1: \documentclass[showpacs,twocolumn,prb,aps,color,superscriptaddress]{revtex4}
2: \usepackage{graphicx}
3:
4: %\documentclass[twocolumn,preprintnumbers,amsmath,amssymb]{revtex4}
5: %\usepackage[dvips]{epsfig}
6: %\usepackage{graphicx}
7: %\usepackage{dcolumn}
8: %\usepackage{bm}
9:
10: \begin{document}
11:
12: \hsize\textwidth\columnwidth\hsize\csname @twocolumnfalse\endcsname
13:
14: \title{Transition from band insulator to Mott insulator in one dimension:\\
15: Critical behavior and phase diagram}
16: \author{Jizhong Lou}
17: \affiliation{Institute of Theoretical Physics, P. O. Box 2735, Beijing 100080, China}
18: \affiliation{Department of Physics, University of Nevada, Las Vegas, Nevada 89154}
19: \author{Shaojin Qin}
20: \affiliation{Institute of Theoretical Physics, P. O. Box 2735, Beijing 100080, China}
21: \affiliation{Department of Physics, Kyushu University, Hakozaki, Higashi-ku, Fukuoka 812-8581,
22: Japan}
23: \author{Tao Xiang}
24: \affiliation{Institute of Theoretical Physics, P. O. Box 2735, Beijing 100080, China}
25: \author{Changfeng Chen}
26: \affiliation{Department of Physics, University of Nevada, Las Vegas, Nevada 89154}
27: \author{Guang-Shan Tian}
28: \affiliation{Department of Physics, Peking University, Beijing 100087, China}
29: \author{Zhaobin Su}
30: \affiliation{Institute of Theoretical Physics, P. O. Box 2735, Beijing 100080, China}
31: \date{\today}
32:
33: \begin{abstract}
34: We report a systematic study of the transition from band insulator (BI) to Mott
35: insulator (MI) in a one-dimensional Hubbard model with an on-site Coulomb interaction
36: $U$ and an alternating periodic site potential $V$. We employ both the zero-temperature
37: density matrix renormalization group (DMRG) method to determine the gap and critical
38: behavior of the system and the finite-temperature transfer matrix renormalization
39: group (TMRG) method to evaluate the thermodynamic properties. We find two critical
40: points at $U = U_c$ and $U = U_s$ that separate the BI and MI phases for a given $V$.
41: A charge-neutral spin-singlet exciton band develops in the BI phase ($U<U_c$)
42: and drops below the band gap when $U$ exceeds a special point $U_e$. The exciton
43: gap closes at the first critical point $U_c$ while the charge and spin gaps persist
44: and coincide between $U_c<U<U_s$ where the system is dimerized. Both the charge and spin
45: gaps collapse at $U = U_s$ when the transition to the MI phase occurs. In the MI
46: phase ($U>U_s$) the charge gap increases almost linearly with $U$ while the spin gap
47: remains zero. These findings clarify earlier published results on the same model
48: and offer new insights into several important issues regarding appropriate scaling
49: analysis of DMRG data and a full physical picture for the delicate nature of the phase
50: transitions driven by electron correlation. The present work provides a comprehensive
51: understanding for the critical behavior and phase diagram for the transition from BI to
52: MI in one-dimensional correlated electron systems with a periodic alternating site potential.
53: \end{abstract}
54:
55: \draft
56: \pacs{71.30.+h, 71.10.Pm, 77.80.-e}
57: \maketitle
58:
59: \section{Introduction}
60:
61: The nature of the insulating ground state of interacting electron systems has
62: been a subject of long-standing interest and debate in condensed matter physics.
63: Because of strong quantum effects caused by spatial confinement and technical
64: advantages for theoretical treatment, one-dimensional (1D) electron systems have
65: been most extensively studied\cite{soly,hald}. Strong correlation effects in 1D
66: lead to the separation (decoupling) of charge and spin degrees of freedom.
67: Starting from a gapless phase with charge-spin separated excitations,
68: interactions can drive the system into new phases of different characteristics
69: with (i) gapful charge excitations only, (ii) gapful spin excitations only, or
70: (iii) co-existing gapful charge and spin excitations. For phases with both charge and
71: spin excitations gapful, the charge and spin degrees of freedom are rarely decoupled.
72: Despite these findings, there remain important unresolved issues regarding
73: the quantum nature of the insulating state in 1D interacting electron systems and the
74: phase transitions driven by the electron correlation. The first issue concerns the
75: establishment of an accurate phase diagram and the critical behavior near the phase
76: boundaries. Secondly, a band insulator (BI) with quasi-particle excitations typically
77: cannot be characterized by charge and spin excitations. A proper characterization scheme
78: needs to be developed. Most importantly, the nature of the correlation-driven transition
79: from BI to Mott insulator (MI) is still not fully understood.
80:
81: There has been considerable recent interest in the study of a prototype
82: one-dimensional model for ferroelectric perovskites for the understanding of the
83: response of strongly correlated electron systems with lattice distortions. These
84: efforts have raised and addressed some fundamental issues in the nature of the quantum
85: phase transition and the related critical behavior in 1D interacting electron systems.
86: Earlier works \cite{Egami,Nagaosa,Ishihara} mainly deal with the effects of strong
87: electron correlation on the electron-lattice interaction and the polarization effects
88: in the insulator. Quantum phase transitions and the characterization of the insulating
89: state are the focus of more recent work.\cite{Ort,Schonhammer,Resta,tsu99,naka,tsuc,wilk,anus,yone,capr,tori} In particular, an issue of fascinating
90: debate is the nature of the transition (or crossover) from the BI phase to the MI phases.
91: Gidopoulos {\it et al.} showed \cite{Gidopoulos} that due to the reversal of
92: inversion symmetry of the ground state from BI to MI, there is a critical point
93: for spin excitations. However, for charge excitations the critical behavior is
94: less clear. Recently, Fabrizio {\it et al.} \cite{Fabrizio} developed an
95: effective field theory for this problem and showed that there are two continuous
96: transitions from BI to MI . One is a spin transition of the Kosterlitz-Thouless type
97: at a critical point $U=U_s$, and the other is an exciton transition at an Ising
98: critical point $U=U_c<U_s$ where the exciton gap closes. Between $U_c$ and $U_s$,
99: the site-parity is spontaneously broken and the system is characterized by a doubly
100: degenerate, dimerized ground state. These results raise interesting questions about the
101: structure of the ground-state phase diagram of 1D interacting electron systems and
102: the characterization of the critical behavior near the transition points from the
103: BI phase to the MI phase.
104:
105: The model Hamiltonian for the system of interest is defined in the Hubbard formalism
106: at half-filling \cite{Ishihara,Ort,Schonhammer,tsu99,avig}
107: \begin{eqnarray}
108: H &=& \sum_{i\sigma }\left[
109: -t\left( c_{i\sigma }^{\dagger }c_{i+1\sigma }+
110: {\rm h.c.}\right) +V(-1)^in_{i\sigma }
111: \right] \nonumber \\
112: & & +U\sum_i\left( n_{i\uparrow }-\frac 12\right)
113: \left( n_{i\downarrow }- \frac 12\right) ,
114: \label{Ham}
115: \end{eqnarray}
116: where $c_{i\sigma }^{\dagger }$ and $n_{i\sigma }$ are the electron creation and
117: number operators at site $i$, $U>0$ is the on-site Coulomb repulsion, and $V$ is the
118: staggered site chemical potential. This model captures the key ingredients of
119: one-dimensional correlated insulators with mixed ionic-covalent characters, such as
120: oxide dielectric materials\cite{rest} and quasi-1D organic charge-transfer
121: complexes.\cite{torr} It incorporates covalency, ionicity, and strong electron
122: correlation.\cite{Egami,Nagaosa}
123:
124: In this paper, we present the results of extensive calculations for Hamiltonian (1)
125: using both the zero-temperature density matrix renormalization group (DMRG)
126: method \cite{White92} and the finite-temperature transfer matrix renormalization
127: group (TMRG) method \cite{Bursill,Wang,Shibata}. These methods have been demonstrated
128: to be highly accurate for 1D interacting electron systems. Our aim is to systematically
129: examine and clarify issues raised in recent work and to provide a comprehensive
130: understanding for the transition from BI to MI in one dimension. We show detailed results
131: on the gap and critical behavior. We find that a charge-neutral spin-singlet exciton
132: band forms in the BI phase and
133: drops below the band gap as $U$ increases beyond a special point $U_e$. With
134: increasing $U$ the excitons then condense and the system enters a dimerized phase,
135: followed by the closure of the quasi-particle (both spin and charge) gap when the
136: system enters the MI phase. We clarify basic concepts on charge and spin excitations
137: studied in recent work. Our results support the conclusion of
138: Fabrizio {\it et al.}\cite{Fabrizio} on the existence of two critical points for the
139: transition from BI to MI. We also present detailed results on the formation of the
140: exciton band and the scaling behavior near the critical points. In addition, we
141: carry out TMRG calculations to study thermodynamic properties to further elucidate the
142: gap and critical behavior of the system.
143:
144: \section{Low-energy excitations}
145: \label{elem}
146:
147: To properly characterize the BI and MI phases and establish the phase diagram, we need to
148: evaluate the behavior of several low-energy excitations, including the charge and spin
149: excitations and an exciton excitation that will be used to characterize the BI phase.
150: We calculate the following three excitation gaps defined on a finite 1D lattice of length
151: $L$ (chosen as an even integer) at half-filling: (i) the singlet exciton gap $\Delta_e(L)$,
152: (ii) the charge gap $\Delta _c(L)$, and (iii) the spin-triplet gap $\Delta _s(L)$,
153: \begin{eqnarray}
154: \Delta _e(L) &=&
155: E_1(\frac L2,\frac L2)-E_0(\frac L2,\frac L2),
156: \label{gape} \\
157: \Delta _c(L) &=&
158: E_0(\frac L2+1,\frac L2)+E_0(\frac L2-1,\frac L2)-
159: 2E_0(\frac L2,\frac L2), \label{gapc} \\
160: \Delta _s(L) &=&E_0(\frac L2+1,\frac L2-1)-E_0(\frac L2,\frac L2),
161: \label{gaps}
162: \end{eqnarray}
163: where $E_0(N_{\uparrow },N_{\downarrow })$ is the lowest energy of the system with
164: $N_{\uparrow }$ up and $N_{\downarrow }$ down spin electrons, and $E_1(N_{\uparrow },
165: N_{\downarrow })$ is the lowest energy of the singlet excitations. For the charge gap
166: $\Delta_c$, there is an alternative definition
167: \begin{equation}
168: \Delta _c(L)=E(\frac L2+1,\frac L2+1)-E(\frac L2,\frac L2).
169: \label{gapcc}
170: \end{equation}
171: For the model studied in this work, these two definitions on the charge gap give the same
172: results in the thermodynamic limit. This is supported by our numerical calculations.
173: However, since the numerical accuracy is much higher in calculating $\Delta_c$ by using
174: Eq. (\ref{gapc}) than by using Eq. (\ref{gapcc}), we use Eq. (\ref{gapc}) in all the
175: reported calculations.
176:
177: Although $\Delta_c$ and $\Delta_s$ are usually considered to be the charge gap and the spin
178: gap, respectively, in the literature, $\Delta_c$ in fact is the chemical potential jump for
179: particles in the system. It measures the chemical potential jump of putting a particle into
180: or taking a particle out of the system. Only when the charge and spin excitations are
181: separated and the spin gap is zero, $\Delta_c$ is equal to the charge excitation gap as,
182: for example, in the standard Hubbard model (Eq. (\ref{Ham}) with $V=0$ or $U\to\infty$).
183: When the charge and spin excitations are not separated, or when the spin gap is not zero,
184: the chemical potential jump is not equal to the charge excitation gap.
185:
186: At half-filling, the first excitation state can be either a spin singlet or a spin triplet
187: and $\Delta_e$ never exceeds $\Delta_s$. When $\Delta_e < \Delta_s$, the first excitation
188: state must be a charge-neutral spin-singlet state; otherwise, $\Delta_e$ equals $\Delta_s$
189: and measures the excitation gap of an exciton band.
190:
191: In the presence of a nonzero $V$ in the Hamiltonian, the MI phase is reached when $U$ is
192: large (including the limit $U\to\infty$). In the MI phase, it is well understood that the
193: charge gap is non-zero, but the spin gap is zero. Therefore $\Delta_c$ is the charge
194: excitation gap. Meanwhile, both $\Delta_s$ and $\Delta_e$ are zero and the gapless
195: elementary excitations in the system are spinons.\cite{ess}
196:
197: In the BI phase, all the elementary excitations are gapful. Let $u^\dag_{k\sigma}$ and
198: $d^\dag_{k\sigma}$ be the creation operators for particles in the upper conducting band
199: and for the holes in the lower valence band. The Hamiltonian for free particles
200: at $U=0$ is
201: \begin{equation}
202: H = \sum_{k\sigma} \varepsilon_k
203: \left( u^\dag_{k\sigma} u_{k\sigma} -
204: v^\dag_{k\sigma} v_{k\sigma} \right),
205: \label{freeham}
206: \end{equation}
207: where $k$ is the momentum, $\sigma$ is the spin index, and
208: $\varepsilon _k=\sqrt{V^2+4t^2\cos ^2k}$.
209: We have $u_{k\sigma}|GS\rangle=v^\dag_{k\sigma}|GS\rangle=0$ for the ground state
210: $|GS\rangle$ at half filling with $u_{k\sigma}$ being the annihilation operator for
211: electrons in the conduction band, and $v^\dag_{k\sigma}$ the creation operator for
212: electrons in the valence band. For particle-hole excitations with one particle and one hole
213: in the system, it is clear that $\Delta_c=\Delta_s=\Delta_e=2V$.
214:
215: When $U$ is small in the BI phase, the particle or hole excitations become dressed
216: quasi-particles, and $\Delta_c$ measures the chemical potential jump of the particles.
217: Since the charge and spin degrees of freedom are not separated, $\Delta_c$ is not exactly
218: the "charge gap" derived from the gapless charge-spin separated excitations in the field
219: theoretical approach. If the particles and holes are not bounded, then
220: $\Delta_c=\Delta_s=\Delta_e$. However, the particle-hole excitations may bound to form
221: excitons, and result in an exciton gap smaller than $\Delta_c$. It is confirmed by our
222: DMRG calculations shown below that $\Delta_e$ equals $\Delta_c$ when $U$ is small, but
223: becomes smaller than $\Delta_c$ when $U$ exceeds a special point $U_e$, indicating the
224: formation of singlet excitons. Meanwhile, the calculations show that $\Delta_s = \Delta_c$
225: is always observed, indicating that there is no triplet exciton formation in the BI phase.
226:
227: The DMRG calculations also show that in the BI phase the exciton gap and quasi-particle
228: excitation gap decrease and approach zero with increasing $U$. The exciton gap closes
229: at the first critical point $U_c$, and the system enters a phase where the excitons condense
230: to form dimerized ground state. At the second critical point $U_s$ the quasi-particle gap
231: also closes ($\Delta_e =\Delta_s = \Delta_c=0 $). When $U$ exceeds $U_s$, $\Delta_e$ and
232: $\Delta_s$ remain zero but $\Delta_c$ increases almost linearly with $U$.
233: Therefore, $U_s$ is the point where the quasi-particle excitation gap in the BI phase
234: collapses, and spinons in the MI phase form. This picture is consistent with that of the
235: recent field theoretical studies\cite{Fabrizio}.
236:
237: From the physical picture outlined above, it is clear that there are three special points
238: $U_e$, $U_c$, and $U_s$ along the $U$ scale. Among them
239: $U_c$ and $U_s$ are two critical points separating the BI and MI phases while $U_e$ is
240: a special point signaling the formation of the spin-singlet exciton band in the BI phase.
241: In the thermodynamic limit for a given $V$ the system can be divided into the following
242: regions: \\
243: \begin{enumerate}
244: \item $0<U<U_{e}$, $\Delta _e = \Delta_c = \Delta_s \ne 0$,
245: gapful quasi-particle excitations only.
246: \item $U_{e}<U<U_{c}$, $ 0< \Delta _e < \Delta_c = \Delta_s$,
247: gapful quasi-particle excitations coexist with singlet particle-hole
248: excitations that bound to form excitons.
249: \item $U=U_{c}$, $ \Delta _e = 0$, $\Delta_c = \Delta_s >0$,
250: the system is critical for exciton excitations.
251: \item $U_{c}<U<U_{s}$, $\Delta _e < \Delta_c = \Delta_s$,
252: excitons condense, and the system is dimerized.
253: \item $U=U_{s}$, $\Delta _e = \Delta_c = \Delta_s=0$,
254: the system is critical for quasi-particle excitations.
255: \item $U>U_{s}$, $\Delta _e = \Delta_s = 0$, $\Delta_c>0$,
256: gapless spinon excitations in the MI phase; the particle-hole
257: picture breaks down.
258: \end{enumerate}
259:
260: In the following we present the calculated results on the gap and critical behavior
261: leading to the establishment of the phase diagram. In all reported calculations,
262: free boundary conditions are used in the zero-temperature DMRG calculation. For a
263: finite size system, we can show rigorously using the variational principle that the
264: lowest energy state of the Hamiltonian (\ref {Ham}) with free boundary conditions in
265: each $(N_{\uparrow },N_{\downarrow }) $ subspace is non-degenerate except for an up-down
266: spin degeneracy\cite{Xiang}. Therefore there is no level crossing with the lowest
267: energy state in the $(N_{\uparrow },N_{\downarrow }) $ subspace and
268: $E_0(N_{\uparrow },N_{\downarrow })$ is an analytic function of $U$ and $V$. From this
269: property and Eqs. (\ref{gapc}) and (\ref {gaps}), we can further show that $\Delta _c(L)$ and
270: $\Delta _s(L)$ are also analytic functions of $U$ and $V$. In this work we focus on two
271: values of the staggered potential $V$=0.3t and 1.0t and study the behavior of the three
272: gaps introduced above in response to the on-site Coulomb repulsion $U$.
273:
274: \section{Excitation Gaps}
275: \label{energygap}
276:
277: For fermion systems, the truncation error of DMRG iterations is generally much
278: smaller than that of spin systems when the same number of optimal states are retained.
279: The efficiency of the finite system DMRG method is related to the truncation error;
280: the bigger the truncation error is, the larger the improvement of the finite lattice
281: sweeping can make. We used both finite and infinite lattice DMRG algorithms in testing
282: calculations. We find that the improvement of the ground-state energy made with the finite
283: lattice sweeping is very small when a large number of states are retained. A better way
284: to increase the accuracy of the results is using the infinity lattice approach by retaining
285: more states.
286:
287: \begin{figure}[h!]
288: \includegraphics[width=8.0cm,angle=0]{fig1_bm.eps}
289: \caption{Energy gaps vs $1/L$ for $V=1.0t$ and (a) $U=2.0 t$, (c) $U=3.0 t$,
290: (e) $U=4.0 t$. Solid lines are $\Delta_e$, dot-dashed lines are $\Delta_s$, and short-dashed
291: lines are $\Delta_c$. The long-dashed line in (c) represents the gap of the second state in
292: the singlet sector $\Delta_c^{(2)}$. The dotted lines in (a) and (c) denote the thermodynamic
293: limit of the corresponding gaps. Panels (b), (d), and (f) show the absolute values of the
294: difference of the charge and exciton gap with the spin gap, where empty circles show
295: $\Delta_c - \Delta_s$ and filled circles show $\Delta_e - \Delta_s$. 500 states are retained
296: in the DMRG calculations.}
297: \label{BDG}
298: \end{figure}
299:
300: Figure \ref{BDG} shows the behaviors of $\Delta _s$, $\Delta _c$ and $\Delta_e$ as a
301: function of $1/L$ for different $U$ at $V=1.0t$. For $V=0.3t$, similar results can be drawn
302: but the $V=1.0t$ case is more accurate because most of the features can be seen when the
303: chain length is short, while in the $V=0.3t$ case, very long chains need to be used to
304: obtain the same results. Figure \ref{BDG} (a) presents the results for $U=2.0 t$ and
305: clearly shows that the three gaps converge to the same finite value in the thermodynamic limit.
306: The difference between these gaps shown in Fig. \ref{BDG} (b) also displays this feature
307: clearly. When the chain length is short, the exciton gap is larger than the spin gap, and
308: level crossing happens at a finite chain length where the exciton gap drops lower thereafter.
309: The exciton gap decreases continually and reaches a minimum when the chain length increases
310: further; it then starts to increase and converge to the value of the spin and charge gaps.
311: For $U=3.0t$, in Fig. \ref{BDG} (c) and (d), the spin gap and charge gap still converge to the
312: same value in the thermodynamic limit, but the exciton gap goes to a different value lower
313: than the other two gaps. In short chains, the exciton gap is still larger than the spin gap,
314: but after they cross each other, the exciton gap decreases monotonically. One can also see that
315: the second state in the singlet sector also crosses the spin and charge gap and converges to
316: the lowest state when $L \to \infty$. In our calculations, we also see that more states in
317: the singlet sector cross the spin and charge gap with increasing chain length. This
318: shows that the whole spectrum of the exciton sector decreases in value and the exciton gap
319: is indeed different from the other two gaps. For the case of $U=4.0t$, shown in
320: Fig. \ref{BDG} (e) and (f), no level crossing for different chain length is detected.
321: All three gaps decrease monotonically when the chain length increases. The exciton gap
322: and the spin gap approach zero at infinite chain length, indicating that there is no gap
323: for the exciton and spin sectors. Meanwhile the charge gap approaches a finite value in
324: the thermodynamic limit.
325:
326: \begin{figure}[h!]
327: \includegraphics[width=8.0cm,angle=0]{fig2_bm.eps}
328: \caption{The spin gap $\Delta_s$ vs $1/L$ for $V=1.0t$ and $U=2.50t$. The
329: number of retained states are m=300 (empty circles),400 (empty squares),
330: 500 (filled circles), and 800 (filled squares).}
331: \label{gaplength}
332: \end{figure}
333:
334: For all the cases we have studied, $\Delta _c$ decreases monotonically with increasing $L$.
335: However, the size dependence of $\Delta _s$ is more complicated. In certain ranges of $U/t$
336: and $V/t$ close to the critical regimes, including the case shown in Fig. \ref{BDG} (a)
337: and (c), $\Delta _s$ and $\Delta_e$ vary non-monotonically and their minima are located at
338: a finite $L=L_{\min }$ rather than at $L=\infty$. In a recent work \cite{Brune}, the authors
339: studied the same model Hamiltonian (\ref{Ham}) using the DMRG method but did not observe such
340: non-monotonically behavior, and suggested that such a behavior may be due to the loss of
341: accuracy in DMRG calculations when the chain length is increased or due to some intrinsic
342: length scale for the spin degree of freedom. We have carefully examined this issue by
343: carrying out extensive scaling analysis. We demonstrate that the non-monotonical behavior
344: is not due to the lack of accuracy of the calculations, instead the behavior is a true feature
345: of Hamiltonian (\ref{Ham}) with the open boundary condition (OBC). Figure \ref{gaplength}
346: shows the chain length dependence of the spin gap for $U=2.5 t$ and $V=1.0t$ calculated by
347: retaining different numbers of optimal states $m=300, 400, 500$, and $800$. One can see that
348: the minimum occurs at $L \sim 30$; at this length the accuracy of the DMRG calculations are
349: still very high. More significantly, the results for different $m$ fall onto
350: the same curve (except for the cases of $L > 100$ and $m=300$). It shows unambiguously
351: the existence of the minimum of $\Delta_s$ in its dependence on the chain length.
352: For the exciton gap $\Delta_e$, the situation is the same. In fact the occurrence of a
353: gap minimum at a finite $L$ is not an uncommon feature for a system with incommensurate
354: low-lying excitations. It suggests that the spin excitations of the model
355: Hamiltonian (\ref{Ham}) maybe incommensurate with a characteristic wave vector defined
356: by $2\pi /L_{\min }$ (or $\pi -2\pi /L_{\min }$) in some area of the phase space.
357:
358: \begin{figure}[h!]
359: \includegraphics[width=8.0cm,angle=0]{fig3_bm.eps}
360: \caption{The difference between the spin gap and the exciton gap near $U_e$.
361: (a) V=1.0t, $U_e \sim 2.264 t$. The dashed fitting line is $0.026(U/t-0.771)(U/t-2.264)$;
362: (b) V=0.3t, $U_e \sim 1.276 t$. The dashed fitting line is $0.0282(U/t-0.584)(U/t-1.276)$.}
363: \label{exc}
364: \end{figure}
365:
366: Comparing Fig. \ref{BDG} (a) and (c), it is clear that there is a special point
367: $U_e$, where the exciton gap begins to deviate from the spin gap (for $V=1.0t$,
368: $2.0t < U_e < 3.0 t$). In both cases, all three excitations are gapful. The system is
369: in the same (BI) phase as the $U =0$ case, where $\Delta_s=\Delta_c=\Delta_e= 2V$ in
370: the thermodynamic limit. Here particle-hole excitations bound into excitons by the
371: Coulomb interaction at $U>U_e$. Fig. \ref{exc} shows the difference $\Delta_s - \Delta_e$
372: for $V=1.0t$ and $0.3t$. The fitting to DMRG results gives the critical value $U_e=2.264 t$
373: for $V=1.0t$ and $U_e=1.276t$ for $V=0.3t$.
374:
375: For finite $L$ we find that $\Delta _c$ is always larger than $\Delta _s$. In the MI phase,
376: $\Delta _c$ is finite but $\Delta _s$ approaches zero in the thermodynamic limit.
377: In the BI phase, $\Delta _s\ $and $\Delta _c$ always approach the same value
378: in the thermodynamic limit. This can be seen either from the asymptotic behaviors of
379: $\Delta _s$ and $\Delta _c$ in the limit $L\rightarrow \infty $ [Fig. \ref{BDG} (a) and
380: (c)] or from the $1/L$ dependence of the difference $\Delta _c-\Delta _s$
381: (Fig. \ref{BDG} (b) and (d)). For all the cases we have studied, we find that
382: $\Delta _c-\Delta _s$ drops monotonically and approaches zero in the limit $1/L\rightarrow 0$
383: even when $\Delta _s$ changes non-monotonically. For a given $V$, this result holds from
384: $U=0$ up to a critical regime where both $\Delta _c$ and $\Delta _s$
385: become smaller than truncation errors. It suggests that $\Delta _c$ and $\Delta _s$ are
386: equal in the thermodynamic limit in the entire BI phase.
387:
388: When $\Delta _s$ changes
389: non-monotonically with $L$, the extrapolation for the spin gap in the limit
390: $1/L\rightarrow 0$ becomes subtle. If the data with $L<L_{\min }$ are used in the
391: extrapolation, the extrapolated value of $ \Delta _s$ will certainly be smaller than
392: the true value. However, if the data with $L>L_{\min }$ is used but $L$ is still not large
393: enough to reach the regime where $\Delta _s$ begins to saturate, the extrapolated value of
394: $ \Delta _s$ will be larger than the true value (this seems to be the case in
395: Ref. \onlinecite{Takada}). For the data shown in Fig. \ref {BDG}(a) and (c),
396: these two kinds of extrapolations result in $\Delta _c>$ $\Delta _s$
397: and $\Delta _c<$ $\Delta _s$, respectively. Both are incorrect. To correctly extrapolate
398: $\Delta _s$ in the limit $1/L\rightarrow 0$, data with $L$ much larger than $L_{\min }$
399: must be used.
400:
401: For the $U=4.0 t$ case shown in Fig. \ref{BDG} (e), $\Delta _c$ is finite
402: but $\Delta _s$ becomes zero in the thermodynamic limit. This indicates that the system
403: is in the same phase as that for $U \to \infty$, namely the MI phase. In the MI phase, the
404: chain length dependence of the three gaps are monotonical; in addition, there is no
405: crossing between $\Delta_s$ and $\Delta_e$ when the chain length varies. In the
406: $L \to \infty$ limit the spin gap $\Delta_s$ is zero in the MI phase suggesting that
407: there is a critical point that separates the MI phase from the BI phase.
408: At this critical point $U_s$, the spin gap vanishes. Because the charge gap is
409: equal to the spin gap in the BI phase, the charge gap will also vanish at the same
410: point $U_s$. However, when $U$ increases further, the charge gap increases with $U$
411: while the spin gap remains zero. Considering that the exciton gap is lower than the spin
412: gap in the BI phase at $U > U_e$, $\Delta_e$ may vanish before the spin gap and charge gap
413: do. In that case, there should be another critical point $U_c$ signaling the collapse of
414: the exciton gap.
415:
416: \begin{figure}[h!]
417: \includegraphics[width=8.0cm,angle=0]{fig4_bm.eps}
418: \caption{The spin (empty circles), charge (solid circles) and exciton (empty squares) gaps
419: as a function of $U$ for $V =0.3t$ (a) and $V=1.0t$ (c). Panels (b) and (d) show the $U$
420: dependence of the spin and exciton gaps in the vicinity of the critical region for (a)
421: and (c) respectively. At $U_e$, the exciton gap deviates from the other two gaps and it
422: collapses first at $U_c \sim 2.225t$ for $V=0.3t$ and $\sim 3.675t$ for $V=1.0t$. The charge
423: and spin gaps collapse at $U_s \sim 2.265t$ for $V=0.3t$ and $\sim 3.71t$ for $V=1.0t$.
424: For $0<U<U_s$ the spin and charge gaps have the same value in the thermodynamic limit.
425: When $U>U_s$, the spin and exciton gap are zero while the charge gap is finite.}
426: \label{bdu}
427: \end{figure}
428:
429: In Fig. \ref{bdu}, we show the $U$ dependence of the three gaps for $V=0.3t$
430: [(a) and (b)] and $V=1.0t$ [(c) and (d)]. For both cases, there are indeed two critical
431: points $U_c$ and $U_s$ although they are very close. When $U<U_e$,
432: $\Delta_e = \Delta_s = \Delta_c$, and the three gaps decrease almost linearly
433: with increasing $U$. At $U_e$, the exciton gap splits off and drops below the other two gaps.
434: At the critical point $U_c$, the exciton gap collapses while the spin and charge gaps still
435: coincide and remain finite until the second critical point $U_s$ where they both collapse.
436: At $U > U_s$, the charge gap increases with increasing $U$ while the spin gap and the
437: exciton gap remain zero in the thermodynamic limit. Although the accuracy of our
438: DMRG calculations do not allow a direct assessment of the behavior of the exciton gap for
439: $U_c < U < U_s$, we believe that $\Delta_e$ is finite in this region (see more detailed
440: discussion on this point in the following section). The extrapolation of the gap behavior
441: leads to $U_c \sim 2.225 t$ and $U_s \sim 2.265 t$ for $V=0.3 t$, while $U_c \sim 3.675 t$,
442: $U_s \sim 3.71 t$ for $V=1.0 t$.
443:
444: It is clear that the $U$ dependence of the three gaps is similar for $V=0.3t$ and $V=1.0t$.
445: It is expected that the same picture is valid for all $V$. Furthermore, $U_e$, $U_c$, and
446: $U_s$ all approach the same point $U_{\infty}=2V$ in the $V \rightarrow \infty$ limit.
447:
448: \section{Critical behavior}
449: \label{critbeh}
450:
451: \begin{figure}[h!]
452: \includegraphics[width=8.0cm,angle=0]{fig5_bm.eps}
453: \caption{The behavior of the charge gap in the vicinity of $U_s$ for $V=0.3t$. (a) The
454: chain length dependence of the charge gap at $U=2.2t$ for different numbers of optimal
455: states retained: $m$=200 (filled circles), 250 (empty squares), 300 (empty circles) and 400
456: (filled squares); (b) The dependence of the charge gap on the number of the retained states
457: at L=300 with $U=2.2t$; (c) The charge gap for $m \to \infty$ with different chain lengths:
458: $L=$ 300 (empty circles), 400 (filled circles), 500 (empty squares) and 600 (filled circles)
459: in the vicinity of the critical region; the fitting lines using eq. (\ref{gapc2}) are also
460: shown; (d) The chain length dependence of $\Delta_{c,min} (L)$; (e) The chain length
461: dependence of $U_{s,L}/t$; (f) The chain length dependence of $\alpha_{c,L}$.}
462: \label{cgcritical}
463: \end{figure}
464:
465: From the gap behavior presented in the previous section, it is clear that there are two
466: critical points $U_c$ and $U_s$ for Hamiltonian (\ref{Ham}) for a given $V$. It is
467: important to study the detailed critical behavior near the critical points for the
468: understanding of the nature of the BI-to-MI transition. In the following we study the critical
469: behavior of the system by examining (i) the evolution of the gap behavior near the critical
470: points and (ii) the behavior of the ground-state energy of the system.
471:
472: \subsection{Analysis of the gap behavior}
473:
474: The $U$ dependence of the gaps shows that the charge instability occurs at the same point
475: as that for the spin transition. To determine the critical points for the charge and spin
476: excitations, we examine when $\Delta _c$ and $\Delta _s $ become zero in the
477: limit $L\rightarrow \infty $. Since the numerical errors are larger than the magnitude of
478: $\Delta _c$ or $\Delta _s$ in the vicinity of the critical points, it is difficult to
479: determine accurately the critical behavior simply from the values of the energy gaps.
480: To resolve this issue, we analyze the scaling behavior of $\Delta_c(L)$ around its minimum
481: with respect to $U$. However, the DMRG results of $\Delta_c(L)$ depend the number of
482: states $m$ retained during the iterations. When the chain length is long enough, the
483: difference due to retaining different number of states show clearly.
484: In Fig. \ref{cgcritical} (a), we show the chain length dependence of the charge gap
485: at $U=2.22t$ and $V=0.3t$ by keeping different $m$. The difference
486: is obvious. This problem can be solved by employing the extrapolation in the limit of
487: $m\rightarrow \infty $. For given $U$ and $V$, the charge gap at chain length $L$ and
488: by keeping $m$ states is $\Delta_c (m, L)$. By extrapolating to the infinite $m$ limit,
489: more accurate result of the charge gap at chain length $L$ can be obtained,
490: \begin{equation}
491: \Delta_c(\infty, L) = \lim_{m \rightarrow \infty} \Delta_c (m,L).
492: \end{equation}
493: In Fig. \ref{cgcritical} (b), we show the $1/m$ dependence and the extrapolation procedure
494: of the $\Delta_c (m,L)$ for L=300 at $U=2.22t$ and $V=0.3t$. The extrapolated result
495: $\Delta_c (\infty, L)$ is considered the exact charge gap $\Delta_c (L)$ at chain length $L$.
496:
497: The extrapolated charge gap $\Delta_c(\infty, L)$ is a function of $U$ and chain length
498: $L$ at a given $V$. By applying the same procedure shown in Fig. \ref{cgcritical} (b), we
499: obtain $\Delta_c(\infty, L)$ for different values of $U$ near the critical point $U_s$ for
500: a serial of selected $L$. In Fig. \ref{cgcritical} (c), we show the $U$ dependence of
501: $\Delta_c(L)$ ($\Delta_c(\infty, L)$) in the vicinity of the critical point $U_s$ at chain
502: length $L=$ 300, 400, 500, and 600, respectively, for $V$=0.3t. A gap minimum at finite
503: chain length is clearly seen. Assuming $\Delta _{c,\min }(L)$ to be the minimum of
504: $\Delta _c(L)$ located at $U_{s,L}$, then around this minimum we can expand
505: $\Delta _c(L)$ to the leading order of the parameter $u=U-U_{s,L}$ as
506: \begin{equation}
507: \Delta _c(L)=\Delta _{c,\min }(L)+\alpha _c\left( L\right) u^2+O\left(
508: u^3\right) . \label{gapc2}
509: \end{equation}
510: Since $\Delta _c(L)$ is an analytic function of $U$, both $\Delta _{c,\min }(L)$ and
511: $\alpha _c\left( L\right) $ should be finite. The critical behavior of the charge
512: excitations is determined by the properties of $ \Delta _{c,\min }(L)$ and
513: $\alpha _c\left( L\right) $ in the limit $ L\rightarrow \infty $.
514: If $\Delta _{c,\min }(L)\rightarrow 0$ in the limit $ L\rightarrow \infty $, the charge
515: excitation is critical at $ U_s=U_{s,\infty }$, which would be consistent with the discussion
516: in the previous section. However, if $\Delta _{c,\min }$ remains finite
517: in the limit $L\rightarrow \infty $, then there is no critical point
518: for charge excitations and the ground state is insulating in the entire parameter space.
519:
520: Figure \ref{cgcritical} (d) shows the calculated $\Delta _{c,\min }(L)$ as a function of $1/L$.
521: The solid curve is a least-square fit of the data and given by
522: $\Delta _{c,\min }(L)\approx 17.894/L-729.785/L^2$. Within numerical
523: errors, we find that $ \Delta _{c,\min }(L)$ is indeed $0$ in the limit $1/L\rightarrow 0$.
524: Figure \ref{cgcritical} (e) shows the $L$ dependence of $U_{s,L}$. It changes almost
525: linearly with $1/L$. Within numerical errors, we find that the data of $U_{s,L}$ are well
526: fitted by $U_{s,L}/t \approx 2.265-22.532/L$. Thus the critical point is at $U_c /t =2.265t$,
527: in full agreement with the value obtained by fitting the gap directly in the previous section.
528:
529: We now turn to the critical behavior of $\alpha _c\left( L\right) $.
530: Figure \ref{cgcritical} (f) shows the $1/L$
531: dependence of $\alpha _c\left( L\right) $ for the case $V=0.3t$.
532: The fitting curve (solid line) is given by $ -0.323+0.0091L$.
533: The divergence of $\alpha _c\left( L\right) $ suggests that
534: the derivative of $\Delta _c(L)$ is singular at $U_s$ and
535: the leading term in $\Delta _c$ in the thermodynamic limit
536: is linear rather than quadratical in $u$, i.e., $\Delta _c(U)\sim \left| U-U_s\right| $.
537:
538: \subsection{Analysis of the ground-state energy}
539: \label{gseng}
540:
541: The ground-state energy as the zero-temperature free energy can also provide evidence
542: for the critical behavior. However, singularities in the ground-state energy are of higher
543: order derivatives with respect to the model parameter for continuous phase transitions.
544: As a result, evidence for critical behavior derived from the ground-state energy
545: is not as strong as that from the gap behavior, this despite the higher accuracy of the
546: ground-state energy than that for the gap in the DMRG calculations.
547:
548: We have calculated the ground-state energy in the critical region by retaining
549: $m$=800 states and up to chain length $L$=1000. For Hamiltonian (\ref{Ham}), with open
550: boundary conditions, the ground-state energy per site $e_0(L)$ satisfies
551: \begin{equation}
552: e_0(L) = \frac{E_0(L)}{L} = \epsilon_0 + \frac{e_b}{L} + \frac{c}{L^2}
553: + O(\frac{1}{L^3}),
554: \end{equation}
555: here $E_0(L)$ is the ground-state energy for a chain of length $L$, $\epsilon_0$ the
556: ground-state energy per site for $L \to \infty$, $e_b$ is the boundary energy
557: (surface energy) due to the free boundary condition, and $c=v \pi$ where $v$ is the spin
558: wave velocity. When $U > U_c$, the system is gapful and $c$ should approach zero when
559: the chain length is much larger than the correlation length $\xi$.
560:
561: \begin{figure}[h!]
562: \includegraphics[width=8.0cm,angle=0]{fig6_bm.eps}
563: \caption{(a), (b), (c): The $U$ dependence of $\epsilon_0$, $e_b$ and $c$ obtained from
564: fitting the $L>200$ data of the ground-state energy; (d) The chain length dependence of
565: c(L) for $U=1.5t$ (dashed line), $U=2.0t$ (dot-dashed line), and $U=2.2t$ (solid line);
566: (e) $A = 10^6 (\epsilon_0(L) - \epsilon_0(100))$ for L=120, 150, 200, 250, 300, and 400
567: (from bottom to top on the $A > 0$ side); (f) $B = 10^3 (e_b(L) - e_b(100))$ for the same
568: chain lengths as in (e) (from top to bottom on the $B < 0$ side).}
569: \label{gse}
570: \end{figure}
571:
572: The $V=0.3 t$, $L> 200$ ground-state energy results are fitted directly by
573: \begin{equation}
574: e_0(L) = \epsilon_0 + e_b/ L + c/L^2,
575: \label{gsfit}
576: \end{equation}
577: and the obtained results are shown in Fig. \ref{gse} (a), (b) and (c).
578: $\epsilon_0$ and $e_b$ are analytic functions of $U$.
579: For $c$, Fig. \ref{gse}(c) shows that it is not only non-zero for $U < U_c$
580: but also has a fairly large value. This is due to the finite chain length effect.
581: The value of $c$ depends very sensitively on the chain length range used for fitting.
582:
583: To analyze the chain length dependence of $c$, we fit the ground-state energy
584: for $L-20$, $L$, and $L+20$ using Eq. (\ref{gsfit}) and vary $L$ from 24 to 980. Each
585: fitting gives the exact result of $\epsilon_0(L)$, $e_b(L)$ and $c(L)$. In
586: Fig. \ref{gse}(d), we show the $L$ dependence of $c(L)$ for $U=1.5 t$, $U=2.0t$
587: and $U=2.2t$. It is clear that for $U=1.5t$ and $2.0t$, $c(L)$ becomes zero when
588: $L \rightarrow \infty$. The result of $U=2.0t$ shows a minimum at $L \sim 90$, and
589: the $c(L)$ begins to approach zero after the minimum. For $U=2.2t$, a minimum is also
590: clearly seen. A comparison of these results leads to the conclusion that for $L \rightarrow
591: \infty$, $c(L)$ will approach zero. The largest chain length used in the calculation
592: is not long enough to obtain correct $c(L)$ values. However, the obtained $\epsilon_0(L)$ and
593: $e_b(L)$ may display the critical behavior of $U_c$ and $U_s$. In Fig. \ref{gse}
594: (e) and (f), we show the results of $\epsilon_0 (L) -\epsilon_0(L_0)$ and
595: $e_b(L) - e_b(L_0)$ for $L_0=100$. These results show the existence of the
596: critical point $U_c$. $L_0$ can be viewed as a characteristic length of the critical region.
597:
598: From the $U$ dependence of $\epsilon_0$, it is possible to examine the type of transition
599: at the critical points. A problem is that the fitting results shown in Fig. \ref{gse}(a)
600: include extra errors induced by the fitting method. To avoid this, we analyze the $U$
601: dependence of the ground-state energy using a different approach.
602:
603: \begin{figure}[h!]
604: \includegraphics[width=8.0cm,angle=0]{fig7_bm.eps}
605: \caption{(a) ${\it d}^2 e_0/{\it d} U^2$ versus $U$ for $L=500$ (filled circles),
606: 600 (filled squares), 800 (empty circles), 1000 (empty squares). The solid line is an
607: extrapolation of the data to the limit $L \to \infty$. Inset: Enlarged figure in the
608: vicinity of the critical point $U_s$. (b) The dependence of the ${\it d}^2 e_0/{\it d} U^2$
609: with $\ln(1/L)$ at $U=2.21t$ (empty squares), $2.225t$ (filled circles), $2.27t$ (empty
610: circles). The solid fitting -0.00997+0.009$\ln(1/L)$ is the least square fitting for
611: the $U=2.225t$ case.}
612: \label{deri}
613: \end{figure}
614:
615: At a finite chain length $L$, the ground-state energy per site $e_0(L)$ is also a function
616: of $U$. Here $e_0(L)$ contains only the errors from the DMRG truncation. We can examine
617: the derivatives of $e_0(L)$ with respect to $U$ and analyze their chain length dependence.
618: In Fig. \ref{deri}(a), we show the second derivative of $e_0(L)$ with $U$ for $L$=500,
619: 600, 800, and 1000. At each chain length, there is a minimum near $U_c$ in the $U$
620: dependence of the second derivative. When the chain length increases , the position of the
621: minimum moves towards larger $U$ and approaches
622: $U_c$ which is the critical point in the thermodynamic limit; meanwhile, the
623: shape of the minimum becomes sharper. Fig. \ref{deri}(b) shows the chain length dependence
624: of the second derivative. It is clear that for $U=2.225t \sim U_c$, the second derivative
625: diverges logarithmically with the chain length, but for other values of $U$ the second
626: derivative does not diverge. These observations suggest that the phase transition at $U_c$ is
627: of the second order. No singularity is found in the first derivative or the second
628: derivative near the critical point $U_s$. This means that the transition at $U_s$ is higher
629: than second order. These results are consistent with those reported by
630: Fabrizio {\it et al.}\cite{Fabrizio}.
631:
632: \section{TMRG study of spin susceptibility and specific heat}
633: \label{secdmrg}
634:
635: To gain more insight into the physics of the BI-to-MI transition, we have studied the
636: thermodynamic properties of the model using the TMRG method \cite{Bursill,Wang,Shibata}
637: which is implemented in the thermodynamic limit and can evaluate very accurately the
638: thermodynamic quantities at low temperature for quasi-1D systems. In our calculations,
639: we kept 250 optimal states. The calculated specific heat $C_v$, charge
640: susceptibility $\chi _c$, and spin susceptibility $\chi _s$ for $ U/t= 1.0, 2.25, 5.0$ and
641: $V/t=0.3$ as a function of temperature are shown in Fig. \ref{tmrg}.
642:
643: \begin{figure}[h!]
644: \includegraphics[width=8.0cm,angle=0]{fig8_bm.eps}
645: \caption{Temperature dependences of (a) the spin susceptibility $\chi_s$, (b) the charge
646: susceptibility $\chi_c$, and (c) the specific heat $C_v$. ($t$ is set to 1).}
647: \label{tmrg}
648: \end{figure}
649:
650: We find that $\chi _c$ decreases exponentially at low temperatures in both the BI and MI
651: phases, while $\chi _s$ shows activated behavior only in the BI phase. In the MI phase,
652: there are two broad peaks in $C_v$, probably due to the charge-spin separation. Near the
653: critical point, U=2.25t, since both the charge and spin energy gaps are very small, the
654: exponential decays in both $\chi_c$ and $\chi_s$ show up only at very low temperatures.
655: These results support the conclusions of the DMRG calculations presented in previous sections.
656:
657: \section{Phase Diagram}
658: \label{phasediag}
659:
660: The overall $U$ dependence of the charge, spin and exciton gaps shown
661: in Fig. \ref{bdu} give a lot of information on the phase diagram
662: of Hamiltonian (\ref{Ham}). The charge and spin gaps coincide in the BI phase.
663: Above $U_s$, $\Delta _c$ increases with $U$ but $ \Delta _s$ remains zero.
664: The exciton gap $\Delta_e$ collapses at $U_c < U_s$, and when $U > U_s$, the exciton gap
665: should also be zero. However, from Fig. \ref{bdu} it is unclear whether the exciton
666: excitations are gapful or gapless in the regime between the two critical points,
667: $U_c < U < U_s$. Even if the exciton excitations are gapful in this regime, the gap
668: would be too small to detect numerically.
669:
670: \begin{figure}[h!]
671: \includegraphics[width=8.0cm,angle=0]{fig9_bm.eps}
672: \caption{(a) Chain length dependence of the dimerization order parameter ${\cal D}$ for
673: $U=2.21t$ (empty squares), $U=2.25t$ (filled circles) and $U=2.29t$ (empty circles).
674: The straight fitting line is $0.0499+3.886/L$. (b) ${\cal D}(L)$ in the critical regime
675: for $L$=200 (empty circles), 300 (filled circles), 400 (empty squares), and 500 (filled
676: squares). The dotted lines indicate the two critical points.}
677: \label{dimerfig}
678: \end{figure}
679:
680: When the exciton gap collapses, the excitons can condense into the ground state \cite{aff}.
681: In this case, the system is expected to be dimerized\cite{Fabrizio}. Here we evaluate
682: the dimerization order parameter
683: \begin{equation}
684: {\cal D} = \frac{1}{L} \sum_{i \sigma} (-1)^i (c^\dagger_{i\sigma}
685: c_{i+1\sigma} + h.c.).
686: \label{dimereq}
687: \end{equation}
688: Figure \ref{dimerfig}(a) shows the chain length dependence of the dimerization operator
689: for different $U$. It is clear that for $U=2.21t$, when $L \rightarrow \infty$,
690: ${\cal D}$ approaches zero. At $U=2.29t$, ${\cal D}$ just starts to fall at the largest
691: chain length we studied; it is expected that it will approach zero as the chain length is
692: long enough. For $U=2.25t$, which is between the two critical points, it seems that
693: ${\cal D}$ will diverge to a nonzero constant. For a large range of chain lengths, the
694: results can be well fitted by a straight line shown in Fig. \ref{dimerfig} (a).
695: The dependence of the finite chain dimerization ${\cal D}(L)$ on $U$ is shown in
696: Fig. \ref{dimerfig}(b) for $L$=200, 300, 400, and 500. These results indicate that in
697: the thermodynamic limit, the ground states are dimerized when $U_c < U < U_s$.
698:
699: The dimerization of the ground state for $U_c < U < U_s$ suggests that the exciton
700: excitations are gapful in this region. So the physical picture on the exciton excitation
701: is emerging: the exciton gap formed in the BI phase collapses at the critical point $U_c$;
702: with further increasing $U$, the (small) exciton gap will first increase, reach a maximum
703: and then decrease and collapse again at $U_s$; at $U > U_s$, the exciton excitations
704: remain critical.
705:
706: \begin{figure}[h!]
707: \includegraphics[width=8.0cm,angle=0]{fig10_bm.eps}
708: \caption{The ground-state phase diagram for Hamiltonian (\ref{Ham}). The empty circles
709: denote the DMRG results. The curves of $U_s$ and $U_c$ are shown with solid lines that are
710: very close to each other. The curve of $U_e$ is shown by the dashed line which does not
711: indicate a phase transition line, but rather denotes a serial of special points. The dotted
712: line is $U=2V$ which is the limit for $V \to \infty$.}
713: \label{phase}
714: \end{figure}
715:
716: When $U\ll V$, first order perturbation leads to
717: \begin{equation}
718: \Delta _s(U)=\Delta _c(U)\approx V-cU,
719: \label{deltas}
720: \end{equation}
721: where $c=V\int_0^\pi d\ k (\sqrt{2}\pi {\varepsilon }_k)^{-1} $ is a constant determined
722: by the single-particle energy dispersion $ \varepsilon _k=\sqrt{V^2+4t^2\cos ^2k}$.
723: Since $\Delta _c(U)$ drops almost linearly with $U/t$ in the BI phase, we can estimate the
724: value of $U_s$ from Eq. (\ref{deltas}) as $U_s\approx V/c$. In the limit $V \to 0$, we have
725: \begin{equation}
726: V/t \approx c_1e^{-c_2t/U_s}, ~~~~or ~~~~~
727: U_s/t \approx -{\frac{c_2}{\log (V/c_1 t)}}\ ,
728: \label{smallv}
729: \end{equation}
730: where $c_1$ and $c_2$ are two constants of order one. Figure \ref{phase} shows the
731: ground-state phase diagram for Hamiltonian (\ref {Ham}). The curve for $U_s$ and $V <1.0t$
732: is obtained from Eq. (\ref{smallv}). The parameters $c_1$ and $c_2$ are fixed
733: by the two $U_s$ values for $V=0.3t$ and $V=1.0t$. When $V\rightarrow 0$, $U_s$ goes to
734: zero but the ratio $U_s/V$ diverges. In the limit $U/t\rightarrow \infty $, $U_s$ is very
735: close to $2V$. The difference between $U_s$ and $2V$ is of order $t$: $U_s-2V\sim t$.
736:
737: \section{Summary}
738: \label{sumdiscus}
739:
740: We have carried out systematic studies using the DMRG and TMRG methods to examine
741: the critical behavior of a one-dimensional Hubbard model with an alternating site potential
742: in the transition from band insulator to Mott insulator. Based on extensive numerical
743: calculations and analytic analysis, we have clarified several important issues raised in
744: recent works and have established the ground-state phase diagram.
745: We have identified two critical points, $U_c$ and $U_s$, that separate the BI and MI phases.
746: When $U>U_s$, the system is in the MI phase where the charge excitations are massive but
747: the spin excitations are critical. When $U<U_c$, the system behaves like a classic band
748: insulator: the charge and spin excitation gaps coincide and a charge-neutral spin-singlet
749: exciton band forms below the band gap when $U$ exceeds a special point $U_e$. Between the
750: two critical points, excitons condense and the ground state is dimerized. These results are
751: consistent with the conclusions of a recent field theoretical study of the same model.
752: The present work provides a detailed account for the critical behavior in the BI-to-MI
753: transition in one dimension for correlated electron systems and establishes a good
754: understanding for its ground-state phase diagram.
755:
756: \begin{acknowledgments}
757: We thank Y. L. Liu, R. Noack , and M. Fabrizio for helpful discussions and M. Tsuchiizu for
758: bringing to our attention Refs. \onlinecite{tsu99} and \onlinecite{tsuc}. This work was
759: supported in part by the Department of Energy at the University of Nevada, Las Vegas, the NSF
760: of China and the Special Funds for Major State Basic research Projects of China.
761: \end{acknowledgments}
762:
763: \begin{references}
764:
765: \bibitem{soly} J. Solyom, Adv. Phys. {\bf 28}, 201 (1979).
766: \bibitem{hald} F. D. M. Haldane, J. Phys. C {\bf 14}, 2585 (1981).
767: \bibitem{Egami}T. Egami, S. Ishihara, and M. Tachiki,
768: {\sl Science} {\bf 261}, 1307 (1993).
769: \bibitem{Nagaosa} N. Nagaosa and J.Takimoto
770: J. Phys. Soc. Jpn. {\bf 55}, 2735 (1986); {\bf 55}, 2745 (1986);
771: N. Nagaosa, {\sl ibid.} {\bf 55}, 2754 (1986); {\bf 55}, 3488 (1986).
772: \bibitem{Ishihara} S. Ishihara, T. Egami, and M. Tachiki,
773: Phys. Rev. B {\bf 49}, 8944 (1994); S. Ishihara, M. Tachiki, and T. Egami,
774: {\sl ibid.} {\bf 49}, 16123 (1994).
775: \bibitem{Ort} G. Ortiz and R. Martin, Phys. Rev. B {\bf 49}, 14202 (1994);
776: G. Ortiz, P. Ordej\'{o}n, R. Martin, and G. Chiappe,
777: {\sl ibid.} {\bf 54}, 13515 (1996).
778: \bibitem{Schonhammer} K. Sch\"{o}nhammer, O. Gunnarson and R.M. Noack,
779: Phys. Rev. B {\bf 52}, 2504 (1995).
780: \bibitem{Resta} R. Resta, and S. Sorella, Phys. Rev. Lett. {\bf 74}, 4738 (1995);
781: {\sl ibid.} {\bf 82}, 370 (1999).
782: \bibitem{tsu99} M. Tsuchiizu and Y. Suzumura, J. Phys. Soc. Jpn. {\bf 68}, 3966 (1999).
783: \bibitem{naka} M. Nakamura, \prb {\bf 61}, 16377 (2000).
784: \bibitem{tsuc} M. Tsuchiizu and A. Furusaki, \prl {\bf 88}, 56402 (2002);
785: M. Tsuchiizu and Y. Suzumura, J. Phys. Soc. Jpn. {\bf 68}, 3966 (1999).
786: \bibitem{wilk} T. Wilkens and R. M. Martin, \prb {\bf 63}, 235108 (2001).
787: \bibitem{anus} Y. Anusooya-Pati and Z. G. Soos, \prb {\bf 63}, 205118 (2001).
788: \bibitem{yone} K. Yonemitsu, \prb {\bf 65}, 85105 (2002).
789: \bibitem{capr} S. Caprara, M. Avignon, O. Navarro, \prb {\bf 61}, 15667 (2000).
790: \bibitem{tori} M. E. Torio, A. A. Aligia, and H. A. Ceccatto,
791: \prb {\bf 64}, 121105 (2001).
792: \bibitem{Gidopoulos} N. Gidopoulos, S. Sorella and E. Tosatti,
793: Eur. Phys. J. B {\bf 14}, 217 (2000).
794: \bibitem{Fabrizio} M. Fabrizio, A. O. Gogolin, and A. A. Nersesyan,
795: Phys. Rev. Lett. {\bf 83}, 2014 (1999).
796: \bibitem{avig} M. Avignon, C. A. Balseiro, C. R. Proetto, and Alascio,
797: \prb {\bf 33}, 205 (1986).
798: \bibitem{rest} R. Resta, Rev. Mod. Phys. {\bf 66}, 899 (1994).
799: \bibitem{torr} J. B. Torrance, J. E. Vazques, J. J. Mayerle, and
800: V. Y. Lee, \prl {\bf 46}, 253 (1981); Y. Anusooya-Pati and Z. G. Soos,
801: Phys. Rev. B {\bf 63}, 205118 (2001); K. Yonemitsu, Phys. Rev. B {\bf 65},
802: 085105 (2002).
803: \bibitem{White92} S. R. White, Phys. Rev. Lett. {\bf 69}, 2863 (1992);
804: Phys. Rev. B {\bf 48}, 10345 (1993).
805: \bibitem{Bursill} R. J. Bursill, T. Xiang, and G. A. Gehring,
806: J. Phys. Cond. Mat. {\bf 8}, L583 (1996).
807: \bibitem{Wang} X. Wang and T. Xiang, Phys. Rev. B {\bf 56}, 5061 (1997).
808: \bibitem{Shibata} N. Shibata, J. Phys. Soc. Jpn. {\bf 66}, 2221 (1997).
809: \bibitem{ess} F. H. L. Essler, V. E. Korepin, and K. Schoutens,
810: Nucl. Phys. B {\bf 384}, 431 (1992).
811: \bibitem{Takada} Y. Takada and M. Kido, J. Phys. Soc. Jpn. {\bf 70}, 21 (2001).
812: \bibitem{Xiang} T. Xiang and N. d'Ambrumenil, Phys. Rev. B {\bf 46}, 11179 (1992).
813: \bibitem{Brune} Ph. Brune, G. I. Japaridze, A. P. Kampf, and M.Sekania, cond-mat/0106007.
814: \bibitem{aff} I. Affleck, Phys. Rev. B {\bf 43}, 3215 (1991).
815: \end{references}
816:
817: \end{document}
818: