ``rotonic''
3289: superconductor. In particular, one may imagine that, once the fermionic
3290: pair field has condensed, it may induce true off-diagonal long-range
3291: order in the rotonic sector through the proximity effect. Naively,
3292: ODLRO appears incompatible with gapless rotons, so one may expect the
3293: explicit pairing term to induce a gap in the roton spectrum. While this
3294: is possible, it is easy to see that it is not inevitable. To see this,
3295: note that, in the rotonic superconductor, the fermionic pair-field in
3296: the explicit pairing term may be replaced simply by its mean-field value,
3297: $\Delta_1 c_{{\bf r}+\hat{\bf
3298: x}_1 \sigma} \epsilon_{\sigma\sigma'}c_{{\bf r}\sigma'} =
3299: \Delta_2 c_{{\bf r}+\hat{\bf
3300: x}_2 \sigma} \epsilon_{\sigma\sigma'}c_{{\bf
3301: r}\sigma'}=\Delta_f$. This leads, in the roton sector, to an
3302: additional term,
3303: \begin{eqnarray}
3304: \label{eq:Hdelta}
3305: H_\Delta & = & - \Delta_f \sum_{j\bf r} (B_{{\bf r},{\bf r}+\hat{\bf
3306: x}_j}^\dagger + {\rm h.c.}).
3307: \end{eqnarray}
3308: Indeed, Eq.~(\ref{eq:Hdelta}), since it embodies a linear coupling to
3309: the bosonic pair field, $B_{{\bf r},{\bf r}+\hat{\bf x}_j}$, will induce
3310: ODLRO in the $B_{\bf r}$ operators. However, this need not itself be a
3311: mechanism to induce a roton gap. Noting that from
3312: Sec.~\ref{sec:quasi-diagonal-long}, the boson pair field correlators are
3313: power-law in form, we expect that for $\Delta_c$ sufficiently large,
3314: perturbation theory in $\Delta_f$ will be regular and convergent, and
3315: the gapless rotons will be preserved. Due to the anisotropic structure
3316: of these correlators, we cannot reliably determine the relevance or
3317: irrelevance of $\Delta_f$ by simple power-counting arguments. We note
3318: that for $\Delta_c>3/2$, the induced ODLRO is expected to be linear in
3319: $\Delta_f$, i.e. the bosonic pair-field susceptibility is finite.
3320: However, the criterion for irrelevance is probably more stringent, e.g.
3321: $\Delta_c>3$. To determine the precise value $\Delta_c^*$ such that for
3322: $\Delta_c>\Delta_c^*$, the explicit pairing term remains irrelevant
3323: after condensation of quasiparticle pairs, would require a more careful
3324: analysis of the structure of the perturbation theory in $\Delta_f$. We
3325: leave this for braver souls, and content ourselves with the observation
3326: that there is a range of stability to this perturbation. We note that,
3327: however, since the RL itself even without the pairing term is always
3328: unstable to either vortex or ``charge'' hopping, the rotonic
3329: superconductor, with true gapless rotons, exists at best as an
3330: intermediate energy scale crossover. Nevertheless, provided the
3331: inevitable roton gap is small (i.e. for small $\Delta_f,t_c, t_v$), we
3332: expect the gap onset will only slightly modify the values for the superfluid
3333: stiffness and its linear temperature derivative determined above.
3334:
3335:
3336:
3337:
3338:
3339:
3340:
3341:
3342:
3343:
3344:
3345:
3346: \subsubsection{Explicit pairing term}
3347: \label{sec:stab-expl-spin}
3348:
3349: The final potential instability of the RFL we consider is from the
3350: explicit (spinon) pairing term $\Delta$. Recall from
3351: Sec.~\ref{sec:u1-vortex-spinon-1},
3352: \begin{eqnarray}
3353: \label{eq:explicit}
3354: {\cal H}_\Delta & = & \sum_j e^{i\beta_j({\bf r})} \Delta_j
3355: [{\cal S}_{\bf r}]^2 f_{{\bf r}+\hat{\bf
3356: x}_j \sigma} \epsilon_{\sigma\sigma'}f_{{\bf r}\sigma'}+ {\rm
3357: h.c.}\\
3358: & = & \sum_j \Delta_j B_{{\bf r},{\bf r}+\hat{\bf x}_j}^\dagger
3359: c_{{\bf r}+\hat{\bf
3360: x}_j \sigma} \epsilon_{\sigma\sigma'}c_{{\bf r}\sigma'}+ {\rm
3361: h.c.},\label{eq:explicitelec}
3362: \end{eqnarray}
3363: where the last line is written in electron variables, or equivalently in
3364: the ``electron gauge'' with $\beta_j=-\pi\overline{e}_j$. This
3365: representation is convenient because it eliminates all unphysical gauge
3366: fluctuations, which, although they do not appear in any physical
3367: properties, may enter inadvertently through approximations. Had we
3368: considered instead a Hamiltonian with $s$-wave pairing, we would have
3369: had
3370: \begin{eqnarray}
3371: \label{eq:Hswave}
3372: {\cal H}_\Delta^{s-{\rm wave}} & = & \Delta^s [{\cal S}_{\bf r}]^2
3373: f_{{\bf r}\sigma}\epsilon_{\sigma\sigma'}f_{{\bf r}\sigma'} + {\rm
3374: h.c.} \nonumber \\
3375: & = & \Delta^s B_{\bf r}^\dagger c_{{\bf
3376: r}\sigma}\epsilon_{\sigma\sigma'}c_{{\bf r}\sigma'} + {\rm
3377: h.c.}
3378: \end{eqnarray}
3379: Since we have already determined the correlations of the boson pair
3380: field operator $B_{\bf r}$ (which exhibits ODQLRO), in
3381: Sec.~\ref{sec:quasi-diagonal-long}, all the operators appearing in
3382: Eq.~(\ref{eq:Hswave}) have well understood properties at this stage, and
3383: so we will discuss this case for simplicity. For the physically more
3384: interesting scenario of $d$-wave pairing, we require instead the
3385: behavior of the bond pair field correlations. As discussed in
3386: Sec.~\ref{sec:odqlro-d-wave}, this behavior is qualitatively identical
3387: to that of the local pair field, with a possible renormalization of
3388: $\Delta_c$. Hence we believe that the results we obtain for local
3389: ($s$-wave) pairing -- in particular that the pairing term is irrelevant
3390: for sufficiently large $\Delta_c$ -- carry over to the $d$-wave case,
3391: provided a possible (and for the moment unknown) $O(1)$ modification of
3392: $\Delta_c$ is made in the relations.
3393:
3394: We would like to determine the ``relevance'' of ${\cal H}_\Delta$ in the
3395: renormalization group (RG) sense, i.e. whether its presence destabilizes
3396: the low energy properties of the RFL. Unfortunately, due to the
3397: extremely anisotropic nature of the rotonic spectrum, and the very
3398: different nature of the low energy electronic quasiparticle states at
3399: the Fermi surface, we do not know how to formulate a proper RG
3400: transformation. However, we do note that correlations of ${\cal
3401: H}_\Delta$ decay as power laws in space and imaginary time in the
3402: theory with $\Delta=0$, since then these correlations factor into
3403: $G^{cp}$ (with power-law behavior described in
3404: Sec.~\ref{sec:quasi-diagonal-long}) and the fermion pair-field
3405: correlator, which has the power law form characteristic of a Fermi
3406: liquid. Clearly, for sufficiently large $\Delta_c$, the Cooper pair
3407: propagator $G^{cp}({\bf r},\tau)$ will decay sufficiently rapidly that
3408: perturbation theory in $\Delta$ does not generate any (primitive)
3409: singularities. However, determining the critical $\Delta_c^*$ above
3410: which this occurs is beyond the scope of the current study. A simple
3411: argument clearly bounds $\Delta_c^*>3/2$. In particular, one may
3412: attempt to integrate out the vortices perturbatively in ${\cal
3413: H}_\Delta$ in a cumulant expansion. The first non-trivial term in
3414: this expansion occurs at second order, and generates an attractive
3415: fermion pair-field to pair-field interaction whose vertex is
3416: simply the Fourier transform of $G^{cp}({\bf r},\tau)$. For
3417: $\Delta_c<3/2$, a simple scaling analysis indicates that this
3418: Fourier transform is divergent at ${\bf q}=\omega_n=0$. Thus for
3419: such values of $\Delta_c$, this attractive Cooper channel
3420: interaction will overwhelm any other repulsive interaction that
3421: might be present at low energies, leading to a Cooper instability
3422: and pairing. This analysis, however, neglects higher cumulant
3423: terms which are certainly present in integrating out the vortices,
3424: and which are presumably crucial in determining the ultimate
3425: limits of stability of the RFL. Nevertheless, and this is all we
3426: shall require at present, it is clear that $\Delta_c^*$ exists and
3427: is not infinite, so that a non-vanishing region of stability also
3428: exists. When $\Delta_c < \Delta_c^*$ the explicit pairing term
3429: will be ``relevant'', destabilizing the RFL phase, presumably
3430: driving it into a conventional superconducting phase with paired
3431: electrons and gapped rotons. Given our lack of knowledge of
3432: $\Delta_c^*$, the additional uncertainty in the decay exponent of ODQLRO
3433: is not particularly damaging.
3434:
3435:
3436:
3437:
3438:
3439: \section{Properties of the RFL phase}
3440:
3441: \label{sec:RFLproperties}
3442:
3443: Having established the existence of the RFL phase, we now discuss
3444: some of it's properties. Since the rotons are effectively
3445: decoupled from the fermions at low energies, much of the physics
3446: of the RFL phase follows directly from the results we established
3447: for the charge sector in Sections III and IV. Specifically, one
3448: expects {\it three} different gapless excitations in the RFL phase
3449: - a gapless longitudinal plasmon, a Bose surface of gapless rotons
3450: and a set of particle-hole excitations across the Fermi surface.
3451: The {\it spin} physics in the RFL phase is then qualitatively
3452: similar to that of a conventional normal metal. Also like a
3453: metal, the RFL has ODQLRO in the Cooper pair field. Unlike an
3454: ordinary metal, however, this power law off-diagonal order exists
3455: with two distinct unrelated powers, one originating from the
3456: two-particle excitations of the Fermi sea, and the other from the
3457: gapless rotons. Because the rotons and electronic quasiparticles
3458: exist as nearly independent excitations in the RFL, one would
3459: expect a sharp electron spectral function despite the critical
3460: rotons. While this is true, as we demonstrate below in detail, the
3461: quasiparticles do scatter appreciably and in a singular manner
3462: from the gapless rotons at particular ``hot spots'' on the Fermi
3463: surface. Nevertheless, even at these hot spots, the decay rate
3464: vanishes more rapidly than the electron's energy.
3465:
3466: Despite the existence of such long lived electron-like quasiparticle
3467: excitations, the electrical transport in the RFL phase is strikingly
3468: non-Drude like, as we now demonstrate. Specifically, in the presence of
3469: impurities the electrical conductivity at low temperatures is dominated
3470: by the quasi-condensate in the charge sector, diverging at low
3471: temperatures as $\sigma \sim T^{-\gamma}$ with $\gamma >1$. But as we
3472: shall see, the {\it Hall conductivity} behaves very differently, being
3473: dominated by the electron-like quasiparticles.
3474:
3475: \subsection{Electrical Conductivity}
3476: \label{subsec:resist}
3477: Here we evaluate the electrical resistance in the RFL phase. As in
3478: Section IVA, it will be necessary to include the effects of the vortex
3479: hopping term, even though vortex hopping is technically ``irrelevant''
3480: when $\gamma >1$. The RFL phase exhibits the same ``emergent
3481: symmetry'' as in the Roton liquid -- the number of vortices on every
3482: row and column being independently conserved -- so inclusion of vortex
3483: hopping is necessary to generate {\it any} dissipative resistance
3484: whatsoever. In order to access the
3485: Hall conductivity we apply a uniform external magnetic field. We
3486: choose a gauge with $A_0=0$, and set,
3487: \begin{equation}
3488: A_j({\sf r},\tau) = A^B_j({\sf r}) + \tilde{A}_j(\tau) ,
3489: \end{equation}
3490: with $\epsilon_{ij} \partial_i A^B_j = B$ the external magnetic field
3491: and $\tilde{A}_j$ a time-dependent source term used to extract the
3492: conductivity tensor. As might be expected it is important to include
3493: the effects of elastic scattering from impurities.
3494:
3495: It is convenient to
3496: employ, as in Sec.~\ref{sec:selectr-bcs-inst}, in which the fermions
3497: have been integrated out, their effects felt only through an
3498: additional contribution to the effective action. In particular, we
3499: choose, as in Sec.~\ref{sec:selectr-bcs-inst}, the electron gauge, and
3500: write
3501: \begin{equation}
3502: \label{eq:actionforRFLsigma}
3503: S_{eff} = \int_\tau \sum_{\sf r} \left[ \hat{\cal L}_{rot} - t_v
3504: \cos(\partial_i \theta+\alpha_i)\right] + S_f(e_i,\beta_0,\tilde{A}_i),
3505: \end{equation}
3506: with $\hat{L}_{rot}$ given in Eq.~(\ref{eq:Lrothat}). Here $S_f$
3507: represents the fermionic terms in the action. We proceed by
3508: shifting $\overline{e}_j \rightarrow \overline{e}_j -A_j^B/\pi$,
3509: which takes $A_j \rightarrow \tilde{A}_j$ in $\hat{L}_{rot}$
3510: without any further changes since $A_j^B$ is time independent.
3511: This has the effect for the fermions of making the magnetic field
3512: appear uniformly in both electron and spinon hopping terms.
3513: Integrating out the fermions then perturbatively in
3514: $\overline{e}_i$, $\beta_0$, and $\tilde{A}_i$ effectively
3515: replaces $S_f \rightarrow \tilde{S}_f = \int_{{\bf k}\omega_n}
3516: s_f$, where
3517: \begin{eqnarray}
3518: \label{eq:SfRFLsigma}
3519: s_f & = & \frac{\Pi_{00}}{2} |\beta_0({\bf k},\omega_n)|^2 + \frac{\pi^2 t_s}{2} T_{nn}|\overline{e}_j|^2 + \frac{t_e}{2}
3520: T_{nn} |\tilde{A}_j|^2 \nonumber \\
3521: & & \hspace{-0.2in} +
3522: \frac{\tilde{\Pi}_{ij}}{2t^2}(\pi t_s \overline{e}_i + t_e
3523: \tilde{A}_i)_{{\bf k},\omega_n}(\pi t_s\overline{e}_j +t_e
3524: \tilde{A}_j)_{-({\bf k}\omega_n)}.
3525: \end{eqnarray}
3526: In Eq.~(\ref{eq:SfRFLsigma}), both $\Pi_{00}$ and $\tilde\Pi_{ij}$ are
3527: functions of ${\bf k}$ and $\omega_n$ (and $B$ through $A_j^B$ which
3528: appears in the electron hopping term), and we have neglected a
3529: cross-term $\Pi_{0i}$ between $\beta_0$ and the spatial gauge fields,
3530: which is negligible in all the limits of interest. We will require the
3531: behavior of $\tilde{\Pi}_{ij}$ in two regimes. Since the external
3532: fields are spatially uniform, we will need $\tilde{\Pi}({\bf
3533: k=0},\omega_n)$ at low frequencies
3534: \begin{eqnarray}
3535: \label{eq:Piuniform}
3536: \tilde{\Pi}_{ij}({\bf k=0},\omega_n) \approx \sigma_{xx}^f |\omega_n|
3537: \delta_{ij} + \sigma_{xy}^f \omega_n \epsilon_{ij},
3538: \end{eqnarray}
3539: where $\sigma_{ij}^f$ is the conductivity tensor for the fermions.
3540: Roton fluctuations are dominated in contrast by $\omega_n \sim k_x \ll
3541: k_y \sim O(1)$ (or the same with $k_x \leftrightarrow k_y$). In this
3542: limit, $\tilde{\Pi}_{ij}$ becomes a non-trivial function of the $O(1)$
3543: component of the wavevector, but has the useful property (due to square
3544: reflection symmetry) of decoupling in the longitudinal and transverse
3545: basis,
3546: \begin{eqnarray}
3547: \label{eq:Pilt}
3548: \tilde\Pi_{ij}({\bf k},\omega_n) \approx \tilde{\Pi}_t \left( \delta_{ij}
3549: - \frac{{\cal K}_i^{\vphantom{*}}{\cal K}_j^*}{{\cal K}^2}\right) +
3550: \tilde\Pi_l
3551: \frac{{\cal K}_i^{\vphantom{*}}{\cal K}_j^*}{{\cal K}^2} .
3552: \end{eqnarray}
3553: In this limit, we note that $\tilde\Pi_t$ is related to $\Pi_{11}$ of
3554: Sec.~\ref{sec:roton-scatt-selectr}\ by $\Pi_{11} = \tilde{\Pi}_t+
3555: T_{nn}$.
3556:
3557: As discussed above, the conductivity is finite only once the vortex
3558: hopping is included to break the row/column symmetries of the RFL.
3559: Hence to extract the conductivity, we must compute the effective
3560: action for $\tilde{A}_j$ to $O(t_v^2)$, which gives the first
3561: non-trivial correction. This requires only Gaussian integrals, with
3562: no further approximations. However, for ease of presentation, it is
3563: convenient, analogously to Sec.~\ref{sec:electr-resist-roton}, to take the
3564: simplifying limit $u_v/v_0^2 + \Pi_{00} \ll 1/\kappa_r$. In this
3565: limit the fluctuations of $\beta_0$ are extremely strong, which in
3566: turn strongly suppresses the fluctuations of $\alpha^t({\bf
3567: k},\omega_n)$ except at ${\bf k=0}$. Furthermore, choosing
3568: the gauge $\vec\nabla\cdot\vec\alpha=0$, we may thereby take
3569: $\alpha_i({\sf r},\tau)$ to be a function of $\tau$ only. Doing so,
3570: we may drop all spatial derivatives of $\alpha_j$ in $\hat{\cal
3571: L}_{rot}$. Furthermore, since fluctuations of $\alpha_j$ are only
3572: temporal, it can be accurately treated in an RPA fashion.
3573:
3574: To carry out the RPA calculation, we first perform the integral over
3575: $e_i$, which gives an effective action in terms of the remaining
3576: $\theta$, $\alpha_i$, and $\tilde{A}_i$ fields,
3577: $S_{eff}\rightarrow S'_{eff}=S'_1(\tilde{A}_i)
3578: + S'_2(\alpha_i,\tilde{A}_i)+ S'_3(\theta, \alpha_i)$, with
3579: \begin{widetext}
3580: \begin{eqnarray}
3581: \label{eq:speff}
3582: S'_1 & = & L^2\int_{\omega_n} \frac{1}{2}\big\{
3583: (\eta t_s + t_e)T_{nn}\delta_{ij}
3584: + \frac{1}{t^2}\left(t_e^2 +
3585: \eta^2 t_s^2+ 2 \eta t_e t_s \right)\tilde{\Pi}_{ij}({\bf
3586: k=0},\omega_n) \big\}
3587: \tilde{A}_i(-\omega_n)\tilde{A}_j(\omega_n), \\
3588: \label{eq:speff2}
3589: S'_2 & = & L^2 \int_{\omega_n} \left\{
3590: \frac{\eta \omega_n^2}{2u_v}\left[\delta_{ij} +\frac{\eta \pi^2
3591: t_s^2}{u_v t^2} \tilde{\Pi}_{ij}({\bf
3592: k=0},\omega_n)\right]\alpha_i(-\omega_n)\alpha_j(\omega_n)
3593: -\frac{\eta
3594: \omega_n}{\pi} \epsilon_{ij} \alpha_i(-\omega_n)
3595: \tilde{A}_j(\omega_n) \right\} \\
3596: \label{eq:speff3}
3597: S'_3 & =&\int_{{\bf k}\omega_n} \frac{1}{2}\left[\frac{\omega_n^2
3598: {\cal K}^2}{\overline{u}_v}+
3599: \kappa_r|{\cal K}_x{\cal K}_y|^2 \right]|\theta({\bf
3600: k},\omega_n)|^2 - t_v \sum_{\sf
3601: r}\int_\tau \cos(\partial_i \theta+\alpha_i),
3602: \end{eqnarray}
3603: \end{widetext}
3604: where $\eta = u_v/(u_v+\pi^2 t_s T_{nn})$, and $\overline{u}_v = u_v +
3605: \pi^2 t_s T_{nn} + \pi^2 t_s^2 \tilde{\Pi}_t/t^2 = \tilde{u}_v + \pi^2
3606: t_s^2 \Pi_{11}/t^2$ as in Sec.~\ref{sec:roton-scatt-selectr}. In
3607: Eqs.~(\ref{eq:speff}),~(\ref{eq:speff2}), $L^2$ is the system volume
3608: (number of sites in the square lattice), arising since
3609: $\alpha_i,\tilde{A}_i$ are spatially constant. In
3610: obtaining Eqs.~(\ref{eq:speff}-~\ref{eq:speff2}), we expanded to linear
3611: order in $\tilde\Pi_{ij}({\bf k=0},\omega_n)$ since we are interested
3612: ultimately in the low-frequency limit, and $\tilde{\Pi}_{ij}$ is linear
3613: in frequency. In Eq.~(\ref{eq:speff3}), we used the fact that $\theta$
3614: coupled to only to the longitudinal part of $e_i$ and hence only to
3615: $\tilde\Pi_t$. Note that the quadratic part of Eq.~(\ref{eq:speff3})
3616: reproduces precisely Eqs.~(\ref{eq:Srotsec5}-~\ref{eq:vrotfull}), in the
3617: limit $v_0 \gg \tilde{v}_1$, as assumed herein.
3618:
3619: We now can integrate out $\theta$ to $O(t_v^2)$. This proceeds
3620: identically as in Sec.~\ref{sec:electr-resist-roton}, with $a_j$
3621: replaced by $-\alpha_j$, and with the renormalized roton liquid
3622: parameters $u_v \rightarrow \overline{u}_v$. Hence we obtain the
3623: correction $S'_3(\theta,\alpha_i) \rightarrow S''_3(\alpha_j)$, with
3624: \begin{eqnarray}
3625: \label{eq:S3pp}
3626: S''_3 & = & \frac{t_v^2}{4} L^2 \int_{\omega_n} \tilde{\cal R}(\omega_n)
3627: |\alpha_j(\omega_n)|^2,
3628: \end{eqnarray}
3629: with $\tilde{\cal R}(\omega_n) \sim
3630: -|\omega_n|^{1+\gamma}$ as in Eq.~(\ref{rotpol}) of
3631: Sec.~\ref{sec:electr-resist-roton}, but $\gamma=2\Delta_v-3$ with the
3632: renormalized $\Delta_v$ given in Eq.~(\ref{eq:deltavrenorm}).
3633:
3634: With this replacement, the remaining quadratic integral over $\alpha_i$
3635: can be easily performed to determine the physical response kernel,
3636: $S'_1+S'_2+S''_3 \rightarrow S_{resp}(\tilde{A}_j)$, with
3637: \begin{eqnarray}
3638: \label{eq:Sresp}
3639: S_{resp} & = & L^2 \int_{\omega_n} \frac{1}{2} \tilde{A}_i(-\omega_n)
3640: \Pi^{RFL}_{ij}(i\omega_n) \tilde{A}_j(\omega_n),
3641: \end{eqnarray}
3642: where
3643: \begin{eqnarray}
3644: \label{eq:PiRFL}
3645: \Pi^{RFL}_{ij}(i\omega_n) & \approx & \tilde{\Pi}_{ij}^f +
3646: \tilde{\Pi}_{ij}^{rot},
3647: \end{eqnarray}
3648: and
3649: \begin{eqnarray}
3650: \label{eq:pif}
3651: \tilde{\Pi}_{ij}^f & = & t_e T_{nn}\delta_{ij} +
3652: \frac{t_e^2+2\eta t_e t_s}{t^2} \tilde\Pi_{ij}({\bf k\! =\!0},\omega_n) ,
3653: \\
3654: \tilde{\Pi}_{ij}^{rot} & = & \left[\frac{u_v}{\pi^2} - \frac{u_v^2 t_v^2
3655: \tilde{\cal R}(\omega_n)}{2\pi^2 \omega_n^2}\right]\delta_{ij}.
3656: \end{eqnarray}
3657: The conductivity is obtained from the Kubo formula as
3658: \begin{equation}
3659: \label{eq:kubo}
3660: \sigma_{ij}(\omega) = \left[\frac{\Pi_{ij}^{RFL}(i\omega_n)-t_e
3661: T_{nn}\delta_{ij}}{\omega_n} \right]_{i\omega_n\rightarrow
3662: \omega+i\delta}.
3663: \end{equation}
3664: As in Sec.~\ref{sec:electr-resist-roton}, we see that the rotonic
3665: contribution to the conductivity
3666: is not perturbative (at low frequencies) in $t_v$, so to capture the
3667: expected behavior, we replace it by,
3668: \begin{equation}
3669: \label{eq:pirotdrude}
3670: \sigma^{rot}_{ij} =
3671: \left[\frac{\tilde{\Pi}_{ij}^{rot}}{\omega_n}\right]_{i\omega_n\rightarrow \omega+i\delta}
3672: \rightarrow \frac{\delta_{ij}}{-i\omega \frac{\pi^2}{u_v} + i
3673: \frac{\pi^2 t_v^2}{2} \tilde{\cal R}_{ret}(\omega)/\omega } .
3674: \end{equation}
3675: This gives
3676: \begin{equation}
3677: \label{eq:condadd}
3678: \sigma_{ij}(\omega,T) = \sigma_{ij}^{rot}(\omega,T)+\frac{t_e^2+2\eta t_e
3679: t_s}{t^2}\sigma_{ij}^f(\omega, T) .
3680: \end{equation}
3681:
3682: Notice that from Eq.~(\ref{eq:condadd}), the conductivity is the
3683: sum of separate fermion and roton contributions, and that the only
3684: effects of the fermions upon the rotonic piece is to modify the
3685: exponent $\gamma$ (or $\Delta_v$) implicit in $\tilde{\cal
3686: R}(\omega)$. Moreover, the fermionic contribution vanishes for
3687: $t_e=0$, as expected on physical grounds since the spinon hopping
3688: term does not transport charge.
3689:
3690:
3691:
3692:
3693: Consider first the dissipative diagonal d.c. (sheet) resistance,
3694: $R(T) = Re[\sigma^{-1}_{xx}]$, which at low temperatures is
3695: completely dominated by the rotonic contribution, given as in
3696: Eq.~(\ref{eq:RLresist}),
3697: \begin{equation}
3698: \label{eq:RFLcond}
3699: R(T) = c_\gamma \frac{(\pi t_v)^2}{v_{rot}^2} \left[ \frac{\pi
3700: T}{v_{rot}}\right]^\gamma ,
3701: \end{equation}
3702: except with $v_{rot}$ and the exponent $\gamma$ renormalized by
3703: the interactions with the fermions. We thereby arrive at the
3704: important conclusion that the resistance in the RFL phase varies
3705: as $R(T) \sim T^\gamma$, with $\gamma >1$. When $\gamma < 1 $ the
3706: RFL phase is unstable to confinement into a Fermi liquid phase,
3707: and right at the quantum confinement transition the resistance is
3708: linear in temperature.
3709:
3710: The Hall conductivity and hence Hall angle on the other hand, are
3711: dominated by the fermionic quasiparticles. Specifically,
3712: $\sigma_{xy} \approx \sigma_{xy}^f$, and in the d.c. limit the
3713: fermionic Hall conductivity can be approximately obtained from a
3714: Drude expression, $\sigma_{xy}^f \approx \omega_c \tau_f (n_f e^2
3715: \tau_f)/m \sim \omega_c \tau_f^2$, where $\omega_c = eB/m$ is the
3716: cyclotron frequency. The fermionic scattering time $\tau_f$ has a
3717: variety of contributions, from elastic impurity scattering to
3718: interactions with rotons, considered in the next section, and
3719: hence may (or may not) have temperature dependence of its own - in
3720: contrast to the diagonal conductivity, $\sigma_{xx} \sim
3721: T^{-\gamma}$, which diverges as $T \rightarrow 0$ due to the
3722: rotonic contribution. These considerations suggest that the
3723: cotangent of the Hall angle, defined in terms of the resistivity
3724: tensor $\rho_{ij}$ as $\cot(\Theta_H) \equiv \rho_{xx} /
3725: \rho_{xy}$, will vary as,
3726: \begin{eqnarray}
3727: \label{eq:thetaH}
3728: \cot(\Theta_H) & \sim & \frac{\sigma_{xx}}{\sigma_{xy}} \sim
3729: \frac{T^{-\gamma}}{\omega_c \tau_f^2} \sim \frac{1}{\omega_c
3730: T^{\gamma}\tau_f^2} .
3731: \end{eqnarray}
3732:
3733:
3734:
3735:
3736:
3737:
3738:
3739:
3740:
3741:
3742:
3743:
3744:
3745:
3746:
3747:
3748:
3749: The complete absence of a roton contribution to the Hall conductivity
3750: $\sigma_{xy}^{RFL}$ is a consequence of the
3751: magnetic field independence of the roton liquid Lagrangian, ${\cal
3752: L}_{RL}$, either in it's original or dual forms, Eqs.~(\ref{LRL}) and
3753: (\ref{LRLdual}), respectively. But as discussed in Sec. IVB, the dual
3754: representation of ${\cal L}_{RL}$ allows for additional terms involving
3755: cosines of the dual field $\phi$, which are present due to the
3756: underlying discreteness of the vortex number operator. While being
3757: irrelevant in the Roton liquid phase, these neglected terms {\it do
3758: depend} on the external magnetic field and if retained will lead to a
3759: non-vanishing roton contribution to the Hall conductivity. But
3760: this contribution is expected to vanish rapidly at low
3761: temperatures. If on the other hand, these terms are {\it
3762: relevant} and drive a superconducting instability at low
3763: temperature, they will likely contribute signifigantly to the Hall
3764: conductivity. A careful analysis of the Hall response {\it above}
3765: the superconducting transition temperature in this situation will
3766: be left for future work.
3767:
3768:
3769:
3770:
3771: \subsection{Electron Lifetime}
3772:
3773:
3774:
3775:
3776: Although the electron spectral function is expected to be sharp at zero
3777: temperature right on the Fermi surface in the ideal RFL ``fixed point''
3778: theory, at finite energies (or temperatures) one expects the electrons
3779: to scatter off the rotons through interactions neglected in the RFL,
3780: causing a decay. Here, we consider two contributions to this electron
3781: decay rate at lowest order in the perturbations to the RFL. We consider
3782: two scattering mechanisms. First, scattering due to coupling of the
3783: quasiparticle current to boson currents through the $\beta_\mu$ and
3784: $e_i$ terms present in the fermionic hamiltonian. Because singular
3785: bosonic current fluctuations are primarily induced by vortex hopping,
3786: non-trivial contributions to the fermion lifetime through this mechanism
3787: occur first at $O(t_v^2)$. Second, we consider scattering of
3788: quasiparticles due to ``superconducting fluctuations'', i.e. fermion
3789: decay mediated by the ``Jospephson coupling'' $\Delta$ to the bosonic
3790: pair field. Since we expect $\Delta \ll t_v$ on physical grounds (see
3791: Sec.~\ref{sec:effective-parameters}), this $O(\Delta^2)$ contribution is
3792: naively much smaller than the former one. However, depending upon the
3793: parameters of the RFL, this need not be the case. This point will be
3794: returned to in the discussion.
3795:
3796:
3797: As discussed above, to compute the spectral function it is
3798: simplest to work in the gauge $\beta^\ell=\pi e^t$, in which we
3799: may assume the electron operator $c_{{\bf r}\sigma} \sim f_{{\bf
3800: r}\sigma}$, and hence study
3801: \begin{equation}
3802: {\cal G}_f ({\bf r}, \tau) = - \langle T_\tau f_{{\bf r} \sigma} (\tau)
3803: f^\dagger_{{\bf 0} \sigma}(0) \rangle .
3804: \end{equation}
3805: In other gauges, it would be necessary to include string factors as
3806: indicated in Eq.~(\ref{electronU1}).
3807:
3808: Neglecting the fluctuating $\beta_\mu$ fields, one has ${\cal
3809: G}_0({\bf k},\omega_n) = (i\omega_n - \epsilon_{\bf k} )^{-1}$, with
3810: dispersion, $\epsilon_{\bf k} = -2t \cos(k_j) - \mu_s$.
3811: Both fluctuations of the $\beta_\mu$ gauge fields, which mediate retarded
3812: interactions amongst the quasiparticles, as well as the explicit
3813: interactions in the pairing term, will induce a self
3814: energy $\Sigma({\bf k},\omega_n)$, defined by
3815: \begin{equation}
3816: \label{eq:selfenergydef}
3817: {\cal G}_0({\bf k},\omega_n) = \frac{1}{i\omega_n -\epsilon_{\bf
3818: k}-\Sigma({\bf k},\omega_n)}.
3819: \end{equation}
3820: We will compute the lowest order corrections to the
3821: electron self energy, obtaining a single-particle lifetime from the
3822: imaginary part of its retarded continuation, $\Sigma_{ret}({\bf
3823: k},\omega) = \Sigma({\bf k},\omega_n \rightarrow -i\omega+ 0^+)$ in
3824: the usual way. At the level of this discussion, the two types of
3825: interactions act in parallel to scatter quasiparticles, so
3826: \begin{eqnarray}
3827: \label{eq:addsigmas}
3828: \Sigma({\bf k},\omega) & = & \Sigma_{cf}({\bf k},\omega) +
3829: \Sigma_{sf}({\bf k},\omega).
3830: \end{eqnarray}
3831: We will focus upon the imaginary part $\Sigma''({\bf k},\omega)$, which
3832: describes broadening of the electron spectral function, $1/\tau_f(T) =
3833: \Sigma''({\bf k},0;T)$. We have
3834: \begin{eqnarray}
3835: \label{eq:matthieson}
3836: {\tau_f}^{-1} & = & (\tau_f^{cf})^{-1}+ (\tau_f^{sf})^{-1}.
3837: \end{eqnarray}
3838: In principle this single-particle ``lifetime'' is
3839: distinct from the momentum scattering rate which is relevant in
3840: discussions of transport quantities. However, we will use the behavior
3841: obtained for $1/\tau_f$ as a crude guide to the quasiparticle transport
3842: as well, leaving a more careful treatment for future study.
3843:
3844: \subsubsection{Scattering mediated by vortex hopping-enhanced current
3845: fluctuations}
3846: \label{sec:scatt-medi-vort}
3847:
3848: Taking into account quadratic fluctuations of the $\beta_\mu$ fields,
3849: the leading self energy correction due to current fluctuations is
3850: \begin{eqnarray}
3851: \Sigma_{cf}({\bf k}, i\omega_n) & = & \sum_{\mu\nu} \int_{{\bf q},
3852: \omega_n^\prime} \!\!\!\!
3853: v_{F\mu} v_{F\nu} {\cal G}_0({\bf k}-{\bf q},\omega_n - \omega_n^\prime)
3854: U_{\mu\nu}( {\bf q}, \omega_n^\prime) \nonumber \\
3855: & & \hspace{-0.75in}= \sum_{\mu\nu} \int_{{\bf q}, \omega_n^\prime}
3856: v_{F\mu} v_{F\nu} \frac{1}{i(\omega_n-\omega_n^\prime)-\epsilon_{{\bf
3857: k-q}}}U_{\mu\nu}( {\bf q}, \omega_n^\prime),
3858: \end{eqnarray}
3859: with the definitions, $v_{Fj} = \partial \epsilon_{\bf k} /
3860: \partial k_j$, $v_{F0} = 1$. Here we have written the self-energy
3861: in terms of the full $\beta-$gauge-field correlator, $U_{\mu\nu}$
3862: to all orders in $t_v$, rather than restricting to $t_v=0$ as we
3863: did in obtaining Eq.~(\ref{eq:Umunu}).
3864:
3865: Introducing a spectral representation,
3866: \begin{equation}
3867: U_{\mu\nu}({\bf k}, \omega_n) = \int^\infty_{-\infty} { d \omega \over \pi}
3868: { U_{\mu\nu}^{\prime\prime}({\bf q}, \omega) \over \omega - i\omega_n } ,
3869: \end{equation}
3870: with $U_{\mu\nu}^{\prime \prime}(\omega) = Im
3871: U_{\mu\nu}^{ret}(\omega) = Im U_{\mu\nu}(\omega_n \rightarrow
3872: -i\omega+ 0^+)$, allows one to analytically continue to obtain the
3873: (retarded) self energy,
3874: \begin{equation}
3875: \Sigma_{ret}({\bf k}, \omega) = \Sigma({\bf k}, \omega_n)|_{i \omega_n
3876: \rightarrow \omega + i0^+ } .
3877: \end{equation}
3878: For positive frequencies, $\omega > 0$, the
3879: imaginary part, $\Sigma^{\prime \prime} = Im \Sigma_{ret}$ is given by,
3880: \begin{eqnarray}
3881: \label{eq:Imself}
3882: \Sigma_{cf}^{\prime \prime}({\bf k}, \omega) & = & \\
3883: & & \hspace{-0.5in}\sum_{\mu\nu} \int_{\bf q} v_{F\mu} v_{F\nu}
3884: U_{\mu\nu}^{\prime \prime}({\bf q},\omega - \epsilon_{{\bf k} +
3885: {\bf q}} ) \Theta(\epsilon_{{\bf k} + {\bf q}}) \Theta(\omega -
3886: \epsilon_{{\bf k} + {\bf q}}).\nonumber
3887: \end{eqnarray}
3888:
3889: The self-energy can now be evaluated by considering progressive orders
3890: in $t_v$. To zeroth order, the expressions for $U_{\mu\nu}^{(0)}({\bf
3891: k},\omega_n)$ may be taken from
3892: Eqs.~(\ref{eq:Umunu}-\ref{eq:lambda0j}). We note that, because they
3893: come from the quadratic RL Lagrangian, they contain only simple poles.
3894: Furthermore, they are explicitly real functions of $i\omega_n$, so
3895: that, upon analytic continuation, their retarded correlator has zero
3896: imaginary part. Since $U_{\mu\nu}^{(0)\prime\prime}=0$, one thus
3897: has $\Sigma_{cf}^{(0)\prime\prime}({\bf k},\omega)=0$.
3898:
3899: Thus there is no broadening of the quasiparticle peak at zeroth
3900: order in the vortex hopping. The first non-trivial contribution
3901: to the quasiparticle lifetime occurs though the $O(t_v^2)$
3902: corrections to the gauge field propagator
3903: $U_{\mu\nu}=U^{(0)}_{\mu\nu}+t_v^2 U^{(2)}_{\mu\nu}+\cdots$. This
3904: correction is obtained by integrating out $\alpha_j$ to second
3905: order in $t_v$ starting from Eq.~(\ref{eq:lvorttilde}) to obtain
3906: $S_{eff}(\lambda_\mu)$ to $O(t_v^2)$ using the cumulant expansion.
3907: We shift $\alpha_i({\sf r},\tau) \rightarrow \alpha_i({\sf
3908: r},\tau) - \upsilon_i({\sf r},\tau)$, to eliminate the linear
3909: terms in $\alpha_i$ in the Lagrangian, with $\upsilon_i$ linear in
3910: $\lambda_\mu$ given explicitly by,
3911: \begin{widetext}
3912:
3913: \begin{eqnarray}
3914: \label{eq:alphashift}
3915: \upsilon_x({\bf k},\omega_n) & = & -\frac{\pi}{(\omega_n^2 +
3916: \omega_{pl}^2)(\omega_n^2+\omega_{rot}^2)} [ v_0^2 {\cal K}_x{\cal
3917: K}_y (v_0^2 {\cal K}_x \lambda_0 +i\omega_n \lambda_x) -
3918: (\omega_n^2 + v_+^2 |{\cal K}_x|^2)(v_0^2 {\cal K}_y \lambda_0
3919: +i\omega_n \lambda_y)], \\
3920: \upsilon_y({\bf k},\omega_n) & = & \frac{\pi}{(\omega_n^2 +
3921: \omega_{pl}^2)(\omega_n^2+\omega_{rot}^2)} [ v_0^2 {\cal K}_x{\cal
3922: K}_y (v_0^2 {\cal K}_y \lambda_0 + i\omega_n \lambda_y) -
3923: (\omega_n^2 + v_+^2 |{\cal K}_y|^2)(v_0^2 {\cal K}_x \lambda_0 +
3924: i\omega_n \lambda_x)].\nonumber
3925: \end{eqnarray}
3926: After this shift the correction to the effective action can be
3927: formally written,
3928: \begin{eqnarray}
3929: \label{eq:deltaSeff}
3930: S_{eff}^{(2)} & = & - \frac{t_v^2}{2}\sum_{ij}\sum_{{\sf
3931: r,r'}}\int_{\tau,\tau'} \big\langle \cos (\alpha_i({\sf r},\tau)
3932: - \upsilon_i({\sf r},\tau)) \cos (\alpha_j({\sf r},\tau)
3933: - \upsilon_j({\sf r},\tau)) \big\rangle_\alpha,
3934: \end{eqnarray}
3935: where the subscript $\alpha$ indicates a Gaussian average over
3936: $\alpha_i$ with respect to the quadratic terms in
3937: Eq.~(\ref{eq:lvorttilde}). Since $\upsilon_i$ is linear in
3938: $\lambda_\mu$, this correction is not quadratic in $\lambda_\mu$,
3939: indicating that the $\beta_\mu$ fluctuations are not Gaussian. To
3940: evaluate the leading order self-energy correction, however, we require
3941: only the two-point function of $\beta_\mu$, and hence may expand
3942: to quadratic order in $\upsilon_i$. One finds
3943: \begin{eqnarray}
3944: \label{eq:u2a}
3945: S_{eff}^{(2)} & = & -\frac{t_v^2}{4} \int_{{\bf k}\omega_n} \left(
3946: |\upsilon_x({\bf k},\omega_n)|^2 \tilde{\cal R}(k_y,\omega_n) +
3947: |\upsilon_y({\bf k},\omega_n)|^2 \tilde{\cal R}(k_x,\omega_n)\right).
3948: \end{eqnarray}
3949: Here $\tilde{\cal R}(k,\omega_n)$
3950: is the two-point function of the vortex hopping operator considered in
3951: Sec.~\ref{sec:vortex-hopping}, i.e.
3952: \begin{eqnarray}
3953: \label{eq:vorthopalpha}
3954: \tilde{\cal R}(k,\omega_n) & = & \sum_x \int_\tau (1-\cos(k x + \omega_n
3955: \tau)) \left\langle e^{i\alpha_y(x,y,\tau)} e^{-i\alpha_y(0,y,0)}
3956: \right\rangle_\alpha \nonumber \\
3957: & \sim & \frac{A(\Delta_v)}{\sin \pi(\Delta_v-1)} (\omega_n^2 +
3958: v_{rot}^2 k^2)^{\Delta_v-1},
3959: \end{eqnarray}
3960: \end{widetext}
3961: where the latter behavior gives the leading non-analytic term for
3962: small $\omega_n, k$, with $A(\Delta_v)$ a constant prefactor. An
3963: analytic quadratic term is also present (and larger than the given
3964: form for $\Delta_v>2$), but does not contribute to the imaginary part
3965: of the self-energy for the same reasons described above for the
3966: $t_v=0$ terms.
3967:
3968: Next, we must analytically continue to obtain the
3969: $U^{(2)\prime\prime}_{\mu\nu}({\bf k},\omega)$. From
3970: Eqs.~(\ref{eq:alphashift}), one immediately sees that
3971: $\upsilon_i^{ret}({\bf k},\omega)$ is purely real. Hence, the
3972: imaginary part comes entirely from the analytic continuation of
3973: ${\tilde R}(k_i,\omega_n)$. One has
3974: \begin{equation}
3975: \label{eq:rimag}
3976: {\cal R}''(k_i,\omega) \sim A(\Delta_v) (\omega^2-v_{rot}^2
3977: k_i^2)^{\Delta_v-1} \Theta(|\omega|-v_{rot}|k_i|){\rm sgn}(\omega).
3978: \end{equation}
3979: $ $From inspection of Eq.~(\ref{eq:u2a}), one thereby sees the imaginary
3980: part of the retarded gauge field correlator, $U^{(2)''}_{\mu\nu}({\bf
3981: k},\omega)$ , is the sum of two terms, which are non-zero only for
3982: $|\omega|>v_{rot}|k_x|$ or $\omega>v_{rot}|k_y|$, respectively.
3983: Thus the momentum integrals in the expression for $\Sigma''_{cf}$ in
3984: Eq.~(\ref{eq:Imself}), are constrained not only by
3985: $0<\epsilon_{\bf
3986: k+q}<\omega$ but also either $\epsilon_{\bf k+q}<
3987: \omega-v_{rot}|q_x|$ or $\epsilon_{\bf k+q}< \omega-v_{rot}|q_y|$, for
3988: the two terms, respectively.
3989:
3990: We focus on the self-energy for momenta exactly on the Fermi
3991: surface, $|{\bf k}|=k_F$. In this case, the constraints clearly
3992: require small ${\bf q}$ for small $\omega$. For such small
3993: wavevectors, one may approximate $\epsilon_{\bf k+q} \approx
3994: \vec{v}_F\cdot\vec{q}+ q^2/2m^*$. For a generic point on the
3995: Fermi surface (taken for simplicity in the quadrant with
3996: $k_x,k_y>0$ , for which $\vec{v}_F$ makes an angle $\theta_F \neq
3997: 0,\pi/2$ to the $x$ axis, the $q^2$ term is negligible, and one
3998: has $\epsilon_{\bf k+q} \approx v_F (q_x \cos\theta_F + q_y
3999: \sin\theta_F)$. Applying the above constraints, one finds that
4000: both $q_x$ and $q_y$ are integrated over a small bounded region in
4001: which both are $O(\omega)$. Thus for such generic points on the
4002: Fermi surface, we should consider the limit of
4003: $U^{(2)\prime\prime}_{\mu\nu}({\bf q},\omega-\epsilon_{\bf k+q})$
4004: in which all arguments are $O(\omega)$. In this limit, inspection
4005: of Eqs.~(\ref{eq:alphashift},\ref{eq:vorthopalpha}) shows that the
4006: correlator satisfies a scaling form,
4007: \begin{equation}
4008: \label{eq:genericscaling}
4009: U^{(2)\prime\prime}_{\mu\nu}({\bf q},\omega-\epsilon_{\bf k+q}) \sim
4010: \omega^{2(\Delta_v-2)} {\cal U}(q_x/\omega, q_y/\omega),
4011: \end{equation}
4012: with a well-behaved limit as ${\bf q}\rightarrow 0$. Inserting this
4013: into Eq.~(\ref{eq:Imself}) and rescaling the momentum integrals by
4014: $\omega$, one sees that $\Sigma''_{cf}({\bf k}_F,\omega) \sim
4015: \omega^{2(\Delta_v-1)}$ for $\theta_F \neq 0,\pi$. Since
4016: $\Delta_v\geq 2$ in the stable region of the RFL, this dependence is
4017: always as weak or weaker than the ordinary $\omega^2$ scattering rate
4018: due to Coulomb interactions in a Fermi liquid.
4019:
4020: The special cases when $\theta_F =0,\pi/2$ require separate
4021: consideration. For these values, the Fermi velocity is along one
4022: of the principal axes of the square lattice, and the linear
4023: approximation for $\epsilon_{\bf k+q}$ is inadequate. Taking for
4024: concreteness $\theta_F=0$, we have instead $\epsilon_{\bf k+q}
4025: \approx v_F q_x + q_y^2/2m^*$, and it is not obviously consistent
4026: to neglect the $q_y^2$ term. For the first term (involving
4027: $\tilde{R}''(q_y,\omega-\epsilon_{\bf k+q})$), the constraints
4028: reduce to $0<v_F q_x + q_y^2/2m^* < \omega-v_{rot} |q_y|$. This
4029: is approximately solved for small $\omega$ by $0< q_y <
4030: (\omega-v_{rot}|q_x|)/v_F$ and $ |q_x|<\omega/v_F$. Thus both
4031: $q_x,q_y$ are again bounded and $O(\omega)$, so the above Fermi
4032: liquid-like scaling applies.
4033:
4034: For the second term (involving
4035: $\tilde{R}''(q_x,\omega-\epsilon_{\bf k+q})$), the constraints reduce
4036: to $0<v_F q_x + q_y^2/2m^* < \omega - v_{rot} |q_x|$. This is solved
4037: by taking $-\omega/v_{rot}<q_x<\omega/(v_{rot}+v_F)$ and ${\rm
4038: max}(0,-v_F q_x) < q_y^2/2m^* < \omega -v_F q_x - v_{rot}|q_x|$.
4039: Hence for this term, $q_x$ is $O(\omega)$ while $q_y$ is
4040: $O(\sqrt{\omega})$. Since with this scaling, $|q_y| \gg \omega \sim
4041: |q_x|$, one has the significant simplification
4042: \begin{equation}
4043: \label{eq:upsilonkybig}
4044: \upsilon_y({\bf q},\omega) \sim \frac{\pi v_0^2 v_1^2}{2v_+^2}
4045: \frac{q_x \lambda_0}{v_{rot}^2q_x^2-\omega^2} \qquad {\rm for}\;|q_y| \gg
4046: \omega\sim |q_x| .
4047: \end{equation}
4048: Since only $\lambda_0$ appears, in this limit,
4049: $U^{(2)\prime\prime}_{00} \gg U^{(2)\prime\prime}_{ii},
4050: U^{(2)\prime\prime}_{0i}$, i.e. the fluctuations of $\beta_0$ are much
4051: stronger than those of $\beta_i$. One may therefore approximate
4052: \begin{eqnarray}
4053: \label{eq:thetaf0}
4054: \left.U^{(2)\prime\prime}_{\mu\nu}({\bf q},\omega)\right|_{\theta_F=0}
4055: & \!\!\! \sim & \!\!\!
4056: \frac{A(\Delta_v)\pi^2}{2} \frac{t_v^2 v_0^4 v_1^4}{v_+^4}
4057: q_x^2(\omega^2-v_{rot}^2q_x^2)^{\Delta_v-3} \nonumber \\
4058: & & \times \Theta(|\omega|-v_{rot}|q_x|) {\rm sgn}\omega \delta_{\mu
4059: 0}\delta_{\nu 0} .
4060: \end{eqnarray}
4061: Inserting this into Eq.~(\ref{eq:Imself}), one may integrate over
4062: $q_y$ (to yield a constant multiplying $\sqrt{|q_x|}$ and rescale $q_x
4063: \rightarrow \omega q_x$, to obtain the result
4064: \begin{equation}
4065: \label{eq:hotspots}
4066: \Sigma''_{cf}({\bf k}_F,\omega) \sim \omega^{2\Delta_v-\frac{5}{2}}\qquad
4067: {\rm for}\, \theta_F=0,\frac{\pi}{2},\cdots .
4068: \end{equation}
4069: At the end point of the RFL, for which $\Delta_v=2$,
4070: this gives the anomalous lifetime $\Sigma''({\bf k}_F,\omega) \sim
4071: \omega^{3/2}$ at these special ``hot spots'' for which the Fermi
4072: velocity is parallel to one of the principle axes of the square
4073: lattice. For the conventional Fermi surface believed to apply to the
4074: cuprates, this corresponds to the points on the Fermi surface closest
4075: to $(\pi,0)$, $(0,\pi)$.
4076:
4077:
4078:
4079:
4080:
4081:
4082:
4083:
4084:
4085:
4086:
4087:
4088: Comparing the scattering rate at these hot spots to elsewhere on the
4089: Fermi surface, one finds
4090: \begin{equation}
4091: { \Sigma^{\prime \prime}_{cf;hot}(\omega) \over \Sigma^{\prime
4092: \prime}_{cf;typ}(\omega) } \sim \sqrt{ m^* v_F^2 \over \hbar
4093: \omega } .
4094: \end{equation}
4095: Since the Fermi surface in the cuprates is particularly ``flat''
4096: near $(\pi,0)$, the effective mass would be large, $m^* \gg m_e$,
4097: and a significantly enhanced scattering rate at such ``hot spots''
4098: would be expected in the RFL phase.
4099:
4100: In parameter regimes where the RFL phase is unstable (via ``charge
4101: hopping") at low temperatures to a conventional superconductor,
4102: the enhanced scattering at the hot spots will be significantly
4103: suppressed upon cooling below the transition temperature. Indeed,
4104: the low energy roton excitations are gapped out in the
4105: superconducting phase, and at temperatures well below $T_c$ will
4106: not be appreciably thermally excited. Unimpeded by the rotons, the
4107: electron lifetime will be greatly enhanced relative to that in the
4108: normal state, particularly at the hot spots on the Fermi surface
4109: with tangents along the $\hat{x}$ or $\hat{y}$ axes, for example
4110: at wavevectors near $(\pi,0)$ in the cuprates.
4111:
4112: \subsubsection{Scattering mediated by ``superconducting fluctuations'',
4113: i.e. boson-fermion pair exchange}
4114: \label{sec:scatt-medi-boson}
4115:
4116: A second mechanism of fermion decay is by the ``pairing term'' ${\cal
4117: H}_\Delta$ in Eq.~(\ref{eq:explicit}). We supposed $|\Delta|$ is small,
4118: and so consider the first perturbative contribution to the quasiparticle
4119: lifetime, at $O(\Delta^2)$. In general, for $d$-wave pairing, this
4120: takes the form
4121: \begin{eqnarray}
4122: \label{eq:sigmadwave}
4123: \Sigma_{sf}({\bf k},\omega_n) & = & \sum_{i,j=1}^2
4124: \Delta_i\Delta_j\int_{{\bf q}\Omega_n}\!\!
4125: \frac{e^{ik_i}\!+\!e^{i(q_i-k_i)}}{2}
4126: \frac{e^{-ik_j}\!+\!e^{-i(q_j-k_j)}}{2}\nonumber \\
4127: & & \times
4128: \frac{G^{cp}_{ij}({\bf
4129: q},\Omega_n)}{-i(\Omega_n-\omega_n)+\epsilon_{\bf q-k}},
4130: \end{eqnarray}
4131: where
4132: \begin{eqnarray}
4133: \label{eq:Gcpijft}
4134: G^{cp}_{ij}({\bf q},\Omega_n) & = & \sum_{\bf r}\int_\tau
4135: G^{cp}_{ij}({\bf r},\tau)e^{i {\bf q}\cdot{\bf r}-i\Omega_n\tau},
4136: \end{eqnarray}
4137: with $G^{cp}_{ij}({\bf r},\tau)$ from Eq.~(\ref{eq:Gcpij}).
4138:
4139: Our uncertainties in the details of the Cooper pair propagator
4140: (originating from ambiguities in the string geometry) do not allow a
4141: thorough calculation of the resulting self energy. However, as we will
4142: now show, we can obtain a rough understanding of its scaling properties
4143: by some simple approximations, which we believe do not significantly
4144: effect the results. In particular, we will assume that the lifetime is
4145: controlled by the small wavevector $|q|\ll \pi$ portion of the Cooper
4146: pair propagator. In this regime, we expect $G^{cp}_{ij}({\bf
4147: q},\Omega_n) \approx G^{cp}({\bf q},\Omega_n)$, the Fourier transform
4148: of the {\sl local} Cooper pair propagator studied in depth in
4149: Sec.~\ref{sec:quasi-diagonal-long}. Making this approximation, one has
4150: $|q|\ll |k|$ for ${\bf k}$ near the quasiparticle Fermi surface, and
4151: one may write
4152: \begin{eqnarray}
4153: \label{eq:sigmadwave2}
4154: \Sigma_{sf}({\bf k},\omega_n) & \approx & |\Delta_{\bf k}|^2
4155: \int_{{\bf q}\Omega_n}
4156: \frac{G^{cp}({\bf
4157: q},\Omega_n)}{-i(\Omega_n-\omega_n)+\epsilon_{\bf q-k}},
4158: \end{eqnarray}
4159: with $\Delta_{\bf k}=|\Delta|(\cos k_x-\cos k_y)$. Note that this form
4160: immediately implies that scattering due to this mechanism is strongly
4161: suppressed upon approaching the nodal regions of the Fermi surface.
4162:
4163: A detailed analysis is now possible based on a spectral representation
4164: of $G^{cp}$, as in Sec.~\ref{sec:scatt-medi-vort}.\cite{BFunpub}\ Unlike
4165: the above case, however (since the power law decay of $G^{cp}({\bf
4166: r},\tau)$ is approximately isotropic) a much simpler scaling analysis
4167: suffices to obtain the qualitative behavior of the lifetime. In particular, the scaling form
4168: of Eq.~(\ref{eq:scaling}) implies that
4169: \begin{eqnarray}
4170: \label{eq:Gcpscalingqw}
4171: G^{cp}({\bf q},\Omega_n) \sim |\Omega_n|^{2\Delta_c-3} \tilde{\cal
4172: G}({\bf q}/\Omega_n),
4173: \end{eqnarray}
4174: up to non-singular additive corrections, for small $|q|$ and $\Omega_n$.
4175: Furthermore, for small ${\bf q}$ and ${\bf k}$ on the Fermi surface, one
4176: may write $\epsilon_{\bf q-k} \approx v_{\scriptscriptstyle F}
4177: q_\parallel + q_\perp^2/2m$, where $q_\parallel,q_\perp$ are the
4178: components of ${\bf q}$ parallel and perpendicular to the local Fermi
4179: velocity at ${\bf k}$, and $v_{\scriptscriptstyle F}$ and $m$ are the
4180: magnitude of the local Fermi velocity and effective mass. The
4181: singularity in the integrand in Eq.~(\ref{eq:sigmadwave2}) is then cut
4182: off by the external frequency $\omega_n$, and rescaling ${\bf
4183: q}\rightarrow {\bf q}\Omega_n$, one sees that the effective mass term
4184: is negligible, and obtains by power counting,
4185: \begin{eqnarray}
4186: \label{eq:Sigmapc}
4187: \Sigma_{sf}({\bf k},\omega_n) \sim |\Delta_{\bf k}|^2
4188: |\omega_n|^{2\Delta_c-1},
4189: \end{eqnarray}
4190: again with possible analytic and sub-dominant corrections. Upon analytic
4191: continuation to obtain the lifetime, only non-analytic terms contribute,
4192: and we expect
4193: \begin{eqnarray}
4194: \label{eq:sigmapppc}
4195: \Sigma''_{sf}({\bf k},\omega) \sim |\Delta_{\bf k}|^2
4196: |\omega|^{2\Delta_c-1} {\rm sign}(\omega),
4197: \end{eqnarray}
4198: for ${\bf k}$ on the Fermi surface. This result can be verified by
4199: more detailed calculations using the spectral representation of
4200: $G^{cp}$.\cite{BFunpub} Moreover, we expect that for
4201: $k_{\scriptscriptstyle B}T \gg \omega$, $\omega$ can be replaced by
4202: $k_{\scriptscriptstyle B}T$ in this formula.
4203:
4204: As expected, for sufficiently large $\Delta_c$, this lifetime vanishes
4205: rapidly at low energies, and in particular for $\Delta_c>3/2$, this
4206: contribution is sub-dominant to the usual Fermi liquid one. However, in
4207: the regime with $\Delta_c<\Delta_v$, this is never the case ($\Delta_c <
4208: 1/\sqrt{2}$), at least within our simple model without dramatic
4209: corrections to the RL exponents. Indeed, for $\Delta_c<1$, as supposed
4210: herein, this ``scattering rate'' is much {\sl larger} than the
4211: quasiparticle energy at low frequency. Taken literally, this implies
4212: increasingly incoherent behavior away from the nodes as temperature is
4213: lowered. Near the nodes, the amplitude of this strong scattering
4214: contribution vanishes rapidly, similar to the idea of ``cold spots''
4215: proposed by Ioffe and Millis.\cite{IoffeMillis} \
4216:
4217: \section{Discussion}
4218:
4219: \label{sec:Discussion}
4220:
4221: We close with a discussion of some theoretical issues concerning
4222: the vortex-quasiparticle formulation of interacting electrons that
4223: we have been employing throughout. In particular, we contrast our
4224: approach with the earlier $Z_2$ formulation\cite{Z2} and mention
4225: the connection with more standard and more microscopic
4226: formulations of correlated electrons. We then address the possible
4227: relevance of the Roton Fermi Liquid phase to the cuprate phase
4228: diagram and the associated experimental phenomenology.
4229:
4230:
4231: \subsection{Theoretical Issues}
4232:
4233: Since the discovery of the cuprate superconductors, the many
4234: attempts to reformulate theories of two-dimensional strongly
4235: correlated electron in terms of new collective or composite
4236: degrees of freedom, have been fueled on the one hand by related
4237: theoretical successes and on the other by cuprate phenomenology.
4238: The remarkable and successful development of {\it
4239: bosonization}\cite{Bosonization} both as a reformulation of
4240: interacting one-dimensional electron models in terms of bosonic
4241: fields and as a means to extract qualitatively new physics outside
4242: of the Fermi liquid paradigm, played an important role in a number
4243: of early theoretical approaches to the cuprates\cite{PWAbook}. The
4244: equally impressive successes of the {\it composite boson} and {\it
4245: composite fermion} approaches in the fractional quantum Hall
4246: effect\cite{FQHEbook} were also influential in early high $T_c$
4247: theories, most notably the {\it anyon} theories\cite{Anyon}. Gauge
4248: theories of the Heisenberg and $t-J$ models were notable early
4249: attempts to reformulate 2d interacting electrons in terms of
4250: ``spin-charge'' separated variables\cite{U1A,U1B} -- electrically
4251: neutral fermionic spin one-half ``spinons'' and charge $e$ bosonic
4252: holons -- and were motivated both by analogies with 1d
4253: bosonization and by resonating valence bond ideas\cite{PWAbook}.
4254: More recent approaches to 2d spin-charge separation have
4255: highlighted the connections with superconductivity\cite{NL,Z2}, by
4256: developing a formulation in terms of vortices, Bogoliubov
4257: quasiparticles and plasmons: the three basic collective
4258: excitations of a 2d superconductor. These latter theories were
4259: primarily attempting to access the pseudo-gap regime by approaching
4260: from the superconducting phase\cite{NL} - focusing on low energy
4261: physics where appreciable pairing correlations were manifest.
4262:
4263: \subsubsection{Vortex-fermion formulation}
4264:
4265:
4266: In this paper, although we are employing a formulation with the same
4267: field content -- $hc/2e$ vortices, fermionic quasiparticles and
4268: collective plasmons -- we are advocating a rather different philosophy.
4269: In particular, we wish to use these same fields to describe higher
4270: energy physics in regimes where the physics is decidedly non-BCS-like
4271: even at short distances, most importantly the cuprate normal state near
4272: optimal doping. The philosophy of this approach is very similar to that
4273: of the $Z_2$ gauge theory proposal of a fractionalized under-doped normal
4274: state. Indeed, as demonstrated in Sec.~\ref{sec:u1-vortex-spinon} and
4275: Apps.~\ref{ap:enslaveZ2}-\ref{ap:enslaveU1}, our $U(1)$ formulation is
4276: completely (unitarily) equivalent to a $Z_2$ gauge theory. However,
4277: because the unitary transformation relating the two formulations is
4278: non-trivial, the $U(1)$ formulation is (much) more convenient for the
4279: types of manipulations and approximations we employ here, largely
4280: because the $U(1)$ gauge fields are continuous variables. The price
4281: paid for the use of the $U(1)$ formulation is that the ``spinon pairing
4282: term'' of the $Z_2$ gauge theory appears non-local in the $U(1)$
4283: Hamiltonian.
4284:
4285: Fortunately, this non-locality and its consequences are readily
4286: understood. In particular, although the $U(1)$ vortex-quasiparticle
4287: Hamiltonian is microscopically equivalent to a $Z_2$ gauge theory, it is
4288: {\sl also} equivalent (as described in App.~\ref{ap:elecform}) to a
4289: theory of electronic (i.e. charge $e$) quasiparticles coupled to charge
4290: $2e$ bosons (with the latter described in dual vortex variables). The
4291: ``two-fluid'' point of view of this vortex-electron formulation is
4292: convenient for understanding the qualitative properties of the RFL phase
4293: and its descendents, although it should be emphasized that the
4294: current-current couplings (embodied by the gauge fields of the $U(1)$
4295: formulation) between the two fluids are not expected to be weak. In the
4296: vortex-electron language, the non-local term simply represents coherent
4297: ``Josephson coupling'' of electron pairs and the bosons. The
4298: non-locality of the pairing term arises simply from the non-local
4299: representation of boson operators in the usual $U(1)$ boson-vortex
4300: duality. The crucial feature which allows us to handle the pairing term
4301: despite its non-locality is the strength of vortex fluctuations in the
4302: bosonic Roton Liquid, which persists even with these strong
4303: current-current couplings. The resulting power-law decay of bosonic
4304: pair-field correlations (ODQLRO) opens up a regime, corresponding to the
4305: RFL phase, in which the pairing term is irrelevant and can be treated
4306: perturbatively. We note in passing that, although it is not relevant
4307: for the cuprates, the above discussion makes clear that a RL phase (for
4308: which fermions need never be introduced) should be possible in a purely
4309: bosonic model, which would be interesting in and of itself.
4310:
4311:
4312: A second important feature of the present formulation is the retention
4313: of the lattice scale structure. Appropriate to the cuprates, we have
4314: carefully defined the theory on a 2d square lattice which we wish to
4315: identify with the microscopic copper lattice. Since our theory obviously
4316: does not similarly retain physics on atomic {\it energy} scales, our
4317: starting ``bare'' Hamiltonian should be viewed as a {\it low energy} but
4318: not spatially coarse grained effective theory (or at most an effective
4319: theory coarse-grained spatially only insofar as to remove e.g. the $O$
4320: $p$-orbitals). It is important to emphasize that the very existence of
4321: the Roton Fermi liquid phase {\it requires} the presence of a square
4322: lattice. In the RFL ground state there is an infinite set of
4323: dynamically generated symmetries, corresponding to a conserved number of
4324: vortices on every row and column of the lattice. If we study the same
4325: model on a different lattice, say the triangular lattice, the quantum
4326: ground state analogous to the RFL phase (obtained, again, by taking the
4327: roton hopping amplitude $\kappa_r \rightarrow \infty$) is highly
4328: unstable, being destroyed by the presence of an arbitrarily small single
4329: vortex hopping amplitude.
4330:
4331: More generally, the importance of employing the dual
4332: vortex-quasiparticle field theory reformulation cannot be
4333: overemphasized. The novel and unusual RFL phase emerges quite simply
4334: when one takes a large amplitude for the roton hopping term, and the
4335: fixed point theory is also quite simple consisting of a Fermi sea of
4336: quasiparticles minimally coupled to an ``electric'' field with Gaussian
4337: dynamics. It is very difficult to see how one could adequately describe
4338: such a phase working with bare electron operators. Indeed, we do not
4339: yet understand the simpler task of constructing a microscopic
4340: (undualized) boson model which enters the RL phase, even without the
4341: complications of fermions. However, previous experience with dual
4342: vortex formulations for bosonic and electronic systems strongly argues
4343: that the RL and RFL theories properly describe physically accessible
4344: (albeit microscopically unknown) models.
4345:
4346: An unfortunate drawback of the present formulation, however, is
4347: the apparent lack of {\it direct} connection between the starting
4348: lattice Hamiltonian and any microscopic electron model. In
4349: particular, it is presently unclear what microscopic electron
4350: physics could be responsible for generating such a large roton
4351: hopping term. In the absence of a microscopic foundation, one
4352: must resort to developing phenomenological implications of the
4353: Roton Fermi Liquid phase and comparing with the experimentally
4354: observed cuprate phenomenology. We take a preliminary look in
4355: this direction in Sec.~\ref{sec:cuprates-rfl-phase} below.
4356:
4357:
4358: \subsubsection{Related approaches}
4359:
4360: It is instructive to contrast the RFL phase with earlier notions of
4361: spin-charge separation. Anderson's original picture of the cuprate
4362: normal state at optimal doping consisted of gas of spinons with a
4363: Fermi surface co-existing and somehow weakly interacting with a gas of
4364: holons\cite{PWA1,PWAbook} In the early gauge theory implementations of
4365: this picture, both the spinons and holons carried a $U(1)$ (or
4366: $SU(2)$) gauge charge\cite{U1A,U1B}. Mean-field phase diagrams were
4367: obtained by pairing the spinons and/or condensing the holons, and a
4368: pseudo-gap regime with d-wave paired spinons yet uncondensed holons was
4369: predicted\cite{U1A,U1B} -- several years before experiment (of course
4370: the d-wave nature of the superconducting state was also predicted by
4371: other approaches\cite{Scalapino}). Within this approach, the normal
4372: state at optimal doping was viewed as an incoherent gas of uncondensed
4373: holons strongly interacting with unpaired spinons. Some efforts were
4374: made to use the neglected gauge field fluctuations to stop the holons
4375: condensing at inappropriately high energy scales, with some
4376: success\cite{Bosongauge}.
4377: A serious worry is that these gauge fluctuations, ignored at the
4378: mean-field level, will almost certainly drive
4379: confinement\cite{Polyakov} (gluing the spinons and the holons
4380: together) and thus at low temperatures invalidate the assumed
4381: stability of the initial mean-field saddle point.
4382:
4383: A few early papers emphasized that spinon pairing would break the
4384: continuous $U(1)$ (or $SU(2)$) gauge symmetry down to a discrete $Z_2$
4385: subgroup\cite{U1toZ2,earlyvison2}, and this might be a way to avoid
4386: confinement. These ideas were later considerably amplified\cite{Z2},
4387: and the stability of genuinely spin-charge separated ground states
4388: established\cite{Frac1,Frac2}. Such ground states are exotic
4389: electrical insulators which support fractionalized excitations
4390: carrying separately the spin and charge of the electron. Each
4391: fragment carries a $Z_2$ gauge charge, but in contrast to the $U(1)$
4392: gauge theory saddle points, the $Z_2$ spinons are electrically neutral
4393: and do not contribute additively to the resistance. Such
4394: fractionalized insulators necessarily support an additional $Z_2$
4395: vortex-like excitation: the vison\cite{vison1}.
4396:
4397: Within the present formulation, these $Z_2$ fractionalized insulators
4398: can be readily accessed by condensing {\it pairs of
4399: vortices}\cite{NL}, rather than condensing rotons. Since
4400: vortex-{\it pairs} only have statistical interactions with charged
4401: particles and not the spinons, the resulting pair-vortex condensate is
4402: an electrical {\it insulator} with deconfined spinon, ``chargon'' and
4403: vison excitations\cite{NL,Z2} -- dramatically different from the
4404: Roton Fermi Liquid. Recent experiments on very under-doped cuprate
4405: samples have failed to find evidence of a gapped
4406: vison\cite{visexp1,visexp2}, suggesting that the pseudo-gap phase is
4407: not fractionalized. But a negative result in these vison detection
4408: experiments does not preclude the RFL phase, which supports mobile
4409: single vortices even at very low temperatures. A different approach supposing
4410: unbound and mobile single vortices is the QED$_3$ theory of
4411: Ref.~\onlinecite{tesanovic}.
4412:
4413: At the most basic level, a roton is a small pattern of swirling
4414: electrical currents. Much recent attention has focused on the
4415: possibility of a phase in the under-doped cuprates with non-vanishing
4416: orbital currents\cite{ddw}, counter rotating about elementary plaquettes
4417: on the two sub-lattices of the square lattice. Such a phase was
4418: initially encountered as a mean-field state within a slave-Fermion gauge
4419: theory approach\cite{stagflux} - the so-called ``staggered-flux phase'',
4420: but has been resurrected as the ``d-density wave'' ordered state of a
4421: Fermi liquid\cite{ddw}. In either case, such a phase within the present
4422: formulation would be described as a ``vortex-antivortex lattice'' - a
4423: checkerboard configuration on the plaquettes of the 2d square lattice.
4424: Ground states with {\it short-ranged} ``orbital antiferromagnetic''
4425: order have also been suggested recently\cite{stagshort}. Surprisingly,
4426: the Roton liquid phase also has appreciable short-ranged orbital order.
4427: A ``snapshot view'' of the orbital current correlations in the RL phase
4428: can be obtained by examining the vortex-density structure function. For
4429: simplicity, consider the charge sector of the RFL theory (the RL) in the
4430: limit of large plasmon velocity, $v_0 \rightarrow \infty$, which
4431: suppresses the longitudinal charge density fluctuations. The transverse
4432: electrical currents are then described by the Hamiltonian $H_{rot}$ in
4433: Eq.~(\ref{eq:Hrot}) with $\tilde{a}_j \rightarrow 0$. The
4434: vortex-density structure function, $S_{NN}$, can be readily obtained
4435: from this Gaussian theory:
4436: \begin{equation}
4437: S_{NN}({\bf k}) \equiv \langle |\hat{N}({\bf k})|^2 \rangle
4438: = \frac{1}{2} \sqrt{\frac{\kappa_r}{u_v}} |{\cal K}_x({\bf k})
4439: {\cal K}_y({\bf k})| \sqrt{{\cal K}^2({\bf k})} ,
4440: \end{equation}
4441: with $|{\cal K}_j({\bf k})| = 2 |\sin(k_j/2)|$ and ${\cal
4442: K}^2({\bf k}) = \sum_j |{\cal K}_j({\bf k})|^2$ as before. The
4443: vortex structure function is analytic except on the $k_x=0$ and
4444: $k_y=0$ axes and is {\it maximum} at ${\bf k} = (\pi,\pi)$,
4445: indicative of short-ranged orbital-antiferromagnetism. As such,
4446: the RFL/RL phase can perhaps be viewed as a quantum-melted
4447: staggered-flux (or vortex-antivortex) state, with only residual
4448: short-ranged orbital current correlations - the rapid motion of
4449: the rotons being responsible for the melting. If the fermions are
4450: paired, the amplitude for the basic roton hopping process which
4451: generates the structure function above vanishes upon approaching
4452: half filling, as is apparent from Eq.~(\ref{rotamp}). A
4453: preliminary analysis\cite{EBL} suggests that the further
4454: neighbor roton hopping processes which do survive at half-filling,
4455: lead to a vortex structure function which vanishes as $|{\bf k}-(\pi,\pi)|$
4456: for wavevectors near $(\pi,\pi)$.
4457:
4458: \subsection{Cuprates and the RFL phase?}
4459: \label{sec:cuprates-rfl-phase}
4460:
4461: In applying BCS theory to low temperature superconductors, one
4462: implicitly assumes that the normal state above $T_c$ is adequately
4463: described by Fermi liquid theory. Within a modern renormalization
4464: group viewpoint\cite{Shankar}, this is tantamount to presuming that
4465: the effective Hamiltonian valid below atomic energy scales (of say
4466: 10eV) is sufficiently ``close'' (in an abstract space of Hamiltonians)
4467: to the Fermi liquid ``fixed point'' Hamiltonian (actually an invariant
4468: or ``fixed'' manifold of Hamiltonians characterized by the marginal
4469: Fermi liquid parameters). In practical terms, ``close'' means that
4470: under a renormalization group transformation which scales down in
4471: energies, the renormalized Hamiltonian arrives at the Fermi liquid
4472: fixed point on energy scales which are still well above $T_c$. BCS
4473: theory then describes the universal crossover flow between the Fermi
4474: liquid fixed point (which is marginally unstable to an attractive
4475: interaction in the Cooper channel) and the superconducting fixed point
4476: which characterizes the universal low energy properties of the
4477: superconducting state well below $T_c$. Four orders of magnitude
4478: between 10eV and $T_c$ gives the RG flows plenty of ``time'' to
4479: accomplish the first step to the Fermi liquid fixed point, and is the
4480: ultimate reason behind the amazing quantitative success of BCS theory.
4481:
4482: \subsubsection{Assumptions underlying the RFL approach}
4483: \label{sec:assumpt-underly-rfl}
4484:
4485: In what follows, our working hypothesis is that the effective
4486: electron Hamiltonian on the 10eV scale appropriate for the 2d
4487: copper-oxygen planes at doping levels within and nearby the
4488: superconducting ``dome'', is ``close'' to the Roton Fermi liquid
4489: ``fixed manifold'' of Hamiltonians (characterized by Fermi/Bose
4490: liquid parameters for the quasiparticles/rotons). More specifically, we
4491: presume that when renormalized down to the scale of say one-half
4492: of an eV, the effective Hamiltonian can be well approximated by a
4493: Hamiltonian {\it on the RFL fixed manifold} - up to {\it small}
4494: perturbations. The important small perturbations are those that
4495: are {\it relevant} over appreciable portions of the fixed
4496: manifold. As established in Sections IV and V, there are three
4497: such perturbations: (i) Single vortex hopping, (ii) ``charge''
4498: hopping and (iii) attractive quasiparticle interactions in the Cooper
4499: channel. When relevant, these three processes destabilize the RFL
4500: phase and cause the RG flows to cross over to different fixed
4501: points which determine the asymptotic low temperature behavior. From
4502: our calculations, we find that at least one, and often more of these
4503: three perturbations is always relevant, regardless of the roton liquid
4504: parameters. Hence the RFL should be regarded as a {\sl critical}, and
4505: usually {\sl multi-critical} phase, rather than a stable one.
4506: For these three processes, the resulting quantum ground states
4507: are, respectively: (i) A (``confined'' and) conventional Fermi
4508: liquid phase, (ii) a conventional superconducting phase with
4509: singlet $d_{x^2-y^2}$ pairing and gapless nodal Bogoliubov
4510: quasiparticles and (iii) a ``rotonic superconductor'' with the
4511: properties of a conventional superconductor but coexisting gapless roton
4512: excitations. The rotonic superconductor, as described in
4513: Sec.~\ref{sec:selectr-bcs-inst}, has further potential instabilities
4514: driven by either single vortex hopping and ``charge'' hopping/explicit
4515: pairing. Both perturbations, if relevant, will generate an energy gap
4516: for the rotons. It seems likely that the domains of relevance of these
4517: two perturbations overlap, so that there no regime of true stability of
4518: the rotonic superconductor. The true ground state of the system in this
4519: regime is then a conventional superconductor, and its ``rotonic'' nature
4520: is evidenced only as an intermediate energy crossover.
4521:
4522:
4523: \subsubsection{Effective parameters}
4524: \label{sec:effective-parameters}
4525:
4526: In order to construct a phase diagram within this scenario, it is
4527: necessary to specify the various parameters (eg. Bose and Fermi
4528: liquid parameters) of the RFL model as a function of the doping
4529: level, $x$. Because the RFL is multi-critical, we cannot rely upon
4530: ``universality'' to validate ad-hoc requirements of smallness of
4531: perturbations, as might be the case e.g. for renormalized perturbations
4532: around a stable fixed point. Instead, we will make some assumptions
4533: based (partly) on physics. First, our main assumption is the basic
4534: validity at high energies of the roton dynamics, and of Fermi liquid-like
4535: quasiparticles. Second, we assume that superconductivity is never
4536: strong, i.e. always occurs below the rotonic/Fermi liquid energy scales.
4537: Mathematically, these two assumptions are encompassed in the
4538: inequalities
4539: \begin{eqnarray}
4540: \label{eq:inequalities}
4541: \kappa_r, u_v, t_s \gg t_v \gg t_c, \Delta_j.
4542: \end{eqnarray}
4543: Reading from left to right, this corresponds to the first and second
4544: assumptions above. In practice, we can at best hope for a factor of a
4545: few between e.g. $t_v$ and $\kappa_r$, so the ``$\gg$'' symbols above
4546: should not be taken too strongly. Although it is not important to our
4547: discussion, it is natural to assume that $v_0 \sim \kappa_r, u_v, t_s$
4548: (since Coulombic energies at the lattice scale are comparable to the
4549: electronic bandwidth). A third assumption, which is not needed for
4550: consistency of the approach, but seems desirable empirically, is that
4551: the fermion dynamics is primarily by ``spinon'' rather than electron
4552: hopping, $t_s \gg t_e$. Many of the parameters of the RFL phase can be
4553: fixed empirically from the observed behavior of the cuprates on or above
4554: the eV scale. For example, the $k-$space location of the quasiparticle
4555: Fermi surface can be chosen to coincide with the electron Fermi surface
4556: as measured via ARPES\cite{ARPES}. The value of other parameters, such
4557: as the bare velocity $v_0$ which appears in ${\cal L}_a$ and sets the
4558: scale of the plasmon velocity, can be roughly estimated from the basic
4559: electronic energy scales, and in any case does not greatly effect the
4560: relevance/irrelevance of the three important perturbations. For our
4561: basic lattice Hamiltonian introduced in Sec.~\ref{sec:model}, the two
4562: most important parameters characterizing the RFL phase are the vortex
4563: core energy, $u_v$ and the roton hopping strength, $\kappa_r$. Indeed,
4564: as shown in Sec.~\ref{sec:inst-roton-liqu}, the scaling dimensions
4565: which determine the relevance of the vortex and ``charge'' hopping
4566: perturbations at the RFL fixed point, denoted $\Delta_v$ and $\Delta_c$
4567: respectively, depended sensitively on the ratio $u_v/\kappa_r$. For
4568: example, ignoring renormalizations from the gapless fermions we found,
4569: \begin{equation}
4570: \Delta_v = {1 \over 2 \Delta_c} = \frac{1}{4\pi}
4571: \sqrt{\frac{u_v}{\kappa_r}} \frac{v_+}{\tilde{v}_0} ,
4572: \end{equation}
4573: with $v^2_+ = \tilde{v}_0^2 + {1 \over 4} u_v \kappa_r = v_0^2 + {1
4574: \over 2} u_v \kappa_r$.
4575:
4576:
4577:
4578: \begin{figure}
4579: \begin{center}
4580: \vskip-2mm
4581: \hspace*{0mm}
4582: \centerline{\fig{2.8in}{fig2.eps}}
4583: \vskip-2mm
4584: \caption{Proposed values for the vortex and ``charge'' hopping scaling
4585: dimensions, $\Delta_v(x)$ and $\Delta_c(x)$, as a function of the
4586: doping $x$. These hopping perturbations are relevant at the RFL
4587: fixed point when their $\Delta <2$. Within the doping range, $x_1 <
4588: x < x_2$, the ``charge'' hopping is {\it more} relevant, and the RFL
4589: phase is unstable to superconductivity. }
4590: \label{fig:scaledim}
4591: \end{center}
4592: \end{figure}
4593:
4594:
4595:
4596:
4597:
4598: Rather than specifying the doping dependence of $u_v, \kappa_r, v_0$
4599: and the Bose/Fermi liquid parameters, though, it is simpler and
4600: suffices for our purposes to specify the final $x-$dependence of
4601: $\Delta_v$ and $\Delta_c$. Shown in Figure 2 is a proposed form for
4602: $\Delta_{v/c}(x)$, primarily chosen to fit the gross features of
4603: the cuprate phase diagram. For example, since the RFL phase is
4604: strongly unstable to the Fermi liquid phase when $\Delta_v \ll 2$, we
4605: have taken $\Delta_v$ decreasing to small values in the strongly
4606: over-doped limit. On the other hand, to account for the observed {\it
4607: non-Fermi liquid} behavior in the normal state near {\it optimal}
4608: doping\cite{PWAbook}, requires that we take $\Delta_v \ge 2$ in
4609: that regime. On the under-doped side of the dome we can use the
4610: observed linear $x$-dependence of the superfluid density to guess
4611: the behavior of $\Delta_v$ for small $x$. In this regime the
4612: (renormalized) vortex core energy in the superconducting state is
4613: presumably tracking the transition temperature, varying linearly
4614: with $x$. It seems plausible that the ``bare'' vortex core energy
4615: $u_v$ in the lattice Hamiltonian, while perhaps significantly
4616: larger, also tracks this $x-$ dependence. This implies $\Delta_v
4617: \sim \sqrt{u_v} \sim \sqrt{x}$ as depicted in the Figure.
4618: Moreover, to recover (conventional) insulating behavior when $x
4619: \rightarrow 0$, requires that vortex condensation (rather than
4620: charge hopping) be more relevant in this limit, i.e. $\Delta_v<
4621: \Delta_c$ (see below).
4622:
4623:
4624: \subsubsection{Phase diagram}
4625: \label{sec:phase-diagram}
4626:
4627: Under the above assumptions, we now discuss in some detail the
4628: resulting phase diagram and predicted behaviors. Consider first the
4629: ground states upon varying $x$. In the extreme over-doped limit with
4630: $\Delta_v \ll 2$, vortex hopping will be a strongly relevant
4631: perturbation at the RFL fixed point. The vortices will condense at
4632: $T=0$, leading to a conventional Fermi liquid ground state. Upon
4633: decreasing $x$ there comes a special doping value ($x_2$ in Fig. 2)
4634: where $\Delta_c$ becomes {\it smaller} than $\Delta_v$. At $x=x_2$
4635: the RFL phase is unstable to {\it both} vortex and ``charge'' hopping
4636: processes, since both $\Delta_v = \Delta_c = 1/\sqrt{2} < 2$. It seems
4637: reasonable to assume that in this situation, the more strongly
4638: relevant process ultimately dominates at low energies. This implies
4639: that $x_2$ demarcates the boundary between a Fermi liquid and a
4640: superconducting ground state, as illustrated in Figure 3. Upon
4641: further decreasing $x$, the ``charge'' hopping becomes even more
4642: strongly relevant, tending to increase $T_c$ until it reaches a
4643: maximum at ``optimal doping'', denoted $x_{opt}$ in Figures 2 and 3.
4644: As one further decreases $x$, $T_c$ should start decreasing. But as
4645: shown in Sec. VA, with decreasing vortex core energy, $u_v \sim x$,
4646: the enhanced vortex density fluctuations generate an increasing
4647: antiferromagnetic exchange interaction $J \sim t_s^2 / u_v$. This
4648: {\it attractive} interaction between fermions will mediate
4649: quasiparticle pairing, with a pairing energy scale growing rapidly as $x
4650: \rightarrow 0$. It is natural to associate this ``quasiparticle pairing''
4651: temperature scale with the crossover temperature into the pseudo-gap
4652: regime\cite{PWAbook}, shown as $T^*$ in Figure 3. As discussed in
4653: Sec.~\ref{sec:selectr-bcs-inst}, under the assumption that quasiparticle
4654: kinetic energy arises primarily through spinon hopping, $t_s \gg t_e$,
4655: the superfluid density associated with the quasiparticle pairing is
4656: small, so that potential true superconductivity as a consequence of this
4657: pairing is suppressed to a low or zero temperature.
4658:
4659: Since $\Delta_v < \Delta_c$ for doping levels with $x < x_1$, vortex
4660: hopping should again dominate over ``charge'' hopping. This is the same
4661: condition which we argued leads to the Fermi liquid state for $x>x_2$
4662: above. However, the physics for small $x$ is more complex, owing to the
4663: strong antiferromagnetic interactions and proximity to the commensurate
4664: filling $x=0$ at which antiferromagnetic order is probable. In the
4665: $U(1)$ vortex-quasiparticle formalism of this paper, this difference
4666: arises from the freedom to choose (as $\Delta_j\rightarrow 0$) the
4667: fermion density to minimize the total (free) energy of the system. In
4668: the majority of this paper we have taken $n_f=\rho_0$, in order to
4669: minimize the vortex kinetic energy. However, if antiferromagnetic
4670: quasiparticle interactions are large, and $x\ll 1$, another possibility
4671: arises. To optimally benefit from the antiferromagnetic interactions,
4672: one may instead choose the fermion density commensurate, $n_f=1$, for
4673: which the Fermi surface is optimally nested and the quasiparticles can
4674: become fully gapped, gaining the maximal ``condensation energy'' from
4675: antiferromagnetic ordering. The cost of this choice is some loss of
4676: vortex kinetic energy as the vortex motion becomes somewhat frustrated
4677: by the resulting dual ``flux'' $\pi x$. Although we assume vortex energy
4678: scales are large, this flux itself is small for small $x$, so eventually
4679: as $x\rightarrow 0$ this rise in kinetic energy becomes smaller than the
4680: lowering of fermionic energy due to antiferromagnetism and such a choice
4681: becomes favorable. The ultimate physics of the remaining vortices in
4682: this limit is difficult to analyze reliably, but it
4683: seems evident that for $\Delta_v<\Delta_c$, the more relevant vortex
4684: hopping will lead to an insulating state. Since the quasiparticles are
4685: then gapped, the remaining uncompensated flux $\pi x$ is expected to
4686: lead to incommensurate charge ordering when the vortices condense -- the
4687: dual analog of the Abrikosov lattice.
4688:
4689: \begin{figure}
4690: \begin{center}
4691: \vskip-2mm
4692: \hspace*{0mm}
4693: \centerline{\fig{2.8in}{fig3.eps}}
4694: \vskip-2mm
4695: \caption{Schematic phase diagram that follows from the doping dependence
4696: of $\Delta_{v,c}$ shown in Fig. 2. With increasing doping $x$, the
4697: ground state evolves from a charge ordered insulator (I), into a
4698: superconductor (dSC) and then back into a Fermi liquid (FL). The
4699: normal state behavior near optimal doping is controlled by the
4700: Roton Fermi liquid (RFL) ground state. Below $T^*$, the
4701: quasiparticles pair. A quantum phase transition between the RFL and
4702: FL ground states occurs at $x_c$ - but is preempted by
4703: superconductivity. At $x=x_c$, a normal state resistance {\it linear}
4704: in temperature is predicted.}
4705: \label{fig:phasediag}
4706: \end{center}
4707: \end{figure}
4708:
4709: \subsubsection{The RFL Normal state}
4710:
4711: Having established the phase diagram which follows from the assumed
4712: doing dependence of $\Delta_v(x)$, we turn to a discussion of the
4713: normal state properties above $T_c$. Under our working assumption, in
4714: the energy range above $T_c$ we can ignore the small (but ultimately
4715: relevant) ``charge'' hopping perturbation, and use the RFL fixed point
4716: Hamiltonian to describe the physics of the normal state. First
4717: consider the special doping value, $x = x_c$, where $\Delta_v=2$ on
4718: the over-doped side of the superconducting dome (see Figures 2 and 3).
4719: For $x < x_c$ the vortex hopping strength $t_v$ decreases upon scaling
4720: down in energy, so that it can be treated perturbatively. As shown in
4721: Sec. VIA, at second order in the vortex hopping, the electrical
4722: conductivity is additive in the roton and quasiparticle conductivities. With
4723: impurities present the quasiparticle conductivity, $\sigma^f$, will saturate
4724: to low temperatures, whereas $\sigma_{rot} \sim T^{-\gamma}$ (with
4725: $\gamma = 2 \Delta_v -3 $) diverges as $T \rightarrow 0$. The low
4726: temperature normal state electrical resistance is thus predicted to
4727: behave as a power law,
4728: \begin{equation}
4729: R(T) \sim t_v^2 T^\gamma ,
4730: \end{equation}
4731: with a {\it vanishing residual resistivity}. Moreover, right {\it at}
4732: $x=x_c$, since $\gamma=1$ a {\it linear} temperature dependence is
4733: predicted. Notably, this transport behavior is a due to the presence
4734: of the quasi-condensate in the RFL phase, and as such is completely
4735: independent of the single particle scattering rate.
4736:
4737: The Hall conductivity, however, will be largely determined by the
4738: fermionic quasiparticle contribution, as detailed in Sec.
4739: \ref{sec:RFLproperties}. Specifically, the cotangent of the Hall
4740: angle, $\cot(\Theta_H) \equiv \rho_{xx}/\rho_{xy}$, was found to
4741: vary as $\cot(\Theta_H) \sim 1/ (T^\gamma \tau_f^2)$ in the RFL
4742: phase, with $\tau_f^{-1}$ the spinon momentum relaxation rate.
4743: Moreover, at the ``hot spots'' on the Fermi surface, $\tau_f^{-1}
4744: \sim T^{\gamma + \frac{1}{2}}$, due to scattering off the gapless
4745: rotons. Under the assumption that these same processes dominate
4746: the temperature dependence of the quasiparticle {\it transport}
4747: scattering rate, we deduce that,
4748: \begin{equation}
4749: \cot(\Theta_H) \sim { 1 \over T^\gamma \tau_f^2 } \sim T^{1
4750: +\gamma} ,
4751: \end{equation}
4752: and right at $x=x_c$, a quadratic dependence $\cot(\Theta_H) \sim
4753: T^2$. In striking contrast to conventional Drude theory which
4754: predicts $\cot(\Theta_H) \sim \tau^{-1} \sim R$ (with $\tau$ the
4755: electron's momentum relaxation time), in the RFL phase the
4756: cotangent of the Hall angle varies with a {\it different power} of
4757: temperature than for the electrical resistance, $R \sim T^\gamma$.
4758: This non-Drude behavior is consistent with the electrical
4759: transport generally observed in the optimally doped
4760: cuprates\cite{expRlin,expHall1,expHall2}, where $R \sim T$ and
4761: $\cot(\Theta_H) \sim T^2$.
4762:
4763: Consider next the thermal conductivity $\kappa$ near optimal doping
4764: within the RFL normal state. One of the most important defining
4765: characteristics of a conventional Fermi liquid is the Wiedemann-Franz
4766: law - the universal low temperature ratio of thermal and electrical
4767: conductivities, $L \equiv \kappa/\sigma T$. In a Fermi liquid,
4768: electron-like Landau quasiparticles carry both the conserved charge
4769: and the heat, and since the energy of the individual quasiparticles
4770: becomes conserved as $T \rightarrow 0$, the Lorenz ratio is universal,
4771: $L^{FL} = L_0 = \pi^2 k_B^2 / 3 e^2$. In contrast, the electrical
4772: conductivity is infinite in a superconductor, but the condensate is
4773: ineffective at carrying heat so that the Lorenz ratio vanishes,
4774: $L^{sc} = 0$. Within the RFL phase, heat can also be transported by
4775: the single fermion excitations, with a contribution to the
4776: thermal conductivity linear in temperature: $\kappa_s = L_0 \sigma^f
4777: T$. At low enough temperatures this will dominate over the phonon
4778: contribution, $\kappa_{phon} \sim T^3$. But the roton excitations,
4779: which have a quasi-one dimensional dispersion at low energies, will
4780: presumably also contribute a linear temperature dependence,
4781: $\kappa_{rot} \sim T$. Thus, the {\it total} thermal conductivity in
4782: the RFL phase is expected to vanish linearly in temperature, $\kappa
4783: \sim T$. But since the roton contribution to the {\it electrical}
4784: conductivity diverges as $T \rightarrow 0$, the RFL phase is predicted
4785: to have a vanishing Lorenz ratio:
4786: \begin{equation}
4787: L^{RFL} = \frac{\kappa}{\sigma T} \sim T^{\gamma} .
4788: \end{equation}
4789: The quasi-condensate in the RFL phase is much more effective at
4790: transporting charge than heat, much as in a superconductor. Electron
4791: doped cuprates near optimal doping, when placed in a strong magnetic
4792: field to quench the superconductivity, do exhibit a small Lorenz ratio
4793: at low temperatures\cite{Taillefer}, $L \approx L_0/5$. But
4794: extracting the zero field Lorenz ratio is problematic, since above
4795: $T_c$ the phonon contribution to $\kappa$ is non-negligible.
4796:
4797: It is instructive to consider the electrical resistance also in the {\it
4798: under-doped regime}, particularly upon cooling below the fermion
4799: pairing temperature. Above this crossover line the predicted electrical
4800: resistance varies with a power of temperature, $R \sim T^\gamma$.
4801: Since, as we assume, the fermion's kinetic energy comes primarily in the
4802: form of spinon hopping, $t\approx t_s \gg t_e$, the resulting superfluid
4803: density is however very small (see Sec.~\ref{sec:selectr-bcs-inst}), and
4804: phase coherent superconductivity does not result, at least not in this
4805: temperature range. Nevertheless, one would expect a dramatic increase
4806: in the fermion conductivity, $\sigma^f$, upon cooling through $T^*$ --
4807: much as seen in superconducting thin films upon cooling through the
4808: materials bulk transition temperature. Since the conductivity is
4809: additive in the roton and spinon contributions, $R(T)^{-1} \sim
4810: \sigma_{rot}(T)+ \sigma^f(T)$, a large and rapid increase in
4811: $\sigma^f(T)$ should be detectable as a drop in the electrical
4812: resistance relative to the ``critical'' power law form, i.e.
4813: \begin{equation}
4814: { R(T) \over T^\gamma } \sim {1 \over 1 + cT^\gamma \sigma^f(T) }
4815: .
4816: \end{equation}
4817: This behavior is generally consistent with that seen in the under-doped cuprates\cite{PWAbook,expHall2}, provided
4818: we take $\gamma \approx 1$.
4819:
4820: \subsubsection{Entering the superconductor}
4821:
4822: We finally discuss the predicted behavior upon cooling from the
4823: RFL normal state into the superconducting phase. The main change
4824: occurs in the spectrum of roton excitations, which become gapped
4825: inside the superconductor. The roton gap, $\Delta_{rot}$, should
4826: be manifest in optical measurements, since the optical
4827: conductivity will drop rapidly for low frequencies, $ \omega <
4828: \Delta_{rot}$. As in BCS theory, the ratio of the (roton) gap to
4829: the superconducting transition temperature $2 \Delta_{rot}/T_c$
4830: should be of order one. This ratio is determined by the RG
4831: crossover flow between the RFL and superconducting fixed points.
4832: It seems likely that this ratio will be {\it non-universal},
4833: depending on the marginal Bose liquid parameters characterizing
4834: the RFL fixed point (this is distinct from changes in this ratio
4835: in strong coupling Eliashberg theories, since here variations in
4836: the ratio are due to marginal parameters of the RFL fixed manifold
4837: even at arbitrarily weak coupling). Nevertheless, it would be
4838: instructive to compute this ratio for the simple near-neighbor RFL
4839: model we have been studying throughout, and to analyze the
4840: behavior of the optical conductivity above the gap.
4841:
4842: Another important consequence of gapped rotons below $T_c$, is that
4843: the electron lifetime should rapidly increase upon cooling into the
4844: superconducting phase. With reduced scattering from the rotons, the
4845: ARPES line width should narrow, most dramatically near the normal state
4846: ``hot spots'' (with tangents parallel to the $x$ or $y$ axes). This
4847: behavior is consistent with the ARPES data in the
4848: cuprates\cite{ARPES}, which upon cooling into the superconductor does
4849: show a significant narrowing of the quasiparticle peak, particularly
4850: so at the Fermi surface crossing near momentum $(\pi,0)$.
4851:
4852:
4853: Despite these preliminary encouraging similarities between various
4854: properties of the RFL phase and the cuprate phenomenology, much more
4855: work is certainly needed before one can establish whether this exotic
4856: non-Fermi liquid ground state might actually underlie the physics of
4857: the high temperature superconductors. We have already emphasized the
4858: strengths of this proposition, but there are, of course, some
4859: experimental features which seem challenging to explain from this point
4860: of view. The ``quasiparticle charge'', i.e. temperature derivative of
4861: the superfluid density $\left.\partial K_s/\partial T\right|_{T=0}$
4862: appears, based upon a small number of experimental data points, to be
4863: large and roughly independent of doping $x$, in apparent conflict with
4864: the RFL prediction. The linear temperature dependence of the
4865: electron lifetime $1/\tau_f \sim k_B T/\hbar$ observed for nodal
4866: quasiparticles near optimal doping in ARPES also does not seem natural
4867: in the RFL. However, it seems likely that it may be possible to explain
4868: a small number of such deviations from the {\sl most naive} RFL
4869: predictions by more detailed considerations. Further investigations of
4870: the RFL proposal should confront other experimental probes, such as
4871: interlayer transport and tunneling.
4872: Towards this end, it will be
4873: necessary to generalize the present approach to three-dimensions.
4874: More detailed predictions, such as for the optical conductivity upon
4875: entering the superconductor and the thermal Hall effect in the RFL
4876: normal state, might also be helpful in this regard. It
4877: would of course be most appealing to identify a {\it new ``smoking gun''}
4878: experiment for the Roton Fermi liquid state, analogous to the
4879: vison-trapping experiment\cite{vison1} for detecting 2d spin-charge
4880: separation. However, given the critical nature of the RFL, with copious
4881: gapless excitations with varieties of quantum numbers, finding such an
4882: incontrovertible experimental signature may be difficult.
4883:
4884:
4885:
4886: \begin{acknowledgments}
4887:
4888: We are grateful to Arun Paramekanti and Ashvin Vishwanath for a number
4889: of helpful conversations, and to Patrick Lee for commenting on a draft
4890: of the manuscript. We are especially indebted to T. Senthil for his
4891: many clarifying remarks, particularly pointing to the importance of
4892: ``spinon pairing'' terms in the Hamiltonian. This work was generously
4893: supported by the NSF; M.P.A.F. under Grants DMR-0210790 and
4894: PHY-9907949, and L.B. by grant DMR-9985255. L.B. also acknowledges
4895: support from the Sloan and Packard foundations.
4896:
4897:
4898: \par
4899:
4900: \end{acknowledgments}
4901:
4902:
4903:
4904:
4905:
4906:
4907:
4908:
4909:
4910:
4911:
4912:
4913: \appendix
4914:
4915:
4916:
4917:
4918: \section{Fermi Liquid Phase} \label{sec:fermiliquid}
4919:
4920: We expect the Fermi Liquid phase to occur upon complete proliferation
4921: and unbinding of vortices. To obtain the Fermi liquid in our
4922: formulation, we therefore in this appendix consider the limit of large
4923: vortex hopping $t_v \rightarrow \infty$ and small vortex energy
4924: $u_0\rightarrow 0$. We have previously demonstrated the equivalence
4925: of the U(1) gauge theory to a $Z_2$ gauge theory, and it is
4926: this latter formulation which is most convenient in this limit.
4927:
4928: To observe the Fermi liquid, we analyze the $Z_2$ gauge theory in its
4929: Hamiltonian form. Although this limit is very straightforward, based
4930: on previous work on $Z_2$ gauge theory, it is instructive to go
4931: through it in some detail here, in order to observe the effects of the
4932: spinon pairing term. The Hamiltonian density can be separated into
4933: pure gauge, charge, and spin parts, ${\cal H}={\cal H}_h+{\cal
4934: H}^{Z_2}_c+{\cal H}^{Z_2}_s$, with
4935: \begin{equation}
4936: {\cal H}_h = -h \sum_j \sigma_j^x({\bf r}),
4937: \end{equation}
4938: where $\sigma^x_j$ is the usual Pauli matrix in the space of states on
4939: link in the $j$ direction coming from site ${\bf r}$ (and hence
4940: anticommutes with $\sigma^z_j$). In the charge sector,
4941: \begin{equation}
4942: {\cal H}^{Z_2}_c = -2t_c \sum_j \sigma_j^z({\bf r}) \cos(\phi_{{\bf r} +
4943: {\bf \hat{x}}_j}-\phi_{\bf r} ) + u_c n_{\bf r}[n_{\bf r}-1],
4944: \end{equation}
4945: where $n_{\bf r}$ is the number operator conjugate to $\phi_{\bf r}$,
4946: satisfying $[\phi_{\bf r},n_{\bf r'}]=i\delta_{\bf rr'}$. For
4947: simplicity, in the spin sector we consider a local $s$-wave pair field
4948: instead of a $d$-wave one. This is not essential, but simplifies the
4949: presentation and still addresses the essential issue of the relevance
4950: of spinon pairing in the Fermi liquid. Hence,
4951: \begin{eqnarray}
4952: {\cal H}^{Z_2}_s & = & - t_s \sigma_j^z({\bf r})
4953: [ f^\dagger_{{\bf r} + {\bf \hat{x}}_j \sigma}
4954: f^{\vphantom\dagger}_{{\bf r} \sigma}+{\rm h.c.}] + \epsilon_0
4955: f^\dagger_{{\bf r}\sigma}f^{\vphantom\dagger}_{{\bf r}\sigma}
4956: \nonumber \\
4957: && +
4958: \Delta(f_{{\bf r}\uparrow}f_{{\bf r}\downarrow} + {\rm h.c.}) + u_f
4959: (f^\dagger_{{\bf r}\sigma}f^{\vphantom\dagger}_{{\bf r}\sigma})^2.
4960: \end{eqnarray}
4961: Note that we have added a local on-site energy $\epsilon_0$ and
4962: interaction $u_f$, allowed in general by symmetry. The Hamiltonian
4963: commutes with the gauge generators,
4964: \begin{equation}
4965: G_{\bf r}=(-1)^{n_{\bf r}+ f^\dagger_{{\bf
4966: r}\sigma}f^{\vphantom\dagger}_{{\bf r}\sigma}} \prod_{j}
4967: \sigma^x_j({\bf r}) \sigma^x_j({\bf r}-{\bf\hat{x}}_j).
4968: \end{equation}
4969: We require $G_{\bf r}=1$ to enforce gauge invariance.
4970:
4971: In the analysis of the large vortex hopping limit above, we obtained
4972: the action for a $Z_2$ gauge theory with zero kinetic term. This
4973: corresponds in the Hamiltonian to large $h$ (in particular we will
4974: take $h \gg t_c, t_s$. In this limit $\sigma_j^x \approx 1$, and the
4975: gauge constraint becomes
4976: \begin{equation}
4977: (-1)^{n_{\bf r}+ f^\dagger_{{\bf
4978: r}\sigma}f^{\vphantom\dagger}_{{\bf r}\sigma}} =1,
4979: \end{equation}
4980: i.e. requiring an even number of bosons and fermions on each site.
4981: Further, for large $h$, the chargon and spinon hopping terms are
4982: strongly suppressed, and can be considered perturbatively. At zeroth
4983: order in $t_c,t_s$, then, the charge and spin sectors are decoupled at
4984: each site and decoupled also from one another except by the gauge
4985: constraint. In the charge sector, for $u_c>0$, it is energetically
4986: favorable to have only either zero or one chargon per site $n_{\bf
4987: r}=0,1$. If $n_{\bf r}=0$, then we must have either zero or two
4988: spinons per site. Note than in this subspace, even for small
4989: $\Delta$, these two local spinon singlet states are non-degenerate:
4990: the energy of the two-spinon state differs from the zero-spinon state
4991: by the energy $\epsilon_0+4u_f$. With non-zero $\Delta$, one obtains
4992: as eigenstates simply two different linear combination of these two
4993: states on each site. The lower energy of the two will be realized in
4994: the ground state, and the upper energy state has no physical
4995: significance. Physically, the lower energy state, which is neutral
4996: ($n_{\bf r}=0$) and is a spin singlet, corresponds to the local
4997: vacuum, i.e. a site with no electron on it. If $n_{\bf r}=1$, then we
4998: must have one spinon on this site, which may have either spin
4999: orientation. This state has thus the quantum numbers of a physical
5000: electron. Fixing the total charge of the system $Q=\sum_{\bf r}
5001: n_{\bf r} \neq 0$ will require some number of electrons in the
5002: system. At zeroth order these are localized, but at O($t_c t_s/h$),
5003: the electrons acquire a hopping between sites, and one obtains a
5004: system clearly in a Fermi liquid phase (it is not non-interacting,
5005: since one has a hard-core constraint in the limit considered).
5006:
5007: The above considerations can be applied for zero or non-zero $\Delta$,
5008: and there are no qualitative differences in the results (the detailed
5009: nature of the vacuum state depends smoothly on $\Delta$, leading to a
5010: weak dependence of the effective electron hopping on the ratio
5011: $\Delta/u_f$) in either case. This strongly suggests that $\Delta$ is
5012: not a ``relevant'' (in the renormalization group sense) perturbation
5013: in the Fermi liquid phase. This notion can be confirmed more formally
5014: by considering the limit of very weak $\Delta \ll u_f$, in which it
5015: may be treated perturbatively. At zeroth order in this perturbation
5016: theory, the vacuum state (on a single site) is just the state with
5017: zero spinons. Formally, the perturbative relevance of $\Delta$ is
5018: determined by the behavior of the two-point function of the pair-field
5019: operator, e.g.
5020: \begin{equation}
5021: C_\Delta(\tau) = \langle f^{\vphantom\dagger}_{\bf
5022: r\uparrow}(\tau)f^{\vphantom\dagger}_{\bf
5023: r\downarrow}(\tau) f^\dagger_{\bf r\downarrow}(0)f^\dagger_{\bf
5024: r\uparrow}(0) \rangle.
5025: \end{equation}
5026: Since the pair-field operator creates a site doubly-occupied by
5027: spinons, the energy of the intermediate states encountered in the
5028: imaginary time evolution from $0$ to $\tau$ is increased by $u_f$, so
5029: that the spinon pair-field correlator decays exponentially,
5030: $C_\Delta(\tau) \sim e^{-4u_f\tau}$. This indicates that the spinon
5031: pair field is strongly irrelevant (formally with infinite scaling
5032: dimension). The importance of this observation in the context of this
5033: paper is that it provides an example in which the spinon pair field --
5034: which naively has a special significance because it alone violates
5035: spinon number conservation -- is irrelevant. This irrelevance is a
5036: consequence of strong vorticity fluctuations, which bind (confine)
5037: charge to the spinons to form electrons. Since charge is conserved,
5038: electron number {\sl must} be conserved in the resulting effective
5039: theory. Similar (but not quite so large) vorticity fluctuations in
5040: the RFL have the effect of rendering the spinon pair field irrelevant.
5041:
5042:
5043:
5044:
5045: \section{Enslaving the $Z_2$ gauge fields} \label{ap:enslaveZ2}
5046:
5047:
5048: To enslave the $Z_2$ gauge fields, we employ
5049: two sequential unitary transformations, $U_{12}=U_1 U_2$, with
5050: \begin{eqnarray}
5051: \label{eq:Z2slavers}
5052: U_1 & \!\!= &\!\!\prod_{\bf r} \left(\prod_{x'=0}^\infty \sigma_1^z({\bf
5053: r}+ x' \hat{\bf x}) \right)^{n^f_{\bf r}}\!\! \!\!=\!\! \prod_{\bf r}
5054: (\sigma_1^z({\bf r}))^{\sum_{x'=0}^\infty n^f_{{\bf r}-x'\hat{\bf
5055: x}}}, \\
5056: U_2 & \!\!= & \!\!\prod_{\sf r} \left(\prod_{x'=0}^\infty \overline{\sigma}_1^x({\sf
5057: r}+ x' \hat{\bf x}) \right)^{N_{\sf r}} \!\! \!\!= \!\!\prod_{\sf r}
5058: (\overline{\sigma}_1^x({\sf r}))^{\sum_{x'=0}^\infty N_{{\sf r}-x'\hat{\bf
5059: x}}}.
5060: \end{eqnarray}
5061: The two operators $U_1$ and $U_2$ are mutually commuting. Roughly,
5062: $U_1$ transforms to a gauge in which $\sigma_1^z=1$, and $U_2$
5063: transforms to a gauge with $\overline{\sigma}_1^x=1$. More
5064: precisely, applying the first unitary transformation, $H_{pl}$ and
5065: $H_N$ are invariant, while the fermion Hamiltonian transforms to
5066: \begin{eqnarray}
5067: \label{eq:unitary1}
5068: U_1^\dagger H^{Z_2}_{f} U_1^{\vphantom\dagger} & = &\left. H^{Z_2}_{f}
5069: \right|_{\sigma_j^z({\bf r}) \rightarrow \sigma_j^{z,{\rm
5070: slave}}({\bf r}; N)},
5071: \end{eqnarray}
5072: where
5073: \begin{eqnarray}
5074: \label{eq:sigmazslave}
5075: \sigma_1^{z,{\rm slave}}({\bf r}) & = & 1, \\
5076: \sigma_2^{z,{\rm slave}}({\bf r}) & = & \prod_{x'=0}^\infty \prod_{\Box({\bf
5077: r+w}+x'\hat{\bf x})} \sigma^z = \prod_{x'=0}^\infty (-1)^{N_{{\bf
5078: r+w}+x'\hat{\bf x}}}. \nonumber
5079: \end{eqnarray}
5080: The vortex kinetic terms also transform
5081: \begin{equation}
5082: \label{eq:unitary1a}
5083: U_1^\dagger H^{Z_2}_{kin} U_1^{\vphantom\dagger} = \left. H^{Z_2}_{kin}
5084: \right|_{\overline{\sigma}_j^x({\sf r}) \rightarrow
5085: \overline{\sigma}_j^x \overline{\sigma}_j^{x,{\rm
5086: slave}}({\sf r}; n)},
5087: \end{equation}
5088: with
5089: \begin{eqnarray}
5090: \label{eq:sigmaxslave}
5091: \overline{\sigma}_1^{x,{\rm slave}}({\sf r}) & = & 1, \\
5092: \overline{\sigma}_2^{x,{\rm slave}}({\sf r}) & = & = \prod_{x'=0}^\infty (-1)^{n^f_{{\sf
5093: r}-\overline{\bf w}-x'\hat{\bf x}}}. \nonumber
5094: \end{eqnarray}
5095: Simultaneously, the first constraint is rendered trivial
5096: \begin{equation}
5097: \label{eq:trivialz2constraint1}
5098: {\cal C}_{\bf r}^{1, {\rm slave}}= U_1^\dagger \tilde{\cal C}_{\bf r}^1
5099: U_1^{\vphantom\dagger} = \prod_{\Box({\bf r})} \overline{\sigma}^x = 1.
5100: \end{equation}
5101: Further transformation with $U_2$ removes $\overline{\sigma}^x_j$
5102: from $H^{Z_2}_{kin}$ and trivializes the remaining constraint,
5103: i.e.
5104: \begin{equation}
5105: \label{eq:trivialz2constraint2}
5106: {\cal C}_{\sf r}^{2, {\rm slave}}= U_2^\dagger \tilde{\cal C}_{\sf r}^2
5107: U_2^{\vphantom\dagger} = \prod_{\Box({\sf
5108: r})} \sigma^z = 1.
5109: \end{equation}
5110: The final transformed Hamiltonian no longer involves any dynamical
5111: gauge fields (whose state is uniquely specified by
5112: Eqs.~(\ref{eq:trivialz2constraint1},~\ref{eq:trivialz2constraint2})),
5113: and is simply given by
5114: \begin{eqnarray}
5115: \label{eq:z2slavefinal}
5116: H_{Z_2}^{\rm slave} & = & U_{12}^\dagger
5117: \tilde{H}_{Z_2} U_{12}^{\vphantom\dagger}
5118: = \left. \tilde
5119: {H}_{Z_2}\right|_{\sigma_j^\mu \rightarrow \sigma_j^{\mu,{\rm
5120: slave}}}.
5121: \end{eqnarray}
5122: The electron destruction operator in the $Z_2$ vortex-spinon
5123: theory, $\tilde{c}_{{\bf r}\sigma}= \tilde{b}_{\bf r} f_{{\bf
5124: r}\sigma}$ with $\tilde{b}_{\bf r}$ in Eq.~(\ref{btilde}),
5125: transforms upon enslaving in an identical fashion,
5126: \begin{equation}
5127: U^\dagger_{12} \tilde{c}_{{\bf r}\sigma} U_{12}^{\vphantom\dagger}
5128: = \left. \tilde{c}_{{\bf r}\sigma}\right|_{\sigma_j^z \rightarrow
5129: \sigma_j^{z,{\rm slave}}}. \label{slaveZ2electron}
5130: \end{equation}
5131:
5132:
5133:
5134: \section{Enslaving the $U(1)$ gauge theory} \label{ap:enslaveU1}
5135:
5136: As for the $Z_2$ case, to enslave the gauge fields in the $U(1)$
5137: formulation we apply two sequential unitary transformations,
5138: $U_{ab} = U_a U_b$, with
5139: \begin{equation}
5140: U_a= e^{i\sum_{{\sf r},{\sf r}^\prime} N_{\sf r} V({\sf r} - {\sf
5141: r}^\prime) (\vec{\nabla}\cdot \vec{\alpha})({\sf r}^\prime)}
5142: e^{i\sum_{{\bf r},{\bf r}^\prime} n^f_{\bf r} V({\bf r} - {\bf
5143: r}^\prime) (\vec{\nabla}\cdot \vec{\beta})({\bf r}^\prime)} ,
5144: \end{equation}
5145: where $\nabla^2 V({\bf r} - {\bf r}^\prime) = \delta_{{\bf r}{\bf
5146: r}^\prime}$ and,
5147: \begin{equation}
5148: U_b = e^{{i \over 2} \sum_{{\bf r},{\sf r}^\prime} n^f_{\bf r}
5149: \Theta({\bf r} - {\sf r}^\prime) N_{{\sf r}^\prime}} ,
5150: \end{equation}
5151: where the lattice Laplacian is
5152: \begin{equation}
5153: \label{eq:laplacian}
5154: \nabla^2 f({\bf r}) = \sum_j \partial_j^2 f({\bf r}-{\bf\hat{x}}_j)
5155: = \sum_j f({\bf r}+{\bf\hat{x}}_j) + f({\bf r}-{\bf\hat{x}}_j)
5156: -2f({\bf r}).
5157: \end{equation}
5158:
5159: Here we have introduced a ``angle'' function $\Theta({\bf r}-{\sf
5160: r}')$, to be determined later.
5161: The two unitary transformations commute with one another. Again,
5162: both $H_{pl}$ and $H_N$ are invariant commuting with $U_a$ and
5163: $U_b$, whereas under application of $U_a$ the vortex kinetic
5164: energy transforms to,
5165: \begin{equation}
5166: U_a^\dagger H_{kin} U_a^{\vphantom\dagger} = \left. H_{kin}
5167: \right|_{e^{i\alpha_j} \rightarrow
5168: e^{i\alpha^t_j} e^{i\alpha_j^{slave}(n^f)} }.
5169: \end{equation}
5170: Here, $\alpha_j^t$ is the transverse part of $\alpha_j$ and
5171: $\alpha_j^{slave}$ satisfies,
5172: \begin{equation}
5173: \epsilon_{ij} \partial_i \alpha_j^{slave}({\bf r} - {\bf w}) = \pi
5174: n^f_{\bf r} ; \hskip1cm \vec{\nabla}\cdot\vec{\alpha}^{slave} = 0
5175: .
5176: \end{equation}
5177: Similarly, the spinon and electron hopping Hamiltonians transform
5178: to,
5179: \begin{eqnarray}
5180: U_a^\dagger H_{s/e} U_a^{\vphantom\dagger} & = &\left. H_{s/e}
5181: \right|_{e^{i\beta_j} \rightarrow
5182: e^{i\beta^t_j} e^{i\beta_j^{slave}(N)}},
5183: \end{eqnarray}
5184: with $\beta_j^t$ the transverse part of $\beta_j$ and
5185: \begin{equation}
5186: \epsilon_{ij} \partial_i \beta_j^{slave}({\sf r} - {\bf w}) = \pi
5187: N_{\sf r} ; \hskip1cm \vec{\nabla}\cdot\vec{\beta}^{slave} = 0
5188: .\label{eq:betaslave}
5189: \end{equation}
5190: Notice that the longitudinal parts of $\alpha$ and $\beta$ have
5191: been eliminated, and the remaining transverse pieces commute with
5192: one another and can be treated as c-numbers. Simultaneously, the
5193: two $U(1)$ constraints are rendered trivial,
5194: \begin{equation}
5195: {\cal G}_f^{slave} = U_a^\dagger {\cal G}_f U_a^{\vphantom\dagger}
5196: = e^{-{i \over \pi} \sum_{\bf r} \Lambda_{\bf r} \epsilon_{ij}
5197: \partial_i \alpha^t_j({\bf r} - {\bf w})} = 1 ,
5198: \end{equation}
5199: \begin{equation}
5200: {\cal G}_v^{slave} = U_a^\dagger {\cal G}_v U_a^{\vphantom\dagger}
5201: = e^{{i \over \pi} \sum_{\sf r} \chi_{\sf r} \epsilon_{ij}
5202: \partial_i \beta^t_j({\sf r} - {\bf w})} = 1 ,
5203: \end{equation}
5204: implying that $\alpha_j^t=\beta^t_j =0$, and fully eliminating
5205: both gauge fields from the full transformed Hamiltonian,
5206: \begin{equation}
5207: U_a^\dagger H U_a^{\vphantom\dagger} = \left. H
5208: \right|_{e^{i\alpha_j},e^{i\beta_j} \rightarrow e^{i\alpha_j^{slave}},
5209: e^{i\beta_j^{slave}}} .
5210: \end{equation}
5211:
5212: Further transformation with $U_b$, which is essentially a
5213: non-singular gauge transformation of both the vortices and
5214: spinons, modifies the enslaved gauge fields (so that they vanish
5215: on the horizontal bonds and are integer multiples of $\pi$ on the
5216: vertical bonds) while leaving the gauge fluxes invariant.
5217: Specifically, we require
5218: \begin{eqnarray}
5219: \label{eq:thetarequire}
5220: \partial_x \Theta({\bf r}) & = &
5221: 2\tilde\beta_x({\bf r}), \qquad \forall {\bf r},\\
5222: \partial_y \Theta({\bf r}) & = & 2\tilde\beta_y({\bf r}) \qquad \forall
5223: {\bf r} \neq x{\bf\hat{x}}+{\bf\overline{w}}, x\geq 0.
5224: \end{eqnarray}
5225: Here $\tilde\beta_j({\bf r})$ is the enslaved gauge field
5226: configuration for a vortex located at ${\sf r}=0$ (i.e. determined
5227: from Eqs.~(\ref{eq:betaslave}) with $N_{\sf r}=\delta_{\sf r,0}$).
5228: The transverseness of $\tilde\beta_j$ implies then
5229: \begin{equation}
5230: \label{eq:ThetaLaplace}
5231: \nabla^2 \Theta({\bf r}+{\bf\overline{w}}) = \psi_x
5232: (\delta_{y,0}-\delta_{y,-1}),
5233: \end{equation}
5234: with an unknown function $\psi_x$ such that $\psi_x=0$ for $x<0$.
5235: Taking a line sum around the origin requires then
5236: \begin{equation}
5237: \label{eq:bcslave}
5238: \partial_y \Theta(x{\bf\hat{x}}+{\bf\overline{w}}) = -2\pi
5239: +2\tilde\beta_y(x{\bf\hat{x}}+{\bf\overline{w}}),
5240: \end{equation}
5241: for $x\geq 0$. Eqs.\ref{eq:ThetaLaplace},\ref{eq:bcslave} are the
5242: lattice analog of Laplace's equation and the condition that
5243: $\Theta$ jumps by $2\pi$ across the positive $x$-axis. These
5244: conditions determine $\psi_x$ and hence $\Theta$. After some
5245: algebra, the solution is expressible as a Fourier integral
5246: \begin{equation}
5247: \label{eq:Thetafour}
5248: \Theta({\bf r}) = \int_{\bf k} \Theta({\bf k}) e^{i{\bf k\cdot r}},
5249: \end{equation}
5250: where
5251: \begin{equation}
5252: \label{eq:Thetasol}
5253: \Theta({\bf k}) = -\frac{{2\pi} e^{i {\bf
5254: k\cdot\overline{w}}}}{{\cal K}^2 F(k_x)} \left[ \frac{{\cal
5255: K}_y^*}{{\cal K}_x^*}- {{\cal
5256: K}_x^*}{{\cal K}_y^*} I(k_x)\right],
5257: \end{equation}
5258: with
5259: \begin{eqnarray}
5260: \label{eq:integrals}
5261: F(k_x) & = & 1 - \frac{|\sin(k_x/2)|}{\sqrt{\sin^2(k_x/2)+1}}, \\
5262: I(k_x) & = & \frac{1}{4\sin(k_x/2) \sqrt{\sin^2(k_x/2)+1}}.
5263: \end{eqnarray}
5264: For large arguments, the asymptotic behavior can be obtained,
5265: \begin{equation}
5266: \Theta({\bf r}) \sim {\rm arctan}[y/x],
5267: \end{equation}
5268: for $\sqrt{x^2+y^2}\gg 1$, with the ${\rm arctan}$ defined on the
5269: interval $[0,2\pi]$. Hence $\Theta({\bf r})$ gives a proper
5270: lattice version of the continuum angle function.
5271:
5272: With this definition, one finds
5273: \begin{equation}
5274: H^{slave} = U_{ab}^\dagger H U_{ab}^{\vphantom\dagger} = \left. H
5275: \right|_{e^{i\alpha_j},e^{i\beta_j} \rightarrow \overline{\sigma}_j^{x,slave},
5276: \sigma_j^{z,slave}} ,
5277: \end{equation}
5278: with $\overline{\sigma}_j^{x,slave}({\sf r};n^f)$ and $
5279: \sigma_j^{z,slave}({\bf r};N)$ as defined in
5280: Eq.~(\ref{eq:sigmaxslave}) and Eq.~(\ref{eq:sigmazslave}),
5281: respectively. Remarkably, the enslaved $U(1)$ Hamiltonian is {\it
5282: identical} to the enslaved $Z_2$ Hamiltonian in
5283: Eq.~(\ref{eq:z2slavefinal}); $H^{slave} \equiv H_{Z_2}^{slave}$.
5284:
5285: We have thereby established the formal equivalence between the
5286: $Z_2$ and $U(1)$ formulations of the vortex-spinon field theory -
5287: the unitarily transformed enslaved versions of the original
5288: Hamiltonians $\tilde{H}_{Z_2}$ and $H$ in
5289: Eqs.~(\ref{HamZ2vs},\ref{HamU1vs}) are identical to one another.
5290: Finally, we can verify that the enslaved versions of the electron
5291: operators in the $U(1)$ and $Z_2$ formulations also coincide. From
5292: the definition of the electron operator in the $U(1)$ formulation,
5293: $c_{{\bf r}\sigma}$ in Eq.~(\ref{electronU1}), one can readily
5294: show that,
5295: \begin{equation}
5296: U_{ab}^\dagger c_{{\bf r}\sigma}
5297: U_{ab}^{\vphantom\dagger} = \left. c_{{\bf r}\sigma}
5298: \right|_{e^{i\beta_j} \rightarrow
5299: \sigma_j^{z,slave}} .
5300: \end{equation}
5301: With the analogous expression for the enslaved $Z_2$ electron
5302: operator in Eq.~(\ref{slaveZ2electron}), and upon comparing the
5303: defining expressions of the electron operators in the $Z_2$ and
5304: $U(1)$ formulations in Eqs.~(\ref{electronZ2}) and
5305: (\ref{electronU1}), respectively, one thereby establishes the
5306: desired formal equivalence: $U^\dagger_{12} \tilde{c}_{{\bf
5307: r}\sigma} U_{12} \equiv U_{ab}^\dagger c_{{\bf r}\sigma}U_{ab}$.
5308:
5309:
5310:
5311: \section{The Vortex-electron formulation} \label{ap:elecform}
5312:
5313: In this Appendix we briefly discuss a third Hamiltonian
5314: formulation of the vortex-fermion field theory. The
5315: vortex-electron Hamiltonian will be expressed in terms of
5316: ``electron operators'', or more correctly operators which create
5317: excitations having a non-vanishing overlap with the bare electron.
5318: In order to transform to this formulation, we start with the
5319: enslaved version of the $U(1)$ vortex-spinon Hamiltonian as
5320: obtained in Appendix \ref{ap:enslaveU1}:
5321: \begin{equation}
5322: U_a^\dagger H U_a^{\vphantom\dagger} = \left. H
5323: \right|_{e^{i\alpha_j},e^{i\beta_j} \rightarrow e^{i\alpha_j^{slave}},
5324: e^{i\beta_j^{slave}}} ,
5325: \end{equation}
5326: where $\alpha_j^{slave}$ and $\beta_j^{slave}$ satisfy,
5327: \begin{equation}
5328: \epsilon_{ij} \partial_i \alpha_j^{slave}({\bf r} - {\bf w}) = \pi
5329: n^f_{\bf r} ; \hskip1cm \vec{\nabla}\cdot\vec{\alpha}^{slave} = 0
5330: ,
5331: \end{equation}
5332: \begin{equation}
5333: \epsilon_{ij} \partial_i \beta_j^{slave}({\sf r} - {\bf w}) = \pi
5334: N_{\sf r} ; \hskip1cm \vec{\nabla}\cdot\vec{\beta}^{slave} = 0 .
5335: \end{equation}
5336: Now consider the unitary transformation,
5337: \begin{equation} U_{el}= e^{{i \over 2} \sum_{{\bf r},{\bf
5338: r}^\prime} n^f_{\bf r} V({\bf r} - {\bf r}^\prime) \epsilon_{ij}
5339: \partial_i e_j({\bf r}^\prime - {\bf w})} ,
5340: \end{equation}
5341: where again $\nabla^2 V({\bf r} - {\bf r}^\prime) = \delta_{{\bf
5342: r}{\bf r}^\prime}$. As is apparent from
5343: Eqs.~(\ref{SvortSphi},\ref{elecspinon}) this transformation takes
5344: one from the spinon operator to the electron operator,
5345: \begin{equation}
5346: U^\dagger_{el} f_{{\bf r}\sigma} U_{el} = {\cal S}_\phi({\bf r})
5347: f_{{\bf r}\sigma} = c_{{\bf r} \sigma} .
5348: \end{equation}
5349: Here ${\cal S}_{\phi}({\bf r}) = \prod_{\bf r}^\infty
5350: e^{-i \pi
5351: e_j^t}$ is defined in Eq.~(\ref{SvortSphi}) and the last equality
5352: follows from Eq.~(\ref{elecspinon}) in the enslaved gauge with
5353: purely transverse gauge field, $\beta^{\ell}=0$. The electrical
5354: charge density in the vortex-spinon formulation transforms to
5355: include the {\it electron} density:
5356: \begin{equation}
5357: U^\dagger_{el} \epsilon_{ij} \partial_i a_j({\bf r} - {\bf w})
5358: U_{el} = \epsilon_{ij} \partial_i a_j({\bf r} - {\bf w}) + \pi
5359: c^\dagger_{{\bf r}\sigma} c_{{\bf r}\sigma} .
5360: \end{equation}
5361:
5362:
5363: The full Hamiltonian density within the vortex-electron
5364: formulation is readily obtained from the enslaved vortex-spinon
5365: Hamiltonian: ${\cal H}_{ve} = U^\dagger_{el} U^\dagger_a {\cal H}
5366: U_a U_{el}$. It can be compactly expressed as,
5367: \begin{eqnarray}
5368: {\cal H}_{ve} &=& {u_v \over 2} \sum_j e^2_j + {v_0^2 \over 2 u_v
5369: } [ \epsilon_{ij} \partial_i a_j + c^\dagger_{{\bf r}\sigma}
5370: c_{{\bf
5371: r}\sigma} - \pi \rho_0 ]^2 \nonumber \\
5372: &+& {\cal H}_v(a_j) + {\cal H}_r(a_j) +
5373: {\cal H}_f ,
5374: \end{eqnarray}
5375: where ${\cal H}_v$ and ${\cal H}_r$ denote the vortex and roton
5376: hopping terms, respectively, and are given explicitly as,
5377: \begin{equation}
5378: {\cal H}_v(a_j) = -t_v \sum_j \cos(a_j) ,
5379: \end{equation}
5380: \begin{equation}
5381: {\cal H}_{r}(a_j) = -{\kappa_r \over 2} \sum_i \cos(\epsilon_{ij}
5382: \partial_i a_j) .
5383: \end{equation}
5384: The fermionic Hamiltonian is
5385: \begin{eqnarray}
5386: \label{eq:Hfelect}
5387: {\cal H}_f & = & - \sum_j t_e c^\dagger_{{\bf r} + {\bf
5388: \hat{x}}_j \sigma}
5389: c^{\vphantom\dagger}_{{\bf r} \sigma} \\
5390: & & \hspace{-0.5in} - \sum_j e^{i\pi \overline{e}_j ({\bf r}) } [t_s
5391: c^\dagger_{{\bf r} + {\bf \hat{x}}_j \sigma}
5392: c^{\vphantom\dagger}_{{\bf r} \sigma} + \Delta_j B_{\bf r}^\dagger c_{{\bf r}+\hat{\bf
5393: x}_j \sigma} \epsilon_{\sigma\sigma'}c_{{\bf r}\sigma'} + {\rm
5394: h.c.}] . \nonumber
5395: \end{eqnarray}
5396: The electric field appearing in the spinon hopping term yields the
5397: same physical effects as the gauge field in the $U(1)$
5398: formulation. Specifically, when the electron hops from one site to
5399: a neighboring site, the factor $e^{i\pi \overline{\epsilon}}$
5400: which shifts the gauge field $a_j$ by $\pi$, effectively hops a
5401: compensating chargon in the opposite direction. In addition, this
5402: minimal coupling form encodes the requisite minus sign when a
5403: spinon is hopped around a vortex and vice versa. The full
5404: Hamiltonian must be supplemented with the constraint that
5405: $\vec{\nabla} \cdot \vec{e} = N$, with integer $N$. We emphasize
5406: that the total spin,
5407: \begin{equation}
5408: \vec{S} = {1 \over 2} \sum_{\bf r} c^\dagger_{{\bf r}\sigma}
5409: \vec{\tau}_{\sigma \sigma^\prime} c^{\vphantom\dagger}_{{\bf
5410: r}\sigma^\prime} ,
5411: \end{equation}
5412: and electric charge,
5413: \begin{equation}
5414: Q = \sum_{\bf r} [ {1 \over \pi} \epsilon_{ij} \partial_i a_j({\bf
5415: r} - {\bf w}) + c^\dagger_{{\bf r}\sigma}
5416: c^{\vphantom\dagger}_{{\bf r}\sigma} ] ,
5417: \end{equation}
5418: are conserved, commuting with ${\cal H}_{ve}$.
5419:
5420:
5421: It is of course also possible to pass to a Euclidean path integral
5422: representation of the partition function associated with the above
5423: vortex-electron Hamiltonian. Specifically, the corresponding
5424: Euclidean Lagrangian can be readily expressed as,
5425: \begin{equation}
5426: {\cal L}_{ve} = i e_j \partial_0 a_j + c^\dagger_{\bf r}
5427: \partial_0 c^{\vphantom\dagger}_{\bf r} + ia_0(\vec{\nabla} \cdot
5428: \vec{e} - N_{\sf r}) + {\cal H}_{ve} ,
5429: \end{equation}
5430: where the time component of the gauge field $a_0({\sf r})$ lives
5431: on the sites of the dual lattice. In the partition function, the
5432: vortex number $N_{\sf r}$ is a continuous field running over the
5433: real numbers, but the integration only contributes when $N$ is
5434: integer. To see why, it is instructive to let $a_\mu \rightarrow
5435: a_\mu - \partial_\mu \theta$, and to integrate the vortex phase
5436: variable $\theta$ over the reals. Since the Hamiltonian is $2\pi$
5437: periodic in $\theta$, upon splitting the integration as
5438: $\partial_0 \theta = 2\pi \ell + \partial_0 \tilde{\theta}$ with
5439: $\tilde{\theta}=[0,2\pi]$, the summation of $exp(i2\pi \ell N)$
5440: over integer $\ell$ vanishes unless the vortex number $N$ is an
5441: integer.
5442:
5443: To obtain a more tractable representation of the Lagrangian, we
5444: introduce a Hubbard-Stratonovich field, $e_0({\bf r})$, to decouple the
5445: Coulomb interaction term above. Here $i e_0$ has the physical meaning of
5446: a dynamical electrostatic potential. In this way the full
5447: Euclidean Lagrangian can be conveniently decomposed as the sum of a
5448: bosonic charge sector and a fermionic spin and charge carrying sector,
5449: ${\cal L}_{ve} = {\cal L}_c + {\cal L}_f$. The full bosonic sector is
5450: given by,
5451: \begin{eqnarray}
5452: {\cal L}_c &=& {u_v \over 2} [e_j^2 + {1 \over {\pi^2 v_0^2}} e_0^2 ] +
5453: ie_j(\partial_0 a_j - \partial_j a_0) - \frac{i}{\pi} e_0 \epsilon_{ij}
5454: \partial_i a_j \nonumber \\
5455: &+& i (\partial_0 \theta - a_0) N + {\cal L}_v + {\cal L}_r ,
5456: \end{eqnarray}
5457: with vortex and roton hopping terms,
5458: \begin{equation}
5459: {\cal L}_v = -t_v \sum_j \cos(\partial_j \theta - a_j) ,
5460: \end{equation}
5461: \begin{equation}
5462: {\cal L}_{r} = -{\kappa_r \over 2} [ \cos(\Delta_{xy} \theta -
5463: \partial_x a_y) + (x \leftrightarrow y)] .
5464: \end{equation}
5465: The Lagrangian density in the fermionic sector is,
5466: \begin{equation}
5467: {\cal L}_f = c^\dagger_{\bf r} (\partial_0 - i e_0) c_{\bf r} +
5468: {\cal H}_f + i e_0 \rho_0 .
5469: \end{equation}
5470:
5471: In addition to the global symmetries corresponding to spin and
5472: charge conservation, the full Lagrangian has a local gauge
5473: symmetry, being invariant under,
5474: \begin{eqnarray}
5475: \label{eq:gauge1}
5476: \theta_{\sf r} & \rightarrow & \theta_{\sf r} + \Theta_{\sf r},
5477: \nonumber \\
5478: a_\mu({\sf r}) & \rightarrow & a_\mu({\sf r}) + \partial_\mu \Theta_{\sf
5479: r},
5480: \end{eqnarray}
5481: with $\Theta_{\sf r}$ an arbitrary function of space and imaginary
5482: time. Because of this gauge invariance we are free to choose an
5483: appropriate gauge.
5484:
5485:
5486:
5487: \section{Polarization tensor}
5488: \label{sec:appendix}
5489:
5490: In this appendix, we calculate the corrections to the polarization
5491: tensor $\Pi_{ij}$ at $O(t_v^2)$ including fluctuations of both
5492: $\tilde{a}$ and $\theta$, for the general case $v_0<\infty$.
5493: Integrating out all dynamical fields to $O(t_v^2)$, one finds that the
5494: effective action as a functional of $A_j$ takes the
5495: form $S_A^{\rm eff} = S_A^0 + S_A^{(2)}$, where
5496: \begin{equation}
5497: \label{eq:sa2}
5498: S_{A}^{(2)} = -\frac{t_v^2}{2}e^{S_A^0} \sum_{\sf r,r'}
5499: \int_{\tau\tau'}
5500: \bigg\langle C_i({\sf r},\tau) C_j({\sf r}^\prime,\tau^\prime) e^{-S_A} \bigg\rangle_{A=0},
5501: \end{equation}
5502: with the shorthand notation, $C_i({\sf r},\tau) = \cos(\partial_i \theta -
5503: \tilde{a}_i)_{{\sf r}\tau}$ and with
5504: the $\langle \cdot\rangle_{A=0}$ indicating a Gaussian average
5505: with respect to the RL action. This can be written as
5506: \begin{equation}
5507: \label{eq:sa2a}
5508: S_A^{(2)} = -\frac{t_v^2}{4} e^{S_A^0} \left( e^{\langle \Gamma_x^2
5509: \rangle_{A=0}} + e^{\langle \Gamma_y^2 \rangle_{A=0}}\right),
5510: \end{equation}
5511: with
5512: \begin{eqnarray}
5513: \label{eq:gammadef}
5514: \Gamma_x & = & \int_{{\bf k},\omega_n}
5515: [ (\psi_{{\bf k}\omega_n} \frac{{\cal
5516: K}_x^*}{{\cal K}} + \frac{i\omega_n {\cal K}_j^* A_j}{\pi{\cal
5517: K}}) a({\bf k},\omega_n) \nonumber \\
5518: & & - \psi_{{\bf k}\omega_n} {\cal K}_y
5519: \theta({\bf k},\omega_n) ] ,
5520: \end{eqnarray}
5521: $\psi_{{\bf k}\omega_n} = e^{i({\bf k}\cdot{\sf r}-\omega_n\tau)}
5522: - e^{i({\bf k}\cdot{\sf r'}-\omega_n\tau')}$, and $\Gamma_y$ obtained
5523: from $\Gamma_x$ by $x\leftrightarrow y$. Evaluating the
5524: expectation value gives
5525: \begin{eqnarray}
5526: \label{eq:sa2b}
5527: S_A^{(2)} & \sim & -t_v^2 \sum_{xx'y}\int_{\tau\tau'}
5528: \frac{1}{[x^2+v_{rot}^2(\tau-\tau')^2]^{\Delta_v}} \\
5529: & \times & e^{\int_{{\bf k}\omega_n} \frac{i\omega_n
5530: u_v}{\pi{\cal K}^2} \psi_{-{\bf k},-\omega_n} (\tilde{\cal
5531: K}_y G_{12} + \tilde{\cal K}_x G_{11}) \tilde{K}_j A_j({\bf
5532: -k},-\omega_n)} \nonumber \\ & & + x\leftrightarrow y. \nonumber
5533: \end{eqnarray}
5534: Since we are interested in the polarization tensor, we may expand the
5535: exponential in Eq.~(\ref{eq:sa2b}) to $O(A^2)$ to obtain $\Pi_{ij} =
5536: \Pi_{ij}^0 + \Pi_{ij}^{(2)}$, with
5537: \begin{eqnarray}
5538: \label{eq:polar2}
5539: \Pi_{ij}^{(2)} & \sim & -\frac{t_v^2\omega_n^2 u_v^2}{2{\cal K}^4}
5540: (\omega_n^2+v_{rot}^2 k_x^2)^{\Delta_v-1} \nonumber \\ & \times &
5541: |\tilde{\cal K}_y G_{12} + \tilde{\cal K}_x G_{11}|^2 \tilde{K}_i
5542: \tilde{K}_j + (x \leftrightarrow y).
5543: \end{eqnarray}
5544: In the limit of interest for the conductivity, $|{\bf k}|\rightarrow
5545: 0$ at fixed frequency, $G_{11} \gg G_{12}$, and we obtain
5546: Eq.~(\ref{eq:polar3}) of the main text.
5547:
5548:
5549:
5550:
5551: \begin{thebibliography}{99}
5552:
5553: \bibitem{PWA1} P.W. Anderson, Science, {\bf 235}, 1196 (1987).
5554:
5555: \bibitem{KRS} S. Kivelson, D.S. Rokhsar, and J. Sethna, Phys. Rev.
5556: {\bf B35}, 8865 (1987).
5557:
5558: \bibitem{SubirReview} S. Sachdev, Annals Phys. {\bf 303}, 226 (2003).
5559:
5560: \bibitem{NL} L. Balents, M.P.A. Fisher and C. Nayak,
5561: Intnl. J.
5562: Mod. Phys. B{\bf 12}, 1033 (1998); Phys. Rev. B{\bf 60}, 1654 (1999);
5563: Phys. Rev. B{\bf 61}, 6307 (2000).
5564:
5565: \bibitem{Z2} T. Senthil and Matthew P.A. Fisher,
5566: Phys. Rev. {\bf B62}, 7850 (2000).
5567:
5568: \bibitem{earlyvison1} For early papers
5569: on the topological vortex in spin-charge separated
5570: states see, N. Read and B. Chakraborty, Phys. Rev. {\bf B40}, 7133 (1989);
5571: S. Kivelson, Phys. Rev. {\bf B39}, 259 (1989).
5572:
5573: \bibitem{earlyvison2} X.G. Wen, Phys. Rev. {\bf B44}, 2664 (1991).
5574:
5575: \bibitem{vison1} T. Senthil and M.P.A. Fisher, Phys. Rev. Lett. {\bf
5576: 86}, 292 (2001).
5577:
5578: \bibitem{vison2} T. Senthil and M.P.A. Fisher, Phys. Rev. B{\bf 63}, 134521 (2001); Phys. Rev. B {\bf 64}, 214511 (2001).
5579:
5580: \bibitem{visexp1} J.C. Wynn, D.A. Bonn, B.W. Gardner, Y.J. Lin, R.X.
5581: Liang, W.N. Hardy, J.R. Kirtley, K.A. Moler, Phys. Rev. Lett. {\bf 87}, 197002 (2001).
5582:
5583: \bibitem{visexp2} D.A. Bonn, J.C. Wynn, B.W. Gardner, Y.J. Lin, R.
5584: Liang, W.N. Hardy,
5585: J.R. Kirtley and K.A. Moler, Nature {\bf 414}, 887-889 (2001)
5586:
5587: \bibitem{EBL} A. Paramekanti, L. Balents and M.P.A. Fisher, Phys.
5588: Rev. B. {\bf 66}, 4526 (2002).
5589:
5590: \bibitem{PWAbook} See ``The Theory of Superconductivity in the
5591: high-$T_c$ Cuprates'',
5592: by P.W. Anderson, Princeton University Press (Princeton N.J., 1997),
5593: and references therein.
5594:
5595: \bibitem{tesanovic} see, e.g. M. Franz, Z. Tesanovic, and O. Vafek,
5596: Phys.Rev. B{\bf 66}, 054535 (2002); I.F. Herbut, Phys.Rev. B{\bf 66},
5597: 094504 (2002).
5598:
5599: \bibitem{Duality} For a discussion of the dual vortex-field
5600: theory for 2d bosons
5601: see, M.P.A. Fisher and D.H. Lee, Phys. Rev. B{\bf 39}, 2756 (1989) and
5602: X.G. Wen and A. Zee, Int. J. Mod. Phys. B {\bf 4}, 437 (1990).
5603:
5604: \bibitem{2dXY}
5605: J. V. Jos\'e, L. P. Kadanoff, S. Kirkpatrick, and D. R. Nelson
5606: Phys. Rev. B {\bf 16}, 1217 (1977).
5607:
5608: \bibitem{BFunpub} L. Balents and M.P.A. Fisher, unpublished.
5609:
5610: \bibitem{IoffeMillis} L.B. Ioffe and A. J. Millis, Phys. Rev. {\bf B58},
5611: 11631 (1998).
5612:
5613: \bibitem{Bosonization} For various approaches to bosonization
5614: see, for example,
5615: V. Emery, in {\em Highly conducting one-dimensional solids}, edited by J.
5616: Devreese, R. Evrard, and V. Van~Doren (Plenum Press, New York, 1979), p.247;
5617: A.W.W. Ludwig, Int. Jour. Mod. Phys. {\bf B8}, 347
5618: (1994); R. Shankar, Acta Phys. Polonica B {\bf 26}, 1835 (1995);
5619: M. P. A. Fisher and L. I. Glazman, Mes. Elec. Transp.,
5620: ed. by L.L. Sohn, L.P. Kouwenhoven, and G. Schon, NATO Series E,
5621: Vol. 345, 331 (Kluwer Academic Publishing, Dordrecht, 1997).
5622:
5623:
5624:
5625: \bibitem{FQHEbook} See, ``Perspectives in Quantum Hall effects'', edited by
5626: S. Das Sarma and A. Pinczuk, Wiley (New York, NY, 1997).
5627:
5628: \bibitem{Anyon} R. B. Laughlin,
5629: Phys. Rev. Lett. {\bf 60}, 2677 (1988);
5630: Phys. Rev. Lett. {\bf 61}, 379 (1988).
5631:
5632: \bibitem{U1A} G. Baskaran, Z. Zou, and P.W. Anderson, Solid State
5633: Commun. {\bf 63}, 873 (1987);
5634: G. Baskaran and P.W. Anderson, Phys. Rev. {\bf B37}, 580 (1988).
5635:
5636: \bibitem{U1B}I. Affleck and J.B. Marston, Phys. Rev. {\bf B37}, 3774
5637: (1988); L. Ioffe and A. Larkin, Phys. Rev. {\bf B 39}, 8988 (1989);
5638: P.A. Lee, N. Nagaosa, T.-K. Ng, and X.G. Wen, Phys. Rev. {\bf B57},
5639: 6003 (1998).
5640:
5641: \bibitem{Scalapino} D.J. Scalapino, E. Loh, Jr., and J.E. Hirsch, {\bf
5642: Phys. Rev. B34}, 8190 (1986).
5643:
5644: \bibitem{Bosongauge} D. H. Kim, D. K. K. Lee, and P. A. Lee, Phys.
5645: Rev. Lett. {\bf 76}, 4801 (1996) and Phys. Rev. B {\it 55},
5646: 591-605 (1997).
5647:
5648:
5649: \bibitem{Polyakov} A. M. Polyakov, {\sl Gauge fields and Strings}
5650: (Harwood Academic Publishers, London, 1987).
5651:
5652: \bibitem{U1toZ2} N. Read and S. Sachdev, Phys. Rev. Lett. {\bf 66}, 1773 (1991);
5653: S. Sachdev and N. Read, Int. J. Mod. Phys. {\bf B5}, 219 (1991).
5654:
5655: \bibitem{Frac1} R. Moessner and S. L. Sondhi
5656: Phys. Rev. Lett. {\bf 86}, 1881 (2001); C. Nayak and K. Shtengel
5657: Phys. Rev. B {\bf 64}, 064422 (2001);
5658: L. Balents, M. P. A. Fisher, and S.M. Girvin Phys. Rev. B {\bf 65}, 224412 (2002).
5659:
5660: \bibitem{Frac2} O. I. Motrunich and T. Senthil, Phys. Rev. Lett.{\bf 89},
5661: 27704 (2002); T. Senthil and O. I. Motrunich, Phys. Rev. B {\bf 66},
5662: 205104 (2002);
5663:
5664:
5665:
5666:
5667: \bibitem{ddw} S. Chakravarty, R. B. Laughlin, D. K. Morr, and C. Nayak
5668: Phys. Rev. B {\bf 63}, 094503 (2001); S. Tewari, H.-Y. Kee, C. Nayak, and S. Chakravarty Phys. Rev. B {\bf 64}, 224516 (2001).
5669:
5670:
5671: \bibitem{stagflux}
5672: See F.C. Zhang, C. Gros, T.M. Rice and H. Shiba,
5673: Sci. Technol. {\bf 1}, 36 (1988).
5674:
5675: \bibitem{stagshort} D. A. Ivanov, P. A. Lee, and X.-G. Wen
5676: Phys. Rev. Lett. {\bf } 84, 3958 (2000).
5677:
5678: \bibitem{Shankar} R. Shankar
5679: Rev. Mod. Phys. {\bf 66}, 129 (1994).
5680:
5681: \bibitem{ARPES} See, A. Damascelli, Z-X. Shen and Z. Hussain,
5682: Rev. Mod. Phys. {\bf 75}, 473 (2003); and references therein.
5683:
5684: \bibitem{expRlin} S. Martin, A. T. Fiory, R. M. Fleming, L. F. Schneemeyer, and J. V. Waszczak, Phys. Rev. B {\bf 41}, 846 (1990); S. J. Hagen, T. W. Jing, Z. Z. Wang, J. Horvath, and N. P. Ong
5685: Phys. Rev. B {\bf 37}, 7928 (1988).
5686:
5687:
5688: \bibitem{expHall1}
5689: J. M. Harris, Y. F. Yan, and N. P. Ong,
5690: Phys. Rev. B {\bf 46}, 14293 (1992).
5691:
5692: \bibitem{expHall2} J. M. Harris, H. Wu, N. P. Ong, R. L. Meng, and C. W. Chu,
5693: Phys. Rev. B {\bf 50}, 3246 (1994).
5694:
5695:
5696: \bibitem{Taillefer} See, R.W. Hill, C. Proust, L. Taillefer, P.
5697: Fournier and R.L. Greene, Nature {\bf 414}, 711 (2001).
5698:
5699:
5700:
5701:
5702:
5703:
5704:
5705:
5706:
5707:
5708:
5709:
5710:
5711:
5712:
5713: \end{thebibliography}
5714:
5715:
5716:
5717:
5718: \end{document}
5719: