cond-mat0103096/tj.tex
1: \documentstyle[aps,prl,manuscript]{revtex}
2: \begin{document}
3: \draft
4: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5: \newcommand{\vektor}[1]{\mbox{\boldmath $#1$}}
6: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7: \title{DMRG and the Two Dimensional t-J Model}
8: \author{I. P. McCulloch${}^{\rm 1}$, A. R. Bishop${}^{\rm 2}$
9: and M. Gulacsi${}^{\rm 1}$}
10: \address{
11: ${}^{\rm 1}$
12: Department of Theoretical Physics, Institute of Advanced Studies \\
13: The Australian National University, Canberra, ACT 0200, Australia \\
14: ${}^{\rm 2}$
15: Theoretical Division, Los Alamos National Laboratory \\
16: Los Alamos, NM 87545, U.S.A.}
17: %% \date{\today}
18: \maketitle
19: \begin{abstract}
20: 
21: We describe in detail the application of the
22: recent non-Abelian Density Matrix Renormalization Group (DMRG) algorithm
23: to the two dimensional t-J model.
24: This extension of the DMRG algorithm allows us to keep the equivalent of
25: twice as many basis states as the conventional DMRG algorithm for the
26: same amount of computational effort, which permits a deeper understanding
27: of the nature of the ground state.
28: 
29: \end{abstract}
30: \pacs{ ~ }
31: 
32: Since the discovery of superconductivity in the rare-earth copper
33: oxides there has been a growing interest in strongly correlated electronic
34: systems. The t-J model proposed by Anderson (1987), and Zhang and
35: Rice (1988) is an example of this interest. The theoretical predictions and
36: implications of the model are possibly relevant and useful for a deeper
37: understanding, particularly, the high temperature superconductors,
38: and in generally the motion of "holes" in an antiferromagnet.
39: 
40: It is conceptually a simple model and has been already widely studied
41: in relation to the high temperature superconductors. However,
42: it belongs to the class of systems which do not obey the condition
43: of Perron-Frobenuis (Yosida 1980). This condition states that if the
44: off-diagonal elements of a matrix are all non-positive and if the
45: matrix is not in a block diagonal form then the ground state
46: eigenvalue is non-degenerate. In the case of the t-J Hamiltonian
47: the off-diagonal elements are not all non-positive. Thus the above
48: theorem can not be applied, which implies that the
49: phenomenon of ground state level crossing is present
50: (Itoyama, McCoy and Perk 1990). As a direct consequence of this,
51: the thermodynamic system will be unstable against phase separation.
52: Indeed, as is known (Blatt and Weisskopf 1979), in three-dimension,
53: the system will collapse to arbitrarily high density. In two-dimensions
54: we have the argument of Emery, Kivelson and Lin (1990) that the model
55: is unstable against phase separation into hole-rich and no-hole phase.
56: 
57: It is known that interchanging particles of different spin
58: leads to a strong coupling between the kinetic and spin degrees of
59: freedom(Barnes 1989). Therefore, in different dimensions the model
60: will represent different cases. Phase separation occurs also in one
61: dimension, but for different
62: values of the band filling and interaction strength than in the higher
63: dimension case. In two dimensions, it was argued that phase separation
64: corresponds to stripe formation (Hellberg and Manousakis 1999). Stripe
65: formation is one of the most controversial issues in the study of high
66: temperature superconductors, where there is a phase separation of the
67: holes which is limited to short range by, e.g., Coulomb forces, and
68: orders in striped structures. Many experiments have
69: found evidence for stripes \footnote{For a recent review, see
70: the proceedings of the {\it Conference on Spectroscopies in Novel
71: Superconductors} (Proceedings, 1998).} and they arise
72: in a number of different theories involving strong correlations
73: (Zaanen and Gunnarsson 1989; Kivelson, Emery and Lin 1990; Zaanen,
74: Horbach and van Saarloos 1996; White and Scalapino 1998).
75: 
76: The first direct proof of the existence of a stripe ground state
77: emerged from t-J model calculations (White and Scalapino 1998).
78: The reason for this is that the t-J model inherits all the exchange
79: hole correlations resulting from the antiparallel spin correlations
80: in the $U \rightarrow \infty$ limit of the Fermi sea. Hence, if stripes
81: do exist, they will be more robust in the t-J model. Stripes have been
82: studied by a number of numerical techniques in the t-J model,
83: unfortunately resulting in conflicting conclusions. A major
84: question is if stripes are part of the known phase separated regime
85: of the t-J model (Hellberg and Manousakis 1999) or they represent
86: a different ground state. This is the focus of our study.
87: 
88: Using the recent developed non-Abelian DMRG (McCulloch and Gulacsi 2000)
89: we study the ground state of the two dimensional t-J model in the
90: vicinity of $J/t = 0.35$ and for different size square lattices up to 24x6.
91: We clearly find, for our choice of boundary conditions, a striped
92: ground state. We have also noticed that in the phase separated regimes,
93: the holes are attracted to the open boundary, while in the stripe
94: regime, the holes are repelled by the boundary. This is evidence
95: that that stripes are different from a usual phase transition
96: phenomena, but also indicates that there are large finite size
97: effects arising from the existence of the open boundary.
98: 
99: The non-Abelian DMRG algorithm
100: is numerically similar to the ``Interaction Round a Face'' (IRF)
101: DMRG (Sierra and Nishino 1997). However, the decomposition of
102: the Hamiltonian into the blocks used by the DMRG algorithm
103: is significantly different between the two algorithms. In
104: IRF-DMRG, the vertex Hamiltonian is first transformed
105: into an IRF model, and then a variant of DMRG is applied
106: to the IRF Hamiltonian. In the non-Abelian case, the
107: standard DMRG algorithm is generalized so that the block
108: operators transform as irreducible representations of an arbitrary
109: compact group (in this case $SU(2)$). This is a true generalization
110: of the original DMRG algorithm (White 1992, 1993), in the sense
111: that if the global symmetry group is chosen to be $U(1)$, as in
112: the usual $S^z$ basis used in DMRG, the original DMRG algorithm
113: is recovered exactly. In this paper, we fill in the details
114: required to apply the non-Abelian algorithm
115: to the 2 dimensional $t-J$ model, in the basis given by particle number,
116: $N$, and total spin, $j$, giving the global symmetry group
117: $U(1) \otimes SU(2)$. We then compare the results of this choice
118: of basis with previous calculations using the usual $S^z$ basis,
119: and comment on the conclusions that can be inferred
120: from the DMRG calculations.
121: 
122: The $t-J$ Hamiltonian is
123: \begin{equation}
124: H = - t \sum_{\langle i,j \rangle, \sigma}
125: (c^\dagger_{i\sigma} c_{j\sigma} + \mbox{H.c.})
126: + J (\sum_{\langle i,j \rangle} \vektor{S}_i \cdot \vektor{S}_j
127: - \frac{1}{4}  n_i n_j)
128: \end{equation}
129: defined on the subspace of no double occupancy, and $\langle i,j \rangle$
130: means summation over nearest neighbour pairs only.  The single site
131: operators act on the usual three dimensional t-J basis of an empty
132: site (hole), a single up spin and a single down spin. The difference
133: for non-Abelian DMRG is that the spin up and spin down states form an
134: $SU(2)$ multiplet of dimension 2, thus reducing the single site basis
135: to two states; a hole (transforming as the $[0,0]$ representation of
136: $U(1) \otimes SU(2)$) and a single spin (transforming as the $[1,1/2]$
137: representation of $U(1) \otimes SU(2)$). All the single site operators
138: are represented as $2 \times 2$ matrices acting on these states. The
139: block basis states used in each step of the DMRG iteration are
140: labelled $\vert \vert n j (\alpha) \rangle$, which denotes the
141: $\alpha$'th state of $n$ particles and total spin $j$. The matrix
142: elements of the single site operators are simply the {\em reduced
143: matrix elements} given by the Wigner-Eckart theorem
144: (Biedenharn and Louck 1981),
145: written here in our $U(1) \otimes SU(2)$ basis,
146: \begin{equation}
147: \langle n^\prime j^\prime m^\prime (\alpha^\prime)
148: \vert T^J_M \vert n j m (\alpha) \rangle \: = \:
149: C^{j J j^\prime}_{m M m^\prime}
150: \langle n^\prime j^\prime (\alpha^\prime) \vert \vert
151: \vektor{T}^J \vert \vert n j (\alpha) \rangle \; ,
152: \label{eq:WignerEckart}
153: \end{equation}
154: for the $M$'th component of an operator $\vektor{T}^J$ transforming
155: as the $[J]$ representation of $SU(2)$.
156: $C^{j^{} J j^{\prime}}_{m^{} M m^{\prime}}$ denotes the
157: Clebsch-Gordan coefficient. We have suppressed the trivial $U(1)$
158: label on the $\vektor{T}^J$ operator. This theorem specifies
159: the relationship between the reduced matrix elements acting on
160: the 2-state $U(1) \otimes SU(2)$ basis, and the 3-state
161: $U(1) \otimes U(1)$ basis of particle number and $z-$component
162: of spin \footnote{In a real calculation using the $U(1) \otimes U(1)$
163: basis, the $j$ index on the left hand side of Eq.\ (\ref{eq:WignerEckart})
164: cannot be used as a good quantum number since the usual DMRG density matrix
165: in this basis does not preserve any non-Abelian global symmetries.}.
166: 
167: In matrix form, choosing basis vectors $(1,0)$ to be a hole,
168: and $(0,1)$ to be a spin, The operators relevant to the t-J model are
169: \begin{eqnarray}
170: c^{[-1,1/2]} & = & \left( \begin{array}{cc} 0 & \sqrt{2} \\
171: 0 & 0 \end{array} \right) \nonumber \\
172: c^{\dagger[1,1/2]} & = & \left( \begin{array}{cc} 0 & 0 \\
173: 1 & 0 \end{array} \right) \nonumber \\
174: s^{[0,1]} & = & \left( \begin{array}{cc} 0 & 0 \\
175: 0 & \sqrt{3/4} \end{array} \right) \nonumber \\
176: n^{[0,0]} & = & \left( \begin{array}{cc} 0 & 0 \\
177: 0 & 1 \end{array} \right) \nonumber \\
178: p^{[0,0]} & = & \left( \begin{array}{cc} 1 & 0 \\
179: 0 & -1 \end{array} \right) \nonumber \; ,
180: \end{eqnarray}
181: where we use square brackets to denote which representation of
182: $U(1) \otimes SU(2)$ the operators transform as. $p^{[0,0]}$ is
183: the usual parity matrix used to enforce the correct commutation
184: relations on the DMRG matrix operators. We can understand the
185: unfamiliar matrix elements of the annihilation operator $c^{[-1,1/2]}$
186: from the Wigner-Eckart theorem: when taking the hermitian conjugate
187: of an operator $T$ that transforms as some representation of $SU(2)$,
188: what we really mean is to find the operator $T^\dagger$ such that
189: $\langle j^{\prime} m^{\prime} \vert T^{\dagger J}_M \vert j^{} m \rangle
190: = (-1)^{J-M} \langle j m \vert T^J_M \vert j^\prime m^\prime \rangle^*$.
191: Due to the action of the Clebsch-Gordan coefficients in the Wigner-Eckart
192: theorem, this is not the same as simply taking the Hermitian conjugate
193: of the reduced matrix operator itself. Our choice is to make the full
194: matrix elements of $c^{[-1,1/2]}$ identical with the usual $S^z$
195: representation - the matrix elements of $c^{\dagger[-1,1/2]}$ then
196: follow. The calculation of the matrix elements of the tensor product
197: of two operators acting on different sites is given by the angular
198: momentum theory as
199: \begin{equation}
200: \begin{array}{c}
201: \langle n^\prime j^\prime (\alpha^\prime_a \alpha^\prime_2
202: n^\prime_1 n^\prime_2 j^\prime_1 j^\prime_2)
203: \vert \vert \left[ \vektor{T}^{[N_1, k_1]} \otimes
204: \vektor{U}^{[N_2, k_2]} \right]^{[N,k]}
205: \vert \vert n j (\alpha_a \alpha_2 n_1 n_2 j_1 j_2) \rangle
206: ~ \\
207: ~ \\
208: = \: \left[
209: \begin{array}{ccc}
210: n_1  & n_2  & n \\
211: N_1  & N_2  & N \\
212: n\prime_1 & n^\prime_2 & n^\prime \\
213: \end{array}
214: \right]_{U(1)} \:
215: \left[
216: \begin{array}{ccc}
217: j_1  & j_2  & j \\
218: k_1  & k_2  & k \\
219: j\prime_1 & j^\prime_2 & j^\prime \\
220: \end{array}
221: \right]_{SU(2)}
222: \\
223: ~ \\
224: \times \: \langle n^\prime_1 j^{\prime}_1 (\alpha^\prime_1) \vert \vert
225: \vektor{T}^{[N_1,k_1]} \vert
226: \vert n_1 j_1 (\alpha_1) \rangle \:
227: \langle n^\prime_2 j^\prime_2 (\alpha^\prime_2)
228: \vert \vert \vektor{U}^{[N_2, k_2]} \vert
229: \vert n_2, j_2 (\alpha_2) \rangle \; .
230: \end{array}
231: \label{eq:TensorProduct}
232: \end{equation}
233: We have written this in a slightly convoluted way so as to make
234: explicit the algebraic structure. The coupling coefficients for
235: $U(1)$ are so trivial that they are not usually represented in this
236: form. Indeed,
237: \begin{equation}
238: \left[
239: \begin{array}{ccc}
240: n_1  & n_2  & n \\
241: N_1  & N_2  & N \\
242: n\prime_1 & n^\prime_2 & n^\prime \\
243: \end{array}
244: \right]_{U(1)}
245: = \;
246: \delta_{n^\prime_1, N_1+n_1} \delta_{n^\prime_2, N_2 + n_2}
247: \delta_{n^\prime, n+N}
248: \delta_{n, n_1+n_2} \delta_{N, N_1+N_2}
249: \delta_{n^\prime, n^\prime_1+n^\prime_2}
250: \; ,
251: \end{equation}
252: so the $U(1)$ component of Eq.\ (\ref{eq:TensorProduct}) reduces
253: to exactly the coupling used in traditional DMRG. The point of
254: departure from traditional DMRG is that the coupling coefficients
255: for $SU(2)$ are not so trivial:
256: \begin{equation}
257: \left[
258: \begin{array}{ccc}
259: j_1  & j_2  & j \\
260: k_1  & k_2  & k \\
261: j\prime_1 & j^\prime_2 & j^\prime \\
262: \end{array}
263: \right]_{SU(2)}
264: = \;
265: \sqrt{2j_1^\prime+1} \sqrt{2j_2^\prime+1}  \sqrt{2j+1} \sqrt{2k+1}
266: \left\{
267: \begin{array}{ccc}
268: j_1  & j_2  & j \\
269: k_1  & k_2  & k \\
270: j\prime_1 & j^\prime_2 & j^\prime \\
271: \end{array}
272: \right\}
273: \; ,
274: \end{equation}
275: where the $\left\{ \cdots \right\}$ term is the Wigner $9j$ coefficient.
276: 
277: We now write the Hamiltonian
278: so that the operators transform as representations of
279: the global symmetry group. For the t-J model, this is
280: \begin{equation}
281: H = -\sqrt{2} t \sum_{\langle i,j \rangle}
282: (c^{\dagger [1,1/2]}_i \cdot c^{[-1,1/2]}_j + \mbox{H.c.})
283: - \sqrt{3} J \sum_{\langle i,j \rangle} S^{[0,1]}_i \cdot S^{[0,1]}_j
284: - \frac{1}{4} J \sum_{\langle i,j \rangle} n^{[0,0]}_i \cdot n^{[0,0]}_j \; .
285: \end{equation}
286: By definition, the Hamiltonian itself transforms as the $[0,0]$
287: identity representation. We note that
288: the interaction between two nearest neighbour sites $(i,j)$ in
289: the non-Abelian representation requires summing 4 distinct terms,
290: $c^\dagger_i c_j$, $c_i c^\dagger_j$, $S_i S_j$ and $n_i n_j$. In
291: the $S^z$ basis, there are 8 terms:
292: $c^\dagger_{\uparrow i} c_{\uparrow j}$,
293: $c_{\uparrow i} c^\dagger_{\uparrow j}$,
294: $c^\dagger_{\downarrow i} c_{\downarrow j}$,
295: $c_{\downarrow i} c^\dagger_{\downarrow j}$,
296:  $S^+_i S^-_j$, $S^z_i S^z_j$, $S^-_i S^+_j$ and $n_i n_j$.
297: Thus, although the matrix elements of the single site operators
298: are more difficult to calculate using the non-Abelian formulation,
299: there are correspondingly fewer matrix elements and operators required.
300: 
301: We have applied this DMRG algorithm to the two dimensional t-J
302: model by unrolling the two dimensional lattice into a one dimensional
303: model with long range interactions, following the 'worm' approach
304: (Liang and Pang 1994). We use periodic boundary conditions in the
305: y direction and open boundary conditions in the (generally longer)
306: x direction. Ideally, we would like to perform the calculations
307: with periodic boundary conditions in both directions, but this would
308: substantially increase the number of states required. Although the
309: resulting one dimensional 'worm' model is reflection symmetric at
310: the midpoint of the lattice, it is difficult to make use of this
311: symmetry in DMRG calculations due to the non-uniform
312: nature of the ground state. As a DMRG sweep progresses from one
313: end of the system towards the centre point,
314: the holes and spins tend to distribute themselves
315: in a slightly asymmetric way between the left and
316: right halves of the system, so that when the centre point
317: is reached the left block basis is biased towards
318: states that have too few holes, and the right block basis
319: is biased towards states that have too many holes
320: (or vice versa). Enforcing reflection symmetry by using
321: only one block plus its spatial reflection thus leads
322: to a catastrophic reduction in the number of admissible
323: superblock states, and a corresponding jump in the energy at
324: that DMRG iteration. In all two dimensional models, constructing
325: the initial state for the finite DMRG sweeps is more problematic
326: than on one dimension. In one dimension, the so-called 'infinite'
327: DMRG algorithm produces quite good results by itself, and usually
328: provides a ground state that is qualitatively very similar to the
329: converged finite DMRG state. In two dimensions however, the reduced
330: translation symmetry of the unrolled one dimensional worm means
331: that a genuine 'infinite' style algorithm would be rather difficult
332: \footnote{Although there are interesting possibilities, see for
333: example, Henelius (1999).}.
334: While there are many possible ways to construct the initial
335: blocks, we use the simple approach of constructing the initial
336: blocks 'in place'; that is, starting from an initial 4 site
337: system consisting of the 2 extreme sites from the left and
338: right ends of the worm and adding two sites at a time,
339: one from each end of the worm, working towards the centre
340: of the system. This means that for most of the warm-up sweeps,
341: there are no interaction terms between the left and right blocks.
342: An alternative procedure is to rotate the system 90 degrees, so
343: that the opposite ends of the worm are connected on the periodic
344: boundary. However this introduces many more interactions between
345: the left and right blocks throughout the calculation, which impacts
346: on the accuracy. Another possibility is to introduce extra interactions
347: for the warm-up sweep only, but we have not yet investigated this option.
348: With no interactions between the two blocks, the eigenstates of the block
349: density matrix coincide with the eigenstates of the block Hamiltonian.
350: Thus there will only be a single non-zero density matrix eigenvalue
351: (more, if the ground state is degenerate). The effect is that, until
352: the first inter-block interaction appears, $m-1$ of the block eigenstates
353: are essentially random vectors. The initial state can be specified
354: further by manipulating the target state as a function of system
355: size. We have done this to obtain various initial
356: condition: a state with all holes uniformly distributed, a phase
357: separated state, and several random states.
358: 
359: We have made calculations for various lattice sizes, keeping up
360: to 1200 basis states per block. Table 1 shows a comparison of the
361: ground state energy as a function of the number of basis
362: states kept, using the $U(1) \otimes U(1)$ and $U(1) \otimes SU(2)$
363: basis, for a typical point in the 'striped' regime (White and
364: Scalapino 1999; Xiang, Lou and Su 2001).
365: The $SU(2)$ symmetry provides a saving of a factor of
366: two in the number of block states required. This is very
367: significant as the computational complexity of the
368: DMRG algorithm scales as at least $O(m^3)$. However, even
369: with 1200 states kept in the $U(1) \otimes SU(2)$ basis
370: (equivalent to around 2500 states in the $U(1) \times U(1)$ basis),
371: the achieved energy is around $0.25\%$ higher than the estimated
372: true ground state energy. This compares very poorly with the
373: accuracies generally achieved by DMRG for one dimensional models.
374: 
375: In DMRG, the ground state wavefunction is iteratively improved,
376: but only locally. This can lead to a situation where the DMRG
377: converges self-consistently to an incorrect state, depending
378: on the initial conditions and the details of the algorithm
379: (Scalapino and White 2000). In many cases,
380: for a small number of states (but still relatively large
381: compared with traditional DMRG studies) we have observed
382: qualitatively different DMRG wavefunctions, depending on
383: how we perform the warm-up sweep. We have also observed
384: different converged wavefunctions even with the same
385: initial condition, simply by varying the the rate at
386: which the number of basis states per block is increased
387: as the DMRG sweeps progress. For example, with the $16 \times 6$
388: system used in Table 1, using 500 basis states in the
389: $U(1) \otimes SU(2)$ basis, the ground state is most likely
390: a two stripe configuration in agreement with
391: (White and Scalapino 1999). However, if we increase the number
392: of retained states at a faster rate so that it takes fewer sweeps
393: to reach the final total, we actually obtain a three stripe
394: configuration. It is not until we increased the number of states
395: to 800 that the three stripe configuration moves out of this local
396: minima and formed the two stripe configuration. Simply doing more
397: DMRG sweeps with 500 states is not effective.
398: 
399: A possible way of dealing with this problem is to compare the
400: energies of the competing DMRG ground states. The problem with
401: this approach is that DMRG only provides a variational upper
402: bound on the energy, and that the goodness of the variational
403: energy (and therefore the truncation error associated with the
404: DMRG state) can depend significantly on the nature of the ground
405: state. Thus, this requires extrapolating the energies of both
406: states to zero truncation error, showing that the difference
407: in energy is statistically significant. This is a very time consuming
408: procedure. In this paper, we only report results where we have a
409: unique ground state, independent of the initial conditions (at
410: least for the initial conditions we have used). However, the
411: problem of multiple candidate ground states depending on the
412: initial conditions deteriorates rather quickly as the system
413: size increases. Indeed, one doesn't need to increase the system
414: size too far before the DMRG fails to converge to a believable
415: ground state at all, at least in a reasonable number of sweeps.
416: 
417: Fig.\ 1 shows the hole density along the x direction for a fixed
418: number of holes, as the system size is increased. As the size grows,
419: the stripes tend to move further apart while keeping approximately
420: the same width, although it is difficult to make any real conclusions
421: about the stripe width from this limited data. In particular, the
422: width of the fluctuations in the real space density may be much
423: larger than the correlation width of the stripe itself, if the
424: stripe is delocalized. That depends on how much effect the boundary
425: conditions have in pinning the stripes. Fig.\ 2 shows the effect
426: of reducing the number of holes to 6, for the $16 \times 6$ case.
427: The ground state we observed is curious in that it breaks spatial
428: reflection symmetry, but it provides evidence that it is
429: energetically favourable to form a 'normal' stripe of hole
430: density 4/6 and one 'proto' stripe, of hole density 2/6, rather
431: than two stripes of equal hole density, or a single stripe. This
432: suggests, as with the results from Fig.\ 1, that that the
433: thermodynamic hole density per stripe is a constant
434: (which depends on J/t).
435: 
436: In Fig.\ 3, we attempt to find the optimal filling per stripe,
437: by increasing the system size and number of holes by 50\%, to a
438: $24 \times 6$ lattice with 12 holes. Since we consistently obtain
439: three stripes, this limits the hole density per unit length of
440: the stripes to $0.5 < d < 1$. This is the largest system that we
441: could study, while still being reasonably certain that the obtained
442: ground state is substantially independent of the initial conditions.
443: Thus, for these parameters and boundary conditions, the two
444: dimensional t-J model almost certainly has a striped ground
445: state, and doping the system changes the density of stripes
446: while the number of holes per stripe remains constant.
447: It is difficult, however, to extrapolate these results to make
448: definite conclusions about the nature of the ground state
449: of the t-J model in the thermodynamic limit. Because of
450: the half-periodic boundary conditions, the hole density is
451: constant in the y direction. Thus any fluctuation in the hole
452: density across the system, pinned by the open boundary
453: in that direction, will appear as a vertical stripe in
454: these calculations. Other possible groundstates of the thermodynamic
455: t-J model, such as diagonal stripes or antiferromagnetic bubbles,
456: are not permitted by construction. We also note that in the phase
457: separated region, the holes are attracted to the open boundary of
458: the finite system. On the other hand, in the striped regime, the
459: holes are repelled by the boundary. Thus the open boundary may
460: well have a significant effect on the nature of the ground state
461: that we observe, and especially, on the critical value of $J/t$
462: that separates stripe formation from phase separation.
463: 
464: It is important to emphasize that stripe formation always implies
465: the presence of antiphase boundaries, i.e., antiphase domain walls
466: in the antiferromagnet. In all approaches, antiphase boundaries
467: are always found in conjunction with stripes: from a simple mean-field
468: calculations (Zaanen and Gunnarsson 1989) to the more sophisticated
469: quantum numerical approaches (White and Scalapino 1998;
470: Morais-Smith {\sl et al.} 1998; Pryadko {\sl et al.} 1999;
471: Martin {\sl et al.} 2000). From our numerical data
472: we cannot conclude that the antiphase boundaries are a consequence
473: of stripes or vice-versa. What we are certain of is that the presence
474: of the antiphase boundaries is a clear evidence of stripe existence.
475: This favours our earlier observation that stripes are different
476: from phase separation.
477: 
478: An interesting explanation of the antiphase boundaries was suggested
479: by Nagaev (1995) \footnote{For a review, see also Markiewicz (1997)
480: and references cited therein.}. This brings us back to the general
481: theory of metal-insulator transition, as formulated by Mott (1984),
482: who pointed out that the number of free carriers should jump
483: discontinuously at the transition. Hence, there has to be a region
484: of phase separation near the metal insulator transition. The existence
485: of phase separation associated with doping away from an antiferromagnetic
486: phase was recognized prior to the discovery of high temperature
487: superconductors (Visscher 1974; Nagaev 1983).
488: 
489: Nagaev refers to this phase as {\sl nanoscale} phase separation
490: (Nagaev 1995; Markiewicz 1997), where phase separation is accompanied
491: by charge separation, which in a perfect isotropic crystal, form an
492: almost periodic structure (Nagaev 1995). Thus,
493: a nanoscale phase separation is realized as a form of charge
494: density wave. In this language the antiphase boundaries appear
495: as the ferromagnetic droplets (Nagaev 1983, 1995):
496: a hole can delocalize over a finite domain by flipping the spins
497: of the neighbouring Cu ion, forming a small ferromagnetic -
498: ferron (Nagaev 1983, 1995) islands. It is interesting to
499: note that if this is the origin of the antiphase boundaries then
500: a second hole can lower its energy by localizing on the same
501: ferron island, i.e., on the same antiphase boundary. This has
502: the appearance of a real space pairing mechanism.
503: 
504: Hence, looking at the stripes as being a nanoscale phase separation,
505: the antiphase boundaries (periodic ferrons) will appear as a consequence
506: of the periodic hole structure (stripes). At mean-field level this
507: can be understood by recalling that the superposition (coexistence) of
508: a charge and spin density-wave will always gives rise to ferromagnetism
509: (Volkov, Kopaev and Rusianov 1973; Gulacsi and Gulacsi 1986).
510: 
511: Work in Australia was supported by the Australian Research Council.  One
512: of us (I. P. McCulloch) acknowledges the hospitality of Los Alamos
513: National Laboratory where much of the numerical work was performed.
514: 
515: \newpage
516: 
517: \section*{References}
518: 
519: Anderson, P. W., 1987, Science {\bf 235}, 1196.
520: 
521: Barnes, S. E., 1989, Phys. Rev. B {\bf 40}, 723.
522: 
523: Biedenharn, L. C. and Louck, J. D., 1981, {\it Angular Momentum in
524: Quantum Physics}, Addison-Wesley, Massachusetts.
525: 
526: Blatt, J. M. and Weisskopf, V. F., 1979, {\it Theoretical
527: Nuclear Physics}, Springer-Verlag, New York.
528: 
529: Emery, V. J., Kivelson, S. A., and Lin, H. Q., 1990,
530: Phys. Rev. Lett. {\bf 64}, 475.
531: 
532: Gulacsi, M. and Gulacsi, Zs., 1986, Phys. Rev. B{\bf 33}, 6147.
533: 
534: Hellberg, C. S. and Manousakis, E., 1999, Phys. Rev. Lett. {\bf 83}, 132.
535: 
536: Henelius, P., 1999,  Phys. Rev. B{\bf 60}, 9561.
537: 
538: Itoyama, H., McCoy, B. M. and Perk, J. H. H., 1990,
539: Int. Jour. Mod. Phys. B{\bf 4}, 295.
540: 
541: Kivelson, S.A., Emery, V.J., and Lin, H. Q., 1990, Phys. Rev.
542: Lett. {\bf 64}, 475.
543: 
544: Liang, S. and Pang, H., 1994, Phys. Rev. B{\bf 49}, 9214.
545: 
546: Markiewicz, R. S., 1997, J. Phys. Chem. Solids {\bf 58}, 1179.
547: 
548: Martin, G. B., Gazza, C., Xavier, J. C., Feiguin, A. and Dagotto,
549: E., 2000, Phys. Rev. Lett. {\bf 84}, 5844.
550: 
551: McCulloch, I. P. and Gulacsi, M., 2000, cond-mat/0012319.
552: 
553: Morais-Smith, M., Dimashko, Y. M., Hasselmann, N. and Caldeiro,
554: A. O., 1998, Phys. Rev. B{\bf 58}, 453.
555: 
556: Mott, N. F., 1984, Phil. Mag. B{\bf 50}, 161.
557: 
558: Nagaev, E. L., 1983, {\it Physics of Magnetic Semiconductors},
559: Mir, Moscow.
560: 
561: Nagaev, E. L., 1995, Physics - Uspekhi {\bf 38}, 497.
562: 
563: Proceedings, 1998, {\it Conference on Spectroscopies in Novel
564: Superconductors}, J. Phys. Chem. {\bf 59}, No. 10 - 12.
565: 
566: Pryadko, L. P., Kivelson, S. A., Emery, V. J., Bazaliy, Y. B. and
567: Demler, E. A., 1999, Phys. Rev. Lett. {\bf 60}, 7541.
568: 
569: Scalapino, D. J. and White, S. R., 2000, cond-mat/0007515.
570: 
571: Sierra, G. and Nishino, T., 1997, Nucl. Phys. B{\bf 495}, 505.
572: 
573: Visscher, P. B., 1974, Phys. Rev. B{\bf 10}, 943.
574: 
575: Volkov, B. A., Kopaev, Y. V. and Rusianov, A. I., 1973, Zh. Eksp.
576: Teor. Fiz. {\bf 65}, 1984 [Sov. Phys. - JETP {\bf 38}, 991 (1974)].
577: 
578: White, S. R., 1992, Phys. Rev. Lett. {\bf 69}, 2863.
579: 
580: White, S. R., 1993, Phys. Rev. B{\bf 48}, 10, 345.
581: 
582: White, S. R. and Scalapino, D. J., 1999, cond-mat/9907243.
583: 
584: Xiang, T., Lou, J. and Su, Z., 2001, cond-mat/0102200
585: 
586: Yosida, K, 1980, {\it Functional Analysis}, Springer-Verlag, Berlin.
587: 
588: Zaanen, J. and Gunnarsson, O., 1989, Phys. Rev. B {\bf 40}, 7391.
589: 
590: Zaanen, J., Horbach, M. L., van Saarloos, W., 1996, Phys. Rev. B
591: {\bf 53}, 8671.
592: 
593: Zhang, F. C. and Rice, T. M., 1988, Phys. Rev. B{\bf 37}, 3759.
594: 
595: 
596: \table{
597: \caption{
598: Comparison of $U(1)$ and $SU(2)$ bases for the number of states versus
599: ground state energy of a $16 \times 6$ t-J system with $J = 0.35$,
600: $t = 1$, 8 holes and cylindrical boundary conditions. The results using
601: the $U(1)$ basis are from reference (White and Scalapino 1999)
602: We also include an estimate of the true energy, extrapolated to zero
603: truncation error.
604: }
605: \[
606: \begin{array}{rr r@{.}l}
607: \hline \hline
608: \mbox{basis} & \multicolumn{1}{c}{m} & \multicolumn{1}{c}{E} \\ \hline
609: U(1)         & 1000 & -52 & 279        \\
610: SU(2)        & 500 & -52 & 284 \\
611: SU(2)        & 800 & -52 & 463 \\
612: SU(2)        & 1200 & -52 & 520 \\
613: \multicolumn{1}{c}{-} & \infty & -52 & 65 \pm 0.05 \\
614: \hline \hline
615: \end{array}
616: \]
617: }
618: 
619: \newpage
620: 
621: \section*{Figure Captions}
622: 
623: Figure 1. Hole density in the $x$ direction for lattice sizes $16 \times 6$, $18 \times 6$
624: and $20 \times 6$, 8 holes, at $J/t = 0.35$.
625: 
626: Figure 2. Hole density in the $x$ direction for lattice size $16 \times 6$, 6 holes,
627: $J/t = 0.35$.
628: 
629: Figure 3. Hole density in the $x$ direction for lattice size $24 \times 6$, 12 holes,
630: $J/t = 0.35$.
631: 
632: \end{document}
633: 
634: