cond-mat0702443/vbs.tex
1: \documentclass[showpacs,twocolumn,floatfix]{revtex4}
2: 
3: \usepackage{amsmath,amssymb,mathrsfs}
4: \usepackage{latexsym}
5: \usepackage{graphicx,psfrag} 
6: \usepackage[dvips]{color}
7: 
8: % boldsymbol (requires amsmath)
9: \newcommand{\bs}[1]{\boldsymbol{#1}}
10: \newcommand{\nn}{\nonumber}
11: \newcommand{\mb}[1]{\mathbf{#1}}
12: % 1/2
13: \newcommand{\half}{$\frac{1}{2}$ }
14: % commutator and anticommutator
15: \newcommand{\comm}[2]{\left[#1,#2\right]}
16: \newcommand{\anticomm}[2]{\left\{#1,#2\right\}}
17: % notation for vacuum, an empty set inside a ket
18: \newcommand{\vac}{\left|\,0\,\right\rangle}
19: % define the bra and ket, and braket
20: \newcommand{\ket}[1]{\left|#1\right\rangle}
21: \newcommand{\bra}[1]{\left\langle#1\right|}
22: \newcommand{\braket}[2]{\bigl\langle#1\bigl|\bigr.#2\bigr\rangle}
23: % simplifies using the up and down arrows to denote spin
24: \newcommand{\up}{\uparrow}
25: \newcommand{\dw}{\downarrow}
26: \newcommand{\cas}{\mathcal{C}^2_{\text{SU($n$)}}}
27: \newcommand{\casi}[2]{\mathcal{C}^2_{\text{SU(3)}}(#1,#2)}
28: \newcommand{\casin}[1]{\mathcal{C}^2_{\text{SU($n$)}}(#1)}
29: \newcommand{\casinn}[2]{\mathcal{C}^2_{\text{SU($#1$)}}(#2)}
30: \newcommand{\dimn}[2]{\text{dim}_{\text{SU($#1$)}}(#2)}
31: 
32: \def\ie{\emph{i.e.},\ }
33: \def\eg{\emph{e.g.}\ }
34: \def\ea{\emph{et al.}}
35: 
36: \def\b{\text{b}}
37: \def\r{\text{r}}
38: \def\g{\text{g}}
39: \def\y{\text{y}}
40: \def\c{\text{c}}
41: \def\m{\text{m}}
42: 
43: \allowdisplaybreaks[1]\begin{document}
44: \title{Valence bond solids for SU($\bs{n}$) spin chains:\\
45: exact models, spinon confinement, and the Haldane gap}
46: \author{Martin Greiter and Stephan Rachel}
47: \affiliation{Institut f\"ur Theorie der Kondensierten Materie,\\
48:   Universit\"at Karlsruhe, Postfach 6980, 76128 Karlsruhe, Germany}
49: \pagestyle{plain}
50: 
51: \begin{abstract}
52:   To begin with, we introduce several exact models for SU(3) spin
53:   chains: (1) a translationally invariant parent Hamiltonian involving
54:   four-site interactions for the trimer chain, with a three-fold
55:   degenerate ground state.  We provide numerical evidence that the
56:   elementary excitations of this model transform under representation
57:   $\bs{\bar 3}$ of SU(3) if the original spins of the model transform
58:   under rep.\ $\bs{3}$.  (2) a family of parent Hamiltonians for
59:   valence bond solids of SU(3) chains with spin reps.\ $\bs{6}$,
60:   $\bs{10}$, and $\bs{8}$ on each lattice site.  We argue that of
61:   these three models, only the latter two exhibit spinon confinement
62:   and hence a Haldane gap in the excitation spectrum.  We generalize
63:   some of our models to SU($n$).  Finally, we use the emerging rules
64:   for the construction of VBS states to argue that models of
65:   antiferromagnetic chains of SU($n$) spins in general possess a
66:   Haldane gap if the spins transform under a representation
67:   corresponding to a Young tableau consisting of a number of boxes
68:   $\lambda$ which is divisible by $n$.  If $\lambda$ and $n$ have no
69:   common divisor, the spin chain will support deconfined spinons and
70:   not exhibit a Haldane gap.  If $\lambda$ and $n$ have a common
71:   divisor different from $n$, it will depend on the specifics of the
72:   model including the range of the interaction.
73: \end{abstract}
74: 
75: \pacs{75.10.Jm, 75.10.Pq, 75.10.Dg, 32.80.Pj}
76: 
77: %75.10.Jm Quantized spin models
78: %75.10.Pq Spin chain models
79: %75.10.Dg Spin Hamiltonians
80: 
81: \maketitle
82: 
83: \section{Introduction}
84: 
85: Quantum spin chains have been a most rewarding subject of study almost
86: since the early days of quantum mechanics, beginning with the
87: invention of the Bethe ansatz in 1931~\cite{bethe31zp205} as a method
88: to solve the $S=\frac{1}{2}$ Heisenberg chain with nearest-neighbor
89: interactions.  The method led to the discovery of the Yang--Baxter
90: equation in 1967~\cite{yang67prl1312, baxter90}, and provides the
91: foundation for the field of integrable
92: models~\cite{KorepinBogoliubovIzergin93}.  Faddeev and
93: Takhtajan~\cite{faddeev-81pla375} discovered in 1981 that the
94: elementary excitations (now called spinons) of the spin-$1/2$
95: Heisenberg chain carry spin $1/2$ while the Hilbert space is spanned
96: by spin flips, which carry spin 1.  The fractional quantization of
97: spin in spin chains is conceptually similar to the fractional
98: quantization of charge in quantized Hall
99: liquids~\cite{laughlin83prl1395, stone92}.  In 1982,
100: Haldane~\cite{haldane83pla464, haldane83prl1153} identified the O(3)
101: nonlinear sigma model as the effective low-energy field theory of
102: SU(2) spin chains, and argued that chains with integer spin possess a
103: gap in the excitation spectrum, while a topological term renders
104: half-integer spin chains gapless~\cite{affleck90proc,Fradkin91}.
105: 
106: The general methods---the Bethe ansatz method and the use of effective
107: field theories including
108: bosonization~\cite{GogolinNersesyanTsvelik98,Giamarchi04}---are
109: complemented by a number of exactly solvable models, most prominently
110: among them the Majumdar--Ghosh (MG) Hamiltonian for the
111: $S=\frac{1}{2}$ dimer chain~\cite{majumdar-69jmp1399}, the AKLT model
112: as a paradigm of the gapped $S=1$ chain~\cite{affleck-87prl799,
113:   affleck-88cmp477}, and the Haldane--Shastry model
114: (HSM)~\cite{haldane88prl635, shastry88prl639, haldane91prl1529,
115:   haldane-92prl2021}.  The HSM is by far the most sophisticated among
116: these three, as it is not only solvable for the ground state, but
117: fully integrable due to its Yangian symmetry~\cite{haldane-92prl2021}.
118: The wave functions for the ground state and single-spinon excitations
119: are of a simple Jastrow form, elevating the conceptual similarity to
120: quantized Hall states to a formal equivalence.  Another unique feature
121: of the HSM is that the spinons are free in the sense that they only
122: interact through their half-Fermi statistics~\cite{haldane91prl937,
123:   ha-93prb12459, essler95prb13357, greiter-05prb224424,greiter06prl},
124: which renders the model an ideal starting point for any perturbative
125: description of spin systems in terms of interacting
126: spinons~\cite{greiter-06prl}.  The HSM has been generalized from SU(2)
127: to SU($n$)~\cite{kawakami92prb1005, kawakami92prbr3191, ha-92prb9359, 
128:   bouwknegt-96npb345, schuricht-05epl987, schuricht-06prb235105}.
129: 
130: For the MG and the AKLT model, only the ground states are known
131: exactly.  Nonetheless, these models have amply contributed to our
132: understanding of many aspects of spin chains, each of them through the
133: specific concepts captured in its ground state~\cite{%use of MG:
134:   jullien-83baps344, affleck89jpcm3047, okamoto-92pla433, eggert96prb9612, 
135:   white-96prb9862,sen-07prb104411,
136:   knabe88jsp627,fannes-89epl633,
137:   freitag-91zpb381,%received 12.02.91
138:   kluemper-91jpal955,%received 03.05.91
139:   kennedy-92prb304,%received 29.07.91
140:   kluemper-92zpb281,%received 13.02.92
141:   kluemper-93epl293,batchelor-94ijmpb3645,
142:   schollwock-96prb3304, kolezhuk-96prl5142, kolezhuk-02prb100401,
143:   normand-02prb104411,lauchli-06prb144426}.
144: The models are specific to SU(2) spin chains.  We will review both 
145: models below.
146: 
147: In the past, the motivation to study SU($n$) spin systems with $n>2$
148: has been mainly formal.  The Bethe ansatz method has been generalized
149: to multiple component systems by
150: Sutherland~\cite{sutherland75prb3795}, yielding the so-called nested
151: Bethe ansatz. In particular, this has led to a deeper understanding
152: of quantum integrability and the applicability of the Bethe
153: ansatz~\cite{choi-82pla83}.  Furthermore, the nested Bethe ansatz was
154: used to study the spectrum of the SU($n$)
155: HSM~\cite{kawakami92prb1005,ha-93prb12459}.  It has also been applied
156: to SU(2) spin chains with orbital degeneracy at the SU(4) symmetric
157: point~\cite{Li-99prb12781,Gu-02prb092404}.
158: %
159: Most recently, Damerau and Kl\"umper obtained highly accurate
160: numerical results for the thermodynamic properties of the SU(4)
161: spin--orbital model~\cite{damerau-06jsm12014}.
162: %
163: SU($n$) Heisenberg models have been studied recently by Kawashima and
164: Tanabe~\cite{kawashima-07prl057202} with quantum Monte Carlo, and by
165: Paramekanti and Marston~\cite{Paramekanti-06cm0608691} using
166: variational wave functions.
167: 
168: The effective field theory description of SU(2) spin chains by Haldane
169: yielding the distinction between gapless half-integer spin chains with
170: deconfined spinons and gapped integer spin chains with confined
171: spinons cannot be directly generalized to SU($n$), as there is no
172: direct equivalent of the CP$^1$ representation used in Haldane's
173: analysis.
174: %
175: The critical behavior of SU($n$) spin chains, however, has been
176: analyzed by Affleck in the framework of effective field
177: theories~\cite{affleck86npb409, affleck88npb582}.
178: 
179: An experimental realization of an SU(3) spin system, and in particular
180: an antiferromagnetic SU(3) spin chain, however, might be possible in
181: an optical lattice of ultracold atoms in the not-too-distant future.
182: The ``spin'' in these systems would of course not relate to the
183: rotation group of our physical space, but rather relate to SU(3)
184: rotations in an internal space spanned by three degenerate ``colors''
185: the atom may assume, subject to the requirement that the number of
186: atoms of each color is conserved.  A possible way to realize such a
187: system experimentally is described in Appendix~\ref{sec:exp}.
188: Moreover, it has been suggested recently that an SU(3) trimer state
189: might be realized approximately in a spin tetrahedron
190: chain~\cite{chen-05prb214428,chen-06prb174424}.
191: 
192: Motivated by both this prospect as well as the mathematical challenges
193: inherent to the problem, we propose several exact models for SU(3)
194: spin chains in this article.  The models are similar in spirit to the
195: MG or the AKLT model for SU(2), and consist of parent Hamiltonians and
196: their exact ground states.  There is no reason to expect any of these
197: models to be integrable, and none of the excited states are known
198: exactly.  We generalize several of our models to SU($n$), and use the
199: emerging rules to investigate and motivate which SU($n$) spin chains
200: exhibit spinon confinement and a Haldane gap.
201: 
202: The article is organized as follows.  Following a brief review of the
203: MG model in Sec.~\ref{sec:mg}, we introduce the trimer model for SU(3)
204: spin chains in Sec.~\ref{sec:trimer}.  This model consists of a
205: translationally invariant Hamiltonian involving four-site
206: interactions, with a three-fold degenerate ground state, in which
207: triples of neighboring sites form SU(3) singlets (or trimers).  In
208: Sec.~\ref{sec:su3rep}, we review the representations of SU(3), which
209: we use to verify the trimer model in Sec.~\ref{sec:trimercont}.  In
210: this section, we further provide numerical evidence that the
211: elementary excitations of this model transform under representation
212: $\bs{\bar 3}$ of SU(3) if the original spins of the model transform
213: under rep.\ $\bs{3}$.  We proceed by introducing Schwinger bosons in
214: Sec.~\ref{sec:schwinger} and a review of the AKLT model in
215: Sec.~\ref{sec:aklt}.  In Sec.~\ref{sec:vbs}, we formulate a family of
216: parent Hamiltonians for valence bond solids of SU(3) chains with spin
217: reps.\ $\bs{6}$, $\bs{10}$, and $\bs{8}$ on each lattice site, and
218: proof their validity.  We argue that only the rep.\ $\bs{10}$ and the
219: rep.\ $\bs{8}$ model, which are in a wider sense generalizations of
220: the AKLT model to SU(3), exhibit spinon confinement and hence a
221: Haldane-type gap in the excitation spectrum.  In Sec.~\ref{sec:sun},
222: we generalize three of our models from SU(3) to SU($n$).
223: %
224: In Sec.~\ref{sec:conf}, we use the rules emerging from the numerous
225: VBS models we studied to investigate which models of SU($n$) spin
226: chains in general exhibit spinon confinement and a Haldane gap.  In
227: this context, we first review a rigorous theorem due to Affleck and
228: Lieb~\cite{affleck-86lmp57} in Sec.\ \ref{sec:afflecklieb}.  In Sec.\
229: \ref{sec:generalcriterion}, we argue that the spinons in SU($n$) spin
230: chains with spins transforming under reps.\ with Young tableaux
231: consisting of a number of boxes $\lambda$ which is divisible by $n$
232: are always confined.  In Sec.\ \ref{sec:examples}, we construct
233: several specific examples to argue that if $\lambda$ and $n$ have a
234: common divisor different from $n$, the model will be confining only if
235: the interactions are sufficiently long ranged.  Specifically, the
236: models we study suggest that if $q$ is the largest common divisor of
237: $\lambda$ and $n$, the model will exhibit spinon confinement only if
238: the interactions extends at least to the $\frac{n}{q}$-th neighbor on
239: the chain.  If $\lambda$ and $n$ have no common divisor, the spinons
240: will be free and chain will not exhibit a Haldane gap.  We briefly
241: summarize the different categories of models in Sec.\ \ref{sec:sum},
242: and present a counter-example to the general rules in Sec.\
243: \ref{sec:counter}.  We conclude with a brief summary of the results 
244: obtained in this article in 
245: Sec.\ \ref{sec:conclusion}.
246: 
247: A brief and concise account of the SU(3) VBS models we elaborate here 
248: has been given previously~\cite{greiter-07prb060401}.
249: 
250: 
251: \section{The Majumdar--Ghosh model}
252: \label{sec:mg}
253: 
254: Majumdar and Ghosh~\cite{majumdar-69jmp1399} noticed in 1969 that on a
255: linear spin $S=\frac{1}{2}$ chain with an even number of sites, the
256: two valence bond solid or dimer states
257: \begin{eqnarray}
258: \big|\psi_{\text{MG}}^{\textrm{even}\rule{0pt}{5pt}\atop 
259: \textrm{(odd)}}\big\rangle 
260: &=&
261: \prod_{i\ \textrm{even}\atop (i\ \textrm{odd})}  
262: \Bigl(c^\dagger_{i\up} c^\dagger_{i+1\dw} - c^\dagger_{i\dw} c^\dagger_{i+1\up}
263: \Bigr) \vac =\nonumber \\
264: &=&\left\{ \begin{array}{lc}
265: \ket{\hbox{\begin{picture}(102,8)(-4,-3)
266: \put(13,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
267: \put(41,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
268: \put(69,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
269: \put(6,0){\circle{4}}
270: \put(20,0){\circle{4}}
271: \put(34,0){\circle{4}}
272: \put(48,0){\circle{4}}
273: \put(62,0){\circle{4}}
274: \put(76,0){\circle{4}}
275: \put(90,0){\circle{4}}
276: %\put(104,0){\circle{4}}
277: %\put(118,0){\circle{4}}
278: \end{picture}}} 
279: &\quad\textrm{``even''}\quad\\
280: %\ket{\hbox{\begin{picture}(132,8)(-4,-3)
281: \ket{\hbox{\begin{picture}(102,8)(-4,-3)
282: %\put(4,0){\makebox(0,0)[r]{\rule{4.pt}{ 0.8pt}}}
283: \put(27,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
284: \put(55,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
285: \put(83,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
286: %\put(111,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
287: \put(6,0){\circle{4}}
288: \put(20,0){\circle{4}}
289: \put(34,0){\circle{4}}
290: \put(48,0){\circle{4}}
291: \put(62,0){\circle{4}}
292: \put(76,0){\circle{4}}
293: \put(90,0){\circle{4}}
294: %\put(104,0){\circle{4}}
295: %\put(118,0){\circle{4}}
296: \end{picture}}} 
297: &\quad\textrm{``odd''}\quad\rule{0pt}{15pt}
298: \end{array}\right.\rule{0pt}{25pt}
299: \label{eq:psimg}
300: \end{eqnarray}
301: where the product runs over all even sites $i$ for one state and over
302: all odd sites for the other, are exact zero-energy ground
303: states~\cite{broek80pla261} of the parent Hamiltonian
304: \begin{equation}
305: H_{\text{MG}} = 
306: \sum_i \left(\boldsymbol{S}_i \boldsymbol{S}_{i+1} +
307: \frac{1}{2}\boldsymbol{S}_i \boldsymbol{S}_{i+2} +\frac{3}{8}\right),
308:  \label{eq:hmg}
309: \end{equation}
310: where
311: \begin{equation}
312: \boldsymbol{S}_i=\frac{1}{2}\,\sum_{\tau,\tau'=\up,\dw}\,
313: c_{i\tau}^{\dagger} \boldsymbol{\sigma}_{\tau\tau'}^{\phantom{\dagger}}
314: c_{i\tau'}^{\phantom{\dagger}}, 
315: \label{eq:s}
316: \end{equation}
317: and $\boldsymbol{\sigma}=(\sigma_x,\sigma_y,\sigma_z)$ is the vector
318: consisting of the three Pauli matrices.
319: 
320: The proof is exceedingly simple.  We rewrite
321: \begin{equation}
322: H_{\textrm{MG}}=\frac{1}{4}\sum_i H_i,\ \
323: H_i=\bigl(\boldsymbol{S}_i +\boldsymbol{S}_{i+1} 
324: +\boldsymbol{S}_{i+2}\bigr)^2 -\frac{3}{4}.
325: \end{equation}
326: Clearly, any state in which the total spin of three neighboring spins
327: is $\frac{1}{2}$ will be annihilated by $H_i$.  (The total spin can
328: only be $\frac{1}{2}$ or $\frac{3}{2}$, as
329: $\bf\frac{1}{2}\otimes\frac{1}{2}\otimes\frac{1}{2}=
330: \frac{1}{2}\oplus\frac{1}{2}\oplus\frac{3}{2}$.)  In the dimer states
331: above, this is always the case as two of the three neighboring spins
332: are in a singlet configuration, and %; the total spin is hence
333: $\bf 0\otimes\frac{1}{2}=\frac{1}{2}$.  Graphically, we may express
334: this as
335: \begin{equation}
336: H_i \ket{\hbox{\begin{picture}(40,8)(0,-3)
337: %\put(7,0){\line(1,0){12}}
338: \put(13,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
339: \put(6,0){\circle{4}}
340: \put(20,0){\circle{4}}
341: \put(34,0){\circle{4}}
342: \end{picture}}}
343: = H_i \ket{\hbox{\begin{picture}(40,8)(0,-3)
344: %\put(7,0){\line(1,0){12}}
345: \put(27,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
346: \put(6,0){\circle{4}}
347: \put(20,0){\circle{4}}
348: \put(34,0){\circle{4}}
349: \end{picture}}}=0.
350: \end{equation}
351:  As $H_i$ is positive definite, the two zero-energy eigenstates of
352: $H_{\textrm{MG}}$ are also ground states. 
353: 
354: Is the Majumdar--Ghosh or dimer state in the universality class
355: generic to one-dimensional spin-\half liquids, and hence a
356: useful paradigm to understand, say, the nearest-neighbor Heisenberg
357: chain?  The answer is clearly no, as the dimer states (\ref{eq:psimg})
358: violate translational symmetry modulo translations by two lattice
359: spacings, while the generic liquid is invariant.  
360: 
361: Nonetheless, the dimer chain shares some important properties of this
362: generic liquid.  First, the spinon excitations---here domain walls
363: between ``even'' and ``odd'' ground states---are deconfined.  (To
364: construct approximate eigenstates of
365: $H_{\textrm{MG}}$, we take momentum superpositions
366: of localized domain walls.)  Second, there are (modulo the overall
367: two-fold degeneracy) only $M+1$ orbitals available for an individual
368: spinon if $2M$ spins are condensed into dimers or valence bond
369: singlets.  This is to say, if there are only a few spinons in a long
370: chain, the number of orbitals available to them is roughly half the
371: number of sites.  This can easily be seen graphically:
372: \begin{center}
373: \begin{picture}(222,25)(-26,-17)
374: \put(-12,0){\makebox(0,0)[r]{\rule{10.pt}{ 0.8pt}}}
375: \put(13,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
376: \put(41,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
377: %\put(55,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
378: %\put(69,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
379: \put(83,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
380: %\put(97,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
381: \put(111,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
382: %\put(125,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
383: \put(139,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
384: %\put(153,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
385: \put(181,0){\makebox(0,0){\rule{10.pt}{ 0.8pt}}}
386: %\put(204,0){\makebox(0,0)[l]{\rule{7.pt}{ 0.8pt}}}
387: \put(-24,0){\circle{4}}
388: \put(-10,0){\circle{4}}
389: \put(6,0){\circle{4}}
390: \put(20,0){\circle{4}}
391: \put(34,0){\circle{4}}
392: \put(48,0){\circle{4}}
393: \put(62,0){\circle{4}}
394: \put(76,0){\circle{4}}
395: \put(90,0){\circle{4}}
396: \put(104,0){\circle{4}}
397: \put(118,0){\circle{4}}
398: \put(132,0){\circle{4}}
399: \put(146,0){\circle{4}}
400: \put(160,0){\circle{4}}
401: \put(174,0){\circle{4}}
402: \put(188,0){\circle{4}}
403: %\put(202,0){\circle{4}}
404: \put(62,1){\makebox(0,0){\vector(0,1){14}}}
405: \put(160,1){\makebox(0,0){\vector(0,1){14}}}
406: \put(14,-12){\makebox(0,0){\small even}}
407: \put(111,-12){\makebox(0,0){\small odd}}
408: \end{picture}
409: \end{center}
410: If we start with an even ground state on the left, the spinon to its
411: right must occupy an even lattice site and vice versa.  The resulting
412: state counting is precisely what one finds in the Haldane--Shastry
413: model, where it is directly linked to the half-Fermi statistics of the
414: spinons~\cite{haldane91prl937}.
415: 
416: The dimer chain is further meaningful as a piece of a general
417: paradigm.  The two degenerate dimer states (\ref{eq:psimg}) can be
418: combined into an $S=1$ chain, the AKLT chain, which serves as a
419: generic paradigm for $S=1$ chains which exhibit the Haldane
420: gap~\cite{haldane83pla464, haldane83prl1153, affleck89jpcm3047},
421: and provides the intellectual background for several of the exact models
422: we introduce further below.  Before doing so, however, we will now
423: introduce the trimer model, which constitutes an SU(3) analog of the MG 
424: model.
425: 
426: 
427: \section{The trimer model}
428: \label{sec:trimer}
429: 
430: \subsection{The Hamiltonian and its ground states} 
431: \label{sec:trimerstated}
432: 
433: Consider a chain with $N$ lattice sites, where $N$ has to be
434: divisible by three, and periodic boundary conditions (PBCs). On each
435: lattice site we place an SU(3) spin which transforms under the
436: fundamental representation $\bs{3}$, \ie the spin can take
437: the values (or colors) blue (b), red (r), or green (g). The trimer
438: states are obtained by requiring the spins on each three neighboring
439: sites to form an SU(3) singlet, which we call a trimer and sketch it
440: by
441: \begin{picture}(38,8)(1,-2)\linethickness{0.8pt}
442: \put(6,0){\circle{4}}\put(8,0){\line(1,0){10}}\put(20,0){\circle{4}}
443: \put(22,0){\line(1,0){10}}\put(34,0){\circle{4}}
444: \end{picture}.
445: The three %orthogonal 
446: linearly independent trimer %ground 
447: states on the chain are given by
448: \begin{eqnarray}\label{trimerstates.3}
449: \ket{\psi_{\text{trimer}}^{(\mu)}}
450: \!&\!=\!&\!\left\{
451: \begin{array}{l}
452: |\hbox{
453: \setlength{\unitlength}{1pt}
454: \begin{picture}(80,8)(4,-2)\linethickness{0.8pt}
455: \put(6,0){\circle{4}}\put(8,0){\line(1,0){10}}\put(20,0){\circle{4}}
456: \put(22,0){\line(1,0){10}}\put(34,0){\circle{4}}\put(48,0){\circle{4}}
457: \put(50,0){\line(1,0){10}}\put(62,0){\circle{4}}\put(64,0){\line(1,0){10}}
458: \put(76,0){\circle{4}}
459: \end{picture}
460: }\rangle\,\equiv\,\ket{\psi_{\text{trimer}}^{(1)}},\\[3mm]
461: |\hbox{
462: \setlength{\unitlength}{1pt}
463: \begin{picture}(80,8)(4,-2)\linethickness{0.8pt}
464: %\put(0,0){\line(1,0){4}}
465: \put(6,0){\circle{4}}\put(20,0){\circle{4}}
466: \put(22,0){\line(1,0){10}}\put(34,0){\circle{4}}\put(36,0){\line(1,0){10}}
467: \put(48,0){\circle{4}}\put(62,0){\circle{4}}\put(64,0){\line(1,0){10}}
468: \put(76,0){\circle{4}}
469: %\put(78,0){\line(1,0){4}}
470: \end{picture}
471: }\rangle\,\equiv\,\ket{\psi_{\text{trimer}}^{(2)}},\\[3mm]
472: |\hbox{
473: \setlength{\unitlength}{1pt}
474: \begin{picture}(80,8)(4,-2)\linethickness{0.8pt}
475: %\put(0,0){\line(1,0){4}}
476: \put(6,0){\circle{4}}\put(8,0){\line(1,0){10}}\put(20,0){\circle{4}}
477: \put(34,0){\circle{4}}\put(36,0){\line(1,0){10}}\put(48,0){\circle{4}}
478: \put(50,0){\line(1,0){10}}\put(62,0){\circle{4}}\put(76,0){\circle{4}}
479: %\put(78,0){\line(1,0){4}}
480: \end{picture}
481: }\rangle\,\equiv\,\ket{\psi_{\text{trimer}}^{(3)}}\,.
482: \end{array}\right.\\[-2mm]\nn
483: \end{eqnarray}
484: Introducing operators $c_{i\sigma}^{\dagger}$ which create a fermion
485: of color $\sigma$ ($\sigma=\b,\r,\g$) at lattice site $i$, the trimer
486: states can be written as
487: \begin{equation}
488:   \label{eq:trimer}
489:   \ket{\psi_{\text{trimer}}^{(\mu)}\!}=\hspace{-15pt}\prod_{\scriptstyle{i} \atop 
490:     \left(\scriptstyle{\frac{i-\mu}{3}\,{\rm integer}}\right)}
491:   \hspace{-5pt}\Bigl(
492:   \sum_{\scriptstyle{(\alpha,\beta,\gamma)}=\atop\scriptstyle{{\pi}(\b,\r,\g)}} 
493:   \hspace{-5pt}\hbox{sign}({\pi})\,
494:   c^{\dagger}_{i\,\alpha}\,c^{\dagger}_{i+1\,\beta}\,c^{\dagger}_{i+2\,\gamma}
495:   \Bigr)\!\vac\!,
496: \end{equation}
497: where $\mu=1,2,3$ labels the three degenerate ground states, and $i$ runs 
498: over the lattice sites subject to the constraint that $\frac{i-\mu}{3}$
499: is integer.  The sum extends over all six permutations $\pi$ of 
500: the three colors b, r, and g, \ie
501: \begin{equation}
502:   \label{eq:permsum}
503:   \begin{split}  
504:   &\sum_{\scriptstyle{(\alpha,\beta,\gamma)}=\scriptstyle{{\pi}(\b,\r,\g)}} 
505:   \hspace{-5pt}\hbox{sign}({\pi})\,
506:   c^{\dagger}_{i\,\alpha}\,c^{\dagger}_{i+1\,\beta}\,c^{\dagger}_{i+2\,\gamma}\\
507:   &\hspace{18pt}=c^{\dagger}_{i\,\b}c^{\dagger}_{i+1\,\r}c^{\dagger}_{i+2\,\g}
508:    +c^{\dagger}_{i\,\r}c^{\dagger}_{i+1\,\g}c^{\dagger}_{i+2\,\b}
509:    +c^{\dagger}_{i\,\g}c^{\dagger}_{i+1\,\b}c^{\dagger}_{i+2\,\r}\\
510:   &\hspace{18pt}-c^{\dagger}_{i\,\b}c^{\dagger}_{i+1\,\g}c^{\dagger}_{i+2\,\r}
511:    -c^{\dagger}_{i\,\g}c^{\dagger}_{i+1\,\r}c^{\dagger}_{i+2\,\b}
512:    -c^{\dagger}_{i\,\r}c^{\dagger}_{i+1\,\b}c^{\dagger}_{i+2\,\g}.\hspace{5pt}
513: \end{split}
514: \end{equation}
515: 
516: The SU(3) generators at each lattice site $i$ are in analogy to
517: \eqref{eq:s} defined as
518: \begin{equation}
519: J^a_i=\frac{1}{2}\,\sum_{\sigma,\sigma'=\b,\r,\g}\,
520: c_{i\sigma}^{\dagger} \lambda^a_{\sigma\sigma'}
521: c_{i\sigma'}^{\phantom{\dagger}}, \quad a=1,\ldots,8,
522: \label{eq:J_a}
523: \end{equation}
524: where the $\lambda^a$ are the Gell-Mann matrices (see
525: App.~\ref{app:conventions}).  The operators \eqref{eq:J_a} satisfy the
526: commutation relations
527: \begin{equation}
528: \comm{J^a_{i}}{J^b_{j}}=\delta_{ij}\;f^{abc}J^c_{i},
529: \quad a,b,c=1,\ldots,8,
530: \label{eq:su3-spincommforsites}
531: \end{equation}
532: (we use the Einstein summation convention) with $f^{abc}$ the
533: structure constants of SU(3) (see App.~\ref{app:conventions}).  We
534: further introduce the total SU(3) spin of $\nu$ neighboring sites
535: $i,\ldots,i+\nu-1$,
536: \begin{equation}\label{eq:defJnu}
537: \bs{J}^{(\nu)}_i = \sum_{j=i}^{i+\nu-1}\,\bs{J}_{j},
538: \end{equation}
539: where $\bs{J}_i$ is the eight-dimensional vector formed by its
540: components \eqref{eq:J_a}.  The parent Hamiltonian for the trimer
541: states \eqref{eq:trimer} is given by
542: \begin{equation}
543:   \label{ham.trimer}
544:   H_{\textrm{trimer}}\, =\, \sum_{i=1}^N
545:   \left(\,\left(\bs{J}^{(4)}_i\right)^{\!4}\,
546:     -\,\frac{14}{3}\left(\bs{J}^{(4)}_i\right)^{\!2}\,+
547:     \,\frac{40}{9}\,\right).
548: \end{equation}
549: The $\bs{J}_i\bs{J}_j$ terms appear complicated in terms
550: of Gell-Mann matrices, but are rather simply when written out
551: using the operator $P_{ij}$, which permutes the SU($n$) spins
552: (here $n=3$) on sites $i$ and $j$:
553: \begin{equation}
554:   \label{eq:permut}
555:   \bs{J}_i\bs{J}_j = \frac{1}{2}\left( P_{ij} - \frac{1}{n} \right).
556: \end{equation}
557: 
558: To verify the trimer Hamiltonian \eqref{ham.trimer}, as well as for
559: the valence bond solid (VBS) models we propose below, we will need a
560: few higher-dimensional representations of SU(3).  We review these in
561: the following section.
562: 
563: \section{Representations of SU(3)}
564: \label{sec:su3rep}
565: 
566: \subsection{Young tableaux and representations of SU(2) }
567: 
568: \begin{figure}[tb]
569: \setlength{\unitlength}{8pt}
570: \begin{picture}(28,10)(0,0)
571: \linethickness{0.3pt}
572: \multiput(0.5,9)(1,0){2}{\line(0,1){1}}
573: \multiput(0.5,9)(0,1){2}{\line(1,0){1}}
574: \put(0.5,9){\makebox(1,1){1}}
575: \put(1,1){\line(0,1){4}}
576: \multiput(1,2)(0,2){2}{\circle*{0.5}}
577: \put(0,5){\makebox(1,1){$S^z$}}
578: \put(1.5,3.5){\makebox(2,1){$\ket{\up}$}}
579: \put(1.5,1.5){\makebox(2,1){$\ket{\dw}$}}
580: 
581: \put(3.5,9){\makebox(1,1){$\otimes$}}
582: 
583: \multiput(6.5,9)(1,0){2}{\line(0,1){1}}
584: \multiput(6.5,9)(0,1){2}{\line(1,0){1}}
585: \put(6.5,9){\makebox(1,1){2}}
586: \put(7,1){\line(0,1){4}}
587: \multiput(7,2)(0,2){2}{\circle*{0.5}}
588: \put(6,5){\makebox(1,1){$S^z$}}
589: \put(7.5,3.5){\makebox(2,1){$\ket{\up}$}}
590: \put(7.5,1.5){\makebox(2,1){$\ket{\dw}$}}
591: 
592: \put(9,9){\makebox(1,1){=}}
593: 
594: \multiput(11.5,8)(1,0){2}{\line(0,1){2}}
595: \multiput(11.5,8)(0,1){3}{\line(1,0){1}}
596: \put(11.5,9){\makebox(1,1){1}}
597: \put(11.5,8){\makebox(1,1){2}}
598: \put(12,2){\line(0,1){2}}
599: \put(12,3){\circle*{0.5}}
600: \put(11,4){\makebox(1,1){$S^z$}}
601: \put(13.8,2.5){\makebox(5,1){$\tfrac{1}{\sqrt{2}}\bigl(
602: \ket{\up\dw}-\ket{\dw\up}\bigr)$}}
603: 
604: \put(15.75,9){\makebox(1,1){$\oplus$}}
605: 
606: \multiput(20,9)(1,0){3}{\line(0,1){1}}
607: \multiput(20,9)(0,1){2}{\line(1,0){2}}
608: \put(20,9){\makebox(1,1){1}}
609: \put(21,9){\makebox(1,1){2}}
610: \put(21,0){\line(0,1){6}}
611: \multiput(21,1)(0,2){3}{\circle*{0.5}}
612: \put(20,6){\makebox(1,1){$S^z$}}
613: \put(21.6,0.5){\makebox(2,1){$\ket{\dw\dw}$}}
614: \put(22.8,2.5){\makebox(5,1){$\tfrac{1}{\sqrt{2}}\bigl(
615: \ket{\up\dw}+\ket{\dw\up}\bigr)$}}
616: \put(21.6,4.5){\makebox(2,1){$\ket{\up\up}$}}
617: \end{picture}
618: \caption{Tensor product of two $S=\frac{1}{2}$ spins with Young
619:   tableaux and weight diagrams of the occurring SU(2) representations.
620:   $S^z$ is the diagonal generator.}
621: \label{fig:su2-1}
622: \end{figure}
623: \begin{figure}[tb]
624: \setlength{\unitlength}{8pt}
625: \begin{picture}(32,8)(2,-1)
626: \linethickness{0.3pt}
627: \put(3,6){\line(1,0){1}}
628: \put(3,5){\line(1,0){1}}
629: \put(3,5){\line(0,1){1}}
630: \put(4,5){\line(0,1){1}}
631: \put(3,5){\makebox(1,1){1}}
632: \put(4.6,5.3){$\otimes$}
633: \put(6,6){\line(1,0){1}}
634: \put(6,5){\line(1,0){1}}
635: \put(6,5){\line(0,1){1}}
636: \put(7,5){\line(0,1){1}}
637: \put(6,5){\makebox(1,1){2}}
638: \put(3,4){$\underbrace{\phantom{iiiiiiiiii}}$}
639: \put(3,2.5){\line(1,0){1}}
640: \put(3,1.5){\line(1,0){1}}
641: \put(3,0.5){\line(1,0){1}}
642: \put(3,0.5){\line(0,1){2}}
643: \put(4,0.5){\line(0,1){2}}
644: \put(3,1.5){\makebox(1,1){1}}
645: \put(3,0.5){\makebox(1,1){2}}
646: \put(2.1,-1){$S=0$}
647: \put(5,1.8){$\oplus$}
648: \put(7,2.5){\line(1,0){2}}
649: \put(7,1.5){\line(1,0){2}}
650: \put(7,1.5){\line(0,1){1}}
651: \put(8,1.5){\line(0,1){1}}
652: \put(9,1.5){\line(0,1){1}}
653: \put(7,1.5){\makebox(1,1){1}}
654: \put(8,1.5){\makebox(1,1){2}}
655: \put(6.8,-0.7){$S=1$}
656: \put(7.6,5.3){$\otimes$}
657: \put(9,6){\line(1,0){1}}
658: \put(9,5){\line(1,0){1}}
659: \put(9,5){\line(0,1){1}}
660: \put(10,5){\line(0,1){1}}
661: \put(9,5){\makebox(1,1){3}}
662: \put(11.6,5.3){$=$}
663: \put(14,6){\line(1,0){1}}
664: \put(14,5){\line(1,0){1}}
665: \put(14,4){\line(1,0){1}}
666: \put(14,3){\line(1,0){1}}
667: \put(14,3){\line(0,1){3}}
668: \put(15,3){\line(0,1){3}}
669: \put(14,5){\makebox(1,1){1}}
670: \put(14,4){\makebox(1,1){2}}
671: \put(14,3){\makebox(1,1){3}}
672: \put(16.1,5.3){$\oplus$}
673: \put(18,6){\line(1,0){2}}
674: \put(18,5){\line(1,0){2}}
675: \put(18,4){\line(1,0){1}}
676: \put(18,4){\line(0,1){2}}
677: \put(19,4){\line(0,1){2}}
678: \put(20,5){\line(0,1){1}}
679: \put(18,5){\makebox(1,1){1}}
680: \put(19,5){\makebox(1,1){3}}
681: \put(18,4){\makebox(1,1){2}}
682: \put(17.5,2.2){$S=\frac{1}{2}$}
683: \put(21.1,5.3){$\oplus$}
684: \put(23,6){\line(1,0){2}}
685: \put(23,5){\line(1,0){2}}
686: \put(23,4){\line(1,0){1}}
687: \put(23,4){\line(0,1){2}}
688: \put(24,4){\line(0,1){2}}
689: \put(25,5){\line(0,1){1}}
690: \put(23,5){\makebox(1,1){1}}
691: \put(23,4){\makebox(1,1){3}}
692: \put(24,5){\makebox(1,1){2}}
693: \put(22.5,2.2){$S=\frac{1}{2}$}
694: \put(26.1,5.3){$\oplus$}
695: \put(28,6){\line(1,0){3}}
696: \put(28,5){\line(1,0){3}}
697: \put(28,5){\line(0,1){1}}
698: \put(29,5){\line(0,1){1}}
699: \put(30,5){\line(0,1){1}}
700: \put(31,5){\line(0,1){1}}
701: \put(28,5){\makebox(1,1){1}}
702: \put(29,5){\makebox(1,1){2}}
703: \put(30,5){\makebox(1,1){3}}
704: \put(28,2.2){$S=\frac{3}{2}$}
705: \put(14.5,4.5){\makebox(0,0){\line(1,2){1.8}}}
706: \put(14.5,4.5){\makebox(0,0){\line(1,-2){1.8}}}
707: \end{picture}
708: \caption{Tensor product of three $S=\frac{1}{2}$ spins with Young
709:   tableaux.  For SU($n$) with $n>2$, the tableau with three boxes on
710:   top of each other exists as well.}
711: \label{fig:su2-2}
712: \end{figure}
713: Let us begin with a review of Young tableaux and the representations
714: of SU(2).  The group SU(2) has three generators $S^a$, $a=1,2,3$,
715: which obey the algebra
716: \begin{equation}
717: \comm{S^a}{S^b}=i\epsilon^{abc}S^c,
718: \label{eq:su2algebra}
719: \end{equation}
720: where repeated indices are summed over and $\epsilon^{abc}$ is the
721: totally antisymmetric tensor.  The representations of SU(2) are
722: classified by the spin $S$, which takes integer or half-integer
723: values.  The fundamental representation of SU(2) has spin
724: $S=\frac{1}{2}$, it contains the two states $\ket{\up}$ and
725: $\ket{\dw}$.  Higher-dimensional representations can be constructed as
726: tensor products of fundamental representations, which is conveniently
727: accomplished using Young tableaux (see
728: \emph{e.g.}~\cite{InuiTanabeOnodera96}).  These tableaux are
729: constructed as follows (see Figs.~\ref{fig:su2-1} and~\ref{fig:su2-2}
730: for examples).  For each of the $N$ spins, draw a box numbered
731: consecutively from left to right.  The representations of SU(2) are
732: obtained by putting the boxes together such that the numbers assigned
733: to them increase in each row from left to right and in each column
734: from top to bottom.  Each tableau indicates symmetrization over all
735: boxes in the same row, and antisymmetrization over all boxes in the
736: same column.  This implies that we cannot have more than two boxes on
737: top of each other.  If $\kappa_i$ denotes the number of boxes in the
738: $i$th row, the spin is given by $S=\frac{1}{2}(\kappa_1-\kappa_2)$.
739: 
740: To be more explicit, let us consider the tensor product
741: $\bf\frac{1}{2}\otimes\frac{1}{2}\otimes\frac{1}{2}$ depicted in
742: Fig.~\ref{fig:su2-2} in detail.  We start with the state
743: $\ket{\tfrac{3}{2},\tfrac{3}{2}}=\ket{\up\up\up}$, and hence find
744: \begin{equation}
745: \ket{\tfrac{3}{2},\tfrac{1}{2}}
746: =\frac{1}{\sqrt{3}}\,S^-\ket{\tfrac{3}{2},\tfrac{3}{2}}=
747: \frac{1}{\sqrt{3}}\bigl(
748: \ket{\up\up\dw}+\ket{\up\dw\up}+\ket{\dw\up\up}\bigr).
749: \label{eq:s=3/2state}
750: \end{equation}
751: The two states with $S=S^z=\frac{1}{2}$ must be orthogonal to
752: \eqref{eq:s=3/2state}.  A convenient choice of basis is
753: \begin{equation}
754: \label{chiralbasis}
755: \begin{split}
756: \ket{\tfrac{1}{2},\tfrac{1}{2},+}&=
757: \frac{1}{\sqrt{3}}\bigl(
758: \ket{\up\up\dw}+\omega\ket{\up\dw\up}+\omega^2\ket{\dw\up\up}\bigr),\\[2mm]
759: \ket{\tfrac{1}{2},\tfrac{1}{2},-}&=
760: \frac{1}{\sqrt{3}}\bigl(
761: \ket{\up\up\dw}+\omega^2\ket{\up\dw\up}+\omega\ket{\dw\up\up}\bigr),
762: \end{split}
763: \end{equation}
764: where $\omega=\exp\bigl(i\frac{2\pi}{3}\bigr)$. The tableaux tell us
765: primarily that two such basis states exist, not what a convenient choice
766: of orthonormal basis states may be.
767: 
768: The irreducible representations of SU(2) can be classified through the
769: eigenvalues of the Casimir operator given by the square of the total
770: spin $\bs{S}^2$.  The special feature of $\bs{S}^2$ is that it
771: commutes with all generators $S^a$ and is hence by Schur's
772: lemma~\cite{Cornwell84vol2} proportional to the identity for any
773: finite-dimensional irreducible representation.
774: The eigenvalues are given by
775: \begin{displaymath}
776:   \bs{S}^2=\mathcal{C}^2_{\text{SU}(2)}=S(S+1).
777: \end{displaymath}
778: 
779: \begin{figure}[t]
780: \includegraphics[scale=0.18]{vbs3.eps}\hspace{5mm}
781: \includegraphics[scale=0.18]{vbs3bar.eps}
782: \caption{(Color online) a) Weight diagram of the fundamental SU(3)
783:   representation $\bs{3}=(1,0)$. b) Weight diagram of the complex conjugate
784:   representation $\bs{\bar{3}}=(0,1)$.
785:   %(reproduced with permission from Ref.~\cite{SchurichtGreiter05colepl}). 
786:   $J^3$ and $J^8$ denote the diagonal generators, $I^+$, $U^+$, and
787:   $V^+$ the raising operators. }
788: \label{fig:33barweight}
789: \end{figure}
790: \begin{figure}[tb]
791: \setlength{\unitlength}{8pt}
792: \begin{picture}(32,8)(2,-1)
793: \linethickness{0.5pt}
794: \put(3,6){\line(1,0){1}}
795: \put(3,5){\line(1,0){1}}
796: \put(3,5){\line(0,1){1}}
797: \put(4,5){\line(0,1){1}}
798: \put(3,5){\makebox(1,1){1}}
799: \put(4.6,5.3){$\otimes$}
800: \put(6,6){\line(1,0){1}}
801: \put(6,5){\line(1,0){1}}
802: \put(6,5){\line(0,1){1}}
803: \put(7,5){\line(0,1){1}}
804: \put(6,5){\makebox(1,1){2}}
805: \put(3,4.4){$\underbrace{\phantom{iiiiiiiiii}}$}
806: \put(3,2.9){\line(1,0){1}}
807: \put(3,1.9){\line(1,0){1}}
808: \put(3,0.9){\line(1,0){1}}
809: \put(3,0.9){\line(0,1){2}}
810: \put(4,0.9){\line(0,1){2}}
811: \put(3,1.9){\makebox(1,1){1}}
812: \put(3,0.9){\makebox(1,1){2}}
813: \put(3,-1){\makebox(1,1){$\bs{\bar{3}}$}}
814: \put(4.6,1.5){$\oplus$}
815: \put(6,2.4){\line(1,0){2}}
816: \put(6,1.4){\line(1,0){2}}
817: \put(6,1.4){\line(0,1){1}}
818: \put(7,1.4){\line(0,1){1}}
819: \put(8,1.4){\line(0,1){1}}
820: \put(6,1.4){\makebox(1,1){1}}
821: \put(7,1.4){\makebox(1,1){2}}
822: \put(6,-1){\makebox(2,1){$\bs{6}$}}
823: \put(7.6,5.3){$\otimes$}
824: \put(9,6){\line(1,0){1}}
825: \put(9,5){\line(1,0){1}}
826: \put(9,5){\line(0,1){1}}
827: \put(10,5){\line(0,1){1}}
828: \put(9,5){\makebox(1,1){3}}
829: \put(11.6,5.3){$=$}
830: \put(14,6){\line(1,0){1}}
831: \put(14,5){\line(1,0){1}}
832: \put(14,4){\line(1,0){1}}
833: \put(14,3){\line(1,0){1}}
834: \put(14,3){\line(0,1){3}}
835: \put(15,3){\line(0,1){3}}
836: \put(14,5){\makebox(1,1){1}}
837: \put(14,4){\makebox(1,1){2}}
838: \put(14,3){\makebox(1,1){3}}
839: \put(14,1){\makebox(1,1){$\bs{1}$}}
840: \put(16.1,5.3){$\oplus$}
841: \put(18,6){\line(1,0){2}}
842: \put(18,5){\line(1,0){2}}
843: \put(18,4){\line(1,0){1}}
844: \put(18,4){\line(0,1){2}}
845: \put(19,4){\line(0,1){2}}
846: \put(20,5){\line(0,1){1}}
847: \put(18,5){\makebox(1,1){1}}
848: \put(19,5){\makebox(1,1){3}}
849: \put(18,4){\makebox(1,1){2}}
850: \put(18,1){\makebox(2,1){$\bs{8}$}}
851: \put(21.1,5.3){$\oplus$}
852: \put(23,6){\line(1,0){2}}
853: \put(23,5){\line(1,0){2}}
854: \put(23,4){\line(1,0){1}}
855: \put(23,4){\line(0,1){2}}
856: \put(24,4){\line(0,1){2}}
857: \put(25,5){\line(0,1){1}}
858: \put(23,5){\makebox(1,1){1}}
859: \put(23,4){\makebox(1,1){3}}
860: \put(24,5){\makebox(1,1){2}}
861: \put(23,1){\makebox(2,1){$\bs{8}$}}
862: \put(26.1,5.3){$\oplus$}
863: \put(28,6){\line(1,0){3}}
864: \put(28,5){\line(1,0){3}}
865: \put(28,5){\line(0,1){1}}
866: \put(29,5){\line(0,1){1}}
867: \put(30,5){\line(0,1){1}}
868: \put(31,5){\line(0,1){1}}
869: \put(28,5){\makebox(1,1){1}}
870: \put(29,5){\makebox(1,1){2}}
871: \put(30,5){\makebox(1,1){3}}
872: \put(28,1){\makebox(3,1){$\bs{10}$}}
873: \end{picture}
874: \caption{Tensor product $\bs{3}\otimes\bs{3}\otimes\bs{3}$ with Young
875:   tableaux.}
876: \label{fig:youngdiagram}
877: \end{figure}
878: 
879: \subsection{Representation theory of SU(3) }
880: 
881: The group SU(3) has eight generators $J^a$, $a=1,\ldots,8$, which
882: obey the algebra
883: \begin{equation}
884: \comm{J^a}{J^b}=f^{abc}J^c,
885: \label{eq:su3algebra}
886: \end{equation}
887: where the structure constants $f^{abc}$ are given in
888: App.~\ref{app:conventions}.  For SU(3) we have two diagonal
889: generators, usually chosen to be $J^3$ and $J^8$, and six generators
890: which define the ladder operators $I^\pm=J^1\pm iJ^2$, $U^\pm=J^6\pm
891: iJ^7$, and $V^\pm=J^4\pm iJ^5$, respectively.  An explicit realization
892: of \eqref{eq:su3algebra} is, for example, given by the $J^a$'s as
893: expressed in terms of Gell-Mann matrices in \eqref{eq:J_a}.  This
894: realization defines the fundamental representation $\bs{3}$ of SU(3)
895: illustrated in Fig.~\ref{fig:33barweight}a.  It is three-dimensional,
896: and we have chosen to label the basis states by the colors blue (b),
897: red (r), and green (g).  The weight diagram depicted in
898: Fig.~\ref{fig:33barweight}a instructs us about the eigenvalues of the
899: diagonal generators as well as the actions of the ladder operators on
900: the basis states.
901: 
902: \begin{figure}[tb]
903: \setlength{\unitlength}{8pt}
904: \begin{picture}(16,4)(0,0)
905: \linethickness{0.3pt}
906: \multiput(0,4)(0,1){2}{\line(1,0){16}}
907: \put(0,3){\line(1,0){11}}
908: \put(0,2){\line(1,0){5}}
909: \multiput(0,2)(1,0){2}{\line(0,1){3}}
910: \multiput(4,2)(1,0){2}{\line(0,1){3}}
911: \put(6,3){\line(0,1){2}}
912: \multiput(10,3)(1,0){2}{\line(0,1){2}}
913: \put(12,4){\line(0,1){1}}
914: \multiput(15,4)(1,0){2}{\line(0,1){1}}
915: \put(11,0.7){\makebox(5,1){$\underbrace{\hspace{38.8pt}}$}}
916: \put(5,0.7){\makebox(6,1){$\underbrace{\hspace{46.7pt}}$}}
917: \put(11,-0.4){\makebox(5,1){$\mu_1\, $boxes}}
918: \put(5,-0.4){\makebox(6,1){$\mu_2\, $columns}}
919: \end{picture}
920: \caption{Dynkin coordinates $(\mu_1,\mu_2)$ for a given Young tableau.
921:   The columns containing three boxes represent additional SU(3)
922:   singlet factors, which yield equivalent representations and hence
923:   leave the Dynkin coordinates $(\mu_1,\mu_2)$ unchanged.}
924: \label{fig:dynkin}
925: \end{figure}
926: 
927: All other representations of SU(3) can be constructed by taking tensor
928: products of reps.\ $\bs{3}$, which is again most conveniently
929: accomplished using Young tableaux (see Fig.~\ref{fig:youngdiagram} for
930: an example).  The antisymmetrization over all boxes in the same column
931: implies that we cannot have more than three boxes on top of each
932: other.  Each tableaux stands for an irreducible representation of
933: SU(3), which can be uniquely labeled by their highest weight or Dynkin
934: coordinates $(\mu_1,\mu_2)$~\cite{Cornwell84vol2,Georgi82}
935: (see Fig.~\ref{fig:dynkin}).  For example, the fundamental
936: representation $\bs{3}$ has Dynkin coordinates (1,0).  Note that all
937: columns containing three boxes are superfluous, as the
938: antisymmetrization of three colors yields only one state.  In
939: particular, the SU(3) singlet has Dynkin coordinates (0,0).  In
940: general, the dimension of a representation $(\mu_1,\mu)$ is given by
941: $\frac{1}{2}(\mu_1+1)(\mu_2+1)(\mu_1+\mu_2+2)$. The labeling using
942: bold numbers refers to the dimensions of the representations alone.
943: Although this labeling is not unique, it will mostly be sufficient for
944: our purposes.  A representation $\bs{m}$ and its conjugated
945: counterpart $\bs{\overline{m}}$ are related to each other by
946: interchange of their Dynkin coordinates.
947: 
948: \subsection{Examples of representations of SU(3)}
949: 
950: We now consider some specific representations of SU(3) in detail.  As
951: starting point we use the fundamental representation $\bs{3}$ spanned
952: by the states $\ket{\b}$, $\ket{\r}$, and $\ket{\g}$.  The second
953: three-dimensional representation $\bs{\bar{3}}$ is obtained by
954: antisymmetrically coupling two reps.\ $\bs{3}$. The Dynkin coordinates
955: of the rep.\ $\bs{\bar{3}}$ are (0,1), \ie the reps.\ $\bs{3}$ and
956: $\bs{\bar{3}}$ are complex conjugate of each other.  An explicit
957: basis of the rep.\ $\bs{\bar{3}}$ is given by the colors yellow (y),
958: cyan (c), and magenta (m),
959: \begin{eqnarray}
960: \ket{\y}\!\!&=&\!\!\frac{1}{\sqrt{2}}\bigl(\ket{\r\g}-\ket{\g\r}\bigr),
961: \nonumber\\
962: \ket{\c}\!\!&=&\!\!\frac{1}{\sqrt{2}}\bigl(\ket{\g\b}-\ket{\b\g}\bigr),\\
963: \ket{\m}\!\!&=&\!\!\frac{1}{\sqrt{2}}\bigl(\ket{\b\r}-\ket{\r\b}\bigr).
964: \nonumber
965: \end{eqnarray}
966: The weight diagram is shown in Fig.~\ref{fig:33barweight}.b.  The
967: generators are given by \eqref{eq:J_a} with $\lambda^a$ replaced by
968: $-(\lambda^a)^*$, where $^*$ denotes complex conjugation of the matrix
969: elements~\cite{Georgi82}. In particular, we find
970: $I^+\ket{\y}=-\ket{\c}$, $U^+\ket{\c}=-\ket{\m}$, and
971: $V^+\ket{\y}=-\ket{\m}$.
972: 
973: \begin{figure}[t]
974: \setlength{\unitlength}{0.4pt}
975: \begin{picture}(320,320)(0,0)
976: \linethickness{0.3pt}
977: \put(0,166){\line(1,0){300}}
978: \put(150,0){\line(0,1){294}}
979: \multiput(50,224)(100,0){3}{\circle*{10}}
980: \multiput(100,137)(100,0){2}{\circle*{10}}
981: \put(150,50){\circle*{10}}
982: \put(320,300){\large\bf 6}
983: \put(155,290){$J^8$}
984: \put(295,150){$J^3$}
985: \put(262,225){$\ket{\b\b}$}
986: \put(5,225){$\ket{\r\r}$}
987: \put(68,245){$\tfrac{1}{\sqrt{2}}\bigl(\ket{\b\r}+\ket{\r\b}\bigr)$}
988: \put(210,115){$\tfrac{1}{\sqrt{2}}\bigl(\ket{\b\g}+\ket{\g\b}\bigr)$}
989: \put(-55,115){$\tfrac{1}{\sqrt{2}}\bigl(\ket{\r\g}+\ket{\g\r}\bigr)$}
990: \put(160,35){$\ket{\g\g}$}
991: \end{picture}
992: \caption{Weight diagram of the representation $\bs{6}=(2,0)$.
993:   The weight diagram of the conjugate representation
994:   $\bs{\bar{6}}=(0,2)$ is obtained by reflection at the
995:   origin~\cite{Cornwell84vol2}.}
996: \label{fig:6weight}
997: \end{figure}
998: 
999: The six-dimensional representation $\bs{6}$ has Dynkin coordinates
1000: (2,0), and can hence be constructed by symmetrically coupling two
1001: reps.\ $\bs{3}$. The basis states of the rep.\ $\bs{6}$ are shown in
1002: Fig.~\ref{fig:6weight}.  The conjugate representation $\bs{\bar{6}}$
1003: can be constructed by symmetrically coupling two reps.\
1004: $\bs{\bar{3}}$.
1005: 
1006: \begin{figure}[t]
1007: \setlength{\unitlength}{0.4pt}
1008: \begin{picture}(460,300)(0,0)
1009: \linethickness{0.3pt}
1010: \put(0,137){\line(1,0){300}}
1011: \put(150,0){\line(0,1){274}}
1012: \multiput(100,224)(100,0){2}{\circle*{10}}
1013: \multiput(50,137)(100,0){3}{\circle*{10}}
1014: \put(150,137){\circle{20}}
1015: \multiput(100,50)(100,0){2}{\circle*{10}}
1016: \put(160,130){\line(3,-2){120}}
1017: \put(320,280){\large\bf 8}
1018: \put(155,270){$J^8$}
1019: \put(295,120){$J^3$}
1020: \put(210,235){$\ket{\b\m}$}
1021: \put(50,235){$\ket{\r\m}$}
1022: \put(263,158){\makebox(3,1){$\ket{\b\c}$}}
1023: \put(37,158){\makebox(3,1){$\ket{\r\y}$}}
1024: \put(210,30){$\ket{\g\c}$}
1025: \put(55,30){$\ket{\g\y}$}
1026: \put(285,63){$\tfrac{1}{\sqrt{2}}\bigl(\ket{\b\y}-\ket{\r\c}\bigr)$}
1027: \put(285,28){$\tfrac{1}{\sqrt{6}}\bigl(\ket{\b\y}+\ket{\r\c}-
1028: 2\ket{\g\m}\bigr)$}
1029: \end{picture}
1030: \caption{Weight diagram of the adjoint representation $\bs{8}=(1,1)$.
1031:   The state with $J^3=J^8=0$ is doubly
1032:   degenerate~\cite{Cornwell84vol2}.  Note that two reps.\ $\bs{8}$ can
1033:   be constructed by combining three fundamental reps.\ $\bs{3}$
1034:   (colors), just as two reps.\ $\bs{\frac{1}{2}}$ can be constructed
1035:   by combining three SU(2) spins (cf. \eqref{chiralbasis}).  The
1036:   states in the diagram span a basis for one of these
1037:   representations.}
1038: \label{fig:8weight}
1039: \end{figure}
1040: 
1041: Let us now consider the tensor product
1042: $\bs{3}\otimes\bs{\bar{3}}=\bs{1}\oplus\bs{8}$.
1043: % included in Fig.~\ref{fig:youngdiagram}.  
1044: The weight diagram of the so-called adjoint representation
1045: $\bs{8}=(1,1)$ is shown in Fig.~\ref{fig:8weight}. The states can be
1046: constructed starting from the highest weight state $\ket{\b\m}$,
1047: yielding $I^-\ket{\b\m}=\ket{\r\m}$, $U^-\ket{\b\m}=-\ket{\b\c}$,
1048: $V^-\ket{\b\m}=\ket{\g\m}-\ket{\b\y}$, and so on. This procedure
1049: yields two linearly independent states with $J^3=J^8=0$.  The
1050: representation $\bs{8}$ can also be obtained by coupling of the reps.\
1051: $\bs{6}$ and $\bs{3}$, as can be seen from the Young tableaux in
1052: Fig.~\ref{fig:youngdiagram}.  On a more abstract level, the adjoint
1053: representation is the representation we obtain if we consider the
1054: generators $J^a$ themselves basis vectors.
1055: In the weight diagram shown in Fig.~\ref{fig:8weight}, the generators
1056: $J^3$ and $J^8$ correspond to the two states at the origin, whereas
1057: the ladder operators $I^\pm$, $U^\pm$, and $V^\pm$ correspond to the
1058: states at the six surrounding points.  In the notation of
1059: Fig.~\ref{fig:8weight}, the singlet orthogonal to $\bs{8}$ is given by
1060: $\tfrac{1}{\sqrt{3}}\bigl(\ket{\b\y}+\ket{\r\c}+\ket{\g\m}\bigr)$.
1061: 
1062: The weight diagrams of four other representations relevant to our
1063: purposes below are shown in Figs.~\ref{fig:10weight}
1064: to~\ref{fig:27weight}.
1065: 
1066: \begin{figure}[t]
1067: \setlength{\unitlength}{0.3pt}
1068: \begin{picture}(420,390)(50,0)
1069: \linethickness{0.3pt}
1070: \put(0,224){\line(1,0){400}}
1071: \put(200,0){\line(0,1){364}}
1072: \multiput(50,311)(100,0){4}{\circle*{13}}
1073: \multiput(100,224)(100,0){3}{\circle*{13}}
1074: \multiput(150,137)(100,0){2}{\circle*{13}}
1075: \put(200,50){\circle*{13}}
1076: \put(420,367){\large\bf 10}
1077: \put(205,357){$J^8$}
1078: \put(395,205){$J^3$}
1079: \put(355,280){$\ket{\b\b\b}$}
1080: \put(-5,280){$\ket{\r\r\r}$}
1081: \put(260,115){$\tfrac{1}{\sqrt{3}}\bigl(\ket{\b\g\g}+\ket{\g\b\g}+
1082: \ket{\g\g\b}\bigr)$}
1083: \put(210,30){$\ket{\g\g\g}$}
1084: \end{picture}
1085: \caption{Weight diagram of the representation $\bs{10}=(3,0)$.
1086:   The weight diagram of the conjugate representation
1087:   $\bs{\overline{10}}=(0,3)$ is obtained by reflection at the
1088:   origin~\cite{Cornwell84vol2}.}
1089:   \label{fig:10weight}
1090: \end{figure}
1091: 
1092: \begin{figure}[t]
1093: \setlength{\unitlength}{0.18pt}
1094: \begin{picture}(550,470)(0,-87)
1095: \linethickness{0.3pt}
1096: \put(0,195){\line(1,0){400}}
1097: \put(200,0){\line(0,1){361}}
1098: \multiput(100,311)(100,0){3}{\circle*{15}}
1099: \multiput(50,224)(100,0){4}{\circle*{15}}
1100: \multiput(150,224)(100,0){2}{\circle{30}}
1101: \multiput(100,137)(100,0){3}{\circle*{15}}
1102: \put(200,137){\circle{30}}
1103: \multiput(150,50)(100,0){2}{\circle*{15}}
1104: \put(420,371){\large\bf 15}
1105: \end{picture}
1106: \begin{picture}(550,550)(-50,0)
1107: \linethickness{0.3pt}
1108: \put(0,282){\line(1,0){500}}
1109: \put(250,0){\line(0,1){448}}
1110: \multiput(50,398)(100,0){5}{\circle*{15}}
1111: \multiput(100,311)(100,0){4}{\circle*{15}}
1112: \multiput(150,224)(100,0){3}{\circle*{15}}
1113: \multiput(200,137)(100,0){2}{\circle*{15}}
1114: \put(250,50){\circle*{15}}
1115: \put(520,454){\large\bf 15'}
1116: \end{picture}
1117: \caption{Weight diagram of the representations $\bs{15}=(2,1)$ and
1118:   $\bs{15'}=(4,0)$.  
1119:   }
1120:   \label{fig:15weight}
1121: \end{figure}
1122: 
1123: \begin{figure}[t]
1124: \setlength{\unitlength}{0.18pt}
1125: \begin{picture}(550,460)(50,0)
1126: \linethickness{0.3pt}
1127: \put(0,224){\line(1,0){500}}
1128: \put(250,0){\line(0,1){448}}
1129: \multiput(150,398)(100,0){3}{\circle*{15}}
1130: \multiput(100,311)(100,0){4}{\circle*{15}}
1131: \multiput(50,224)(100,0){5}{\circle*{15}}
1132: \multiput(100,137)(100,0){4}{\circle*{15}}
1133: \multiput(150,50)(100,0){3}{\circle*{15}}
1134: \multiput(200,311)(100,0){2}{\circle{30}}
1135: \multiput(150,224)(100,0){3}{\circle{30}}
1136: \multiput(200,137)(100,0){2}{\circle{30}}
1137: \put(250,224){\circle{40}}
1138: \put(520,440){\large\bf 27}
1139: \end{picture}
1140: \caption{Weight diagram of the self-conjugate representation
1141:   $\bs{27}=(2,2)$. The state with $J^3=J^8=0$ is three-fold
1142:   degenerate~\cite{Cornwell84vol2}.}
1143:   \label{fig:27weight}
1144: \end{figure}
1145: 
1146: It is known that the physical properties of SU(2) spin chains
1147: crucially depend on whether on the lattice sites are integer or
1148: half-integer spins.  A similar distinction can be made for SU(3)
1149: chains, as elaborated in Sec.~\ref{sec:conf}.  The distinction integer
1150: or half-integer spin for SU(2) is replaced by a distinction between
1151: three families of irreducible representations of SU(3): either the
1152: number of boxes in the Young tableau is divisible by three without
1153: remainder (\eg $\bs{1}$, $\bs{8}$, $\bs{10}$, $\bs{27}$), 
1154: %, $\bs{28}$, $\bs{35}$), 
1155: with remainder one (\eg $\bs{3}$, $\bs{\bar 6}$, $\bs{15}$, $\bs{15'}$),
1156: %, $\bs{\overline{21}}$, $\bs{\overline{24}}$), or
1157: or with remainder two 
1158: (\eg $\bs{\bar 3}$, $\bs{6}$, $\bs{\overline{15}}$, $\bs{\overline{15}'}$). 
1159: %$\bs{21}$, $\bs{24}$).  
1160: 
1161: While SU(2) has only one Casimir operator, SU(3) has two.  The
1162: quadratic Casimir operator is defined as
1163: \begin{equation}
1164: \bs{J}^2=\sum_{a=1}^8 J^aJ^a,
1165: \label{eq:casimir}
1166: \end{equation}
1167: where the $J^a$'s are the generators of the representation. As
1168: $\bs{J}^2$ commutes with all generators $J^a$ it is proportional to
1169: the identity for any finite-dimensional irreducible representation.
1170: The eigenvalue in a representation with Dynkin
1171: coordinates $(\mu_1,\mu_2)$ is~\cite{Cornwell84vol2}
1172: \begin{equation}
1173: \bs{J}^2=\mathcal{C}^2_{\text{SU($3$)}}(\mu_1,\mu_2)=
1174: \frac{1}{3}\bigl(\mu_1^2 + \mu_1\mu_2 + \mu_2^2 + 3\mu_1 + 3\mu_2\bigr).
1175: \label{eq:casimirev}
1176: \end{equation}
1177: We have chosen the normalization in \eqref{eq:casimir} according to
1178: the convention
1179: \begin{displaymath}
1180:   \mathcal{C}^2_{\text{SU($n$)}}(\text{adjoint representation})=n,
1181: \end{displaymath}
1182: which yields $\mathcal{C}^2_{\text{SU($3$)}}(1,1)=3$ for the
1183: representation $\bs{8}$.  Note that the quadratic Casimir operator
1184: cannot be used to distinguish between a representation and its
1185: conjugate.  This distinction would require the cubic Casimir 
1186: operator~\cite{Cornwell84vol2}, which we will not need for any of
1187: the models we propose below.
1188: 
1189: 
1190: \section{The trimer model (continued)}
1191: \label{sec:trimercont}
1192: \addtocounter{subsection}{1}
1193: 
1194: \subsection{Verification of the model}
1195: \label{sec:trimerverified}
1196: 
1197: We will now proceed with the verification of the trimer Hamiltonian
1198: \eqref{ham.trimer}.  Since the spins on the individual sites transform
1199: under the fundamental representation $\bs{3}$, the SU(3) content of
1200: four sites is
1201: \begin{equation}\label{rep3power4}
1202: \mb{3}\otimes\mb{3}\otimes\mb{3}\otimes\mb{3}\,=\,
1203: 3\cdot\mb{3}\,\oplus\,2\cdot\mb{\bar 6}\,\oplus\,3\cdot\mb{15}\,
1204: \oplus\,\mb{15'},
1205: \end{equation}
1206: \ie we obtain representations $\bs{3}$, $\bs{\bar 6}$, and two
1207: non-equivalent 15-dimensional representations with Dynkin coordinates
1208: $(2,1)$ and $(4,0)$, respectively.  All these representations can be
1209: distinguished by their eigenvalues of the quadratic Casimir operator,
1210: which is given by $\left(\bs{J}^{(4)}_i\right)^2$ if the four spins
1211: reside on the four neighboring lattice sites \hbox{$i,\ldots,i+3$.}
1212: 
1213: For the trimer states \eqref{trimerstates.3}, the situation
1214: simplifies as we only have the two possibilities 
1215: \begin{displaymath}
1216: \begin{array}{rcl}
1217: &
1218: \begin{picture}(55,8)(0,-2)\linethickness{0.8pt}
1219: \put(8,0){\line(1,0){10}}\put(6,0){\circle{4}}\put(20,0){\circle{4}}
1220: \put(22,0){\line(1,0){10}}\put(34,0){\circle{4}}\put(48,0){\circle{4}}
1221: %\put(50,0){\line(1,0){4}}
1222: \end{picture}
1223: ~&~\hat =~~~\mb{1}\,\otimes\,\mb{3}~=~\mb{3},\label{eq:one}\\[1mm]
1224: &
1225: \begin{picture}(55,8)(0,-2)\linethickness{0.8pt}
1226: %\put(0,0){\line(1,0){4}}
1227: \put(8,0){\line(1,0){10}}\put(6,0){\circle{4}}
1228: \put(20,0){\circle{4}}\put(36,0){\line(1,0){10}}\put(34,0){\circle{4}}
1229: \put(48,0){\circle{4}}
1230: %\put(50,0){\line(1,0){4}}
1231: \end{picture}~&~
1232: \hat =~~~\mb{\bar 3} \,\otimes\,\mb{\bar 3}~=~\mb{3}\,\oplus\,\mb{\bar 6},
1233: \end{array}
1234: \end{displaymath}
1235: which implies that the total SU(3) spin on four neighboring sites can
1236: only transform under representations $\mb{3}$ or $\mb{\bar{6}}$. The
1237: eigenvalues of the quadratic Casimir operator for these
1238: representations are $4/3$ and $10/3$, respectively.  The auxiliary
1239: operators
1240: \begin{equation}
1241:   \label{eq:auxop}
1242:   H_i=\left(\Bigl(\bs{J}^{(4)}_i\Bigr)^2-\frac{4}{3}\right)
1243:   \left(\Bigl(\bs{J}^{(4)}_i\Bigr)^2-\frac{10}{3}\right)
1244: \end{equation} 
1245: hence annihilate the trimer states for all values of $i$, while they
1246: yield positive eigenvalues for $\bs{15}$ or $\bs{15'}$, \ie all other
1247: states.  Summing $H_i$ over all lattice sites $i$ yields
1248: \eqref{ham.trimer}.  We have numerically confirmed by exact
1249: diagonalization of \eqref{ham.trimer} for chains with $N=9$ and 12
1250: lattice sites that the three states \eqref{eq:trimer} are the only
1251: ground states.
1252: 
1253: Note that the representation content of five neighboring sites in 
1254: the trimer chains is just the conjugate of the above, as
1255: \begin{displaymath}
1256: \begin{array}{rcl}
1257: &
1258: \begin{picture}(65,8)(0,-2)\linethickness{0.8pt}
1259: \put(8,0){\line(1,0){10}}\put(6,0){\circle{4}}\put(20,0){\circle{4}}
1260: \put(22,0){\line(1,0){10}}\put(34,0){\circle{4}}\put(48,0){\circle{4}}
1261: \put(50,0){\line(1,0){10}}\put(62,0){\circle{4}}
1262: %\put(64,0){\line(1,0){4}}
1263: \end{picture}
1264: ~&~\hat =~~\mb{1}\,\otimes\,\mb{\bar 3}~=~\mb{\bar 3}\,,\\[1mm]
1265: &
1266: \begin{picture}(65,8)(0,-2)\linethickness{0,8pt}
1267: %\put(0,0){\line(1,0){4}}
1268: \put(6,0){\circle{4}}\put(20,0){\circle{4}}
1269: \put(36,0){\line(1,0){10}}\put(34,0){\circle{4}}\put(48,0){\circle{4}}
1270: \put(22,0){\line(1,0){10}}\put(62,0){\circle{4}}
1271: %\put(64,0){\line(1,0){4}}
1272: \end{picture}
1273: ~&~\hat =~~\mb{3}\,\otimes\,\mb{1}\,\otimes\,\mb{3}~=~\mb{\bar 3}\,\oplus\,
1274: \mb{6}.\hspace{12pt}
1275: \end{array}
1276: \end{displaymath}
1277: Since the quadratic Casimirs of conjugate representations have
1278: identical eigenvalues, $\mathcal{C}^2_{\text{SU($3$)}}(\mu_1,\mu_2)=
1279: \mathcal{C}^2_{\text{SU($3$)}}(\mu_2,\mu_1)$, we can construct another
1280: parent Hamiltonian for the trimer states \eqref{eq:trimer} by simply
1281: replacing $\bs{J}^{(4)}_i$ with $\bs{J}^{(5)}_i$ in
1282: \eqref{ham.trimer}.  This Hamiltonian will have a different spectrum.
1283: In comparison to the four-site interaction Hamiltonian
1284: \eqref{ham.trimer}, however, it is more complicated while bearing no
1285: advantages.  We will not consider it further.  
1286: 
1287: \subsection{Elementary excitations}
1288: \label{sec:trimerexcitations}
1289: 
1290: \begin{figure}[t]
1291: \begin{minipage}[b]{.45\linewidth}
1292: \includegraphics[width=\linewidth]{loc_exc_3.eps}
1293: \\ (a)
1294: \end{minipage}\hfill
1295: \begin{minipage}[b]{.495\linewidth}
1296: \includegraphics[width=\linewidth]{loc_exc_bar3.eps}
1297: \\ (b)
1298: \end{minipage}
1299: \caption{Couplings used in the numerical studies to create (a) the
1300:   localized rep.\ $\bs{3}$ trial state and (b) the localized rep.\
1301:   $\bs{\bar{3}}$ trial state.}
1302: \label{fig:loc}
1303: \end{figure}
1304: 
1305: Let us now turn to the low-lying excitations of \eqref{ham.trimer}.
1306: In analogy with the MG model, it is evident that the SU(3) spinon
1307: or ``coloron'' excitations correspond to domain walls between the
1308: degenerate ground states.  For the trimer model, however, there
1309: are two different kinds of domain walls, as illustrated by:
1310: \begin{eqnarray}
1311: \setlength{\unitlength}{2pt}
1312: &\begin{picture}(90,12)(4,-2)
1313: \linethickness{0.8pt}
1314: \put(6,0){\circle{4}}
1315: \put(8,0){\line(1,0){10}}
1316: \put(20,0){\circle{4}}
1317: \put(22,0){\line(1,0){10}}
1318: \put(34,0){\circle{4}}
1319: \put(48,0){\circle{4}}
1320: \put(47.5,8){\makebox(1,1){$\mb{3}$}}
1321: \put(62,0){\circle{4}}
1322: \put(64,0){\line(1,0){10}}
1323: \put(76,0){\circle{4}}
1324: \put(78,0){\line(1,0){10}}
1325: \put(90,0){\circle{4}}
1326: \end{picture}&
1327: \label{eq:trial3}\\[3mm]
1328: &\begin{picture}(104,12)(4,-2)
1329: \linethickness{0.8pt}
1330: \put(6,0){\circle{4}}
1331: \put(8,0){\line(1,0){10}}
1332: \put(20,0){\circle{4}}
1333: \put(22,0){\line(1,0){10}}
1334: \put(34,0){\circle{4}}
1335: \put(48,0){\circle{4}}
1336: \put(50,0){\line(1,0){10}}
1337: \put(54.5,8){\makebox(1,1){$\mb{\bar 3}$}}
1338: \put(62,0){\circle{4}}
1339: \put(76,0){\circle{4}}
1340: \put(78,0){\line(1,0){10}}
1341: \put(90,0){\circle{4}}
1342: \put(92,0){\line(1,0){10}}
1343: \put(104,0){\circle{4}}
1344: \end{picture}&
1345: \label{eq:trial3bar}
1346: \end{eqnarray}
1347: The first domain wall \eqref{eq:trial3} connects ground state $\mu$ to
1348: the left to ground state $\mu+1$ to the right, where $\mu$ is defined
1349: modulo 3 (see \eqref{eq:trimer}), and consists of an individual SU(3)
1350: spin, which transforms under representation $\bs{3}$.  The second
1351: domain wall \eqref{eq:trial3bar} connects ground state $\mu$ with 
1352: ground state $\mu+2$.  It consists of two antisymmetrically coupled
1353: spins on two neighboring sites, and hence transforms under
1354: representation $\bs{\bar{3}}$.  As we take momentum superpositions of
1355: the localized domain walls illustrated above, we expect one of them,
1356: but not both, to constitute an approximate eigenstate of the trimer
1357: model.  The reason we do not expect both of them to yield a valid
1358: excitation is that they can decay into each other, \ie if the rep.\
1359: $\bs{3}$ excitation is valid the rep.\ $\bs{\bar{3}}$ domain wall
1360: would decay into two rep.\ $\bs{3}$ excitations, and vice versa.  The
1361: question which of the two excitations is the valid one, \ie whether
1362: the elementary excitations transform under $\bs{3}$ or $\bs{\bar{3}}$
1363: under SU(3) rotations, can be resolved through numerical studies.  We
1364: will discuss the results of these studies now.
1365: 
1366: \begin{table}[t]
1367: \begin{center}
1368: \begin{tabular}{c@{\hspace{10pt}}|@{\hspace{10pt}}c@{\hspace{10pt}}c@{\hspace{10pt}}c@{\hspace{10pt}}c}
1369: \hline\hline
1370: mom&$E_\text{tot}$&&\%&~over-~\\
1371: $[2\pi/N]$&exact&trial&~~off~~&lap\\
1372: \hline
1373: 0    &2.9735& 4.5860&54.2&0.9221\\
1374: 1, 12&6.0345&10.2804&70.4&0.5845\\
1375: 2, 11&9.0164&17.2991&91.9&0.0\\
1376: 3, 10&6.6863&13.1536&96.7&0.0\\
1377: 4, 9 &3.0896& 5.0529&63.5&0.8864\\
1378: 5, 8 &4.8744& 7.5033&53.9&0.8625\\
1379: 6, 7 &8.5618&16.6841&94.9&0.1095\\
1380: \hline\hline
1381: \end{tabular}
1382: \end{center}
1383: \caption{Energies of the rep.\ $\bs{3}$ trial states \eqref{eq:trial3}
1384:   in comparison to the 
1385:   exact excitation energies of the trimer model \eqref{ham.trimer} and
1386:   their overlaps for an SU(3) spin chain with $N=13$ sites.}
1387: \label{tab:rep3}
1388: \end{table}
1389: 
1390: \begin{figure}[t]
1391: \begin{center}
1392: \includegraphics[scale=0.35,angle=270]{13prb.eps}
1393: \caption{Dispersion of the rep.\ $\bs{3}$ trial states
1394:   \eqref{eq:trial3} in comparison to the exact excitation energies of
1395:   \eqref{ham.trimer} for a chain with $N=13$.  The lines are a guide
1396:   to the eye.}
1397: \label{fig:spectrum3}
1398: \end{center}
1399: \end{figure}
1400: 
1401: The rep.\ $\bs{3}$ and the rep.\ $\bs{\bar{3}}$ trial states require
1402: chains with \mbox{$N=3\cdot\text{integer}+1$} and
1403: \mbox{$N=3\cdot\text{integer}+2$} sites, respectively; we chose $N=13$
1404: and $N=14$ for our numerical studies.  To create the localized domain
1405: walls \eqref{eq:trial3} and \eqref{eq:trial3bar}, we numerically
1406: diagonalized auxiliary Hamiltonians with appropriate couplings, as
1407: illustrated in Fig.~\ref{fig:loc}.  From these localized excitations,
1408: we constructed momentum eigenstates by superposition, and compared them
1409: to the exact eigenstates of our model Hamiltonian \eqref{ham.trimer}
1410: for chains with the same number of sites.  The results are shown in
1411: Tab.~\ref{tab:rep3} and Fig.~\ref{fig:spectrum3} for the rep.\
1412: $\bs{3}$ trial state, and in Tab.~\ref{tab:rep3bar} and
1413: Fig.~\ref{fig:spectrum3bar} for the rep.\ $\bs{\bar 3}$ trial state.
1414: 
1415: \begin{table}[t]
1416: \begin{center}
1417: \begin{tabular}{c@{\hspace{10pt}}|@{\hspace{10pt}}c@{\hspace{10pt}}c@{\hspace{10pt}}c@{\hspace{10pt}}c}
1418: \hline\hline
1419: mom&$E_\text{tot}$&&\%&~over-~\\
1420: $[2\pi/N]$&exact&trial&~~off~~&lap\\
1421: \hline
1422: 0   &2.1013&2.3077&9.8&0.9953\\
1423: 1, 13&4.3677&4.8683&11.5&0.9864\\
1424: 2, 12&7.7322&8.7072&12.6&0.9716\\
1425: 3, 11&6.8964&7.7858&12.9&0.9696\\
1426: 4, 10&3.2244&3.5415&9.8&0.9934\\
1427: 5, 9 &2.2494&2.4690&9.7&0.9950\\
1428: 6, 8 &5.4903&6.1016&11.1&0.9827\\
1429: 7   &7.4965&8.5714&14.3&0.9562\\
1430: \hline\hline
1431: \end{tabular}
1432: \caption{Energies of the rep.\ $\bs{\bar 3}$ trial states \eqref{eq:trial3bar}
1433:   in comparison to the 
1434:   exact excitation energies of the trimer model \eqref{ham.trimer} and
1435:   their overlaps for an SU(3) spin chain with $N=14$ sites.}
1436: \label{tab:rep3bar}
1437: \end{center}
1438: \end{table}
1439: 
1440: \begin{figure}[t]
1441: \begin{center}
1442: \includegraphics[scale=0.35,angle=270]{14prb.eps}
1443: \caption{Dispersion of the rep.\ $\bs{\bar{3}}$ trial states
1444:   \eqref{eq:trial3bar} in comparison to the exact excitation energies
1445:   of \eqref{ham.trimer} for a chain with $N=14$.  The lines are a
1446:   guide to the eye.}
1447: \label{fig:spectrum3bar}
1448: \end{center}
1449: \end{figure}
1450: 
1451: The numerical results clearly indicate that the rep.\ $\bs{\bar 3}$
1452: trial states \eqref{eq:trial3bar} are valid approximations to the
1453: elementary excitations of the trimer chain, while the rep.\ $\bs{3}$
1454: trial states \eqref{eq:trial3} are not.  We deduce that the elementary
1455: excitations of the trimer chain \eqref{ham.trimer} transform under
1456: $\bs{\bar{3}}$, that is, under the representation conjugated to the
1457: original SU(3) spins localized at the sites of the chain.  Using the
1458: language of colors, one may say that if a basis for the original spins
1459: is spanned by blue, red, and green, a basis for the excitations is
1460: spanned by the complementary colors yellow, cyan, and magenta.  This
1461: result appears to be a general feature of SU(3) spin chains, as it was
1462: recently shown explicitly to hold for the Haldane--Shastry model as
1463: well~\cite{ bouwknegt-96npb345, schuricht-05epl987, schuricht-06prb235105}.
1464: 
1465: Note that the elementary excitations of the trimer chain are
1466: deconfined, meaning that the energy of two localized representation
1467: $\bs{\bar{3}}$ domain walls or colorons %spinons
1468: \eqref{eq:trial3bar} does not depend on the distance between them.
1469: The reason is simply that domain walls connect one ground state with
1470: another, without introducing costly correlations in the region between
1471: the domain walls.  In the case of the MG model and the trimer model
1472: introduced here, however, there is still an energy gap associated with
1473: the creation of each coloron, %spinon,
1474: which is simply the energy cost associated with the domain wall.
1475: 
1476: In most of the remainder of this article, we will introduce a family
1477: of exactly soluble valence bond models for SU(3) chains of various
1478: spin representations of the SU(3) spins at each lattice site.  To
1479: formulate these models, we will first review Schwinger bosons for
1480: both SU(2) and SU(3) and the AKLT model.
1481: 
1482: 
1483: \section{Schwinger Bosons}
1484: \label{sec:schwinger}
1485: 
1486: Schwinger bosons~\cite{schwinger65proc,Auerbach94} constitute a way
1487: to formulate spin-$S$ representations of an SU(2) algebra.  The spin
1488: operators
1489: \begin{equation}
1490: \begin{array}{rcccl}
1491: S^x + iS^y &=& S^+ &=& a^\dagger b, \\\rule{0pt}{12pt} 
1492: S^x - iS^y &=& S^- &=& b^\dagger a, \\\rule{0pt}{12pt}
1493: && S^z  &=& \frac{1}{2}(a^\dagger a - b^\dagger b), 
1494: \label{eq:schw}
1495: \end{array}
1496: \end{equation}
1497: are given in terms of boson creation and annihilation
1498: %where $a^\dagger,b^\dagger$ (and $a,b$) are boson creation (and annihilation)
1499: operators which obey the usual commutation relations
1500: \begin{equation}
1501: \begin{array}{c}
1502: \comm{a}{a^\dagger}=\comm{b}{b^\dagger}=1, \\
1503: \comm{a^{\phantom{\dagger}}\!\!}{b}=\comm{a}{b^\dagger}
1504: =\comm{a^\dagger}{b}=\comm{a^\dagger}{b^\dagger}=0\rule{0pt}{16pt}.
1505: \end{array}
1506: \label{eq:schwb}
1507: \end{equation}
1508: It is readily verified with (\ref{eq:schwb}) that $S^x$, $S^y$, and
1509: $S^z$ satisfy \eqref{eq:su2algebra}.
1510: The spin quantum number $S$ is given by half the number of bosons,  
1511: \begin{equation}
1512: 2S=a^\dagger a + b^\dagger b,
1513: \label{eq:schwt}
1514: \end{equation}
1515: and the usual spin states (simultaneous eigenstates of $\boldsymbol{S}^2$
1516: and $S^z$) are given by
1517: \begin{equation}
1518: \ket{S,m} = \frac{(a^\dagger)^{S+m}}{\sqrt{(S+m)!}} \frac{(b^\dagger)^{S-m}}
1519: {\sqrt{(S-m)!}} \vac.
1520: \label{eq:schs}
1521: \end{equation}
1522: In particular, the spin-\half states are given by
1523: \begin{equation}
1524: \ket{\up}=c_{\up}^\dagger \vac =a^\dagger \vac ,\qquad 
1525: \ket{\dw}=c_{\dw}^\dagger \vac =b^\dagger \vac ,
1526: \label{eq:schwfun}
1527: \end{equation}
1528: \ie $a^\dagger$ and $b^\dagger$ act just like the fermion creation
1529: operators $c^\dagger_\up$ and $c^\dagger_\dw$ in this case.  The
1530: difference shows up only when two (or more) creation operators act on
1531: the same site or orbital.  The fermion operators create an
1532: antisymmetric or singlet configuration (in accordance with the Pauli
1533: principle), 
1534: \begin{equation}
1535: \ket{0,0} = c_{\up}^\dagger c_{\dw}^\dagger \vac ,
1536: \label{eq:schwfer}
1537: \end{equation}
1538: while the Schwinger bosons create a totally symmetric or
1539: triplet (or higher spin if we create more than two bosons) configuration,
1540: \begin{eqnarray}
1541: \ket{1,1} &=& \textstyle{\frac{1}{\sqrt{2}}} (a^\dagger)^{2}\vac 
1542: ,\nonumber\\[1mm]
1543: \ket{1,0} &=& a^\dagger b^\dagger \vac , \\[1mm]
1544: \ket{1,-1} &=& \textstyle{\frac{1}{\sqrt{2}}} (b^\dagger)^{2}\vac 
1545: .\nonumber
1546: \end{eqnarray}
1547: 
1548: The generalization to SU($n$) proceeds without incident.  We content
1549: ourselves here by writing the formalism out explicitly for SU(3).  In
1550: analogy to \eqref{eq:schw}, we write the SU(3) spin operators
1551: \eqref{eq:J_a} 
1552: \begin{equation}
1553: \begin{array}{rcccl}
1554: J^1 + iJ^2 &=& I^+ &=& b^\dagger r, \\\rule{0pt}{12pt} 
1555: J^1 - iJ^2 &=& I^- &=& r^\dagger b, \\\rule{0pt}{12pt} 
1556: &&            J^3 &=& \frac{1}{2}(b^\dagger b - r^\dagger r),\\\rule{0pt}{12pt}
1557: J^4 + iJ^5 &=& V^+ &=& b^\dagger g, \\\rule{0pt}{12pt} 
1558: J^4 - iJ^5 &=& V^- &=& g^\dagger b, \\\rule{0pt}{12pt} 
1559: J^6 + iJ^7 &=& U^+ &=& r^\dagger g, \\\rule{0pt}{12pt} 
1560: J^6 - iJ^7 &=& U^- &=& g^\dagger r, \\\rule{0pt}{12pt} 
1561: &&             J^8 &=& 
1562: \frac{1}{2\sqrt{3}}(b^\dagger b + r^\dagger r - 2g^\dagger g),
1563: \label{eq:schwsu3}
1564: \end{array}
1565: \end{equation}
1566: in terms of the boson annihilation and creation operators 
1567: $b,b^\dagger$ (blue), $r,r^\dagger$ (red), and $g,g^\dagger$ (green)
1568: satisfying
1569: \begin{equation}
1570: %\begin{array}{c}
1571: \comm{b}{b^\dagger}=\comm{r}{r^\dagger}=\comm{g}{g^\dagger}=1 %, \\
1572: %\text{all other commutators}=0.\rule{0pt}{16pt}
1573: %\end{array}
1574: \label{eq:schwbsu3}
1575: \end{equation}
1576: while all other commutators vanish.  Again, it is readily verified
1577: with \eqref{eq:schwbsu3} that the operators $J^a$ satisfy
1578: \eqref{eq:su3algebra}.  The basis states spanning the fundamental
1579: representation $\bs{3}$ may in analogy to \eqref{eq:schwfun} be
1580: written using either fermion or boson creation operators:
1581: \begin{equation}
1582: \begin{array}{c}
1583: \ket{\b}=c_{\b}^\dagger \vac=b^\dagger \vac ,\\\rule{0pt}{12pt} 
1584: \ket{\r}=c_{\r}^\dagger \vac=r^\dagger \vac ,\\\rule{0pt}{12pt}
1585: \ket{\g}=c_{\g}^\dagger \vac=g^\dagger \vac .
1586: \end{array}
1587: \label{eq:schwfunsu3}
1588: \end{equation}
1589: We write this abbreviated
1590: \begin{equation}
1591: \begin{array}{rcccccccl}
1592: \bs{3}&=&(1,0)&=&
1593: \setlength{\unitlength}{8pt}
1594: \begin{picture}(1,1)(0,0.1)
1595: \put(0,1){\line(1,0){1}}
1596: \put(0,0){\line(1,0){1}}
1597: \put(0,0){\line(0,1){1}}
1598: \put(1,0){\line(0,1){1}}
1599: \end{picture}
1600: &\hat =&c_{\alpha}^\dagger\vac &=& \alpha^\dagger \vac. %\nonumber
1601: \end{array}
1602: \end{equation}
1603: The fermion operators can be used to combine spins transforming under
1604: the fundamental representation $\bs{3}$ antisymmetrically, and hence 
1605: to construct the representations 
1606: %labeled by Young tableaux in which the boxes are arranged in a
1607: %vertical column, in particular
1608: \begin{equation}
1609: \begin{array}{rcccccl}
1610: \bs{\bar 3}&=&(0,1)&=&
1611: \setlength{\unitlength}{8pt}
1612: \begin{picture}(1,2)(0,0.5)
1613: \put(0,2){\line(1,0){1}}
1614: \put(0,1){\line(1,0){1}}
1615: \put(0,0){\line(1,0){1}}
1616: \put(0,0){\line(0,1){2}}
1617: \put(1,0){\line(0,1){2}}
1618: \end{picture}
1619: &\hat =&c_{\alpha}^\dagger c_{\beta}^\dagger \vac,\\\rule{0pt}{12pt}
1620: \bs{1}&=&(0,0)&=&
1621: \setlength{\unitlength}{8pt}
1622: \begin{picture}(1,3)(0,1)
1623: \put(0,3){\line(1,0){1}}
1624: \put(0,2){\line(1,0){1}}
1625: \put(0,1){\line(1,0){1}}
1626: \put(0,0){\line(1,0){1}}
1627: \put(0,0){\line(0,1){3}}
1628: \put(1,0){\line(0,1){3}}
1629: \end{picture}
1630: &\hat =&c_{\b}^\dagger c_{\r}^\dagger c_{\g}^\dagger \vac.
1631: \end{array}
1632: \vspace{6pt}
1633: \label{eq:schwketsu3}
1634: \end{equation}
1635: The Schwinger bosons, by contrast, combine fundamental representations
1636: $\bs{3}$ symmetrically, and hence yield representations labeled by
1637: Young tableaux in which the boxes are arranged in a horizontal row,
1638: like
1639: \begin{equation}
1640: \begin{array}{rcccccl}
1641: \bs{6}&=&(2,0)&=&
1642: \setlength{\unitlength}{8pt}
1643: \begin{picture}(2,1)(0,0.1)
1644: \put(0,1){\line(1,0){2}}
1645: \put(0,0){\line(1,0){2}}
1646: \put(0,0){\line(0,1){1}}
1647: \put(1,0){\line(0,1){1}}
1648: \put(2,0){\line(0,1){1}}
1649: \end{picture}
1650: &\hat =&\alpha^\dagger \beta^\dagger \vac,\\\rule{0pt}{14pt}
1651: \bs{10}&=&(3,0)&=&
1652: \setlength{\unitlength}{8pt}
1653: \begin{picture}(3,1)(0,0.1)
1654: \put(0,1){\line(1,0){3}}
1655: \put(0,0){\line(1,0){3}}
1656: \put(0,0){\line(0,1){1}}
1657: \put(1,0){\line(0,1){1}}
1658: \put(2,0){\line(0,1){1}}
1659: \put(3,0){\line(0,1){1}}
1660: \end{picture}
1661: &\hat =&\alpha^\dagger \beta^\dagger \gamma^\dagger \vac,\\\rule{0pt}{14pt}
1662: \bs{15'}&=&(4,0)&=&
1663: \setlength{\unitlength}{8pt}
1664: \begin{picture}(4,1)(0,0.1)
1665: \put(0,1){\line(1,0){4}}
1666: \put(0,0){\line(1,0){4}}
1667: \put(0,0){\line(0,1){1}}
1668: \put(1,0){\line(0,1){1}}
1669: \put(2,0){\line(0,1){1}}
1670: \put(3,0){\line(0,1){1}}
1671: \put(4,0){\line(0,1){1}}
1672: \end{picture}
1673: &\hat =&\alpha^\dagger \beta^\dagger \gamma^\dagger \delta^\dagger \vac, 
1674: %\ \text{\sl etc.},
1675: \end{array}
1676: \label{eq:schwbossu3}
1677: \end{equation}
1678: where $\alpha, \beta, \gamma, \ldots \in \{b,r,g\}$.
1679: Unfortunately, it is not possible to construct representations
1680: like 
1681: \begin{displaymath}
1682: \bs{8}\ =\ (1,1)\ =\
1683: \setlength{\unitlength}{8pt}
1684: \begin{picture}(2,1.2)(0,0.7)
1685: \put(0,2){\line(1,0){2}}
1686: \put(0,1){\line(1,0){2}}
1687: \put(0,0){\line(1,0){1}}
1688: \put(0,0){\line(0,1){2}}
1689: \put(1,0){\line(0,1){2}}
1690: \put(2,1){\line(0,1){1}}
1691: \end{picture}
1692: \vspace{2pt}
1693: \end{displaymath}
1694: by simply taking products of anti-commuting or commuting creation 
1695: or annihilation operators.
1696:  
1697: 
1698: \section{The AKLT model}
1699: \label{sec:aklt}
1700: 
1701: Using the SU(2) Schwinger bosons introduced in the previous section, we
1702: may rewrite the Majumdar--Ghosh states \eqref{eq:psimg} as
1703: \begin{equation}
1704: \big|\psi_{%\scriptscriptstyle
1705: \text{MG}}^{\textrm{even}\rule{0pt}{5pt}\atop 
1706: \textrm{(odd)}} \big\rangle \, =
1707: \underbrace{\displaystyle
1708: \prod_{i\ \textrm{even}\atop (i\ \textrm{odd})}  
1709: \Bigl(a^\dagger_{i} b^\dagger_{i+1} - b^\dagger_{i} a^\dagger_{i+1}\Bigr)
1710: }_{\displaystyle
1711: \equiv\Psi_{%\scriptscriptstyle
1712: \text{MG}}^{\textrm{even}\rule{0pt}{5pt}\atop 
1713: \textrm{(odd)}} \big[a^\dagger,b^\dagger\big]
1714: } \,\vac
1715: \label{eq:mgschw}
1716: \end{equation}
1717: This formulation was used by Affleck, Kennedy, Lieb, and
1718: Tasaki~\cite{affleck-88cmp477, affleck-87prl799} to propose a family
1719: of states for higher spin representations of SU(2).  In particular, they 
1720: showed that the valance bond solid (VBS) state
1721: \begin{eqnarray}
1722: \ket{\psi_{%\scriptscriptstyle
1723: \text{AKLT}}} \, &=&
1724: \prod_i\,\Bigl(a^\dagger_{i} b^\dagger_{i+1} - b^\dagger_{i} a^\dagger_{i+1}\Bigr)
1725: \,\vac =\nonumber\\
1726: &=&
1727: \Psi_{%\scriptscriptstyle
1728: \textrm{MG}}^{\scriptscriptstyle\textrm{even}}
1729: \big[a^\dagger,b^\dagger\big] \,\cdot\, 
1730: \Psi_{%\scriptscriptstyle
1731: \textrm{MG}}^{\scriptscriptstyle\textrm{odd}}
1732: \big[a^\dagger,b^\dagger\big] \,\vac =\nonumber\\ 
1733: \label{eq:psiaklt}
1734: &=&
1735: \setlength{\unitlength}{1pt}
1736: \Big|
1737: \begin{picture}(128,20)(-7,4)
1738: \linethickness{0.8pt}
1739: \multiput(0,10)(14,0){9}{\circle{4}}
1740: \multiput(0,2)(14,0){9}{\circle{4}}
1741: \thicklines
1742: \multiput(2,10)(28,0){4}{\line(1,0){10}}
1743: \multiput(16,2)(28,0){4}{\line(1,0){10}}
1744: %\put(-6,2){\line(1,0){4}}
1745: %\put(114,10){\line(1,0){4}}
1746: \thinlines
1747: \put(37,-3){\line(1,0){10}}
1748: \put(37,-3){\line(0,1){18}}
1749: \put(37,15){\line(1,0){10}}
1750: \put(47,-3){\line(0,1){18}}
1751: \put(42,-3){\line(0,-1){9}}
1752: \put(42,-20){\makebox(0,0)[c]{\small projection onto spin $S=1$}}
1753: \end{picture}\Big\rangle\\[5mm]
1754: & &\nonumber
1755: \end{eqnarray}
1756: is the exact zero-energy ground state of the spin-1 extended 
1757: Heisenberg Hamiltonian
1758: \begin{equation}
1759: H_{\text{AKLT}} = 
1760: \sum_i \left(\boldsymbol{S}_i \boldsymbol{S}_{i+1} +
1761: \frac{1}{3}\bigl(\boldsymbol{S}_i \boldsymbol{S}_{i+1}\bigr)^2 
1762: +\frac{2}{3}\right) 
1763: \label{eq:haklt}
1764: \end{equation}
1765: with periodic boundary conditions.  Each term in the sum
1766: (\ref{eq:haklt}) projects onto the subspace in which the total spin of
1767: a pair of neighboring sites is $S=2$.  The Hamiltonian
1768: (\ref{eq:haklt}) thereby lifts all states except (\ref{eq:psiaklt}) to
1769: positive energies.  The VBS %valance bond solid
1770: state (\ref{eq:psiaklt}) is a generic paradigm as it shares all the
1771: symmetries, but in particular the Haldane spin
1772: gap~\cite{haldane83pla464, haldane83prl1153, affleck89jpcm3047}, of
1773: the spin-1 Heisenberg chain.  It even offers a particularly simple
1774: understanding of this gap, or of the linear confinement potential
1775: between spinons responsible for it, as illustrated by the cartoon:
1776: %(see Sec.~\ref{sec:10VBS}):
1777: \begin{center}
1778: \setlength{\unitlength}{1pt}
1779: \begin{picture}(196,45)(0,0)
1780: \linethickness{0.8pt}
1781: \multiput(0,35)(14,0){14}{\circle{4}}
1782: \multiput(14,25)(14,0){14}{\circle{4}}
1783: \thicklines
1784: \multiput(2,35)(28,0){2}{\line(1,0){10}}
1785: \multiput(72,35)(28,0){2}{\line(1,0){10}}
1786: \multiput(142,35)(28,0){2}{\line(1,0){10}}
1787: \multiput(16,25)(28,0){6}{\line(1,0){10}}
1788: \put(184,25){\line(1,0){10}}
1789: %\put(-6,25){\line(1,0){4}}
1790: \thinlines
1791: \put(56,36){\makebox(0,0){\vector(0,1){14}}}
1792: \put(126,36){\makebox(0,0){\vector(0,1){14}}}
1793: \put(91,11){\vector(-1,0){35}}
1794: \put(91,11){\vector(1,0){35}}
1795: \put(91,0){\makebox(0,0){\small energy cost $\propto$ distance}}
1796: \end{picture}\\[5mm]
1797: \end{center}
1798: Our understanding~\cite{greiter02prb134443,greiter02prb054505} of the
1799: connection between the confinement force and the Haldane gap is that
1800: the confinement effectively imposes an oscillator potential for the
1801: relative motion of the spinons.  We then interpret the zero-point
1802: energy of this oscillator as the Haldane gap in the excitation
1803: spectrum.%, as illustrated in Fig.\,\ref{fig:osci}.
1804: 
1805: The AKLT state can also be written as a matrix
1806: product~\cite{kluemper-91jpal955,kluemper-92zpb281,kluemper-93epl293}.
1807: We first rewrite the valence bonds
1808: \begin{displaymath}
1809: \Bigl(a^\dagger_{i} b^\dagger_{i+1} - b^\dagger_{i} a^\dagger_{i+1}\Bigr)=
1810: \Bigl(a^\dagger_{i}, b^\dagger_{i}\Bigr)
1811: \left(\!\begin{array}{c} b^\dagger_{i+1}\\-a^\dagger_{i+1}\end{array}\!\right),
1812: \end{displaymath}
1813: and then use the outer product to combine the two vectors at each site
1814: into a matrix
1815: %\vspace{2pt}
1816: \begin{equation}
1817:   \label{eq:akltmatrix}
1818:   \begin{array}{rcl}
1819:     M_i &\equiv& 
1820:   \left(\!\begin{array}{c} b^\dagger_{i}\\-a^\dagger_{i}\end{array}\!\right)
1821:   \Bigl(a^\dagger_{i}, b^\dagger_{i}\Bigr) \;\vac_i 
1822: %  \\ &=&\left(\!\begin{array}{cc} 
1823: %      a^\dagger_{i} b^\dagger_{i} & {b^\dagger_{i}}^2\\
1824: %      -{a^\dagger_{i}}^2 & a^\dagger_{i} b^\dagger_{i}
1825: %    \end{array}\!\right)\vac_i 
1826:   \\[18pt] &=&\left(\!\begin{array}{cc} 
1827:        \ket{1,0}_i& \sqrt{2}\ket{1,-1}_i\\[4pt]
1828:       -\sqrt{2}\ket{1,1}_i &-\ket{1,0}_i 
1829:     \end{array}\!\right).\\
1830:   \end{array}
1831: \end{equation}
1832: %\vspace{2pt}
1833: %where the index $i$ on the states denotes the site.  
1834: Assuming PBCs, \eqref{eq:psiaklt} may then be written as the trace of
1835: the matrix product
1836: \begin{equation}
1837:   \label{eq:akltm}
1838:   \ket{\psi_{\text{AKLT}}}=\text{tr}\biggl( \prod_i M_i \biggr).
1839: \end{equation}
1840: 
1841: In the following section, we will propose several exact models of
1842: VBSs for SU(3).
1843: 
1844: 
1845: \section{SU(3) valence bond solids}
1846: \label{sec:vbs}
1847: 
1848: To begin with, we use SU(3) Schwinger bosons
1849: introduced in Sec.~\ref{sec:schwinger} to rewrite the trimer states
1850: \eqref{eq:trimer} as
1851: \begin{eqnarray}
1852:   \ket{\psi_\text{trimer}^{(\mu)}\!}
1853:   &=&\hspace{-5pt}\prod_{\scriptstyle{i} \atop 
1854:     \left(\scriptstyle{\frac{i-\mu}{3}\,{\rm integer}}\right)}
1855:   \hspace{-5pt}\Bigl(
1856:   \sum_{\scriptstyle{(\alpha,\beta,\gamma)}=\atop\scriptstyle{{\pi}(b,r,g)}} 
1857:   \hspace{-5pt}\hbox{sign}({\pi})\,
1858:   \alpha^{\dagger}_{i}\,\beta^{\dagger}_{i+1}\gamma^{\dagger}_{i+2}
1859:   \Bigr)\vac \nonumber \\[2pt]
1860:   &\equiv&\,\Psi^{\mu}\!\left[
1861:     b^{\dagger},r^{\dagger},g^{\dagger}\right]\,\vac,
1862:   \label{eq:trimerschw}
1863: \end{eqnarray}
1864: where, as in \eqref{eq:trimer}, $\mu=1,2,3$ labels the three
1865: degenerate ground states, $i$ runs over the lattice sites subject to
1866: the constraint that $\frac{i-\mu}{3}$ is integer, and the sum extends
1867: over all six permutations $\pi$ of the three colors b, r, and g.
1868: This formulation can be used directly to construct VBSs for SU(3) 
1869: spin chains with spins transforming under representations $\bs{6}$ 
1870: and $\bs{10}$ on each site.  
1871: 
1872: \subsection{The representation $\bs{6}$ VBS}
1873: 
1874: We obtain a representations $\bs{6}$ VBS from two trimer states by
1875: projecting the tensor product of two fundamental representations
1876: $\bs{3}$ onto the symmetric subspace, \ie onto the $\bs{6}$ in the
1877: decomposition $\bs{3}\,\otimes\,\bs{3}=\bs{\bar{3}}\,\oplus\,\bs{6}$.
1878: Graphically, this is illustrated as follows:
1879: \begin{equation}
1880: \setlength{\unitlength}{1pt}
1881: \begin{picture}(128,48)(4,-22)
1882: \linethickness{0.8pt}
1883: \multiput(0,10)(14,0){9}{\circle{4}}
1884: \multiput(14,0)(14,0){9}{\circle{4}}
1885: \thicklines
1886: \put(2,10){\line(1,0){10}}
1887: \put(16,10){\line(1,0){10}}
1888: \put(44,10){\line(1,0){10}}
1889: \put(58,10){\line(1,0){10}}
1890: \put(86,10){\line(1,0){10}}
1891: \put(100,10){\line(1,0){10}}
1892: \put(16,0){\line(1,0){10}}
1893: \put(30,0){\line(1,0){10}}
1894: \put(58,0){\line(1,0){10}}
1895: \put(72,0){\line(1,0){10}}
1896: \put(100,0){\line(1,0){10}}
1897: \put(114,0){\line(1,0){10}}
1898: \thinlines
1899: % \put(65,-7){\line(1,0){10}}
1900: % \put(65,-7){\line(0,1){24}}
1901: % \put(65,17){\line(1,0){10}}
1902: % \put(75,-7){\line(0,1){24}}
1903: % \put(70,-15){\line(0,1){8}}
1904: \put(65,-5){\framebox(10,20)}
1905: \put(70,-15){\line(0,1){10}}
1906: \put(70,-22){\makebox(1,1){\small projection onto rep.\ $\bs{6}=(2,0)$}}
1907: \put(70,25){\makebox(1,1){\small one site}}
1908: \end{picture}
1909: \label{fig:6state}
1910: \end{equation}
1911: This construction yields three linearly independent $\mb{6}$ VBS
1912: states, as there are three ways to choose two different trimer states
1913: out of a total of three.  These three VBS states are readily written
1914: out using \eqref{eq:trimerschw},
1915: \begin{equation}
1916: \label{eq:6VBS}
1917: \ket{\psi_{\bs{6}\, \text{VBS}}^{(\mu)}}=
1918: \Psi^{\mu}\!\left[b^{\dagger},r^{\dagger},g^{\dagger}\right]\cdot
1919: \Psi^{\mu+1}\!\left[b^{\dagger},r^{\dagger},g^{\dagger}\right]\,
1920: \vac
1921: \end{equation}
1922: for $\mu=1$, 2, or 3.
1923: If we pick four neighboring sites on a chain with any of these states, 
1924: the total SU(3) spin of those may contain the representations
1925: \begin{displaymath}
1926: \setlength{\unitlength}{1pt}
1927: \begin{picture}(56,12)(0,2)
1928: \linethickness{0.8pt}
1929: \multiput(0,10)(14,0){4}{\circle{4}}
1930: \multiput(0,0)(14,0){4}{\circle{4}}
1931: \thicklines
1932: \put(2,10){\line(1,0){10}}
1933: \put(16,10){\line(1,0){10}}
1934: %\put(44,10){\line(1,0){4}}
1935: %\put(-2,0){\line(-1,0){4}}
1936: \put(16,0){\line(1,0){10}}
1937: \put(30,0){\line(1,0){10}}
1938: \end{picture}
1939: \hat =~~~\bs{3}\,\otimes\,\bs{3}~=~\bs{\bar 3}\,\oplus\,\bs{6}
1940: \end{displaymath}
1941: or the representations
1942: \begin{displaymath}
1943:   \setlength{\unitlength}{1pt}
1944: \begin{picture}(56,12)(0,2)
1945: \linethickness{0.8pt}
1946: \multiput(0,10)(14,0){4}{\circle{4}}
1947: \multiput(0,0)(14,0){4}{\circle{4}}
1948: \thicklines
1949: \put(2,10){\line(1,0){10}}
1950: \put(16,10){\line(1,0){10}}
1951: \put(2,0){\line(1,0){10}}
1952: \put(30,0){\line(1,0){10}}
1953: \end{picture}
1954: \hat =~~~\bs{\bar 3}\,\otimes\,\bs{\bar 3}\,\otimes\,\bs{3}~
1955: =~2\cdot\bs{\bar 3}\,\oplus\,\bs{6}\,\oplus\,\bs{\overline{15}},
1956: \end{displaymath}
1957: \ie the total spin transforms under $\bs{\bar{3}}$, $\bs{6}$, or
1958: $\bs{\overline{15}}=(1,2)$, all of which are contained in the product
1959: \begin{multline}
1960:   \label{eq:6666}
1961:   \bs{6}\,\otimes\,\bs{6}\,\otimes\,\bs{6}\,\otimes\,\bs{6}=\\
1962:   3\cdot\bs{\overline{3}}\,\oplus\,6\cdot\bs{6}\,\oplus\,
1963:   7\cdot\bs{\overline{15}}\,\oplus\,3\cdot\bs{\overline{15}'}\,\oplus\,
1964:   3\cdot\bs{21}\,\\ \oplus\,8\cdot\bs{24}\,\oplus\,
1965:   6\cdot\bs{\overline{42}}\,\oplus\,\bs{45}\,\oplus\,
1966:   6\cdot\bs{60}\,\oplus\,3\cdot\bs{63}
1967: \end{multline}
1968: and hence possible for a representation $\bs{6}$ spin chain in
1969: general.  The corresponding Casimirs are given by
1970: $\casi{0}{1}=\frac{4}{3}$, $\casi{2}{0}=\frac{10}{3}$, and
1971: $\casi{1}{2}=\frac{16}{3}$.  This leads us to propose the parent
1972: Hamiltonian
1973: \begin{equation}
1974:   \label{eq:ham6}
1975:   H_{\bs{6}\,\text{VBS}}=\sum_{i=1}^N H_i
1976: \end{equation}
1977: with 
1978: \begin{equation}
1979:   \label{eq:hi6}
1980:   H_i=\left(\!\left(\!\bs{J}^{(4)}_i\!\right)^2 \! - \frac{4}{3}\right)
1981:   \!\left(\!\left(\!\bs{J}^{(4)}_i\!\right)^2 \! - \frac{10}{3}\right)
1982:   \!\left(\!\left(\!\bs{J}^{(4)}_i\!\right)^2 \! - \frac{16}{3}\right).
1983: \end{equation}
1984: Note that the operators $J_i^a$, $a=1,\ldots,8$, are now given by
1985: $6\times 6$ matrices, as the Gell-Mann matrices only provide the
1986: generators \eqref{eq:J_a} of the fundamental representation $\bs{3}$.
1987: Since the representations $\bs{\bar 3}$, $\bs{6}$, and
1988: $\bs{\overline{15}}$ possess the smallest Casimirs in the expansion
1989: \eqref{eq:6666}, $H_i$ and hence also $H_{\bs{6}\,\text{VBS}}$ are
1990: positive semi-definite (\ie have only non-negative eigenvalues).  The
1991: three linearly independent states \eqref{eq:6VBS} are zero-energy
1992: eigenstates of \eqref{eq:ham6}.  
1993: 
1994: To verify that these are the only ground states, we have numerically
1995: diagonalized \eqref{eq:ham6} for $N=6$ and $N=9$ sites.  For $N=9$, we
1996: find zero-energy ground states at momenta $k=0,3,\ \text{and}\ 6$ (in
1997: units of $\frac{2\pi}{N}$ with the lattice constant set to unity).
1998: Since the dimension of the Hilbert space required the use of a LANCZOS
1999: algorithm, we cannot be certain that there are no further ground
2000: states.  We therefore diagonalized \eqref{eq:ham6} for $N=6$ as well,
2001: where we were able to obtain the full spectrum.  We obtained five zero
2002: energy ground states, two at momentum $k=0$ and one each at
2003: $k=2,3,4$.  One of the ground states at $k=0$ and
2004: the $k=2,4$ ground states constitute the space of momentum eigenstates
2005: obtained by Fourier transform of the space spanned by the three
2006: $\bs{6}$ VBS states \eqref{eq:6VBS}.  The remaining two states at
2007: $k=0,3$ are the momentum eigenstates formed by superposition of the
2008: state
2009: %\pagebreak
2010: \begin{equation}
2011: \setlength{\unitlength}{1pt}
2012: \begin{picture}(120,36)(-14,-12)
2013: \linethickness{0.8pt}
2014: \put(0,0){\circle{4}}
2015: \put(0,10){\circle{4}}
2016: \put(98,0){\circle{4}}
2017: \put(98,10){\circle{4}}
2018: \multiput(14,10)(14,0){6}{\circle{4}}
2019: \multiput(14,0)(14,0){6}{\circle{4}}
2020: \thicklines
2021: \put(2,0){\line(1,0){5}}
2022: \put(7,0){\line(1,0){5}}
2023: \put(30,0){\line(1,0){10}}
2024: \put(58,0){\line(1,0){10}}
2025: \put(16,1){\line(4,3){10}}
2026: \put(44,1){\line(4,3){10}}
2027: \put(72,1){\line(4,3){10}}
2028: \put(86,0){\line(1,0){5}}
2029: \put(91,0){\line(1,0){5}}
2030: %\put(100,1){\line(4,3){3}}
2031: \put(-7,5){\line(4,3){5}}
2032: \qbezier(15,12)(28,25)(41,12)
2033: \qbezier(43,12)(56,25)(69,12)
2034: \qbezier(99,12)(101,14)(103,15)
2035: \thinlines
2036: \put(7,-8){\dashbox{2}(84,32)}
2037: \put(-14,-15){\makebox(0,0){\small $i=$}}
2038: \put(0,-15){\makebox(0,0){\small $6$}}
2039: \put(14,-15){\makebox(0,0){\small $1$}}
2040: \put(28,-15){\makebox(0,0){\small $2$}}
2041: \put(42,-15){\makebox(0,0){\small $3$}}
2042: \put(56,-15){\makebox(0,0){\small $4$}}
2043: \put(70,-15){\makebox(0,0){\small $5$}}
2044: \put(84,-15){\makebox(0,0){\small $6$}}
2045: \put(98,-15){\makebox(0,0){\small $1$}}
2046: \end{picture}
2047: \label{fig:finite6state}
2048: \end{equation}
2049: and the same translated by one lattice spacing.  It is readily seen
2050: that these two states are likewise zero energy eigenstates of
2051: \eqref{eq:ham6} for $N=6$ sites.  The crucial difference, however, is
2052: that the $\bs{6}$ VBS states \eqref{eq:6VBS} remain zero-energy
2053: eigenstates of \eqref{eq:ham6} for all $N$'s divisible by three, while
2054: the equivalent of \eqref{fig:finite6state} for larger $N$ do not.  We
2055: hence attribute these two additional ground states for $N=6$ to the
2056: finite size, and conclude that the three states
2057: \eqref{eq:6VBS} are the only zero-energy ground states of
2058: \eqref{eq:ham6} for general $N$'s divisible by three.
2059: 
2060: Excitations of the $\bs{6}$ VBS model are given by domain walls
2061: between two of the ground states \eqref{eq:6VBS}.  As in the trimer
2062: model, two distinct types of domain walls exist, which transform
2063: according to representations $\bs{\bar{3}}$ and $\bs{3}$:
2064: \begin{equation}\label{eq:6VBSexc}
2065: \setlength{\unitlength}{1pt}
2066: \begin{picture}(220,45)(4,-15)
2067: \linethickness{0.8pt}
2068: \multiput(0,10)(14,0){3}{\circle{4}}
2069: \multiput(42,10)(14,0){2}{\circle*{4}}
2070: \multiput(70,10)(14,0){6}{\circle{4}}
2071: \multiput(154,10)(14,0){1}{\circle*{4}}
2072: \multiput(168,10)(14,0){3}{\circle{4}}
2073: \multiput(14,0)(14,0){15}{\circle{4}}
2074: \thicklines
2075: \put(2,10){\line(1,0){10}}
2076: \put(16,10){\line(1,0){10}}
2077: \put(44,10){\line(1,0){10}}
2078: \put(72,10){\line(1,0){10}}
2079: \put(114,10){\line(1,0){10}}
2080: \put(128,10){\line(1,0){10}}
2081: \put(86,10){\line(1,0){10}}
2082: \put(170,10){\line(1,0){10}}
2083: \put(184,10){\line(1,0){10}}
2084: \put(16,0){\line(1,0){10}}
2085: \put(30,0){\line(1,0){10}}
2086: \put(58,0){\line(1,0){10}}
2087: \put(72,0){\line(1,0){10}}
2088: \put(100,0){\line(1,0){10}}
2089: \put(114,0){\line(1,0){10}}
2090: \put(142,0){\line(1,0){10}}
2091: \put(156,0){\line(1,0){10}}
2092: \put(184,0){\line(1,0){10}}
2093: \put(198,0){\line(1,0){10}}
2094: \put(12,-12){\makebox(0,0){\small $\Psi^1\!\cdot\!\Psi^2$}}
2095: \put(101,-12){\makebox(0,0){\small $\Psi^2\!\cdot\!\Psi^3$}}
2096: \put(198,-12){\makebox(0,0){\small $\Psi^1\!\cdot\!\Psi^2$}}
2097: \put(49,25){\makebox(0,0){$\bs{\bar 3}$}}
2098: \put(154,25){\makebox(0,0){$\bs{3}$}}
2099: \thinlines
2100: %
2101: \put(23,-5){\dashbox{2}(52,20)}
2102: \put(121,-6){\dashbox{2}(52,20)}
2103: \put(135,-4){\dashbox{1}(52,20)}
2104: %
2105: \end{picture}
2106: \end{equation}
2107: %\nopagebreak
2108: It is not clear which excitation has the lower energy, and it appears
2109: likely that both of them are stable against decay.  Let us first look
2110: at the rep.\ $\bs{\bar 3}$ excitation.  The four-site Hamiltonian
2111: \eqref{eq:hi6} annihilates the state for all $i$'s except the four
2112: sites in the dashed box in \eqref{eq:6VBSexc}, which contains
2113: the representations
2114: \begin{multline}%{displaymath}
2115:   \bs{\bar 3}\,\otimes\,\bs{\bar 3}\,\otimes\,\bs{\bar 3}\,
2116:   \otimes\,\bs{3}\,\otimes\,\bs{3}\\ =
2117:   6\cdot\bs{\bar 3}\,\oplus\,5\cdot\bs{6}\,\oplus\,6\cdot
2118:   \bs{\overline{15}}\,\oplus\,\bs{\overline{15}'}\,\oplus\,2\cdot\bs{24}\,
2119:   \oplus\,\bs{\overline{42}}
2120:   \nonumber
2121: \end{multline}%{displaymath}
2122: \ie the representations $\bs{\overline{15}'}=(0,4)$, $\bs{24}=(3,1)$
2123: twice, and $\bs{\overline{42}}=(2,3)$ with Casimirs
2124: $\frac{28}{3}$, $\frac{25}{3}$ and $\frac{34}{3}$, respectively, in
2125: addition to representations annihilated by $H_i$.  For the
2126: rep.\ $\bs{3}$ excitation sketched on the right in \eqref{eq:6VBSexc},
2127: there are two sets of four neighboring sites not annihilated by $H_i$
2128: as indicated by the dashed and the dotted box.  Each set contains the
2129: representations
2130: \begin{displaymath}
2131:   \bs{\bar 3}\,\otimes\,\bs{3}\,\otimes\,\bs{3}\,\otimes\,\bs{3}\,=
2132:   \,3\cdot\bs{\bar 3}\,\oplus\,3\cdot\bs{6}\,\oplus\,2\cdot\bs{\overline{15}}\,
2133:   \oplus\,\bs{24}
2134: \end{displaymath}
2135: \ie only the rep.\ $\bs{24}$ in addition to representations
2136: annihilated by $H_i$.  For our parent Hamiltonian \eqref{eq:ham6}, it
2137: hence may well be that the rep.\ $\bs{3}$ anti-coloron %spinon
2138: has the lower energy, but it is all but clear that the rep.\ $\bs{\bar
2139:   3}$ has sufficiently higher energy to decay.  For general
2140: representation $\bs{6}$ spin chains, it may depend on the specifics of
2141: the model which excitation is lower in energy and whether the
2142: conjugate excitation decays or not.
2143: 
2144: Since the excitations of the rep.\ $\bs{6}$ VBS chain are merely
2145: domain walls between different ground states, there is no confinement
2146: between them.  We expect the generic antiferromagnetic rep.\ $\bs{6}$
2147: chain to be gapless, even though the model we proposed here has a gap
2148: associated with the energy cost of creating a domain wall.
2149: 
2150: \subsection{The representation 10 VBS}\label{sec:10VBS}
2151: 
2152: Let us now turn to the $\bs{10}$ VBS chain, which is a direct
2153: generalization of the AKLT chain to SU(3).  By combining the three
2154: different trimer states \eqref{eq:trimerschw} for $\mu=1,2,\
2155: \text{and}\ 3$ symmetrically,
2156: \begin{eqnarray}
2157: \ket{\psi_{\bs{10}\,\text{VBS}}}\!&\!=\!&\!
2158: \Psi^1\!\left[b^{\dagger}\!,r^{\dagger}\!,g^{\dagger}\right]
2159: \!\cdot\!\Psi^2\!\left[b^{\dagger}\!,r^{\dagger}\!,g^{\dagger}\right]
2160: \!\cdot\!\Psi^3\!\left[b^{\dagger}\!,r^{\dagger}\!,g^{\dagger}\right]\vac
2161: \nonumber \\[5pt]
2162: \!&\!=\!&\!
2163: \prod_i\Bigl(
2164: \sum_{\scriptstyle{(\alpha,\beta,\gamma)}=\atop\scriptstyle{{\pi}(b,r,g)}} 
2165: \hspace{-5pt}\hbox{sign}({\pi})\,
2166: \alpha^{\dagger}_{i}\,\beta^{\dagger}_{i+1}\gamma^{\dagger}_{i+2}
2167: \Bigr)\vac , 
2168: \label{state.10VBS}
2169: \end{eqnarray}
2170: we automatically project out the rep.\ $\bs{10}$ in the decomposition
2171: $\bs{3}\otimes\bs{3}\otimes\bs{3}=\bs{1}\oplus2\!\cdot\bs{8}\oplus\bs{10}$
2172: generated on each lattice site by the three trimer chains.  This
2173: construction yields a unique state, as illustrated:
2174: \begin{equation}
2175: \setlength{\unitlength}{1pt}
2176: \begin{picture}(180,56)(4,-6)
2177: \linethickness{0.8pt}
2178: \multiput(0,35)(14,0){12}{\circle{4}}
2179: \multiput(14,25)(14,0){12}{\circle{4}}
2180: \multiput(28,15)(14,0){12}{\circle{4}}
2181: \thicklines
2182: \put(2,35){\line(1,0){10}}
2183: \put(16,35){\line(1,0){10}}
2184: \put(44,35){\line(1,0){10}}
2185: \put(58,35){\line(1,0){10}}
2186: \put(86,35){\line(1,0){10}}
2187: \put(100,35){\line(1,0){10}}
2188: \put(16,25){\line(1,0){10}}
2189: \put(30,25){\line(1,0){10}}
2190: \put(58,25){\line(1,0){10}}
2191: \put(72,25){\line(1,0){10}}
2192: \put(100,25){\line(1,0){10}}
2193: \put(114,25){\line(1,0){10}}
2194: \put(30,15){\line(1,0){10}}
2195: \put(44,15){\line(1,0){10}}
2196: \put(72,15){\line(1,0){10}}
2197: \put(86,15){\line(1,0){10}}
2198: \put(114,15){\line(1,0){10}}
2199: \put(128,15){\line(1,0){10}}
2200: %
2201: \put(128,35){\line(1,0){10}}
2202: \put(142,35){\line(1,0){10}}
2203: \put(142,25){\line(1,0){10}}
2204: \put(156,25){\line(1,0){10}}
2205: \put(156,15){\line(1,0){10}}
2206: \put(170,15){\line(1,0){10}}
2207: %
2208: \thinlines
2209: \put(51,10){\framebox(10,30){}}
2210: \put(56,2){\line(0,1){8}}
2211: \put(56,-5){\makebox(1,1){\small projection onto $\bs{10}=(3,0)$}}
2212: \put(56,48){\makebox(1,1){\small one site}}
2213: \put(121,10){\dashbox{2}(24,30)}
2214: \end{picture}
2215: \label{fig:10state}
2216: \end{equation}
2217: In order to construct a parent Hamiltonian, note first that the total
2218: spin on two (neighboring) sites of a rep.\ $\bs{10}$ chain is given by
2219: \begin{equation}
2220: \bs{10}\otimes\bs{10}
2221: =\bs{\overline{10}}\oplus\bs{27}\oplus\bs{28}\oplus\bs{35}.
2222: \label{twosites.10}
2223: \end{equation}
2224: On the other hand, the total spin of two neighboring sites for
2225: the $\bs{10}$ VBS state can contain only the representations
2226: \begin{equation}
2227:   \bs{\bar{3}}\otimes\bs{\bar{3}}\otimes\bs{3}\otimes\bs{3}=
2228:   2\cdot\bs{1}\oplus 4\cdot\bs{8}
2229:   \oplus\bs{10}\oplus\bs{\overline{10}}\oplus\bs{27},
2230:   \label{twosites.10VBS}
2231: \end{equation}
2232: as can be seen easily from the dashed box in the cartoon above.  (Note
2233: that this result is independent of how many sites we include in the
2234: dashed box.)  After the projection onto rep.\ $\bs{10}$ on each lattice
2235: site, we find that only reps.\ $\bs{\overline{10}}=(0,3)$ and
2236: $\bs{27}=(2,2)$ occur for the total spin of two neighboring sites for the
2237: $\bs{10}$ VBS state.  With the Casimirs $\casi{0}{3}=6$ and
2238: $\casi{2}{2}=8$ we obtain the parent Hamiltonian
2239: \begin{equation}
2240:   \label{ham.10VBS}
2241:   H_{\bs{10}\,\text{VBS}}
2242:   =\sum_{i=1}^N\,\left(\bigl(\bs{J}_i\bs{J}_{i+1}\bigr)^2
2243:     +5\,\bs{J}_i\bs{J}_{i+1}+6\right),
2244: \end{equation}
2245: the operators $J_i^a$, $a=1,\ldots,8$, are now $10\times 10$ matrices,
2246: and we have used $\bs{J}^2_i=6$.  $H_{\bs{10}\,\text{VBS}}$ is
2247: positive semi-definite and annihilates the $\bs{10}$ VBS state
2248: \eqref{state.10VBS}.  We assume that \eqref{state.10VBS} is the only
2249: ground state of \eqref{ham.10VBS}.
2250: 
2251: The Hamiltonian \eqref{ham.10VBS} provides the equivalent of the AKLT
2252: model~\cite{affleck-87prl799,affleck-88cmp477}, whose unique ground state is
2253: constructed from dimer states by projection onto spin 1, for SU(3)
2254: spin chains.  Note that as in the case of SU(2), it is sufficient to
2255: consider linear and quadratic powers 
2256: %(\ie $\bs{S}_i\bs{S}_{i+1}$ or $\bs{J}_i\bs{J}_{i+1}$) 
2257: of the total spin of only two neighboring
2258: sites.  This is a general feature of the corresponding SU($n$) models,
2259: as we will elaborate in the following section.
2260: 
2261: Since the $\bs{10}$ VBS state \eqref{state.10VBS} is unique, we cannot
2262: have domain walls connecting different ground states.  We hence expect
2263: the coloron and anti-coloron excitations to be confined in pairs, as
2264: illustrated below.  The state between the excitations is no longer
2265: annihilated by \eqref{ham.10VBS}, as there are pairs of neighboring
2266: sites containing higher-dimensional representations, as indicated by
2267: the dotted box below.  As the number of such pairs increases linearly
2268: with the distance between the excitation, the confinement potential
2269: depends linearly on this distance.
2270: \begin{equation}
2271: \setlength{\unitlength}{1pt}
2272: \begin{picture}(210,70)(0,0)
2273: \linethickness{0.8pt}
2274: \multiput(0,45)(14,0){15}{\circle{4}}
2275: \multiput(42,45)(14,0){2}{\circle*{4}}
2276: \multiput(154,45)(14,0){1}{\circle*{4}}
2277: \multiput(14,35)(14,0){15}{\circle{4}}
2278: \multiput(28,25)(14,0){15}{\circle{4}}
2279: \multiput(182,25)(14,0){3}{\circle{4}}
2280: \thicklines
2281: \put(2,45){\line(1,0){10}}\put(16,45){\line(1,0){10}}
2282: \put(44,45){\line(1,0){10}}\put(72,45){\line(1,0){10}}
2283: \put(86,45){\line(1,0){10}}\put(114,45){\line(1,0){10}}
2284: \put(128,45){\line(1,0){10}}\put(170,45){\line(1,0){10}}
2285: \put(184,45){\line(1,0){10}}
2286: \put(16,35){\line(1,0){10}}\put(30,35){\line(1,0){10}}
2287: \put(58,35){\line(1,0){10}}\put(72,35){\line(1,0){10}}
2288: \put(100,35){\line(1,0){10}}\put(114,35){\line(1,0){10}}
2289: \put(142,35){\line(1,0){10}}\put(156,35){\line(1,0){10}}
2290: \put(184,35){\line(1,0){10}}\put(198,35){\line(1,0){10}}
2291: \put(30,25){\line(1,0){10}}\put(44,25){\line(1,0){10}}
2292: \put(72,25){\line(1,0){10}}\put(86,25){\line(1,0){10}}
2293: \put(114,25){\line(1,0){10}}\put(128,25){\line(1,0){10}}
2294: \put(156,25){\line(1,0){10}}\put(170,25){\line(1,0){10}}
2295: \put(198,25){\line(1,0){10}}\put(212,25){\line(1,0){10}}
2296: \thinlines
2297: \put(105,11){\vector(-1,0){49}}
2298: \put(105,11){\vector(1,0){49}}
2299: \put(105,0){\makebox(0,0){\small energy cost $\propto$ distance}}
2300: %
2301: \put(49,56){\makebox(0,0){$\bs{\bar 3}$}}
2302: \put(154,56){\makebox(0,0){$\bs{3}$}}
2303: \put(49,67){\makebox(0,0){\small coloron}}
2304: \put(154,67){\makebox(0,0){\small anti-coloron}}
2305: \put(93,20){\dashbox{1}(24,30)}
2306: \end{picture}\label{10VBSstate.exc}
2307: \end{equation}
2308: In principle, it would also be possible to create three colorons (or
2309: three anti-colorons) rather than a coloron--anti-coloron pair, but as
2310: all three excitations would feel strong confinement forces, we expect
2311: the coloron--anti-coloron pair to constitute the dominant low energy
2312: excitation.  The confinement force between the pair induces a
2313: linear oscillator potential for the relative motion of the
2314: constituents.  The zero-point energy of this oscillator gives rise to
2315: a Haldane-type energy gap
2316: (see~\cite{greiter02prb134443,greiter02prb054505} for a similar
2317: discussion in the two-leg Heisenberg ladder), which is independent of
2318: the model specifics.  We expect this gap to be a generic feature of
2319: rep.\ $\bs{10}$ spin chains with short-range antiferromagnetic
2320: interactions.
2321: 
2322: 
2323: \subsection{The representation 8 VBS}
2324: \label{sec:8VBS}
2325: 
2326: To construct a representation $\bs{8}$ VBS state, consider first a chain
2327: with alternating representations $\bs{3}$ and $\bs{\bar{3}}$ on
2328: neighboring sites, which we combine into singlets.  This can be done in
2329: two ways, yielding the two states  
2330: \begin{equation}
2331: \setlength{\unitlength}{1pt}
2332: \begin{picture}(210,22)(0,-12)
2333: \linethickness{0.8pt}
2334: \multiput(5,0)(30,0){3}{\circle{4}}
2335: \multiput(20,0)(30,0){3}{\circle{6}}
2336: \multiput(138,0)(30,0){3}{\circle{4}}
2337: \multiput(153,0)(30,0){3}{\circle{6}}
2338: %
2339: \thicklines
2340: \multiput(7,0)(30,0){3}{\line(1,0){10}}
2341: \multiput(156,0)(30,0){2}{\line(1,0){10}}
2342: %
2343: \put(136,0){\line(-1,0){4}}
2344: \put(216,0){\line(1,0){3}}
2345: %
2346: \thinlines
2347: \put(109,0){\makebox(0,0){and}}
2348: \put(224,-3){\makebox(0,0){.}}
2349: \put(35,-12){\makebox(0,0){\small $\bs{3}$}}
2350: \put(50,-12){\makebox(0,0){\small $\bs{\bar{3}}$}}
2351: \end{picture}
2352: \nonumber
2353: %\label{fig:33bar}
2354: \end{equation}
2355: We then combine a $\bs{3}$--$\bs{\bar{3}}$ state with an identical one
2356: shifted by one lattice spacing.  This yields representations
2357: $\bs{3}\otimes\bs{\bar{3}}=\bs{1}\oplus\bs{8}$ at each site.  The
2358: $\bs{8}$ VBS state is obtained by projecting onto %the adjoint
2359: representation $\bs{8}$.  Corresponding to the two
2360: $\bs{3}$--$\bs{\bar{3}}$ states illustrated above, we obtain two
2361: linearly independent $\bs{8}$ VBS states, $\Psi^\text{L}$ and
2362: $\Psi^\text{R}$, which may be visualized as
2363: \begin{equation}
2364: \setlength{\unitlength}{1pt}
2365: %\begin{picture}(210,50)(0,-20)
2366: \begin{picture}(210,60)(0,-24)
2367: \linethickness{0.8pt}
2368: \multiput(5,10)(30,0){3}{\circle{4}}
2369: \multiput(20,10)(30,0){3}{\circle{6}}
2370: \multiput(138,10)(30,0){3}{\circle{4}}
2371: \multiput(153,10)(30,0){3}{\circle{6}}
2372: \multiput(5,0)(30,0){3}{\circle{6}}
2373: \multiput(20,0)(30,0){3}{\circle{4}}
2374: \multiput(138,0)(30,0){3}{\circle{6}}
2375: \multiput(153,0)(30,0){3}{\circle{4}}
2376: %
2377: \thicklines
2378: \multiput(7,10)(30,0){3}{\line(1,0){10}}
2379: \multiput(156,10)(30,0){2}{\line(1,0){10}}
2380: \multiput(22,0)(30,0){2}{\line(1,0){10}}
2381: \multiput(141,0)(30,0){3}{\line(1,0){10}}
2382: %
2383: \put(136,10){\line(-1,0){4}}
2384: \put(216,10){\line(1,0){3}}
2385: \put(2,0){\line(-1,0){3}}
2386: \put(82,0){\line(1,0){4}}
2387: %
2388: \thinlines
2389: \put(109,5){\makebox(0,0){and}}
2390: \put(224,2){\makebox(0,0){.}}
2391: %
2392: \put(44,-6){\framebox(12,22)}
2393: \put(50,-6){\line(0,-1){9}}
2394: \put(50,-22){\makebox(1,1){\small projection onto $\bs{8}=(1,1)$}}
2395: \put(50,25){\makebox(1,1){\small one site}}
2396: \put(162.5,-6){\dashbox{2}(26,22)}
2397: \end{picture}
2398: \label{fig:8VBS}
2399: \end{equation}
2400: These states transform into each other under space reflection or color 
2401: conjugation (interchange of $\bs{3}$ and $\bs{\bar{3}}$).
2402: % $(\bs{3},\bs{\bar{3}})\to (\bs{\bar{3}},\bs{3})$.  
2403: 
2404: 
2405: It is convenient to formulate the corresponding state vectors as a
2406: matrix product.  Taking (\b,\r,\g ) and (\y,\c,\m ) as bases for the
2407: reps.\ $\bs{3}$ and $\bs{\bar{3}}$, respectively, the singlet bonds in
2408: $\Psi^\text{L}$ above can be written
2409: \begin{multline}
2410:   \Bigl(\ket{\b}_i \ket{\y}_{i+1} + \ket{\r}_i \ket{\c}_{i+1}
2411:   + \ket{\g}_i \ket{\m}_{i+1}\Bigr)\\ =
2412:   \Bigl(\ket{\b}_i,\ket{\r}_i,\ket{\g}_i\Bigr)
2413:   \left(\!
2414:     \begin{array}{c} \ket{\y}_{i+1}\\\ket{\c}_{i+1}\\\ket{\m}_{i+1}\end{array}
2415:   \!\right)\!.
2416: \nonumber
2417: \end{multline}
2418: We are hence led to consider matrices composed of the outer product
2419: of these vectors on each lattice site,
2420: \begin{displaymath}
2421:   %\tilde 
2422:   M_i^{\bs{1}\oplus\bs{8}} 
2423:   = \left(\!
2424:     \begin{array}{c} \ket{\y}_{i}\\\ket{\c}_{i}\\\ket{\m}_{i}\end{array}
2425:     \!\right)\!
2426:   \Bigl(\ket{\b}_i,\ket{\r}_i,\ket{\g}_i\Bigr).
2427: \end{displaymath}\\
2428: In the case of the AKLT model reviewed above, the Schwinger bosons take
2429: care of the projection automatically, and we can simply assemble
2430: these matrices into a product state.  For the $\bs{8}$ VBS,
2431: however, we need to enforce the projection explicitly.  This is
2432: most elegantly accomplished using the Gell-Mann matrices, yielding the
2433: projected matrix
2434: \begin{equation}
2435:   \label{eq:m8}
2436:   M_i%^{\bs{8}}
2437:   =\frac{1}{2}\sum_{a=1}^8 \lambda^a\, 
2438:   \text{tr}\Bigl(\lambda^a M_i^{\bs{1}\oplus\bs{8}}\Bigr).
2439: \end{equation}
2440: Here we have simply used the fact that the eight Gell-Mann matrices,
2441: supplemented by the unitary matrix, constitute a complete basis for
2442: the space of all complex $3\times 3$ matrices.  By omitting the unit
2443: matrix in the expansion \eqref{eq:m8}, we effectively project out
2444: the singlet state.  
2445: Written out explicitly, we obtain 
2446: \begin{widetext}
2447: \begin{equation}
2448:   \label{eq:m8explictly}
2449:   M_i%^{\bs{8}}
2450:   =\left(\!\begin{array}{ccc} 
2451:   \frac{2}{3}\ket{\b\y}_i-\frac{1}{3}\ket{\r\c}_i-\frac{1}{3}\ket{\g\m}_i
2452:   &\ket{\r\y}_i&\ket{\g\y}_i\\[4pt]
2453:   \ket{\b\c}_i
2454:   &-\frac{1}{3}\ket{\b\y}_i+\frac{2}{3}\ket{\r\c}_i-\frac{1}{3}\ket{\g\m}_i
2455:   &\ket{\g\c}_i\\[4pt]
2456:   \ket{\b\m}_i&\ket{\r\m}_i
2457:   &-\frac{1}{3}\ket{\b\y}_i-\frac{1}{3}\ket{\r\c}_i+\frac{2}{3}\ket{\g\m}_i 
2458:   \\[4pt]    
2459:     \end{array}\!\right).
2460: \end{equation}
2461: \end{widetext}
2462: Assuming PBCs, the $\bs{8}$ VBS state $\Psi^\text{L}$ (illustrated in
2463: on the left \eqref{fig:8VBS}) is hence given by the trace of the
2464: matrix product
2465: \begin{equation}
2466:   \label{eq:8VBSL}
2467:   \ket{\psi_{\bs{8}\,\text{VBS}}^\text{L}}
2468:   =\text{tr}\biggl( \prod_i M_i%^{\bs{8}}
2469:   \biggr).
2470: \end{equation}
2471: To obtain the state $\Psi^\text{R}$ (illustrated on the right in
2472: \eqref{fig:8VBS}) we simply have to transpose the matrices in the
2473: product,
2474: \begin{equation}
2475:   \label{eq:8VBSR}
2476:   \ket{\psi_{\bs{8}\,\text{VBS}}^\text{R}}
2477:   =\text{tr}\biggl( \prod_i M_i%^{\bs{8}}
2478:   ^\text{T} \biggr).
2479: \end{equation}
2480: 
2481: Let us now formulate a parent Hamiltonian for these states.
2482: If we consider two lattice sites on an SU(3) chain with a representation
2483: $\bs{8}$ on each lattice site in general, we find the full SU(3) content
2484: \begin{equation}
2485: \bs{8}\,\otimes\,\bs{8}\,=\,\bs{1}\,\oplus\,2\cdot\bs{8}\,
2486: \oplus\,\bs{10}\,\oplus\,\bs{\overline{10}}\,\oplus\,\bs{27}
2487: \end{equation}
2488: with $\bs{10}=(3,0)$, $\bs{\overline{10}}=(0,3)$, and $\bs{27}=(2,2)$.
2489: On the other hand, for the $\bs{8}$ VBS states only the
2490: representations $\bs{3}\otimes\bs{\bar 3}\,=\,\bs{1}\oplus\bs{8}$ can
2491: occur for the total spin of two neighboring sites, as the two sites
2492: always contain one singlet (see dashed box in \eqref{fig:8VBS} on the
2493: right above).  With the Casimirs $\casi{0}{0}=0$ and $\casi{1}{1}=3$
2494: for representations $\bs{1}$ and $\bs{8}$, respectively, we construct
2495: the parent Hamiltonian
2496: \begin{equation}
2497: \label{ham.8VBS}
2498: H_{\bs{8}\,\text{VBS}}=\sum_{i=1}^N\left(\bigl(\bs{J}_i\bs{J}_{i+1}\bigr)^2+
2499: \frac{9}{2}\,\bs{J}_i\bs{J}_{i+1} + \frac{9}{2}\right),
2500: \end{equation}
2501: where the operators $J_i^a$, $a=1,\ldots,8$, are now $8\times 8$
2502: matrices, and we have used the Casimir $\bs{J}_i^2=3$ on each site.
2503: $H_{\bs{8}\,\text{VBS}}$ is positive semi-definite, and annihilates the
2504: states $\Psi^\text{L}$ and $\Psi^\text{R}$.  We have numerically
2505: verified for chains with $N=3$, 4, 5, and 6 lattice sites that
2506: $\Psi^\text{L}$ and $\Psi^\text{R}$ are the only ground states of
2507: \eqref{ham.8VBS}.
2508: 
2509: Naively, one might assume the $\bs{8}$ VBS model to support deconfined
2510: spinons or colorons, which correspond to domain walls between the
2511: two ground states $\Psi^\text{L}$ and $\Psi^\text{R}$.  A closer look
2512: at the domain walls, however, shows that this is highly unlikely,
2513: as each domain wall is a bound state of either two anti-colorons or two
2514: colorons, as illustrated below.
2515: \begin{equation}
2516: \setlength{\unitlength}{1pt}
2517: %\begin{picture}(210,50)(0,-20)
2518: \begin{picture}(210,64)(0,-24)
2519: \linethickness{0.8pt}
2520: \multiput(5,10)(30,0){7}{\circle{4}}
2521: \multiput(20,10)(30,0){7}{\circle{6}}
2522: \multiput(35,0)(30,0){7}{\circle{6}}
2523: \multiput(20,0)(30,0){7}{\circle{4}}
2524: \multiput(65,10)(30,0){1}{\circle*{4}}
2525: \multiput(50,0)(30,0){1}{\circle*{4}}
2526: \multiput(170,10)(30,0){1}{\circle*{6}}
2527: \multiput(155,0)(30,0){1}{\circle*{6}}
2528: %
2529: \thicklines
2530: \multiput(7,10)(30,0){2}{\line(1,0){10}}
2531: \multiput(83,10)(30,0){3}{\line(1,0){10}}
2532: \multiput(187,10)(30,0){1}{\line(1,0){10}}
2533: \multiput(22,0)(30,0){1}{\line(1,0){10}}
2534: \multiput(68,0)(30,0){3}{\line(1,0){10}}
2535: \multiput(172,0)(30,0){2}{\line(1,0){10}}
2536: %
2537: \thinlines
2538: %\put(50,0){\line(3,2){13}}
2539: \put(22,-15){\makebox(0,0){\small $\Psi^\text{L}$}}
2540: \put(112,-15){\makebox(0,0){\small $\Psi^\text{R}$}}
2541: \put(202,-15){\makebox(0,0){\small $\Psi^\text{L}$}}
2542: %
2543: \put(50,22){\makebox(0,0){\small $\bs{3}$}}
2544: \put(65,22){\makebox(0,0){\small $\bs{3}$}}
2545: \put(155,22){\makebox(0,0){\small $\bs{\bar{3}}$}}
2546: \put(170,22){\makebox(0,0){\small $\bs{\bar{3}}$}}
2547: %
2548: \put(57.5,34){\makebox(0,0){\small anti-colorons}}
2549: \put(162.5,34){\makebox(0,0){\small colorons}}
2550: \end{picture}
2551: \label{fig:8VBSdomainwall}
2552: \end{equation}
2553: There is no reason to assume that the domain wall depicted above as
2554: two anti-colorons in fact corresponds to a single coloron, as it
2555: appears to be the case for the trimer chain.  There we created a
2556: domain wall corresponding to a single coloron by removing one of the
2557: rep.\ $\bs{3}$ spins from a trimer, leaving the remaining rep.\
2558: $\bs{3}$ spins coupled antisymmetrically as in the ground state.
2559: %There, however, we created a domain wall corresponding to a single
2560: %coloron by removing one of the rep.\ $\bs{3}$ spins from a trimer.  
2561: If we were to combine the two reps.\ $\bs{{3}}$ into a rep.\
2562: $\bs{\bar{3}}$ in \eqref{fig:8VBSdomainwall}, we would not reproduce a
2563: correlation present in the ground state, but enforce a new
2564: correlation.  The correct interpretation of the domain wall between
2565: $\Psi^\text{L}$ and $\Psi^\text{R}$ is hence that of a bound state between
2566: two linearly confined anti-colorons.  The origin of the confining
2567: potential is illustrated below.
2568: \begin{equation}
2569: \setlength{\unitlength}{1pt}
2570: \begin{picture}(175,70)(0,-30)
2571: \linethickness{0.8pt}
2572: \multiput(5,10)(30,0){6}{\circle{4}}
2573: \multiput(20,10)(30,0){5}{\circle{6}}
2574: \multiput(35,0)(30,0){5}{\circle{6}}
2575: \multiput(20,0)(30,0){6}{\circle{4}}
2576: \multiput(125,10)(30,0){1}{\circle*{4}}
2577: \multiput(50,0)(30,0){1}{\circle*{4}}
2578: %
2579: \thicklines
2580: \multiput(7,10)(30,0){4}{\line(1,0){10}}
2581: \multiput(143,10)(30,0){1}{\line(1,0){10}}
2582: \multiput(22,0)(30,0){1}{\line(1,0){10}}
2583: \multiput(68,0)(30,0){4}{\line(1,0){10}}
2584: %
2585: \thinlines
2586: \put(22,-15){\makebox(0,0){\small $\Psi^\text{L}$}}
2587: \put(157,-15){\makebox(0,0){\small $\Psi^\text{R}$}}
2588: %
2589: \put(50,23){\makebox(0,0){\small $\bs{3}$}}
2590: \put(125,23){\makebox(0,0){\small $\bs{3}$}}
2591: \put(50,34){\makebox(0,0){\small anti-coloron}}
2592: \put(125,34){\makebox(0,0){\small anti-coloron}}
2593: %
2594: \put(87.5,-14){\vector(-1,0){37.5}}
2595: \put(87.5,-14){\vector(1,0){37.5}}
2596: \put(87.5,-25){\makebox(0,0){\small energy cost $\propto$ distance}}
2597: \end{picture}
2598: \label{fig:8VBS3bar3bar}
2599: \end{equation}
2600: As in the $\bs{10}$ VBS, the confinement induces a linear oscillator
2601: potential for the relative motion of the anti-colorons. The zero-point
2602: energy of this oscillator corresponds to a Haldane-type gap in the
2603: spectrum.  The ground state
2604: wave function of the oscillator is symmetric, and hence corresponds to
2605: a symmetric combination of $\bs{3}\otimes\bs{3}$, \ie rep.\ $\bs{6}$.
2606: The antisymmetric combination $\bs{\bar{3}}$ corresponds to the first
2607: excited state of the oscillator, which we expect to cost more than
2608: twice the energy of the symmetric state~\cite{greiter02prb054505}.
2609: This statement holds for the pair of colorons in
2610: \eqref{fig:8VBSdomainwall} as well.
2611: 
2612: \begin{figure}[t]
2613: \includegraphics[angle=270,scale=0.53]{rep8spec.eps}
2614: %%BoundingBox: 176 43 574 530 (new)
2615: \begin{picture}(0,0)
2616: \put(13,5){\makebox(0,0){$k~[2\pi/N]$}}
2617: \put(-120,110){\rotatebox{90}{$E_{\bs{8}\,\text{VBS}}$}}
2618: \put(105,61){\makebox(0,0)[r]{{\scriptsize singlet}}}
2619: \put(105,53){\makebox(0,0)[r]{{\scriptsize rep.\,$\bs{8}$}}}
2620: \put(105,46){\makebox(0,0)[r]{{\scriptsize rep.\,$\bs{10}$}}}
2621: \put(105,38){\makebox(0,0)[r]{{\scriptsize rep.\,$\bs{27}$}}}
2622: \end{picture}
2623: \caption{%(Color online)
2624:   Spectrum of the $\bs{8}$ VBS Hamiltonian \eqref{ham.8VBS} for $N=8$
2625:   sites obtained by exact diagonalization.  (The lines are merely
2626:   guides to the eye.)  The ``magnon'' excitation transforming under
2627:   rep.\ $\bs{8}$ of SU(3) has the lowest energy, followed by a singlet
2628:   excitation, as expected from the discussion in the text.  The well
2629:   defined modes at low energies provide strong evidence of
2630:   coloron--anti-coloron bound states as compared to deconfined domain
2631:   walls, and hence support our conclusion that the $\bs{8}$ VBS
2632:   exhibits a Haldane gap due to spinon confinement.}
2633: \label{fig:8VBSspec}
2634: \end{figure}
2635: 
2636: The domain walls, however, are not the only low energy excitations.
2637: In either of the ground states, we can create coloron--anti-coloron
2638: bound states, which make no reference to the other ground state, as
2639: illustrated below.
2640: \begin{equation}
2641: \setlength{\unitlength}{1pt}
2642: \begin{picture}(175,70)(0,-30)
2643: \linethickness{0.8pt}
2644: \multiput(5,10)(30,0){6}{\circle{4}}
2645: \multiput(20,10)(30,0){6}{\circle{6}}
2646: \multiput(35,0)(30,0){5}{\circle{6}}
2647: \multiput(20,0)(30,0){5}{\circle{4}}
2648: \multiput(125,0)(30,0){1}{\circle*{6}}
2649: \multiput(50,0)(30,0){1}{\circle*{4}}
2650: 
2651: %
2652: \thicklines
2653: \multiput(7,10)(30,0){6}{\line(1,0){10}}
2654: \multiput(142,0)(30,0){1}{\line(1,0){10}}
2655: \multiput(22,0)(30,0){1}{\line(1,0){10}}
2656: \multiput(68,0)(30,0){2}{\line(1,0){10}}
2657: %
2658: \thinlines
2659: \put(22,-15){\makebox(0,0){\small $\Psi^\text{L}$}}
2660: \put(157,-15){\makebox(0,0){\small $\Psi^\text{L}$}}
2661: %
2662: \put(50,23){\makebox(0,0){\small $\bs{3}$}}
2663: \put(125,23){\makebox(0,0){\small $\bs{\bar{3}}$}}
2664: \put(50,34){\makebox(0,0){\small anti-coloron}}
2665: \put(125,34){\makebox(0,0){\small coloron}}
2666: %
2667: \put(87.5,-14){\vector(-1,0){37.5}}
2668: \put(87.5,-14){\vector(1,0){37.5}}
2669: \put(87.5,-25){\makebox(0,0){\small energy cost $\propto$ distance}}
2670: \end{picture}
2671: \label{fig:8VBS33bar}
2672: \end{equation}
2673: The oscillator model tells us again that the ``symmetric'' combination
2674: of $\bs{3}\otimes\bs{\bar{3}}$, \ie rep.\ $\bs{8}$, has the lowest
2675: energy, which we expect to be comparable, if not identical, to the
2676: energy required to create each of the domain walls above.  
2677: The %antisymmetric combination, \ie the
2678: singlet $\bs{1}$ will have an energy comparable to that of a domain
2679: wall transforming under either $\bs{\bar{3}}$ or $\bs{3}$.  In any
2680: event, we expect the $\bs{8}$ VBS model to display a Haldane gap due to
2681: coloron confinement.
2682: 
2683: The excitation spectrum of \eqref{ham.8VBS} for a chain with $N=8$
2684: sites and PBCs is shown in Fig.~\ref{fig:8VBSspec}.  The spectrum shows
2685: that the lowest excitation transforms under rep.~$\bs{8}$, as expected
2686: from \eqref{fig:8VBS33bar}, with a singlet and then another rep~$\bs{8}$
2687: following at slightly higher energies.  It is tempting to interpret 
2688: those three levels as the lowest levels of the coloron--anti-coloron
2689: oscillator \eqref{fig:8VBS33bar}, but then there should be another 
2690: singlet at a comparable spacing above.  The fact that the spacings
2691: between these excitations are significantly smaller than the energy
2692: of the first exited state, however, would be consistent with such an 
2693: interpretation, as the spinons in VBS models always have a local energy 
2694: cost associated with their creation, which is specific to these models
2695: and not related to the universal Haldane gap stemming from confinement
2696: forces.
2697: 
2698: Most importantly, the spectrum provides strong evidence in favor of
2699: our assumption that the domain walls are not elementary excitations,
2700: but bound states of either two colorons or two anti-colorons, and
2701: hence that the lowest energy excitations of finite chains are
2702: coloron--anti-coloron bound states as illustrated in
2703: \eqref{fig:8VBS33bar}.  The assumption is crucial for our conclusion
2704: that the model exhibits a Haldane gap.  If the low energy sector of
2705: the model was determined by two deconfined domain walls, we would see
2706: a continuum of states in the spectrum, similar to the spectrum seen in
2707: spin $S=\frac{1}{2}$ chains of SU(2).  The well defined low-energy
2708: modes in Fig.~\ref{fig:8VBSspec}, however, look much more like the
2709: spinon--spinon bound state excitations seen in $S=1$ chains
2710: or two-leg $S=\frac{1}{2}$ Heisenberg ladders.  In particular, if we
2711: assume that the individual domain walls transform under reps.\
2712: $\bs{6}$ and $\bs{\bar{6}}$, we expect excitations transforming under
2713: the representations contained in
2714: $\bs{6}\otimes\bs{\bar{6}}=\bs{1}\oplus\bs{8}\oplus\bs{27}$ to be
2715: approximately degenerate.  Fig.~\ref{fig:8VBSspec} shows clearly that
2716: such a multiplet is not present a the lowest energies.
2717: 
2718: 
2719: 
2720: 
2721: \section{SU($\bs{n}$) models}
2722: \label{sec:sun}
2723: 
2724: In this section, we generalize three of the models proposed for SU(3)
2725: spin chains, the trimer model, the symmetric representation $\bs{10}$ VBS, 
2726: and the matrix product state $\bs{8}$ VBS to the case of 
2727: SU($n$) spin chains.
2728: 
2729: \subsection{The $\bs{n}$-mer model}
2730: 
2731: Consider an  SU($n$) spin chain with $N$ sites, where $N$ is a multiple of $n$,
2732: with a spin transforming according 
2733: to the fundamental representation $\bs{n}$ of  SU($n$) at each lattice site,
2734: \begin{equation}
2735: \begin{array}{rcccccccl}
2736: \bs{n}&=&(1,0,\ldots,0)&=&
2737: \setlength{\unitlength}{8pt}
2738: \begin{picture}(1,1)(0,0.1)
2739: \put(0,1){\line(1,0){1}}
2740: \put(0,0){\line(1,0){1}}
2741: \put(0,0){\line(0,1){1}}
2742: \put(1,0){\line(0,1){1}}
2743: \end{picture}
2744: &\hat =&c_{\sigma}^\dagger\vac, 
2745: \end{array}
2746: \end{equation}
2747: where $\sigma$ denotes a ``flavor'', $\sigma\in\{{\rm f}_1,\ldots,{\rm
2748:   f}_n\}$, and $c_{\sigma}^\dagger$ creates a fermion of flavor
2749: $\sigma$.
2750:   
2751: The SU($n$) generators at site $i$ are in analogy to
2752: \eqref{eq:s} and \eqref{eq:J_a} defined as
2753: \begin{equation}
2754: J^a_i=\frac{1}{2}\,\sum_{\sigma,\sigma'={\rm f}_1,\ldots,{\rm f}_n}\,
2755: c_{i\sigma}^{\dagger} V^a_{\sigma\sigma'}
2756: c_{i\sigma'}^{\phantom{\dagger}}, \quad a=1,\ldots,n^2-1,
2757: \label{eq:J_an}
2758: \end{equation}
2759: where the $V^a$ denote the $n^2\!-\!1$ SU($n$) Gell-Mann
2760: matrices~\cite{macfarlane-68cmp77}.  The generators are normalized
2761: through the eigenvalue the quadratic Casimir operator takes in the
2762: adjoint representation, $\bs{J}^2=\casin{1,0,\ldots,0,1}=n$.
2763: 
2764: To determine the eigenvalues of the quadratic Casimir for general
2765: representations of SU($n$), a significant amount of representation
2766: theory is required~\cite{Humphreys87}.  We content ourselves here
2767: by providing the formulas up to $n=6$ in App.~\ref{quadraticcasimirs}.
2768: 
2769: In analogy to the trimer states \eqref{eq:trimer}, we construct 
2770: the $n$-mer states of an SU($n$) spin chain %are simply constructed 
2771: by combining sets of $n$ neighboring spins into a singlet,
2772: \begin{equation}
2773:   \label{eq:nmer}
2774:   \ket{\psi_{n\text{-mer}}^{(\mu)}\!}=\hspace{-15pt}\prod_{\scriptstyle{i} \atop 
2775:     \left(\scriptstyle{\frac{i-\mu}{n}\,{\rm integer}}\right)}
2776:   \hspace{-5pt}\Bigl(
2777:   \sum_{\scriptstyle{(\sigma_1,\ldots,\sigma_n)}=
2778:     \atop\scriptstyle{{\pi}({\rm f}_1,\ldots,{\rm f}_n)}} 
2779:   \hspace{-10pt}\hbox{sign}({\pi})\,
2780:   \prod_{\kappa=1}^n c^{\dagger}_{i-1+\kappa\;\sigma_\kappa}
2781:   \Bigr)\!\vac\!,
2782: \end{equation}
2783: where $\mu=1,\ldots,n$ labels the $n$ degenerate ground states, and
2784: $i$ runs over the lattice sites subject to the constraint that
2785: $\frac{i-\mu}{n}$ is integer.  The sum extends over all $n!$
2786: permutations $\pi$ of the $n$ flavors ${\rm f}_1,\ldots,{\rm f}_n$.
2787: 
2788: In order to identify a parent Hamiltonian, consider the total SU($n$)
2789: spin on $n+1$ neighboring sites for the $n$-mer states.  Following the
2790: rules of combining representations labeled by Young tableaux (see \eg
2791: \cite{Cornwell84vol2,Georgi82}), it is not difficult to see that the
2792: total spin will only contain representations given by tableaux with
2793: $n+1$ boxes and two columns, \ie tableaux of the form
2794: \begin{displaymath}
2795: \setlength{\unitlength}{8pt}
2796: \begin{picture}(5,7.5)
2797: \put(0,0){\line(0,1){7.5}}
2798: \put(1,0){\line(0,1){7.5}}
2799: \put(2,3.5){\line(0,1){4}}
2800: %
2801: \put(0,0){\line(1,0){1}}
2802: \put(0,1){\line(1,0){1}}
2803: \put(0,2.5){\line(1,0){1}}
2804: \put(0,3.5){\line(1,0){2}}
2805: \put(0,4.5){\line(1,0){2}}
2806: \put(0,6.5){\line(1,0){2}}
2807: \put(0,7.5){\line(1,0){2}}
2808: \put(4.5,5.5){\makebox(0,0)
2809: {$\left.\begin{array}{c}~\\~\end{array}\right\}{\small ~\nu~\text{rows}}$}}
2810: \end{picture}
2811: \end{displaymath}
2812: with $1\le\nu\le\frac{n+1}{2}$.  The 
2813: eigenvalues of the quadratic Casimir operator for these
2814: representations are
2815: \begin{equation}
2816: f_n(\nu)=\frac{1}{2n}\Bigl(n^2(2\nu -1) -2n(\nu -1)^2 -1 \Bigr).
2817: \end{equation}
2818: An educated guess for a parent Hamiltonian for the $n$-mer chain
2819: hence appears to be
2820: \begin{equation}
2821: H_{\text{trial}}=\sum_{i=1}^N H_i\quad
2822: \label{eq:MGSUni}
2823: \end{equation}
2824: with
2825: \begin{equation}
2826: H_i=\prod_{\nu=1}^{\lfloor\frac{n+1}{2}\rfloor}
2827: \left(\!\left(\bs{J}_i^{(n+1)}\right)^2 - f_n(\nu)\right),
2828: \label{eq:MGSUn}
2829: \end{equation}
2830: where $\lfloor \; \rfloor$ denotes the floor function, \ie $\lfloor
2831: x\rfloor$ is the largest integer $l\le x$, and we use the notation
2832: introduced in \eqref{eq:defJnu}.  
2833: 
2834: This construction yields the MG model~\cite{majumdar-69jmp1399} for
2835: SU(2), the trimer model \eqref{ham.trimer} for SU(3), and a valid
2836: parent Hamiltonian for the four degenerate 4-mer states for SU(4).
2837: For $n\ge 5$, however, the decomposition of the tensor product
2838: $\bs{n}^{\otimes (n+1)}$ contains irreducible representations
2839: corresponding to Young tableaux with more than two columns, whose
2840: Casimirs are equal or smaller than a number of Casimirs included in
2841: the list $f_n(\nu)$, $\nu=1,2,\ldots,\lfloor\frac{n+1}{2}\rfloor$.  If
2842: the Casimir of such an ``undesired'' representation not included in
2843: the list is smaller than an odd number of Casimirs included in the
2844: list, we obtain negative eigenvalues for $H_i$, and it is not a priori
2845: clear any more that the Hamiltonian \eqref{eq:MGSUni} is positive
2846: semi-definite.  An obvious cure to this problem is to write
2847: \begin{equation}
2848:   H_{n\text{-mer}}=\sum_{i=1}^N H_i^2,\quad
2849:   \label{eq:hisquare}
2850: \end{equation}
2851: with $H_i$ as in \eqref{eq:MGSUn}.  This does, however, not cure
2852: potential problems arising from undesired representations which 
2853: %happen to 
2854: share the eigenvalues of the Casimir with one of the
2855: representations from the list, as it happens to be the case for $n=5$.
2856: The Hamiltonian \eqref{eq:hisquare} likewise annihilates these
2857: representations, giving rise to a remote possibility that the $n$-mer
2858: states \eqref{eq:nmer} are not the only ground states of
2859: \eqref{eq:hisquare}.  The potential relevance of these problems has to
2860: be investigated for each $n$ separately.
2861: 
2862: \subsection{The representation $\bs{(n,0,\ldots,0)}$ VBS}
2863: 
2864: As a generalization of the AKLT model for SU(2) and the ${\bs{10}}$ VBS
2865: model for SU(3) discussed above, we now consider a VBS for an SU($n$) chain
2866: of spins transforming under the symmetric representation
2867: \begin{equation}
2868:   \label{eq:n000}
2869:   \begin{array}{rcccl}
2870:     (n,0,\ldots,0)&=&
2871:     \setlength{\unitlength}{8pt}
2872:     \begin{picture}(5,2)(0,3.15)
2873:       \put(0,4){\line(1,0){5}}
2874:       \put(0,3){\line(1,0){5}}
2875:       \put(0,3){\line(0,1){1}}
2876:       \put(1,3){\line(0,1){1}}
2877:       \put(4,3){\line(0,1){1}}
2878:       \put(5,3){\line(0,1){1}}
2879:       \put(2.5,1.5){\makebox(0,0)%
2880:         {$\underbrace{\quad\text{~~~}\qquad}_{\text{\small $n$~boxes}}$}}
2881:    \end{picture}
2882:     &\hat =& b_{\sigma_1}^\dagger b_{\sigma_2}^\dagger\ldots b_{\sigma_n}^\dagger
2883:     \vac,\\[2pt] \nonumber
2884: \end{array}
2885: \end{equation}
2886: where each $b_{\sigma}^\dagger$, $\sigma\in\{{\rm f}_1,\ldots,{\rm
2887:   f}_n\}$, is an SU($n$) Schwinger boson.  The VBS state is obtained
2888: by combining $n$ $n$-mer states \eqref{eq:nmer}, one for each
2889: $\mu=1,\dots,n$, in that the total spin on each lattice site is
2890: projected onto the symmetric representation $(n,0,\ldots,0)$.  This
2891: yields
2892: \begin{equation}
2893:   \label{eq:n000VBS}
2894:   \ket{\psi_{(n,0,\ldots,0)\, \text{VBS}}}=
2895:   \prod_{i} \Bigl(\!\!
2896:   \sum_{\scriptstyle{(\sigma_1,\ldots,\sigma_n)}=
2897:     \atop\scriptstyle{{\pi}({\rm f}_1,\ldots,{\rm f}_n)}} 
2898:   \hspace{-12pt}\hbox{sign}({\pi})
2899:   \prod_{\kappa=1}^n b^{\dagger}_{i-1+\kappa,\sigma_\kappa}
2900:   \Bigr)\!\vac\!.
2901: \end{equation}
2902: 
2903: Let us now construct a parent Hamiltonian for the symmetric VBS
2904: \eqref{eq:n000VBS}.  The total SU($n$) spin of two neighboring sites
2905: of a representation $(n,0,\ldots,0)$ spin chain in general
2906: contains all the representations corresponding to Young tableaux 
2907: with $2n$ boxes and at most two rows, \ie all tableaux of the form
2908: \begin{displaymath}
2909: \setlength{\unitlength}{8pt}
2910: \begin{picture}(9,5)(0,1.5)
2911: \put(0,6){\line(1,0){9}}
2912: \put(0,5){\line(1,0){9}}
2913: \put(0,4){\line(1,0){5}}
2914: \put(0,4){\line(0,1){2}}
2915: \put(1,4){\line(0,1){2}}
2916: \put(4,4){\line(0,1){2}}
2917: \put(5,4){\line(0,1){2}}
2918: \put(6,5){\line(0,1){1}}
2919: \put(8,5){\line(0,1){1}}
2920: \put(9,5){\line(0,1){1}}
2921: \put(2.5,2.5){\makebox(0,0)%
2922: {$\underbrace{\quad\text{~~~}\qquad}_{\text{\small $n-\nu$~columns}\qquad\ }$}}
2923: \put(7,2.5){\makebox(0,0)%
2924: {$\underbrace{\quad\text{~\,~}\quad}_{\quad\text{\small $2\nu$ boxes}}$}}
2925: %\put(17.5,22.5){\makebox(0,0){$\cdots$}}
2926: %\put(17.5,27.5){\makebox(0,0){$\cdots$}}
2927: \end{picture}
2928: \end{displaymath}
2929: The eigenvalues of the quadratic Casimir operator for these
2930: representations are given by
2931: \begin{equation}
2932:   \label{eva:aklt}
2933:   g_n(\nu)\equiv \casin{2\nu,n\!-\!\nu,0,\dots,0}=2n^2-4n+\nu(\nu+1). 
2934: \end{equation}
2935: 
2936: On the other hand, the total SU($n$) spin of two neighboring sites of
2937: the representation $(n,0,\ldots,0)$ VBS \eqref{eq:n000VBS} has to be
2938: contained in the product
2939: \begin{equation}
2940:   \label{eq:sunprod}
2941:   \setlength{\unitlength}{8pt} 
2942:   \begin{picture}(1.2,1.7)(0,0.6)
2943:     \multiput(0,0)(1,0){2}{\line(0,1){2}}
2944:     \multiput(0,0)(0,1){3}{\line(1,0){1}}
2945:   \end{picture}
2946:   ^{\otimes n-1}\otimes\
2947:   \begin{picture}(1,1)(0,0.1)
2948:     \multiput(0,0)(1,0){2}{\line(0,1){1}}
2949:     \multiput(0,0)(0,1){2}{\line(1,0){1}}
2950:   \end{picture}
2951:   \ \otimes\
2952:   \begin{picture}(1,1)(0,0.1)
2953:     \multiput(0,0)(1,0){2}{\line(0,1){1}}
2954:     \multiput(0,0)(0,1){2}{\line(1,0){1}}
2955:   \end{picture}
2956: \end{equation}
2957: As we project the spin on each lattice site onto the representation
2958: $(n,0,\ldots,0)$, only two these representations remain:
2959: \begin{displaymath}
2960:  \setlength{\unitlength}{8pt}
2961: \begin{picture}(19,2.5)(0,.6)
2962: \linethickness{0.3pt}
2963: \multiput(0,0)(1,0){2}{\line(0,1){2}}
2964: \multiput(4,0)(1,0){2}{\line(0,1){2}}
2965: \multiput(6,1)(1,0){2}{\line(0,1){1}}
2966: \multiput(0,1)(0,1){2}{\line(1,0){7}}
2967: \put(0,0){\line(1,0){5}}
2968: \put(8,0.5){\makebox(3,1){and}}
2969: \multiput(12,0)(1,0){2}{\line(0,1){2}}
2970: \multiput(16,0)(1,0){3}{\line(0,1){2}}
2971: %\multiput(12,0)(1,0){7}{\line(0,1){2}}
2972: \multiput(12,0)(0,1){3}{\line(1,0){6}}
2973: \end{picture}
2974: \end{displaymath}
2975: The eigenvalues of the quadratic Casimir operator are given by
2976: $g_n(0)=2n(n-2)$ and $g_n(1)=2(n-1)^2$, respectively.  Hence, using
2977: $\bs{J}^2_i=n(n-1)$, we obtain the parent Hamiltonian 
2978: %\begin{multline}
2979: \begin{equation}
2980:   \label{ham.aklt.sun}
2981:   \begin{split}
2982:   &H_{{\scriptstyle (n,0,\ldots,0)}\, {\scriptstyle \text{VBS}}}\\
2983: %  H_{\text{SU($n$)\hspace{2pt}VBS}}
2984:   &=\sum_{i=1}^N
2985:   \Bigl(\bigl(\bs{J}_i\bs{J}_{i+1}\bigr)^2 + (2n-1)\,\bs{J}_i\bs{J}_{i+1}+
2986:   n(n-1)\Bigr).
2987:   \end{split}
2988: \end{equation}
2989: %\end{multline}
2990: Since $g_n(0)\ge 0$ for $n\ge 2$ and $g_n(\nu)$ is a strictly
2991: increasing function of $\nu$, the Hamiltonian \eqref{ham.aklt.sun} is
2992: positive semi-definite.  For $n=2$, we recover the AKLT model
2993: \eqref{eq:haklt}; for $n=3$, we recover the $\bs{10}$ VBS model
2994: \eqref{ham.10VBS}.
2995: 
2996: \subsection{An example of a matrix product state}
2997: \label{sec:mp}
2998: 
2999: In principle, a matrix product VBS can be formulated on all SU($n$)
3000: chains with spins transforming under the symmetric combination of any
3001: representation $\bs{m}$ and its conjugate representation
3002: $\bs{\overline{m}}$.  Unless the rep.~$\bs{m}$ is self-conjugate,
3003: %\ie $\bs{m}=\bs{\overline{m}}$, 
3004: we obtain two inequivalent states, which transform into each other
3005: under space reflection.  The construction of these is analogous to the
3006: $\bs{8}$ VBS introduced above, and likewise best illustrated as
3007: \begin{equation}
3008: \setlength{\unitlength}{1.333pt}
3009: \begin{picture}(170,52)(0,-24)
3010: \linethickness{0.8pt}
3011: \multiput(5,10)(30,0){3}{\makebox(0,0){\small $\bs{m}$}}
3012: \multiput(20,10.6)(30,0){3}{\makebox(0,0){\small $\bs{\overline{m}}$}}
3013: \multiput(5,0.6)(30,0){3}{\makebox(0,0){\small $\bs{\overline{m}}$}}
3014: \multiput(20,0)(30,0){3}{\makebox(0,0){\small $\bs{m}$}}
3015: \multiput(128,10)(30,0){2}{\makebox(0,0){\small $\bs{m}$}}
3016: \multiput(143,10.6)(30,0){1}{\makebox(0,0){\small $\bs{\overline{m}}$}}
3017: \multiput(128,0.6)(30,0){2}{\makebox(0,0){\small $\bs{\overline{m}}$}}
3018: \multiput(143,0)(30,0){1}{\makebox(0,0){\small $\bs{m}$}}
3019: %
3020: \thicklines
3021: \multiput(9,10)(30,0){3}{\line(1,0){7}}
3022: \multiput(24,0)(30,0){2}{\line(1,0){7}}
3023: \multiput(147,10)(30,0){1}{\line(1,0){7}}
3024: \multiput(132,0)(30,0){1}{\line(1,0){7}}
3025: %
3026: \put(1,0){\line(-1,0){3}}
3027: \put(84,0){\line(1,0){3}}
3028: \put(124,10){\line(-1,0){3}}
3029: \put(162,0){\line(1,0){3}}
3030: %
3031: \thinlines
3032: \put(104,5){\makebox(0,0){and}}
3033: \put(169,2){\makebox(0,0){.}}
3034: %
3035: \put(44,-6){\framebox(12,22)}
3036: \put(50,-6){\line(0,-1){7}}
3037: \put(50,-20){\makebox(1,1){\small projection onto the symmetric}}
3038: \put(50,-28){\makebox(1,1){\small combination in 
3039: $\bs{m}\otimes\bs{\overline{m}}$}}
3040: \put(50,23){\makebox(1,1){\small one site}}
3041: \put(122.5,-6){\dashbox{2}(26,22)}
3042: \end{picture}
3043: \vspace{10pt}
3044: \label{fig:mmbarVBS}
3045: \end{equation}
3046: The thick lines indicate that we combine pairs of neighboring
3047: representations $\bs{m}$ and $\bs{\overline{m}}$ into singlets.  On
3048: each lattice site, we project onto the symmetric combination of
3049: $\bs{m}$ and $\bs{\overline{m}}$, as indicated.  By ``symmetric
3050: combination'' we mean that if representations $\bs{m}$ and
3051: $\bs{\overline{m}}$ of SU($n$) have Dynkin coordinates
3052: $(\mu_1,\mu_2,\ldots,\mu_{n-1})$ and
3053: $(\mu_{n-1},\mu_{n-2},\ldots,\mu_1)$, respectively, we combine them
3054: into the representation with Dynkin coordinates
3055: $(\mu_1\!+\!\mu_{n-1},\mu_2\!-\!\mu_{n-2},\ldots,\mu_{n-1}\!+\!\mu_1)$.
3056: In other words, we align the columns of both tableaux horizontally,
3057: and hence obtain a tableau with twice the width, without ever adding a
3058: single box vertically to a column of the tableaux we started with.
3059: %
3060: The states \eqref{fig:mmbarVBS} we obtain are translationally
3061: invariant and we expect the parent Hamiltonians to involve
3062: nearest-neighbor interactions only.
3063: 
3064: In this subsection, we will formulate the simplest SU($n$) model
3065: of this general family.  We take $\bs{m}$ to be the representation
3066: formed by antisymmetrically combining $\kappa\le \frac{n}{2}$ fundamental 
3067: representations,
3068: \begin{displaymath}
3069: \bs{m}\;=\;[\kappa]\;\equiv\; (0,\ldots,0,1,0,\ldots,0)\;=\;
3070: \setlength{\unitlength}{8pt}
3071: \begin{picture}(7,3)(0.7,5.2)
3072: \put(1,3.5){\line(0,1){4}}
3073: \put(2,3.5){\line(0,1){4}}
3074: %
3075: \put(1,3.5){\line(1,0){1}}
3076: \put(1,4.5){\line(1,0){1}}
3077: \put(1,6.5){\line(1,0){1}}
3078: \put(1,7.5){\line(1,0){1}}
3079: \put(5,5.5){\makebox(0,0){$\left.\begin{array}{c}~\\~\end{array}\right\}
3080: {\small ~\kappa~\text{boxes}}$\ ,}}
3081: \put(-7,3.6){\line(0,1){0.8}}
3082: \put(-7,2.6){\makebox(0,0){\small $\kappa$-th entry}}
3083: \end{picture}\\[20pt]
3084: \end{displaymath}
3085: which implies that we consider a model with the representation
3086: corresponding to a Young tableaux with a column with $n-\kappa$ boxes
3087: to the left of a column with $\kappa$ boxes at each lattice site:\\
3088: %[\kappa ,n\!-\!\kappa]\;\equiv\;
3089: %(0,\ldots,0,1,0,\ldots,0,1,0,\ldots,0)\;=\; 
3090: \setlength{\unitlength}{8pt}
3091: \begin{picture}(8,8)(-14,0)
3092: \put(-7,3.8){$[\kappa ,n\!-\!\kappa]\;\equiv$}
3093: \put(0,0){\line(0,1){7.5}}
3094: \put(1,0){\line(0,1){7.5}}
3095: \put(2,3.5){\line(0,1){4}}
3096: %
3097: \put(0,0){\line(1,0){1}}
3098: \put(0,1){\line(1,0){1}}
3099: \put(0,2.5){\line(1,0){1}}
3100: \put(0,3.5){\line(1,0){2}}
3101: \put(0,4.5){\line(1,0){2}}
3102: \put(0,6.5){\line(1,0){2}}
3103: \put(0,7.5){\line(1,0){2}}
3104: \put(4.5,5.5){\makebox(0,0){$\left.\begin{array}{c}~\\~\end{array}\right\}
3105: {\small ~\kappa~\text{rows}}$}}
3106: %\put(9,3.6){.}
3107: \end{picture}%\\[26pt]
3108: 
3109: The construction of the parent Hamiltonian is similar to the $n$-mer
3110: model above.  The total spin on two neighboring lattice sites can only
3111: assume representations contained in $\bs{m}\otimes\bs{\overline{m}}$,
3112: \ie representations corresponding to tableaux of the form
3113: \begin{displaymath}
3114: [\nu ,n-\nu]\;=\;
3115: \setlength{\unitlength}{8pt}
3116: \begin{picture}(8.5,4)(-0.3,3.9)
3117: \put(0,0){\line(0,1){7.5}}
3118: \put(1,0){\line(0,1){7.5}}
3119: \put(2,4){\line(0,1){3.5}}
3120: %
3121: \put(0,0){\line(1,0){1}}
3122: \put(0,1){\line(1,0){1}}
3123: \put(0,3){\line(1,0){1}}
3124: \put(0,4){\line(1,0){2}}
3125: \put(0,5){\line(1,0){2}}
3126: \put(0,6.5){\line(1,0){2}}
3127: \put(0,7.5){\line(1,0){2}}
3128: \put(4.5,5.75){\makebox(0,0)
3129: {$\left.\begin{array}{c}~\\[-0.6pt]~\end{array}\right\}
3130: {\small ~\nu~\text{rows}}$}}
3131: \end{picture}\\[26pt] 
3132: \end{displaymath}
3133: with $0\le\nu\le\kappa$.  The eigenvalues of the quadratic
3134: Casimir operator for these representations are
3135: \begin{equation}
3136:   h_n(\nu)=\nu\,(n-\nu+1).
3137: \end{equation}
3138: The obvious proposal for a parent Hamiltonian is hence 
3139: \begin{equation}
3140: H=\sum_{i=1}^N H_i, \quad %\text{with}\quad
3141: H_i=\prod_{\nu=0}^{\lfloor\frac{n}{2}\rfloor}
3142: \left(\!\left(\bs{J}_i^{(2)}\right)^2 - h_n(\nu)\right),
3143: \label{eq:mmbarham}
3144: \end{equation}
3145: where $\lfloor x\rfloor$ denotes again the floor function.  This
3146: Hamiltonian singles out the matrix product state \eqref{fig:mmbarVBS}
3147: as unique ground states for $n\le 5$, but suffers from the same
3148: shortcomings as \eqref{eq:MGSUni} with \eqref{eq:MGSUn} for $n\ge 6$.
3149: 
3150: 
3151: \section{Spinon confinement and the Haldane gap}
3152: \label{sec:conf}
3153: 
3154: \subsection{The Affleck--Lieb theorem}
3155: \label{sec:afflecklieb}
3156: 
3157: For the generic SU(2) spin chain, Haldane~\cite{haldane83pla464,
3158:   haldane83prl1153} identified the O(3) nonlinear sigma model as the
3159: effective low energy field theory of SU(2) spin chains, and argued
3160: that chains with integer spin possess a gap in the excitation spectrum,
3161: %due to spinon confinement, 
3162: while a topological term renders
3163: half-integer spin chains gapless~\cite{affleck90proc,Fradkin91}.  The exact
3164: models for SU(2) spin chains we reviewed above, the MG and the AKLT
3165: chain, serve a paradigms to illustrate the general principle.
3166: Unfortunately, the effective field theory description of Haldane
3167: yielding the distinction between gapless half integer spin chains with
3168: deconfined spinons and gapped integer spin chains with confined
3169: spinons cannot be directly generalized to SU($n$) chains, as there is
3170: no direct equivalent of the CP$^1$ representation used in Haldane's
3171: analysis.
3172: 
3173: Nonetheless, there is a rigorous and rather general result for
3174: antiferromagnetic chains of SU($n$) spins transforming under a
3175: representation corresponding to a Young tableau with a number of boxes
3176: not divisible by $n$: Affleck and Lieb~\cite{affleck-86lmp57} showed
3177: that if the ground state is non-degenerate, and the Hamiltonian
3178: consists nearest-neighbor interactions only, than the gap in the
3179: excitation spectrum vanishes as $1/N$ (where $N$ is the number of
3180: sites) in the thermodynamic limit.  This result is fully consistent
3181: with the picture suggested by the models introduced above.  Like the
3182: MG model, the trimer model and the representation $\bs{6}$ VBS have
3183: degenerate ground states and interactions which extend beyond the
3184: nearest neighbor, which implies that the theorem is not directly
3185: applicable.
3186: 
3187: On physical grounds, however, the statement that a given model is
3188: gapless (\ie the excitation gap vanishes in the thermodynamic limit)
3189: implies that the spinons are deconfined.  The reason is simply that if
3190: there was a confinement force between them, the zero-point energy
3191: associated with the quantum mechanical oscillator of the relative
3192: motion between the spinons would inevitably yield an energy gap.  The
3193: MG, the trimer and the $\bs{6}$ VBS model constitute pedagogically valuable
3194: paradigms of deconfined spinons.  Since the excitations in these
3195: models are literally domain walls between different ground states, no
3196: long-range forces can exist between them.
3197: 
3198: %\subsection{A category of models with spinon confinement}
3199: \subsection{A general criterion for spinon confinement}
3200: \label{sec:generalcriterion}
3201: 
3202: More importantly, however, the exact models we introduced above
3203: provide information about the models of SU($n$) spin chains with
3204: representations corresponding to Young tableaux with a number of boxes
3205: divisible by $n$, \ie models for which the Affleck--Lieb theorem is
3206: not applicable.  We have studied two SU(3) models belonging to this
3207: family 
3208: %, which we refer to as the singlet family since it contains the
3209: %singlet rep.~$\bs{1}$, 
3210: in Sec.~\ref{sec:vbs}, the rep.~$\bs{10}$ VBS
3211: and the rep.~$\bs{8}$ VBS, and found that both have confined
3212: spinons or colorons and hence display a Haldane-type gap in the
3213: spectrum.  
3214: 
3215: In this section, we will argue that 
3216: %\begin{quote}
3217: %\emph{
3218: models of antiferromagnetic chains of SU($n$) spins transforming under
3219: a representation corresponding to a Young tableau consisting of a
3220: number of boxes $\lambda$ divisible by $n$ generally possess a
3221: Haldane-type gap due to spinon confinement forces.
3222: %}
3223: %\end{quote}
3224: 
3225: We should caution immediately that our argument is based on several
3226: assumptions, which we consider reasonable, but which we are ultimately
3227: unable to prove.
3228: 
3229: The first, and also the most crucial, is the assumption that the
3230: question of whether a given model supports free spinon excitations can
3231: be resolved through study of the corresponding VBS state.  This
3232: assumption definitely holds for SU(2) spin chains, where the MG model
3233: for $S=\frac{1}{2}$ indicates that the spinons are free, while the
3234: AKLT model for $S=1$ serves as a paradigm for spinon confinement and
3235: hence the Haldane gap.  The general conclusions we derived from our
3236: studies of the SU(3) VBSs above rely on this assumption.  The
3237: numerical results we reported on the rep.~$\bs{8}$ VBS provide
3238: evidence that this assumption holds at least for this model.
3239: 
3240: Let us consider an SU($n$) spin chain with spins transforming under
3241: a representation corresponding to Young tableau consisting of $L$ columns
3242: with $\kappa_1\le\kappa_2\le\ldots\le\kappa_L<n$ boxes each,
3243: %\pagebreak
3244: \begin{equation}
3245:   \label{eq:rep}
3246:   [\kappa_1,\kappa_2,\ldots,\kappa_L]\;\equiv\;
3247: \setlength{\unitlength}{8pt}
3248: \begin{picture}(9,4)(0,3)
3249: \put(0,0){\line(1,0){1}}
3250: \multiput(0,1)(0,1){2}{\line(1,0){2}}
3251: %\put(2,2.25){\line(1,0){2.75}}
3252: \put(2,2.45){\line(1,0){2.75}}
3253: \put(4.75,3.75){\line(1,0){1.75}}
3254: \put(6.5,5){\line(1,0){2.25}}
3255: \put(7.75,6){\line(1,0){3}}
3256: \put(0,7){\line(1,0){10.75}}
3257: \put(0,6){\line(1,0){2}}
3258: \multiput(0,0)(1,0){2}{\line(0,1){7}}
3259: \put(2,1){\line(0,1){6}}
3260: %\put(4.75,3.75){\line(0,-1){1.5}}
3261: \put(4.75,3.75){\line(0,-1){1.3}}
3262: \put(6.5,5){\line(0,-1){1.25}}
3263: \multiput(9.75,6)(1,0){2}{\line(0,1){1}}
3264: \multiput(7.75,5)(1,0){2}{\line(0,1){2}}
3265: \put(0.55,-1){\makebox(0,0){\small $\kappa_L$}}
3266: %\put(9.35,4){\makebox(0,0){\small $\kappa_3\kappa_2\kappa_1$}}
3267: \put(8.25,4){\makebox(0,0){\small $\kappa_3$}}
3268: \put(10,4.3){\makebox(0,0){\small $\kappa_2\kappa_1$}}
3269: %\put(13,-0.8){\makebox(0,0){\small boxes}}
3270: \end{picture} \\[27pt] \phantom{oooops} 
3271: \end{equation}
3272: with a total number of boxes
3273: \begin{displaymath}
3274:   \lambda=\sum_{l=1}^L \kappa_l
3275: \end{displaymath}
3276: divisible by $n$.  We denote the Dynkin coordinates of this
3277: representation by $(\mu_1,\mu_2,\ldots,\mu_{n-1})$, which implies
3278: \begin{displaymath}
3279:   \sum_{i=1}^{n-1}\mu_i=L.
3280: \end{displaymath}
3281: Note that this representation is, by definition, given by the 
3282: maximally symmetric component of the tensor product of the individual
3283: columns,
3284: \begin{equation}
3285:   \label{eq:sym}
3286:   [\kappa_1,\kappa_2,\ldots,\kappa_L]=\mathcal{S}
3287:   \big([\kappa_1]\otimes [\kappa_2]\otimes\ldots\otimes [\kappa_L ]\big).
3288: \end{equation}
3289: For convenience, we denote the $\left({n \atop \kappa}\right)$
3290: dimensional representation $[\kappa]$ in this subsection as
3291: $\bs{\kappa}_l\equiv [\kappa_l]=
3292: \setlength{\unitlength}{1.25pt}
3293: \begin{picture}(10,3)(0,-2)
3294: \put(5,0){\circle{8}}
3295: \put(5,0){\makebox(0,0){\small $l$}}
3296: \end{picture}$.
3297: 
3298: Since $\lambda$ is divisible by $n$, it will always be possible to
3299: obtain a singlet from the complete sequence of representations
3300: $\bs{\kappa}_1,\bs{\kappa}_2,\ldots,\bs{\kappa}_L$ by combining them
3301: antisymmetrically.  To be precise, when we write that we combine
3302: representations $\bs{\kappa_1}$ and $\bs{\kappa_2}$ antisymmetrically, we
3303: mean we obtain a new representation $[\kappa_1+\kappa_2]$ by
3304: stacking the two columns with $\kappa_1$ and $\kappa_2$ boxes on
3305: top of each other if $\kappa_1+\kappa_2<n$, and a new representation 
3306: $[\kappa_1+\kappa_2-n]$ if $\kappa_1+\kappa_2\ge n$.  In equations,
3307: we write this as
3308: \begin{equation}
3309:   \label{eq:asym}
3310:   \mathcal{A}
3311:   \big([\kappa_1]\otimes [\kappa_2]\big)\equiv
3312:   \begin{cases}
3313:     [\kappa_1+\kappa_2]&\text{for}\ \kappa_1+\kappa_2<n\\
3314:     [\kappa_1+\kappa_2-n]&\text{for}\ \kappa_1+\kappa_2\ge n
3315:   \end{cases}
3316: \end{equation}
3317: Following the notation used above, we indicate the
3318: antisymmetric combination of representations $\bs{\kappa}_l$ by a line
3319: connecting them.  In particular, we depict the singlet formed by
3320: combining $\bs{\kappa}_1,\bs{\kappa}_2,\ldots,\bs{\kappa}_L$ on
3321: different lattice sites as
3322: \begin{equation}
3323: \label{eq:kappasinglet}
3324:   \setlength{\unitlength}{1.25pt}
3325: \begin{picture}(90,0)(0,-1)
3326: \linethickness{0.8pt}
3327: \multiput(5,0)(20,0){3}{{\circle{8}}}
3328: \multiput(85,0)(20,0){1}{{\circle{8}}}
3329: \put(5,0){\makebox(0,0){\small 1}}
3330: \put(25,0){\makebox(0,0){\small 2}}
3331: \put(45,0){\makebox(0,0){\small 3}}
3332: \put(65,0){\makebox(0,0){\small \ldots}}
3333: \put(85,0){\makebox(0,0){\small $L$}}
3334: %
3335: \thicklines
3336: \multiput(15,0)(20,0){2}{\makebox(0,0){\line(2,0){12}}}
3337: \put(49,0){\line(1,0){8}}
3338: \put(81,0){\line(-1,0){8}}
3339: %
3340: \end{picture}
3341: \vspace{10pt}
3342: \end{equation}
3343: The understanding here is that we combine them in the order indicated
3344: by the line, \ie in \eqref{eq:kappasinglet} we first combine
3345: $\bs{\kappa}_1$ and $\bs{\kappa}_2$, then we combine the result with
3346: $\bs{\kappa}_3$, and so on.  Depending on the order of the
3347: representations $\bs{\kappa}_l$ on the line, we obtain different, but
3348: not necessarily orthogonal, singlets.  We assume that it is irrelevant
3349: whether we combine the representations starting from the left or from
3350: the right of the line, as the resulting state will not depend on it.
3351: 
3352: In general, it will be possible to construct a number $\le\lambda/n$
3353: of singlets out of various combinations of the $\bs{\kappa}$'s, one
3354: for each block of $\bs{\kappa}$'s for which the values of $\kappa$ add
3355: up to $n$ as we combine the representation in the order described above.  
3356: In this case, we will be able to construct one VBS for
3357: each singlet, and subsequently combine them at each site symmetrically
3358: to obtain the desired representation \eqref{eq:rep}.  The argument
3359: for spinon confinement we construct below will hold for each of the
3360: individual VBSs, and hence for the combined VBS as well.  It is hence
3361: sufficient for our purposes to consider situations where the entire
3362: sequence $\bs{\kappa}_1,\bs{\kappa}_2,\ldots,\bs{\kappa}_L$
3363: is needed to construct a singlet.
3364: 
3365: A possible VBS ``ground state'' for a representation corresponding to a
3366: Young tableau with $L=4$ columns is depicted below.
3367: \begin{equation}
3368:   \setlength{\unitlength}{1.25pt}
3369: %  \setlength{\unitlength}{1.3pt}
3370: \begin{picture}(160,70)(0,-15)
3371: \linethickness{0.8pt}
3372: \multiput(0,40)(20,0){9}{{\circle{7}}}
3373: \multiput(0,40)(20,0){9}{\makebox(0,0){\small 1}}
3374: \multiput(0,30)(20,0){9}{{\circle{7}}}
3375: \multiput(0,30)(20,0){9}{\makebox(0,0){\small 2}}
3376: \multiput(0,20)(20,0){9}{{\circle{7}}}
3377: \multiput(0,20)(20,0){9}{\makebox(0,0){\small 3}}
3378: \multiput(0,10)(20,0){9}{{\circle{7}}}
3379: \multiput(0,10)(20,0){9}{\makebox(0,0){\small 4}}
3380: %
3381: \thicklines
3382: \multiput(3.5,38.25)(20,0){8}{\line(2,-1){13}}
3383: \multiput(3.5,28.25)(20,0){8}{\line(2,-1){13}}
3384: \multiput(3.5,18.25)(20,0){8}{\line(2,-1){13}}
3385: %
3386: \thinlines
3387: \put(54,4){\framebox(12,42)}
3388: \put(60,4){\line(0,-1){7}}
3389: \put(60,-10){\makebox(1,1){\small projection onto representation}}
3390: %\put(60,-18){\makebox(1,1){\small $[\kappa_1,\kappa_2,\ldots,\kappa_L]$}}
3391: \put(60,-18){\makebox(1,1){\small $[\kappa_1,\kappa_2,\kappa_3,\kappa_4]$}}
3392: \put(60,55){\makebox(1,1){\small one site}}
3393: \end{picture}
3394: %\vspace{10pt}
3395: %\nonumber
3396: \label{eq:VBS1}
3397: \end{equation}
3398: In general, there are as many inequivalent VBS ``ground states'' as
3399: there are inequivalent ways to order the representations
3400: $\bs{\kappa}_1,\bs{\kappa}_2,\ldots,\bs{\kappa}_L$, \ie
3401: the number of inequivalent VBS ``ground states'' is given by
3402: \begin{displaymath}
3403:   \frac{L!}{\mu_1!\cdot\mu_2!\cdot\ldots\cdot\mu_{n-1}!}.
3404: \end{displaymath}
3405: To give an example, the following VBS
3406: \begin{equation}
3407:   \setlength{\unitlength}{1.25pt}
3408: %  \setlength{\unitlength}{1.3pt}
3409: \begin{picture}(160,70)(0,-15)
3410: \linethickness{0.8pt}
3411: \multiput(0,40)(20,0){9}{{\circle{7}}}
3412: \multiput(0,40)(20,0){9}{\makebox(0,0){\small 2}}
3413: \multiput(0,30)(20,0){9}{{\circle{7}}}
3414: \multiput(0,30)(20,0){9}{\makebox(0,0){\small 1}}
3415: \multiput(0,20)(20,0){9}{{\circle{7}}}
3416: \multiput(0,20)(20,0){9}{\makebox(0,0){\small 3}}
3417: \multiput(0,10)(20,0){9}{{\circle{7}}}
3418: \multiput(0,10)(20,0){9}{\makebox(0,0){\small 4}}
3419: %
3420: \thicklines
3421: \multiput(3.5,38.25)(20,0){8}{\line(2,-1){13}}
3422: \multiput(3.5,28.25)(20,0){8}{\line(2,-1){13}}
3423: \multiput(3.5,18.25)(20,0){8}{\line(2,-1){13}}
3424: %
3425: \thinlines
3426: \put(54,4){\framebox(12,42)}
3427: \put(60,4){\line(0,-1){7}}
3428: \put(60,-10){\makebox(1,1){\small projection onto representation}}
3429: %\put(60,-18){\makebox(1,1){\small $[\kappa_1,\kappa_2,\ldots,\kappa_L]$}}
3430: \put(60,-18){\makebox(1,1){\small $[\kappa_1,\kappa_2,\kappa_3,\kappa_4]$}}
3431: \put(60,55){\makebox(1,1){\small one site}}
3432: \end{picture}
3433: %\vspace{10pt}
3434: %\nonumber
3435: \label{eq:VBS2}
3436: \end{equation}
3437: is inequivalent to the one above if and only if $\kappa_1\ne\kappa_2$.
3438: Note that all these VBS ``ground states'' states are translationally
3439: invariant.  We expect some of these states, but not all of them,
3440: degenerate in energy for the appropriate parent Hamiltonian, and have
3441: accordingly written ``ground states'' in quotation marks.  For example,
3442: if we form a SU(4) VBS with representation 
3443: $[\kappa_1,\kappa_2,\kappa_3]=[1,1,2]$, the
3444: combination 
3445: \begin{equation}
3446:   \setlength{\unitlength}{1.25pt}
3447: \begin{picture}(50,0)(0,-1)
3448: \linethickness{0.8pt}
3449: \multiput(5,0)(20,0){3}{{\circle{8}}}
3450: \put(5,0){\makebox(0,0){\small 1}}
3451: \put(25,0){\makebox(0,0){\small 3}}
3452: \put(45,0){\makebox(0,0){\small 2}}
3453: %
3454: \thicklines
3455: \multiput(15,0)(20,0){2}{\makebox(0,0){\line(2,0){12}}}
3456: \end{picture}\nonumber
3457: \vspace{4pt}
3458: \end{equation}
3459: might yield a state with a lower energy for the appropriate
3460: Hamiltonian than the state formed by combining
3461: \begin{equation}
3462:   \setlength{\unitlength}{1.25pt}
3463: \begin{picture}(50,0)(0,-1)
3464: \linethickness{0.8pt}
3465: \multiput(5,0)(20,0){3}{{\circle{8}}}
3466: \put(5,0){\makebox(0,0){\small 1}}
3467: \put(25,0){\makebox(0,0){\small 2}}
3468: \put(45,0){\makebox(0,0){\small 3}}
3469: %
3470: \thicklines
3471: \multiput(15,0)(20,0){2}{\makebox(0,0){\line(2,0){12}}}
3472: \end{picture}. \nonumber
3473: \vspace{4pt}
3474: \end{equation}
3475: Simple examples where we have only two inequivalent VBS ground states
3476: are provided by the matrix product states discussed in
3477: Secs.~\ref{sec:8VBS} and \ref{sec:mp}.
3478: 
3479: We will now argue that the elementary excitations of the corresponding
3480: VBS models are always confined.  To this end, we first observe that
3481: any domain wall between translationally invariant ``ground states''
3482: consists of a total of $m\cdot n$ representations $\bs{\kappa}_l$ ($m$
3483: integer).  To illustrate this, consider a domain wall between the
3484: ``ground states'' depicted in \eqref{eq:VBS2} and \eqref{eq:VBS1}:
3485: \begin{equation}
3486:   \setlength{\unitlength}{1.25pt}
3487: %  \setlength{\unitlength}{1.3pt}
3488: \begin{picture}(170,49)(20,-2)
3489: \linethickness{0.8pt}
3490: \multiput(20,40)(20,0){2}{{\circle{7}}}
3491: \multiput(20,40)(20,0){2}{\makebox(0,0){\small 2}}
3492: \multiput(80,40)(20,0){6}{{\circle{7}}}
3493: \multiput(80,40)(20,0){6}{\makebox(0,0){\small 1}}
3494: \multiput(20,30)(20,0){3}{{\circle{7}}}
3495: \multiput(20,30)(20,0){3}{\makebox(0,0){\small 1}}
3496: \multiput(100,30)(20,0){5}{{\circle{7}}}
3497: \multiput(100,30)(20,0){5}{\makebox(0,0){\small 2}}
3498: \multiput(20,20)(20,0){4}{{\circle{7}}}
3499: \multiput(20,20)(20,0){4}{\makebox(0,0){\small 3}}
3500: \multiput(120,20)(20,0){4}{{\circle{7}}}
3501: \multiput(120,20)(20,0){4}{\makebox(0,0){\small 3}}
3502: \multiput(20,10)(20,0){5}{{\circle{7}}}
3503: \multiput(20,10)(20,0){5}{\makebox(0,0){\small 4}}
3504: \multiput(140,10)(20,0){3}{{\circle{7}}}
3505: \multiput(140,10)(20,0){3}{\makebox(0,0){\small 4}}
3506: %
3507: %{\color{blue}
3508: \put(60,40){{\circle*{7}}}
3509: \put(80,30){{\circle*{7}}}
3510: \put(100,20){{\circle*{7}}}
3511: \put(120,10){{\circle*{7}}}
3512: %}
3513: {\color{white}
3514: \put(60,40){\makebox(0,0){\small\bf 2}}
3515: \put(80,30){\makebox(0,0){\small\bf 2}}
3516: \put(100,20){\makebox(0,0){\small\bf 3}}
3517: \put(120,10){\makebox(0,0){\small\bf 4}}
3518: }
3519: % 
3520: \thicklines
3521: \multiput(23.5,38.25)(20,0){2}{\line(2,-1){13}}
3522: \multiput(23.5,28.25)(20,0){2}{\line(2,-1){13}}
3523: \multiput(23.5,18.25)(20,0){2}{\line(2,-1){13}}
3524: %
3525: %{\color{red}
3526: \put(63.5,28.25){\line(2,-1){13}}
3527: \put(63.5,18.25){\line(2,-1){13}}
3528: %}
3529: \multiput(83.5,38.25)(20,0){5}{\line(2,-1){13}}
3530: \multiput(83.5,28.25)(20,0){5}{\line(2,-1){13}}
3531: \multiput(83.5,18.25)(20,0){5}{\line(2,-1){13}}
3532: \end{picture}
3533: %\vspace{10pt}
3534: \nonumber
3535: \end{equation}
3536: In the example, the domain wall consists of reps.~$\bs{\kappa}_2$,
3537: $\bs{\kappa}_2$, $\bs{\kappa}_3$, and $\bs{\kappa}_4$.  If the
3538: translationally invariant states on both sites are true ground states,
3539: the domain wall is likely to correspond to two elementary excitations:
3540: a rep.~$\bs{\bar{\kappa}}_1$ spinon consisting of an antisymmetric
3541: combination of a $\bs{\kappa}_2$, a $\bs{\kappa}_3$, and a
3542: $\bs{\kappa}_4$, as indicated by the line in the drawing, and another
3543: rep.~$\bs{\kappa}_2$ spinon.  The reason we assume that
3544: $\bs{\kappa}_2$, $\bs{\kappa}_3$, and $\bs{\kappa}_4$ form a
3545: rep.~$\bs{\bar{\kappa}}_1$ is simply that this combination is present
3546: in both ground states on either side, and hence bound to be the
3547: energetically most favorable combination.  The second
3548: $\bs{\kappa}_2$, however, is not part of this elementary excitation,
3549: as combining it antisymmetrically with the others (\ie the
3550: $\bs{\bar{\kappa}}_1$) would induce correlations which are not present
3551: in the ground state.  We hence conclude that the second
3552: $\bs{\kappa}_2$ is an elementary excitation as well.  The domain wall
3553: depicted above consists of a spinon transforming under
3554: rep.~$\bs{\bar{\kappa}}_1$ and an anti-spinon transforming under
3555: rep.~$\bs{\kappa}_2$.
3556: 
3557: The next step in our argument is to notice that the spinon and the
3558: complementary particle created simultaneously which may either be its
3559: anti-particle or some other spinon, are confined through a linear
3560: potential.  To see this, we pull them apart and look at the state
3561: inbetween (color online):
3562: \begin{equation}
3563:   \setlength{\unitlength}{1.25pt}
3564: %  \setlength{\unitlength}{1.3pt}
3565: \begin{picture}(180,73)(0,-10)
3566: \linethickness{0.8pt}
3567: \multiput(0,40)(20,0){1}{{\circle{7}}}
3568: \multiput(0,40)(20,0){1}{\makebox(0,0){\small 2}}
3569: \multiput(40,40)(20,0){8}{{\circle{7}}}
3570: \multiput(40,40)(20,0){8}{\makebox(0,0){\small 1}}
3571: \multiput(0,30)(20,0){2}{{\circle{7}}}
3572: \multiput(0,30)(20,0){2}{\makebox(0,0){\small 1}}
3573: \multiput(40,30)(20,0){4}{{\circle{7}}}
3574: \multiput(40,30)(20,0){4}{\makebox(0,0){\small 2}}
3575: \multiput(140,30)(20,0){3}{{\circle{7}}}
3576: \multiput(140,30)(20,0){3}{\makebox(0,0){\small 2}}
3577: \multiput(0,20)(20,0){7}{{\circle{7}}}
3578: \multiput(0,20)(20,0){7}{\makebox(0,0){\small 3}}
3579: \multiput(160,20)(20,0){2}{{\circle{7}}}
3580: \multiput(160,20)(20,0){2}{\makebox(0,0){\small 3}}
3581: \multiput(0,10)(20,0){8}{{\circle{7}}}
3582: \multiput(0,10)(20,0){8}{\makebox(0,0){\small 4}}
3583: \multiput(180,10)(20,0){1}{{\circle{7}}}
3584: \multiput(180,10)(20,0){1}{\makebox(0,0){\small 4}}
3585: %
3586: %{\color{blue}
3587: \put(20,40){{\circle*{7}}}
3588: \put(120,30){{\circle*{7}}}
3589: \put(140,20){{\circle*{7}}}
3590: \put(160,10){{\circle*{7}}}
3591: %}
3592: {\color{white}
3593: \put(20,40){\makebox(0,0){\small\bf 2}}
3594: \put(120,30){\makebox(0,0){\small\bf 2}}
3595: \put(140,20){\makebox(0,0){\small\bf 3}}
3596: \put(160,10){\makebox(0,0){\small\bf 4}}
3597: }
3598: % 
3599: %{\color{blue}
3600: \thicklines
3601: \put(123.5,28.25){\line(2,-1){13}}
3602: \put(143.5,18.25){\line(2,-1){13}}
3603: %}
3604: \thicklines
3605: \multiput(3.5,38.25)(20,0){1}{\line(2,-1){13}}
3606: \multiput(3.5,28.25)(20,0){2}{\line(2,-1){13}}
3607: \multiput(3.5,18.25)(20,0){2}{\line(2,-1){13}}
3608: %
3609: {\color{red}
3610: \multiput(43.5,38.25)(20,0){3}{\line(2,-1){13}}
3611: \multiput(43.5,28.25)(20,0){4}{\line(2,-1){13}}
3612: \multiput(43.5,18.25)(20,0){4}{\line(2,-1){13}}
3613: %
3614: % TWO ALTERNTIVES
3615: %
3616: \put(90,25){\makebox(0,0){\line(2,3){15.9}}}
3617: %\qbezier(43.5,31.75)(50,35)(70,35)
3618: %\qbezier(96.4,38.25)(90,35)(70,35)
3619: }
3620: %
3621: \multiput(123.5,38.25)(20,0){3}{\line(2,-1){13}}
3622: \multiput(123.5,28.25)(20,0){3}{\line(2,-1){13}}
3623: \multiput(123.5,18.25)(20,0){3}{\line(2,-1){13}}
3624: %
3625: \thinlines
3626: \put(20,56){\makebox(0,0){\small $\bs{\kappa}_2$-spinon}}
3627: \put(140,56){\makebox(0,0){\small $\bs{\bar{\kappa}}_1$-spinon}}
3628: %\put(20,56){\makebox(0,0){\small $[\kappa_2]$-spinon}}
3629: %\put(140,56){\makebox(0,0){\small $[n\!-\!\kappa_1]$-spinon}}
3630: \put(80,-2){\vector(-1,0){60}}
3631: \put(80,-2){\vector(1,0){60}}
3632: \put(80,-8){\makebox(0,0){\small energy cost $\propto$ distance}}
3633: \end{picture}
3634: %\vspace{10pt}
3635: \label{eq:kappaconf}
3636: \end{equation}
3637: The state between spinon and anti-spinon is not translationally
3638: invariant.  In the example, the unit cell of this state is depicted in
3639: red and consists of two regular bonds with three ``singlet lines''
3640: between the sites, one stronger bond with four lines (which cross in
3641: the cartoon), and one weaker bond with only two lines.  If we assume
3642: that the two states \eqref{eq:VBS2} and \eqref{eq:VBS1} on both sides
3643: are true ground states, it is clear that the irregularities in the
3644: strength of the bond correlations will cause the state between the
3645: spinon and anti-spinon to have a higher energy than either of them.
3646: This additional energy cost will induce a linear confinement potential
3647: between the spinons, and hence a linear oscillator potential for the
3648: relative motion of the particles.  As in the models studied above, the
3649: Haldane gap corresponds to the zero-point energy of this oscillator.
3650: 
3651: If one of the ``ground states'' to the left or to the right of the
3652: domain wall is not a true ground state, but a translationally
3653: invariant state corresponding to a higher energy than the ground
3654: state, there will be a confining force between this domain wall and
3655: another domain wall which transforms the intermediate ``ground state''
3656: with a higher energy back into a true ground state.  This force will
3657: be sufficient to account for a Haldane gap, regardless of the strength
3658: or existence of a confinement force between the constituent particles
3659: of each domain wall.
3660: 
3661: The lowest-lying excitations of a representation
3662: $[\kappa_1,\kappa_2,\ldots,\kappa_L]$ spin chain, however, will in
3663: general not be domain walls, but spinons created by breaking up one of
3664: the singlets \eqref{eq:kappasinglet} in a ground state.  This is 
3665: illustrated below for the ground state \eqref{eq:VBS1}:
3666: \begin{equation}
3667:   \setlength{\unitlength}{1.25pt}
3668: %  \setlength{\unitlength}{1.3pt}
3669: \begin{picture}(170,49)(20,-2)
3670: \linethickness{0.8pt}
3671: \multiput(20,40)(20,0){2}{{\circle{7}}}
3672: \multiput(20,40)(20,0){2}{\makebox(0,0){\small 1}}
3673: \multiput(80,40)(20,0){6}{{\circle{7}}}
3674: \multiput(80,40)(20,0){6}{\makebox(0,0){\small 1}}
3675: \multiput(20,30)(20,0){3}{{\circle{7}}}
3676: \multiput(20,30)(20,0){3}{\makebox(0,0){\small 2}}
3677: \multiput(100,30)(20,0){5}{{\circle{7}}}
3678: \multiput(100,30)(20,0){5}{\makebox(0,0){\small 2}}
3679: \multiput(20,20)(20,0){4}{{\circle{7}}}
3680: \multiput(20,20)(20,0){4}{\makebox(0,0){\small 3}}
3681: \multiput(120,20)(20,0){4}{{\circle{7}}}
3682: \multiput(120,20)(20,0){4}{\makebox(0,0){\small 3}}
3683: \multiput(20,10)(20,0){5}{{\circle{7}}}
3684: \multiput(20,10)(20,0){5}{\makebox(0,0){\small 4}}
3685: \multiput(140,10)(20,0){3}{{\circle{7}}}
3686: \multiput(140,10)(20,0){3}{\makebox(0,0){\small 4}}
3687: %
3688: %{\color{blue}
3689: \put(60,40){{\circle*{7}}}
3690: \put(80,30){{\circle*{7}}}
3691: \put(100,20){{\circle*{7}}}
3692: \put(120,10){{\circle*{7}}}
3693: %}
3694: {\color{white}
3695: \put(60,40){\makebox(0,0){\small\bf 1}}
3696: \put(80,30){\makebox(0,0){\small\bf 2}}
3697: \put(100,20){\makebox(0,0){\small\bf 3}}
3698: \put(120,10){\makebox(0,0){\small\bf 4}}
3699: }
3700: % 
3701: \thicklines
3702: \multiput(23.5,38.25)(20,0){2}{\line(2,-1){13}}
3703: \multiput(23.5,28.25)(20,0){2}{\line(2,-1){13}}
3704: \multiput(23.5,18.25)(20,0){2}{\line(2,-1){13}}
3705: %
3706: %{\color{red}
3707: \put(63.5,28.25){\line(2,-1){13}}
3708: \put(63.5,18.25){\line(2,-1){13}}
3709: %}
3710: \multiput(83.5,38.25)(20,0){5}{\line(2,-1){13}}
3711: \multiput(83.5,28.25)(20,0){5}{\line(2,-1){13}}
3712: \multiput(83.5,18.25)(20,0){5}{\line(2,-1){13}}
3713: \end{picture}
3714: %\vspace{10pt}
3715: \nonumber
3716: \end{equation}
3717: In the example, we have created a spinon transforming under
3718: rep.~$\bs{\bar{\kappa}}_1$ and its anti-particle, a spinon
3719: transforming under rep.~$\bs{\kappa}_1$.  This is, however, irrelevant
3720: to the argument---we may break the singlet in any way we like.
3721: The important feature is that we obtain, by construction, at least two
3722: excitations, and that these are confined.  In our specific example, the 
3723: confining potential is equivalent to the confining potential in
3724: \eqref{eq:kappaconf} above (color online):
3725: \begin{equation}
3726:   \setlength{\unitlength}{1.25pt}
3727: %  \setlength{\unitlength}{1.3pt}
3728: \begin{picture}(180,78)(0,-15)
3729: \linethickness{0.8pt}
3730: \multiput(0,40)(20,0){1}{{\circle{7}}}
3731: \multiput(0,40)(20,0){1}{\makebox(0,0){\small 1}}
3732: \multiput(40,40)(20,0){8}{{\circle{7}}}
3733: \multiput(40,40)(20,0){8}{\makebox(0,0){\small 1}}
3734: \multiput(0,30)(20,0){2}{{\circle{7}}}
3735: \multiput(0,30)(20,0){2}{\makebox(0,0){\small 2}}
3736: \multiput(40,30)(20,0){4}{{\circle{7}}}
3737: \multiput(40,30)(20,0){4}{\makebox(0,0){\small 2}}
3738: \multiput(140,30)(20,0){3}{{\circle{7}}}
3739: \multiput(140,30)(20,0){3}{\makebox(0,0){\small 2}}
3740: \multiput(0,20)(20,0){7}{{\circle{7}}}
3741: \multiput(0,20)(20,0){7}{\makebox(0,0){\small 3}}
3742: \multiput(160,20)(20,0){2}{{\circle{7}}}
3743: \multiput(160,20)(20,0){2}{\makebox(0,0){\small 3}}
3744: \multiput(0,10)(20,0){8}{{\circle{7}}}
3745: \multiput(0,10)(20,0){8}{\makebox(0,0){\small 4}}
3746: \multiput(180,10)(20,0){1}{{\circle{7}}}
3747: \multiput(180,10)(20,0){1}{\makebox(0,0){\small 4}}
3748: %
3749: %{\color{blue}
3750: \put(20,40){{\circle*{7}}}
3751: \put(120,30){{\circle*{7}}}
3752: \put(140,20){{\circle*{7}}}
3753: \put(160,10){{\circle*{7}}}
3754: %}
3755: {\color{white}
3756: \put(20,40){\makebox(0,0){\small\bf 1}}
3757: \put(120,30){\makebox(0,0){\small\bf 2}}
3758: \put(140,20){\makebox(0,0){\small\bf 3}}
3759: \put(160,10){\makebox(0,0){\small\bf 4}}
3760: }
3761: % 
3762: %{\color{blue}
3763: \thicklines
3764: \put(123.5,28.25){\line(2,-1){13}}
3765: \put(143.5,18.25){\line(2,-1){13}}
3766: %}
3767: \thicklines
3768: \multiput(3.5,38.25)(20,0){1}{\line(2,-1){13}}
3769: \multiput(3.5,28.25)(20,0){2}{\line(2,-1){13}}
3770: \multiput(3.5,18.25)(20,0){2}{\line(2,-1){13}}
3771: %
3772: {\color{red}
3773: \multiput(43.5,38.25)(20,0){3}{\line(2,-1){13}}
3774: \multiput(43.5,28.25)(20,0){4}{\line(2,-1){13}}
3775: \multiput(43.5,18.25)(20,0){4}{\line(2,-1){13}}
3776: %
3777: % TWO ALTERNTIVES
3778: %
3779: \put(90,25){\makebox(0,0){\line(2,3){15.9}}}
3780: %\qbezier(43.5,31.75)(50,35)(70,35)
3781: %\qbezier(96.4,38.25)(90,35)(70,35)
3782: }
3783: %
3784: \multiput(123.5,38.25)(20,0){3}{\line(2,-1){13}}
3785: \multiput(123.5,28.25)(20,0){3}{\line(2,-1){13}}
3786: \multiput(123.5,18.25)(20,0){3}{\line(2,-1){13}}
3787: %
3788: \thinlines
3789: \put(20,56){\makebox(0,0){\small $\bs{\kappa}_1$-spinon}}
3790: \put(140,56){\makebox(0,0){\small $\bs{\bar{\kappa}}_1$-spinon}}
3791: \put(80,-2){\vector(-1,0){60}}
3792: \put(80,-2){\vector(1,0){60}}
3793: \put(80,-8){\makebox(0,0){\small energy cost $\propto$ distance}}
3794: \end{picture}
3795: %\vspace{10pt}
3796: \nonumber
3797: \end{equation}
3798: We leave it to the reader to convince him- or herself that the
3799: conclusions regarding confinement drawn from the simple examples
3800: studied here hold in general.
3801: 
3802: \subsection{Models with confinement through interactions 
3803: extending beyond nearest neighbors }
3804: \label{sec:examples}
3805: 
3806: Let us briefly summarize the results obtained.  The SU($n$) models we
3807: have studied so far fall into two categories.  The models belonging to
3808: the first---the trimer chain, the $\bs{6}$ VBS, and the $n$-mer
3809: chain---have $n$ degenerate ground states, which break translational
3810: invariance up to translations by $n$ lattice spacings.  The Young
3811: tableaux describing the representations of SU($n$) at each site
3812: consist of a number of boxes $\lambda$ which is smaller than $n$, and
3813: hence obviously not divisible by $n$.  The models support deconfined
3814: spinon excitations, and hence do not possess a Haldane gap in the
3815: spectrum.  The Hamiltonians of these models require interactions
3816: between $n+1$ neighboring sites along the chain.  Even though the
3817: Affleck--Lieb theorem is not directly applicable to the models we
3818: constructed above, it is applicable to SU($n$) spin chains with spins
3819: transforming under the same representations.  Like the VBS models, the
3820: theorem suggests that there is no Haldane gap in this family of
3821: models.
3822: 
3823: 
3824: The models belonging to the second category---the $\bs{10}$ VBS, the
3825: $\bs{8}$ VBS, the representation $(n,0,\ldots,0)$ VBS, and the
3826: $\bs{m}$-$\bs{\overline{m}}$ matrix product state---have
3827: translationally invariant ground states.  The ground states are unique
3828: for some models, like the $\bs{10}$ and the $(n,0,\ldots,0)$ VBS, and
3829: degenerate for others, like the $\bs{8}$ VBS.  The Young tableaux
3830: describing the representations of SU($n$) at each site consist of a
3831: number of boxes $\lambda$ which is divisible by $n$.  The
3832: Affleck--Lieb theorem is not applicable to models of this category.
3833: The spinon excitations for this category of models are subject to
3834: confinement forces, which give rise to a Haldane gap.  The parent
3835: Hamiltonians for these models require interactions between
3836: nearest-neighbor sites only.
3837: 
3838: 
3839: At first glance, this classification might appear complete.  Further
3840: possibilities arise, however, in SU($n$) spin chains where number of
3841: boxes $\lambda$ the Young tableau consists of and $n$ have a common
3842: divisor different from $n$, which obviously requires that $n$ is not a
3843: prime number.  In this case, it is possible to construct VBS models in
3844: which the ground state breaks translational invariance only up to
3845: translations by $n/q$ lattice spacings, where $q$ denotes the largest
3846: common divisor of $\lambda$ and $n$ such that $q<n$.  This implies
3847: that the ground state of the appropriate, translationally invariant
3848: Hamiltonian will be $n/q$-fold degenerate.  In the examples we will
3849: elaborate on below, the parent Hamiltonians for these models require
3850: interactions between $\frac{n}{q}+1$ sites, a feature we conjecture to
3851: hold in general.  The spinon excitations of these models are confined,
3852: even though the Affleck--Lieb theorem states that the nearest-neighbor
3853: Heisenberg chain of SU($n$) spins transforming under these
3854: representations is gapless.  (We implicitly assume here that the 
3855: ground states of the SU($n$) nearest-neighbor Heisenberg chains are
3856: non-degenerate.)  Let us illustrate the general features of
3857: this third category of models with a few simple examples.
3858: 
3859: 
3860: (1) Consider an SU(4) chain with spins transforming under the
3861: 10-dimensional representation
3862: \begin{displaymath}
3863:   (2,0,0)\ =\
3864:   \setlength{\unitlength}{8pt}
3865:   \begin{picture}(2,1)(0,0.15)
3866:     \multiput(0,0)(1,0){3}{\line(0,1){1}}
3867:     \multiput(0,0)(0,1){2}{\line(1,0){2}}
3868:   \end{picture}\ \ .
3869: \end{displaymath}
3870: Following the construction principle of the $\bs{6}$ VBS of SU(3), we
3871: find that the two degenerate VBS states illustrated through
3872: \begin{equation}
3873: \setlength{\unitlength}{1pt}
3874: \begin{picture}(162,48)(4,-22)
3875: \linethickness{0.8pt}
3876: \multiput(0,10)(14,0){12}{\circle{4}}
3877: \multiput(0,0)(14,0){12}{\circle{4}}
3878: \thicklines
3879: \multiput(2,10)(14,0){3}{\line(1,0){10}}
3880: \multiput(58,10)(14,0){3}{\line(1,0){10}}
3881: \multiput(114,10)(14,0){3}{\line(1,0){10}}
3882: \put(-2,0){\line(-1,0){4}}
3883: \multiput(2,0)(14,0){1}{\line(1,0){10}}
3884: \multiput(30,0)(14,0){3}{\line(1,0){10}}
3885: \multiput(86,0)(14,0){3}{\line(1,0){10}}
3886: \multiput(142,0)(14,0){1}{\line(1,0){10}}
3887: \put(156,0){\line(1,0){4}}
3888: \thinlines
3889: %\put(37,-7){\framebox(10,24)}
3890: \put(37,-5){\framebox(10,20)}
3891: \put(107,-5){\dashbox{2}(38,20)}
3892: \put(42,-5){\line(0,-1){8}}
3893: \put(42,-20){\makebox(1,1){\small projection onto rep.\ $(2,0,0)$}}
3894: \put(42,23){\makebox(1,1){\small one site}}
3895: \end{picture}
3896: \label{eq:200state}
3897: \end{equation}
3898: are exact zero-energy ground states of 
3899: \begin{equation}
3900:   \label{eq:SU4model}
3901: %  H_{\text{SU}(4)}^{\text{10-VBS}}
3902:   H_{(2,0,0)\,\text{VBS}}
3903:   =\sum_{i=1}^N
3904:   \left(\Bigl(\bs{J}_i^{(3)}\Bigr)^4-12\Bigl(\bs{J}_i^{(3)}\Bigr)^2+
3905:     \frac{135}{4}\right),
3906: \end{equation}
3907: which contains next-nearest neighbor interactions.  
3908: % Note that \eqref{eq:200state} breaks translational invariance only up
3909: % to translations by two lattice spacings.
3910: 
3911: The example illustrates the general rule stated above. The largest
3912: common divisor of $n=4$ and $\lambda=2$ is $q=2$.  This implies
3913: $\frac{n}{q}=2$ and hence two degenerate VBS states which break
3914: translational invariance up to translations by two lattice spacings.
3915: The parent Hamiltonian requires interaction between three neighboring
3916: sites.
3917: 
3918: According to the Affleck--Lieb theorem, models of antiferromagnetic
3919: SU(4) chains of representation $(2,0,0)$ with nearest-neighbor
3920: Heisenberg interactions and non-degenerate ground states are gapless
3921: in the thermodynamic limit, which implies that the spinons are
3922: deconfined.  In all the models we have studied in previous sections,
3923: the conclusions drawn from the Affleck--Lieb theorem were consistent
3924: with those drawn from our exact models.  For the present model,
3925: however, they are not consistent.
3926: 
3927: Specifically, we strongly conjecture that the spinons in the $(2,0,0)$
3928: VBS are confined.  This conjecture is based on the reasonable
3929: assumption that the elementary excitations of the model transform as
3930: either the fundamental representation $\bs{4}=(1,0,0)$ of SU(4) or its
3931: conjugate representation $\bs{\bar 4}=(0,0,1)$.  This assumption
3932: implies that a single domain wall in one of the $4$-mer chains used to
3933: construct the VBS state \eqref{eq:200state} shifts this chain by one
3934: lattice spacing.  If we assume a ground state to the left of the
3935: spinon, the state on to the right will have a higher energy for the
3936: next-nearest neighbor Hamiltonian \eqref{eq:SU4model}, as illustrated
3937: below.  
3938: \begin{equation}
3939: \setlength{\unitlength}{1pt}
3940: \begin{picture}(170,60)(4,-25)
3941: \linethickness{0.8pt}
3942: \multiput(0,10)(14,0){13}{\circle{4}}
3943: \multiput(0,0)(14,0){13}{\circle{4}}
3944: \multiput(0,10)(14,0){3}{\circle*{4}}
3945: \multiput(140,0)(14,0){3}{\circle*{4}}
3946: \thicklines
3947: \multiput(2,10)(14,0){2}{\line(1,0){10}}
3948: \multiput(44,10)(14,0){3}{\line(1,0){10}}
3949: \multiput(100,10)(14,0){3}{\line(1,0){10}}
3950: \multiput(156,10)(14,0){1}{\line(1,0){10}}
3951: \put(-2,0){\line(-1,0){4}}
3952: \multiput(2,0)(14,0){1}{\line(1,0){10}}
3953: \multiput(30,0)(14,0){3}{\line(1,0){10}}
3954: \multiput(86,0)(14,0){3}{\line(1,0){10}}
3955: \multiput(142,0)(14,0){2}{\line(1,0){10}}
3956: \put(170,10){\line(1,0){4}}
3957: \thinlines
3958: \put(84,-14){\vector(-1,0){70}}
3959: \put(84,-14){\vector(1,0){70}}
3960: \put(84,-22){\makebox(0,0){\small energy cost $\propto$ distance}}
3961: \put(14,22){\makebox(0,0){$\bs{\bar 4}$}}
3962: \put(14,32){\makebox(0,0){\small spinon}}
3963: \put(154,22){\makebox(0,0){$\bs{\bar 4}$}}
3964: \put(154,32){\makebox(0,0){\small spinon}}
3965: \put(65,-5){\dashbox{1}(38,20)}
3966: \end{picture}
3967: \label{eq:200spinons}
3968: \end{equation}
3969: To recover the ground state, a second domain wall is required
3970: nearby, which is bound to the first by a linear potential.
3971: 
3972: Our conclusion is not in contradiction with the rigorous result of
3973: Affleck and Lieb, as they explicitly restrict themselves to models
3974: with nearest-neighbor interactions.  If we had only nearest-neighbor
3975: interactions, the energy expectation value in the region between the
3976: domain walls would not be higher than in the ground state, as one can
3977: see easily from the cartoon above---the sequence of alternating links
3978: would merely shift from (strong, weak, strong, weak) to (strong,
3979: strong, weak, weak).
3980: 
3981: (2) The situation is similar for an SU(6) chain with spins transforming
3982: under the 56-dimensional representation
3983: \begin{displaymath}
3984:   (3,0,0,0,0)\ =\
3985:   \setlength{\unitlength}{8pt}
3986:   \begin{picture}(3,1)(0,0.15)
3987:     \multiput(0,0)(1,0){4}{\line(0,1){1}}
3988:     \multiput(0,0)(0,1){2}{\line(1,0){3}}
3989:   \end{picture}\ \ .
3990: \end{displaymath}
3991: With $n=6$ and $\lambda=3$, we have again $\frac{n}{q}=2$.
3992: Accordingly, we find that the two VBS states illustrated through
3993: \begin{equation}
3994: \setlength{\unitlength}{1pt}
3995: \begin{picture}(162,58)(4,-22)
3996: \linethickness{0.8pt}
3997: \multiput(0,20)(14,0){12}{\circle{4}}
3998: \multiput(0,10)(14,0){12}{\circle{4}}
3999: \multiput(0,0)(14,0){12}{\circle{4}}
4000: \thicklines
4001: \multiput(2,20)(14,0){5}{\line(1,0){10}}
4002: \multiput(86,20)(14,0){5}{\line(1,0){10}}
4003: \put(-2,10){\line(-1,0){4}}
4004: \put(-2,0){\line(-1,0){4}}
4005: \multiput(2,10)(14,0){1}{\line(1,0){10}}
4006: \multiput(30,10)(14,0){5}{\line(1,0){10}}
4007: \multiput(114,10)(14,0){3}{\line(1,0){10}}
4008: \put(156,10){\line(1,0){4}}
4009: \multiput(2,0)(14,0){3}{\line(1,0){10}}
4010: \multiput(58,0)(14,0){5}{\line(1,0){10}}
4011: \multiput(142,0)(14,0){1}{\line(1,0){10}}
4012: \put(156,0){\line(1,0){4}}
4013: \thinlines
4014: %\put(37,-7){\framebox(10,24)}
4015: \put(37,-5){\framebox(10,30)}
4016: \put(107,-5){\dashbox{2}(38,30)}
4017: \put(42,-5){\line(0,-1){8}}
4018: \put(42,-20){\makebox(1,1){\small projection onto rep.\ $(3,0,0,0,0)$}}
4019: \put(42,33){\makebox(1,1){\small one site}}
4020: \end{picture}
4021: \label{eq:30000state}
4022: \end{equation}
4023: are exact ground states of a parent Hamiltonian containing up to
4024: next-nearest-neighbor interactions only, and that the spinon excitations
4025: are confined.
4026: 
4027: (3) As a third example, consider an SU(6) spin chain with spins
4028: transforming under the 21-dimensional representation
4029: \begin{displaymath}
4030:   (2,0,0,0,0)\ =\
4031:   \setlength{\unitlength}{8pt}
4032:   \begin{picture}(2,1)(0,0.15)
4033:     \multiput(0,0)(1,0){3}{\line(0,1){1}}
4034:     \multiput(0,0)(0,1){2}{\line(1,0){2}}
4035:   \end{picture}\ \ .
4036: \end{displaymath}
4037: This implies $\frac{n}{q}=3$.  We find that the three VBS states
4038: illustrated by
4039: \begin{equation}
4040: \setlength{\unitlength}{1pt}
4041: \begin{picture}(162,48)(4,-22)
4042: \linethickness{0.8pt}
4043: \multiput(0,10)(14,0){12}{\circle{4}}
4044: \multiput(0,0)(14,0){12}{\circle{4}}
4045: \thicklines
4046: \multiput(2,10)(14,0){5}{\line(1,0){10}}
4047: \multiput(86,10)(14,0){5}{\line(1,0){10}}
4048: \put(-2,0){\line(-1,0){4}}
4049: \multiput(2,0)(14,0){2}{\line(1,0){10}}
4050: \multiput(44,0)(14,0){5}{\line(1,0){10}}
4051: \multiput(128,0)(14,0){2}{\line(1,0){10}}
4052: %\multiput(170,10)(14,0){2}{\line(1,0){10}}
4053: \put(156,0){\line(1,0){4}}
4054: %\put(198,10){\line(1,0){4}}
4055: \thinlines
4056: %\put(37,-7){\framebox(10,24)}
4057: \put(37,-5){\framebox(10,20)}
4058: \put(93,-5){\dashbox{2}(52,20)}
4059: \put(42,-5){\line(0,-1){8}}
4060: \put(42,-20){\makebox(1,1){\small projection onto rep.\ $(2,0,0,0,0)$}}
4061: \put(42,23){\makebox(1,1){\small one site}}
4062: \end{picture}
4063: \label{eq:20000state}
4064: \end{equation}
4065: are exact ground states of a parent Hamiltonian involving the
4066: quadratic Casimir of total spin of four neighboring sites,
4067: \begin{equation}
4068:   \label{eq:ham20000}
4069:   H_{(2,0,0,0,0)\,\text{VBS}}=\sum_{i=1}^N H_i
4070: \end{equation}
4071: with 
4072: \begin{equation}
4073:   \label{eq:hi20000}
4074:   H_i=\left(\!\left(\!\bs{J}^{(4)}_i\!\right)^2 \! - \frac{32}{3}\right)
4075:   \!\left(\!\left(\!\bs{J}^{(4)}_i\!\right)^2 \! - \frac{44}{3}\right)
4076:   \!\left(\!\left(\!\bs{J}^{(4)}_i\!\right)^2 \! - \frac{50}{3}\right).
4077: \end{equation}
4078: These VBS states break translational symmetry only up to translations
4079: by three lattice spacings.  The spinons of this model are again
4080: confined through a linear potential, as illustrated below.
4081: \begin{equation}
4082: \setlength{\unitlength}{1pt}
4083: \begin{picture}(176,60)(4,-25)
4084: \linethickness{0.8pt}
4085: \multiput(0,10)(14,0){15}{\circle{4}}
4086: \multiput(0,0)(14,0){15}{\circle{4}}
4087: \multiput(0,10)(14,0){5}{\circle*{4}}
4088: \multiput(126,0)(14,0){5}{\circle*{4}}
4089: \thicklines
4090: \put(-2,0){\line(-1,0){4}}
4091: \multiput(2,10)(14,0){4}{\line(1,0){10}}
4092: \multiput(72,10)(14,0){5}{\line(1,0){10}}
4093: \multiput(156,10)(14,0){2}{\line(1,0){10}}
4094: \put(184,10){\line(1,0){4}}
4095: \multiput(2,0)(14,0){2}{\line(1,0){10}}
4096: \multiput(44,0)(14,0){5}{\line(1,0){10}}
4097: \multiput(128,0)(14,0){4}{\line(1,0){10}}
4098: %\multiput(156,0)(14,0){1}{\line(1,0){10}}
4099: %\put(198,0){\line(1,0){4}}
4100: \thinlines
4101: \put(91,-14){\vector(-1,0){63}}
4102: \put(91,-14){\vector(1,0){63}}
4103: \put(91,-22){\makebox(0,0){\small energy cost $\propto$ distance}}
4104: \put(28,22){\makebox(0,0){$\bs{\bar 6}$}}
4105: \put(28,32){\makebox(0,0){\small spinon}}
4106: \put(154,23){\makebox(0,0){$\bs{\bar 6}$}}
4107: \put(154,33){\makebox(0,0){\small spinon}}
4108: \put(107,-5){\dashbox{1}(52,20)}
4109: \end{picture}
4110: \label{eq:20000spinons}
4111: \end{equation}
4112: The conclusions we have drawn for this VBS model rest on the
4113: assumption that the quadratic Casimirs of the representations
4114: contained in the tensor product shown in the dashed box in
4115: \eqref{eq:20000state} as well as in the tensor product one obtains if
4116: one shifts this box by one lattice spacing to the left or to the right
4117: are smaller than the largest Casimir contained in the tensor product
4118: shown in the dotted box in \eqref{eq:20000spinons}.  We have verified the
4119: validity of this assumption for the (2,0,0,0,0) VBS model we
4120: considered here, but not shown it to hold for similar models with
4121: larger $n$ or $\lambda$.
4122: 
4123: (4) Finally, consider an SU(6) spin chain
4124: with spins transforming under the 70-dimensional representation
4125: \begin{displaymath}
4126:   (1,1,0,0,0)\ =\
4127:   \setlength{\unitlength}{8pt}
4128:   \begin{picture}(2,1)(0,-0.35)
4129:     \multiput(0,0)(1,0){3}{\line(0,1){1}}
4130:     \multiput(0,0)(1,0){2}{\line(0,-1){1}}
4131:     \multiput(0,0)(0,1){2}{\line(1,0){2}}
4132:     \multiput(0,-1)(0,1){1}{\line(1,0){1}}
4133:   \end{picture}\ \ .
4134: \end{displaymath}
4135: Thus we have once again $\frac{n}{q}=2$.  In a notation
4136: similar to the one introduced for the $\bs{8}$ VBS, 
4137: \begin{displaymath}
4138: %  (0,1,0,0,0)\ =\
4139:   \setlength{\unitlength}{8pt}
4140:   \begin{picture}(1.8,1)(0,0.15)
4141:     \multiput(0,0)(1,0){2}{\line(0,1){1}}
4142:     \multiput(0,0)(0,1){2}{\line(1,0){1}}
4143:   \end{picture}\hat =
4144:   \setlength{\unitlength}{1pt}
4145:   \begin{picture}(10,10)(0,0)
4146:     \put(6,2.5){\circle{4}}
4147:   \end{picture},\ \
4148:   \setlength{\unitlength}{8pt}
4149:   \begin{picture}(1.8,1)(0,-0.35)
4150:     \multiput(0,0)(1,0){2}{\line(0,1){1}}
4151:     \multiput(0,0)(1,0){2}{\line(0,-1){1}}
4152:     \multiput(0,-1)(0,1){3}{\line(1,0){1}}
4153:   \end{picture}\hat =
4154:   \setlength{\unitlength}{1pt}
4155:   \begin{picture}(11,10)(0,0)
4156:     \put(6,2.5){\circle{6}}
4157:   \end{picture} ,
4158: \end{displaymath}
4159: the two degenerate VBSs are illustrated by
4160: \begin{equation}
4161: \setlength{\unitlength}{1pt}
4162: \begin{picture}(162,48)(4,-22)
4163: \linethickness{0.8pt}
4164: \multiput(5,10)(60,0){3}{\circle{4}}
4165: \multiput(50,10)(60,0){3}{\circle{4}}
4166: \multiput(20,0)(60,0){3}{\circle{4}}
4167: \multiput(35,0)(60,0){3}{\circle{4}}
4168: \multiput(20,10)(60,0){3}{\circle{6}}
4169: \multiput(35,10)(60,0){3}{\circle{6}}
4170: \multiput(5,0)(60,0){3}{\circle{6}}
4171: \multiput(50,0)(60,0){3}{\circle{6}}
4172: %
4173: \thicklines
4174: \multiput(7,10)(60,0){3}{\line(1,0){10}}
4175: \multiput(38,10)(60,0){3}{\line(1,0){10}}
4176: \multiput(23,10)(60,0){3}{\line(1,0){9}}
4177: \multiput(37,0)(60,0){3}{\line(1,0){10}}
4178: \multiput(8,0)(60,0){3}{\line(1,0){10}}
4179: \multiput(53,0)(60,0){2}{\line(1,0){9}}
4180: \put(2,0){\line(-1,0){4}}
4181: \put(173,0){\line(1,0){4}}
4182: %
4183: \thinlines
4184: \put(44,-6){\framebox(12,22)}
4185: \put(50,-6){\line(0,-1){9}}
4186: \put(50,-22){\makebox(1,1){\small projection onto rep.\ $(1,1,0,0,0)$}}
4187: \put(50,25){\makebox(1,1){\small one site}}
4188: \put(104.5,-6){\dashbox{2}(41,22)}
4189: %\put(104.5,-6){\dashbox{2}(41,24)}
4190: %\put(89.5,-8){\dashbox{2}(41,24)}
4191: \end{picture}
4192: \label{eq:11000state}
4193: \end{equation}
4194: are exact ground states of a parent Hamiltonian involving the
4195: quadratic Casimir of the total spin of three neighboring sites,
4196: \begin{equation}
4197:   \label{eq:ham11000}
4198:   H_{(1,1,0,0,0)\,\text{VBS}}=\sum_{i=1}^N H_i
4199: \end{equation}
4200: with 
4201: \begin{equation}
4202:   \label{eq:hi11000}
4203:   H_i=\left(\!\left(\!\bs{J}^{(3)}_i\!\right)^2 \! - 20\right)
4204:   \!\left(\!\left(\!\bs{J}^{(3)}_i\!\right)^2 \! - 70\right)
4205:   \!\left(\!\left(\!\bs{J}^{(3)}_i\!\right)^2 \! - 540\right).
4206: \end{equation}
4207: The states \eqref{eq:11000state} are not the only VBSs one can form.
4208: Other possibilities like
4209: \begin{equation}
4210: \setlength{\unitlength}{1pt}
4211: \begin{picture}(174,12)(4,2)
4212: \linethickness{0.8pt}
4213: \multiput(5,10)(60,0){3}{\circle{6}}
4214: \multiput(50,10)(60,0){3}{\circle{6}}
4215: \multiput(20,0)(60,0){3}{\circle{6}}
4216: \multiput(35,0)(60,0){3}{\circle{6}}
4217: \multiput(20,10)(60,0){3}{\circle{4}}
4218: \multiput(35,10)(60,0){3}{\circle{4}}
4219: \multiput(5,0)(60,0){3}{\circle{4}}
4220: \multiput(50,0)(60,0){3}{\circle{4}}
4221: %
4222: \thicklines
4223: \multiput(8,10)(60,0){3}{\line(1,0){10}}
4224: \multiput(37,10)(60,0){3}{\line(1,0){10}}
4225: \multiput(22,10)(60,0){3}{\line(1,0){11}}
4226: \multiput(38,0)(60,0){3}{\line(1,0){10}}
4227: \multiput(7,0)(60,0){3}{\line(1,0){10}}
4228: \multiput(52,0)(60,0){2}{\line(1,0){11}}
4229: \put(3,0){\line(-1,0){4}}
4230: \put(172,0){\line(1,0){4}}
4231: %
4232: \thinlines
4233: \put(74.5,-6){\dashbox{2}(41,22)}
4234: \end{picture}
4235: \nonumber
4236: \end{equation}
4237: or
4238: \begin{equation}
4239: \setlength{\unitlength}{1pt}
4240: \begin{picture}(178,16)(4,-2)
4241: \linethickness{0.8pt}
4242: \multiput(5,10)(30,0){6}{\circle{4}}
4243: \multiput(20,10)(30,0){6}{\circle{6}}
4244: \multiput(5,0)(30,0){6}{\circle{6}}
4245: \multiput(20,0)(30,0){6}{\circle{4}}
4246: %
4247: \thicklines
4248: \multiput(7,10)(30,0){6}{\line(1,0){10}}
4249: \multiput(23,10)(60,0){3}{\line(1,0){10}}
4250: \multiput(8,0)(30,0){6}{\line(1,0){10}}
4251: \multiput(52,0)(60,0){2}{\line(1,0){10}}
4252: \put(2,0){\line(-1,0){4}}
4253: \put(172,0){\line(1,0){4}}
4254: %
4255: \thinlines
4256: \put(104.5,-6){\dashbox{2}(41,24)}
4257: \put(89.5,-8){\dashbox{2}(41,24)}
4258: \end{picture},
4259: \nonumber
4260: \end{equation}
4261: however, contain additional representations for the total SU(6) spin
4262: of three neighboring sites, and are hence expected to possess
4263: higher energies.
4264: 
4265: Spinon excitations transforming under the 6-dimensional rep.\
4266: $(0,0,0,0,1)$ are linearly confined to spinons transforming under the
4267: 15-dimensional rep.\ $(0,0,0,1,0)$:
4268: \begin{equation}
4269: \setlength{\unitlength}{1pt}
4270: \begin{picture}(180,60)(4,-25)
4271: \linethickness{0.8pt}
4272: \multiput(5,10)(60,0){3}{\circle{4}}
4273: \multiput(50,10)(60,0){3}{\circle{4}}
4274: \multiput(20,0)(60,0){3}{\circle{4}}
4275: \multiput(35,0)(60,0){3}{\circle{4}}
4276: \multiput(20,10)(60,0){3}{\circle{6}}
4277: \multiput(35,10)(60,0){3}{\circle{6}}
4278: \put(185,10){\circle{6}}
4279: \multiput(5,0)(60,0){3}{\circle{6}}
4280: \multiput(50,0)(60,0){3}{\circle{6}}
4281: %
4282: \multiput(5,10)(30,0){1}{\circle*{4}}
4283: \multiput(20,10)(15,0){2}{\circle*{6}}
4284: \multiput(170,0)(15,0){1}{\circle*{6}}
4285: \multiput(155,0)(30,0){2}{\circle*{4}}
4286: %
4287: \thicklines
4288: \multiput(7,10)(60,0){3}{\line(1,0){10}}
4289: \put(172,10){\line(1,0){10}}
4290: \multiput(52,10)(60,0){2}{\line(1,0){11}}
4291: \multiput(23,10)(60,0){3}{\line(1,0){9}}
4292: \multiput(8,0)(60,0){3}{\line(1,0){10}}
4293: \multiput(37,0)(60,0){3}{\line(1,0){10}}
4294: \multiput(53,0)(60,0){2}{\line(1,0){9}}
4295: \put(173,0){\line(1,0){10}}
4296: \put(2,0){\line(-1,0){4}}
4297: \put(188,10){\line(1,0){4}}
4298: %
4299: \thinlines
4300: \put(20,25){\makebox(0,0){\small $(0,0,0,0,1)$}}
4301: \put(20,36){\makebox(0,0){\small spinon }}
4302: \put(170,25){\makebox(0,0){\small $(0,0,0,1,0)$}}
4303: \put(170,36){\makebox(0,0){\small spinon}}
4304: \put(74.5,-6){\dashbox{2}(41,22)}
4305: \put(95,-14){\vector(-1,0){75}}
4306: \put(95,-14){\vector(1,0){75}}
4307: \put(95,-22){\makebox(0,0){\small energy cost $\propto$ distance}}
4308: \end{picture}
4309: \label{eq:11000spinons}
4310: \end{equation}
4311: The VBS configuration we have drawn between the two spinons in
4312: \eqref{eq:11000spinons} constitutes just one of several possibilities.
4313: We expect, however, that this possibility corresponds to the lowest
4314: energy among them for the Hamiltonian \eqref{eq:ham11000}.  This
4315: concludes our list of examples.
4316: 
4317: The models introduced in this subsection are interesting in that they
4318: provide us with examples where spinon confinement, and with the
4319: confinement the existence of a Haldane gap, is caused by interactions
4320: extending beyond nearest neighbors.  The conclusion drawn from the
4321: Affleck--Lieb theorem for SU($n$) models with spins transforming under
4322: representations we have labeled here as the ``third category'' hence
4323: appear to hold for models with nearest-neighbor interactions only, to
4324: which the theorem is applicable.  For these models, the theorem states
4325: that the spectrum is gapless, which according to our understanding
4326: implies that the models support deconfined spinon excitations.  The
4327: examples we have studied, by contrast, suggest that models with
4328: longer-ranged interactions belonging to this category exhibit
4329: confinement forces between the spinon excitations and hence possess a
4330: Haldane gap.
4331: 
4332: It is worth noting that even though in the examples we elaborated here
4333: $\lambda <n$, we expect our conclusions to hold for models with
4334: $\lambda >n$ as well.  To see this, let $m>0$ be an integer such that
4335: $n m<\lambda <n(m+1)$.  We can now construct a first VBS with spinon
4336: confinement using $n m$ boxes of the Young tableau and combine it with
4337: a second by projection on each side with a second VBS constructed from
4338: the remaining $\lambda'= \lambda-n m$ boxes.  Since the spinons
4339: of the first VBS are always confined and hence gapped, the final VBS
4340: will support deconfined spinons if and only if the second VBS will
4341: support them, which in turn will depend on the largest common divisor
4342: $q'$ of $\lambda'$ and $n$ as well as the range of the interaction.
4343: Since the largest common divisor $q$ of $n$ and $\lambda$ is
4344: equal to $q'$, there is no need to think in terms of $\lambda'$ and $q'$.
4345: The conclusions regarding confinement and the Haldane gap will not
4346: depend on the distinction between $\lambda$ and $\lambda'$.
4347: 
4348: \subsection{The different categories of models}
4349: \label{sec:sum}
4350: 
4351: In this section, we used the rules
4352: emerging from the numerous examples we studied to argue that models of
4353: SU($n$) spin chains in general fall into three categories.  The
4354: classification depends on the number of boxes $\lambda$ the Young
4355: tableau corresponding to the representation of the individual spins
4356: consists off, as follows:
4357: \begin{enumerate}
4358: \item If $\lambda$ and $n$ have no common divisor, the models will
4359:   support free spin excitations and hence not exhibit a Haldane gap.
4360: 
4361:   The general reasoning here is simply that the VBS states in this
4362:   category break translational invariance up to translations by $n$
4363:   lattice spacings, and that there are (at least) $n$ degenerate
4364:   VBS ground states to each model.  Spinons transforming under
4365:   representations of Young tableaux with an arbitrary number of boxes
4366:   can be accommodated in domain walls between these different ground
4367:   states.  Consequently, the spinons are deconfined.  
4368: 
4369: \item If $\lambda$ is divisible by $n$, the general argument we have
4370:   constructed in Sec.\ \ref{sec:generalcriterion} indicates that the
4371:   model will exhibit spinon confinement and hence a Haldane gap.
4372: 
4373: \item \label{three}
4374:   If $\lambda$ and $n$ have a common divisor different from $n$,
4375:   the examples studied in Sec.\ \ref{sec:examples} suggest that the
4376:   question of whether the spinons are confined or not depends on the
4377:   details of the interactions.  If $q$ is the largest common divisor
4378:   of $\lambda$ and $n$, interactions ranging to the $\frac{n}{q}$-th
4379:   neighbor were required for spinon confinement in the models we
4380:   studied.  The Affleck-Lieb theorem~\cite{affleck-86lmp57}, on the
4381:   other hand, tells us that SU($n$) chains with nearest-neighbor
4382:   Heisenberg interactions belonging to this category are gapless if
4383:   the ground states are non-degenerate.
4384: \end{enumerate}
4385: 
4386: Note that the second category is really just the special case $q=n$
4387: of the third:  with $\frac{n}{q}=1$, nearest-neighbor interactions
4388: are already sufficient for spinon confinement and a Haldane gap.
4389: 
4390: The conclusion that interactions ranging to the $\frac{n}{q}$-th
4391: neighbor are required for spinon confinement and a Haldane gap,
4392: however, is not universally valid.  A counter-example is provided by
4393: the extended VBSs (XVBSs) introduced by Affleck, Arovas, Marston, and
4394: Rabson~\cite{affleck-91npb467}.  In this model, each site effectively
4395: takes the role of two neighboring sites when a VBS is constructed, and
4396: nearest-neighbor interactions are already sufficient to cause spinon
4397: confinement.  We briefly review this model in the following
4398: subsection.
4399: 
4400: 
4401: \subsection{A counter-example: the SU(4) representation $\bs{6}$
4402: extended VBS}
4403: \label{sec:counter}
4404: 
4405: In an article devoted to quantum antiferromagnets with spins transforming under
4406:  the self-conjugate representations of SU($2n$),
4407: Affleck \ea~\cite{affleck-91npb467} introduced an extended VBS for the
4408: six-dimensional SU(4) representation $(0,1,0)$.  This %six-dimensional 
4409: representation is, with regard to the number of boxes the corresponding
4410: Young tableau consists of, not distinguishable from the symmetric 
4411: representation $\bs{10}$ considered in Sec.~\ref{sec:examples}, as both
4412: tableaux consist of two boxes:
4413: \begin{equation}
4414: \setlength{\unitlength}{0.8pt}  
4415: \begin{picture}(116,36)(-6,-6)
4416: \put(-6,10){\line(1,0){10}} 
4417: \put(-6,20){\line(1,0){10}} 
4418: \put(-6,10){\line(0,1){10}} 
4419: \put(4,10){\line(0,1){10}} 
4420: \put(14,15){\makebox(1,1){$\otimes$}}
4421: \put(24,10){\line(1,0){10}} 
4422: \put(24,20){\line(1,0){10}} 
4423: \put(24,10){\line(0,1){10}} 
4424: \put(34,10){\line(0,1){10}} 
4425: \put(47,15){\makebox(1,1){$=$}}
4426: \put(60,10){\line(1,0){20}} 
4427: \put(60,20){\line(1,0){20}} 
4428: \put(60,10){\line(0,1){10}} 
4429: \put(70,10){\line(0,1){10}} 
4430: \put(80,10){\line(0,1){10}} 
4431: \put(90,15){\makebox(1,1){$\oplus$}}
4432: % \put(100,10){\line(1,0){10}} 
4433: % \put(100,20){\line(1,0){10}} 
4434: % \put(100,30){\line(1,0){10}} 
4435: % \put(100,10){\line(0,1){20}} 
4436: % \put(110,10){\line(0,1){20}} 
4437: \put(100,5){\line(1,0){10}} 
4438: \put(100,15){\line(1,0){10}} 
4439: \put(100,25){\line(1,0){10}} 
4440: \put(100,5){\line(0,1){20}} 
4441: \put(110,5){\line(0,1){20}} 
4442: \put(-1,-2){\makebox(1,1){$\bs{4}$}}
4443: \put(29,-2){\makebox(1,1){$\bs{4}$}}
4444: \put(70,-2){\makebox(1,1){$\bs{10}$}}
4445: \put(105,-7){\makebox(1,1){$\bs{6}$}}
4446: \end{picture}
4447: \end{equation}
4448: Since the two boxes are combined antisymmetrically for rep.~$\bs{6}$,
4449: a VBS constructed along the lines of Sec.~\ref{sec:examples} would no
4450: longer provide a paradigm for antiferromagnetic spin chains of the
4451: corresponding representation in general.  Affleck
4452: \ea~\cite{affleck-91npb467} have constructed an extended VBS, % (XVBS),
4453: which is illustrated in the following cartoon:
4454: \begin{equation}
4455: \label{gs.su4.xvbs}
4456: \setlength{\unitlength}{1pt}
4457: \begin{picture}(170,20)(0,0)
4458:  \multiput(0,10)(30,0){6}{\circle{4}}
4459:  \multiput(8,10)(30,0){6}{\circle{4}}
4460:  \multiput(4,10)(30,0){6}{\circle{16}}
4461: % \multiput(-4,4)(30,0){6}{\dashbox{2}(16,12)}
4462: % \multiput(-4,4)(30,0){6}{\framebox(16,12)}
4463: \thicklines
4464:  \multiput(2,10)(60,0){3}{\line(1,0){4}}
4465:  \multiput(10,10)(30,0){5}{\line(1,0){18}}
4466:  \put(-8,10){\line(1,0){6}}
4467:  \put(160,10){\line(1,0){7}}
4468: \end{picture}
4469: \end{equation}
4470: Here each small circle represents a fundamental representation
4471: $\bs{4}$ of SU(4) (a box in the Young tableau), and each large circle a
4472: lattice site.  The lines connecting four dots indicate that these four
4473: fundamental representations are combined into an SU(4) singlet.  The
4474: total spin of two neighboring sites in this state may assume the
4475: representations
4476: \begin{equation}
4477:   \bs{4}\otimes\bs{\bar 4}=\bs{1}\oplus\bs{15},
4478: \end{equation}
4479: while combining two reps.~$\bs{6}$ on neighboring sites in general
4480: yields
4481: \begin{equation}
4482:   \bs{6}\otimes\bs{6}=\bs{1}\oplus\bs{15}\oplus\bs{20}.
4483: \end{equation}
4484: To construct a parent Hamiltonian for the XVBS \eqref{gs.su4.xvbs}, it
4485: is hence sufficient to sum over projectors onto rep.~$\bs{20}$ on all
4486: pairs of neighboring sites.  Using our conventions (Affleck \ea
4487: ~\cite{affleck-91npb467} have normalized the eigenvalues of the
4488: quadratic Casimir operator to $\casinn{4}{\textrm{adjoint rep.}}=8$),
4489: the parent Hamiltonian takes the form
4490: \begin{equation}
4491:   \label{ham.su4.xvbs}
4492:   H_{\textrm{XVBS}}=\sum_{i=1}^N \left( \bs{J}_i\bs{J}_{i+1} +
4493:      \frac{1}{3}\left(\bs{J}_i\bs{J}_{i+1}\right)^2+\frac{5}{12} \right),
4494: \end{equation}
4495: where the operators $J_i^a$, $a=1,\ldots,15$, are given by $6\times 6$
4496: matrices.  Note that the ground state is two fold degenerate, as it 
4497: breaks translational symmetry modulo translations by two lattice spacings.
4498: 
4499: We conjecture that the lowest lying excitation is a
4500: bound state consisting of two fundamental reps $\bs{4}$, which most
4501: likely are combined antisymmetrically into a rep.~$\bs{6}$:
4502: \begin{equation*}
4503:   \setlength{\unitlength}{1pt}
4504: \begin{picture}(170,20)(0,0)
4505:  \multiput(0,10)(30,0){6}{\circle{4}}
4506:  \multiput(8,10)(30,0){6}{\circle{4}}
4507:  \multiput(68,10)(22,0){2}{\circle*{4}}
4508:  \multiput(4,10)(30,0){6}{\circle{16}}
4509: % \multiput(-4,4)(30,0){6}{\dashbox{2}(16,12)}
4510: % \multiput(52,0)(60,0){3}{\line(1,0){10}}
4511: \thicklines
4512:  \multiput(32,10)(90,0){2}{\line(1,0){4}}
4513:  \multiput(10,10)(30,0){5}{\line(1,0){18}}
4514:  \put(-8,10){\line(1,0){6}}
4515:  \put(160,10){\line(1,0){7}}
4516: \end{picture}
4517: \end{equation*}
4518: The SU(4) rep.~$\bs{6}$ XVBS provides us with an example where a
4519: nearest-neighbor Hamiltonian is sufficient to induce spinon
4520: confinement and a Haldane gap, even though the largest common divisor
4521: of $n=4$ and $\lambda=2$ is $q=2$, \ie interactions including
4522: $\frac{n}{q}=2$ neighbors would be required following the rules
4523: derived from the examples in Sec.~\ref{sec:examples}.  The reason for
4524: this discrepancy is that in the XVBS model considered here, each site
4525: effectively takes the role of two neighboring sites.  
4526: Affleck \ea~\cite{affleck-91npb467,marston-89prb11538} conjecture
4527: that the ground state of the SU(4) rep.~$\bs{6}$ nearest-neighbor 
4528: Heisenberg model is, like the XVBS reviewed here, two-fold degenerate,
4529: which implies that the Affleck-Lieb theorem is not applicable.
4530: 
4531: This example is valuable in showing that it is advisable to explicitly
4532: construct the VBS for a given representation of SU($n$) in order to
4533: verify the applicability of the general rules motivated above.
4534: 
4535: \section{Conclusion}
4536: \label{sec:conclusion}
4537: 
4538: In the first part of this article, we have formulated several exact
4539: models of SU(3) spin chains.  We introduced a trimer model and
4540: presented evidence that the elementary excitations of the model
4541: transform under the SU(3) representations conjugate to the
4542: representation of the original spin on the chain.  We then introduced
4543: three SU(3) valence bond solid chains with spins transforming under
4544: representations $\bs{6}$, $\bs{10}$, and $\bs{8}$, respectively.  We
4545: argued that of these four models, the coloron excitations are confined
4546: only in the $\bs{10}$ and the $\bs{8}$ VBS models, and that only those
4547: models with confined spinons exhibit a Haldane gap.  We subsequently
4548: generalized three of our models to SU($n$), and investigated again
4549: which models exhibit spinon confinement.  
4550: 
4551: Finally, we used the rules emerging from the numerous examples we
4552: studied to argue that models of SU($n$) spin chains in general fall
4553: into three categories with regard to spinon confinement and the
4554: Haldane gap.  These are summarized in the previous subsection.
4555: The results rely crucially on the assumption that the conclusions we
4556: obtained for the VBS models we studied are of general validity.
4557: This assumption certainly holds for the corresponding SU(2) models,
4558: and appears reasonable on physical grounds.  Ultimately, however,
4559: it is only an assumption, or at best a hypotheses.
4560: 
4561: On a broader perspective, we believe that the models we have studied
4562: provide further indication that SU($n$) spin chains are an
4563: equally rich and rewarding subject of study as SU(2) spin chains have 
4564: been since Bethe.
4565: 
4566: 
4567: 
4568: 
4569: \section*{ACKNOWLEDGMENTS}
4570: 
4571: We would like to thank Peter Zoller and Peter W\"olfle, but 
4572: especially Dirk Schuricht and Ronny Thomale for many highly inspiring 
4573: discussions of various aspects of this work.  We further wish to
4574: thank Dirk Schuricht for valuable suggestions on the manuscript.  
4575: One of us (SR) was supported by a Ph.D.\ scholarship of the Cusanuswerk.
4576: 
4577: %\newpage
4578: \appendix
4579: \section{A proposal for an experimental realization of SU(3) spin 
4580: chains in an optical lattice}
4581: \label{sec:exp}
4582: 
4583: In this appendix, we wish to describe a proposal for an experimental
4584: realization of SU(3) spin chains. The most eligible candidate for such
4585: experiments are ultra-cold gases in optical lattices. Recently, these
4586: systems have become an interesting playground for the realization of
4587: various problems of condensed matter physics, such as the phase
4588: transition from a superfluid to a Mott
4589: insulator~\cite{Jaksch-98prl3108,Greiner-02n39}, the fermionic Hubbard
4590: model~\cite{Honerkamp-04prl170403,Honerkamp-04prb094521,rapp-06cm0607138},
4591: and SU(2) spin chains~\cite{Duan-03prl090402,Garcia-04prl250405}.  In
4592: particular, the Hamiltonians for spin lattice models may be engineered
4593: with polar molecules stored in optical lattices, where the spin is
4594: represented by a single-valence electron of a heteronuclear
4595: molecule~\cite{micheli-06np341,brennen-06qp0612180}.
4596: 
4597: In a most naive approach, one might expect to realize an SU(3) spin
4598: (at a site in an optical lattice) by using atoms with three internal
4599: states, like an atom with spin $S=1$.  If we now were to interpret the
4600: $S^z=+1$ state as SU(3) spin ``blue'', the $S^z=0$ state as ''red'',
4601: and the $S^z=-1$ state as ``green'', however, the SU(3) spin would not
4602: be conserved.  The SU(2) algebra would allow for the process
4603: $\ket{+1,-1}\rightarrow\ket{0,0}$, which in SU(3) language corresponds
4604: to the forbidden process $\ket{\b,\g}\rightarrow\ket{\r,\r}$.
4605: 
4606: A more sophisticated approach is hence required.  One way to obtain a
4607: system with three internal states in which the number of particles in
4608: each state (\ie of each color) is conserved 
4609: %(proposed to us by P.\ Zoller) 
4610: is to manipulate an atomic system with total
4611: angular momentum $F=3/2$ (where
4612: $\bs{F}=\bs{S}_{\text{el}}+\bs{L}_{\text{orb}}+\bs{S}_{\text{nuc}}$
4613: includes the internal spin of the electrons, the orbital angular
4614: momentum, and the spin of the nucleus) to simulate an SU(3) spin.  The
4615: important feature here is that the atoms have four internal states,
4616: corresponding to
4617: $F^z=-\frac{3}{2},-\frac{1}{2},+\frac{1}{2},+\frac{3}{2}$.  
4618: %, which is the case for \eg $^{137}$Ba.  
4619: For such atoms, one has to suppress the
4620: occupation of one of the ``middle'' states, say the $F^z=-\frac{1}{2}$
4621: state, by effectively lifting it to a higher energy while keeping the
4622: other states approximately degenerate.  This can be accomplished
4623: through a combination of an external magnetic field and two carefully
4624: tuned lasers, which effectively push down the energies of the
4625: $F^z=-\frac{3}{2}$ and the $F^z=+\frac{1}{2}$ states by coupling these
4626: states to states of (say) the energetically higher $F=5/2$ multiplet
4627: (see Fig.~\ref{fig:exp}).
4628: \begin{figure}
4629: \begin{center}
4630: \includegraphics[width=0.38\textwidth]{exper2.eps}
4631: \caption{Effective lifting of the $F^z=-\frac{1}{2}$ state.}
4632: \label{fig:exp}
4633: \end{center}
4634: \end{figure}
4635: At sufficiently low temperatures, we are hence left with a system with
4636: three internal states $F^z=-\frac{3}{2},+\frac{1}{2},+\frac{3}{2}$,
4637: which we may identify with the colors ``blue'', ``red'', and ``green''
4638: of an SU(3) spin.  In leading order, the number of particles of each
4639: color is now conserved, as required by SU(3) symmetry.
4640: For example, conservation of $F^z$ forbids processes in which a
4641: ``blue'' and a ``green'' particle turn into two ``red'' ones,
4642: $\ket{\b,\g}\rightarrow\ket{\r,\r}$.  Higher order processes of the
4643: kind $\ket{\b,\g,\g}\rightarrow\ket{\r,\r,\r}$ are still possible, but
4644: negligible if the experiment is conducted at sufficiently short time
4645: scales.
4646: 
4647: If one places fermionic atoms with an artificial SU(3) spin engineered
4648: along the lines of this or a related proposal in an optical lattice
4649: and allows for a weak hopping of the atoms on the lattice, one has
4650: developed an experimental realization of an SU(3) Hubbard model.
4651: If the energy cost $U$ of having two atoms on the same lattice site is
4652: significantly larger than the hopping $t$, and the density is one atom
4653: per site, the system will effectively constitute an SU(3)
4654: antiferromagnet.  The dimension of this antiferromagnet will depend on
4655: the optical lattice, which can be one-, two-, or three-dimensional.
4656: 
4657: %\newpage
4658: In principle, the above proposal can be generalized to SU($n$), even
4659: though the experimental obstacles are likely to grow ``exponentially''
4660: with $n$.  Besides, it is far from clear that such an endeavor is
4661: worthwhile, as all the non-trivial properties of SU($n$) are already
4662: present in SU(3) chains (while SU(2) constitutes a special case).
4663: 
4664: %\vfill\newpage
4665: 
4666: \section{Gell-Mann matrices}
4667: \label{app:conventions}
4668: 
4669: The Gell-Mann matrices are given by~\cite{Cornwell84vol2,Georgi82}
4670: \begin{displaymath}
4671: %\begin{array}{r@{}c@{}l@{}r@{}c@{}l@{}r@{}c@{}l}
4672: \begin{array}{c@{}c@{}c@{\hspace{6pt}}c@{}c@{}c@{\hspace{6pt}}c@{}c@{}c}
4673: \lambda^1&=&\!\left(\begin{array}{ccc}0&1&0\\1&0&0\\0&0&0\end{array}
4674: \right)\!\!,&
4675: \lambda^2&=&\!\left(\begin{array}{ccc}0&-i&0\\i&0&0\\0&0&0\end{array}
4676: \right)\!\!,&
4677: \lambda^3&=&\!\left(\begin{array}{ccc}1&0&0\\0&-1&0\\0&0&0\end{array}
4678: \right)\!\!,\\[27pt]
4679: \lambda^4&=&\!\left(\begin{array}{ccc}0&0&1\\0&0&0\\1&0&0\end{array}
4680: \right)\!\!,&
4681: \lambda^5&=&\!\left(\begin{array}{ccc}0&0&-i\\0&0&0\\i&0&0\end{array}
4682: \right)\!\!,&
4683: \lambda^6&=&\!\left(\begin{array}{ccc}0&0&0\\0&0&1\\0&1&0\end{array}
4684: \right)\!\!,\\[27pt]
4685: %\end{array}
4686: %\end{displaymath}
4687: %\begin{displaymath}
4688: %\begin{array}{c@{}c@{}c@{\hspace{6pt}}c@{}c@{}c@{\hspace{6pt}}c@{}c@{}c}
4689: \lambda^7&=&\!\left(\begin{array}{ccc}0&0&0\\0&0&-i\\0&i&0\end{array}
4690: \right)\!\!,&
4691: \lambda^8&=&
4692: \multicolumn{3}{l}{{\displaystyle\frac{1}{\sqrt{3}}}
4693: \!\left(\begin{array}{ccc}1&0&0\\0&1&0\\0&0&-2\end{array}\right)\!\!.}
4694: \end{array}
4695: \end{displaymath}
4696: They are normalized as
4697: $\mathrm{tr}\left(\lambda^a\lambda^b\right)=2\delta_{ab}$ and satisfy
4698: the commutation relations
4699: $\comm{\lambda^a}{\lambda^b}=2f^{abc}\lambda^c.$ The structure
4700: constants $f^{abc}$ are totally antisymmetric and obey Jacobi's
4701: identity %$f^{abc}f^{cde}+f^{bdc}f^{cae}+f^{dac}f^{cbe}=0$.
4702: \begin{displaymath}
4703: f^{abc}f^{cde}+f^{bdc}f^{cae}+f^{dac}f^{cbe}=0.
4704: \end{displaymath} 
4705: Explicitly, the non-vanishing structure constants are given by $f^{123}=i$,
4706: $f^{147}=f^{246}=f^{257}=f^{345}=-f^{156}=-f^{367}=i/2$,
4707: $f^{458}=f^{678}=i\sqrt{3}/2$, and 45 others obtained by permutations of 
4708: the indices.\\[-0.1pt]
4709: 
4710: \section{Eigenvalues of the quadratic Casimir operator}
4711: \label{quadraticcasimirs}
4712: 
4713: The eigenvalues of the quadratic Casimir operator for representations
4714: $\casin{\mu_1,\mu_2,\ldots,\mu_{n-1}}$ of SU($n$) up to $n=6$ are given by:
4715: \begin{itemize}
4716: \item[] $ \casinn{2}{\mu}=\frac{1}{4}\left(\mu^2+2\mu\right)=
4717:        \frac{\mu}{2}\left(\frac{\mu}{2}+1\right)$ 
4718: \item[] $\casinn{3}{\mu_1,\mu_2}=\frac{1}{3} \bigl( \mu_1^2 + \mu_1\mu_2 +
4719:        \mu_2^2 + 3\mu_1 + 3\mu_2 \bigr)$
4720: \item[] $\casinn{4}{\mu_1,\mu_2,\mu_3} =
4721:        \frac{1}{8}\left(3\mu_1^2+4\mu_2^2+
4722:          3\mu_3^2+4\mu_1\mu_2\right.$\\[3pt]
4723:        $ \phantom{\casinn{2}{\mu}} \left.
4724:       + 2\mu_1\mu_3 +4\mu_2\mu_3 + 12\mu_1+ 16\mu_2+12\mu_3 \right)$
4725: \item[] $\casinn{5}{\mu_1,\mu_2,\mu_3,\mu_4} = \frac{1}{5}\left( 2\mu_1^2 
4726:       + 3\mu_2^2 + 3\mu_3^2 + 2\mu_4^2 \right.$\\[3pt] 
4727: $\phantom{\casinn{2}{\mu}} + 3\mu_1\mu_2 + 4\mu_2\mu_3 + 3\mu_3\mu_4 + 
4728: 2\mu_1\mu_3 + \mu_1\mu_4$\\[3pt]
4729: $\phantom{\casinn{2}{\mu}} \left. + 2\mu_2\mu_4 + 10\mu_1
4730: + 15\mu_2 + 15\mu_3 + 10\mu_4 \right)$
4731: \item[] $\casinn{6}{\mu_1,\mu_2,\mu_3,\mu_4,\mu_5}= $\\[3pt]
4732: $\phantom{\casinn{2}{\mu}}\!\!\!\frac{1}{12}\left( 
4733:   5\mu_1^2 + 8\mu_2^2 + 9\mu_3^2 + 8\mu_4^2 + 5\mu_5^2 \right.$\\[3pt]
4734: $\phantom{\casinn{2}{\mu}}+ 8\mu_1\mu_2 + 12\mu_2\mu_3 + 12\mu_3\mu_4 
4735:   + 8\mu_4\mu_5  $\\[3pt]
4736: $\phantom{\casinn{2}{\mu}} + 4\mu_1\mu_4 + 6\mu_1\mu_3  + 8\mu_2\mu_4 + 6\mu_3\mu_5$\\[3pt]
4737: $\phantom{\casinn{2}{\mu}} + 4\mu_2\mu_5 + 2\mu_1\mu_5 + 30\mu_1 + 48\mu_2$\\[3pt]
4738: $\phantom{\casinn{2}{\mu}}\left. + 54\mu_3 + 48\mu_4 + 30\mu_5 \right)$\\
4739: \end{itemize}
4740: The general method to obtain these and further eigenvalues for $n>6$
4741: requires a discussion of representation theory~\cite{Humphreys87} at a level
4742: which is beyond the scope of this article.
4743: 
4744: The dimensionality of a representation $(\mu_1,\mu_2,\ldots,\mu_{n-1})$
4745: is determined by the so-called Hook formula~\cite{Cornwell84vol2} 
4746: %for SU($n$)
4747: \begin{equation}
4748:   \label{eq:hook-formula}
4749:   \text{dim}=\frac{\prod_{i<j}^n\left( \lambda_i - \lambda_j + j - i  \right)}
4750:              {\prod_{i<j}^n\left( j - i  \right)},
4751: \end{equation}
4752: where $\lambda_i=\sum_{j=i}^{n-1}\mu_j$ for $i=1,\ldots,n$. In particular, 
4753: it yields for $n=2,3,4$:
4754: 
4755: %\pagebreak
4756: \begin{itemize}
4757: \item[] $\dimn{2}{\mu}=\mu+1$
4758: \item[] $\dimn{3}{\mu_1,\mu_2}=\frac{1}{2}(\mu_1+1)(\mu_2+1)(\mu_1+\mu_2+2)$
4759: %\pagebreak
4760: \item[] \parbox{\linewidth}{$\dimn{4}{\mu_1,\mu_2,\mu_3}=\frac{1}{12}(\mu_1+1)(\mu_2+1)
4761:          (\mu_3+1)$\\[3pt]
4762:         $\phantom{\text{dimdim}}(\mu_1+\mu_2+2)(\mu_2+\mu_3+2)
4763:          (\mu_1+\mu_2+\mu_3+3)$}
4764: \end{itemize}
4765: 
4766: \begin{thebibliography}{10}
4767: 
4768: \bibitem{bethe31zp205}
4769: H. Bethe, Z. Phys. {\bf 71},  205  (1931), translated in D. C. Mattis, ed.,
4770:   \emph{The Many-Body Problem} (World Scientific, Singapore, 1993).
4771: 
4772: \bibitem{yang67prl1312}
4773: C.~N. Yang, Phys. Rev. Lett. {\bf 19},  1312  (1967).
4774: 
4775: \bibitem{baxter90}
4776: R.~J. Baxter, {\em Exactly Solved Models in Statistical Mechanics} (Academic
4777:   Press, London, 1990).
4778: 
4779: \bibitem{KorepinBogoliubovIzergin93}
4780: V.~E. Korepin, N.~M. Bogoliubov, and A.~G. Izergin, {\em Quantum Inverse
4781:   Scattering Method and Correlation Functions} (Cambridge University Press,
4782:   Cambridge, 1997).
4783: 
4784: \bibitem{faddeev-81pla375}
4785: L.~D. Faddeev and L.~A. Takhtajan, Phys. Lett.~A {\bf 85},  375  (1981).
4786: 
4787: \bibitem{laughlin83prl1395}
4788: R.~B. Laughlin, Phys. Rev. Lett. {\bf 50},  1395  (1983).
4789: 
4790: \bibitem{stone92}
4791: {\em Quantum Hall Effect}, edited by M. Stone (World Scientific, Singapur,
4792:   1992).
4793: 
4794: \bibitem{haldane83pla464}
4795: F.~D.~M. Haldane, Phys. Lett.~A {\bf 93},  464  (1983).
4796: 
4797: \bibitem{haldane83prl1153}
4798: F.~D.~M. Haldane, Phys. Rev. Lett. {\bf 50},  1153  (1983).
4799: 
4800: \bibitem{affleck90proc}
4801: I. Affleck,  in {\em Fields, strings and critical phenomena}, Vol.~XLIX of {\em
4802:   Les {H}ouches lectures}, edited by E. Br\'ezin and J. Zinn-Justin (Elsevier,
4803:   Amsterdam, 1990).
4804: 
4805: \bibitem{Fradkin91}
4806: E. Fradkin, {\em Field Theories of Condensed Matter Systems}, No.~82 in {\em
4807:   Frontiers in Physics} (Westview Press, Bouler, 1991).
4808: 
4809: \bibitem{GogolinNersesyanTsvelik98}
4810: A.~O. Gogolin, A.~A. Nersesyan, and A.~M. Tsvelik, {\em Bosonization and
4811:   Strongly Correlated Systems} (Cambridge University Press, Cambridge, 1998).
4812: 
4813: \bibitem{Giamarchi04}
4814: T. Giamarchi, {\em Quantum Physics in One Dimension} (Oxford University Press,
4815:   Oxford, 2004).
4816: 
4817: \bibitem{majumdar-69jmp1399}
4818: C.~K. Majumdar and D.~K. Ghosh, J.~Math. Phys. {\bf 10},  1399  (1969).
4819: 
4820: \bibitem{affleck-87prl799}
4821: I. Affleck, T. Kennedy, E.~H. Lieb, and H. Tasaki, Phys. Rev. Lett. {\bf 59},
4822:   799  (1987).
4823: 
4824: \bibitem{affleck-88cmp477}
4825: I. Affleck, T. Kennedy, E.~H. Lieb, and H. Tasaki, Commun. Math. Phys. {\bf
4826:   115},  477  (1988).
4827: 
4828: \bibitem{haldane88prl635}
4829: F.~D.~M. Haldane, Phys. Rev. Lett. {\bf 60},  635  (1988).
4830: 
4831: \bibitem{shastry88prl639}
4832: B.~S. Shastry, Phys. Rev. Lett. {\bf 60},  639  (1988).
4833: 
4834: \bibitem{haldane91prl1529}
4835: F.~D.~M. Haldane, Phys. Rev. Lett. {\bf 66},  1529  (1991).
4836: 
4837: \bibitem{haldane-92prl2021}
4838: F.~D.~M. Haldane, Z.~N.~C. Ha, J.~C. Talstra, D. Bernard, and V. Pasquier,
4839:   Phys. Rev. Lett. {\bf 69},  2021  (1992).
4840: 
4841: \bibitem{haldane91prl937}
4842: F.~D.~M. Haldane, Phys. Rev. Lett. {\bf 67},  937  (1991).
4843: 
4844: \bibitem{ha-93prb12459}
4845: Z.~N.~C. Ha and F.~D.~M. Haldane, Phys. Rev.~B {\bf 47},  12459  (1993).
4846: 
4847: \bibitem{essler95prb13357}
4848: F.~H.~L. E{\ss}ler, Phys. Rev.~B {\bf 51},  13357  (1995).
4849: 
4850: \bibitem{greiter-05prb224424}
4851: M. Greiter and D. Schuricht, Phys. Rev.~B {\bf 71},  224424  (2005).
4852: 
4853: \bibitem{greiter06prl}
4854: M. Greiter, Statistical Phases and Momentum Spacings for One-Dimensional
4855:   Anyons, {s}ubmitted to Phys. Rev. Lett.
4856: 
4857: \bibitem{greiter-06prl}
4858: M. Greiter and D. Schuricht, Many-spinon states and the secret significance of
4859:   Young tableaux, {s}ubmitted to Phys. Rev. Lett.
4860: 
4861: \bibitem{kawakami92prb1005}
4862: N. Kawakami, Phys. Rev.~B {\bf 46},  1005  (1992).
4863: 
4864: \bibitem{kawakami92prbr3191}
4865: N. Kawakami, Phys. Rev.~B {\bf 46},  R3191  (1992).
4866: 
4867: \bibitem{ha-92prb9359}
4868: Z.~N.~C. Ha and F.~D.~M. Haldane, Phys. Rev.~B {\bf 46},  9359  (1992).
4869: 
4870: \bibitem{bouwknegt-96npb345}
4871: P. Bouwknegt and K. Schoutens, Nucl. Phys.~B {\bf 482},  345  (1996).
4872: 
4873: \bibitem{schuricht-05epl987}
4874: D. Schuricht and M. Greiter, Europhys. Lett. {\bf 71},  987  (2005).
4875: 
4876: \bibitem{schuricht-06prb235105}
4877: D. Schuricht and M. Greiter, Phys. Rev. B {\bf 73},  235105  (2006).
4878: 
4879: \bibitem{jullien-83baps344}
4880: R. Jullien and F.~D.~M. Haldane, Bull. Am. Phys. Soc. {\bf 28},  344  (1983).
4881: 
4882: \bibitem{affleck89jpcm3047}
4883: I. Affleck, J. Phys.: Condens. Matter {\bf 1},  3047  (1989).
4884: 
4885: \bibitem{okamoto-92pla433}
4886: K. Okamoto and K. Nomura, Phys. Lett. A {\bf 169},  433  (1992).
4887: 
4888: \bibitem{eggert96prb9612}
4889: S. Eggert, Phys. Rev. B {\bf 54},  R9612  (1996).
4890: 
4891: \bibitem{white-96prb9862}
4892: S.~R. White and I. Affleck, Phys. Rev. B {\bf 54},  9862  (1996).
4893: 
4894: \bibitem{sen-07prb104411}
4895: D. Sen and N. Surendran, Phys. Rev. B {\bf 75},  104411  (2007).
4896: 
4897: \bibitem{knabe88jsp627}
4898: S. Knabe, J. Stat. Phys. {\bf 52},  627  (1988).
4899: 
4900: \bibitem{fannes-89epl633}
4901: M. Fannes, B. Nachtergaele, and R.~F. Werner, Europhys. Lett. {\bf 10},  633
4902:   (1989).
4903: 
4904: \bibitem{freitag-91zpb381}
4905: W.-D. Freitag and E. M\"uller-Hartmann, Z. Phys. B {\bf 83},  381  (1991).
4906: 
4907: \bibitem{kluemper-91jpal955}
4908: A. Kl\"umper, A. Schadschneider, and J. Zittartz, J. Phys. A {\bf 24},  L955
4909:   (1991).
4910: 
4911: \bibitem{kennedy-92prb304}
4912: T. Kennedy and H. Tasaki, Phys. Rev. B {\bf 45},  304  (1992).
4913: 
4914: \bibitem{kluemper-92zpb281}
4915: A. Kl\"umper, A. Schadschneider, and J. Zittartz, Z. Phys. B {\bf 87},  281
4916:   (1992).
4917: 
4918: \bibitem{kluemper-93epl293}
4919: A. Kl\"umper, A. Schadschneider, and J. Zittartz, Europhys. Lett. {\bf 24},
4920:   293  (1993).
4921: 
4922: \bibitem{batchelor-94ijmpb3645}
4923: M.~T. Batchelor and C.~M. Yung, Int. J. Mod. Phys. B {\bf 8},  3645  (1994).
4924: 
4925: \bibitem{schollwock-96prb3304}
4926: U. Schollw\"ock, T. Jolicoeur, and T. Garel, Phys. Rev. B {\bf 53},  3304
4927:   (1996).
4928: 
4929: \bibitem{kolezhuk-96prl5142}
4930: A. Kolezhuk, R. Roth, and U. Schollw\"ock, Phys. Rev. Lett. {\bf 77},  5142
4931:   (1996).
4932: 
4933: \bibitem{kolezhuk-02prb100401}
4934: A. Kolezhuk and U. Schollw\"ock, Phys. Rev. B {\bf 65},  100401(R)  (2002).
4935: 
4936: \bibitem{normand-02prb104411}
4937: B. Normand and F. Mila, Phys. Rev. B {\bf 65},  104411  (2002).
4938: 
4939: \bibitem{lauchli-06prb144426}
4940: A. L\"auchli, G. Schmid, and S. Trebst, Phys. Rev. B {\bf 74},  144426  (2006).
4941: 
4942: \bibitem{sutherland75prb3795}
4943: B. Sutherland, Phys. Rev.~B {\bf 12},  3795  (1975).
4944: 
4945: \bibitem{choi-82pla83}
4946: T.~C. Choy and F.~D.~M. Haldane, Phys. Lett.~A {\bf 90},  83  (1982).
4947: 
4948: \bibitem{Li-99prb12781}
4949: Y.~Q. Li, M. Ma, D.~N. Shi, and F.~C. Zhang, Phys. Rev.~B {\bf 60},  12781
4950:   (1999).
4951: 
4952: \bibitem{Gu-02prb092404}
4953: S.~J. Gu and Y.~Q. Li, Phys. Rev.~B {\bf 66},  092404  (2002).
4954: 
4955: \bibitem{damerau-06jsm12014}
4956: J. Damerau and A. Kl\"umper, J. Stat. Mech.  12014  (2006).
4957: 
4958: \bibitem{kawashima-07prl057202}
4959: N. Kawashima and Y. Tanabe, Phys. Rev. Lett. {\bf 98},  057202  (2007).
4960: 
4961: \bibitem{Paramekanti-06cm0608691}
4962: A. Paramekanti and J.~B. Marston, cond-mat/0608691.
4963: 
4964: \bibitem{affleck86npb409}
4965: I. Affleck, Nucl. Phys. B {\bf 265},  409  (1986).
4966: 
4967: \bibitem{affleck88npb582}
4968: I. Affleck, Nucl. Phys.~B {\bf 305},  582  (1988).
4969: 
4970: \bibitem{chen-05prb214428}
4971: S. Chen, C. Wu, S.~C. Zhang, and Y. Wang, Phys. Rev. B {\bf 72},  214428
4972:   (2005).
4973: 
4974: \bibitem{chen-06prb174424}
4975: S. Chen, Y. Wang, W.~Q. Ning, C. Wu, and H.~Q. Lin, Phys. Rev. B {\bf 74},
4976:   174424  (2006).
4977: 
4978: \bibitem{affleck-86lmp57}
4979: I. Affleck and E.~H. Lieb, Lett. Math. Phys. {\bf 12},  57  (1986).
4980: 
4981: \bibitem{greiter-07prb060401}
4982: M. Greiter, S. Rachel, and D. Schuricht, Phys.~Rev.~B {\bf 75},  060401(R)
4983:   (2007).
4984: 
4985: \bibitem{broek80pla261}
4986: P.~M. van~den Broek, Phys. Lett.~A {\bf 77},  261  (1980).
4987: 
4988: \bibitem{InuiTanabeOnodera96}
4989: T. Inui, Y. Tanabe, and Y. Onodera, {\em Group Theory and Its Applications in
4990:   Physics} (Springer, Berlin, 1996).
4991: 
4992: \bibitem{Cornwell84vol2}
4993: J.~F. Cornwell, {\em Group theory in physics} (Academic Press, London, 1984),
4994:   Vol.~II.
4995: 
4996: \bibitem{Georgi82}
4997: H. Georgi, {\em Lie Algebras in Particle Physics} (Addison-Wesley, Redwood
4998:   City, 1982).
4999: 
5000: \bibitem{schwinger65proc}
5001: J. Schwinger,  in {\em Quantum Theory of Angular Momentum}, edited by L.
5002:   Biedenharn and H. van Dam (Academic Press, New York, 1965).
5003: 
5004: \bibitem{Auerbach94}
5005: A. Auerbach, {\em Interacting electrons and quantum magnetism} (Springer, New
5006:   York, 1994).
5007: 
5008: \bibitem{greiter02prb134443}
5009: M. Greiter, Phys. Rev. B {\bf 65},  134443  (2002).
5010: 
5011: \bibitem{greiter02prb054505}
5012: M. Greiter, Phys. Rev. B {\bf 66},  054505  (2002).
5013: 
5014: \bibitem{macfarlane-68cmp77}
5015: A.~J. Macfarlane, A. Sudbery, and P.~H. Weisz, Commum. Math. Phys. {\bf 11},
5016:   77  (1968).
5017: 
5018: \bibitem{Humphreys87}
5019: J.~E. Humphreys, {\em Introduction to Lie Algebras and Representation Theory}
5020:   (Springer, New York, 1987).
5021: 
5022: \bibitem{affleck-91npb467}
5023: I. Affleck, D.~P. Arovas, J.~B. Marston, and D.~A. Rabson, Nucl. Phys. {\bf
5024:   B366},  467  (1991).
5025: 
5026: \bibitem{marston-89prb11538}
5027: J.~B. Marston and I. Affleck, Phys. Rev. B {\bf 39},  11538  (1989).
5028: 
5029: \bibitem{Jaksch-98prl3108}
5030: D. Jaksch, C. Bruder, J.~I. Cirac, C.~W. Gardiner, and P. Zoller, Phys. Rev.
5031:   Lett. {\bf 81},  3108  (1998).
5032: 
5033: \bibitem{Greiner-02n39}
5034: M. Greiner, O. Mandel, T. Esslinger, T.~W. H\"ansch, and I. Bloch, Nature {\bf
5035:   415},  39  (2002).
5036: 
5037: \bibitem{Honerkamp-04prl170403}
5038: C. Honerkamp and W. Hofstetter, Phys. Rev. Lett. {\bf 92},  170403  (2004).
5039: 
5040: \bibitem{Honerkamp-04prb094521}
5041: C. Honerkamp and W. Hofstetter, Phys. Rev. B {\bf 70},  094521  (2004).
5042: 
5043: \bibitem{rapp-06cm0607138}
5044: A. Rapp, G. Zar\'{a}nd, C. Honerkamp, and W. Hofstetter, cond-mat/0607138.
5045: 
5046: \bibitem{Duan-03prl090402}
5047: L.-M. Duan, E. Demler, and M.~D. Lukin, Phys. Rev. Lett. {\bf 91},  090402
5048:   (2003).
5049: 
5050: \bibitem{Garcia-04prl250405}
5051: J.~J. Garc\'{\i}a-Ripoll, M.~A. Martin-Delgado, and J.~I. Cirac, Phys. Rev.
5052:   Lett. {\bf 93},  250405  (2004).
5053: 
5054: \bibitem{micheli-06np341}
5055: A. Micheli, G.~K. Brennen, and P. Zoller, Nature Phys. {\bf 2},  341  (2006).
5056: 
5057: \bibitem{brennen-06qp0612180}
5058: G.~K. Brennen, A. Micheli, and P. Zoller, quant-ph/0612180.
5059: 
5060: \end{thebibliography}
5061: 
5062: \end{document}
5063: 
5064: