1: \documentclass[twocolumn,amsmath,amssymb,showpacs,prb]{revtex4}
2: \usepackage{graphicx}
3: \usepackage{epsf}
4: %\usepackage{showkeys}
5: \newcommand{\be}{\begin{equation}}
6: \newcommand{\ee}{\end{equation}}
7: \newcommand{\bea}{\begin{eqnarray}}
8: \newcommand{\eea}{\end{eqnarray}}
9: \newcommand{\rD}{\mbox{D}}
10: \newcommand{\rR}{{\rm R}}
11: \newcommand{\rL}{{\rm L}}
12: \newcommand{\p}{\partial}
13: \newcommand{\s}{\sigma}
14: \newcommand{\rF}{{\rm F}}
15: \newcommand{\rf}{{\rm f}}
16: \newcommand{\la}{\langle}
17: \newcommand{\ra}{\rangle}
18: \newcommand{\rd}{\mbox{d}}
19: \newcommand{\ri}{\mbox{i}}
20: \newcommand{\re}{\mbox{e}}
21: \newcommand{\rc}{{\rm c}}
22: \newcommand{\rs}{{\rm s}}
23: \newcommand{\rt}{{\rm t}}
24: \def\nn{\nonumber\\}
25: \def\np{$N_\perp$}
26: \def\up{\uparrow}
27: \def\da{\downarrow}
28: \def\r#1{(\ref{#1})}
29: \def\tp{\tilde{t}_\perp}
30: \def\tr{\rm Tr}
31: \def\z0{{\cal Z}}
32: \def\tm{\bar{t}_0}
33: \def\dt{\delta t}
34: \def\bl{{\bf l}}
35: \def\bm{{\bf m}}
36: \def\bn{{\bf n}}
37: \def\tg{\tilde{g}}
38: \def\zh{\widehat{Z}}
39: \begin{document}
40:
41: \title{A strange metal with a small Fermi surface and strong collective excitations}
42: \author{F.H.L. Essler$^1$ and A.M. Tsvelik$^2$}
43: \affiliation{$^1$ Department of Physics, University of Oxford, 1 Keble
44: Road, Oxford OX1 3NP, UK;\\
45: $^2$ Department of Physics, Brookhaven National Laboratory,
46: Upton, NY 11973-5000, USA}
47:
48:
49: \begin{abstract}
50: We develop a theory of a hybrid
51: state, where quasi-particles coexist with strong collective modes, taking as a
52: starting point a model of infinitely many one-dimensional Mott insulators
53: weakly coupled by interchain tunneling. This
54: state exists at an intermediate temperature range and undergoes an
55: antiferromagnetic phase transition at temperatures much smaller than
56: the Mott-Hubbard gap. The most peculiar feature of the hybrid state is
57: that its Fermi surface volume is unrelated to the electron density.
58: We present a self-consistent derivation of the low energy effective
59: action for our model.
60:
61: \end{abstract}
62:
63: \pacs{71.10.Pm, 72.80.Sk}
64:
65: \maketitle
66:
67: \section{Introduction}
68:
69: The presence of strong collective modes interacting with quasi-particles
70: is a distinctive feature of many strongly interacting systems such as
71: 'bad' metals, weakly doped Mott insulators (such as the cuprates) and
72: heavy fermion materials. This interaction is believed to result in a
73: variety of unusual phenomena observed in these systems such as the
74: violation of the Mott-Regel limit in the temperature dependence of the
75: electrical resistivity of bad metals or the absence of quasi-particle
76: peaks in the spectral function of the cuprates. The lack of
77: non-perturbative techniques in dimensions higher than one makes a
78: detailed theoretical description of these phenomena quite challenging.
79: One sucessful approach has been developed by Chubukov, Schmalian and Abanov,
80: who have studied the so-called spin-fermion model put forward by
81: D. Pines \cite{pines}. This model is semi-phenomenological and
82: postulates the existence of a strong, coherent, collective mode, which
83: interacts with quasi-particles located in the vicinity of a Fermi
84: surface. This model is reviewed comprehensively in
85: Refs [\onlinecite{chub2}] and has proven quite successful in explaining
86: various properties of the cuprates. However, a derivation from a
87: microscopic Hamiltonian is lacking.
88:
89: In this paper we provide a microscopic derivation of a model in the
90: same class as the spin-fermion model of Pines and Chubukov. Namely, we
91: continue to develop a theory of a hybrid state combining features of a
92: Landau Fermi liquid and a Mott insulator suggested in
93: Ref.[\onlinecite{gf2}]. This state is characterized by the coexistence
94: of well-developed collective modes with quasi-particles. The latter
95: ones have a small Fermi surface (SFS), the volume of which is
96: unrelated to the total number of electrons.
97: By definition, the Fermi surface (FS) is small if its volume is less
98: than the maximum volume allowed by Luttinger's theorem
99: \cite{Lutt,AGD,Dz,AltChub}. The existence of such a state does not
100: contradict Luttinger's theorem since the latter, contrary to popular
101: belief, does not fix the volume of FS. Instead the theorem states that
102: the electron density $n$ is proportional to the volume of phase space
103: enclosed by the surface where the single electron Green's function
104: changes its sign
105: \bea
106: n = \frac{2}{(2\pi)^d}\int_{G(\omega = 0,{\bf k}) > 0}\rd^d k \ .
107: \label{Lut}
108: \eea
109: When the Green's function has zeroes, the Fermi surface constitutes
110: only a part of this surface, namely the one where $G(0,{\bf k})
111: \rightarrow \infty$. Hence Luttinger's theorem (\ref{Lut}) does
112: not even require the existence of a Fermi surface: the Green's
113: function may only have zeroes and no poles, as it is the case
114: for superconductors \cite{AGD} and certain one-dimensional systems,
115: in which the spectral gap is generated dynamically (for the latter
116: case a general proof is outlined in Ref. [\onlinecite{ET03}]).
117:
118: A metallic state with a small FS would necessarily be associated with
119: a Green's function that has both poles and zeroes at $\omega =0$. In
120: our previous work [\onlinecite{gf2}] we suggested a model for such a
121: state based on the quasi-one-dimensional Hubbard model at half
122: filling. The transverse hopping was treated in a Random Phase
123: approximation (RPA). In order to understand the conditions of stability of
124: such an exotic metal, one has to go beyond RPA, which is the main
125: subject of the present paper.
126: Experimental indications of the existence of SFS states come from ARPES
127: measurements in underdoped cuprates \cite{arpes} and from the Hall
128: effect measurements in heavy fermion materials \cite{Sielke}.
129:
130: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
131: %\section{The Window of Opportunity}
132: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
133:
134: Before turning to the calculations, we shall give a qualitative account
135: of the physics we are after. Our starting point is
136: an {\sl ensemble of uncoupled, Mott insulating chains}. The
137: relevant energy scale is the 1D Mott gap $m$. We consider finite
138: temperatures $T$ such that $T\ll m$. The physics is purely one
139: dimensional.
140:
141: We then turn on a small {\sl long range interchain tunneling}
142: with characteristic energy scale $t_\perp$. Clearly, at zero
143: temperature the hopping between chains will be essential and
144: induce a 3D ordered state. On the other hand, in the window
145: \be
146: t_\perp\ll T,\tilde{t}_\perp({\bf k} )\ll m\ ,
147: \ee
148: we will recover the physics of 1D Mott insulating chains. Here
149: $\tilde{t}_\perp({\bf k})$ denotes the Fourier transform of the
150: interchain tunneling. Furthermore, as $T\ll m$ we may to a good
151: approximation work with zero temperature quantities in many
152: instances.
153:
154:
155: The crucial point is that while $t_\perp$ remains much smaller than the
156: Mott gap $m$, the Fourier transform $\tilde{t}_\perp({\bf k})$ can become
157: comparable to $m$ in some region of the Brillouin zone, i.e. we may
158: have a situation where
159: \be
160: t_\perp\ll T\ll \tilde{t}({\bf 0})\approx m\ .
161: \label{window}
162: \ee
163:
164: In this case an interesting {\sl ``hybrid'' state} combining 1D with
165: 3D features develops. In particular, the low energy sector corresponds
166: to a 3D metal with a small Fermi surface and quasi-particles interacting
167: with well-developed collective modes. The existence of the regime
168: (\ref{window}) is ensured by making the interchain tunneling long ranged.
169:
170: The dimensional crossover from a quasi one dimensional Mott insulator
171: to an anisotropic 3D Fermi liquid as a function of the strength
172: $t_\perp$ of the interchain hopping is sketched in Fig. \ref{fig:MIFL}.
173:
174: \begin{figure}[ht]
175: \begin{center}
176: \epsfxsize=0.45\textwidth
177: \epsfbox{MItoFL.eps}
178: \end{center}
179: \caption{Cartoon Phase Diagram for $T\ll m$ for weakly coupled 1D Mott
180: insulators, where $m$ is the 1D Mott gap. }
181: \label{fig:MIFL}
182: \end{figure}
183:
184: The purpose of the present work is to derive an effective theory for
185: the low-energy degrees of freedom in the ``1D Mott insulator/3D Fermi
186: liquid hybrid'' regime and to analyze its instabilities towards 3D
187: order at sufficiently low temperatures.
188:
189:
190: %%%%%%%%%%%%%%%%%%%%%
191: \section{The model}
192: %%%%%%%%%%%%%%%%%%%%%
193: The model we study is the Hubbard model with a strongly anisotropic hopping:
194: \bea
195: H&=&-t\sum_{n,{\bf l},\sigma}\left[{c^{\dagger}_{n,{\bf l},\sigma}}
196: c_{n+1,{\bf l},\sigma}+{\rm h.c.}\right]
197: +U\sum_{n,{\bf l}} n_{j,{\bf l},\uparrow}n_{j,{\bf l},\downarrow}\nn
198: &&+ \sum_{{\bf l},{\bf m},n,p,\sigma}t_{{\bf
199: l}{\bf m}}^{np}\ {c^{\dagger}_{n,{\bf l},\sigma}} c_{p,{\bf
200: m},\sigma}\ .
201: \label{Hamiltonian}
202: \eea
203: For definiteness we consider the chain direction to be $z$, so that
204: ${\bf l}=(l_x,l_y), {\bf m}=(m_x,m_y)$ label Hubbard chains and
205: $n,p$ label the sites along a given chain.
206:
207: As we have mentioned before, the hopping integrals in the transverse
208: directions are supposed to be
209: small in comparison to $t$. In the limit $t_{\perp} =0$
210: and at half filling the model has a Mott-Hubbard gap $m$. We work in a
211: regime where the magnitude of this gap is much smaller than the
212: bandwidth $W \approx 4t$. In our previous paper \cite{gf2}
213: the transverse hopping was treated in a Random Phase approximation
214: (RPA). In order to suppress corrections to RPA (at least in some
215: temperature interval) we assume that the transverse hopping is long
216: ranged (see below).
217:
218:
219: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
220: \subsection{Uncoupled, Mott insulating chains}
221: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
222: Let us briefly discuss the low-energy physics for uncoupled chains. In
223: order to ease notations we suppress the chain index $({\bf l})$.
224: Keeping only low-energy modes around the two Fermi points $\pm k_F$,
225: the electron annihilation operators are decomposed as
226: \be
227: c_{n,\sigma}=\sqrt{a_0}\left[\exp(ik_Fx)\ R_\sigma(x)+
228: \exp(-ik_Fx)\ L_\sigma(x)\right],
229: \ee
230: where $a_0$ is the lattice spacing, $x=ja_0$ and $k_F=\pi/2a_0$.
231: The fermionic creation operators for left and right moving Fermions
232: are bosonized, using the following conventions
233: \bea
234: L^\dagger_\sigma(x)&=&\frac{\eta_\sigma}{\sqrt{2\pi}}e^{if_\sigma\pi/4}
235: \exp\left(-\frac{i}{2}\bar{\phi}_c\right)
236: \exp\left(-\frac{if_\sigma}{2}\bar{\phi}_s\right),\nn
237: R^\dagger_\sigma(x)&=&\frac{\eta_\sigma}{\sqrt{2\pi}}e^{if_\sigma\pi/4}
238: \exp\left(\frac{i}{2}\phi_c\right)
239: \exp\left(\frac{if_\sigma}{2}\phi_s\right),
240: \eea
241: where $\eta_a$ are Klein factors with $\{\eta_a,\eta_b\}=2\delta_{ab}$ and where
242: $f_\uparrow=1$, $f_\da=-1$.The chiral boson fields $\phi_a$ and
243: $\bar{\phi}_a$ fulfil the following commutation relations
244: \be
245: [\phi_a(x),\bar{\phi}_a(y)]=2\pi i\ ,\quad a=c,s.
246: \ee
247: In terms of the chiral fields $\phi_a$ and $\bar{\phi}_a$ we define
248: canonical Bose fields $\Phi_a$ and their dual fields $\Theta_a$ by
249: \be
250: \Phi_a=\phi_a+\bar{\phi}_a\ ,\quad
251: \Theta_a=\phi_a-\bar{\phi}_a\ .
252: \ee
253: The low-energy effective Hamiltonian density for a single chain takes
254: the following bosonic form
255: \bea
256: {\cal H}_s &=& \frac{v_s}{16\pi}\left[(\p_x\Theta_s)^2 +
257: (\partial_x\Phi_s)^2\right]-g\ {\bf J}\cdot\bar{\bf J}, \label{spinboson}\nn
258: {\cal H}_c &=& \frac{v_c}{16\pi}\left[(\p_x\Theta_c)^2 +
259: (\p_x\Phi_c)^2\right] +g\ {\bf I}\cdot\bar{\bf I}.
260: \label{SGM}
261: \eea
262: Here $I^\alpha$ and $\bar{I}^\alpha$ ($J^\alpha$ and $\bar{J}^\alpha$)
263: are the chiral components of the SU(2) pseudospin (spin) currents
264: \bea
265: I^z&=&-\frac{1}{4\pi}\partial_x\phi_c\ ,\quad
266: I^+=\frac{\eta_\up\eta_\da}{2\pi}\ e^{i\phi_c}\ ,\nn
267: J^z&=&-\frac{1}{4\pi}\partial_x\phi_s\ ,\quad
268: J^+=i\frac{\eta_\up\eta_\da}{2\pi}\ e^{i\phi_s}\ .
269: \eea
270: The current-current interaction in the spin sector of (\ref{SGM}) is
271: marginally irrelevant and we will ignore it in what follows. We note
272: that doing so enhances the symmetry in the spin sector from $SU(2)$
273: (spin rotational symmetry) to $SU(2)\times SU(2)$ (rotational symmetry
274: of the left and right sectors).
275:
276: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
277: \subsubsection{Single-Particle Green's Function}
278: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
279:
280: The single-particle Green's function for the half-filled Hubbard model
281: was obtained in the framework of the formfactor approach in Refs
282: [\onlinecite{gf2,LukZam01}]. In particular, when the charge and spin
283: velocities are equal we have
284: \bea
285: G_0(\omega, \pm k_F+q) = \frac{Z_0}{\ri\omega \mp vq}\left[1 -
286: \frac{m}{\sqrt{m^2 + \omega^2 + (vq)^2}}\right]\!,\nn
287: \label{gt0}
288: \eea
289: where $Z_0 \approx 0.921862$. In order to obtain the above expression
290: for $G_0$ we took into account only processes involving the emission
291: of a single massive holon and a cascade of gapless spinons.
292:
293:
294: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
295: \subsubsection{Spin Sector}
296: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
297: The spin operators
298: $S^\alpha_n=\frac{1}{2}c^\dagger_{n,a}\sigma^\alpha_{ab}c_{n,b}$ are expressed
299: in terms of the left and right moving Fermi fields by
300: \bea
301: S^\alpha_j &\simeq& (-1)^j\frac{a_0}{2}\left[R^\dagger_a(x)\
302: \sigma^\alpha_{ab}\ L_b(x)+{\rm h.c.}\right]\nn
303: &&+\frac{a_0}{2}\left[R^\dagger_a(x)\
304: \sigma^\alpha_{ab}\ R_b(x)+R\rightarrow L\right]\nn
305: &\equiv& a_0\left[(-1)^j n^\alpha(x)+J^\alpha(x)\right].
306: \eea
307: The bosonized expressions for the staggered components of the spin
308: operators are
309: \bea
310: R^\dagger_a(x)\ \sigma_{ab}^\alpha\ L_b(x)\simeq \frac{1}{\pi i\sqrt{2a_0}}
311: \exp\left(\frac{i}{2}\Phi_c\right)\ {\rm tr}\left(g\sigma^\alpha\right),
312: \eea
313: where we have replaced the product of Klein factors by their
314: expectation value
315: \be
316: \langle \eta_\up\eta_\da\rangle= -i\ ,
317: \ee
318: and where the matrix field $g$ is expressed in terms of the spin boson
319: $\Phi_s$ and its dual field $\Theta_s$ by
320: \bea
321: g = \sqrt{\frac{a_0}{2}}\left(
322: \begin{array}{cc}
323: \exp\left(\frac{i}{2}\Phi_s\right) & i\exp\left(-\frac{i}{2}\Theta_s\right)\\
324: i\exp\left(\frac{i}{2}\Theta_s\right) & \exp\left(-\frac{i}{2}\Phi_s\right)
325: \end{array}\right).
326: \label{gfab}
327: \eea
328: At $T =0$ we have
329: \bea
330: \la g_{\alpha\beta}(\tau,x) g_{\gamma\delta}^\dagger(0,0)\ra =
331: \delta_{\alpha\delta}\delta_{\beta\gamma}\frac{a_0}{2\sqrt{v^2\tau^2 +
332: x^2}}.
333: \label{prop}
334: \eea
335: Using (\ref{gfab}) one can easily calculate multi-point correlation
336: functions of $g$.
337:
338: The action (\ref{spinboson}) describing the collective spin
339: excitations on each chain is equivalent to the SU$_1$(2)
340: Wess-Zumino-Novikov-Witten (WZNW) model once we drop the marginally
341: irrelevant interaction of spin currents. The WZWN action for the
342: matrix field $g(\tau,x)$ is given by
343: \bea
344: W[g]&=&\frac{1}{16\pi}\int d^2x\ {\rm Tr}
345: \left[\partial^\mu g^\dagger\partial_\mu g\right]\nn
346: &&+\frac{\epsilon_{\mu\nu\lambda}}{24\pi}\int_B d^3x\
347: {\rm Tr}\left[g^\dagger\partial^{\mu}g g^\dagger\partial^{\nu}g
348: g^\dagger\partial^{\lambda}g\right],
349: \label{wzw}
350: \eea
351: where $x_1=v\tau$, $x_2=x$, $\partial_\mu=\frac{\partial}{\partial x^\mu}$,
352: $B$ is a three dimensional half-space ($x_3\leq 0$)
353: whose boundary at $x_3=0$ conincides
354: with the two-dimensional $(v\tau,x)$-plane and $g$ is an arbitrary
355: extrapolation of the function defined on the two-dimensional space $x_3=0$,
356: which approaches $1$ at $x_3\to-\infty$. The action (\ref{wzw}) is
357: invariant under $SU(2)\times SU(2)$ transformations
358: $g\rightarrow Ug\widetilde{U}^\dagger$. The marginally irrelevant
359: interactions of spin currents breaks this symmetry down to the
360: diagonal spin-rotational $SU(2)$ $g\rightarrow UgU^\dagger$.
361: The form (\ref{wzw}) of the action for the spin degrees of freedom is
362: significantly more complicated than the free boson representation
363: (\ref{spinboson}). The latter is very convenient for calculations in one
364: dimension, but may be less useful when one considers interchain
365: coupling due to the fact that the dual field $\Theta_s$ is non-local
366: with respect to $\Phi_s$. The formulation in terms of the matrix field $g$
367: has the advantage of the fundamental field being the order parameter
368: itself. In fact, $W[g]$ is the Ginzburg-Landau action for a 1D
369: spin-1/2 antiferromagnet.
370:
371: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
372: \subsubsection{Three Point Function}
373: \label{3-point}
374: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
375:
376: An important ingredient in our analysis are three-point functions of
377: the form $\langle \tr[g({\bf z})\sigma^\alpha]\ R^\dagger_a({\bf {\bf
378: z_1}})\ L_b({\bf {\bf z_2}})\rangle$. The large-distance asymptotics
379: of these correlators can be evaluated by using the results of
380: \cite{LukZam01}
381: \begin{widetext}
382: \bea
383: \langle \tr[g({\bf z})\sigma^+]\ R^\dagger_\da({\bf {\bf z_1}})\
384: L_\up({\bf {\bf z_2}})\rangle
385: &=&-i\frac{\langle\eta_\da\eta_\up\rangle}{2\pi}
386: \langle \tr[g({\bf z})\sigma^+]\ e^{-\frac{i}{2}\phi_s({\bf z_1})}\
387: e^{\frac{i}{2}\bar{\phi}_s({\bf z_2})}\rangle_s
388: \langle e^{\frac{i}{2}\phi_c({\bf z_1})}\ e^{\frac{i}{2}\bar{\phi}_c({\bf z_2})}\rangle_c\nn
389: &\simeq& i\frac{\langle\eta_\da\eta_\up\rangle}{2\pi}
390: \langle \tr[g({\bf z})\sigma^+]\ e^{-\frac{i}{2}\phi_s({\bf z_1})}\
391: e^{\frac{i}{2}\bar{\phi}_s({\bf z_2})}\rangle_s
392: \frac{Z_1\sqrt{a_0}}{\pi}e^{\frac{3i\pi}{4}}\ K_0(mr_{12})\nn
393: &\simeq& \widehat{Z}\langle\eta_\da\eta_\up\rangle\
394: K_0(mr_{12})
395: \langle \tr[g({\bf z})\sigma^+]\ \tr[g({\bf z_+})\sigma^-]\rangle_s,\nn
396: \label{3point}
397: \eea
398: where ${\bf z_{1,2}}=(\tau_{1,2},x_{1,2})$, $r_{12}=|{\bf z_1}-{\bf
399: z_2}|$ and ${\bf
400: z_+}=(\frac{\tau_1+\tau_2}{2},\frac{x_1+x_2}{2})$. The constant
401: $\widehat{Z}$ is related to the normalisation $Z_0$ of the single-particle
402: Green's function by \cite{LukZam01}
403: \end{widetext}
404: \be
405: \widehat{Z}=-\frac{Z_0}{\pi^\frac{3}{2}}\sqrt{\frac{m}{va_0}}.
406: \ee
407: The calculation we have just carried out can be summarized by the
408: following approximate relations
409: \bea
410: R^\dagger_a({\bf z_1})\sigma^\alpha_{ab}L_b({\bf z_2})
411: &\longrightarrow&{i\widehat{Z}}\ K_0(mr_{12})
412: \ {\tr}[g({\bf z_+})\sigma^\alpha]\ ,\nn
413: L^\dagger_a({\bf z_1})\sigma^\alpha_{ab}R_b({\bf z_2})
414: &\longrightarrow&{i\widehat{Z}}\ K_0(mr_{12})
415: \ {\tr}[g({\bf z_+})\sigma^\alpha]\ .\nn
416: \label{single}
417: \eea
418: The approximation (\ref{single}) fails at small distances.
419: In order to remove the logarithmic singularity of $K_0$ one needs to
420: include terms corresponding to the multiple production of solitons and
421: antisolitons. At energies much smaller than the Mott gap, the fusion
422: (\ref{single}) gives rise to the spin-fermion vertex depicted on
423: Fig. \ref{vertex}.
424: \begin{figure}[ht]
425: \begin{center}
426: \epsfxsize=0.4\textwidth
427: \epsfbox{fusion3.eps}
428: \end{center}
429: \caption{The fermion-spinon interaction generated by fusion
430: (\ref{single}). }
431: \label{vertex}
432: \end{figure}
433:
434: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
435: \subsection{Long range interchain hopping}
436: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
437:
438: In order to have a small parameter in our theory we consider the
439: interchain hopping to be long-ranged, such that the Fourier transform
440: of the hopping matrix elements strongly depends on the wave
441: vector. This is a well-known trick (see, for example
442: Ref. [\onlinecite{gorkov}]) and results in a controlled ``loop''
443: expansion, where every integration over the transverse momenta leads
444: to a small factor $\kappa_0^2$ in three dimensions, where $\kappa_0$
445: is the inverse range of the interchain tunnelling. The interchain
446: hopping may be taken long ranged both along and perpendicular to the
447: chain direction. In order to simplify the calculations, we will
448: constrain our discussion to the case where
449: $t_{{\bf l}{\bf m}}^{np}=t({\bf l}-{\bf m})\delta_{n,p}$,
450: i.e. the interchain hopping has no component along the chain direction
451: and depends only on the distance between chains. The Fourier transform
452: of the interchain tunneling is then given by
453: \be
454: \tilde{t}_{\perp}({\bf k}_\perp)=\sum_{{\bf m}}
455: t_{{\bf l}{\bf m}}\ \exp(i{\bf k}_\perp\cdot (\bl-\bm)a_\perp).
456: \ee
457: In the following we choose the interchain hopping such that it respects the
458: particle-hole symmetry
459: \bea
460: c_{n,{\bf l},\sigma}\longleftrightarrow (-1)^{n+l_x+l_y}
461: c_{n,{\bf l},\sigma}^{\dagger}\ ,
462: \eea
463: which implies that
464: \be
465: \tp({\bf k}_\perp + {\bf Q}) = - \tp({\bf k}_\perp),
466: \ee
467: where
468: \be
469: {\bf Q} = \left(\frac{\pi}{a_\perp},\frac{\pi}{a_\perp}\right)
470: \ee
471: is the antiferromagnetic wave vector in the direction transverse to
472: the chains. It is straightforward to generalize our following analysis
473: to non particle-hole symmetric cases. The basic assumptions underlying
474: our model are then summarized in the following inequalities:
475: \bea
476: W \gg m\sim |\tilde{t}_{\perp}({\bf 0})|=|\tilde{t}_{\perp}({\bf Q})|
477: \gg
478: \tilde{t}_{\perp}({\bf p}_\perp)\ .
479: \label{assump}
480: \eea
481: Here $W=4t$ and $m$ are the band width and Mott gap for uncoupled chains
482: respectively and $|{\bf p}_\perp a_\perp|,|({\bf p}_\perp-{\bf Q}) a_\perp|
483: \gg \kappa_0$. The small parameter $\kappa_0$ characterizes
484: the support of $\tilde{t}_\perp({\bf k}_\perp)$ in momentum space.
485: The precise form of the momentum dependence of $\tp$ is supposedly
486: unimportant, but in order to simplify the concrete calculations we
487: shall use the following model:
488: \bea
489: \tp({\bf k}_\perp) & =&
490: -\frac{t_0}{1 + |{\bf k}_\perp a_\perp|^2
491: \kappa_0^{-2}}, ~~
492: |{\bf k}_{\perp}a_\perp| \ll 1.
493: \label{dependence}
494: \eea
495: Within the model (\ref{dependence}) the integration over the transverse
496: wave vectors may be replaced by integration over $t \equiv
497: t_{\perp}({\bf k}_\perp)$
498: \bea
499: a_\perp^2\int\frac{d^2k_\perp}{4\pi^2}
500: f(t)\approx
501: \frac{\kappa_0^2t_0}{4\pi}\int_{\frac{\kappa_0^2t_0}{4\pi^2}}^{t_0}
502: \frac{dt}{t^2}\left[f(t)+f(-t)\right].
503: \label{dos1}
504: \eea
505: Some readers may find that our approach is similar to Dynamical
506: Mean Field theory in an {\it infinitely dimensional} space. This is not
507: the case: the difference comes from the fact that in our model the
508: transverse density of states is constant on the zone boundary. This
509: feature strengthens the influence of fermionic coherent modes and
510: utterly changes the physics (see the discussion in the Conclusions).
511:
512: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
513: \section{Perturbation Theory in the interchain tunneling}
514: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
515:
516: As we have already mentioned, the perturbation theory in the
517: interchain tunneling can be reorganized in terms of a loop
518: expansion. Every integration over the transverse momenta generates a
519: small factor $\kappa_0^2$.
520: We will refer to the leading order ${\cal O}(\kappa_0^0)$ in this
521: expansion as ``Random Phase Approximation'' (RPA).
522:
523: The RPA expression for the single particle Green's function $G$ was
524: derived in Ref. [\onlinecite{gf2}] and is given by
525: \bea
526: G(\omega,\pm k_F+q, {\bf k}_\perp) =
527: \frac{G_0(\omega,\pm k_F+ q)}{1-\tilde{t}_{\perp}({\bf k}_\perp)
528: G_0(\omega,\pm k_F+q)}\ .
529: \label{G2}
530: \eea
531: Here $G_0$ is the single-particle Green's function for an individual
532: chain (\ref{gt0}). In a purely one-dimensional Mott insulator the electron
533: is a composite particle and as a result the spectral function is
534: incoherent. Coherent electronic excitations reappear as soon as the
535: interchain tunneling is turned on. They can be understood as
536: antiholon-spinon bound states and occur at energies {\sl below} the
537: Mott gap. When $t_{\perp}$ exceeds a certain critical value, the
538: dispersion of the coherent mode crosses the chemical potential and a
539: Fermi surface is formed. As a consequence of particle-hole symmetry,
540: at half-filling this Fermi surface consists of electron- and hole-like
541: pockets of equal volume. A sketch of such a Fermi surface is shown in
542: Fig. \ref{fig:surface}.
543: \begin{figure}[ht]
544: \begin{center}
545: \epsfxsize=0.4\textwidth
546: \epsfbox{surface.eps}
547: \end{center}
548: \caption {The Brillouin zone with the electron (red ovals) and
549: hole-like (blue semi-ovals) Fermi pockets of a two-dimensional
550: lattice. The noninteracting Fermi surface is shown as a dashed line.}
551: \label{fig:surface}
552: \end{figure}
553: A convenient measure for the strength of the interchain coupling is
554: given by the quantity
555: \be
556: {\cal Z}\equiv\frac{Z_0t_0}{m}\ ,
557: \label{Z}
558: \ee
559: where $t_0$ is defined in (\ref{dependence}). The RPA form
560: (\ref{G2}) for the
561: Green's function features a pole corresponding to a coherent
562: quasi-particle mode. This mode crosses the chemical potential when
563: ${\cal Z}$ exceeds the critical value
564: \be
565: {\cal Z}_c=3.33019\ldots\ ,
566: \label{Zc}
567: \ee
568: and a Fermi surface is present for all ${\cal Z}>{\cal Z}_c$.
569:
570: Having in hand the expression for the chain single-particle Green's
571: function, we may use it to define a dressed interchain hopping
572: $\tilde{T}_{R,L}(\omega,q,{\bf k}_\perp)$ by summing the diagrams shown in
573: Fig. \ref{fig:TR}.
574: \begin{figure}[ht]
575: \begin{center}
576: \epsfxsize=0.4\textwidth
577: \epsfbox{dressedT.eps}
578: \end{center}
579: \caption{The dressed interchain hopping.}
580: \label{fig:TR}
581: \end{figure}
582: This results in
583: \bea
584: T_{R,L}(\omega,q,{\bf k}_\perp)&=&\frac{\tilde{t}_\perp({\bf
585: k}_\perp)}{1-{\tilde t}({\bf k}_\perp)\ G_0(\omega,\pm k_F+q)}\ ,
586: \eea
587: where the $+$ sign corresponds to $R$ and the $-$ sign to $L$
588: respectively. We note that the dressed interchain hopping is equal
589: to the propagator of the Hubbard-Stratonovich field that can be
590: introduced to decouple the interchain hopping \cite{Boies}.
591:
592: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
593: \subsection{The Spin Sector in RPA}
594: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
595:
596: In the RPA, the spin sector remains one-dimensional and critical. This
597: can be seen as follows. Let us consider the real-space correlator
598: between the staggered components of spins on different chains ${\bf
599: l}$ and ${\bf m}$
600: \bea
601: \langle n^{+}_{j,\bl}(t)\ n^{-}_{1,\bm}(0)\rangle.
602: \eea
603: Within perturbation theory in the interchain hopping, we need at least
604: one right moving and left moving fermion operator each on chains $l$
605: and $m$ in order to obtain a nonzero expectation value in the spin sector.
606: The only ways to achieve this are shown in Fig.\ref{fig:spindiag}.
607: Here the dashed lines denote the bare interchain hopping, ellipses
608: enclosing two black (white) circles represent the purely 1D Green's
609: function of right (left) moving electrons on a given chain and the
610: ellipses enclosing two circles and a hexagon stand for the three-point
611: function (\ref{3point}).
612: Clearly all such diagrams involve at least one integration over the
613: transverse momentum. Hence, within the RPA spin-spin correlation
614: functions remain entirely one-dimensional and spins on different
615: chains remain uncorrelated.
616: \begin{figure}[ht]
617: \begin{center}
618: \epsfxsize=0.45\textwidth
619: \epsfbox{spindiag.eps}
620: \end{center}
621: \caption{Real-space diagrams that contribute to the two-point function of
622: staggered magnetizations between chains $(l)$ and $(m)$.}
623: \label{fig:spindiag}
624: \end{figure}
625:
626: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
627: \subsection{Excitation Spectrum in RPA}
628: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
629: From the discussion above it is clear that for sufficiently strong
630: interchain hopping ${\cal Z}>{\cal Z}_c$ the RPA leads to two types of
631: gapless excitations
632: \begin{itemize}
633: \item{}Fermionic particle and hole excitations over the Fermi surface
634: with anisotropic 3D dispersions.
635: \item{}Collective excitations of the spin degrees of freedom. These
636: are of a purely 1D nature and do not have a dispersion in
637: the direction transverse to the chains.
638: \end{itemize}
639: If one goes beyond the RPA, interactions between these two types of
640: excitations will be generated. In the following we determine the form
641: of these interactions and study their effects. To go beyond the
642: RPA in principle requires the knowledge of the two-particle Green's
643: function of uncoupled Mott-insulating chains. However, if one
644: restricts one's attention to the regime of energies small compared to
645: the Mott gap, the three-point function \r{3point} (which corresponds to a
646: particular limit of the two-particle Green's function) suffices.
647: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
648: \section{Interchain Exchange and Estimate of the Transition
649: Temperature}
650: \label{sec:exch}
651: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
652: Although it is obvious that corrections to RPA are of higher order in
653: the small parameter $\kappa_0$, they will diverge at small
654: temperatures. Therefore RPA works only at finite temperatures and for
655: its consistency the transition temperature (below which the system is
656: three-dimensionally ordered) must be much smaller than the
657: Mott-Hubbard gap $m$. It is therefore important to estimate the
658: corresponding corrections to RPA and their temperature dependence.
659: As the first step in taking into account corrections to RPA we have
660: to estimate the interchain RKKY interaction. As we have shown in
661: section \r{3-point}, there is a three-point ``vertex'' that couples
662: the spin degrees of freedom to the fermionic quasi-particles. In second
663: order perturbation theory in this interaction, an interchain exchange
664: interaction between the spin degrees of freedom is generated. The
665: corresponding action is given by
666: \bea
667: S_{\rm xc}&=&\int\prod_{j=1}^2 d\tau_j dx_j\sum_{\bl\neq \bm}
668: J_{\bl\bm}(x_1-x_2,\tau_1-\tau_2)\nn
669: &&\qquad\times\ \tr\left[g_\bl(\tau_1,x_1)\
670: g^{\dagger}_\bm(\tau_2,x_2)\right].
671: \label{Jint}
672: \eea
673: The Fourier transform of the leading order (in $\kappa_0$) exchange
674: matrix element is given by the ``bubble'' diagram shown in
675: Fig. \ref{fig:J}, where the doubles lines are the dressed interchain
676: hoppings for left and right moving fermions and the squares denote
677: the elements of the matrix field.
678: \begin{figure}[ht]
679: \begin{center}
680: \epsfxsize=0.25\textwidth
681: \epsfbox{exchange2.eps}
682: \end{center}
683: \caption{Leading order ${\cal O}(\kappa_0^2)$ contribution to the
684: interchain exchange.}
685: \label{fig:J}
686: \end{figure}
687: The result is
688: \begin{widetext}
689: \bea
690: \tilde{J}(\omega,q,{\bf q_\perp})&=&\widehat{Z}^2
691: \int\frac{d\omega^\prime dq^\prime}{(2\pi)^2}
692: a_\perp^2\int{d^2k_\perp}{}
693: \left[\frac{v}{m^2+\omega'^2+(vq')^2}\right]^2\
694: {T}_R\left(\omega^\prime+\frac{\omega}{2},q'+\frac{q}{2},{\bf
695: k_\perp}\right)\nn
696: &&\qquad\qquad\qquad\times {T}_L\left(\omega^\prime-\frac{\omega}{2},q'-\frac{q}{2},{\bf
697: q_\perp-k_\perp}\right)\Bigr].
698: \label{xch}
699: \eea
700: \end{widetext}
701: To make the calculations easier, we use the $k$-dependence
702: (\ref{dependence}) so that the sum over the transverse wave
703: vectors is replaced by integration over $t$ according to (\ref{dos1}).
704: Since $\tilde{t}({\bf k}_\perp)$ is peaked near zero and ${\bf Q}$, there
705: are two interesting wave vectors: ${\bf q_\perp}=0$ and ${\bf
706: q}_{\perp} = {\bf Q}$.
707: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
708: \subsection{Case 1: ${\cal Z}<{\cal Z}_c$}
709: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
710: In this case the coherent electron modes still have a gap and no Fermi
711: surface is formed in the RPA. Using (\ref{dos1})
712: to carry out the summation over
713: transverse momenta we obtain
714: \bea
715: \tilde{J}(0,0,{\bf 0}) \approx{\cal C}_1
716: \int_0^\infty ds\ s\frac{\arctan(\z0 G(s))}{(1+s^2)^2\ G(s)}\equiv
717: \alpha_0{\cal C}_1,
718: \label{J0}
719: \eea
720: where ${\cal Z}=Z_0t_0/m$ is defined in (\ref{Z}) and
721: \be
722: {\cal C}_1=\zh^2{\kappa_0^2t_0}{}
723: \frac{v}{mZ_0}=\frac{{\cal Z}\kappa_0^2}{\pi^3a_0}m\ ,
724: \ee
725: and
726: \bea
727: G(s) = s^{-1}\left[1 - \frac{1}{\sqrt{s^2 + 1}}\right].
728: \eea
729: As expected, this interaction is of order of $\kappa_0^2 t_0$. The
730: numerical factor $\alpha_0$ ranges between $0$ for ${\cal Z}=0$ and
731: $2.81$ for ${\cal Z}\to 3.33019$. The exchange at momentum transfer
732: ${\bf Q}$ is
733: \bea
734: \tilde{J}(0,0,{\bf Q}) &\approx&-\frac{{\cal C}_1}{2}
735: \int_0^\infty ds\ s
736: \frac{\ln\left[\frac{1+\z0 G(s)}{1-\z0 G(s)}\right]}{(1+s^2)^2\
737: G(s)}\nn
738: &=&-\alpha_Q{\cal C}_1,
739: \label{JQ}
740: \eea
741: where $\alpha_Q$ varies between $0$ for for ${\cal Z}=0$ and
742: $3.07$ for ${\cal Z}\to {\cal Z}_c$.
743:
744: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
745: \subsection{Case 2: ${\cal Z}>{\cal Z}_c$}
746: \label{ssec:case2}
747: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
748: In this case a Fermi surface in the form of electron and hole pockets
749: is present. The presence of zero energy modes does not really affect
750: the exchange at zero momentum transfer, which is given by
751: \bea
752: \tilde{J}(0,0,{\bf 0}) \approx {\cal C}_1
753: \int_0^\infty ds\ \frac{sf(s)}{(1+s^2)^2\ G(s)}
754: =\alpha'_0{\cal C}_1\ ,
755: \label{J02}
756: \eea
757: where
758: \bea
759: f(s)&=&2\arctan(\xi(s)\z0 G(s))-\arctan(\z0 G(s))\ ,\nn
760: \xi(s)&=&{\rm min}(1,[{\cal Z}G(s)]^{-1})\ .
761: \eea
762: We find that $\alpha'_0$ starts at $1.405$ for $\z0\to 3.33019$, then
763: goes through a maximum of approximately $1.48$ around $\z0\approx 4.18$
764: and then diminishes slowly. Hence the exchange at wave number zero
765: play a subdominant role.
766:
767: Let us now turn to the exchange at wavenumber ${\bf Q}$. We find that
768: there is a logarithmic divergence in (\ref{xch}), which is related
769: to the Fermi surface formation and will be discussed later in more
770: detail. We regularize the divergence by temperature. This may be done
771: by replacing the $\omega'$-integral in (\ref{xch}) by a sum over Matsubara
772: frequencies and substituting the finite-temperature Green's function
773: for their $T=0$ analogs in the dressed interchain hoppings
774: $\tilde{T}_{R,L}$ in (\ref{xch}). At low temperatures the single particle
775: Green's function is given by \cite{ET03}
776: \bea
777: &&G_R(\omega_n,q)=\int_{-\infty}^\infty dx\frac{A_R(x,q)}{i\omega_n-x}\
778: ,\nn
779: &&A_R(\omega,q)=\frac{Z_0}{4\pi^2 m}
780: \sqrt{\frac{2m}{T}}\left[\frac{m}{|\omega-vq|}\right]^{\frac{3}{2}}\nn
781: &&\quad\times {\rm Re}\left[\sqrt{-2i}
782: B\left(\frac{1}{4}-\frac{i}{2}\frac{\omega^2-v^2q^2-m^2}{2\pi
783: T|\omega-vq|},\frac{1}{2}\right)\right].
784: \label{finitetgf}
785: \eea
786: The singular piece of $\tilde{J}^{\rm sing}(0,0,{\bf Q})$
787: diverges logarithmically with temperature and is estimated as
788: \bea
789: \tilde{J}^{\rm sing}(0,0,{\bf Q}) \approx -{\cal C}_1\
790: \ln\left[\frac{m}{T}\right] \int_{s_-}^{s_+}ds\
791: \frac{s[G(s)]^{-1}}{(1+s^2)^2} ,
792: \label{jsing}
793: \eea
794: where $s_\pm$ are solutions to the equation
795: \bea
796: 1-\z0 G(s_\pm)=0\ .
797: \eea
798: In order to establish the exchange at nonzero values of $\omega$ and
799: $q$ we have calculated $\tilde{J}(\omega,q,{\bf Q})$ numerically at
800: small temperatures. Rather than using the finite-$T$ Green's function
801: (\ref{finitetgf}) we work with the $T=0$ expression (\ref{gt0}) and
802: replace $\omega$ by the discrete Matsubara frequencies. For small
803: temperatures this is a reasonable approximation. We find that
804: $|\tilde{J}(\omega,q,{\bf Q})|$ is largest at $\omega=0=q$.
805:
806: In addition to the singular piece (\ref{jsing}) there also is a
807: regular contribution to the exchange. As long as we are close to the
808: transition, i.e.
809: \be
810: \frac{{\cal Z}-{\cal Z}_c}{{\cal Z}_c}\ll 1\ ,
811: \label{region}
812: \ee
813: we may estimate the regular contribution to the exchange by its value
814: at the critical strength ${\cal Z}_c$ of the interchain tunneling. The
815: latter is given by (\ref{JQ})
816: \bea
817: \tilde{J}^{\rm reg}(0,0,{\bf Q}) \approx -3.071\ {\cal C}_1.
818: \label{jreg}
819: \eea
820: In the regime (\ref{region})
821: \be
822: \tilde{J}^{\rm sing}(0,0,{\bf Q}) \approx -1.318\
823: \sqrt{{\cal Z}-{\cal Z}_c}\
824: {\cal C}_1\ \ln\left[\frac{m}{T}\right] .
825: \label{Jsing}
826: \ee
827: The total exchange constant at wave number ${\bf Q}$ is then estimated
828: as
829: \bea
830: \tilde{J}(0,0,{\bf Q}) \approx -{\cal C}_1\left[3.071+
831: 1.318\
832: \sqrt{{\cal Z}-{\cal Z}_c}\ \ln\left[\frac{m}{T}\right]\right].
833: \label{jtot}
834: \eea
835:
836: Having determined the exchange constant we are now in a position to
837: estimate the temperature at which a magnetic instability develops.
838: In the absence of interchain hopping the correlation functions of the
839: matrix field at $T=0$ are given by (\ref{prop}).
840: At $T>0$ we have
841: \bea
842: \langle g_{\alpha\beta}(\tau,x) g^+_{\gamma\delta}(0,0)\rangle =
843: \delta_{\alpha\delta}\delta_{\beta\gamma}
844: \frac{\pi Ta_0/v}{2|\sinh\left(\frac{\pi T}{v}[x+iv\tau]\right)|},
845: \label{correlT}
846: \eea
847: if we neglect the marginally irrelevant current-current interaction in
848: the spin-sector of the Hamiltonian (\ref{SGM}) describing the 1D Mott
849: insulating chains. If one takes it into account in renormalization group
850: improved perturbation theory one obtains \cite{Barzykin00a}
851: \be
852: \langle {\rm tr}(g\sigma^\alpha)\ {\rm
853: tr}(g^\dagger\sigma^\beta)\rangle=
854: \delta_{\alpha\beta}
855: \frac{\sqrt{\ln\left[\frac{\Lambda}{T}\right]}\ \pi Ta_0v^{-1}}
856: {|\sinh\left(\frac{\pi T}{v}[x+iv\tau]\right)|},
857: \ee
858: where $\Lambda$ is a high-energy cutoff, which we may take to be of
859: the order of the hopping integral along the chain direction.
860: Carrying out an analogous calculation for the dimerization operator we find
861: \be
862: \langle {\rm tr}(g)\ {\rm tr}(g^\dagger)\rangle=
863: \delta_{\alpha\beta}
864: \frac{\left[\ln\left[\frac{\Lambda}{T}\right]\right]^{-\frac{3}{2}}\ \pi Ta_0v^{-1}}
865: {|\sinh\left(\frac{\pi T}{v}[x+iv\tau]\right)|}.
866: \label{dimer}
867: \ee
868: Upon Fourier transformation and analytical contibuation one finds the
869: following result for the dynamical magnetic susceptibility of the
870: uncoupled chain system \cite{sb83}
871: \bea
872: \chi_{\rm 1d}(\omega, q) &=& -
873: \frac{a_0\sqrt{\ln\left[\frac{\Lambda}{T}\right]}
874: }{2T}
875: \frac{\Gamma\left[\frac{1}{4}- i\frac{\omega - vq}{4\pi T}\right]
876: \Gamma\left[\frac{1}{4}- i\frac{\omega + vq}{4\pi T}\right]}
877: {\Gamma\left[\frac{3}{4}- i\frac{\omega - vq}{4\pi T}\right]
878: \Gamma\left[\frac{3}{4}- i\frac{\omega + vq}{4\pi T}\right]}.\nn
879: \label{susco}
880: \eea
881: The dimerization susceptibility is equal to (\ref{susco}) apart from the
882: prefactor, in which $\sqrt{\ln[\Lambda/T]}$ is
883: replaced by $\left[\ln\left[\Lambda/T\right]\right]^{-3/2}$. Hence the
884: staggered spin susceptibility is always more singular than the
885: dimerization susceptibility and as a result the dominant instability
886: of the spin sector is towards N\'eel order. The enhancement of the
887: spin susceptibility as compared to the dimerization susceptibility is
888: caused by the leading irrelevant operator in the Hamiltonian, namely
889: the interaction of spin currents. If we were to add an interaction to
890: the underlying lattice Hamiltonian in order to eliminate this
891: interaction, the symmetry between dimerization and the staggered
892: components of the spins would be broken by some other irrelevant operator.
893: The the dynamical susceptibility of the coupled chains system can be
894: determined by a expansion in the interchain coupling of the type
895: discussed in \cite{marc,katanin}. The leading term is given by the
896: Random Phase Approximation
897: \bea
898: \chi_{\rm 3d}(\omega, q,{\bf p}_\perp) &=& \frac{\chi_{\rm 1d}(\omega,q)}
899: {1-2\tilde{J}(\omega,q,{\bf p}_\perp)\chi_{\rm 1d}(\omega,q)}.
900: \label{chiRPA}
901: \eea
902: Given the expression (\ref{chiRPA}) for the dynamical susceptibility we
903: may obtain an estimate for the transition temperature $T_c$ below
904: which three dimensional magnetic long range order develops. $T_c$ is
905: defined as the temperature at which a zero frequency pole develops in
906: $\chi_{\rm 3d}$. Given that $\chi_{\rm 1d}(0,q)$ is peaked at $q=0$ and
907: $\tilde{J}$ is peaked at $q=0$ and ${\bf p}={\bf Q}$ we obtain the
908: following condition fixing $T_c$
909: \bea
910: 1 - 2\tilde{J}(0,0,{\bf Q})\chi_{\rm 1d}(0,0) =0.
911: \eea
912: Replacing $\tilde{J}(0,0,{\bf Q})$ by (\ref{jtot}) we arrive at the
913: following equation determining the transition temperature $T_c$
914: \bea
915: \frac{T_c}{m}&\approx& 2.887\kappa_0^2
916: \sqrt{\ln\left[\frac{\Lambda}{T_c}\right]}\nn
917: &&\times \left[1+0.429\sqrt{{\cal Z}-{\cal Z}_c}\
918: \ln\left[\frac{m}{T_c}\right]\right].
919: \label{tc1}
920: \eea
921: Let us consider the two limiting cases in which either the regular
922: (\ref{jreg}) or the singular part (\ref{Jsing}) of the exchange dominates and
923: drives the transition. The first case occurs if we are very close to
924: the point where the Fermi pockets are first formed and ${\cal Z}-{\cal
925: Z}_c\ll(\ln(\kappa_0^2))^{-2}$. Then the transition temperature is
926: roughly equal to
927: \be
928: \frac{T_c}{m}\approx 2.887\ \kappa_0^2
929: \sqrt{\ln\left[\frac{\Lambda}{m\kappa_0^2}\right]}.
930: \ee
931: The second case occurs if ${\cal Z}-{\cal
932: Z}_c\gg(\ln(\kappa_0^2))^{-2}$ and then
933: \be
934: \frac{T_c}{m}\approx 1.239\ \kappa_0^2\delta
935: \sqrt{\ln\left[\frac{\Lambda}{m\kappa_0^2\delta}\right]}
936: \ln\left[\frac{1}{\kappa_0^2\delta}\right],
937: \label{tc2}
938: \ee
939: where $\delta=\sqrt{{\cal Z}-{\cal Z}_c}$.
940:
941:
942: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
943: \section{Effective Theory at Low Energies; the Residual Interactions}
944: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
945:
946: Now we are in position of writing the low-energy effective action for
947: the metallic state. This effective action describes interactions of the
948: low-energy modes i.e. the coherent fermions and the order parameter
949: field $g_{ab}$. In the previous section we calculated the interchain
950: coupling for the $g$-field. It contains a part coming from states far
951: from the chemical potential and the part with logarithmic divergences
952: coming from the states close to the Fermi surface. We can isolate the
953: first piece and include it into the effective action of $g$
954: \bea
955: S_{\rm sp} &=& \sum_{\bf n} W[g_{\bf n}]\nn
956: &+&\sum_{\bm,\bl}J_{\bm\bl}\int\! d\tau\ \! dx
957: \tr\left[g_{\bm}(\tau,x)g_\bl^{\dagger}(\tau,x)\right],
958: \label{spin}
959: \eea
960: where to first approximation
961: \bea
962: J_{\bn\bl} \approx \frac{Z_0^2}{\pi^2 ma_0}t^2_{\bn\bl}\ .
963: \eea
964: This part of the action plays the role of the sigma model in the
965: spin-fermion model by Pines and Chubukov (see \cite{chub2} and
966: references therein). Taken in isolation, this model has an instability
967: at some temperature $T_c$ which can be estimated from the RPA
968: expression for the dynamical magnetic susceptibility in complete
969: analogy with our calculation in section \ref{ssec:case2}.
970: Since the coherent fermions are low-energy excitations, they cannot be
971: simply integrated out, but their interaction with $g$-field should
972: be added to the action. The Fermi surface of the coherent fermions is
973: determined by the equation
974: \bea
975: G_0(0,q)\ \tilde{t}_{\perp}({\bf k}_\perp) = 1,
976: \eea
977: and consists of four pockets (two electron-like ones and two hole-like
978: ones) as is shown in Fig.\ref{fig:surface}. Let us consider the
979: situation where the scale of the interchain hopping $t_0$ is very
980: slightly larger than the minimal value $\bar{t}_0$ required for
981: the formation of a Fermi surface
982: \be
983: t_0=\tm+\dt=\sqrt{\frac{11+5\sqrt{5}}{2}}\frac{m}{Z_0}+\dt\ .
984: \ee
985: The electron and hole pockets are then shallow and anisotropic and the Fermi
986: surface is determined by the equation
987: \be
988: E(q,{\bf k}_\perp)=0\ ,
989: \label{e=0}
990: \ee
991: where
992: \bea
993: E(q,{\bf k}_\perp)&=&A_\parallel m\frac{(q-q_0)^2}{q_0^2}+A_\perp m
994: \frac{|{\bf k}_\perp a_\perp|^2}{\kappa_0^2}-E_0\ ,\label{E}\\
995: E_0&\approx&0.352 \delta t\ ,\quad
996: \frac{vq_0}{m}=\left[\frac{1+\sqrt{5}}{2}\right]^\frac{1}{2}\approx 1.27202,\nn
997: A_\parallel&\approx& 0.543\ ,\quad
998: A_\perp\approx 1.27202\ .
999: \eea
1000: The electron pockets are formed at $(k_F+q,{\bf k}_\perp)$ and $(-k_F-q,{\bf
1001: k}_\perp)$ whereas the hole pockets are located at $(k_F-q,{\bf
1002: Q}+{\bf k}_\perp)$ and $(-k_F+q,{\bf Q}+{\bf k}_\perp)$, where $q$
1003: and ${\bf k}_\perp$ are determined from (\ref{e=0}). Let us
1004: denote the annihilation operator of the coherent fermions by
1005: $\Psi(\tau,q,{\bf k}_\perp)$. The soft modes occur in the vicinity of the
1006: electron and hole pockets and it is convenient to decompose
1007: $\Psi(\tau,q,{\bf k}_\perp)$ accordingly. We denote by $R_e(\tau,q,{\bf p}_\perp)$
1008: and $L_e(\tau,q,{\bf p}_\perp)$ the annihilation operator in the vicinity of
1009: the electron pockets and $(q,{\bf p}_\perp)$ is the deviation from $(\pm k_F,{\bf
1010: 0})$. Similarly we denote by $R_h(\tau,q,{\bf p}_\perp)$ and
1011: $L_h(\tau,q,{\bf p}_\perp)$ the annihilation operator in the vicinity of the
1012: hole pockets and $(q,{\bf p}_\perp)$ is the deviation from $(\pm k_F,{\bf Q})$.
1013:
1014: From Eq.(\ref{E}) we determine the particle density associated with a
1015: single pocket is
1016: \bea
1017: n \approx 0.027 a_{\perp}^{-2} q_0\kappa_0^2({\cal Z} - {\cal Z}_c)^{3/2}
1018: \eea
1019: The liquid of quasi-particles becomes degenerate at temperatures of
1020: order of $E_0$. Comparing it with the transition temperatures
1021: (\ref{tc1}), (\ref{tc2}) we conclude that the degenerate metallic state
1022: exists only at
1023: \bea
1024: {\cal Z}-{\cal Z}_c \gg \kappa_0^2\ .
1025: \eea
1026: corresponding to $na_{\perp}^2 \gg 0.027 q_0\kappa_0^5$.
1027: Close to the Fermi surface the Green's function (\ref{G2}) can be
1028: approximated as
1029: \bea
1030: G(\omega,\pm k_F+q,{\bf k}_\perp) &\approx& \frac{Z_2}{i\omega - E(\pm
1031: q,{\bf k}_\perp)}\ ,\\
1032: G(\omega,\pm k_F+ q,{\bf k}_\perp +{\bf Q}) &\approx& \frac{Z_2}
1033: {i\omega + E(\mp q,{\bf k}_\perp)}\ ,
1034: \label{Geh}
1035: \eea
1036: where
1037: \be
1038: Z_2\approx\frac{vq_0}{\tm} \approx 0.352\ .
1039: \ee
1040: The expressions (\ref{Geh}) exhibit the particle-hole symmetry
1041: characteristic of our model at half-filling. As usual, we include the
1042: residue $Z$ in the coupling constant and replace the fermionic action
1043: by the action of four components of free fermions. The effective
1044: action describing the fermions is then given by
1045: \begin{widetext}
1046: \bea
1047: S_{\rm f} &=& a_\perp^2\int
1048: \frac{d\tau\ d^3{\bf k}}{(2\pi)^3}
1049: \left[R^*_{a,\alpha}(\tau,{\bf k})\left(\partial_{\tau} -
1050: E^R_a({\bf k})\right)R_{a,\alpha}(\tau,{\bf k})
1051: + L^*_{a,\alpha}(\tau,{\bf k})
1052: \left(\partial_{\tau} - E^L_a({\bf k})\right)
1053: L_{a,\alpha}(\tau,{\bf k})\right],
1054: \label{fermion}
1055: \eea
1056: \end{widetext}
1057: where ${\bf k}=(q,{\bf k}_\perp)$, $\alpha=\uparrow,\downarrow$, $a=e,h$
1058: and
1059: \be
1060: E_{e}^{R,L}({\bf k})=E(\pm q,{\bf k}_\perp)\ ,\quad
1061: E_{h}^{R,L}({\bf k})=-E(\mp q,{\bf k}_\perp).
1062: \ee
1063: The fermion-spin vertex is described by the action
1064: \bea
1065: S_{\rm int}&=&a_\perp^4\int\frac{d\tau\ d^3{\bf k}\ d^3{\bf k'}}
1066: {(2\pi)^6}{\cal L}_{\rm int}\ ,
1067: \eea
1068: where
1069: \bea
1070: {\cal L}_{\rm int}&=&
1071: I_{{\bf k},{\bf k'}}\sum_{a=e,h}R^*_{a,\alpha}(\tau,{\bf k})L_{a,\beta}(\tau,{\bf k'})
1072: g_{\alpha\beta}(\tau,{\bf k} - {\bf k'})\nn
1073: &+&I_{{\bf k},{\bf Q}+{\bf k'}}R^*_{e,\alpha}(\tau,{\bf k})L_{h,\beta}(\tau,{\bf k'})
1074: g_{\alpha\beta}(\tau,{\bf k} - {\bf Q}-{\bf k'})\nn
1075: &+&I_{{\bf Q}+{\bf k},{\bf k'}}R^*_{h,\alpha}(\tau,{\bf k})L_{e,\beta}(\tau,{\bf k'})
1076: g_{\alpha\beta}(\tau,{\bf k} + {\bf Q}-{\bf k'}) \nn
1077: &+&{\rm h.c.}\ ,\nn
1078: I_{{\bf k},{\bf k'}} &=&2\pi \frac{v \zh Z_2}{m^2}
1079: \tilde{t}_{\perp}({\bf k})\tilde{t}_{\perp}({\bf k'})\ .
1080: \label{int}
1081: \eea
1082: All wave vectors in the above formulae lie close to the
1083: non-interacting Fermi surface and therefore their longitudinal
1084: components are small in comparison to $\pi$: $|q| \ll \pi a_0$. The
1085: entire approach is valid only when the volume inside of the Fermi
1086: surfaces is small. One then can neglect the momentum dependence of
1087: the exchange constant in Eq.(\ref{int}). The sign of the exchange
1088: constant depends on the ``pocket index'' $a, b$
1089: \bea
1090: I_{ab} \approx \gamma m\left(
1091: \begin{array}{cc}
1092: 1 & -1\\
1093: -1 & 1
1094: \end{array}
1095: \right),
1096: \eea
1097: where $\gamma$ is a constant. The interaction can be cast in the form
1098:
1099: \begin{widetext}
1100: \bea
1101: \label{interaction}
1102: {\cal L}_{int} = \gamma m \Bigl[
1103: \sum_aR^*_{a,\alpha}({\bf k})L_{a,\beta}({\bf k}')g_{\alpha\beta}({\bf k}-{\bf k}')
1104: - \Bigl\{R^*_{e,\alpha}({\bf k})L_{h,\beta}({\bf k}')
1105: +R^*_{h,\alpha}({\bf k})L_{e,\beta}({\bf k}')\Bigr\}
1106: g_{\alpha\beta}({\bf k}-{\bf k}'-{\bf Q})\Bigr] + \text{h.c.}
1107: \eea
1108: \end{widetext}
1109: The value of the coupling constant $\gamma$ can be extracted from
1110: Eqns (\ref{jsing}) and (\ref{DOS}) by noting that it is this
1111: interaction which gives rise to the logarithmic singularity in
1112: $J(Q)$
1113: \bea
1114: \tilde{J}^{\rm sing}(0,0,{\bf Q}) \approx -2\gamma^2m^2\frac{\rho(0)}{a_0}
1115: \ln\left[\frac{\delta t}{T}\right].
1116: \eea
1117: Here $\rho(0)$ is the density of states per species at the Fermi
1118: surface of coherent fermions
1119: \bea
1120: \rho(0) &=& \lim_{\omega\to 0}
1121: a_0\int \frac{d^3{\bf k}}{(2\pi)^3}\left[-\frac{1}{\pi}{\rm Im}
1122: G(\omega,k_F+q,{\bf k}_\perp)\right]\nn
1123: &\approx&0.539\ \frac{a_0}{(2\pi)^2}\frac{\kappa_0^2}{v}
1124: \left[\frac{\delta t}{t_0}\right]^\frac{1}{2}.
1125: \label{DOS}
1126: \eea
1127: The result is that close to the transition we have
1128: \be
1129: \gamma\propto \sqrt{\frac{t}{m}}\ ,
1130: \ee
1131: where $t\gg m$ is the hopping along the chains and the constant of
1132: proportionality is of order 1. Though $\gamma$ is never small, the
1133: small parameter $\kappa_0^2$ appears every time one integrates over
1134: the transverse momentum. Hence the magnitude of $\gamma$ is not a
1135: problem.
1136: The effective action describing the metallic side of the
1137: Mott-insulator to metal transition is given by Eqns (\ref{spin}),
1138: (\ref{fermion}) and (\ref{interaction}).
1139: We find it instructive to write it down also in position space
1140: \begin{widetext}
1141: \bea
1142: S_{\rm f}&=&\int d\tau d^3{\bf x}\ \Psi^\dagger_\alpha(\tau,{\bf x})
1143: \left\{(I\otimes I)\p_{\tau} + (I\otimes\tau^z)
1144: \left[E_0+\frac{\partial_x^2}{2M_\parallel}+\frac{\vec{\nabla}_\perp^2}{2M_\perp}
1145: \right]\right\}\Psi_\beta(\tau,{\bf x})\ ,\nn
1146: S_{\rm int}&=&\frac{\gamma m}{2}\int d\tau d^3{\bf x}\ \Psi^\dagger_\alpha(\tau,{\bf x})\left(
1147: \left\{\tau^+\otimes[\exp(-2iq_0x\tau^z)-\tau^x\exp(-i{\bf Q}\cdot{\bf x}_\perp)]\right\}
1148: g_{\alpha\beta}(\tau,{\bf x}) + {\rm h.c.}\right)\Psi_\beta(\tau,{\bf x})\ .
1149: \eea
1150: \end{widetext}
1151: Here we have taken the continuum limit in the directions perpendicular
1152: to the chains and introduced a field
1153: $\Psi^+_\alpha = (\phi^* R^+_{e,\alpha}, \phi R^+_{h,\alpha},
1154: \phi L^+_{e,\alpha}, \phi^* L^+_{h,\alpha})$, where
1155: $\phi=\exp(iq_0x)$. We employ a tensor-product notation, where
1156: the first space is associated with the ``right/left'' index and the
1157: second space with the ``e/h'' index. The Fermi surfaces of electrons and holes
1158: are shifted to the origin and superimposed. The spin action $S[g]$ is
1159: given by Eq.(\ref{spin}). Alternatively, one may use the Abelian
1160: representation given by Eq.(\ref{SGM}), with $g$ defined by
1161: (\ref{gfab}).
1162:
1163: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1164: \subsection{Marginal Fermi liquid}
1165: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1166: As we shall now demonstrate, at temperatures higher than the N\'eel
1167: temperature $T_c$ this metal is, in fact a Marginal Fermi Liquid
1168: \cite{MFL}. The following discussion closely parallels the analysis
1169: given by Chubukov {\it et al} for the spin-fermion model (see, for
1170: example Ref.[\onlinecite{chub2}]). Let us consider diagrams for the
1171: Green's function of right moving electrons. We expand around uncoupled
1172: chains and take both the spin-fermion coupling and interchain
1173: spin-spin exchange into account perturbatively. The elements of the
1174: diagram technique for the fermionic degrees of freedom are as usual,
1175: whereas the building blocks in the spin sector are the connected
1176: 2n-point spin correlators for a single chain. In diagrams that do not
1177: contain closed fermionic loops or the interchain exchange such as the
1178: ones in Figs 7a and 7b, the spin correlations are independent of the
1179: transverse wave vector. This means that each fermion Green's function
1180: is integrated over $k_{\perp}$. This integral does not differ
1181: significantly from the integral over all momenta and as a result
1182: is independent of $q_{\parallel}$, corresponding to a Green's function
1183: that is local in real space:
1184: \be
1185: \int \frac{\rd {\bf k}_{\perp}^2}{(2\pi)^2} \frac{1}
1186: {\ri\omega -E^L_e(q,{\bf k}_\perp)}\approx {\rm const}\
1187: \ri\kappa_0^2\ \mbox{sgn}(\omega)\ .
1188: \label{local}
1189: \ee
1190: As (\ref{local}) is independent of $q$, we may integrate the spin
1191: correlator in the diagram of Fig. 7a over $q$. This makes the spin
1192: correlator local. As a result the contributions to the self energy
1193: which do not contain closed fermionic loops or interchain spin
1194: exchange are approximately momentum independent. Then the self energy
1195: calculation becomes essentially a local problem like
1196: the problem of electron-phohon interactions in metals and
1197: superconductors (the Eliashberg theory)\cite{sima}. In fact, such an
1198: approach works under less stringent conditions, namely, when the spin
1199: excitations in the transverse direction are much slower than the
1200: quasi-particles. Therefore the diagrams generating a ${\bf
1201: k}_{\perp}$ dependence of the spin-spin
1202: correlators, such as the ones in Fig. 7c) and
1203: Fig. 7d) do not affect the result for the electron self energy even
1204: close to the transition. Once such diagrams are neglected, we get an
1205: expansion where a factor $\kappa_0^2$ is associated with each
1206: fermionic line (originating from the integration over ${\bf
1207: k}_{\perp}$, as in Eq.(\ref{local}). Since $\Sigma$ depends only on
1208: frequency, making these lines fat does not change the result
1209: (\ref{local}) and no self-consistency is required.
1210: \begin{figure}[ht]
1211: \begin{center}
1212: \epsfxsize=0.45\textwidth
1213: \epsfbox{diagram.eps}
1214: \end{center}
1215: \caption{Diagrams for the quasi-particle self energy of right-moving
1216: electrons. The lines with arrows represent the fermionic Green's
1217: functions of right and left moving electrons and holes. The 2n-point
1218: vertices denote cumulants of the matrix fields $g$ and $g^\dagger$.}
1219: \label{fig:M}
1220: \end{figure}
1221: The contribution from the diagram Fig. 7a contains the correlation
1222: function
1223: \be
1224: \bigl\langle\!\bigl\langle
1225: e^{\frac{i}{2}\Theta(\tau_1)}\
1226: e^{-\frac{i}{2}\Theta(\tau_2)}
1227: \bigr\rangle\!\bigr\rangle
1228: \simeq \frac{\Bigl(\ln\Bigl[\frac{\Lambda}{T}\Bigr]\Bigr)^{1/2}\pi Tv^{-1}}
1229: {|\sin(\pi T [\tau_1-\tau_2])|}.
1230: \ee
1231: The contribution of the diagram in Fig. 7a) to the self-energy is then
1232: \bea
1233: && \Sigma^{(a)}(\omega) \propto \kappa_0^2\Bigl(\ln\Bigl[\frac{\Lambda}{T}\Bigr]\Bigr)^\frac{1}{2}
1234: \int \rd \tau \frac{\re^{\ri\omega \tau}\ T}{\tau|\sin(\pi T\tau)|} \nonumber\\
1235: &&\sim \ri\kappa_0^2
1236: \omega\ln\left[\frac{E_0}{\mbox{max}\{\omega,T\}}\right]
1237: \Bigl(\ln\Bigl[\frac{\Lambda}{T}\Bigr]\Bigr)^\frac{1}{2}.
1238: \label{mfl}
1239: \eea
1240: Here $E_0^{-1}$ serves as a short-time cutoff in all integrals.
1241: The diagram in Fig 7b) involves cumulants of the type
1242: \bea
1243: &&\la\!\la
1244: e^{\frac{i}{2}\Theta(\tau_1)}
1245: e^{-\frac{i}{2}\Theta(\tau_2)}
1246: e^{\frac{i}{2}\Theta(\tau_3)}
1247: e^{-\frac{i}{2}\Theta(\tau_4)}\ra\!\ra
1248: \eea
1249: and gives a contribution
1250: \bea
1251: &&\Sigma^{(b)}(\omega)\propto \kappa_0^6
1252: \int d\tau_2 d\tau_3 d\tau_4
1253: \frac{e^{i\omega \tau_{14}}}{\tau_{12}\tau_{23}\tau_{34}}\nn
1254: &\times&
1255: \left[\left|\frac{\tau_{13}\tau_{24}}
1256: {\tau_{12}\tau_{14}\tau_{23}\tau_{34}}\right|
1257: - \frac{1}{|\tau_{12}\tau_{34}|} -
1258: \frac{1}{|\tau_{14}\tau_{23}|}\right] \propto \omega^2
1259: %&&\Sigma^{(b)}_2(\omega) \propto \kappa_0^4\int
1260: %\frac{d\tau_2 d\tau_3 d\tau_4}{\tau_{12}\tau_{23}\tau_{34}}
1261: %\frac{e^{i\omega \tau_{14}}}{|\tau_{13}\tau_{24}|}\propto
1262: %\omega^2\ ,
1263: \label{4point}
1264: \eea
1265: to the self energy. Various other contributions are zero because some
1266: local cumulants vanish
1267: \bea
1268: &&\la\!\la
1269: e^{\frac{i}{2}\Phi(\tau_1)}
1270: e^{-\frac{i}{2}\Theta(\tau_2)}
1271: e^{\frac{i}{2}\Theta(\tau_3)}
1272: e^{-\frac{i}{2}\Phi(\tau_4)}\ra\!\ra=0.
1273: \eea
1274: Equation \r{4point} shows that contributions from higher cumulants
1275: can be neglected at small frequencies. As a result, the only essential
1276: contribution to the self energy comes from the diagram Fig. 7a and is
1277: given by Eq. (\ref{mfl}).
1278:
1279:
1280: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1281: \section{ The ordered state}
1282: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1283: As we have discussed in the previous sections, the system undergoes
1284: an antiferromagnetic transition at a temperature much smaller than
1285: the Mott-Hubbard gap $T_c \sim m\kappa_0^2$. Once $T_c$ becomes small
1286: compared to the quasi-particle Fermi energy $E_0$, one can distinguish
1287: between metallic and insulating behavior. As we have demonstrated,
1288: the corresponding metal is rather unusual, being in fact a Marginal
1289: Fermi Liquid. Below $T_c$ however, the system becomes either an
1290: insulator (for zero doping) or an ordinary Fermi liquid. Indeed, at
1291: zero doping the electron and hole Fermi surfaces are nested and the
1292: ordering occurs at the antiferromagnetic wave vector in the transverse
1293: direction (recall that the chains run along the $z$-axis) such that
1294: \be
1295: \langle g_{\bl,\alpha\beta}(x)\rangle=\vec{\s}_{\alpha\beta}\cdot{\bf
1296: M}(-1)^{l_x+l_y}\ .
1297: \label{LRO}
1298: \ee
1299: Here the components of $\vec{\s}$ are the Pauli matrices
1300: and ${\bf M}$ is the ordering vector.
1301: In the mean-field approximation the fermionic spectrum is gapped
1302: \bea
1303: \omega^2_a = (E^R_a({\bf k}))^2 + \gamma^2 m^2 |{\bf M}|^2\ , \ a=e,h.
1304: \eea
1305:
1306: At non-zero doping our approach still holds provided the chemical
1307: potential lies inside the Mott-Hubbard gap. There is no nesting any
1308: longer and the magnetic ordering does not open a gap in the
1309: quasi-particle spectrum. As usual, magnetic fluctuations interact
1310: with quasi-particles through gradient vertices and these interactions
1311: are weak.
1312:
1313: %%%%%%%%%%%%%%%%%%%%%%
1314: \section{Conclusions}
1315: %%%%%%%%%%%%%%%%%%%%%%
1316: The main result of this paper is a formulation of a self-consistent
1317: description
1318: of the hybrid state of 3D quasi-particles interacting with magnetic
1319: collective modes. The derivation is done for
1320: a toy model of half filled Hubbard chains weakly coupled
1321: through a long range interchain hopping. A certain artificiality of
1322: the model was necessary to ensure the self-consistency of our approach
1323: through the presence of a small parameter $\kappa_0^2$.
1324: We have also neglected the long-range component
1325: of the Coulomb interaction, which plays an important role in
1326: determining the character of the metal-insulator transition. In
1327: reality a long-range interaction may lead to an instability of the
1328: small FS phase, though for small Mott-Hubbard gap its influence is
1329: diminished by the presence of a large dielectric constant. In this
1330: case the first order MI transition line may terminate below the
1331: antiferromagnetic transition line of Fig. 1.
1332:
1333: The resulting low energy effective theory is of the Eliashberg type:
1334: the interaction between quasi-particles and collective modes leads to
1335: strong retardation effects resulting in a strongly frequency
1336: dependent quasi-particle self energy. In the present model this takes
1337: place not just at the 'hot spots', as in the spin-fermion model of
1338: Chubukov and Pines, but on the entire quasi-particle FS. This makes the
1339: present model a candidate for the description of 'bad' metals. The
1340: fact that in our model the electron self energy is of the Marginal
1341: Fermi Liquid form is not universal and is determined by the particular
1342: spin fluctuation spectrum.
1343:
1344: As we discussed in the introduction, our theory provides an example
1345: of a state where the number of carriers is unrelated to the volume
1346: of the FS. Though this idea is well established (see e.g. the textbook
1347: [\onlinecite{AGD}]),
1348: its microscopic realization was restricted to superconductors (the
1349: example given in Ref.[\onlinecite{AGD}]). Our model provides another
1350: example. It also demonstrates that one does not need exotic ground
1351: states to have a small FS, as was suggested in Refs
1352: [\onlinecite{senthil1}], [\onlinecite{senthil2}]. The
1353: small FS phenomenology can be generalized beyond our model. In general
1354: there is no {\it a priori} reason for the Fermi surface even to be
1355: closed; for instance, Ref.[\onlinecite{rice}] describes a state with a
1356: truncated Fermi surface observed in ARPES experiments on
1357: undoped cuprates \cite{arpes}.
1358:
1359:
1360: \acknowledgments
1361:
1362: We are grateful to Andrei Chubukov for numerous discussions and
1363: interest to the work. AMT is grateful to S. Buehler-Paschen for
1364: information about Hall effect measurements. AMT acknowledges support
1365: by the US Department of Energy under contract number DE-AC02-98 CH
1366: 10886. FHLE is supported by the EPSRC under grant GR/R83712/01 and the
1367: Institute for Strongly Correlated and Complex Systems at BNL.
1368:
1369: \begin{thebibliography}{99}
1370:
1371:
1372: \bibitem{pines}
1373: D. Pines, Z. Phys. B{\bf 103}, 129 (1997).
1374:
1375: \bibitem{chub2}
1376: A. Abanov, A.V. Chubukov and J. Schmalian, Adv. Phys. {\bf 52}, 119 (2003);
1377: A. V. Chubukov, D. Pines and J. Schmalian, cond-mat/0201140.
1378:
1379: \bibitem{gf2}
1380: F. H. L. Essler and A. M. Tsvelik, Phys. Rev B{\bf 65}, 115117 (2002).
1381:
1382: \bibitem{Lutt}
1383: J.M. Luttinger, Phys. Rev. {\bf 119}, 1153 (1960).
1384:
1385: \bibitem{AGD}
1386: A.A. Abrikosov, L.P. Gorkov and I.E. Dzyaloshinski,
1387: {\sl Methods of Quantum Field Theory in Statistical Physics},
1388: Dover (New York) 1975, page 168.
1389:
1390: \bibitem{Dz}
1391: I.E. Dzyaloshinski, Phys. Rev. B{\bf 68}, 085113 (2003).
1392:
1393: \bibitem{AltChub}
1394: B. L. Altshuler, A. V. Chubukov, A. Dashevskii, A. M. Finkel'stein,
1395: D. K. Morr, Europhys. Lett. {\bf 41}, 401 (1998).
1396:
1397: \bibitem{ET03}
1398: F. H. L. Essler and A. M. Tsvelik, Phys. Rev. Lett. {\bf 90}, 126401
1399: (2003).
1400:
1401: \bibitem{arpes}
1402: M. R. Norman {\it et al.}, Nature (London), {\bf 392}, 157 (1998).
1403:
1404: \bibitem{Sielke}
1405: S. Buehler-Paschen {\it et al.}, submitted to Nature.
1406:
1407: \bibitem{Boies}
1408: D. Boies, C. Bourbonnais and A.-M.S. Tremblay, Phys. Rev. Lett. {\bf
1409: 74}, 968 (1995).
1410:
1411: \bibitem{marc}
1412: M. Bocquet, F.H.L. Essler, A.M. Tsvelik and A. Gogolin,
1413: Phys. Rev. B{\bf 64}, 094425 (2001);
1414: M. Bocquet, Phys. Rev. B{\bf 65}, 184415 (2002).
1415:
1416: \bibitem{katanin}
1417: V.Y. Irkhin and A.A. Katanin, Phys. Rev. {\bf B61}, 6757 (2000).
1418:
1419: \bibitem{LukZam01}
1420: S. Lukyanov and A. B. Zamolodchikov, Nucl. Phys. {\bf B607}, 437 (2001).
1421:
1422: \bibitem{gorkov}
1423: M. Dzero and L. P. Gor'kov, Phys. Rev. B{\bf 69}, 092501 (2004).
1424:
1425: \bibitem{Barzykin00a}
1426: V. Barzykin, J. Phys. Cond. Mat. {\bf 12}, 2053 (2000).
1427:
1428: \bibitem{sb83}
1429: H. J. Schulz and C. Bourbonnais, Phys. Rev. B{\bf 27}, 5856 (1983).
1430:
1431: \bibitem{sima}
1432: G. M. Eliashberg, Sov. Phys. JETP {\bf 11}, 696 (1960);
1433: D. J. Scalapino in {\it Superconductivity}, Vol. 1, p. 449, ed. by
1434: D. Parks, Dekker Inc. N.Y. 1969; F. Marsiglio and J. P. Carbotte, in
1435: {\it The Physics of Conventional and Unconventional Sperconductors},
1436: ed. by K. H. Bennemann and J. B. Ketterson, Springer-Verlag.
1437: %\bibitem{DMFT}
1438: %A. Georges, G. Kotliar, W. Krauth and M.J. Rozenberg,
1439: %Rev. Mod. Phys. {\bf 68}, 13 (1996),\\
1440: %W. Metzner and D. Vollhardt, Phys. Rev. Lett. {\bf 62}, 324 (1989).
1441:
1442: \bibitem{MFL}
1443: C. M. Varma, P. B. Littlewood, S. Schmitt-Rink, E. Abrahams, and
1444: A. E. Ruckenstein, Phys. Rev. Lett. {\bf 63}, 1996 (1989).
1445:
1446:
1447: %\bibitem{giamarchi} S. Biermann, A. Georges, A. Lichtenstein and
1448: %T. Giamarchi, Phys. Rev. Lett. {\bf 87}, 276405 (2001).
1449:
1450:
1451: \bibitem{senthil1}
1452: T. Senthil, S. Sachdev and M. Vojta, Phys. Rev. Lett. {\bf 90}, 216403 (2003).
1453:
1454: \bibitem{senthil2}
1455: T. Senthil, M. Vojta and S. Sachdev, Phys. Rev. B{\bf 69}, 035111 (2004).
1456:
1457: \bibitem{rice}
1458: C. Honerkamp, M. Salmhofer, N. Furukawa and T. M. Rice,
1459: Phys. Rev. B{\bf 63}, 035109 (2001).
1460:
1461: \end{thebibliography}
1462: \end{document}
1463: