1: \documentclass[twocolumn,prb,floatfix,amssymb,aps]{revtex4}
2: %\documentclass[preprint]{revtex4}
3: \usepackage{graphicx}
4:
5: \begin{document}
6:
7: \hsize\textwidth\columnwidth\hsize\csname@twocolumnfalse\endcsname
8:
9: \title{Insulating charge density wave for a half-filled SU(N)
10: Hubbard model with an attractive on-site interaction in one
11: dimension}
12: \author{Jize Zhao and Kazuo Ueda}
13: \affiliation{Institute for Solid State Physics, University of
14: Tokyo, Kashiwa, Chiba 277-8581, Japan}
15: \author{Xiaoqun Wang}
16: \affiliation{Department of Physics, Renmin University of China,
17: Beijing 100872, China}
18: \date{\today}
19: \begin{abstract}
20: We study a one-dimensional SU(N) Hubbard model with an attractive
21: on-site interaction and $N>2$ at half-filling on the bipartite
22: lattice using density-matrix renormalization-group method and a
23: perturbation theory. We find that the ground state of the SU(N)
24: Hubbard model is a charge density wave state with two-fold
25: degeneracy. All the excitations are found to be gapful, resulting
26: in an insulating ground state, on contrary to that in the SU(2)
27: case. Moreover, the charge gap is equal to the Cooperon gap, which
28: behaves as $-2Nt^2/(N-1)U$ in the strong coupling regime. However,
29: the spin gap $\Delta_{s}$ and the quasiparticle gap $\Delta_{1}$
30: as well open exponentially in the weak coupling region, while in
31: the strong coupling region, they linearly depend on $U$ such that
32: $\Delta_{s}\sim -U(N-1)$ and $\Delta_{1}\sim -U(N-1)/2$.
33: \end{abstract}
34:
35: \pacs{71.10.Fd, 71.10.Pm, 71.30.+h}
36: \maketitle
37:
38: \section{INTRODUCTION}
39: Correlation effects in electronic systems have been of long-term
40: interest in the condensed matter physics. In recent years,
41: important progress has been made experimentally in ultra-cold atomic
42: systems where strong correlation leads to some novel physical
43: phenomena. In particular, interacting fermionic atoms can be trapped
44: in an optical lattice\cite{kohl05,matrin}. More interestingly, the interaction
45: in the ultra-cold atomic systems is tunable through the Feshbach resonance, which
46: allows for a full exploration of various fundamental properties of
47: strongly correlated models. Moreover, the nuclear
48: spin of the atoms can be larger than the electronic spin.
49: One could expect richer physics
50: induced by the degree of freedom of higher spins\cite{HO1, WU1, WU2, TU1}.
51: Therefore, one would like naturally to generalize the Hubbard model
52: with two spin components to the one with $N$-components.
53:
54: In this paper, we investigate low energy properties of a one-dimensional
55: half-filled SU(N) Hubbard model\cite{AFFLECK1} with an attractive on-site
56: interaction and $N>2$. This is a generalization of our previous
57: work in which we have focused on the SU(4) case \cite{ZHAO1}.
58: The Hamiltonian of the one-dimensional SU(N) Hubbard model is
59: represented by
60: \begin{equation}
61: \mathcal{H}=-t\sum_{i=1}^{L}\sum_{\sigma}(\hat{c}^{\dagger}_{i\sigma}\hat{c}_{i+1\sigma}+h.c.)+\\
62: \frac{U}{2}\sum_{i=1}^{L}\sum_{\sigma\ne\sigma^{'}}\hat{n}_{i\sigma}\hat{n}_{i\sigma^{'}},
63: \label{HAM}
64: \end{equation}
65: where $t>0$ is the hopping matrix set as the energy unit, $U<0$
66: the coupling constant of the attractive on-site interaction, $L$
67: the number of lattice sites, $\sigma$ and $\sigma^{'}$ the spin
68: indices, which take the values $(N-1)/2$, $(N-1)/2-1$, $\cdots$,
69: $1-(N-1)/2$, $-(N-1)/2$, respectively.
70: $\hat{c}^{\dagger}_{i\sigma}$ and $\hat{c}_{i\sigma}$ denote the
71: creation and annihilation operators, respectively, for a particle
72: with spin $\sigma$ at the site $i$ and
73: $\hat{n}_{i\sigma}=\hat{c}^{\dagger}_{i\sigma} \hat{c}_{i\sigma}$
74: indicates the operator of the particle number. The Hamiltonian (1)
75: has U(1)$\otimes$SU(N) symmetry\cite{ASSARAF1}. The generator for
76: the U(1) symmetry is $\mathcal{N}=\sum_{i\sigma}\hat{n}_{i\sigma}$
77: and the U(1) symmetry implies that the particle number
78: $\mathcal{N}$ is conserved. The generator for the SU(N) symmetry
79: is $S^{A}=\sum_{i\sigma\sigma^{'}}
80: \hat{c}^{\dagger}_{i\sigma}\mathcal{T}^{A}_{\sigma\sigma^{'}}\hat{c}_{i\sigma^{'}}$,
81: where $\mathcal{T}^{A}$ are the generators of the SU(N) group in
82: its fundamental representation. These symmetries are useful for
83: simplifying numerical calculations and classifying excitations for
84: the present system.
85:
86: Although this model is exactly solvable for $N=2$\cite{LIEB1}, it
87: seems that physical features obtained for $N=2$ are not immediately
88: applicable to more general cases with $N>2$\cite{CHOY1}. Recently, this model
89: with $N>2$ and $U>0$ has been studied by using several analytic as
90: well as numerical approaches. At half-filling, a
91: renormalization-group analysis \cite{AFFLECK1} shows that both the
92: charge and spin coupling constants are renormalized to a large
93: value, resulting in gaps in both sectors. An analytic perturbative
94: renormalization group treatment in the fermionic representation
95: alternatively gives rise to a gapful spectrum for all $N>2$ and
96: $U>0$\cite{SZIRMAI1}. Moreover, by employing bosonization method
97: and quantum Monte Carlo simulations, Assaraf et al found for a
98: $1/N$ filling that the spin excitation is gapless for any positive
99: $U$, whereas the charge excitation is gapful only for $U>U_c$
100: where $U_c\neq 0$. However, Buchta et al obtained gapless spin as
101: well as gapful charge excitations from accurate density-matrix
102: renormalization group(DMRG) calculations with $N=3, 4$ and $5$ for
103: any $U>0$.
104:
105: For the attractive interaction, on the other hand, it is known
106: that the one-dimensional attractive half-filled SU(2) Hubbard
107: model is described by a Luther-Emery liquid model, in which the
108: charge excitation is gapless, whereas the spin excitation is
109: gapful. By the hidden SU(2) transformation, the SU(2) Hubbard
110: model with $U$ can be mapped to the one with $-U$. However, for
111: the SU(N) Hubbard model with $N>2$ such a mapping does not exist
112: so that one cannot obtain any insight into the low-energy
113: properties for the attractive case through the mapping from the
114: repulsive case. In this paper, we will show that the SU(N) Hubbard
115: model at half filling with an attractive interaction belongs to a
116: different universality class from the SU(2) one and all the
117: excitations are gapful. We expect that our findings
118: are not only of fundamental interest, but also useful for
119: experimentalists, since the attractive SU(N) Hubbard model may be
120: possibly realized by experiments in ultra-cold atomic systems.
121:
122: The rest of this paper is organized as follows.
123: In Sec. II, we analyze low-energy properties of the Hamiltonian (\ref{HAM})
124: in the strong coupling limit $(N-1)|U|\gg t$ first and then in the weak coupling
125: limit $(N-1)|U|\ll t$ by the perturbation treatments. Particularly we discuss
126: the dependence of charge gap, Cooperon gap, spin gap and quasiparticle gap on $U$ in both regimes.
127: In Sec. III, we present our numerical results which are obtained from the DMRG calculations
128: and compare them with analytic behavior in both weak and strong coupling regime.
129: A summary is finally given in Sec. IV.
130:
131: \section{PERTURBATION CALCULATIONS}
132: In this section, we study the low-energy properties of the
133: Hamiltonian (\ref{HAM}) by a perturbation theory.
134: For this purpose, we rewrite it as
135: \begin{eqnarray}
136: \mathcal{H}=\mathcal{H}_t+\mathcal{H}_u,
137: \end{eqnarray}
138: where
139: $\mathcal{H}_t=-t\sum_{i\sigma}(\hat{c}^{\dagger}_{i\sigma}\hat{c}_{i+1\sigma}+h.c.)$
140: is the hopping term and
141: $\mathcal{H}_{u}=(U/2)\sum_{i\sigma\ne\sigma^{'}}\hat{n}_{i\sigma}\hat{n}_{i\sigma^{'}}$
142: the on-site interaction. We start with the on-site interaction
143: part $\mathcal{H}_{u}$. Since the on-site interaction is
144: attractive, $N$ particles with different $\sigma$ tend to stay on
145: one site and form a SU(N) singlet. The energy of the SU(N) singlet
146: is $UN(N-1)/2$. In the half-filling case where $N/2$ particles per
147: site, the ground states of $\mathcal{H}_u$ are highly degenerate,
148: involving half of the lattice sites occupied by the SU(N) singlets
149: and the other half being empty. In the strong coupling region,
150: i.e. $|U|(N-1)\gg t$, $\mathcal{H}_{u}$ is taken as the zeroth
151: order Hamiltonian, while $\mathcal{H}_t$, being the order of $Nt$,
152: is regarded as a perturbation. Up to the second order, we obtain
153: the first order effective Hamiltonian
154: \begin{eqnarray*}
155: \mathcal{H}_{eff}^{(1)}=P\mathcal{H}_tP=0 ,
156: \end{eqnarray*}
157: where $P$ is a projection operator which restricts the effective
158: Hamiltonian in the subspace spanned by the ground states of
159: $\mathcal{H}_{u}$. The second order effective Hamiltonian
160: reads
161: \begin{eqnarray}
162: \mathcal{H}_{eff}^{(2)} & = & P\mathcal{H}_{t}\frac{1}{E_0-\mathcal{H}_u}(1-P)\mathcal{H}_tP \nonumber\\
163: & = & \frac{2t^2}{(N-1)U}P\sum_{i}\hat{n}_{i}P\nonumber\\
164: & & - \frac{2t^2}{N(N-1)U}P\sum_{i}\hat{n}_{i}\hat{n}_{i+1}P,
165: \label{HEFF}
166: \end{eqnarray}
167: where $\hat{n}_{i}=\sum_{\sigma}\hat{n}_{i\sigma}$ the number
168: operator at the site $i$. The degeneracy of the ground state of
169: $\mathcal{H}_u$ is thus lifted by $\mathcal{H}_t$. At half
170: filling, one can easily obtain the ground state energy correction per site
171: which is given as $Nt^2/(N-1)U$ by the first term of Eq.
172: (\ref{HEFF}). Since the second term introduces an effective
173: repulsion interaction between the particles at nearest
174: neighbor(NN) sites, it results in a charge density wave (CDW)
175: state in which the SU(N) singlet and the empty site occur
176: alternatively to exhibit a long range order resulting from
177: translational symmetry breaking. It turns out that the ground
178: state is two-fold degenerate. The configuration of the
179: ground state is schematically shown in Fig. \ref{fig0}(a).
180: \begin{figure}[ht]
181: \includegraphics[width=7.8cm,angle=0]{C-CO-Excitation.eps}
182: \caption{Schematic illustrations of the configurations, and the
183: empty circle here represents one SU(N) singlet formed by $N$
184: particles with different $\sigma$. (a) For the CDW ground state.
185: Shifting all the circles by one site, one can obtain the other one
186: of two-fold degenerate ground states. (b) For the charge
187: excitation: the SU(N) singlet at site $i+2$ shifts to the site
188: $i+1$. (c) For the Cooperon excitation: one SU(N) singlet is added
189: to the site i+1.} \label{fig0}
190: \end{figure}
191:
192: It is well-known that both the charge and Cooperon excitations for
193: the attractive SU(2) Hubbard model are gapless. However for the
194: SU(4) case, it has been shown\cite{ZHAO1} that the charge and
195: Cooperon excitations are gapful and equal to each other at
196: half-filling. In the following, one can see that this conclusion
197: is also valid for other attractive SU(N) Hubbard models with $N>2$.
198: In particular, the charge gap and Cooperon gap can be
199: easily derived from the effective Hamiltonian (\ref{HEFF}). The
200: charge gap $\Delta_c$ is defined as the lowest excitation in the
201: SU(N) singlet subspace as follows
202: \begin{eqnarray}
203: \Delta_c=E_1(L,NL/2,0)-E_0(L,NL/2,0), \label{DELTAC}
204: \end{eqnarray}
205: where $E_0(L,M,S)$ is the ground state energy in the spin-$S$
206: channel with $L$ sites and $M$ particles, and $E_n(L,M,S)$ the
207: $n$-th excitation energy. The gap for Cooperon excitations for the
208: SU(N) Hubbard model is given from the energy
209: difference between states by adding $N$ particles or $N$ holes to
210: the system, which is defined as
211: \begin{eqnarray}
212: \Delta_N & = & \frac{1}{2}\left[E_0(L,\frac{NL}{2}+N,0)
213: +E_0(L,\frac{NL}{2}-N,0)\right]\nonumber\\
214: & & -E_0(L,\frac{NL}{2},0).
215: \end{eqnarray}
216: The charge gap $\Delta_c$ is given by shifting one of SU(N)
217: singlets in the ground state configuration to its nearest neighbor
218: site as shown in Fig. \ref{fig0}(b). Then one has
219: \begin{eqnarray}
220: \Delta_c=-\frac{2Nt^2}{(N-1)U}. \label{DCHARGE}
221: \end{eqnarray}
222: Similarly,
223: one obtains the SU(N) Cooperon gap
224: \begin{eqnarray}
225: \Delta_N=-\frac{2Nt^2}{(N-1)U}.
226: \end{eqnarray}
227: which is shown in Fig. \ref{fig0}(c).
228: One can see that $\Delta_c=\Delta_N$ for all $N>2$, which was
229: previous shown for the $N=4$ case\cite{ZHAO1}. In the large-$N$
230: limit,
231: \begin{eqnarray}
232: \Delta_c=\Delta_N=-\lim_{N\rightarrow\infty}\frac{2Nt^2}{(N-1)U}=-\frac{2t^2}{U}.
233: \end{eqnarray}
234: We note that the difference for the low-lying excitations between
235: $N=2$ and $N>2$ results from the effective interaction
236: between the singlets at the NN sites as involved in effective
237: Hamiltonian (\ref{HEFF}). For the $N=2$ case, one has the
238: effective Hamiltonian\cite{EMERY1}
239: \begin{eqnarray}
240: \mathcal{H}^{(2)}_{eff,su(2)} & = &
241: \frac{2t^2}{U}P\sum_{i\sigma}\hat{n}_{i\sigma}P \label{HEFFSU2}\\
242: & & - \frac{t^2}{U}P\sum_{\langle{ij}\rangle\sigma}
243: (\hat{n}_{i\sigma}\hat{n}_{j\sigma}-\hat{c}^{+}_{i\sigma}\hat{c}^{+}_{i\bar{\sigma}}
244: \hat{c}_{j\bar{\sigma}}\hat{c}_{j\sigma})P. \nonumber
245: \end{eqnarray}
246: Compared to the effective Hamiltonian (\ref{HEFF}), one has
247: additionally a pair hopping term, which involves the same
248: amplitude as the NN repulsion term and eventually destroys the CDW
249: long range order for $N=2$. On the other hand, for
250: $N>2$ cases, although a similar hopping term emerges at the $N$-th
251: order perturbation, it has a smaller amplitude than the NN
252: repulsion term so that one can has a stable CDW ground state.
253:
254: Now we turn to study the spin and quasiparticle excitations.
255: The spin gap is defined in correspondence to the
256: lowest excitation with different spin quantum number from the
257: ground state. The quasiparticle gap is defined as a
258: energy change by adding one particle or hole to the system. Since
259: the ground state is CDW, in order to obtain these two gaps, we
260: resort to the following Hartree-Fock(HF) approximation.
261:
262: \begin{eqnarray}
263: \hat{n}_{i\sigma}\hat{n}_{i\sigma^{'}}\simeq n_{i\sigma}\langle{\hat{n}_{i\sigma^{'}}}\rangle+
264: \langle{\hat{n}_{i\sigma}}\rangle\hat{n}_{i\sigma^{'}}-\langle{\hat{n}_{i\sigma}}\rangle
265: \langle{\hat{n}_{i\sigma^{'}}}\rangle,
266: \end{eqnarray}
267: where $\langle{\hat{n}_{i\sigma}}\rangle=n_0 +(-1)^{i}\delta{n}$,
268: and $\delta{n}$ is the order parameter, $\langle\cdots\rangle$ is
269: the average over the ground state. The HF Hamiltonian then reads
270: \begin{eqnarray}
271: \mathcal{H}^{HF}&=&-t\sum_{i\sigma}(\hat{c}_{i\sigma}^{\dagger}\hat{c}_{i+1\sigma}+h.c.)
272: +\frac{U}{2}\sum_{i\sigma\ne\sigma^{'}}(\hat{n}_{i\sigma}\langle{\hat{n}_{i\sigma^{'}}}\rangle\nonumber\\
273: &+&
274: \langle{\hat{n}_{i\sigma}}\rangle\hat{n}_{i\sigma^{'}}-\langle{\hat{n}_{i\sigma}}\rangle
275: \langle{\hat{n}_{i\sigma^{'}}}\rangle). \label{HHF}
276: \end{eqnarray}
277: At half-filling, one has $n_0=\frac{1}{2}$ and
278: $\sum_{l\sigma}\hat{n}_{l\sigma}=\frac{LN}{2}$. Eq. (\ref{HHF})
279: can be further simplified as
280: \begin{eqnarray}
281: \mathcal{H}^{HF}&=&-t\sum_{i\sigma}(\hat{c}^{\dagger}_{i\sigma}\hat{c}_{i+1\sigma}+h.c.)\nonumber\\
282: &+&U(N-1)\delta{n}\sum_{i\sigma}(-1)^{i}\hat{n}_{i\sigma}+{\rm
283: const}.
284: \end{eqnarray}
285: Introducing $\hat{a}_{l\sigma}=\hat{c}_{2l\sigma}$,
286: $\hat{b}_{l\sigma}=\hat{c}_{2l+1\sigma}$ and taking a Fourier
287: transformation
288: $\hat{a}_{l\sigma}=\sqrt{\frac{2}{L}}\sum_{k}\hat{a}_{k\sigma}e^{ikl}$,
289: $\hat{b}_{l\sigma}=\sqrt{\frac{2}{L}}\sum_{k}\hat{b}_{k\sigma}e^{ikl}$,
290: we obtain
291: \begin{eqnarray}
292: \mathcal{H}^{HF}&=&-t\sum_{k\sigma}((1+e^{-ik})\hat{a}^{\dagger}_{k\sigma}\hat{b}_{k\sigma}+h.c.)\\
293: &+&U(N-1)\delta
294: n\sum_{k\sigma}(\hat{a}^{\dagger}_{k\sigma}\hat{a}_{k\sigma}-\hat{b}^{\dagger}_{k\sigma}\hat{b}_{k\sigma})+{\rm
295: const}.\nonumber
296: \end{eqnarray}
297: Diagonalizing this HF Hamiltonian, we can obtain two bands for
298: each spin species with the following dispersion
299: \begin{eqnarray}
300: \omega_{k\sigma}=\pm \sqrt{\Delta_{1}^{2}+4t^2\cos^{2}{\frac{k}{2}}},
301: \end{eqnarray}
302: where $\Delta_{1}=-(N-1)U\delta n$ is the quasiparticle gap. Under
303: the HF approximation, one has that
304: $\Delta_{c}=\Delta_{s}=2\Delta_{1}$. By solving the
305: self-consistent equation for $\delta n=\langle
306: \hat{a}^{\dagger}_{l\sigma}\hat{a}_{l\sigma}-\hat{b}^{\dagger}_{l\sigma}\hat{b}_{l\sigma}\rangle$
307: one can determine the order parameter $\delta n$. In the weak
308: coupling limit $t\ll |U|(N-1)$, we can solve it approximately and
309: obtain
310: \begin{eqnarray}
311: \Delta_{1}\simeq 2\pi te^{\frac{2\pi t}{U(N-1)}}.
312: \end{eqnarray}
313: In the strong coupling limit, one can put $\delta{n}\simeq 0.5$, and then
314: obtain the quasiparticle gap
315: \begin{eqnarray}
316: \Delta_{1}\simeq -U(N-1)/2 ,
317: \label{D1}
318: \end{eqnarray}
319: and the spin gap
320: \begin{eqnarray}
321: \Delta_{s}\simeq -U(N-1) .
322: \label{DS}
323: \end{eqnarray}
324:
325: In the HF approximation, since $\Delta_c=2\Delta_1$, one obtains
326: $\Delta_c\simeq -U(N-1)$ which is inconsistent with the results
327: Eq. (\ref{DCHARGE}) obtained from the strong coupling perturbation
328: theory. This is because in the presence of the strong attractive
329: on-site interaction between particles, the single particle picture is not
330: valid for the charge excitations, in which $N$ particles forming a
331: SU(N) singlet shift as a whole to the NN site, as illustrated in
332: Fig. \ref{fig0}(b). However, we would expect that the HF results are
333: still qualitatively correct for weak coupling limit.
334:
335: \section{NUMERICAL CALCULATIONS}
336: In this section, we present our numerical results for $N=3, 4, 5$
337: and $6$ cases and compare them with those analytic predictions
338: from the perturbation theories and the HF approximation. Our
339: numerical results are obtained from a large scale DMRG
340: computation.\cite{WHITE1,PESCHEL1,SCHOLLWOCK1} It is well known
341: that DMRG method is the most powerful numerical tool for accurate
342: exploration on low-energy properties of one dimensional systems at
343: zero temperature. However, it is nontrivial to apply this method
344: to the SU(N) Hubbard model. On one hand, there is a large number
345: of degrees of freedom per site. For instance, the degree of
346: freedom per site is 32 when $N=5$. On the other hand, although a
347: open boundary condition (OBC) in DMRG calculations may lead to
348: more accurate results than a periodic boundary condition (PBC),
349: while the emergence of edge states under the OBC makes the
350: calculation of various excitations practically more difficult.
351:
352: For these reasons, we have made some additional considerations to
353: the standard DMRG algorithm as follows. When $N=3$, i.e. SU(3)
354: case, one has 8 degrees of freedom per site. In this case, we use
355: the PBC with keeping about $2000\sim 3500$ optimal states in the
356: DMRG procedure and two sites are added in order to enlarge the
357: chain length at each DMRG step. The maximal truncation error is of
358: the order $10^{-5}$. Similarly, the charge gap for $N=4$ is
359: calculated with at most 1600 states kept. The other excitations
360: are calculated with the OBC. At each step we add one lattice site
361: to the chain, breaking the lattice site into two pseudo sites. To
362: obtain accurate results, we preselect some specific chain lengths,
363: and perform sweeping at these preselected lengths, the final
364: results are got by extrapolating the results at the preselected
365: lengths to the thermodynamic limit. To avoid cumbersome edge
366: excitations when the OBC is employed, the preselected lengths are
367: odd instead of even. Correspondingly the definitions of the
368: excitations are changed. For example, the Cooperon gap for $N=4$
369: is redefined as $\Delta_{N}=E_0(L,2L+6,0)-E_0(L,2L+2,0)-6U$, where
370: the particle-hole symmetry is taken into account and odd $L$ are
371: used.
372:
373: \subsection{Energy correction and degeneracy of the ground states}
374: The analysis based on the strong coupling perturbation theory in
375: the above section shows that the hopping term, lifting the
376: degeneracy of $\mathcal{H}_u$, results in an energy correction per
377: site $Nt^{2}/(N-1)U$. To verify this prediction, we calculate the
378: energy corrections for $N=3, 4$ and $5$ using DMRG method. Both
379: numerical and analytic results are shown in Fig. \ref{fig1}.
380: \begin{figure}[ht]
381: \includegraphics[width=7.0cm,angle=0]{gs-energycorrection.eps}
382: \caption{Ground state energy correction per site obtained from
383: both the DMRG calculation and the strong coupling perturbation
384: theory $\mathcal{H}^{(2)}_{eff}$ for N=3, N=4 and N=5 are shown as
385: a function of $U$. } \label{fig1}
386: \end{figure}
387: One can see that the energy correction is negative and decreases
388: monotonically with respect to increasing of $U$. In the strong
389: coupling region, the results given by the perturbation theory
390: agree well with the DMRG data. For $N=3$, the deviation between
391: analytic and numerical results is within the numerical accuracy
392: for $U<-2$. With increasing of $N$, the deviation becomes smaller
393: and smaller. This implies that the perturbation theory gives rise
394: to better results for larger $N$ for the strong coupling regime.
395: In the weak coupling region, the analytic results from the
396: effective Hamiltonian (\ref{HEFF}) severely deviates from the DMRG
397: results. This is reasonable since in the weak coupling region,
398: $\mathcal{H}_t$ is no longer a small perturbation. Nevertheless,
399: when $N$ increases, the valid region of the strong coupling
400: perturbation theory increases simultaneously.
401:
402: \begin{figure}[ht]
403: \includegraphics[width=7.0cm,angle=0]{N3-u-1.0-u-4.0.eps}
404: \caption{Two lowest excitation gaps in the SU(N) singlet subspace
405: for $N=3$ case with two different values of $U$. In the thermodynamic
406: limit, the first excitation energy becomes degenerate with the ground
407: state, while the energy difference between the second excitation
408: and the ground state remains finite, indicating a gapful excitation.}
409: \label{fig2}
410: \end{figure}
411:
412: We have examined the symmetry properties of the ground state. In
413: the thermodynamic limit, the ground state is two-fold degeneracy. For
414: a finite $L$, the ground state is a SU(N) singlet, and there is
415: another SU(N) singlet state just above the ground state. When the
416: chain length $L$ is increased, the energy difference between these
417: two states decreases and eventually vanishes in the large-$L$
418: limit. This is consistent with the prediction from the effective
419: Hamiltonian (\ref{HEFF}). Moreover, for a given chain length $L$, we find that the
420: energy difference between two states for the Hamiltonian
421: (\ref{HAM}) decrease rapidly with increasing $|U|$, since the
422: effective Hamiltonian Eq. (\ref{HEFF}) becomes more accurate for
423: larger $|U|$. In Fig. \ref{fig2} and \ref{fig3}, we demonstrate
424: this feature with two different values of $U$ under OBC for $N=3$
425: and $4$, respectively. Note that the energy difference for the
426: second excitation, which is finite in the large-$L$ limit,
427: corresponds to edge excitations rather than bulk excitations.
428: \begin{figure}[ht]
429: \includegraphics[width=7.0cm,angle=0]{N4-u-0.8-u-2.0.eps}
430: \caption{Two lowest excitation gaps in the SU(N) singlet subspace
431: for $N=4$ case with two different values of $U$. In the
432: thermodynamic limit, the first excitation energy becomes degenerate with
433: the ground state, while the energy difference between the second
434: excitation and the ground states remains finite, indicating a gapful
435: excitation.} \label{fig3}
436: \end{figure}
437:
438: \subsection{Charge gaps and Cooperon gaps}
439: The prediction that the charge gap $\Delta_c$ is equal to the
440: Cooperon gap $\Delta_N$ is confirmed numerically for N=4\cite{ZHAO1}.
441: However, the strong coupling perturbation theory gives rise to
442: $\Delta_c=\Delta_N$ for general $N$, it is still interesting to
443: examine numerically whether this equation is valid for both even and odd N and all
444: regimes of the coupling constant $U$\cite{BUCHTA1}. Fig.
445: \ref{fig4} demonstrates the charge and Cooperon gaps for $N=3$ as
446: comparison to $N=4$ as a function of $U$.
447: \begin{figure}[ht]
448: \includegraphics[width=7.0cm,angle=0]{charge-cooperon-N3-4.eps}
449: \caption{The Cooperon and charge gaps are shown for $N=3$ and $4$.
450: The asymptotic behavior from the effective Hamiltonian(\ref{HEFF})
451: (curves) are also shown for comparison.} \label{fig4}
452: \end{figure}
453: One can see that both charge and Cooperon gaps for both $N=3$ and
454: $N=4$ are finite for all $U<0$. When $|U|$ is increased, these
455: gaps increase first in the weak coupling region. After they reach
456: their maxima, they decrease following the asymptotic behavior
457: $-\frac{2Nt^2}{(N-1)U}$ resulting from the strong coupling perturbation theory. We
458: would mention that even beyond the valid range of the perturbation
459: theory, our numerical data still show that the charge gaps are
460: equal to the corresponding Cooperon gaps for both odd $N (=3)$ and
461: even $N (=4)$. Therefore, one may conclude that
462: $\Delta_c=\Delta_N$ for all $N$ and $U<0$.
463:
464: Moreover we calculate the Cooperon gaps for $N=5$ and $6$ in order
465: to explore the large-$N$ behavior of the charge as well as
466: Cooperon gaps. The results of the Cooperon gap for $N=5$ and $6$
467: are shown in Fig. \ref{fig6}.
468: \begin{figure}[ht]
469: \includegraphics[width=7.0cm,angle=0]{N5-N6-cooperon.eps}
470: \caption{The Cooperon gaps and the corresponding asymptotic behaviors as a function of $U$ for N=5 and 6
471: are shown.}
472: \label{fig6}
473: \end{figure}
474: Comparing Fig. \ref{fig4} and \ref{fig6}, one can see that the
475: larger $N$ is, the closer Cooperon gaps to that derived from the
476: effective Hamiltonian (\ref{HEFF}). For $N=3$ as shown in Fig.
477: \ref{fig4}, the deviation of the asymptotic behavior from the DMRG
478: data is visible even at $U=-10$, while for $N=4$ case, the
479: deviation is within the numerical accuracy up to $U\simeq -5$.
480: For $N=5$ and $6$, the good agreement can be seen up
481: to $U\simeq -3$ and $U\simeq -2$, respectively. Furthermore, as
482: $N$ increases, the height of the peak of the Cooperon gaps
483: increases and the position of the peak shifts toward $U=0$. These
484: features can be easily understood since the criteria for the
485: strong coupling perturbation $t\ll |U|(N-1)$ depends on $|U|(N-1)$
486: rather than $U$ only. The DMRG results here also verify that the
487: charge and Cooperon gap behaves asymptotically as
488: $-\frac{2t^2}{U}$ in the large-$N$ limit.
489:
490: \subsection{Quasiparticle gap and spin gap}
491: In the following, we discuss the quasiparticle and spin gaps. In
492: order to improve numerical efficiency, we need to carry out DMRG
493: calculations in a subspace which can be obtained by decomposing
494: the irreducible representations of SU(N) into the irreducible
495: representation of SO(3)\cite{HAMERMESH1,YAMASHITA1}. Numerically,
496: one can determine the irreducible representations for short chains
497: by varying $z$-component. Table \ref{TABLE1} shows the
498: decomposition of some irreducible representations of SU(3) to
499: SO(3).
500: \begin{table} \caption{Decomposition of
501: some irreducible representations of SU(3) to SO(3).}
502: \begin{ruledtabular}
503: \begin{tabular}{cccc}
504: $SU(3)$&$SO(3)$&$\nu$\\
505: \hline
506: $[1^3]$&$0$&1\\
507: $[2^{1}1^{1}]$&$2\oplus{1}$&8\\
508: $[3^1]$&$3\oplus{1}$&10\\
509: \hline
510: $[1^1]$&$1$&3\\
511: $[2^2]$&$2\oplus{0}$&6 \label{TABLE1}
512: \end{tabular}
513: \end{ruledtabular}
514: \end{table}
515:
516: \begin{table}[ht]
517: \caption{Decomposition of some irreducible representations of SU(4) to SO(3).}
518: \begin{ruledtabular}
519: \begin{tabular}{cccc}
520: $SU(4)$&$SO(3)$&$\nu$\\
521: \hline
522: $[1^4]$&$0$&1\\
523: $[2^2]$&$4\oplus{2}\oplus{2}\oplus{0}$&20\\
524: $[2^{1}1^{2}]$&$3\oplus{2}\oplus{1}$&15\\
525: $[4^1]$&$0\oplus{2}\oplus{3}\oplus{4}\oplus{6}$&35\\
526: $[3^{1}1^{1}]$&$1\oplus{1}\oplus{2}\oplus{3}\oplus{3}\oplus{4}\oplus{5}$&45\\
527: \hline
528: $[1^1]$&$\frac{3}{2}$&4\\
529: $[2^{2}1^{1}]$&$\frac{7}{2}\oplus\frac{5}{2}\oplus\frac{3}{2}\oplus\frac{1}{2}$&20\\
530: $[3^{1}1^{2}]$&$\frac{9}{2}\oplus\frac{7}{2}\oplus\frac{5}{2}\oplus\frac{5}{2}\oplus\frac{3}{2}\oplus\frac{1}{2}$&36
531: \label{TABLE2}
532: \end{tabular}
533: \end{ruledtabular}
534: \end{table}
535:
536: For $N=3$ that the spin excitation belongs to the
537: representation $[2^{1}1^{1}]$ and it is 8-fold degenerate,
538: while the quasiparticle excitation belongs to the representation
539: $[1^{1}]$ and it is 3-fold degenerate. Table \ref{TABLE2} shows
540: the decomposition of some irreducible representations of SU(4) to
541: SO(3). In this case, the spin excitation belongs to the
542: representation $[2^{1}1^{2}]$ and its degeneracy is 15-fold,
543: whereas the quasiparticle excitation belongs to the representation
544: $[1^1]$ and its degeneracy 4-fold. Fig. \ref{fig7} shows the quasiparticle
545: and spin gaps for $N=3$ and $4$ as well.
546: \begin{figure}[ht]
547: \includegraphics[width=7.0cm,angle=0]{N3-N4-sp-spin-gap.eps}
548: \caption{The quasiparticle gap and spin gap for N=3, 4 are shown in the figure.
549: For N=3, the quasiparticle gap($\small{\bigcirc}$) and spin gap($\square$) are shown by the black symbols.
550: For N=4, the quasiparticle gap($\triangle$) and spin gap($\times$) are shown by red symbols.}
551: \label{fig7}
552: \end{figure}
553: In weak coupling region, one can see that both the quasiparticle
554: and spin gaps open exponentially, as predicted by the HF approximation,
555: although we cannot determine precise critical behavior for the
556: opening of the gaps from our numerical data due to limited
557: numerical accuracy. In the strong coupling region, both the
558: quasiparticle gap and spin gap depend linearly on $U$ as seen in
559: the inset. This is also consistent with the HF results. Moreover, one
560: also see that the ratio $\Delta_{1}/|U|$ approaches 1 for $N=3$
561: asymptotically, while it approaches 1.5 for $N=4$. These asymptotic
562: behaviors qualitatively coincide with Eq. (\ref{D1}) and
563: (\ref{DS}) and the relation $\Delta_s=2\Delta_1$ derived from the
564: HF approximation. For finite $U$, however, the relation is just approximately
565: valid since there remains residual interaction between
566: quasiparticles.
567:
568: \section{CONCLUSIONS}
569: In this paper, we have investigated low-energy properties of the
570: SU(N)(N$>$2) Hubbard model with attractive on-site interaction at
571: half-filling. By using the perturbation theory and DMRG method, we
572: show that the ground state is a CDW state with two-fold
573: degeneracy. The CDW long range order reflects the broken translational
574: symmetry. On one hand, the strong coupling perturbation theory
575: predicts that both the charge excitations and Cooperon excitations
576: are gapful and equal to each other with the asymptotic behavior
577: $-2Nt^{2}/(N-1)U$. Combining this with numerical results for
578: $N=3, 4, 5$ and $6$, we conclude that the charge gap is equal to the
579: Cooperon gap for all $U<0$ and $N>2$. In the large N limit, both
580: the charge and Cooperon gaps behaves as $-2t^{2}/U$.
581: Considering the CDW ground state, we
582: obtain qualitatively correct behavior for the spin and
583: quasiparticle gaps from HF approximation, which are confirmed by
584: our numerical data. In the weak coupling region, they open
585: exponentially, whereas in the strong coupling region, they depend
586: linearly on $U$ with the spin gap $\Delta_{s}\sim -(N-1)U$ and
587: quasiparticle gap $\Delta_{1}\sim -(N-1)U/2$.
588:
589: Our findings indicate that at half filling, the SU(N)(N$>$2)
590: Hubbard model belongs to a different universality class from the
591: SU(2) case. This can be easily understood by observing the
592: difference between the effective Hamiltonians for the SU(N)(N$>$2)
593: and the SU(2) cases. In particular, the effective Hamiltonian
594: which for the SU(2) case involves an effective hopping term has
595: the same order as the effective repulsion interaction for
596: particles between nearest-neighbor sites. The hopping term
597: destroys the CDW order. For $N>2$, the repulsive
598: interaction dominates over the effective hopping term.
599:
600: J. Zhao would thank H. Tsunetsugu for helpful discussion. X.Q.
601: Wang is supported in part by NFC2005CB32170X, and NSFC10425417
602: $\&$ C10674142.
603:
604: \begin{references}
605: \bibitem{kohl05}M. K\"{o}hl, H. Moritz, T. St\"{o}ferle, K. G\"{u}nter,
606: and T. Esslinger, Phys. Rev. Lett {\bf 94}, 080403 (2005).
607: \bibitem{matrin}M.W. Zwierlein, C.H. Schunck, A. Schirotzek and W. Ketterle,
608: Nature (London) {\bf 442}, 54 (2006).
609: \bibitem{HO1} Tin-Lun Ho and Sungkit Yip, Phys. Rev. Lett. {\bf 82}, 247 (1999).
610: \bibitem{WU1} Congjun Wu, Jiang-ping Hu, and Shou-cheng Zhang, Phys. Rev. Lett. {\bf 91}, 186402 (2003).
611: \bibitem{WU2} Congjun Wu, Mod. Phys. Lett. B {\bf 20}, 1707 (2006)
612: \bibitem{TU1} Hong-Hao Tu, Guang-Ming Zhang, and Lu Yu, Phys. Rev. B {\bf 74}, 174404 (2006)
613: \bibitem{AFFLECK1} I. Affleck and J.B. Marston, Phys. Rev. B {\bf 37}, 3774 (1988);
614: J.B. Marston and I. Affleck, Phys. Rev. B {\bf 39}, 11538 (1989).
615: \bibitem{ZHAO1} J.Z. Zhao, K. Ueda and X.Q. Wang, Phys. Rev. B {\bf 74}, 233102 (2006).
616: \bibitem{ASSARAF1}R. Assaraf, P. Azaria, M. Caffarel, and P. Lecheminant, Phys. Rev. B {\bf 60}, 2299 (1999).
617: \bibitem{LIEB1} E.H. Lieb and F.Y. Wu, Phys. Rev. Lett. {\bf 20}, 1445 (1968).
618: \bibitem{CHOY1} T.C. Choy, Phys. Lett. {\bf 80A}, 49 (1980); F.D.M. Haldane, Phys. Lett. A {\bf 80A}, 281 (1980);
619: T.C. Choy and F.D.M. Haldane, Phys. Lett. {\bf 90A}, 83 (1982).
620: \bibitem{SZIRMAI1} E. Szirmai and J. S\'olyom, Phys. Rev. B {\bf 71}, 205108 (2005).
621: \bibitem{BUCHTA1} K. Buchta, $\ddot{O}$. Legeza, E. Szirmai, J. S$\acute{o}$lyom, cond-mat/0607374.
622: \bibitem{EMERY1} V.J. Emery, Phys. Rev. B {\bf 14}, 2989 (1976).
623: \bibitem{WHITE1} S.R. White, Phys. Rev. B {\bf 48}, 10345 (1993).
624: \bibitem{PESCHEL1} I. Peschel, X. Wang, M. Kaulke and K. Hallberg, {\it Density Matrix Renormalization},
625: LNP{\bf 528}, Springer-Verlag, 1999.
626: \bibitem{SCHOLLWOCK1} U. Schollw$\ddot{o}$ck, Rev. Mod. Phys. {\bf 77}, 259 (2005)
627: \bibitem{HAMERMESH1} M. Hamermesh, Group Theory (Addison-Wesley, Reading, MA, 1962).
628: \bibitem{YAMASHITA1} Y. Yamashita, N. Shibata and K. Ueda, Phys. Rev. B {\bf 58}, 9114(1998)
629: \end{references}
630: \vfill
631: \end{document}
632: