cond-mat0308513/gm.tex
1: \documentclass[aps,prl,twocolumn,showpacs,balancelastpage]{revtex4}
2: %\documentclass[aps,prl,preprint,showpacs]{revtex4}
3: 
4: \usepackage{amsmath,amssymb,amsfonts}
5: \usepackage{bm}
6: \usepackage{epsfig}
7: 
8: %% macros
9: \newcommand{\nop}[1]{#1}
10: \newcommand{\bra}[1]{\langle{#1}|}
11: \newcommand{\ket}[1]{|{#1}\rangle}
12: \newcommand{\braket}[2]{\langle{#1}|{#2}\rangle}
13: \newcommand{\expval}[1]{\langle{#1}\rangle}
14: \newcommand{\gwfexp}[1]{\langle{#1}\rangle}%_{\text{G}}^{\phantom{+}}}
15: \newcommand{\ham}{\hat{H}}%{\hat{H}_{\text{G}}}
16: 
17: \begin{document}
18: 
19:   \title{Gossamer metals}
20: 
21:   \author{Marcus Kollar}
22: 
23: %  \email[Email address: ]{kollar@itp.uni-frankfurt.de}
24:   
25:   \affiliation{In\-sti\-tut f\"{u}r Theore\-ti\-sche Phy\-sik,
26:     Jo\-hann-Wolf\-gang-Goethe-Uni\-ver\-si\-t\"{a}t Frank\-furt,
27:     Ro\-bert-Mayer-Stra\ss{}e~8, D-60054~Frankfurt am~Main, Germany.}
28: 
29:   \date{August 26, 2003}
30: 
31:   \begin{abstract}
32:     Laughlin's construction of exact gossamer ground states is applied
33:     to normal metals. We show that for each variational parameter
34:     $0\leq g\leq1$, the paramagnetic or ferromagnetic Gutzwiller wave
35:     function is the exact ground state of an extended Hubbard model
36:     with correlated hopping, with arbitrary particle density,
37:     non-interacting dispersion, and lattice dimensionality.  The
38:     susceptibility and magnetization curves are obtained, showing that
39:     the Pauli susceptibility is enhanced by correlations.  The
40:     elementary quasiparticle excitations are gapless, except for a
41:     half-filled band at $g=0$, where a Mott transition from metal to
42:     insulator occurs.
43:   \end{abstract}
44: 
45:   \pacs{71.27.+a, 71.10.Fd, 71.30.+h}
46: 
47:   \maketitle
48:   
49:   
50:   Progress in the understanding of many-body effects in strongly
51:   correlated electron systems, such as quantum magnets, narrow-band
52:   transition metal compounds, fractional quantum Hall systems, or
53:   high-temperature superconductors, has depended on a variety of
54:   theoretical tools.  Important information about the electronic
55:   structure can often be obtained from \emph{ab initio} calculations,
56:   which are however less reliable if interactions between electrons
57:   are dominant over their kinetic energy.  On the other hand, the
58:   study of idealized model systems, containing only the presumably
59:   relevant degrees of freedom, can provide insight into microscopic
60:   physical mechanisms. However, since such models are rarely exactly
61:   solvable, analytical and numerical calculations usually involve
62:   approximations or extrapolations. In view of these limitations,
63:   support for proposed physical notions has occasionally come from an
64:   inverse strategy: starting from a correlated many-body wavefunction
65:   one constructs a hopefully ``reasonable'' model Hamiltonian for
66:   which it is the exact ground state.  Correlated quantum phases may
67:   then be classified according to their elementary excitations or
68:   correlation functions. This approach has been useful in particular
69:   for the understanding of the fractional quantum Hall effect,
70:   spin-Peierls or Haldane-gap antiferromagnets, and quantum
71:   rotors~\cite{Arovas92}.
72:   
73:   Recently, Laughlin~\cite{Laughlin02a} developed a new approach to
74:   high-temperature superconductivity, viewing the insulating state as
75:   a superconductor with very low superfluid density.  Pursuing the
76:   above strategy, he proposed that the ground-state wavefunction of
77:   such a ``gossamer superconductor'' is obtained from the BCS
78:   mean-field product state by applying the Gutzwiller correlation
79:   operator ($0\leq g\leq1$),
80:   \begin{align}
81:     \hat{K}(g)
82:     =
83:     g^{\sum_{i}\!\hat{D}_{i}}
84:     =
85:     \prod_{\scriptstyle i}
86:     \left[1-(1-g)\hat{D}_{i}\right]
87:     ,\label{eq:K}
88:   \end{align}    
89:   where $\hat{D}_{i}=\hat{n}_{i\uparrow}\hat{n}_{i\downarrow}$ is the
90:   operator for double occupation at lattice site $i$, and
91:   constructed a corresponding model Hamiltonian.  Elementary
92:   excitations~\cite{Laughlin02a}, the transition from superconductor
93:   to Mott insulator~\cite{Zhang02a}, magnetic
94:   instabilities~\cite{Bernevig03a}, and related mean-field
95:   Hamiltonians~\cite{Bernevig03a,Yu02ff} were also studied in this
96:   context.
97:   
98:   The purpose of this letter is the application of Laughlin's gossamer
99:   paradigm to normal metals, i.e., itinerant electrons on a lattice
100:   without broken (discrete) translational symmetries. (In particular,
101:   antiferromagnetic or superconducting phases are excluded.) It is
102:   well known that a metallic system can be driven into an insulating
103:   state by strong electronic correlations.  This type of transition
104:   from metal to insulator, the Mott transition, occurs for example in
105:   transition metal oxides, and has been analyzed by a variety of
106:   theoretical methods \cite{Gebhard97a}.  These include the
107:   variational Gutzwiller wavefunction (GWF) \cite{Gutzwiller63ff}
108:   obtained by acting with (\ref{eq:K}) on an uncorrelated Fermi sea.
109:   In general the GWF describes a correlated metal, except for the
110:   insulating state with one immobile particle at each lattice site
111:   that results at $g=0$ for a half-filled band. When used as a
112:   variational wavefunction for the Hubbard model and evaluated within
113:   the Gutzwiller approximation \cite{Gutzwiller63ff}, this
114:   Brinkman-Rice (BR) transition \cite{Brinkman70a} occurs at finite
115:   critical Hubbard interaction $U_c^{\text{BR}}$.  While the
116:   Gutzwiller approximation becomes exact in the limit of infinite
117:   dimensions \cite{Metzner89a}, the BR transition is shifted to
118:   $U_c^{\text{BR}}=\infty$ in finite dimensions
119:   \cite{vanDongen89a,Gulacsi91ff}.  However, the reliability of these
120:   variational results is limited, as the true ground state of the
121:   Hubbard model in infinite dimensions may behave rather differently
122:   \cite{Georges96a}; for example the number of doubly occupied sites
123:   in general does not vanish at the transition as in the BR scenario.
124:   Furthermore, the analysis of elementary excitations is hampered by
125:   the fact that the true ground state is lower in energy, and on these
126:   grounds the GWF has been criticized as inadequate for describing the
127:   Mott transition \cite{Millis91a}.  Some of these difficulties are
128:   resolved for models with exact GWF ground states, which we now
129:   proceed to construct.
130: 
131: 
132:   \emph{Metallic gossamer ground state.} %
133:   In general a gossamer ground state is built as follows
134:   \cite{Laughlin02a}.  Starting from an uncorrelated product wave
135:   function $\ket{\phi}$ and operators
136:   $\hat{b}_{\bm{k}\sigma}^{\phantom{+}}$ such that
137:   $\hat{b}_{\bm{k}\sigma}^{\phantom{+}}\ket{\phi}=0$ for all $\bm{k}$
138:   and $\sigma$, one applies an invertible many-body correlator
139:   $\hat{K}$ to obtain a correlated wavefunction $\ket{\psi}$ $=$
140:   $\hat{K}\ket{\phi}$, and defines
141:   $\hat{\tilde{b}}_{\bm{k}\sigma}^{\phantom{+}}$ $=$
142:   $\hat{K}\hat{b}_{\bm{k}\sigma}^{\phantom{+}}\hat{K}^{-1}$. Then
143:   $\ket{\psi}$ is an exact ground state of the hermitian Hamiltonian
144:   $\ham$ $=$ $\sum_{\bm{k}\sigma}\tilde{E}_{\bm{k}\sigma}
145:   \hat{\tilde{b}}_{\bm{k}\sigma}^{+}
146:   \hat{\tilde{b}}_{\bm{k}\sigma}^{\phantom{+}}$
147:   for arbitrary $\tilde{E}_{\bm{k}\sigma}\geq0$, since $\ham\geq0$
148:   and $\ham\ket{\psi}=0$.
149:   
150:   In the present context we use the Gutzwiller correlator (\ref{eq:K})
151:   as in Refs.~\onlinecite{Laughlin02a,Zhang02a,Bernevig03a,Yu02ff},
152:   which is invertible for $g\neq0$, $\hat{K}(g)^{-1}=\hat{K}(g^{-1})$,
153:   but start from a product state containing spin-up and spin-down
154:   fermions, characterized by the occupation numbers
155:   $n_{\bm{k}\sigma}^{0}$ (with $n_{\bm{k}\sigma}^{0}$ $=$ $0$ or $1$),
156:   \begin{align}
157:     \ket{\phi}
158:     &=
159:     \!\!\!\!
160:     \prod_{~~~~~\bm{k}\sigma~(n_{\bm{k}\sigma}^{0}=1)~}
161:     \hat{c}_{\bm{k}\sigma}^{+}
162:     \;\ket{0}
163:     \,.\label{eq:phi}
164:   \end{align}
165:   This state is annihilated by the operators
166:   $\hat{b}_{\bm{k}\sigma}^{\phantom{+}}$ $=$
167:   $(1-n_{\bm{k}\sigma}^{0})\hat{c}_{\bm{k}\sigma}^{\phantom{+}}$ $+$
168:   $n_{\bm{k}\sigma}^{0}\hat{c}_{\bm{k}\sigma}^{+}$.  After some
169:   algebra, we can rewrite the Hamiltonian $\ham$ as
170:   \begin{align}
171:     \ham
172:     &=
173:     \hat{H}_t
174:     +
175:     \hat{H}_h
176:     +
177:     \hat{H}_U
178:     +
179:     \hat{H}_X
180:     +
181:     \hat{H}_Y
182:     +
183:     \text{const}
184:     \,,\label{eq:H_tXYU}
185:     \\
186:     \hat{H}_t
187:     &=
188:     \sum_{i\neq j,\sigma}
189:     \nop{T}_{ij\sigma}
190:     \hat{c}_{i\sigma}^{+}    
191:     \hat{c}_{j\sigma}^{\phantom{+}}
192:     \,,%\label{eq:H_t}
193:     ~~%\\
194:     \hat{H}_h
195:     =%&=
196:     -h
197:     \sum_{i}
198:     (
199:     \hat{n}_{i\uparrow}    
200:     -
201:     \hat{n}_{i\downarrow}    
202:     )
203:     \,,%\label{eq:H_h}
204:     \label{eq:H_t+h}
205:     \\
206:     \hat{H}_X
207:     &=
208:     \sum_{i\neq j,\sigma}
209:     X_{ij\sigma}
210:     (\hat{n}_{i\bar{\sigma}}
211:     +\hat{n}_{j\bar{\sigma}})
212:     \hat{c}_{i\sigma}^{+}    
213:     \hat{c}_{j\sigma}^{\phantom{+}}
214:     \,,\label{eq:H_X}
215:     \\
216:     \hat{H}_Y
217:     &=
218:     \sum_{i\neq j,\sigma}
219:     Y_{ij\sigma}
220:     \hat{n}_{i\bar{\sigma}}
221:     \hat{n}_{j\bar{\sigma}}
222:     \hat{c}_{i\sigma}^{+}    
223:     \hat{c}_{j\sigma}^{\phantom{+}}
224:     \,,%\label{eq:H_Y}
225:     ~~%\\
226:     \hat{H}_U
227:     =%&=
228:     U\sum_i
229:     \hat{n}_{i\uparrow}    
230:     \hat{n}_{i\downarrow}    
231:     \,,%\label{eq:H_U}
232:     \label{eq:H_Y+U}
233:   \end{align}
234:   with the constant term depending only on $g$ and the total particle
235:   density $n$ $=$ $\hat{n}_{\uparrow}+\hat{n}_{\downarrow}$, which is
236:   fixed; we mostly consider densities $n$ $\leq$ $1$ since
237:   $\ket{\psi(0)}$ $=$ $0$ otherwise.  Here $\nop{T}_{ij\sigma}$ is the
238:   Fourier transform of $\nop{E}_{\bm{k}\sigma}$ $=$
239:   $(1-2n_{\bm{k}\sigma}^0)\tilde{E}_{\bm{k}\sigma}$, and the other
240:   parameters are given by
241:   \begin{align}
242:     h
243:     &=
244:     -\frac{1}{2L}
245:     \sum_{\bm{k}\sigma}
246:     \sigma\,
247:     (1-(1+g^2)n_{\bm{k}\sigma}^0))
248:     \tilde{E}_{\bm{k}\sigma}
249:     \,,\label{eq:h}
250:     \\
251:     U
252:     &=
253:     \frac{1-g^2}{g^2L}
254:     \sum_{\bm{k}\sigma}
255:     (1-(1-g^2)n_{\bm{k}\sigma}^0))
256:     \tilde{E}_{\bm{k}\sigma}
257:     \,,\label{eq:U}
258:     \\
259:     X_{ij\sigma}
260:     &=
261:     \frac{1-g}{2g}
262:     \left[
263:     (1-g)\nop{T}_{ij\sigma}
264:     +
265:     (1+g)\tilde{T}_{ij\sigma}
266:     \right]
267:     ,\label{eq:X}
268:     \\
269:     Y_{ij\sigma}
270:     &=
271:     \frac{(1-g)^2}{2g^2}
272:     \left[
273:     (1+g^2)\nop{T}_{ij\sigma}
274:     +
275:     (1-g^2)\tilde{T}_{ij\sigma}
276:     \right]
277:     ,\label{eq:Y}
278:   \end{align}
279:   where $\tilde{T}_{ij\sigma}$ is the Fourier transform of
280:   $\tilde{E}_{\bm{k}\sigma}$, and $L$ is the number of lattice sites.
281: 
282:   A model with arbitrary non-interacting dispersion
283:   $\epsilon_{\bm{k}}$ can now be obtained as follows.  For given band
284:   dispersion $\nop{E}_{\bm{k}\sigma}$ we construct the Fermi sea via
285:   $n_{\bm{k}\sigma}^0$ $=$ $\Theta(-\nop{E}_{\bm{k}\sigma})$ and let
286:   $\tilde{E}_{\bm{k}\sigma}$ = $|\nop{E}_{\bm{k}\sigma}|$ $\geq$ $0$.
287:   Then we put $\nop{E}_{\bm{k}\sigma}$ $=$ $\epsilon_{\bm{k}}$ $-$
288:   $\epsilon_{\text{F}\!\sigma}$ and adjust the Fermi energies
289:   $\epsilon_{\text{F}\!\sigma}$ so that the starting wavefunction
290:   $\ket{\phi}$ is a Fermi sea with desired densities $n_\sigma$ $=$
291:   $\frac{1}{L}\sum_{\bm{k}}n_{\bm{k}\sigma}^0$.
292:   In the following we assume $\sum_{\bm{k}}\epsilon_{\bm{k}}$ $=$
293:   $0$ for convenience,
294:   hence $\epsilon_{0\sigma}$ $\equiv$
295:   $\frac{1}{L}\sum_{\bm{k}}\epsilon_{\bm{k}}n_{\bm{k}\sigma}^0$ $\leq$
296:   $0$.
297:   
298:   For each $0\leq g\leq1$ the Gutzwiller wavefunction $\ket{\psi(g)}$
299:   $=$ $\hat{K}(g)\ket{\phi}$ is the exact ground state of the extended
300:   Hubbard Hamiltonian (\ref{eq:H_tXYU}).  It contains the kinetic
301:   energy $\hat{H}_t$ of a single band $\epsilon_{\bm{k}}$, which is
302:   independent of $g$, and an optional Zeeman term $\hat{H}_h$, absent
303:   for $\hat{n}_{\uparrow}$ $=$ $\hat{n}_{\downarrow}$.  For $g<1$,
304:   $\ham$ contains interactions that involve at most two sites: a
305:   repulsive on-site interaction $\hat{H}_U$ and correlated hopping
306:   terms $\hat{H}_X$ and $\hat{H}_Y$, whose amplitudes are related by
307:   $gY_{ij\sigma}$ $=$ $(1-g)^2(\nop{T}_{ij\sigma}+X_{ij\sigma})$.
308:   Note that $X_{ij\sigma}$, $Y_{ij\sigma}$, and $U$ all diverge in the
309:   limit $g\to0$.  Similar interaction terms appear in models with
310:   superconducting gossamer ground states
311:   \cite{Laughlin02a,Zhang02a,Bernevig03a,Yu02ff}, but here those
312:   states cannot be lower in energy than $\ket{\psi(g)}$.  Apart from
313:   $g$, the magnetic field and the strength and range of the
314:   interactions depend on the chosen band dispersion
315:   $\epsilon_{\bm{k}}$ and the densities $n_\sigma$. To illustrate the
316:   behavior of the amplitude $\tilde{T}_{ij\sigma}$ appearing in
317:   (\ref{eq:X})-(\ref{eq:Y}) we now discuss several examples.
318: 
319: 
320:   \emph{One-dimensional systems.} %
321:   The dispersion for a one-dimensional ring with nearest-neighbor
322:   hopping $-t<0$ is $\epsilon_k=-2t\cos(k)$.
323:   For the Fourier transform of
324:   $\tilde{E}_{k\sigma}=|\epsilon_k-\epsilon_{\text{F}\!\sigma}|$ we
325:   find
326:   \begin{align}
327:     \tilde{T}_{j\pm1,j\sigma}
328:     &=
329:     t\,
330:     \big[
331:     2n_\sigma-1
332:     +
333:     \tfrac{1}{\pi}
334:     \sin(\pi n_\sigma)
335:     \big]
336:     \,,\label{eq:tildeT_1}
337:     \\
338:     \tilde{T}_{j+r,j\sigma}
339:     &=
340:     \frac{4t}{\pi(r^2-1)}
341:     \big[
342:     r\sin(\pi n_\sigma r)\cos(\pi n_\sigma)
343:     \\
344:     &~~~
345:     +
346:     \cos(\pi n_\sigma r)\sin(\pi n_\sigma)
347:     \big]
348:     \,,~~|r|\geq2
349:     ,\label{eq:tildeT_2}
350:   \end{align}
351:   which falls off algebraically at large distances.  At half-filling
352:   ($n_\sigma$ $=$ $1/2$) it is on the order of $1/r^2$ and alternates
353:   in sign for even $r$, while vanishing for odd $r$.  This long-range
354:   behavior of $\tilde{T}_{ij\sigma}$ is rather generic.  As another
355:   example we consider ``$1/r$'' hopping, $\nop{T}_{j+r,j\sigma}$ $=$
356:   $it(-1)^{r}/r$ with dispersion $\epsilon_k$ $=$ $tk$, for which the
357:   corresponding Hubbard model was solved by Gebhard \emph{et al.} (see
358:   Ref.~\onlinecite{Gebhard97a} for a review).  We obtain
359:   \begin{align}
360:     \tilde{T}_{j+r,j\sigma}
361:     &=
362:     \frac{(-1)^rt}{\pi r^2}
363:     \big[
364:     1-i\pi(2n_\sigma-1)r-e^{-2\pi in_\sigma r}
365:     \big]
366:     \,,\label{eq:tildeT_lr}
367:   \end{align}
368:   again with contributions proportional to $1/r$ (absent for
369:   half-filling) and $1/r^2$.
370:   Similar power-law behavior is typically found in dimensions $D=2,3$.
371: 
372: 
373:   \emph{Infinite-dimensional systems.} %
374:   Nearest-neigh\-bor hopping $t=1/\sqrt{2D}$ on a hypercubic lattice
375:   with dispersion $\epsilon_{\bm{k}}$ $=$ $-2t\sum_{\alpha=1}^D\cos
376:   k_\alpha$ yields the density of states $\rho_{\text{hc}}(\epsilon)$
377:   $=$ $\exp(-\epsilon^2/2)/\sqrt{2\pi}$ in the limit $D\to\infty$
378:   \cite{Metzner89a,Georges96a}.
379:   In order to construct the corresponding amplitude
380:   $\tilde{T}_{ij\sigma}$ further assumptions about its symmetry are
381:   necessary.  Following Ref.~\onlinecite{Bluemer03a} we assume that it
382:   depends only on the ``taxi-cab'' distance $||\bm{R}||$ $=$
383:   $\sum_{\alpha=1}^D|R_\alpha|$ and use the appropriate scaling
384:   $\tilde{T}_{ij\sigma}$ $=$
385:   $\tilde{T}_{r,\sigma}^*/\smash[b]{\sqrt{2^r\scriptstyle\binom{D}{r}}}$
386:   where $r$ $=$ $||\bm{R}_i-\bm{R}_j||$ $\geq$ $0$. We then obtain
387:   \begin{align}
388:     \tilde{T}_{r\sigma}^*
389:     &= 
390:     \int%\limits_{-\infty}^{\infty}
391:     |\epsilon-\epsilon_{\text{F}\!\sigma}|
392:     \,
393:     \rho_{\text{hc}}(\epsilon)
394:     \,
395:     \frac{\text{He}_r(\epsilon)}{\sqrt{r!}}
396:     \,d\epsilon
397:     \,,
398:     \\
399:     \tilde{T}_{2r+1,\sigma}^*
400:     &=
401:     (1-2n_\sigma)
402:     \,
403:     \delta_{r,0}
404:     \,,
405:     \\
406:     \tilde{T}_{2r+2,\sigma}^*
407:     &=
408:     2\,
409:     \rho_{\text{hc}}(\epsilon_{\text{F}\!\sigma})
410:     \frac{\text{He}_{2r}(\epsilon_{\text{F}\!\sigma})}{\sqrt{(2r+2)!}}
411:     \,,
412:   \end{align}
413:   where $\text{He}_{n}(x)$ are Hermite polynomials. At half-filling we
414:   find $\tilde{T}_{2r\sigma}^*\sim r^{-5/4}$, corresponding to an
415:   effective correlated hopping range $\sum_r\tilde{T}_{r\sigma}^*{}^2$
416:   of order unity.  For other densities of states $\rho(\epsilon)$, in
417:   particular those with finite bandwidth, it is also possible
418:   construct a corresponding dispersion $\epsilon_{\bm{k}}$
419:   \cite{Bluemer03a}, and then derive $\nop{T}_{ij\sigma}$ and
420:   $\tilde{T}_{ij\sigma}$ in a similar fashion.
421: 
422: 
423:   \emph{Response to external magnetic field.} %
424:   Returning to the case of arbitrary dispersion and densities, we note
425:   that according to the equation of state (\ref{eq:h}) the
426:   ground-state magnetization $m$ $=$
427:   $\hat{n}_{\uparrow}-\hat{n}_{\downarrow}$ is nonzero if an external
428:   magnetic field $h$ is present.  For the homogeneous susceptibility
429:   $\chi$ we obtain
430:   \begin{align}
431:     \chi(h)^{-1}
432:     &=
433:     \frac{\partial h}{\partial m}
434:     =
435:     \frac{1}{4}
436:     \sum_\sigma
437:     \frac{
438:       1-(1-g^2)n_{\sigma}
439:     }{
440:       \rho(\epsilon_{\text{F}\!\sigma})
441:     }
442:     \,.\label{eq:chi_h}
443:   \end{align}
444:   \begin{figure}[t]
445:     \centerline{\epsfig{file=figure1.eps,width=\hsize}}
446:     \caption{Magnetization curves $m(h)$
447:       for a one-dimensional ring with nearest-neighbor hopping at
448:       half-filling. Inset: same for $1/r$ hopping.
449:       \label{fig:m}}
450:   \end{figure}
451:   In the limit of zero field this reduces to $\chi(0)$ $=$
452:   $\chi_0/[1-(1-g^2)n/2]$.  As expected the system behaves like a
453:   correlated paramagnet, i.e., the interactions enhance the Pauli
454:   susceptibility $\chi_0$ $=$ $2\rho(\epsilon_{\text{F}})$ of the
455:   uncorrelated system.  However, it should be kept in mind that the
456:   interaction parameters (\ref{eq:U})-(\ref{eq:Y}) do not remain
457:   constant when the parameters $h$ or $m$ are varied.
458:   Fig.~\ref{fig:m} shows the magnetization as a function of magnetic
459:   field for a one-dimensional ring at half-filling. Interestingly, for
460:   nearest-neighbor hopping the upward curvature of these magnetization
461:   curves is very similar to Bethe-ansatz results for the pure Hubbard
462:   model \cite{Lieb68a,Takahashi69a}, where a metal-insulator
463:   transition occurs at $U_c=0^+$.  By contrast, for $1/r$ hopping the
464:   magnetization curves are strictly linear, $m$ $=$ $\chi(0)h$, due to
465:   the constant density of states.
466: 
467: 
468:   \emph{Metal-insulator transition.} %
469:   For an unpolarized half-filled band ($n$ $=$ $1$,
470:   $\epsilon_{\text{F}\!\sigma}$ $=$ $0$), the ground-state wave
471:   function $\ket{\psi(g)}$ describes a metal for $g>0$ and an
472:   insulator for $g=0$. In the insulating state there are no doubly
473:   occupied sites, the discontinuity of $n_{\bm{k}\sigma}$ at the Fermi
474:   surface vanishes, and the kinetic energy $\gwfexp{\hat{H}_t}$ is
475:   zero.  This Mott metal-insulator transition in the ground state of
476:   $\ham$ occurs at infinite interactions (\ref{eq:U})-(\ref{eq:Y}), in
477:   contrast to the variational BR transition, or numerical results for
478:   the Hubbard model in infinite dimensions \cite{Georges96a}.
479:   
480:   Nevertheless we may, somewhat artificially, shift the transition to
481:   finite interactions as follows. Clearly $\ket{\psi(g)}$ remains the
482:   ground state when we multiply $\ham$ by a positive $g$-dependent
483:   factor, although qualitatively different Hamiltonians may then
484:   result in the limit $g\to0$.  For example, for the Hamiltonian
485:   $\ham^{(1)} = g \ham$ the $X$ term has a finite limit, while
486:   $\ham^{(2)} = g^2 \ham$ yields a vanishing $X$ term and finite $Y$
487:   and $U$ terms; in both cases the quadratic kinetic energy vanishes
488:   at $g=0$.  In particular we may conclude that for any dispersion
489:   $\epsilon_{\bm{k}}$ the Hamiltonian
490:   \begin{align}
491:     \ham'
492:     &=
493:     \sum_{i\neq j,\sigma}
494:     Y_{ij}'
495:     \hat{n}_{i\bar{\sigma}}
496:     \hat{n}_{j\bar{\sigma}}
497:     \hat{c}_{i\sigma}^{+}    
498:     \hat{c}_{j\sigma}^{\phantom{+}}
499:     +
500:     U'
501:     \sum_i
502:     \hat{n}_{i\uparrow}    
503:     \hat{n}_{i\downarrow}    
504:     \,,\label{eq:H_prime}
505:   \end{align}
506:   where $Y_{ij}'$ is the Fourier transform of $\epsilon_{\bm{k}}
507:   (1-n_{\bm{k}\sigma}^0)$, has the exact ground state
508:   $\ket{\psi(g=0)}$ at half-filling if $U'$ $\geq$ $U_c'$, with
509:   critical interaction $U_c'$ $=$ $-\sum_\sigma \epsilon_{0\sigma}$
510:   $\equiv$ $|\epsilon_0|$.  Interestingly, the uncorrelated kinetic
511:   energy also sets the energy scale of the BR transition in the
512:   Gutzwiller approximation \cite{Brinkman70a}, where $U_c^{\text{BR}}$
513:   $=$ $8|\epsilon_0|$.  Although $\ham'$ is not a standard Hubbard
514:   Hamiltonian, it nonetheless appears to be the simplest model with a
515:   BR-type transition to an exact insulating ground state at finite
516:   Hubbard interaction.
517: 
518:   \begin{figure}[t]
519:     \centerline{\epsfig{file=figure2.eps,width=\hsize}}
520:     \caption{Expectation values of parts of $\ham$,
521:       (\ref{eq:H_t+h})-(\ref{eq:H_Y+U}), for nearest-neighbor hopping
522:       in $D$ $=$ $1$ (arrows) and $D$ $=$ $\infty$ at half-filling.
523:       Inset: double occupancy for $1/r$ hopping in $D$ $=$ $1$, as
524:       compared to the pure Hubbard model.\label{fig:H_tXYU}}
525:   \end{figure}
526: 
527:   \emph{Ground-state expectation values.} %
528:   The separate expectation values of the kinetic energy $\hat{H}_t$
529:   and the interaction terms $\hat{H}_X$, $\hat{H}_Y$, and $\hat{H}_U$
530:   can be calculated from the quantities $n_{\bm{k}\sigma}$ $=$
531:   $\gwfexp{\hat{c}_{\bm{k}\sigma}^{+}
532:     \hat{c}_{\bm{k}\sigma}^{\phantom{+}}}$, $d$ $=$ $\frac{1}{L}
533:   \sum_{i} \gwfexp{\hat{n}_{i\uparrow} \hat{n}_{i\downarrow}}$,
534:   $x_{ij\sigma}$ $=$ $\gwfexp{\hat{n}_{i\bar{\sigma}}
535:     \hat{c}_{i\sigma}^{+} \hat{c}_{j\sigma}^{\phantom{+}}}$, and
536:   $y_{ij\sigma}$ $=$ $\gwfexp{\hat{n}_{i\bar{\sigma}}
537:     \hat{n}_{j\bar{\sigma}} \hat{c}_{i\sigma}^{+}
538:     \hat{c}_{j\sigma}^{\phantom{+}}}$.  Using the methods of
539:   \cite{Metzner87ff,Kollar01ff} it can be shown that for the GWF the
540:   Fourier transforms of the latter are given by
541:   \begin{align}
542:     x_{\bm{k}\sigma}
543:     &=
544:     n_{\bm{k}\sigma}^{0}
545:     \bigg[
546:       n_{\bar{\sigma}}
547:       -\frac{1-n_{\bm{k}\sigma}}{1-g}
548:     \bigg]
549:     -
550:     \frac{(1-n_{\bm{k}\sigma}^{0})g\,n_{\bm{k}\sigma}}{1-g}
551:     -
552:     d
553:     \,,\label{eq:xks-alternate}
554:     \\
555:     y_{\bm{k}\sigma}
556:     &=
557:     n_{\bm{k}\sigma}^{0}
558:     \bigg[
559:       n_{\bar{\sigma}}
560:       -\frac{1-n_{\bm{k}\sigma}}{(1-g)^2}
561:     \bigg]
562:     +
563:     \frac{(1-n_{\bm{k}\sigma}^{0})
564:         g^2n_{\bm{k}\sigma}}{(1-g)^2}
565:     -
566:     d
567:     \,,\!\!
568:     \label{eq:yks-alternate}
569:   \end{align}
570:   i.e., only the momentum-space distribution $n_{\bm{k}\sigma}$ and
571:   the double occupancy $d$ are needed. They may be evaluated
572:   numerically by Monte-Carlo methods, but are also available in closed
573:   form under certain circumstances.  For one-dimensional systems with
574:   symmetric Fermi sea ($n_{k\sigma}^{0}$ $=$ $n_{-k\sigma}^{0}$) both
575:   quantities have been calculated analytically
576:   \cite{Metzner87ff,Kollar01ff}.  In dimension $D=2,3$ high-order
577:   perturbative methods can be used \cite{Gulacsi91ff}.  The Gutzwiller
578:   approximation \cite{Gutzwiller63ff}, with piecewise constant
579:   momentum distribution $n_{\bm{k}\sigma}$, is recovered in infinite
580:   dimensions \cite{Metzner89a}.
581:   
582:   The expectation values of the various parts of $\ham$ are shown in
583:   Fig.~\ref{fig:H_tXYU} for nearest-neighbor hopping in $D$ $=$ $1$
584:   and $D$ $=$ $\infty$ at half-filling. We note that
585:   $\gwfexp{\hat{H_X}}$ approaches a constant for $g\to0$, while
586:   $\gwfexp{\hat{H_Y}}$ and $\gwfexp{\hat{H_U}}$ diverge. This behavior
587:   occurs for all $D$, since $d\sim g^2\ln(1/g)$ in one dimension
588:   \cite{Metzner87ff}, $d=o(g)$ in all finite dimensions
589:   \cite{vanDongen89a,Gulacsi91ff}, and $d$ $\sim$ $g$ in infinite
590:   dimensions. We may thus conclude that the penalty that $\hat{H}_U$
591:   imposes on double occupancies is compensated by assisted hopping due
592:   to the nonstandard three-body interaction $\hat{H}_Y$.
593:   
594:   The effect of correlated hopping is also apparent when comparing to
595:   the pure Hubbard ring with $1/r$ hopping, which features a
596:   metal-insulator transition at $U_c$ $=$ $2\pi t$ with continuous
597:   nonzero double occupancy $d$ \cite{Gebhard97a}.  For comparison with
598:   previous studies of variational wavefunctions in the vicinity of
599:   this transition \cite{Gebhard94a,Dzierzawa95a}, $d$ vs.\ $U$ is
600:   shown in the inset of Fig.~\ref{fig:H_tXYU}.  The results for both
601:   models with $1/r$ hopping agree for weak interactions, but the
602:   energy gain from correlated hopping leads to a larger number of
603:   doubly occupied sites for strong coupling in the model
604:   (\ref{eq:H_tXYU}), as expected.
605: 
606: 
607:   \emph{Quasiparticle excitations.} %
608:   The known ground state of $\ham$ suggests that it might also be
609:   possible to calculate dynamical properties of the model, such as the
610:   spectral function. Unfortunately the construction of exact excited
611:   states is not straightforward, be it with one added or removed
612:   particle, or with charge or spin excitations. We therefore proceed
613:   by considering the variational states
614:   \cite{Laughlin02a,Buenemann03a}
615:   \begin{align}
616:     \ket{\bm{k}\sigma}
617:     =
618:     \hat{K}\hat{b}_{\bm{k}\sigma}^{+}\ket{\phi}
619:     =
620:     \left\{
621:       \begin{array}{ll}
622:         \hat{K}\hat{c}_{\bm{k}\sigma}^{+}\ket{\phi}
623:         &
624:         \text{~if~}n_{\bm{k}\sigma}^{0}=0
625:         \\
626:         \hat{K}\hat{c}_{\bm{k}\sigma}^{\phantom{+}}\ket{\phi}        
627:         &
628:         \text{~if~}n_{\bm{k}\sigma}^{0}=1
629:       \end{array}
630:     \right.
631:     \,,
632:   \end{align}
633:   whose mean energy is
634:   \begin{align}
635:     \nop{E}_{\bm{k}\sigma}^{\pm}
636:     =
637:     \frac{
638:       \bra{\bm{k}\sigma}
639:       \ham
640:       \ket{\bm{k}\sigma}}{\braket{\bm{k}\sigma}{\bm{k}\sigma}}
641:     =
642:     \frac{
643:       \braket{\psi}{\psi}}{
644:       \braket{\bm{k}\sigma}{\bm{k}\sigma}}
645:       \tilde{E}_{\bm{k}\sigma}\;
646:     \,,
647:   \end{align}
648:   where the commutator relations
649:   ${[\hat{b}_{\bm{k}\sigma}^{\phantom{+}},}$
650:   ${\hat{b}_{\bm{k'}\sigma'}^{+}]}$ $=$
651:   $\delta_{\bm{k}\bm{k}'}\delta_{\sigma\sigma'}$ and
652:   ${[\hat{b}_{\bm{k}\sigma}^{\phantom{+}},}$
653:   ${\hat{b}_{\bm{k'}\sigma'}^{\phantom{+}}]}$ $=$ $0$ were used. The
654:   states $\ket{\bm{k}\sigma}$ are mutually orthogonal and their energy
655:   is thus an upper bound to the quasiparticle energy for momentum
656:   $\bm{k}$ and spin $\sigma$.  The variational energy to add a
657:   particle (i.e., $n_{\bm{k}\sigma}^{0}=0$) is
658:   \begin{align}
659:     \nop{E}_{\bm{k}\sigma}^{+}
660:     &=
661:     \frac{
662:       |\epsilon_{\bm{k}}-\epsilon_{\text{F}\!\sigma}|
663:     }{
664:       1-(1+g)[(1-g)n_{\bar{\sigma}}+(1+g)n_{\bm{k}\sigma}]
665:     }
666:     \,,
667:   \end{align}
668:   while for the removal of a particle (with $n_{\bm{k}\sigma}^{0}=1$)
669:   \begin{align}
670:     \nop{E}_{\bm{k}\sigma}^{-}
671:     &=
672:     \frac{
673:       g^2
674:       |\epsilon_{\bm{k}}-\epsilon_{\text{F}\!\sigma}|
675:     }{
676:       (1+g)[(1-g)n_{\bar{\sigma}}+(1+g)n_{\bm{k}\sigma}]-1-2g
677:     }
678:     \,.
679:   \end{align}
680:   \begin{figure}[t]
681:     \centerline{\epsfig{file=figure3.eps,width=\hsize}}
682:     \caption{Quasiparticle excitations in a one-dimensional
683:       ring with nearest-neighbor hopping $t>0$.\label{fig:E}}
684:   \end{figure}
685:   Clearly the quasiparticle excitations are gapless, since
686:   $\nop{E}_{\bm{k}\sigma}^{\pm}$ $\to$ $0$ close to the Fermi surface.
687:   Fig.~\ref{fig:E} shows these energies for one-dimensional
688:   nearest-neighbor hopping at half-filling.
689: 
690:   
691:   \emph{Conclusion.} %
692:   We have constructed and characterized a new class of itinerant
693:   electron models for which the metallic Gutzwiller wavefunction is an
694:   exact ground state, due to the interplay of Hubbard interaction and
695:   correlated hopping.  For a half-filled band a Mott metal-insulator
696:   transition similar to the Brinkman-Rice scenario occurs,
697:   illustrating Mott's original idea of a quantum phase transition
698:   entirely due to charge correlations without magnetic ordering.
699:   Further study of the elementary excitations in these models should
700:   be fruitful.
701: 
702:   This work was supported in part by the DFG via Forschergruppe
703:   FOR~412.
704: 
705:   \begin{thebibliography}{99}
706:     
707:   \bibitem{Arovas92}%
708:     For a review, see D.~P.\ Arovas and S.~M.\ Girvin, in: {\em Recent
709:       Progress in Many-Body Theories}, Vol.\ 3, edited by T.~L.\ 
710:     Ainsworth, C.~E.\ Campbell, B.~E.\ Clements, and E.\ Krotschek
711:     (Plenum Press, New York, 1992), p.\ 315.
712: 
713:   \bibitem{Laughlin02a}%
714:     R.~B.\ Laughlin, \eprint{cond-mat/0209269}.
715:     
716:   \bibitem{Zhang02a}%
717:     F.~C.\ Zhang, \eprint{cond-mat/0209272}.
718: 
719:   \bibitem{Bernevig03a}%
720:     B.~A.\ Bernevig, R.~B.\ Laughlin, and D.~I.\ Santiago,
721:     \eprint{cond-mat/0303045}.
722: 
723:   \bibitem{Yu02ff}%
724:     Y.\ Yu, \eprint{cond-mat/0211131}; \eprint{cond-mat/0303501}.    
725: 
726:   \bibitem{Gebhard97a}%
727:     F.\ Gebhard, \textit{The Mott metal-insulator transition: models
728:       and methods} (Springer, Berlin 1997).
729: 
730:   \bibitem{Gutzwiller63ff}%
731:     M.~C.\ Gutzwiller, 
732:     Phys.\ Rev.\ Lett.\ {\bf 10}, 159 (1963);
733:     Phys.\ Rev.\ {\bf 134}, A 923 (1964);
734:     Phys.\ Rev.\ {\bf 137}, A 1726 (1965).
735:     
736:   \bibitem{Brinkman70a}%
737:     W.~F.\ Brinkman and T.~M.\ Rice,
738:     Phys.\ Rev.\ B {\bf 2}, 4302 (1970).
739:       
740:   \bibitem{Metzner89a}%
741:     W.\ Metzner and D.\ Vollhardt,
742:     Phys.\ Rev.\ Lett.\ {\bf 62}, 324 (1989).
743: 
744:   \bibitem{vanDongen89a}%
745:     P.~G.~J.\ van Dongen, F.\ Gebhard, and D.\ Vollhardt, Z.\ Phys.\ B
746:     {\bf 76}, 199 (1989).
747: 
748:   \bibitem{Gulacsi91ff}%
749:     Z.\ Gul\'{a}csi and M.\ Gul\'{a}csi,
750:     Phys.\ Rev.\ B {\bf 44}, 1475 (1991).
751:     Z.\ Gul\'{a}csi,
752:     M.\ Gul\'{a}csi, and
753:     B.\ Jank\'{o},
754:     Phys.\ Rev.\ B {\bf 47}, 4168 (1993).
755: 
756:   \bibitem{Georges96a}%
757:     For a review, see A.\ Georges, G.\ Kotliar, W.\ Krauth, and M.~J.\ 
758:     Rozenberg, Rev.\ Mod.\ Phys.\ {\bf 68}, 13 (1996).
759: 
760:   \bibitem{Millis91a}%
761:     A.~J.\ Millis and S.~N.\ Coppersmith,
762:     Phys.\ Rev.\ B {\bf 43}, 13770 (1991).    
763: 
764:   \bibitem{Metzner87ff}%
765:     W.\ Metzner and D.\ Vollhardt,
766:     Phys.\ Rev.\ Lett.\ {\bf 59}, 121 (1987);
767:     Phys.\ Rev.\ B {\bf 37}, 7382 (1988);
768:     \emph{ibid.}~{\bf 39}, 12339 (1989).
769: 
770:   \bibitem{Kollar01ff}%
771:     M.\ Kollar and D.\ Vollhardt, Phys.\ Rev.\ B {\bf 63}, 045107
772:     (2001); \emph{ibid.}~{\bf 65}, 155121 (2002).
773:         
774:   \bibitem{Gebhard94a}%
775:     F.\ Gebhard and A.\ Girndt, Z.\ Phys.\ B {\bf 93}, 445 (1994).
776: 
777:   \bibitem{Dzierzawa95a}%
778:     M.\ Dzierzawa, D.\ Baeriswyl, M.\ Di Stasio, Phys.\ Rev.\ B {\bf
779:       51}, 1993 (1995).
780: 
781:   \bibitem{Bluemer03a}%
782:     N.\ Bl\"{u}mer and P.~G.~J.\ van Dongen,
783:     \eprint{cond-mat/0303204}.  N.\ Bl\"{u}mer, Ph.~D.\ thesis
784:     (Universit\"{a}t Augsburg, 2002).
785: 
786:   \bibitem{Lieb68a}%
787:     E.~H.\ Lieb and F.~Y.\ Wu,
788:     Phys.\ Rev.\ Lett.\ {\bf 20}, 1445 (1968).
789: 
790:   \bibitem{Takahashi69a}%
791:     M.~Takahashi, Progr.\ Theor. Phys. {\bf 42}, 1098 (1969).
792:     
793:   \bibitem{Buenemann03a}%
794:     J.\ Buenemann, F.\ Gebhard, and R.\ Thul, Phys.\ Rev.\ B {\bf 67},
795:     075103 (2003).
796: 
797:   \end{thebibliography}
798: 
799: \end{document}
800: