nlin0603028/BBS.tex
1: \documentclass[12pt]{iopart}
2: \usepackage{amssymb}
3: \usepackage[mathscr]{eucal}
4: %\usepackage{showkeys}
5: \usepackage{epic}
6: \usepackage{eepic}
7: \usepackage[active]{srcltx}
8: \vfuzz2pt \hfuzz2pt
9: %  Version 27.1.06
10: %\textheight 260mm \textwidth 160mm \voffset=-30mm %\hoffset=-20mm
11: %\oddsidemargin=-2mm \evensidemargin=-2mm
12: \usepackage{amsfonts}
13: \usepackage{bm}
14: 
15: \newtheorem{prop}{Proposition}  \newtheorem{remark}{Remark}
16: \newtheorem{defin}{Definition}
17: \newtheorem{lem}{Lemma}   \newtheorem{theorem}{Theorem}
18: 
19: \newcommand{\nc}{\newcommand}
20: \nc{\BBS}{{\rm BBS}\ } \nc{\g}{\gamma}  \nc{\lm}{\lambda}
21: \nc{\la}{\lambda} \nc{\bh}{{\bf h}}  \nc{\av}{\prod_{s\,\in\,\ZN}}
22: \nc{\cR}{{\cal R}} \nc{\kp}{{\varkappa}} \nc{\om}{\omega}
23: \nc{\qt}{\tilde{q}} \nc{\tp}{\tilde{p}} \nc{\rt}{\tilde{r}}
24: \nc{\ty}{\tilde{y}} \nc{\tx}{\tilde{x}} \nc{\tQ}{{\widetilde Q}}
25: \nc{\trh}{\tilde{\rho}}\nc{\ny}{\nonumber}\nc{\lk}{\left(}
26: \nc{\rk}{\right)} \nc{\Rb}{\right]} \nc{\Lb}{\left[}
27: \nc{\rb}{\right\}} \nc{\lb}{\left\{} \nc{\hs}{\hspace*{1cm}}
28: \nc{\hx}{\hspace*{3mm}} \nc{\hq}{\hspace*{6mm}} \nc{\JStP}{{\it J. Stat.
29: Phys.}}  \nc{\IJMP}{{\it Intern. J. Mod. Phys.}}
30: 
31: \nc{\al}{\alpha}  \nc{\ZN}{\mathbb{Z}_N}
32: \nc{\bg}{\boldsymbol{\gamma}} \nc{\bdr}{\boldsymbol{\rho}}
33: \nc{\bu}{{\bf u}} \nc{\bv}{{\bf v}} \nc{\bV}{{\bf V}}
34: \nc{\rnk}{r_{n,k}} \nc{\lns}{\la_{n,s}} \nc{\lnl}{\la_{n,l}}
35: \nc{\FD}{{\cal F}} \nc{\lnk}{\la_{n,k}} \nc{\xnl}{x_{n,l}}
36: \nc{\Psr}{\Psi_{\bdr_n}} \nc{\ap}{a_{n+1}} \nc{\bp}{b_{n+1}}
37: \nc{\cp}{c_{n+1}} \nc{\dpp}{d_{n+1}}
38: \nc{\bep}{\bu_{n+1}^{-1}(\ap-\bp\bv_{n+1})} \nc{\Xop}{\mathbf{X}}
39: \nc{\Zop}{\mathbf{Z}}  \nc\s{{\gamma}}
40: \def\r#1{(\ref{#1})}
41: 
42: \nc{\ra}{\rangle} \nc{\BAR}{\begin{array}} \nc{\EAR}{\end{array}}
43: \nc{\bdm}{\begin{displaymath}} \nc{\edm}{\end{displaymath}}
44: \nc{\be}{\begin{equation}} \nc{\ee}{\end{equation}}
45: \nc{\ba}{\begin{array}} \nc{\ea}{\end{array}}
46: \nc{\bea}{\begin{eqnarray}} \nc{\eea}{\end{eqnarray}}
47: 
48: \nc\sip{\gamma^\prime}\nc\ma{a'}\nc\mb{b'}\nc\mc{c\,'}\nc\md{d\,'}
49: \nc\pra{a''}\nc\prb{b''}\nc\prc{c\,''}\nc\prd{d\,''} \nc{\hu}{{\bf
50: u}}\nc{\hh}{{\hat h}}\nc{\bl}{\boldsymbol{\lambda}}\nc{\hv}{{\bf v}}
51: \nc\si{{\mathrm{s}}}
52: 
53: 
54: \begin{document}
55:  \title{Baxter--Bazhanov--Stroganov model: Separation of Variables and Baxter Equation}
56: \author{G von Gehlen$^\dag$,~ N Iorgov$^\ddag$,~ S Pakuliak$^{\sharp\flat}$~
57:  and V Shadura$^\ddag$}
58: \address{$^\dag$\ Physikalisches Institut der Universit\"at Bonn,
59: Nussallee 12, D-53115 Bonn, Germany}
60: \address{$^\ddag$ Bogolyubov Institute for Theoretical Physics, Kiev 03143,
61: Ukraine}
62: \address{$^\sharp$\ Bogoliubov Laboratory of Theoretical Physics,
63: Joint Institute for Nuclear Research, Dubna 141980, Moscow region,
64: Russia}
65: \address{$^\flat$\ Institute of Theoretical and Experimental Physics,
66: Moscow 117259, Russia}
67: \ead{gehlen@th.physik.uni-bonn.de, iorgov@bitp.kiev.ua, pakuliak@theor.jinr.ru,
68: shadura@bitp.kiev.ua}
69: %
70: \begin{abstract}
71: The Baxter-Bazhanov-Stroganov model (also known as the $\tau^{(2)}$ model)
72: has attracted much interest because it provides a tool for solving the integrable
73: chiral $\ZN$-Potts model. It can be formulated as a face spin model or via cyclic
74: $L$-operators. Using  the latter formulation and the
75: Sklyanin-Kharchev-Lebedev approach, we give the explicit derivation of
76: the eigenvectors of the component $B_n(\la)$ of the monodromy
77: matrix for the fully inhomogeneous chain of finite length. For the periodic chain
78: we obtain the Baxter T-Q-equations via separation of variables.
79: The functional relations for the transfer matrices of the $\tau^{(2)}$ model
80: guarantee non-trivial solutions to the Baxter equations.
81: For the $N=2$ case, which is free fermion point of a generalized Ising model,
82: the Baxter equations are solved explicitly.
83: \end{abstract}
84: \hspace*{2.5cm}{\small \today}\hspace*{12mm}   \submitto{\JPA}
85: \vspace*{-12mm} \pacs{75.10Hk, 75.10Jm, 05.50+q, 02.30Ik}
86: 
87: 
88: \section{Introduction}
89: The aim of this paper is the explicit construction of eigenvectors
90: of the transfer-matrix for the finite-size inhomogenous periodic
91: Baxter--Bazhanov--Stroganov model (\BBS model) also known as the
92: $\tau^{(2)}$-model \cite{B_tau,BaxInv,BS,BBP}. This is a $N$-state spin
93: lattice model, intimately related to the integrable chiral Potts model.
94: The connection between the 6-vertex model, the \BBS model
95: and the chiral Potts model gives the possibility
96: to formulate a system of functional relations \cite{BS,BBP} for the transfer
97: matrices of these models. Solving these systems is the basic method for
98: calculating the eigenvalues of the transfer matrix of the chiral Potts
99: model \cite{BaxtEV}, and under some analyticity assumptions, to derive the
100: free energy of this model \cite{BaxtEV} and its order parameter \cite{BaxtOP}.
101: 
102: 
103: 
104: In general for the \BBS model there is no Bethe pseudovacuum state and so the
105: algebraic Bethe ansatz cannot be used. Therefore, in order to achieve our goal, we shall use the
106: formulation of the BBS model in terms of cyclic $L$-operators
107: first introduced by Korepanov \cite{Kore} and Bazhanov-Stroganov \cite{BS}
108: and adapt the Sklyanin--Kharchev--Lebedev method of
109: separation of variables (SoV) \cite{Skly1,KarLeb1,KarLeb2,KhLS}
110: for solving the BBS eigenvector problem. The fusion equations will
111: provide the existence of solutions to the Baxter equations.
112: 
113: The paper is organized as follows. After defining the \BBS model as a statistical face spin model,
114: we give the vertex formulation of the model in terms of a cyclic $L$-operator
115: and conclude the Introduction explaining the two basic steps involved in the SoV
116: method. Section 2 deals with the solution of the auxiliary problem,
117: leaving for Section 3 the lengthy inductive proof of the main
118: formula. The short Section 4 derives the action of the diagonal component $D$ of the
119: $L$-operator on the auxiliary eigenstates. Then in Section 5 we come
120: to the periodic model and in deriving the Baxter T-Q-equations
121: we show the role of the fusion equations for solving these Baxter
122: equations. In Section 6 we apply these results to the homogenous $N=2$ case and calculate the
123: eigenvalues and eigenvectors of the homogenous $N=2$ \BBS
124: model which is the free fermion point of  a generalized Ising model. Section 7 gives our
125: conclusions. In an Appendix we show a strong simplification
126: occurring if the \BBS model is homogenous.
127: 
128: 
129: \subsection{The \BBS model}
130: 
131: Following the notation of a recent paper of Baxter \cite{B_tau}, we define the BBS model
132: as a statistical model of short-range interacting spins placed at the vertices of a
133: rectangular lattice. We label the spin variables
134: $\si_{x,y}$  by a pair $(x,y)$ of integers: $x=1,\ldots,n,$ and
135: $y=1,\ldots,m $. Each spin variable $\si_{x,y}$ takes  $N$ values ($N\ge 2$): $0$,
136: $1,\ldots$, $N-1$. The model shall have $\mathbb{Z}_N$-symmetry and
137: we may extend the range of the spins $\si_{x,y}$ to all integers identifying
138: two values if their difference is a multiple of $N$. The model has a chiral restriction on the
139: values of vertically neighboring spins:
140:  \begin{equation}
141:  \label{adj-rule}
142:   \si_{x,y}-\si_{x,y+1}=0 \mbox{\ or\ } 1 \quad {\rm mod}\ N\ .
143: \end{equation}
144: In the following we will consider the spin variables on two adjacent rows:
145: $(k,\,l)$ and  $(k,\,l+1)$, where $l$ is fixed and $k=1,\ldots,n$.
146: Let us denote $\si_{k,\,l}=\g_{k}$ and $\si_{k,\,l+1}=\g^\prime_{k}$.
147: The model depends on the parameters $t_q$ and  $\ma_k\,, \mb_k\,,
148: \mc_k\,, \md_k$, $\pra_k\,,\prb_k\,, \prc_k\,, \prd_k$,
149: $k=1,2,\ldots,n$. Each square plaquette of the lattice has the
150: Boltzmann weight (see Fig.\ref{tri})
151: \be\label{bbs_weight}
152:  W_\tau( \g_{k-1},
153: \g_{k};\sip_{k-1}, \sip_{k}) =  \sum_{m_{k-1}=0}^1
154: \omega^{m_{k-1}(\sip_{k} - \s_{k-1})} (-\omega t_q)^{ \s_{k} -
155: \sip_{k}- m_{k-1}}\times \ee
156: \[
157: \qquad \qquad \qquad\qquad \times F'_{k-1}( \s_{k-1}- \sip_{k-1},
158: m_{k-1})
159:  F''_{k}( \s_{k}- \sip_{k}, m_{k-1}),
160:  \]
161: where $\omega=e^{2\pi{\rm i}/N}$, and
162: \[\fl
163: F'_{k}(0,0)=1, \qquad F'_{k}(0,1)=-\omega t_q\,\frac{\mc_k}{\mb_k},
164: \qquad F'_{k}(1,0) = \frac{\md_k}{\mb_k},\qquad F'_{k}(1,1)=-\omega
165: \frac{\ma_k}{\mb_k},
166: \]
167: and expressions for $ F''_{k}( \s_{k}- \sip_{k}, m_{k-1})$ are
168: obtained from $ F'_k ( \s_{k}- \sip_{k}, m_{k})$ by substitutions:
169: $\ma_k\,,$ $\mb_k\,,$ $\mc_k\,,$ $\md_k \rightarrow
170:  \pra_k\,,$ $\prb_k\,, $ $\prc_k\,,$ $ \prd_k\,$.
171: 
172: \begin{figure}[ht]
173: \begin{center}
174: \renewcommand{\dashlinestretch}{30}
175: \unitlength=0.04pt
176: \begin{picture}(8622,4539)(0,-10)
177: \drawline(2715,3612)(5415,3612)(5415,912)
178:     (2715,912)(2715,3612)
179: \drawline(15,3612)(2715,3612)(2715,4512)
180: \drawline(5415,4512)(5415,3612)(8115,3612)
181: \drawline(15,912)(2715,912)(2715,12)
182: \drawline(5415,12)(5415,912)(8115,912)
183: \dashline{60.000}(15,3612)(2715,912)(5415,3612)(8115,912)
184: \dashline{60.000}(15,912)(2715,3612)(5415,912)(8115,3612)
185: \put(4600,2217){\makebox(0,0)[cc]{$m_{k-1}$}}
186: \put(7200,2217){\makebox(0,0)[cc]{$m_k$}}
187: \put(1900,2217){\makebox(0,0)[cc]{$m_{k-2}$}} \put(3100,462)
188: {\makebox(0,0)[cc]{$\g_{k-1}$}} \put(5685,462)
189: {\makebox(0,0)[cc]{$\g_k$}}
190: \put(3100,4017){\makebox(0,0)[cc]{$\g'_{k-1}$}}
191: \put(5730,4017){\makebox(0,0)[cc]{$\g'_k$}} \put  (15,912)
192: {\makebox(0,0)[cc]{$\bullet$}}
193: \put(2715,3612){\makebox(0,0)[cc]{$\bullet$}} \put(2715,912)
194: {\makebox(0,0)[cc]{$\bullet$}}
195: \put(5415,3612){\makebox(0,0)[cc]{$\bullet$}} \put(5415,912)
196: {\makebox(0,0)[cc]{$\bullet$}}
197: \put(8115,3612){\makebox(0,0)[cc]{$\bullet$}} \put(8115,912)
198: {\makebox(0,0)[cc]{$\bullet$}} \put  (15,3612)
199: {\makebox(0,0)[cc]{$\bullet$}}
200: \put(1365,2262){\makebox(0,0)[cc]{$\bullet$}}
201: \put(4065,2262){\makebox(0,0)[cc]{$\bullet$}}
202: \put(6765,2262){\makebox(0,0)[cc]{$\bullet$}}
203: \end{picture}
204: \end{center}
205: \caption{\footnotesize{The triangle with vertices markes by the spin
206: variables $\g_{k-1}$, $\g'_{k-1}$, $m_{k-1}$ correspond to the
207: function $F'_{k-1}(\g_{k-1}-\g'_{k-1},m_{k-1})$ in (\ref{bbs_weight});
208: the triangle $\g_{k}$, $\g'_{k}$, $m_{k-1}$ to
209: $F''_{k}(\g_{k}-s'_{k},m_{k-1})$. }} \label{tri}\end{figure}
210: 
211: We will consider the periodic boundary condition: $\s_{n+1}=\s_{1}$,
212: $\s'_{n+1}=\s'_{1}$, where $n$  is the number of sites on the
213: lattice along the horizontal axis. The transfer-matrix of the
214: periodic \BBS model is $N^{n}\times N^{n}$ matrix with matrix elements
215: \begin{equation}
216: {\bf t}_{n}({\bg},{\bg}')=\prod_{k=2}^{n+1}W_\tau(
217: \s_{k-1},
218:    \s_{k};\sip_{k-1}, \sip_{k}),
219:  \label{tm}
220: \end{equation}
221: labelled by the sets of spin variables ${\bg}=\{\s_1,\s_2,\ldots,\s_n\}$
222:  and ${\bg}'=\{\s'_1,\s'_2,\ldots,\s'_n\}$ of two neighbour rows.
223: 
224: 
225: Considering $m_k$, $k=1,\ldots,n$, in
226: (\ref{bbs_weight}) as auxiliary spin variables which take the two
227: values $0$ and $1$, we can rewrite transfer-matrix \r{tm} in a vertex
228: formulation associating a statistical weight not to the plaquettes but
229: to vertices each of them relating four spins: $m_{k-1}$, $m_k$,
230: $\s_{k}$, $\sip_{k}$ (see Fig.~1). Then the weight associated with
231: the $k$th vertex is
232: \[
233:  \ell_k(t_q;m_{k-1}, m_k;\s_{k}, \sip_{k})=
234: \omega^{m_{k-1}\sip_k-m_{k}\s_k}(-\omega t_q) ^{\s_{k}-
235: \sip_{k}-m_{k-1}}\times
236: \]
237: \be
238:  \label{tau_2} \,\qquad \qquad \qquad \qquad  \times F''_{k}(
239: \s_{k}- \sip_{k}, m_{k-1}) F'_{k }( \s_{k }- \sip_{k }, m_{k}).
240: \ee
241: and  the transfer-matrix (\ref{tm}) can be rewritten as
242: \be\label{tm_1}
243: {\bf t}_{n}({\bg},{\bg}')=\sum_{m_1, ...,\, m_n} \prod_{k=2}^{n+1}
244: \ell_k(t_q;m_{k-1}, m_k;\s_{k}, \sip_{k}).
245: \ee
246: 
247: \subsection{The $L$-operator formulation of  \BBS model}
248: 
249: For our construction of the \BBS model eigenvectors we will
250: use a description of  this model
251: as a quantum chain model as introduced in \cite{Kore} and \cite{BS}.
252: To the each site $k$ of the quantum chain we associate
253: the cyclic $L$-operator acting in a two-dimensional auxiliary space
254: \\[-6mm]
255: \be
256: \label{bazh_strog}L_k(\lm)=\left( \ba{ll}
257: 1+\lm \kp_k \hv_k,\ &  \lm \hu_k^{-1} (a_k-b_k \hv_k)\\ [3mm]
258: \hu_k (c_k-d_k \hv_k) ,& \lm a_k c_k + \hv_k {b_k d_k}/{\kp_k }
259: \ea \right)\!\!,\hspace*{4mm} k=1,2,\ldots,n.
260: \ee
261: At each site $k$ we define ultra-local Weyl elements $\hu_k$ and $\hv_k$
262: obeying the commutation rules and normalization
263:  \be\label{comrel_uv}\fl \bu_j
264: \bu_k=\bu_k \bu_j\,, \quad\;\bv_j \bv_k=\bv_k
265: \bv_j\,,\quad\; \bu_j
266: \bv_k=\om^{\delta_{j,k}}\bv_k\bu_j\,,\quad\; \om=e^{2\pi i/N},
267: \quad \bu_k^N=\bv_k^N=1.
268: \ee
269: In \r{bazh_strog},
270: $\lm$ is the spectral parameter and we have five
271: parameters $\:\kp_k,\;a_k,\;b_k,\;c_k,\;d_k\,$ per site. At each site $k$ we
272: define a $N$-dimensional linear space (quantum space) ${\cal V}_k$
273: with the basis $ |\g\ra_k$, $\g\in \ZN$ and natural
274: scalar product $\: _k\langle\g'|\g\ra_k=\delta_{\g',\g}.$ In
275: ${\cal V}_k$ the Weyl elements $\bu_k$ and $\bv_k$ act by the
276: formulas:
277: \be\label{uv} \bu_k |\g\ra_k=\om^\g |\g\ra_k\, ,\qquad
278: \bv_k |\g\ra_k = |\g+1\ra_k\, . \ee
279: 
280: 
281: The correspondence between the lattice \BBS model and its quantum chain analog is established
282: through the relation
283: \begin{equation}\label{identific}
284: \ell_k(t_q; m_{k-1}, m_k;\s_{k}, \sip_{k})= \, _k\langle\g| L_k(\lm)_{m_{k-1},m_k}|\g'\ra_k
285: \end{equation}
286: and the following connection between the parameters of these models
287: \[\fl
288: \lambda=-\omega t_q\,,\quad
289: \kp_k=\frac{\md_k}{\mb_k} \frac{\prd_k}{\prb_k} \,,\quad
290: a_k=\frac{\prc_k}{\prb_k}\,,\quad b_k=\omega
291:       \frac{\pra_k}{\prb_k}\frac{\md_k}{\mb_k}\, ,\quad
292: c_k=\frac{\prc_k}{\prb_k}\,,\quad
293: d_k=\frac{\pra_k}{\prb_k}\frac{\md_k}{\mb_k}\,.
294: \]
295: We extend the action of the
296: operators $\bu_k,\;\bv_k\:$ to ${\cal V}^{(n)}={\cal V}_1\otimes{\cal
297: V}_2\otimes\cdots\otimes{\cal V}_n$ defining this action to be
298: trivial in all $\:{\cal V}_s\,$ with $\,s\ne k$.
299: The monodromy matrix for the quantum chain with $n$ sites
300: is defined as \be \label{mm}
301: {T}_n(\lm)\,=\,L_1(\lm)\,L_2(\lm)\,\cdots\, L_n(\lm)= \lk \ba{ll}
302: A_n(\lm)& B_n(\lm)\\ [2mm]
303: C_n(\lm)& D_n(\lm) \ea \rk. \ee  The transfer-matrix (\ref{tm_1}) is
304: obtained taking the trace in the auxiliary space
305:  \be \label{tr_matl} {\bf t}_{n}(\lambda)\:=\:\tr\:
306: T_n(\lm)\:=\:A_n(\lm)+D_n(\lm)\,.
307: \ee
308: This  quantum chain is
309: integrable because the $L$-operators \r{bazh_strog} are
310: intertwined by the twisted 6-vertex $R$-matrix at root of unity
311: \be
312: {R}(\la,\nu)\;=\;\lk\begin{array}{cccc} \la-\om\nu & 0 & 0 & 0 \\
313: [0.3mm] 0 & \om(\la-\nu) & \la (1-\om) & 0 \\ [0.3mm] 0 & \nu (1-\om)
314: & \la-\nu & 0 \\ [0.3mm] 0 & 0 & 0 & \la-\om\nu \end{array}\rk\!,
315: \ee
316: \be R(\la,\nu)\,  L^{(1)}_k\, (\la) L^{(2)}_k(\nu)=
317: L^{(2)}_k(\nu)\, L^{(1)}_k(\la)\, R(\la,\nu), \label{rll} \ee
318: where
319: $\:L^{(1)}_k(\la)= L_k(\la)\otimes\mathbb{I}$,
320: $L^{(2)}_k(\la)=\mathbb{I} \otimes L_k(\la)$. Relation
321: (\ref{rll}) leads to  $\:[{\bf t}_{n}(\lambda),{\bf t}_{n}(\mu)]=0\;$
322: and so $\;{\bf t}_n(\la)\,$ is the generating
323:  function for the commuting set of non-local and
324: non-hermitian Hamiltonians ${\bf H}_0,\ldots,{\bf H}_n$
325: of the model:
326: \be \label{tr_mat_bs} {\bf t}_n(\lm) \:=\:{\bf
327: H}_0\,+{\bf H}_1\lm\,+\cdots\,+{\bf H}_{n-1}\lm^{n-1}\,+{\bf
328: H}_{n}\lm^{n}.
329: \ee
330: The lowest and highest Hamiltonians can be easily written explicitly in
331: terms of the global $\ZN$-charge rotation operator $\;\bV_n$
332: \be \label{h0}\fl {\bf
333: H}_0\,=\,1\,+\bV_n\prod_{k=1}^n\frac{b_k d_k}{\kp_k};\qquad
334: {\bf H}_n\,=\,\prod_{k=1}^n\: a_k
335: c_k\,+\,\bV_n\prod_{k=1}^n \kp_k;\qquad
336: \bV_n\;=\;\bv_1 \bv_2 \cdots \bv_n. \label{Znc}
337: \ee
338: It also follows from the intertwining relation (\ref{rll}) that $B_n(\lm)$
339: is the generating function for another commuting set of operators
340: ${\bf h}_1,\ldots,{\bf h}_n$: \be\label{Bpol} \left[ B_n(\lambda),
341: B_n(\mu)\right]=0,\qquad B_n(\lm)= {\bf h}_1 \lm + {\bf h}_2 \lm^2
342: +\cdots + {\bf h}_{n}\lm^{n}\,  . \ee
343: 
344: 
345: 
346: The great interest in the \BBS chain model is due to its relation to the
347: integrable chiral Potts model. In \cite{BS,Kore} it was observed that besides the
348: intertwining relations \r{rll},
349: the $L$-operators  \r{bazh_strog} satisfy a second intertwining relation
350: in the Weyl quantum space:\\[-4mm]
351: \begin{eqnarray}
352: \lefteqn{\sum_{\beta_1\beta_2,j}\!
353:    \mathsf{S}_{\al_1\al_2;\beta_1\beta_2}(p,p',q,q')
354:    \;L_{i_1j}^{\beta_1\gamma_1}(\la;p,p')\;L_{j\,i_2}^{\beta_2,\gamma_2}
355:    (\la;q,q')}\ny\\ [-2mm] \hs\hx=\;\sum_{\beta_1\beta_2,j}\!\!\!
356:     \; L_{i_1j}^{\al_2\beta_2}(\la;q,q')\;L_{j\,i_2}^{\al_1\beta_1}
357:     (\la;p,p')\;\:\mathsf{S}_{\beta_1,\beta_2;\gamma_1,\gamma_2}(p,p',q,q'),
358:     \label{SLL}\end{eqnarray}
359: if the parameters are chosen as
360: \be \fl \kp_k=\frac{y_{q_k}
361: y_{q'_k}}{\mu_{q_k} \mu_{q'_k}};\qquad a_k=x_{q_k};\qquad
362: b_k=\frac{y_{q'_k}}{\mu_{q_k} \mu_{q'_k}};\qquad c_k=\om
363: \,x_{q'_k};\qquad d_k=\frac{y_{q_k}}{\mu_{q_k} \mu_{q'_k}},
364: \label{cp_param}\ee
365: where $x_{q_k},\;y_{q_k},\;\mu_{q_k}$
366: (analogously for $x_{q'_k}$ etc.) satisfy the chiral Potts model
367: constraints \be \fl x_{q_k}^N+\,y_{q_k}^N=\mbox{\sf
368: k}\,(x_{q_k}^Ny_{q_k}^N+1);\quad
369:   \mbox{\sf k}\,x_{q_k}^N=1-\mbox{\sf k}'\,\mu_{q_k}^{-N};\quad
370:   \mbox{\sf k}\,y_{q_k}^N=1-\mbox{\sf k}'\,\mu_{q_k}^N;\quad
371:   \mbox{\sf k}^2+{\mbox{\sf k}'}^2=1. \label{cpcon}\ee
372: Here $\mbox{\sf k}\,$ and $\,\mbox{\sf k}'\,$ are temperature parameters.
373: In \r{SLL} we have written the spin matrix elements of the
374: $L$-operators with the parametrization \r{cp_param} as
375: $\;L_{i,j}^{\al,\beta}(\la,q,q')\;$ where $i,j=0,1$ are the
376: components in the auxiliary space and greek indices
377: $\al,\beta=0,\ldots,N-1$ denote the components in the quantum space
378: \r{uv}, suppressing the site index $k$. The matrix $\:\mathsf{S}\:$
379: turns out to be the product of four chiral Potts--Boltzmann weights
380: \cite{BS} \be \fl\mathsf{S}_{\al_1\al_2,\beta_1\beta_2}(p,p',q,q')
381: \;=\;W_{pq'}(\al_1-\al_2)
382:    W_{p'q}(\beta_2-\beta_1)\bar{W}_{pq}(\beta_2-\al_1)\bar{W}_{p'q'}
383:    (\beta_1-\al_2).\label{BSS}\ee
384: In the parametrization \r{cp_param} of the \BBS model there are
385: various functional relations to the chiral Potts model transfer
386: matrix which have been used to obtain explicit solutions for the
387: chiral Potts eigenvalues \cite{BBP}. Only further restricting the
388: parameters to the ``superintegrable chiral Potts'' case:
389: \be
390: a_k\;=\;\om^{-1}\,b_k\;=\;c_k\;=\;d_k\;=\;\kp_k\;=\:1 \label{super}
391: \ee
392: allows to solve the \BBS model by algebraic Bethe ansatz, see e.g.
393: \cite{AMCP,TarasovL,BaxSkew}.
394:  In this form  Baxter in 1989  first obtained the \BBS model as
395: an ``inverse'' of the superintegrable chiral Potts model, see
396: Equations (8.13),(8.14) of \cite{BaxInv}.
397:  In this paper we shall
398: follow \cite{B_tau} in not using restrictions like \r{cpcon} on
399: the parameters $\kp_k,$ $a_k,$ $b_k,$ $c_k,$ $d_k$. We only shall exclude
400: the superintegrable case \r{super}.
401: 
402: It was shown in the paper \cite{Bugrij} that the $N=2$ \BBS model is equivalent
403: to the generalized Ising model at the free fermion point. The results of our paper
404: permit to obtain the transfer-matrix eigenvectors for this model.  Recently,
405: the interesting
406: paper \cite{Lisovy} appeared, where  these eigenvectors was constructed  using
407: the Grassmann functional integral.
408: 
409: \subsection{Functional Bethe ansatz and SoV}\label{SoV}
410: The construction of common eigenvectors of the set commuting
411: integrals \r{tr_mat_bs} will be solved in two main steps, which
412: generally can be formulated as follows:
413: 
414: First, for the given quantum integrable chain type model one has to
415: find an {\it auxiliary} integrable model such that: 1) the
416: eigenvectors for original model can be expressed as a linear
417: combination of the eigenvectors for the auxiliary model
418:  and 2) the coefficients of this decomposition should factorize into
419:  products of the single variable functions (phenomenon of ``separation of variables").
420: 
421: Second, the auxiliary problem should be chosen in such a way that
422: the construction of its eigenvectors is a simple iteration process:
423: eigenvectors for the auxiliary model of  size $n$ have to be
424: obtained from the eigenvectors for the model of size $n-1$.
425: 
426: An example of realization of the first step in case of the Toda
427: chain model was proposed in the paper \cite{Gu81}. The auxiliary
428: model for the periodic Toda chain was the open Toda chain. In the
429: paper \cite{Sk85} on the Toda example this approach has been
430: formalized as  ``the functional Bethe ansatz" \cite{Skly1}. A complete
431: realization of this step for the periodic Toda chain model can be
432: found in the paper \cite{KarLeb1}.
433: 
434: Regarding  a recurrent procedure for
435: eigenvectors of the auxiliary problem, probably the first reference
436: to this possibility may be found in the series of lectures
437: \cite{Skly2}. The main idea of this approach can be formulated as
438: follows. Consider an integrable quantum chain model of size $n$. The
439: monodromy matrix is the product of $n$ $L$-operators. We decompose a
440: system into two subsystem of the sizes $n_1$ and $n_2$ such that
441: $n=n_1+n_2$. Suppose we can solve the eigenvalue and eigenvector
442: problems for the subsystems. In  \cite{Skly2} E.~Sklyanin
443: claims that there is a relation between the eigenvectors of the
444: original system and the eigenvectors of its smaller subsystems.
445: For the open Toda chain model this Sklyanin approach has been realized in the
446: paper \cite{KarLeb2}.
447: 
448: In our case of the \BBS chain model, the auxiliary model is governed by the
449: set of commuting integrals \r{Bpol}. So we first solve the problem
450: of finding the eigenvectors for these integrals. Then we shall show
451: that  the eigenvectors for the operators \r{tr_mat_bs} can be
452: constructed  as linear combinations of the eigenvectors of the set
453: \r{Bpol}. The multi-variable coefficients of this decomposition
454: admit the separation of variables and can be written as products of
455: single variable functions, each satisfying a Baxter difference
456: equation. We shall obtain this Baxter equation for generic $N$ and
457: solve it explicitly for $N=2$ corresponding to the free fermion
458: point of the generalized Ising model \cite{Bugrij}.
459: Note that   the eigenvectors of
460: the commuting set of operators which come from the generating polynomial
461:  $A_n(\lambda)$ ($[A_n(\lambda),A_n(\mu)]=0$) were found in paper \cite{Iorgov}.
462: 
463: \section{Eigenvectors of $B_n(\lm)$.}
464: \subsection{Consequences of the $RTT$ relations}
465: We start with the second step of the program described in Section \ref{SoV}.
466: Following \cite{Skly1,Skly2,KarLeb2},
467: we first construct eigenvectors of  $B_n(\lm)$.
468: According to (\ref{Bpol}), any common eigenvector of the commuting set of
469: operators $\,{\bf h}_1,\ldots,{\bf h}_{n}\,$
470: is eigenvector of $B_n(\lm)$ and the eigenvalue is a polynomial in $\lm$.
471: Factorizing this polynomial we get\\[-6mm]
472: \be\label{eig_b}
473: B_n(\lambda)\,\Psi_{\bl}\: =\:\lm\lm_0\prod_{k=1}^{n-1}\,(\lm-\lm_k)\,\Psi_{\bl};
474: \qquad {\bl}=(\lambda_0, \lambda_1,\ldots,\lambda_{n-1}),\ee
475: where we labelled the eigenvector $\Psi_{\bl}$ by the normalizing factor $\lm_0$
476: and the $n-1$ non-vanishing zeros $\lm_1,\lm_2,\ldots,\lambda_{n-1}$ of
477: the eigenvalue polynomial.
478: 
479: Now the intertwining relations \r{rll} tell us
480: that if $\Psi_{\bl}$ is eigenvector of $B_n(\lm)$, then by applying repeatedly
481: the operators $A(\lm_j)$ and $D(\lm_k)$  $(j,k=1,\ldots,n-1)$ and $\bV_n$ \r{Znc},
482: we can generate a whole set of $N^n$ eigenvectors of $B_n(\lm)$.
483: 
484: The   intertwining relations (\ref{rll}) give
485: \bea\fl\label{BA}
486: (\lambda-\omega\mu)A_n(\lambda)B_n(\mu)&=&\omega(\lambda-\mu)
487: B_n(\mu)A_n(\lambda)+\mu(1-\omega)A_n(\mu)B_n(\lambda)\\[1mm]
488: \fl(\lambda-\omega\mu)D_n(\mu)B_n(\lambda)&=&
489: \omega(\lambda-\mu)B_n(\lambda)D_n(\mu)+\lambda(1-\omega)D_n(\lambda)B_n(\mu)\label{BD}.\eea
490: Fixing $\;\lm=\lm_k$, $\;k=1,\ldots, n-1\:$ in  (\ref{BA}) and
491: acting by it on $\Psi_{\bl}$ we obtain
492: \be
493: B_n(\mu)\left(A_n(\lambda_k)\Psi_{\bl} \right)=\mu\lambda_0
494: ( \mu-{\omega}^{-1} {\lambda_k}) {\textstyle \prod_{s\neq k}} ( \mu-\lambda_s)\:
495: \left(A_n(\lm_k)\Psi_{\bl}\right)\,.\ee
496: This means that up to a constant
497: \be\label{alm}
498: A_n(\lambda_k)\Psi_{\bl} \sim \Psi_{\lambda_{\,0}, \ldots ,\,
499:  {\omega}^{-1} \lambda_k , \ldots ,\, \lambda_{n-1} }\,.
500: \ee
501: Similarly, from \r{BD} we can get
502: \be\label{dlm}
503: D_n(\lambda_k)\Psi_{\bl} \sim \Psi_{ {\omega}^{-1}{\lambda_{\,0}} , \ldots ,
504:  \,{\omega}{ \lambda_k} , \ldots , \,\lambda_{n-1} }.
505: \ee
506: (The proportional factors will be obtained later in \r{Almk} and \r{Dlmk}).
507: Furthermore, acting by (\ref{BA}) on $\,\Psi_{\bl}$ and extracting
508: coefficients of $\:\lm^{n+1}\mu^n\,$ we have
509: \be\label{vvv}
510: \bV_n \Psi_{\bl}\sim \Psi_{ {\omega}^{-1}{\lambda_{\,0}} , \ldots ,
511:  \,{ \lambda_k} , \ldots , \,\lambda_{n-1} }\,.
512: \ee
513: 
514: We see that the operators $A_n(\lm)$ and $D_n(\lm)$
515: at the eigenvalue zeros $\lm_k$ of $B_n(\lm)$, together with the charge rotation operator
516: $\bV_n=\bv_1 \bv_2 \cdots \bv_n$ act as cyclic ladder operators on the eigenvectors
517: of $B_n(\lm)$.
518: So the eigenvalues of $B_n(\lm)$ can be written
519: \be\label{iBlm}
520: B_n(\lm)\Psi_{\bdr_n} =\lm\, r_{n,0}\,\om^{-\rho_{n,0}}
521: \prod_{k=1}^{n-1}\lk\lm+r_{n,k}\om^{-\rho_{n,k}}\rk\Psi_{\bdr_n}\,,
522: \ee
523: where $r_{n,s}$, $s=0,\ldots,n-1$ is a set of constants
524: and we shall use the phases
525: \be    \bdr_n=(\rho_{n,0},\ldots,\rho_{n,n-1})\in(\ZN)^n\label{br}\ee
526: as new labels of the eigenvectors.
527: The fact that the eigenvectors of the operator
528: $B_n(\lambda)$ can be considered as dependent only on the integer phases
529: of the roots
530: \be\label{roots}
531: \lambda_{n,k}=-r_{n,k}\om^{-\rho_{n,k}}
532: \ee is a common property
533: of the root of unity integrable models. The amplitudes $r_{n,k}$ of these roots are
534: fixed by some``classical" procedure
535: which will be described below. In some cases this procedure becomes
536: an classical integrable system naturally incorporated into the quantum system (see
537: \cite{GPS} and references therein).
538: 
539: \subsection{One- and two-site eigenvectors for the auxiliary problem}
540: 
541: We now start to solve the auxiliary problem, which is
542: to compute the eigenvectors of $B_n(\lm)$ in the basis ${\cal V}^{(n)}$.
543: We shall adapt the recursive procedure of Kharchev and
544: Lebedev \cite{KarLeb2} to the \BBS chain model.
545: 
546: In our root of unity case a very important role will be played by the
547: cyclic function $\:w_p(\g)\:$ \cite{BB} which
548: depends on a $\ZN$-variable $\g$ and on a point $p=(x,y)$
549: restricted to the Fermat curve $x^N+y^N=1$. We define $w_p(\g)$ by the difference equation
550: \be  \fl
551: \frac{w_p(\g)}{w_p(\g-1)}\:=\:\frac{y}{1\,-\,\om^\g\,x}\,;\qquad
552: x^N\,+\,y^N\,=\,1\,;\qquad \g\in\ZN;\hq\quad w_p(0)=1\,. \label{Fermat} \ee
553: %
554: The Fermat curve restriction guarantees the cyclic property $\;w_{p}(\g+N)=w_{p}(\g).$
555: The function $\:w_p(\g)\:$ is a root of unity analog of the $q$-gamma function.
556: 
557: It is convenient to change the bases in the spaces ${\cal V}_k$.
558: Instead of $|\g\ra_k$, $\g\in \ZN$, we will use the vectors
559: \be\label{psik}
560: \psi^{(k)}_{\rho}=
561: \sum_{\g\in \ZN} w_{p_k}(\g-\rho)|\g\ra_k,\qquad \rho\in \ZN\,,
562: \ee
563: which are eigenvectors of the upper off-diagonal matrix element
564: $\:\lm\bu_k^{-1} (a_k-b_k \bv_k)\,$ of the operator $L_k$:
565: \bea \fl \la\,\bu_k^{-1} (a_k-b_k \bv_k)\, \psi^{(k)}_{\rho}\ny\\
566:  \hspace*{-1cm}=\: \la\,a_k\!\!\sum_{\g\in
567: \ZN} w_{p_k}(\g-\rho)\om^{-\g}|\g\ra_k-\la\,b_k\!\!\sum_{\g\in
568: \ZN} w_{p_k}(\g-\rho)\om^{-(\g+1)}|\g+1\ra_k
569: \ny\\
570: \hspace*{-1cm}=\: \la\,a_k\!\!\sum_{\g\in
571: \ZN} w_{p_k}(\g-\rho)\om^{-\g}|\g\ra_k-\la\,b_k\!\!\sum_{\g\in
572: \ZN} w_{p_k}(\g-\rho-1)\om^{-\g}|\g\ra_k
573: \ny\\
574:  \hspace*{-1cm} =\:\la\,\sum_{\g\in \ZN} w_{p_k}(\g-\rho)\left[
575: \lk a_k -\frac{b_k}{y_k}\rk\om^{-\g}+b_k\frac{x_k}{y_k}\om^{-\rho}\right]|\g\ra_k
576: =\la\,r_k\,\om^{-\rho}\,\psi_\rho^{(k)}.
577:   \label{efb} \eea
578: In the first step we used \r{uv} and to obtain the last line we used \r{Fermat} with
579:  $y_k=b_k/a_k$, $r_k=x_k a_k$. The Fermat curve restriction for $p_k=(x_k,y_k)$ gives
580:  $r_k^N=a_k^N-b_k^N$. We see that if $r_k=0$ (in particular, in the superintegrable
581:  case \r{super}) it leads to
582: $x_k=0,\;y_k=1$. In this case \r{psik} does not give a new basis in ${\cal V}_k$.
583: This is a reason why we exclude values of the parameters which lead to
584: the degeneration of the cyclic  function $w_p(\gamma)$.
585: 
586:  This sequence of operations applied in \r{efb} will be performed rather often in the
587:  following derivations. The application of $\bv_k$ shifts the
588:  spin index. This is compensated by the shift of the summation variable, which
589:  results in an opposite shift of the argument of $w_p$. This in turn is
590:  removed using \r{Fermat}.
591: 
592: The operator $\bv_k$ shifts the index of $\:\psi_\rho^{(k)}$:
593: \be \bv_k\psi_\rho^{(k)}=\sum_{\g\in
594: \ZN} w_{p_k}(\g-\rho)|\g+1\ra_k=
595: \sum_{\g\in\ZN} w_{p_k}(\g-\rho-1)|\g\ra_k= \psi_{\rho+1}^{(k)}\,.
596: \label{vpsi}\ee
597: Using \r{efb} for $k=1$ and comparing to (\ref{iBlm}), we write one-site
598: eigenvector as $\Psi_{\rho_{1,0}}:= \psi^{(1)}_{\rho_{1,0}}$. With $\;r_{1,0}\,=r_1\:$
599: we have  \be\label{AB1}
600: B_1(\lm)\Psi_{\rho_{1,0}}=\lm\: r_{1,0}\: \om^{-\rho_{1,0}}\, \Psi_{\rho_{1,0}}\,,
601: \quad
602: A_1(\lm)\Psi_{\rho_{1,0}}=\Psi_{\rho_{1,0}}+\lm\kp_1 \Psi_{\rho_{1,0}+1}\,.
603: \ee
604: 
605: The construction of the two-site eigenvectors $\,\Psi_{\bdr_2}\,$ will show us
606: the first step of the recursive method.
607: In accordance with (\ref{iBlm}) we are looking for
608: eigenvectors $\Psi_{\bdr_2}$,
609: $\:(\bdr_2\equiv(\rho_{2,0},\rho_{2,1})\in{\mathbb Z}_N\times\ZN)$
610: of the two-site operator $B_2(\la)$, which should satisfy
611: \be\label{B2}
612: B_2(\la)\Psi_{\bdr_2}=\la\,r_{2,0}
613:         \om^{-\rho_{2,0}}(\la+r_{2,1}\om^{-\rho_{2,1}})\Psi_{\bdr_2}\,.
614: \ee
615: We suppose that $\Psi_{\bdr_2}$ can be
616: written as a linear combinations of products of one-site eigenvectors
617: \be \label{Psi2}
618: \Psi_{\bdr_2}\;=\;
619: \sum_{\rho_1,\,\rho_2\in \ZN}Q(\rho_1,\rho_2| \bdr_2)\:
620: \psi^{(1)}_{\rho_1}\!\otimes \psi_{\rho_2}^{(2)}\,.
621: \ee
622: Using (\ref{AB1}) and (\ref{efb}), the matrix $Q$ can be calculated as follows:
623: \bea
624: \fl B_2(\la)\Psi_{\bdr_2}= \left(A_1(\lm)\, \lm\,\bu_2^{-1} (a_2-b_2 \bv_2)+
625: B_1(\lm) (\lm a_2 c_2+b_2\,d_2\bv_2/\kp_2)\right) \Psi_{\bdr_2}=\ny\\ [2mm]
626: \hspace*{-1.5cm}   =
627: \sum_{\rho_1, \,\rho_2 } \lb Q(\rho_1,\rho_2|\bdr_2 ) \lk \la r_2\om^{-\rho_2}
628: +\la^2 a_2c_2 r_1\om^{-\rho_1}\rk +Q(\rho_1-1,\rho_2|\bdr_2 )\la^2
629: \kp_1 r_2\om^{-\rho_2}\right.
630: \ny \\ [-2mm]
631:  \hspace*{1.5cm}  +  \left.Q(\rho_1,\rho_2-1|\bdr_2 )
632: \frac{b_2d_2}{\kp_2}\la r_1\om^{-\rho_1}\!\rb
633: \psi^{(1)}_{\rho_1}\!\otimes \psi^{(2)}_{\rho_2}.\label{B2calc}
634: \eea
635: Comparing powers of the spectral parameter $\lm$ in (\ref{B2calc}) and
636: in (\ref{B2}), together with (\ref{Psi2}), we get
637: \be
638: \hspace*{-1.5cm}\lk r_{2,0}\,\om^{-\rho_{2,0}}-
639: a_2\,c_2\,r_1\,\om^{-\rho_1}\rk Q( \rho_1,\rho_2|\bdr_2) \;=
640: \;  \kp_1 r_2\om^{-\rho_2} Q( \rho_1-1,\rho_2|\bdr_2),\label{twop1}
641: \ee
642: \be
643: \hspace*{-1.5cm}\lk r_{2,0}r_{2,1}\om^{-\rho_{2,0}-\rho_{2,1}}
644: -r_2\om^{-\rho_2}\rk Q( \rho_1,\rho_2|\bdr_2 )=
645: \frac{b_2d_2}{\kp_2}r_1\om^{-\rho_1}Q(\rho_1,\rho_2-1|\bdr_2 ).\label{diff2}
646: \ee
647: The difference equations (\ref{twop1}) and (\ref{diff2}) have the solution
648: \be \label{twop3}
649:   Q(\rho_1,\rho_2|\bdr_2)\!=\!\frac{\om^{-(\rho_{2,0}+\rho_{2,1}-\rho_1)
650:  (\rho_{2,0} -\rho_2)}}
651: {w_{p_{\,2,\,0}}(\rho_{2,0}-\rho_1-1) w_{\tilde p_{\,2}}(\rho_{2,0}+\rho_{2,1}-\rho_2-1 )},
652: \ee
653: where $p_{2,\,0}=(x_{2,\,0},y_{2,\,0})$,
654: $ \tilde p_{2}=( \tilde x_{2},\tilde y_{2})$ and
655:  \be
656:  \hspace*{-1.5cm} x_{2,\,0}=a_2\,c_2\frac {r_1}{r_{2,\,0}}\,, \hq
657:  y_{2,\,0}=\kp_1\,\frac {r_2}{r_{2,\,0}}\,,\hq
658:  \tilde x_{2}= \frac {r_2}{r_{2,\,0}\, r_{2,\,1}}\,, \hq
659: \tilde y_{2}= \frac{b_2\,d_2}{\kp_2}\frac {r_1}{r_{2,\,0}\, r_{2,\,1}}\,.\label{ptwo}
660:  \ee
661: The parameters $r_{2,0}$ and $r_{2,1}$ are determined from the
662: condition that the points $p_{2,0}$ and $\tilde p_{2}$ belong to the Fermat curve.
663: 
664: One can proceed this way to construct $n$-site eigenvectors of the auxiliary
665: problem. In fact, in order to see the general structure emerging, one has to go to
666: the four-site case. We shall not do this here, but rather in Section~\ref{eigenvectors}
667: we shall prove the general result by induction. This proof will use recursive relations
668: between amplitudes $r_{n,k}$, $k=0,1,\ldots,n-1$ formulated in the
669: Subsection~\ref{subsec24}. Fermat parameters $p=(x,y)$ of
670: the cyclic functions $w_{p}(\rho)$ appearing in our construction will
671: depend on these amplitudes. Compatibility conditions between recursive relations
672: for the amplitudes and Fermat curve equation $x^N+y^N=1$ can be formulated
673: as a ``classical'' BBS chain model \cite{GPS}.
674: This model will be formulated in the next Subsection
675: using an averaging procedure for the cyclic $L$-operators \r{bazh_strog} given in
676: \cite{Tarasov}.
677: 
678: 
679: \subsection{Determination of the parameters  $r_{m,s}$}\label{recinh}
680: 
681: 
682: Let us define the ``classical'' counterpart
683: $\,{\cal O}(\lm^N)\,$ of a quantum cyclic operator $\,O(\lm)\,$ using
684: averaging procedure \cite{Tarasov}
685: \be {\cal O}(\lm^N)\;=\;\langle\,O\,\rangle(\lm^N)=\;{\textstyle \av} O(\om^s\lm)
686: \label{ave}\ee
687: and apply this procedure to the entries of the quantum $L$-operator \r{bazh_strog}.
688: Denote the result by $\mathcal{L}_k(\lm^N)$
689: \be\label{Lclass}
690: \fl \mathcal{L}_k(\lm^N)\;=\;\left(\ba{cc} \langle\, L_{00}\,\rangle&
691:          \langle\,L_{01}\,\rangle \\ [2mm]
692:          \langle\,L_{10}\,\rangle & \langle\,L_{11}\,\rangle\ea\right)
693: =\left( \ba{cc}1-\epsilon \kp_k^N \lm^N &\;\; -\epsilon\lm^N(a_k^N-b_k^N)\\ [2mm]
694:             c_k^N-d_k^N &\;\; b_k^N d_k^N/\kp_k^N-\epsilon \lm^N a_k^N c_k^N\ea \right)\ee
695: %
696: where $\epsilon=(-1)^N$, and call it as the ``classical'' $\mathcal{L}$-operator
697: of the classical BBS model. Accordingly, the classical
698: monodromy ${\cal T}_m$ for the $m$-site chain is
699: \be \label{Tclass}
700: \mathcal{T}_m(\lm^N)\:=\;{\cal L}_1(\lm^N)\: {\cal L}_2(\lm^N) \cdots\:
701: {\cal L}_m(\lm^N)\:=\:\left(\ba{cc}
702: {\cal A}_m(\lm^N) &  {\cal B}_m(\lm^N)\\ [1mm]
703: {\cal C}_m(\lm^N) &  {\cal D}_m(\lm^N)
704: \ea\right)\ee
705: where the entries are polynomials of $\lm^N$.
706:  By Proposition~1.5 from \cite{Tarasov}, these
707: polynomials coincide with averages $\langle{A}_m\rangle$, $\langle{B}_m\rangle$,
708: $\langle{C}_m\rangle$ and $\langle{D}_m\rangle$ of the entries of \r{mm}.
709: This proposition provides a tool for finding the $N$-th powers of the amplitudes $r_{m,s}$:
710: applying \r{ave} to \r{iBlm} we obtain
711: \be
712:  \lefteqn{{\cal B}_m(\lm^N)=(-\epsilon)^m \lm^N r_{m,0}^N \prod_{s=1}^{m-1}(\lm^N-
713: \epsilon\, r_{m,s}^N)\,.}\label{Brec}
714: \ee
715: This relation together with \r{Lclass} and \r{Tclass} allows to find $r_{m,s}^N$ in terms of
716: the parameters $a_k^N$, $b_k^N$, $c_k^N$, $d_k^N$ and $\kp_k^N$, $k=1,\ldots,m$.
717: The problem of finding the amplitudes $r_{m,s}$ is reduced
718: to the problem of solving a $(m-1)$-th degree algebraic relation.
719: As shown in the Appendix, in
720: the case of the homogeneous \BBS chain model
721: the problem is reduced to solving a quadratic equation only.
722: The described procedure gives the amplitudes $r_{m,s}$ up to some roots of unity.
723: In fact we can fix these phases arbitrarily because this leads just to relabeling of the
724: eigenvectors. In what follows we suppose that we fixed some solution
725: $\{r_{m,s}\}$ in terms of the parameters $a_k^N$, $b_k^N$, $c_k^N$, $d_k^N$ and $\kp_k^N$.
726: 
727: Let us give a recursive description for ${\cal B}_m(\lm^N)$.
728: From \r{Tclass}, we immediately read off the recursion relations
729: \[\fl
730: {\cal A}_m(\lm^N)= (1-\epsilon \kp_m^N \lm^N ) {\cal A}_{m-1}(\lm^N)
731: +(c_m^N-d_m^N) {\cal B}_{m-1}(\lm^N)\,,
732: \]
733: \be\label{BAB}\fl
734: {\cal B}_m(\lm^N)= -\epsilon\lm^N(a_m^N-b_m^N) {\cal A}_{m-1}(\lm^N)
735: +(b_m^N d_m^N/\kp_m^N-\epsilon \lm^N a_m^N c_m^N) {\cal B}_{m-1}(\lm^N)\,.
736: \ee
737: Combining these two relations we get
738: \be\label{calAB}
739: {\cal A}_m(\lm^N)\,=\,
740: \frac{\epsilon\,\kp_m^N\lm^N-1}{\epsilon\,\lm^N\,(a_m^N\,-b_m^N)}\: {\cal B}_m(\lm^N)
741: \,+\,\frac{\det\, {\cal L}_m(\lm^N)}{\epsilon\,\lm^N\,(a_m^N\,-b_m^N)}\: {\cal B}_{m-1}(\lm^N),
742: \ee
743: where \\ [-8mm] \be \det \:{\cal L}_m(\lm^N)\;=\;(d_m^N\,-\epsilon\,\lm^N\,c_m^N\,\kp_m^N)\,
744: (b_m^N\,-\epsilon\,\lm^N\,a_m^N\,\kp_m^N)/\kp_m^N. \label{detlN}\ee
745: Substituting ${\cal A}_{m-1}$ from this equation with $m$ replaced by $m-1$ into
746: \r{BAB}, we
747: obtain a three-term recursion for $\:{\cal B}_m(\lm^N)$:
748: \bea \fl {\cal B}_m(\lm^N)&=&\lk\frac{r_m^N}{r_{m-1}^N}(1-\epsilon\,\kp_{m-1}^N\lm^N)+
749: \,b_m^N\, d_m^N/\kp_m^N\,-\epsilon\, \lm^N a_m^N\, c_m^N\rk {\cal
750: B}_{m-1}(\lm^N)+ \ny\\
751: \fl &&\hspace*{-20mm}+\:\frac{r_m^N}{r_{m-1}^N\kp_{m-1}^N}(b_{m-1}^N -\epsilon
752: \lm^N a_{m-1}^N \kp_{m-1}^N ) (\epsilon \lm^N c_{m-1}^N
753: \kp_{m-1}^N - d_{m-1}^N)\: {\cal B}_{m-2}(\lm^N), \quad m\ge 2,
754: \label{relBm} \eea
755: where we abbreviated
756: \be r_m^N\;=\;a_m^N\,-\,b_m^N. \label{rmN}\ee
757: To define $\:{\cal B}_m(\lm^N)$ by \r{relBm} we have to provide the initial values
758: \[
759: {\cal B}_1(\lm^N)=-\epsilon\, \lm^N r_1^N; \hq\qquad
760: {\cal B}_0(\lm^N)=0.
761: \]
762: 
763: 
764: 
765: 
766: \subsection{Fermat curve points appearing in the construction of
767:                       the eigenvectors of $B_n(\lm)$}
768:                       \label{subsec24}
769: 
770: As we have seen in the case of the two-site chain,
771: the formulas \r{Psi2}, \r{twop3} for the eigenvectors are given in terms of the points
772: $p_{2,\,0}$ and $ \tilde p_{2}$ on the Fermat curve. The coordinates of these points are
773: fixed by the values of amplitudes  $r_{2,0}$, $r_{2,1}$, see \r{ptwo}.
774: In the $n$-site case  four types of such points will appear:
775: \be  \fl \tilde p_m=(\tilde
776: x_m,\tilde y_m);\quad
777: p_{m,s}=(x_{m,s}, y_{m,s});\quad\tilde p_{m,s}=(\tilde
778: x_{m,s},\tilde y_{m,s});\quad
779: p^{m,s}_{m',s'}=(x^{m,s}_{m',s'},y^{m,s}_{m',s'}).\label{pxy}\ee
780: The coordinates of these points are expressed in the terms of amplitudes
781: $r_{m,s}$, $m=1,\ldots,$ $n$, $s=0,\ldots,$ $m-1$ (defined as {\it some} solutions of
782: equations \r{Brec}, $m=1,\ldots,$ $n$) by
783: \be \fl x^{m,s}_{m',s'}=r_{m,s}/r_{m',s'},\quad x_{m,s}=a_m \kp_m
784: r_{m,s}/b_m,\quad \tilde x_{m,s}=d_m/(\kp_m c_m r_{m,s}),\quad s,s'\ge 1.
785: \label{xxx}\ee
786: The corresponding $y^{m,s}_{m',s'}$, $\tilde y_{m,s}$ are defined
787: by the only condition on $p^{m,s}_{m',s'}$, $\tilde p_{m,s}$ to belong to the Fermat curve.
788: The coordinates $y_{m-1,l}$, $1\le l\le m-2$ are defined by
789: \[\fl
790: \frac{\tilde r_{m-1} r_{m,0}\,r_{m-1}}{\tilde r_{m-2}\,r_{m-1,0} r_{m}\,
791: b_{m-1}\,c_{m-1}\, y_{m-1,l}\,\tilde y_{m-1,l}}\;
792: \]
793: \be \label{rel_other}
794: \qquad\qquad \times \prod_{s\ne l}^{m-2} \frac{y^{m-1,l}_{m-1,s}}{y^{m-1,s}_{m-1,l}}\;
795: \;\frac{\prod_{k=1}^{m-1} \;y^{m,k}_{m-1,l}}{\prod_{s=1}^{m-3}\; y_{m-2,s}^{m-1,l}}\:=\:1\,,
796: \qquad l=1,\ldots,m-2,
797: \ee
798: where
799: \be \rt_m\,=\,r_{m,0}\,r_{m,1}\,\ldots\, r_{m,m-1}\,.\label{rtilde}\ee
800: The coordinates of the points $p_{m,0}$ and ${\tilde p}_m$  are defined by
801: \be x_{m,0} r_{m,0}=r_{m-1,0} a_m c_m,\qquad y_{m,0}
802: r_{m,0}=\kp_1\kp_2\cdots\kp_{m-1} r_m\ ,\label{xynull}\ee
803: \be \tilde x_m \tilde r_m =r_m,\qquad \tilde y_m \tilde r_m =b_m
804: d_m \tilde r_{m-1}/\kp_m\,.\label{xyt}\ee
805: Formulas \r{rel_other}--\r{xyt}  are result from the solution of the
806: eigenvector problem \r{iBlm}, see Section~3.
807: 
808: The condition on the points $p_{m-1,l}$ ($1\le l\le m-2$), $p_{m,0}$ and ${\tilde p}_m$ defined
809: by \r{rel_other}, \r{xynull}, \r{xyt} to belong to the Fermat curve gives
810: \bea \fl \lefteqn{r_{m-1}^N  \kp_{m-1}^N r_{m,0}^N
811:  \prod_{k=1}^{m-1} (r_{m-1,l}^N-r_{m,k}^N)=
812: r_{m}^N r_{m-2,0}^N\:  (b_{m-1}^N -a_{m-1}^N \kp_{m-1}^N r_{m-1,l}^N)\,\times}
813: \ny\\ [-3mm]\fl &&\label{relotherN}\hs \times\, (r_{m-1,l}^N c_{m-1}^N  \kp_{m-1}^N-
814: d_{m-1}^N) \prod_{s=1}^{m-3} (r_{m-1,l}^N-r_{m-2,s}^N), \hq
815: l=1,2,\ldots,m-2, \eea
816: \be\label{rel0N}
817: r_{m,0}^N= r_{m-1,0}^N a_m^N c_m^N+\kp_1^N\kp_2^N\cdots\kp_{m-1}^N r_m^N,
818: \ee
819: \be\label{reltN}
820: \rt_m^N\:\equiv\:r_{m,0}^N r_{m,1}^N \cdots r_{m,m-1}^N\:
821:              =\:r_m^N+b_m^N d_m^N \rt_{m-1}^N/\kp_m^N\,.\ee
822: In order to relate these relations to the recurrent formulas of the classical BBS model
823: \r{relBm} we observe that
824: the relations \r{rel0N} (resp. \r{reltN}) follow from the relations obtained by
825: the consideration of the highest (resp. lowest) terms in $\lm$ in \r{relBm}
826: starting from $m=2$. Then, fixing in \r{relBm} $\lm^N$ successively at the $m-2$
827: non-vanishing zeros of ${\cal B}_{m-1}$,
828: i.e. putting $\lm^N\,=\,\epsilon r_{m-1,l}^N$,  $l=1,\ldots,m-2,$
829: we obtain \r{relotherN}. Thus the points $p_{m-1,l}$, $p_{m,0}$ and ${\tilde p}_m$
830: defined by \r{rel_other}, \r{xynull}, \r{xyt} belong to the Fermat curve automatically.
831: 
832: At the end of this section we would like to mention that the amplitudes $r_{m,s}$ can be found directly
833: (i.e. not using the results from the previous subsection) from the relations
834: \r{rel_other}, \r{xynull}, \r{xyt} considered as equations with respect to
835: $r_{m,s}$ and the coordinates of the Fermat curve points \r{pxy}.
836: These equations can be solved recursively starting from $m=2$ and taking
837: $N$-th powers of these relations (see \r{relotherN}, \r{rel0N}, \r{reltN}).
838: Then the explicit formulas for the eigenvectors from the next section allow
839: to obtain the Tarasov Proposition~1.5 in \cite{Tarasov}  as a corollary.
840: 
841: 
842: \section{Inductive proof of the general solution of the auxiliary problem}
843: \label{eigenvectors}
844: 
845: Recall from (\ref{AB1}) that the vector
846: $\Psi_{\rho_{1,0}}:= \psi^{(1)}_{\rho_{1,0}}\in {\cal V}_1$ is
847: eigenvector for $B_1(\lm)$:
848: \[B_1(\lm) \Psi_{\rho_{1,0}} = \lm\: r_{1,0}\: \om^{-\rho_{1,0}}\, \Psi_{\rho_{1,0}},\]
849: and
850: recall from \r{iBlm}, \r{br} that the eigenvectors $\Psi_{\bdr_n}$ of $B_n(\la)$
851: were labeled by the vector $\bdr_n=(\rho_{n,0},\ldots,\rho_{n,n-1})\in(\ZN)^n$.
852: Let us further define:
853: \be  \trh_n={\textstyle \sum_{k=0}^{n-1}}\;\;\rho_{n,k};\qquad\quad \bdr_n'=
854: (\rho_{n,1},\ldots,\rho_{n,n-1})\in(\ZN)^{n-1}. \ee
855: $\bdr_n^{\pm k}\;$ denotes the vector $\;\bdr_n\;$ in which $\rho_{n,k}$ is
856: replaced by $\rho_{n,k}\pm 1$, i.e.\\[1mm]
857: $\hs\hs\hx\;\:\bdr_n^{\pm k}=(\rho_{n,0},\ldots,\rho_{n,k}\pm 1,\ldots,\rho_{n,n-1}),\hq
858: k=0,1,\ldots,n$.\\[-3mm]
859: 
860: The following Theorem~\ref{th1} gives a procedure to obtain the eigenvectors
861: $\Psi_{\bdr_n}\:\in\:{\cal V}^{(n)}$, $n\ge 2$,
862: of $\;B_n(\lm)\;$ from eigenvectors
863: $\Psi_{\bdr_{n-1}}\in\,{\cal V}^{(n-1)}\;$ of $\:B_{n-1}(\lm)\,$
864: and single site vectors $\psi^{(n)}_{\rho_n}\:\in\:{\cal V}_{n}$
865: defined by (\ref{psik}). We start from $\Psi_{\rho_{1,0}}$.
866: As result of the first step of the induction we obtain the two-site
867: result \r{Psi2} with \r{twop3},\r{ptwo}.
868: 
869: The following theorem is valid provided  $r_m^N\ne 0$,
870: the polynomials ${\cal B}_m(\lm^N)/\lm^N$, $m=2,\ldots,n$,
871: have nonzero simple zeros and $\det {\cal T}_n(\epsilon r_{m,s}^N)\ne 0$
872: (cf. the definition of the $B$-representation in \cite{Tarasov}).
873: 
874: \begin{theorem}\label{th1}
875: The vector
876: \be
877: \Psi_{\bdr_n}=\sum_{\bdr_{n-1}\in (\ZN)^{n-1}\atop \rho_n \in \ZN}
878:  Q(\bdr_{n-1},\rho_n|\bdr_n) \Psi_{\bdr_{n-1}}\otimes
879: \psi^{(n)}_{\rho_n} \label{PSI}\ee
880: where \\ [-13mm]
881: \bea
882: Q(\bdr_{n-1},\rho_n|\bdr_{n})&=&\frac{\om^{(\tilde \rho_n-
883: \tilde \rho_{n-1}) (\rho_n-\rho_{n,0})}}
884: {w_{p_{n0}}(\rho_{n,0}-\rho_{n-1,0}-1) w_{\tilde p_n}(\tilde
885: \rho_n- \rho_n-1)}\: \times \ny \\ [2mm]  \label{QQQ} &&\hspace*{-2cm}\times\:
886: \frac{ \prod_{l=1}^{n-2} \prod_{k=1}^{n-1}
887: w_{p_{n-1,l}^{n,k}}(\rho_{n-1,l}-\rho_{n,k})} {\prod_{j,l=1\atop
888: (j\ne l)}^{n-2} w_{p_{n-1,j}^{n-1,l}}(\rho_{n-1,j}-\rho_{n-1,l})}
889: \prod_{l=1}^{n-2} \frac{w_{p_{n-1,l}}(-\rho_{n-1,l})}{w_{\tilde
890: p_{n-1,l}}(\rho_{n-1,l})}\eea
891: is eigenvector of $\;B_n(\la)$:\\ [-6mm]
892: %
893: \be\label{Blm}
894: B_n(\lm)\Psi_{\bdr_n} =\lm\, r_{n,0}\om^{-\rho_{n,0}}
895: \prod_{k=1}^{n-1}\lk\lm+r_{n,k}\om^{-\rho_{n,k}}\rk\Psi_{\bdr_n}.
896: \ee
897: The Fermat curve points $\;\tp_n,\:p_{n,l},\:\tp_{n,l},\:p_{n',l}^{n,k}$
898: and $\:r_{n,k},\:$ entering (\ref{QQQ})
899: are related to the parameters of the model $\:a_s,b_s,c_s,d_s,\varkappa_s\:$
900: by equations (\ref{xxx}), \r{rel_other}, \r{xynull}, \r{xyt}.
901: 
902: $A_n(\lm)$ acts on $\Psi_{\bdr_n}$ as follows:
903: \bea
904: \lefteqn{A_n(\lm)\Psi_{\bdr_n}\:=\:\prod_{s=1}^{n-1} \lk
905: 1-\frac{\lm}{\lm_{n,s}}\rk\:\Psi_{\bdr_n} +\lm \kp_1\cdots\kp_n\:
906: \prod_{s=1}^{n-1} (\lm-\lm_{n,s})\,\Psi_{\bdr_n^{+0}}\;+}\ny\\ &&
907: \hs\hs\hs+\;\sum_{k=1}^{n-1}\; \lk\prod_{s\ne k}
908: \frac{\lm-\lm_{n,s}}{\lm_{n,k}-\lm_{n,s}}\!\rk\!
909: \frac{\lm}{\lm_{n,k}}\;\varphi_k(\bdr'_n)\;\Psi_{\bdr_n^{+k}},
910: \label{Alm}
911: \eea
912: where\\ [-11mm]
913: \be \label{phik} \varphi_k(\bdr_n')\;=\;-\frac{\tilde r_{n-1}}{r_n}\;\om^{-\tilde
914: \rho_{n}+\rho_{n,0}}\;F_n(\lm_{n,k}/\om)\;\prod_{s=1}^{n-2}y_{n-1,s}^{n,k}\,
915:  \ee
916:  with\\ [-10mm]
917: \be F_n(\lm)\:=\:\lk\, b_n\,+\om a_n \,\kp_n \lm\rk\,
918: \lk \,\la\, c_n\,+d_n/\kp_n\, \rk. \label{qdet}\ee
919: {\it\bf Corollary.}
920: In particular, at the $\:n-1\:$ zeros $\:\la_{n,k}\:$ of the
921: eigenvalue polynomial of $\;B_n(\la)\;$
922: \be \la_{n,k}\:=\:-r_{n,k}\om^{-\rho_{n,k}},\hs\hs k=1,\ldots,n-1\ee
923: the operator $\:A_n\:$ acts as shift operator for the $k$-th index of $\;\Psi_{\rho_n}$:
924: \be
925:  A_n\lk\lm_{n,k}\rk\Psi_{\bdr_n}\,=\, \varphi_k(\bdr_n')\;\Psi_{\bdr_n^{+k}}\,.\label{Almk}
926:  \ee
927: Further, the term in \r{Alm} of highest degree in $\lm$ gives:
928: $\bV_n=\bv_1\bv_2\ldots\bv_n\;$ is a shift operator for the zeroth index of $\Psi_{\bdr_n}$:
929: \be\label{vvvPsi} \bV_n\,\Psi_{\bdr_n}\,=\,\Psi_{\bdr_n^{+0}}\,. \ee
930: \end{theorem}
931: \noindent {\it\bf Proof.}
932: We shall prove the Theorem~\ref{th1}  by induction,
933: showing that if it is valid for $n-1$ site eigenvectors, then it
934: follows for $n$ site eigenvectors. Namely,
935: we assume the correctness of the following formulas
936: \bea\label{Blm1}
937: B_{n-1}(\lm)\Psi_{\bdr_{n-1}} &=& \lm\, r_{n-1,0}\om^{-\rho_{n-1,0}}
938: \prod_{l=1}^{n-2}\lk\lm-\lm_{n-1,l}\rk\Psi_{\bdr_{n-1}},\\
939: A_{n-1}(\lm)\Psi_{\bdr_{n-1}}&=&\sum_{l=1}^{n-2} \left(\prod_{s\ne l}
940: \frac{\lm-\lm_{n-1,s}}{\lm_{n-1,l}-\lm_{n-1,s}}\right)
941: \frac{\lm}{\lm_{n-1,l}}\,\varphi_l(\bdr'_{n-1}) \Psi_{\bdr_{n-1}^{+l}}+
942:      \ny\\  \label{Alm1}
943:      &&\hspace*{-27mm} +\;\prod_{s=1}^{n-2} \left(1-\frac{\lm}
944:      {\lm_{n-1,s}}\right)\Psi_{\bdr_{n-1}}
945:      +\lm \kp_1\cdots\kp_{n-1} \prod_{s=1}^{n-2}
946:      (\lm-\lm_{n-1,s})\cdot \Psi_{\bdr_{n-1}^{+0}},
947: \eea
948: where
949: $\lm_{n-1,l}\,=\,-r_{n-1,l}\om^{-\rho_{n-1,l}}$ and the formulas for
950: $\varphi_l(\bdr_{n-1}')$ given by (\ref{phik}) with $n$
951: replaced by $n-1$.
952: 
953: 
954: \vspace{2mm}
955: 
956: \noindent
957: {\sf Formula \r{Blm} for $\:B_n(\lm)\Psi_{\bdr_n}$}:\\ [2mm]
958: In order to prove the eigenvalue formula (\ref{Blm}) we use the following relation
959: \be\label{recB}
960: B_n(\lm)=A_{n-1}(\lm)\, \lm\,\bu_n^{-1} (a_n-b_n \bv_n)+
961: B_{n-1}(\lm) \left(\lm a_n c_n+\frac{b_n d_n}{\kp_n} \bv_n\right)
962: \ee
963: which follows directly from \r{bazh_strog} and \r{mm}. We apply its left-hand
964: side to the left-hand side of \r{PSI} and its right-hand side to the
965: right-hand side of \r{PSI}. On the right, we use \r{Blm1}, \r{Alm1}, \r{efb}, \r{vpsi}.
966: According to \r{Alm1}, $A_{n-1}$
967: introduces shifts in the indices $\bdr_{n-1}$ of $\:\Psi_{\bdr_{n-1}}\,$, while the
968: second term involving $\bv_n$ shift the index of $\:\psi^{(n)}_{\rho_n}$. Since we are
969: looking for an eigenstate, by shifting the summation indices we restore the original
970: indices. However, this leaves a change in the matrix $\:Q(\bdr_{n-1},\rho_n|\bdr_n)\,$.
971: Now the difference equation \r{Fermat} for the $w_p(\g)$ functions appearing in
972: $\:Q(\bdr_{n-1},\rho_n|\bdr_n)\,$ is used, producing several factors under the summation
973: which together we call $\;R\,$:
974: \[
975: B_n(\lm)\Psi_{\bdr_n}=\sum_{\bdr_{n-1}\in (\ZN)^{n-1}\atop \rho_n \in \ZN}
976:  Q(\bdr_{n-1},\rho_n|\bdr_n)\, R\, \Psi_{\bdr_{n-1}}\otimes \psi^{(n)}_{\rho_n},
977: \]
978: After some calculation we obtain
979: \[\fl
980: R\;=\;\left\{\sum_{l=1}^{n-2} \left(\prod_{s\ne l}
981: \frac{\lm-\lm_{n-1,s}}{\om\lm_{n-1,l}-\lm_{n-1,s}}\right)
982: \frac{\lm}{\om\lm_{n-1,l}}\;\varphi_l({\bdr'}^{-l}_{n-1})\,
983: \frac{Q(\bdr^{-l}_{n-1},\rho_n|\bdr_n)}{Q(\bdr_{n-1},\rho_n|\bdr_n)}\;\:+\:\right.
984: \]
985: \[\fl\qquad\left.
986: +\prod_{s=1}^{n-2} \left(1-\frac{\lm}{\lm_{n-1,s}}\right)
987: +\lm \kp_1\cdots\kp_{n-1} \prod_{s=1}^{n-2} (\lm-\lm_{n-1,s})
988: \frac{Q(\bdr^{-0}_{n-1},\rho_n|\bdr_n)}{Q(\bdr_{n-1},\rho_n|\bdr_n)}\right\}
989: \lm r_n\om^{-\rho_n}\;+
990: \]
991: \[
992: + \lm r_{n-1,0}\om^{-\rho_{n-1,0}}
993: \prod_{l=1}^{n-2}\lk\lm-\lm_{n-1,l}\rk
994: \left(\lm a_n c_n +\frac{b_n d_n}{\kp_n}\cdot
995: \frac{Q(\bdr_{n-1},\rho_n-1|\bdr_n)}{Q(\bdr_{n-1},\rho_n|\bdr_n)}\right).
996: \]
997: We have to show that\\ [-8mm]
998: \be\label{R}
999: R=\lm\, r_{n,0}\om^{-\rho_{n,0}}
1000: \prod_{k=1}^{n-1}(\lm-\lm_{n,k});\qquad \lm_{n,k}=-r_{n,k}\om^{-\rho_{n,k}}.
1001: \ee
1002: This  will prove (\ref{Blm}).
1003: Using the definitions (\ref{xxx}) of $\:x_{m,s}\:$ and $\:\tilde x_{m,s}\:$,
1004: we can rewrite \r{qdet} for the argument $\lm=\lm_{n,k}/\om$ as follows:
1005: \be F_n(\lm_{n,k}/\om)\:=\:\lm_{n,k}b_n c_n\om^{-1}\lk 1-x_{n,k}\om^{-\rho_{n,k}}\rk\:
1006: \lk 1-\tilde x_{n,k}\om^{\rho_{n,k}+1}\rk.\label{qdetmk}\ee
1007: Taking into account the expression (\ref{QQQ}) for $Q(\bdr_{n-1},\rho_n|\bdr_n)$,
1008: the definition for $w_p(\g)$ and the relations
1009: \r{xxx},\r{rel_other},\r{rtilde},\r{xynull},\r{qdetmk} we obtain:
1010: \[\fl
1011: \frac{Q(\bdr^{-l}_{n-1},\rho_n|\bdr_n)}{Q(\bdr_{n-1},\rho_n|\bdr_n)}=
1012: \om^{\rho_n-\rho_{n,0}}
1013: \prod_{k=1}^{n-1} \frac{w_{p^{n,k}_{n-1,l}}(\rho_{n-1,l}-\rho_{n,k}-1)}
1014: {w_{p^{n,k}_{n-1,l}}(\rho_{n-1,l}-\rho_{n,k})}
1015: \cdot
1016: \frac{w_{p_{n-1,l}}(-\rho_{n-1,l}+1)} {w_{p_{n-1,l}}(-\rho_{n-1,l})}
1017: \times\]
1018: \[\fl\;\times
1019: \frac{w_{\tp_{n-1,l}}(\rho_{n-1,l})}{w_{\tp_{n-1,l}}(\rho_{n-1,l}-1)}\,
1020: \prod_{s\ne l}\left(
1021: \frac{w_{p^{n-1,l}_{n-1,s}}(\rho_{n-1,s}-\rho_{n-1,l})}
1022: {w_{p^{n-1,l}_{n-1,s}}(\rho_{n-1,s}-\rho_{n-1,l}+1)}
1023: \cdot
1024: \frac{w_{p^{n-1,s}_{n-1,l}}(\rho_{n-1,l}-\rho_{n-1,s})}
1025: {w_{p^{n-1,s}_{n-1,l}}(\rho_{n-1,l}-\rho_{n-1,s}-1)}
1026: \right)=
1027: \]\[
1028: =\frac{\om}
1029: {\varphi_l({\bdr'}^{-l}_{n-1})}
1030: \frac{r_{n,0}\,\om^{-\rho_{n,0}}} {r_n\om^{-\rho_n}}
1031: \prod_{k=1}^{n-1} (\lm_{n-1,l}-\lm_{n,k})
1032: \prod_{s\ne l}\frac{\om\lm_{n-1,l}-\lm_{n-1,s}}{\lm_{n-1,l}-\lm_{n-1,s}},
1033: \]
1034: \[\fl
1035: \frac{Q(\bdr^{-0}_{n-1},\rho_n|\bdr_n)}{Q(\bdr_{n-1},\rho_n|\bdr_n)}=
1036: \om^{\rho_n-\rho_{n,0}}
1037: \frac{w_{p_{n0}}(\rho_{n,0}-\rho_{n-1,0}-1)}
1038: {w_{p_{n0}}(\rho_{n,0}-\rho_{n-1,0})}=
1039: \frac{r_{n,0}\om^{-\rho_{n,0}}-r_{n-1,0}\om^{-\rho_{n-1,0}}a_n c_n}
1040: {\varkappa_1 \varkappa_2\cdots \varkappa_{n-1} r_n\om^{-\rho_n}},
1041: \]
1042: \[\fl
1043: \frac{Q(\bdr_{n-1},\rho_n-1|\bdr_n)}{Q(\bdr_{n-1},\rho_n|\bdr_n)}=
1044: \om^{\trh_{n-1}-\trh_n} \frac{w_{\tp_n}(\trh_n-\rho_n-1)}{w_{\tp_n}(\trh_n-\rho_n)}=
1045: \frac{\varkappa_n}{b_n d_n}\cdot
1046: \frac{\rt_n\om^{-\trh_n}-r_n \om^{-\rho_n}}
1047: {\rt_{n-1}\om^{-\trh_{n-1}}}.
1048: \]
1049: Substituting these expressions into $R$ gives
1050: \[\fl
1051: R\;=\;\left\{\sum_{l=1}^{n-2} \left(\prod_{s\ne l}
1052: \frac{\lm-\lm_{n-1,s}}{\lm_{n-1,l}-\lm_{n-1,s}}\right)
1053: \frac{\lm}{\lm_{n-1,l}}
1054: \frac{r_{n,0}\,\om^{-\rho_{n,0}}} {r_n\om^{-\rho_n}}
1055: \prod_{k=1}^{n-1} (\lm_{n-1,l}-\lm_{n,k})+\right.
1056: \]
1057: \[\fl\quad\left.
1058: +\prod_{s=1}^{n-2} \left(1-\frac{\lm}{\lm_{n-1,s}}\right)
1059: +\lm\prod_{s=1}^{n-2} (\lm-\lm_{n-1,s})\cdot
1060: \frac{r_{n,0}\om^{-\rho_{n,0}}-r_{n-1,0}\om^{-\rho_{n-1,0}}a_n c_n}
1061: { r_n\om^{-\rho_n}}\right\} \lm r_n\om^{-\rho_n}+
1062: \]
1063: \[
1064: + \lm r_{n-1,0}\om^{-\rho_{n-1,0}}
1065: \prod_{l=1}^{n-2}\lk\lm-\lm_{n-1,l}\rk\cdot
1066: \left(\lm a_n c_n +\frac{\rt_n\om^{-\trh_n}-r_n \om^{-\rho_n}}
1067: {\rt_{n-1}\om^{-\trh_{n-1}}}\right).
1068: \]
1069: After appropriate cancellations this becomes
1070: \[\fl
1071: R\;=\;\sum_{l=1}^{n-2} \left(\prod_{s\ne l}
1072: \frac{\lm-\lm_{n-1,s}}{\lm_{n-1,l}-\lm_{n-1,s}}\right)
1073: \frac{\lm^2}{\lm_{n-1,l}}
1074: r_{n,0}\,\om^{-\rho_{n,0}}
1075: \prod_{k=1}^{n-1} (\lm_{n-1,l}-\lm_{n,k})\:+
1076: \]
1077: \be\label{RR}
1078: \fl\qquad\qquad\quad +\:\lm^2\prod_{s=1}^{n-2} (\lm-\lm_{n-1,s})\cdot
1079: r_{n,0}\om^{-\rho_{n,0}}\,+ \lm\, \rt_n\om^{-\trh_n}
1080:  \prod_{l=1}^{n-2}\left(1-\frac{\lm}{\lm_{n-1,l}}\right)\,.
1081: \ee
1082: To prove (\ref{R}) we note that the coefficients at $\lm^n$ in both
1083: expressions (\ref{R}) and (\ref{RR}) are $r_{n,0}\om^{-\rho_{n,0}}$
1084: and coefficients at $\lm$ also coincide being
1085: $\rt_n\om^{-\trh_n}$.
1086: Therefore the difference of these two expressions
1087: has the form $\lm^2 P(\lm)$ where $P(\lm)$ is a polynomial of degree $n-3$.
1088: Using the explicit expressions (\ref{R}) and (\ref{RR}) we convince
1089: ourselves that $P(\lm_{n-1,j})=0$ and therefore $P(\lm)\equiv 0$.
1090: This completes the proof that $\Psi_{\bdr_n}$ defined by \r{PSI}, (\ref{QQQ})
1091: is eigenvector of $B_n(\lm)$ with eigenvalue (\ref{R}).
1092: 
1093: \vspace{2mm}
1094: 
1095: \noindent
1096: {\sf Formula \r{Almk} for $\:A_n(\lm_{n,k})\Psi_{\bdr_n}$}: \\ [2mm]
1097: Next we show the validity of (\ref{Almk}), (\ref{phik}).
1098: We will need the relation
1099: \be \fl \,\bu_n^{-1} (a_n-b_n \bv_n)
1100: A_n(\lm)\;=\;(1+\lm\kp_n\om^{-1}\bv_n) B_n(\lm)/\lm
1101:          \,-\,\bv_n\,F_n(\lm/\om)\:B_{n-1}(\lm)/\lm \label{recAB} \ee
1102: which can be obtained by eliminating
1103: $\:A_{n-1}\:$ between \r{recB} and
1104: \be   A_n(\lm)=(1+\lm\kp_n\bv_n) A_{n-1}(\lm)+\bu_n(c_n-d_n \bv_n)
1105:                              B_{n-1}(\lm).\label{recA}\ee
1106: Let us apply (\ref{recAB}) to
1107: $\:\Psi_{\bdr_n}\:$ for $\la=-r_{n,k}\om^{-\rho_{n,k}}$,
1108: i.e. at the zeros of $\:B_n(\la)$.
1109: This gives
1110: \bea \fl \bu_n^{-1} (a_n-b_n \bv_n) A_n(\lm_{n,k}) \Psi_{\bdr_n}&=&
1111: -F_n(\lm_{n,k}/\om)/\lm_{n,k}\;\times\ny\\ [1mm]
1112: \fl &\times&\!\!\! \sum_{\bdr_{n-1}\in (\ZN)^{n-1}\atop \rho_n \in
1113: \ZN}\!\!
1114:  Q(\bdr_{n-1},\rho_n|\bdr_n)\; B_{n-1}(\lm_{n,k})\Psi_{\bdr_{n-1}}\otimes
1115: \psi^{(n)}_{\rho_n+1}.\label{ABm} \eea
1116: From (\ref{Blm1}) we know how to apply
1117: $B_{n-1}$ to $\Psi_{\bdr_{n-1}}$:
1118: %
1119: \bea\label{Blmnk}
1120: \fl \lefteqn{B_{n-1}(\lnk)\Psi_{\bdr_{n-1}} =\la_{n,k}\,
1121: r_{n-1,0}\om^{-\rho_{n-1,0}} \prod_{s=1}^{n-2}\lk
1122: -\rnk\om^{-\rho_{n,k}}+r_{n-1,s}\om^{-\rho_{n-1,s}}\rk\Psi_{\bdr_{n-1}}}\ny\\ &=&
1123: \la_{n,k}\,
1124: \rt_{n-1}\om^{-\trh_{n-1}} \lk\prod_{s=1}^{n-2}y_{n-1,s}^{n,k}\:
1125:    \frac{w_{p_{n-1,s}^{n,k}}\lk\rho_{n-1,s}-\rho_{n,k}-1\rk}
1126: {w_{p_{n-1,s}^{n,k}}\lk\rho_{n-1,s}-\rho_{n,k}\rk}\rk\: \Psi_{\bdr_{n-1}}\,.
1127: \eea
1128: %
1129: Using (\ref{efb}) we find the action of the inverse of the operator
1130: $\bu_n^{-1} (a_n-b_n \bv_n)$ on $\psi^{(n)}_{\rho_n}$\,:
1131: \be\label{invpsi}
1132: (\bu_n^{-1} (a_n-b_n \bv_n))^{-1} \psi^{(n)}_{\rho_n}=
1133: r_n^{-1} \om^{\rho_n} \psi^{(n)}_{\rho_n}.
1134: \ee
1135: Shifting the summation over $\rho_n$ in \r{ABm} and then applying
1136: (\ref{invpsi}) we obtain
1137: \bea  \fl A_n(\lm_{n,k}) \Psi_{\bdr_n}&=&
1138: -r_n^{-1} F_n(\lm_{n,k}/\om)/\lm_{n,k}  \;\times\ny\\ [1mm]
1139: &&\hspace*{-15mm}\times\!\!\!\sum_{\bdr_{n-1}\in (\ZN)^{n-1}\atop \rho_n \in
1140: \ZN}
1141:  Q(\bdr_{n-1},\rho_n-1|\bdr_n)\,\om^{\rho_n}
1142:  \; B_{n-1}(\lm_{n,k})\:\Psi_{\bdr_{n-1}}\otimes \psi^{(n)}_{\rho_n}. \eea
1143: Finally, using (\ref{Blmnk}) and observing that
1144: \[\fl
1145: Q(\bdr_{n-1},\rho_n\!-1|\bdr_n)\:\om^{\rho_n\!-\trh_{n-1}}\,
1146: \prod_{s=1}^{n-2}\frac{w_{p_{n-1,s}^{n,k}}
1147: \lk\rho_{n-1,s}\!-\rho_{n,k}-1\rk}
1148: {w_{p_{n-1,s}^{n,k}}\lk\rho_{n-1,s}\!-\rho_{n,k}\rk}\,=\,\om^{-\trh_{n}+\rho_{n,0}}
1149: \;Q(\bdr_{n-1},\rho_n|\bdr_n^{+k}) \]
1150: we come to (\ref{Almk}).
1151: 
1152: 
1153: \vspace{2mm}
1154: 
1155: \noindent
1156: {\sf Formula \r{vvvPsi} for $\;\bV_n\Psi_{\bdr_n}$}:\\ [2mm]
1157: Using $\:\bV_{n-1} \Psi_{\bdr_{n-1}}\,=\,\Psi_{\bdr_{n-1}^{+0}}$
1158: and $\:\bv_{n} \psi^{(n)}_{\rho_n}\,=\,\psi^{(n)}_{\rho_n+1}\:$  we have
1159: \bea\fl\lefteqn{\bV_{n-1}\,\bv_{n}\Psi_{\bdr_{n}}=
1160: \!\!\!\!\sum_{\bdr_{n-1}\in(\ZN)^{n-1}\atop \rho_{n} \in \ZN}
1161:  \!\!\!Q(\bdr_{n-1},\rho_{n}|\bdr_{n})
1162:  \Psi_{\bdr_{n-1}^{+0}}\otimes\,\psi^{(n)}_{\rho_n+1}}\ny\\ [-6.5mm]
1163: && \hspace*{3cm}=\!\!\!\!\sum_{\bdr_{n-1}\in(\ZN)^{n-1}\atop \rho_n \in \ZN}
1164:  \!\!Q(\bdr_{n-1}^{-0},\rho_n-1|\bdr_n) \Psi_{\bdr_{n-1}}
1165:  \otimes\,\psi^{(n)}_{\rho_n}, \eea
1166: where in the second line we have shifted the summation variables
1167: $\rho_{n-1,0}$ and $\rho_n$. Now considering the explicit form
1168: (\ref{QQQ}) for $\:Q(\bdr_{n-1},\rho_n|\bdr_n)\:$ we read off that
1169: \be
1170: Q(\bdr_{n-1}^{-0},\rho_n-1|\bdr_n)\:=\:Q(\bdr_{n-1},\rho_n|\bdr_n^{+0})\hs
1171: \ee
1172: which gives (\ref{vvvPsi}).
1173: 
1174: \vspace{2mm}
1175: 
1176: \noindent
1177: {\sf Formula \r{Alm} for $A_n(\lm)\Psi_{\bdr_n}$}:\\ [2mm]
1178: The operator $A_n(\la)$ is a polynomial in $\la$ of $n$th order.
1179: From \r{bazh_strog} and \r{mm} we immediately read off its
1180: the highest and lowest coefficients:   \be
1181: A_n(\la)\:=\:1\:+\:\ldots\:+\:\la^n\kp_1\kp_2\ldots\kp_n\,\bV_n\,.
1182: \label{An}  \ee
1183: Using \r{vvvPsi} we know how these terms act on $\Psi_{\bdr_n}$
1184: and if in addition we use the action of $\:A_n\,$ at the $n-1$ particular
1185: values given in \r{Almk}, we can
1186: reconstruct the action of the whole
1187: polynomial $A_n(\la)$ on $\:\Psi_{\bdr_n}\:$ uniquely.
1188: Comparing this to \r{Alm} we see that formula \r{Alm} is the one which
1189: satisfies all these data.
1190: Therefore by uniqueness formula (\ref{Alm}) is one which we are looking for.
1191: 
1192: This completes the proof of Theorem~\ref{th1}.
1193: 
1194: 
1195: \section{Action of $D_{n}$ on the eigenstates of $B_{n}$.}
1196: 
1197: In order to obtain the action of $\,D_n(\lm)\,$ on $\,\Psi_{\bdr_n}$ we use the notion of
1198: the quantum determinant ${\det}_q\, T_n(\lm)$ of the monodromy matrix.
1199: Since the rank of the matrix $R(\om\lm,\lm)$ is $1$, the definition of the quantum determinant
1200: is given by
1201: \be\label{qdett}\fl
1202: R(\om\lm,\lm)\; T_n^{(1)}(\om\lm)\,T_n^{(2)}(\lm)\,=
1203: T_n^{(2)}(\lm)\, T_n^{(1)}(\om\lm)\, R(\om\lm,\lm)\,=:
1204: {\det}_q\, T_n(\lm) \cdot\, R(\om\lm,\lm).
1205: \ee
1206: Explicitly we have
1207: \be\label{qdetexpl}
1208: {\det}_q\, T_n(\lm)=A_n(\om \lm)D_n(\lm)-C_n(\om \lm) B_n(\lm).
1209: \ee
1210: Using (\ref{mm}) and (\ref{qdett}) we obtain
1211: the factorization property of the quantum determinant
1212: \[ {\det}_q\, T_n(\lm)={\det}_q\, L_1(\lm)\cdot
1213: {\det}_q\, L_2(\lm)\cdots\:{\det}_q\,L_n(\lm), \]
1214: For a single $L$-operator, using \r{qdet}, \r{qdetexpl} gives
1215: $\;{\det}_q\,L_m(\lm)\:=\:\bv_m\,F_m(\lm)\,$. So
1216: \be  A_n(\om \lm)D_n(\lm)-C_n(\om \lm) B_n(\lm)
1217:        \:=\:\bV_n\cdot\prod_{m=1}^n\;F_m(\lm).\label{qdetall}
1218:        \ee
1219: Acting by both sides of this identity on $\Psi_{\bdr_n},$
1220: fixing $\lm=\lm_{n,k}$ (i.e. at the zeroes of the eigenvalue polynomial of $B_n(\lm)$)
1221: and using the inverse of
1222: (\ref{Almk}) with (\ref{phik}), we
1223: see that, very similar to $\:A_n(\la_{n,k}),$ also $\:D_n(\la_{n,k})\:$
1224: acts as a shift operator on $\Psi_{\bdr_n}$, compare \r{dlm}:
1225: \be\fl
1226: D_n(\lm_{n,k})\Psi_{\bdr_n}=\tilde\varphi_k(\bdr_n')\,\Psi_{\bdr_n^{+0,-k}};\qquad
1227:  \tilde\varphi_k(\bdr'_n)=
1228: -\frac{r_n}{\tilde r_{n-1}}\frac{\om^{\trh_n-\rho_{n,0}-1}}
1229: {\prod_{s=1}^{n-2}y_{n-1,s}^{n,k}}\;\prod_{m=1}^{n-1}\;F_m(\lm_{n,k}). \label{Dlmk}
1230: \ee
1231: Note that $D_n(\lm_{n,k})$ shifts $\rho_{n,k}$ in the opposite
1232: direction as $A_n(\lm_{n,k})$ (see \r{phik} and \r{Almk})
1233: and $D_n(\lm_{n,k})$ also
1234: shifts $\rho_{n,0}$. The shift in $\rho_{n,0}$
1235: is due to the operator $\bV_n$ at the right-hand side of (\ref{qdetall}).
1236: Apart from the shifts just mentioned, applying the inverse of $A_n(\om \la)$ has cancelled
1237: in \r{Dlmk} the last factor $m=n$ of the quantum determinant (\ref{qdetall}).
1238: Analogously to (\ref{Alm}), using the particular values (\ref{Dlmk}) and reading off
1239: the coefficients of $\la^0$ and $\la^n$ directly from
1240: (\ref{mm}), we obtain the following interpolation formula
1241: for $D_n(\lm)\Psi_{\bdr_{n}}$:
1242: %
1243: \bea
1244: \fl D_n(\lm)\Psi_{\bdr_{n}}&=&\sum_{k=1}^{n-1} \lk\prod_{s\ne k}
1245: \frac{\lm-\lm_{n,s}}{\lm_{n,k}-\lm_{n,s}}\rk
1246: \frac{\lm}{\lm_{n,k}}\,\tilde\varphi_k(\bdr'_{n})\Psi_{\bdr_{n}^{+0,-k}}+\ny\\
1247: \fl &+&\prod_{s=1}^{n-1} \lk 1-\frac{\lm}{\lm_{n,s}}\rk\cdot
1248: \prod_{m=1}^n\frac{b_m d_m}{\kp_m} \cdot
1249: \Psi_{\bdr_{n}^{+0}} +\lm \prod_{m=1}^n a_m
1250: c_m \cdot \prod_{s=1}^{n-1} (\lm-\lm_{n,s})\cdot \Psi_{\bdr_{n}}.
1251: \label{Dlm}\eea
1252: 
1253: 
1254: \section{Periodic model. Baxter equation and functional relations}
1255: \label{periodic}
1256: \subsection{The Baxter equations}
1257: 
1258: After having determined the eigenvalues and eigenvectors of the auxiliary system, we
1259: now perform the first step of the program exposed in Subsection \ref{SoV}, i.e.
1260: the calculation of the eigenvalues and eigenvectors of the inhomogenous
1261: $n$-site {\it periodic} \BBS chain model with the transfer matrix \r{tr_matl},
1262: \r{tr_mat_bs}. Following the ideas  of the papers
1263: \cite{Gu81,Skly1,Sk85,Skly2,KarLeb1} we are looking for eigenvectors of ${\bf t}_n(\lm)$ as
1264: linear combinations of the eigenvectors $\Psi_{{\bdr}_{n}}$ of the auxiliary system.
1265: 
1266: It is convenient to go by Fourier transform in $\rho_{n,0}$
1267: from $\Psi_{\bdr_n}$ to a basis of eigenvectors of $\bV_n$
1268: (and therefore of the Hamiltonians $\:{\bf H}_0$ and $\:{\bf H}_n$, see \r{tr_mat_bs}, \r{Znc})
1269: \be \tilde\Psi_{\rho,\bdr'_n}\:=\:
1270: {\textstyle \sum_{\rho_{n,0}\in\ZN}}\;\om^{-\rho\cdot\rho_{n,0}}\:
1271:      \Psi_{\bdr_n};\qquad\bV_n \:\tilde \Psi_{\rho,\bdr'_n}\;=\;\om^\rho\: \tilde
1272:      \Psi_{\rho,\bdr'_n}.\ee
1273: A shift of $\,\rho_{n,0}\,$ in $\Psi_{\bdr_n}$ is replaced by a multiplication of $\tilde
1274: \Psi_{\rho,\bdr'_n}$ by powers of $\om$.
1275: %
1276: So from (\ref{Alm}) and (\ref{Dlm}), the action of $\:{\bf t}_n(\la)$ on
1277: $\tilde\Psi_{\rho,\bdr'_{n}}$ becomes
1278: \bea \fl \lefteqn{ {\bf t}_n(\la)\tilde\Psi_{\rho,{\bdr'}_{n}}
1279: =\,\sum_{k=1}^{n-1}\lk\prod_{s\ne k} \frac{\lm\!-\lm_{n,s}}{\lm_{n,k}\!-\lm_{n,s}}\rk
1280: \!\frac{\lm}{\lm_{n,k}}\lk\varphi_k(\bdr'_{n})\tilde\Psi_{\rho,{\bdr'}_{n}^{+k}}\!
1281:            +\om^\rho \tilde\varphi_k(\bdr'_{n})\tilde\Psi_{\rho,{\bdr'}_{n}^{-k}}\rk} \ny\\
1282: \fl &&\hspace*{30mm}+\:
1283: \lb\lk 1+\om^\rho\prod_{m=1}^n\frac{b_m d_m}{\kp_m}\rk\:
1284:    \prod_{s=1}^{n-1} \lk 1-\frac{\la}{\la_{n,s}}\rk\,+\right.\ny\\
1285: \fl && \hspace*{50mm}\left.+ \:\la \lk \om^\rho\,\prod_{m=1}^n\:\kp_m
1286: +\prod_{m=1}^n\: a_mc_m \rk
1287: \prod_{s=1}^{n-1}(\lm-\lm_{n,s}) \rb
1288: \tilde\Psi_{\rho,\bdr'_{n}},
1289: \label{ApD}\eea
1290: where we have taken into account that
1291: $\varphi_k(\bdr'_{n})$ and $\tilde\varphi_k(\bdr'_{n})$ are
1292: independent of $\rho_{n,0}$.
1293: Of course, since $\:{\bf t}_n(\la)\,$ commutes with $\bV_n$, in (\ref{ApD}) there is
1294: no coupling between sectors of different $\rho$ and we
1295: get separate equations for the different ``charge'' quantum numbers $\rho$ which
1296: often will not be indicated explicitly.
1297: 
1298: Let $\Phi_{\rho,{\bf E}}$ be eigenvector of ${\bf t}_n(\lm)$ with eigenvalue
1299: \be  t_n(\lm|\,\rho,\,{\bf E})\;=\;
1300: E_0+E_1\lm+\cdots+E_{n-1}\lm^{n-1}+E_{n}\lm^{n}\,, \label{tnh}\ee
1301: where $\;\:{\bf E}\;=\;\{E_1,\ldots,E_{n-1}\}\;$ and from \r{Znc}
1302: the values of $E_0$ and $E_n$
1303: are
1304: \be\label{iE0En}
1305:  E_0\,=1+\om^\rho \prod_{m=1}^n\frac{b_m d_m}{\kp_m},\qquad
1306: E_n\,=\prod_{m=1}^n a_m c_m\,+\om^{\rho}\,\prod_{m=1}^n\kp_m\,.\ee
1307: %
1308: We are looking for $\:\Phi_{\rho,{\bf E}}\:$ to be of the form
1309: \be \Phi_{\rho,{\bf E}}\:=\:
1310: \sum_{\bdr'_n} \; Q(\bdr'_{n}|\,\rho,{\bf E})\;\:\tilde\Psi_{\rho,{\bdr'}_{n}}.
1311: \label{EV}\ee
1312: From (\ref{ApD}) we get a difference equation for $\:Q (\bdr'_{n}|\,\rho,{\bf E})\:$
1313: with respect to variables
1314: $\:\bdr'_{n}\:$ which depends on $\;\lm$:
1315: \bea \fl
1316: t_n(\lm|\rho,{\bf E})\:Q (\bdr'_{n}|\,\rho,{\bf E})\:&=&\:\sum_{k=1}^{n-1}
1317: \frac{\lm}{\om\lm_{n,k}}\,\varphi_k({\bdr'}^{-k}_{n})\;
1318: Q({\bdr'}^{-k}_{n}|\,\rho,{\bf E})
1319:         \prod_{s\ne k}\frac{\lm-\lm_{n,s}}{\om\lm_{n,k}-\lm_{n,s}}\:\: +
1320: \ny\\  &&\hspace*{-10mm}+\:
1321: \sum_{k=1}^{n-1}\frac{\om\lm}{\la_{n,k}}\,\om^\rho
1322: \tilde\varphi_k({\bdr'}^{+k}_{n})\; Q({\bdr'}^{+k}_{n}|\,\rho,{\bf E})
1323: \prod_{s\ne k}\frac{\lm-\lm_{n,s}}{\om^{-1}\la_{n,k}-\la_{n,s}}\,+\ny\\
1324:  &&\hspace*{-10mm}+\:
1325: \lb\lk 1+\om^\rho\prod_{m=1}^n\frac{b_m d_m}{\kp_m}\rk
1326: \prod_{s=1}^{n-1} \lk 1-\frac{\la}{\la_{n,s}}\rk\:+\right.\ny\\
1327:  &&\hspace*{-3mm}  \left.+\,\la
1328:  \lk \om^\rho\prod_{m=1}^n \kp_m+\prod_{m=1}^n a_mc_m \rk
1329: \prod_{s=1}^{n-1}(\la-\la_{n,s}) \rb Q(\bdr'_n|\,\rho,{\bf E}).
1330: \eea
1331: 
1332: Substituting sequentially $\lm=\lm_{n,k}$, $k=1,2,\ldots,n-1$, we
1333: obtain a system of difference equations with respect to the variables
1334: $\bdr'_{n}$:
1335: \bea  \fl \lefteqn{ t_n(\lm_{n,k}|\rho,\,{\bf E})\:\:
1336: Q (\bdr'_{n}|\,\rho,{\bf E})=
1337: \lk\prod_{s\ne k} \frac{\lm_{n,k}-\lm_{n,s}}{\om\lm_{n,k}-\lm_{n,s}}\rk
1338: \om^{-1}\,\varphi_k({\bdr'}^{-k}_{n})\: Q ({\bdr'}^{-k}_{n}|\,\rho,{\bf E})\;+}\ny\\
1339: \fl &&\hspace*{10mm}+\:\lk\prod_{s\ne k}
1340: \frac{\lm_{n,k}-\lm_{n,s}}{\om^{-1}\lm_{n,k}-\lm_{n,s}}\rk
1341: \om^{\rho+1} \tilde\varphi_k({\bdr'}^{+k}_{n})\: Q({\bdr'}^{+k}_{n}|\,\rho,{\bf E}),
1342: \hq k=1,\ldots,n-1.
1343: \label{pdiff}\eea
1344: In analogy to \cite{Gu81,Sk85,KarLeb1} we can decouple these difference equations
1345: using a Sklyanin's measure, namely, by
1346:  introducing  $\;\tQ (\bdr'_{n}|\,\rho,{\bf E})\;$ defined as
1347: \be Q (\bdr'_{n}|\,\rho,{\bf E})=\frac{\tQ (\bdr'_{n}|\,\rho,{\bf E})}
1348: {\prod_{s,s'=1\atop (s\ne s')}^{n-1}
1349: w_{p_{n,s}^{n,s'}}(\rho_{n,s}-\rho_{n,s'})}.\ee
1350: Rewriting \r{pdiff} in terms of $\tQ$ produces factors $R_\pm$
1351: in both terms of the right hand side:
1352: \bea  \fl \lefteqn{ t_n(\lm_{n,k}|\rho,\,{\bf E})\:\:
1353: \tQ (\bdr'_{n}|\,\rho,{\bf E})=
1354: \lk\prod_{s\ne k} \frac{\lm_{n,k}-\lm_{n,s}}{\om\lm_{n,k}-\lm_{n,s}}\rk
1355: \om^{-1}\,\varphi_k({\bdr'}^{-k}_{n})\:R_-\:\tQ({\bdr'}^{-k}_{n}|\,\rho,{\bf E})\;+}\ny\\
1356: \fl &&\hspace*{10mm}+\:\lk\prod_{s\ne k}
1357: \frac{\lm_{n,k}-\lm_{n,s}}{\om^{-1}\lm_{n,k}-\lm_{n,s}}\rk
1358: \om^{\rho+1} \tilde\varphi_k({\bdr'}^{+k}_{n})\:R_+\:\tQ({\bdr'}^{+k}_{n}|\,\rho,{\bf E}),
1359: \hq k=1,\ldots,n-1.
1360: \label{pdiffR}\eea
1361: where \bea \fl R_+\;&=&\prod_{s=1\atop{s\neq k}}^{n-1}\;
1362:   \frac{w_{p_{n,s}^{n,k}}(\rho_{n,s}-\rho_{n,k})}
1363:   {w_{p_{n,s}^{n,k}}(\rho_{n,s}-\rho_{n,k}\,- 1)}\:\cdot\:
1364:   \frac{w_{p_{n,k}^{n,s}}(\rho_{n,k}-\rho_{n,s})}
1365:      {w_{p_{n,k}^{n,s}}(\rho_{n,k}-\rho_{n,s}+ 1)}\ny\\
1366:   \fl \;&=&\prod_{s=1\atop{s\neq k}}^{n-1}\;\frac{y_{n,s}^{n,k}}{1\,-x_{n,s}^{n,k}\:
1367:   \om^{\rho_{n,s}-\rho_{n,k}}}\;
1368:    \:\frac{1\,-x_{n,k}^{n,s}\:\om^{\rho_{n,k}-\rho_{n,s}+1}}{y_{n,k}^{n,s}}\;=\;
1369:    \prod_{s=1\atop{s\neq k}}^{n-1}\;\:\frac{y_{n,s}^{n,k}}{y_{n,k}^{n,s}}\;
1370:    \frac{\lm_{n,s}}{\lm_{n,k}}\;\:\frac{\om\,\lm_{n,s}-\lm_{n,k}}{\lm_{n,k}-\lm_{n,s}}
1371:    \,, \ny\eea
1372: and analogously $\:R_-$. We see that passing from $\:Q\:$ to $\:\tQ\:$ the brackets
1373: containing differences of terms $\:\lm_{n,l}\:$ in \r{pdiffR}
1374: are cancelled and so the equations decouple.
1375: This means that in terms of $\;\tQ\,,\;$ the difference equations (\ref{pdiff}) admit the
1376: {\it separation of variables}:
1377: \be\tQ (\bdr'_{n}|\,{\bf E})\;=\;\prod_{k=1}^{n-1}\: \tilde
1378: q_k(\rho_{n,k}).
1379: \ee
1380: Inserting the explicit expressions for $\varphi_k({\bdr'}^{-k}_{n})$
1381: and $\tilde\varphi_k({\bdr'}^{+k}_{n})$
1382: we obtain Baxter type difference equations for the functions $\qt_k(\rho_{n,k})$:
1383: \be\fl
1384: t_n(\lm_{n,k}|\rho,{\bf E})\;\qt_k(\rho_{n,k})\;=\;
1385: \Delta_+(\lm_{n,k})\;\qt_k(\rho_{n,k}+1)\;+\;\Delta_-(\om\lm_{n,k})\;
1386: \qt_k(\rho_{n,k}-1)\label{BAX}\ee
1387: with\\ [-9mm]
1388: \be \fl\Delta_+(\lm)\:=\:(\om^\rho/\chi_k)\,(\lm/\om)^{1-n}\:
1389:          \prod_{m=1}^{n-1}\,F_m(\lm/\om);\qquad
1390: \Delta_-(\lm)\:=\:\chi_k\:(\lm/\om)^{n-1}\:F_n(\lm/\om); \label{Dpm}\ee
1391: \be
1392: \chi_k\;=\;\frac{r_{n,0}\:\rt_{n-1}}{r_n\:\rt_n}\;\:(\prod_{s=1\atop s\ne
1393: k}^{n-1}\; y_{n,k}^{n,s}/y^{n,k}_{n,s})\:\prod_{s=1}^{n-2}\:y^{n,k}_{n-1,s}\,.
1394: \label{chi} \ee
1395: In what follows we will mainly use $t(\lm)$ instead $t_n(\lm|\rho,{\bf E})$.
1396: In fact the system of linear homogeneous equations (\ref{BAX}) with respect
1397: to $\qt_k(\rho_{n,k})$, $\rho_{n,k}\in\ZN$, is not completely defined.
1398: Since $E_1,$ $E_2,\ldots,$ $E_{n-1}$ are unknown, the coefficients $t(\lm_{n,k})$
1399: are also unknown.
1400: The requirement on the system of homogeneous equations (\ref{BAX}) for some
1401: fixed $k$, $k=1,$ $2,\ldots,$ $n-1$,
1402: to have a nontrivial solution leads to the requirement that the matrix of coefficients
1403: must be degenerate.
1404: The latter gives a relation for the values $E_0,$ $E_1,\ldots,$
1405: $E_n$ entering $\:t(\lm)$.
1406: Taking all such relations corresponding to all $k=1,$ $2,\ldots,$ $n-1$,
1407: and using the values of $E_0$ and $E_n$ given in (\ref{iE0En}),
1408: at least in principle we can find the
1409: possible values of $E_1,\ldots,$ $E_{n-1}$. This fixes $\,t(\lm)$.
1410: Then for every $k$, $k=1,$ $2,\ldots,$ $n-1$, we solve (\ref{BAX})
1411: to find $\qt_k(\rho_{n,k})$ for $\rho_{n,k}\in\ZN\,$
1412: (These difference equations have three terms and
1413: cannot be solved in terms of the functions $w_p$).
1414: This gives us finally  $Q(\bdr'_{n}|\,\rho,{\bf E})$ and therefore the
1415: eigenvectors of the periodic \BBS model:
1416: \[
1417: \Phi_{\rho,{\bf E}}=
1418: \sum_{\bdr_n=(\rho_{n,0},\ \bdr'_n)}
1419: \om^{-\rho\cdot\rho_{n,0}}\, Q (\bdr'_{n}|\,\rho,{\bf E})\, \Psi_{\bdr_n}.
1420: \]
1421: 
1422: \subsection{Role of the functional relations}
1423: Now we will show that mentioned requirement on the systems of
1424: homogeneous equations (\ref{BAX}) for all $k$ to have a nontrivial
1425: solution is equivalent to functional relations \cite{BS,BBP,B_tau}
1426: of the $\tau^{(2)}$-model. We define $\tau^{(0)}(\lm)=0$,
1427: $\tau^{(1)}(\lm)=1$, $\tau^{(2)}(\lm)=t(\lm)$ (see \r{tnh}, \r{iE0En}) and
1428: \be\fl \tau^{(j+1)}(\lm)=\tau^{(2)}(\om^{j-1}\lm)\,\tau^{(j)}(\lm)-
1429: \om^{\rho}\,z(\om^{j-1}\lm)\, \tau^{(j-1)}(\lm),\qquad j=2,3,\ldots,
1430: N, \label{rectau} \ee
1431: where \\ [-11mm]
1432: \be\label{zDelta}
1433:  z(\lm)=\om^{-\rho} \Delta_+(\lm)\Delta_-(\lm)\:=\:\prod_{m=1}^{n} F_m(\lm/\om).
1434: \ee
1435: The appearance of the monodromy determinant (\ref{qdet}) in the fusion relation
1436: is  a direct consequence of the fusion procedure (see \cite{KS82,KR87}).
1437: 
1438: The fusion hierarchy can be used to find eigenvalues of the transfer matrices
1439: in lattice integrable models. A key tools here is, in addition to
1440: (\ref{rectau}) to demand a ``truncation'' identity which allows to express $\tau^{(j)}(\lm)$
1441: for some value $j$ through $\tau^{(i)}(\lm)$ with $i<j$. A combination of the
1442: fusion hierarchy and truncation identity allows one to obtain an equation
1443: for $\tau^{(2)}(\lm)=t(\lm)$. This method was applied to many integrable
1444: models, in particular, to the RSOS models in  \cite{BR89}
1445: and to the root of unity  lattice vertex models
1446: in  \cite{N2003}. The functional relations for the $\tau^{(2)}$-model for $N=3$
1447: and the superintegrable  case  were first guessed in \cite{AMCP}
1448: and have been solved to some extend in \cite{McCR90}.
1449: 
1450: The goal of the present section is to prove that the relations to determine the
1451: values $E_1,\ldots,E_{n-1}$ entering $t(\lm)$ also have the form of a truncation identity.
1452: We formulate this statement as follows:
1453: \begin{theorem}\label{th2}
1454: The polynomial $\tau^{(N+1)}(\lm)$ satisfies the ``truncation'' identity
1455: \be\label{reltau}
1456: \tau^{(N+1)}(\lm)-
1457: \om^{\rho}\,z(\lm)\, \tau^{(N-1)}(\om\lm)=
1458: {\cal A}_n(\lm^N)+{\cal D}_n(\lm^N)
1459: \ee
1460: if and only if the system of homogeneous equations (\ref{BAX}) for all $k$
1461: has a nontrivial solution.
1462: \end{theorem}
1463: Note, the classical polynomial $\;{\cal A}_n(\lm^N)+{\cal D}_n(\lm^N)\;$ corresponds
1464: to $\;\alpha_q+\bar \alpha_q\;$ in \cite{B_tau}.
1465: 
1466: \medskip
1467: \noindent {\it\bf Proof.}
1468: Let $t(\lm)$ be a polynomial \r{tnh}, \r{iE0En} such that
1469: the systems of homogeneous equations (\ref{BAX}) for all $k$
1470: have a nontrivial solution.
1471: We shall show that the polynomial $P(\lm)=\tau^{(N+1)}(\lm)-
1472: \om^{\rho}\,z(\lm)\, \tau^{(N-1)}(\om\lm)$ at the left-hand side of
1473: (\ref{reltau}) is equal to ${\cal A}_n(\lm^N)+{\cal D}_n(\lm^N)$
1474: With this aim we introduce the matrix
1475: \[
1476: \Gamma(\lm)\:=\:\lk\ba{cc}\tau^{(2)}(\lm)& \om^\rho z(\lm)\\ -1& 0\ea \rk.
1477: \]
1478: Then it is easy to verify from \r{rectau} by induction that
1479: \[
1480: \Gamma(\om^{j-1}\, \lm)\: \cdots\: \Gamma(\om \lm)\: \Gamma(\lm)\:=\:\lk\ba{cc}
1481: \tau^{(j+1)}(\lm)& \om^\rho z(\lm)\, \tau^{(j)}(\om \lm)\\
1482: -\tau^{(j)}(\lm)\;& -\om^\rho z(\lm)\, \tau^{(j-1)}(\om \lm)\ea \rk
1483: \]
1484: and we see that
1485: \be\label{Gprod}
1486: P(\lm)\;=\;\tr\: \Gamma(\om^{N-1}\, \lm)\: \cdots\: \Gamma(\om \lm)\: \Gamma(\lm)\,.\ee
1487: This relation shows the invariance of $P(\lm)$ under cyclic shifting $\lm\to\om\lm$.
1488: It means that in fact $P(\lm)$ depends only on $\lm^N$. We denote ${\cal P}(\lm^N)=P(\lm)$.
1489: Thus we have to show that $\;{\cal P}(\lm^N)={\cal A}_n(\lm^N)+{\cal D}_n(\lm^N)$.
1490: The direct calculation gives that the coefficients of $\lm^0$ and $\lm^{Nn}$
1491: at both sides of this equation coincide.
1492: In order to calculate the coefficient in front of $\lm^0$ in the trace
1493: of the  product of $\Gamma$-matrices (\ref{Gprod}) one has to substitute
1494: \be\label{Gzero}
1495: \Gamma(\lm)\to \lk\ba{cc}1 +\om^\rho \prod_{m=1}^n b_m d_m/\kp_m \hq&
1496: \om^\rho \prod_{m=1}^n b_m d_m/\kp_m\\ -1& 0\ea \rk
1497: \ee
1498: where only the lowest terms in $\lm$ in the matrix elements were kept.
1499: Therefore the lowest term in $\lm$ in $P(\lm)$ is
1500: \[
1501: \tr \lk\ba{cc}1 +\om^\rho \prod_{m=1}^n \frac{b_m d_m}{\kp_m}&
1502: \om^\rho \prod_{m=1}^n \frac{b_m d_m}{\kp_m}\\ -1& 0\ea \rk^N.
1503: \]
1504: Finding the eigenvalues of the matrix (\ref{Gzero}) one can easily calculate
1505: the latter trace which is the lowest term ${\cal P}(0)$ and identify it
1506: with the lowest term
1507: \[
1508: {\cal A}_n(0)+{\cal D}_n(0)=1+\prod_{m=1}^n \frac{b_m^N d_m^N}{\kp_m^N}
1509: \]
1510: of the polynomial ${\cal A}_n(\lm^N)+{\cal D}_n(\lm^N)$ calculated by means of relations
1511: (\ref{Lclass}) and  (\ref{Tclass}).
1512: The case of the coefficients in front of $\lm^{Nn}$ can be treated analogously.
1513: 
1514: 
1515: Therefore to prove the Theorem~\ref{th2} it remains to prove
1516: \be\label{Plmnk}
1517: P(\lm_{n,k})={\cal A}_n(\lm_{n,k}^N)+{\cal D}_n(\lm_{n,k}^N), \qquad
1518: k=1, 2,\ldots, n-1\,,
1519: \ee
1520: where $\lm_{n,k}$ are given by \r{roots} and
1521:  $\lm_{n,k}^N=\epsilon r^N_{n,k}$ are zeros of the polynomial
1522: (\ref{Brec}).
1523: Let us fix some $k$ and $\rho_{n,k}$ and denote the matrix of the coefficients
1524: of (\ref{BAX}) with respect to the variables
1525: $\;\tilde q_k(\rho_{n,k})$, $\;\tilde q_k(\rho_{n,k}-1),\: \ldots$,
1526: $\:\tilde q_k(\rho_{n,k}-N+1)\;$ by $\;\:{\cal M}$:\\ [-3mm]
1527: \be\label{matM} {\cal M}\:=\:\lk\BAR{cccccc} \!t_0\;&-\Delta_1^-&0&\ldots&0&-\Delta^+_0\\
1528:   -\Delta^+_1&\!t_1\;&-\Delta^-_2&\ldots&0&0\\  0&-\Delta^+_2&t_2&\ldots&0&0\\
1529:  \multicolumn{6}{c}{\ldots \qquad \ldots\qquad \ldots}\\
1530:  -\Delta^-_0&0&0&\ldots&-\Delta^+_{N-1}&t_{N-1}\! \EAR\rk, \ee
1531: where we abbreviated:
1532: $\;t_j=t(\om^j\lm_{n,k});\;\;\Delta^\pm_j=\Delta_\pm(\om^j\lm_{n,k})\,.$
1533: In order (\ref{BAX}) to have a nontrivial solution,
1534: the matrix $\cal M$ must be degenerate.
1535: Let $\cal M'$ be the matrix which has the same matrix elements as $\cal M$
1536: except for ${\cal M}'_{1,N}=0$ and ${\cal M}'_{N,1}=0$.
1537: Then, using the recursive definition (\ref{rectau})
1538: of $\tau^{(j)}(\lm)$ and \r{zDelta}, it is easy to show
1539: that the principal minor corresponding to the first $j$, $j\le N$ rows
1540: of the matrix $\:{\cal M'}\:$ gives $\:\tau^{(j+1)}(\lm_{n,k})$.
1541: Calculating the determinant of the matrix \r{matM} and equating it to zero
1542: we obtain
1543: \bea \det {\cal M}&=&\tau^{(N+1)}(\lm_{n,k})-
1544: \om^{\rho}z(\lm_{n,k})\tau^{(N-1)}(\om\lm_{n,k})\ny\\
1545: &-&\!\!\av \Delta_+(\lm_{n,k} \om^s)
1546: -\av \Delta_-(\lm_{n,k} \om^s)\,=\,0\,. \label{detM}\eea
1547: Further,
1548: \[\fl\av \Delta_-(\lm_{n,k} \om^s)=\chi_k^N (-1)^{n-1} r_{n,k}^{N(n-1)}\:
1549:        \det {\cal L}_n(\lm_{n,k}^N)\;=\;
1550:        \epsilon\frac{{\cal B}_{n-1}(\lm_{n,k}^N)}{r_n^N \lm_{n,k}^N}\;
1551:        \det{\cal L}_n(\lm_{n,k}^N), \]
1552: where we have used \r{Dpm}, \r{detlN},  \r{chi}, \r{xxx}, \r{Brec} and
1553: \[ \chi_k^N\,=\,\frac{(-1)^n\: r_{n-1,0}^N}{r_n^N\; (r_{n,k}^N)^{n-1}}\;\:
1554: \prod_{s=1}^{n-2}\;(r_{n-1,s}^N-r_{n,k}^N)\:=\:
1555: \frac{(-1)^{n-1} {\cal B}_{n-1}(\lm_{n,k}^N)}
1556: {r_n^N\: (r_{n,k}^N)^{n}}\,.
1557: \]
1558: Evaluating (\ref{calAB}) at $m=n$ and $\,\lm=\lm_{n,k}\,$ so that
1559: $\,{\cal B}_{n}(\lm_{n,k}^N)=0\,,$ finally we obtain
1560: \[
1561: \av \Delta_-(\lm_{n,k} \om^s)={\cal A}_n(\lm_{n,k}^N).
1562: \]
1563: Substituting $\lm=\lm_{n,k}$ into
1564: \bea \fl \quad\det\,{\cal T}_n(\lm^N)={\cal A}_n(\lm^N)\,{\cal D}_n(\lm^N)-
1565: {\cal B}_n(\lm^N)\,{\cal C}_n(\lm^N) =
1566:   \prod_{m=1}^n \det {\cal L}_m(\lm^N)\ny\\[-3mm]\fl\qquad\qquad
1567:    = \prod_{m=1}^n \av F_m(\lm\,\om^{s-1}) = \av z(\lm\,\om^s) =\av \lk\,\Delta_+(\lm\,\om^s)
1568:          \cdot\Delta_-(\lm\,\om^s)\,\rk.\eea
1569: we get\\[-11mm]
1570: \[
1571: \av \Delta_+(\lm_{n,k} \om^s)={\cal D}_n(\lm_{n,k}^N).
1572: \]
1573: Using (\ref{detM}) we obtain (\ref{Plmnk}).
1574: 
1575: Conversely, if we have the polynomials $\tau^{(N-1)}(\lm)$ and $\tau^{(N+1)}(\lm)$
1576: built from $\tau^{(2)}(\lm)=t(\lm)$ (see \r{tnh}, \r{iE0En}) by the recursion \r{rectau}
1577: and satisfying \r{reltau} we get \r{detM} at particular values of $\lm$.
1578: This means that the Baxter equations \r{BAX} have nontrivial solutions.
1579: 
1580: This completes the proof of Theorem~\ref{th2}.
1581: 
1582: \section{Periodic homogeneous \BBS model for $N=2$}
1583: \label{perN2}
1584: \subsection{Solution of the Baxter equations}
1585: 
1586: In this section we consider in more detail
1587:  the case of the $N=2$ periodic homogeneous \BBS model, where $\;\om=-1$.
1588: By homogenous we mean that the parameters $a$, $b$, $c$, $d$ and $\kp$ are taken to be the same
1589:  for all sites. As it was shown in \cite{Bugrij}, for
1590: $N=2$ and with arbitrary homogeneous parameters this model is a particular case (``free
1591: fermion point'') of the generalized Ising model.
1592: 
1593: We will find the eigenvalues $t_n(\lm|\rho,{\bf E})$ of the transfer-matrix
1594: ${\bf t}_n(\lm)$ using a functional relation (see also \cite{CPMBaxSol}, where
1595: a similar calculation is presented). We use the short notation $\,t(\lm)\,$ for
1596: $\:t_n(\lm|\rho,{\bf E})$.
1597: From the previous section we have
1598: \be\label{tlm}\fl
1599: t(\lm)=1+(-1)^\rho\frac{b^n d^n}{\kp^n}+E_1\lm+\cdots+E_{n-1}\lm^{n-1}
1600: +\lm^{n} (a^n c^n+(-1)^\rho \kp^n).
1601: \ee
1602: Using \r{rectau} for $j=2$ and eliminating $\tau^{(3)}$ by \r{reltau}
1603: we get the functional relation
1604: \be\label{frN2}
1605: t(\lm)\:t(-\lm)= (-1)^\rho (z(\lm)+z(-\lm))+{\cal A}_n(\lm^2)+{\cal D}_n(\lm^2)
1606: \ee  which we shall use to find $t(\lm)$. Equivalently, we could have obtained
1607: \r{frN2} by multiplying together the two
1608: Baxter equations \r{BAX} for $\lm_{n,k}=\pm r_{n,k}$.
1609: 
1610: Postponing for a moment the derivation (which will be supplied after \r{acac}),
1611: let us anticipate that \r{frN2} can be rewritten as
1612: \be\label{ttABC}
1613: t(\lm)\:t(-\lm)=(-1)^n \prod_q
1614: \left ( A(q)\,\lm^2\,-\,C(q)\,+2{\rm i}\,B(q) \lm\right),
1615: \ee
1616: where\\ [-10mm]
1617: \bea  \lefteqn{ A(q)\,=\,a^2\,c^2\,-2\kp\, ac\,\cos q\,+\kp^2;\hs
1618: B(q)\,=\,(ad-bc)\sin q\,;}\ny\\ [2mm] &&
1619: \hs\hs C(q)=1-2\frac{bd}{\kp}\cos q + \frac{b^2d^2}{\kp^2}\,,\eea
1620: $q$ is running over the set $\;\pi (2s+1-\rho)/n\,$, $\,s=0,1,\ldots,n-1$.
1621: Factorizing (\ref{ttABC}) we get
1622: \be t(\lm)\,t(-\lm)=(-1)^n\prod_q A(q)(\la\,-\la_q)(\la\,+\la_{-q}) \label{abc} \ee
1623: with   \\ [-10mm]
1624: \be  \lm_q=\frac{1}{A(q)}(\sqrt{D(q)}\,-{\rm i} B(q)),\qquad
1625: D(q)\;=\;A(q)\,C(q)\,-B(q)^2.        \label{acb}\ee
1626: We fix the sign of $\:\sqrt{D(q)}\:$ by the conditions
1627: \bea  \sqrt{D(q)}&=&\sqrt{D(-q)},\qquad q\ne 0,\pi;
1628: \label{dq1}\\ [2mm]
1629:  \sqrt{D(0)}&=&(\kp\,-ac)(1\,-bd/\kp);\quad \sqrt{D(\pi)}\;=\;(\kp\,+ac)(1\,+bd/\kp).
1630: \label{dq}\eea
1631: In what follows we shall need the relations
1632: \be\label{prodA} \prod_{q} A(q)\,=\, \prod_{q}
1633: (\kp-e^{{\rm i}q}ac)(\kp-e^{-{\rm i}q}ac)= (a^n c^n+(-1)^\rho\kp^n)^2\,,\ee
1634: \be\label{eqDB}
1635: \prod_q (\sqrt{D(q)}-{\rm i} B(q)) =
1636: \prod_q (\kp-e^{{\rm i}q}ac)(1-e^{{\rm i}q} bd/\kp).
1637: \ee
1638: The last relation can be obtained by grouping terms with
1639: opposite signs of $q$ (modulo $2\pi$) and using the definition of $\sqrt{D(q)}$.
1640: Using (\ref{prodA}) we get
1641: \bea
1642: t(\lm)t(-\lm)&=&(-1)^n (a^n c^n+(-1)^\rho\kp^n)^2
1643: \prod_{q} (\lm-\lm_q)\cdot \prod_{q}(\lm+\lm_{-q})\ny\\
1644: &=&(-1)^n (a^n c^n+(-1)^\rho\kp^n)^2 \prod_{q} (\lm^2-\lm_q^2),
1645: \label{ttlm}
1646: \eea
1647: where we made the change $q\to -q$ in second product.
1648: From (\ref{tlm}) it follows that $t(\lm)$ can be presented as
1649: \[
1650: t(\lm)=(a^n c^n+(-1)^\rho\kp^n)\prod_{s=1}^{n}(\lm-\Lambda_s)
1651: \]
1652: with zeroes $\Lambda_s$. Therefore
1653: \[
1654: t(\lm)\:t(-\lm)=(-1)^n(a^n c^n+(-1)^\rho\kp^n)^2\:\prod_{s=1}^{n}\:(\lm^2-\Lambda_s^2).
1655: \]
1656: Comparing with (\ref{ttlm}) we obtain
1657: \be\label{tNSR}
1658: t(\lm)=(a^n c^n+(-1)^\rho\kp^n)\prod_{q} (\lm\pm\lm_q),
1659: \ee
1660: where the signs are not yet fixed. To fix these signs we compare
1661: the $\la$-independent term in (\ref{tlm})
1662: with the corresponding term in (\ref{tNSR}). The latter can be found using
1663: \bea  \fl \lefteqn{(a^n c^n+(-1)^\rho\kp^n) \prod_{q} \lm_q\;=\;
1664: (a^n c^n+(-1)^\rho\kp^n)^{-1} \prod_{q} (\sqrt{D(q)}-{\rm i} B(q))}\ny\\
1665: &&\hspace*{-15mm}=\;(a^n c^n+(-1)^\rho\kp^n)^{-1}
1666: \prod_{q} (\kp-e^{{\rm i}q}ac)(1-e^{{\rm i}q} bd/\kp)\;=\;
1667: (-1)^\rho+ b^nd^n/\kp^n\,,\label{acac}\eea
1668: where we took into account (\ref{eqDB}).
1669: Therefore the number of minus signs in (\ref{tNSR}) must be even for the sector
1670: $\rho=0$ and odd for $\rho=1$.
1671: Thus we have obtained $2^n$ eigenvalues
1672: ($2^{n-1}$ each for both $\rho=0$ and $\rho=1$).
1673: These eigenvalues provide the existence of nontrivial solutions of the system
1674: \r{BAX} of homogeneous equations. These solutions give
1675: the eigenvectors (\ref{EV}), a basis in the space of states of the
1676: periodic \BBS model for $N=2$.
1677: 
1678: We conclude this section supplying the derivation of (\ref{ttABC}) from (\ref{frN2}):
1679: Using
1680: \[\fl
1681: \delta_+(\lm):={(b+a \kp \lm)(d- c\kp \lm)}/{\kp},\qquad
1682: \delta_-(\lm):=\delta_+(-\lm)={(b-a \kp \lm)(d+ c\kp \lm)}/{\kp}
1683: \]
1684: we easily verify the relations
1685: $\delta_+^n(\lm)=z(\lm)$, $\delta_-^n(\lm)=z(-\lm)$,
1686: $\delta_+(\lm)\delta_-(\lm)= \delta(\lm^2)$,
1687: where $z(\lm)$ and $\delta(\lm^2)$ are given by (\ref{zDelta}) and
1688: (\ref{deltalm}) respectively.
1689: Taking into account
1690: ${\cal A}_n(\lm^2)+{\cal D}_n(\lm^2)=\tr \left({\cal L}(\lm^2)\right)^n=
1691: x_+^n(\lm^2)+x_-^n(\lm^2)$ and the relations
1692: \[\fl
1693: \tau(\lm^2)= \tr{\cal L}(\lm^2)= x_+(\lm^2)+ x_-(\lm^2),\qquad
1694: \delta(\lm^2)= \det{\cal L}(\lm^2)= x_+(\lm^2) x_-(\lm^2),
1695: \]
1696: we can rewrite the functional relation  (\ref{frN2}) as
1697: \bea
1698: \fl \qquad\quad t(\lm)t(-\lm)&=& (-1)^\rho (z(\lm)+z(-\lm))+{\cal A}_n(\lm^2)+{\cal D}_n(\lm^2)\ny\\
1699: &=& (-1)^\rho \left(\delta_+^n+\delta_-^n\right)+x_+^n+x_-^n \ny\\ [3mm]
1700: &=&(-1)^\rho\left( x_+^n+(-1)^\rho\delta_+^n\right)
1701: \left((x_-/\delta_+)^n+ (-1)^\rho\right)\ny\\ [4mm]
1702: &=&(-1)^\rho\prod_{q}\; (x_+ -e^{{\rm i}q} \delta_+)(x_-/\delta_+-e^{{\rm i}q})\ny\\
1703: &=&(-1)^{n} \prod_{q}\; (\: e^{{\rm i}q} \delta_+-\tau(\lm^2)+
1704: e^{-{\rm i}q} \delta_-\:)\ny\\ [-2mm]
1705: &=&(-1)^n \prod_{q}\;
1706: \lb\lk(a^2c^2+\kp^2)\lm^2-\frac{b^2d^2}{\kp^2}-1\rk
1707: +2\lk\frac{bd}{\kp}-\lm^2\kp\, a\, c\rk\cos q \right.\ny\\ [-2mm]
1708:  &&\hspace*{3cm}\left.-2{\rm i}\,\lm (a\,d\,-b\,c)\,\sin q\rb,
1709: \eea
1710: which confirms (\ref{ttABC}).
1711: 
1712: \subsection{Relation to the standard Ising model notations}
1713: 
1714: 
1715: In the homogeneous $N=2$ case we have $\om=-1$ and $\hu_k^{-1}=\hu_k$,
1716: so the cyclic $L$-operator \r{bazh_strog} reduces to
1717: \be
1718: L_k(\lm)=\lk\ba{cc}
1719: 1+\lm\, \kp\, \hv_k &  \lm \,\hu_k\, (a\,-b \hv_k)\\
1720: \hu_k\, (c\,-d\, \hv_k) & \lm a c + \hv_k\, b\, d/\kp\ea \rk.\ee
1721: Let us make the special choice of the parameters $\;d=bc/a\;$ and $\:\lm=b/(a\kp)$. Then
1722: \be \fl L_k(\lm)=(1+{\bf v}_k\,b/a)\lk\ba{cc}
1723: 1 &  \hu_k {b}/{\kp}\\c \hu_k & {b c}/{\kp }\ea \rk\,=\,
1724: (1+{\bf v}_k \,b/a)
1725: \lk \ba{c} 1\\ c \hu_k \ea \rk
1726: \lk \ba{cc} 1\; &  \hu_k {b}/\kp\ea \rk.\ee
1727:  and the transfer-matrix is
1728: \be\fl
1729: {\bf t}_n(\lm)=\tr\; L_1(\lm)\,L_2(\lm)\cdots L_n(\lm)=
1730: \prod_{k=1}^n (1+{\bf v}_k \cdot {b}/{a})\cdot
1731: \prod_{k=1}^n (1+\hu_{k-1}\hu_{k}\cdot bc/\kp).
1732: \ee
1733: Recall that due to the periodic boundary conditions $\hu_{n+1}\equiv\hu_{k}$.
1734: Using
1735: \[\fl
1736: \exp (K_1 \hu_{k-1}\hu_{k})=\cosh K_1 (1\!+\hu_{k-1}\hu_{k} \tanh K_1);\;
1737: \exp (K^*_2 \hv_k)=\cosh K^*_2 (1\!+\hv_k \tanh K^*_2),
1738: \]
1739: and writing $\:\hu_k\,=\:\sigma^z_k\,$ and $\:\hv_k\,=\:\sigma^x_k$,
1740: it is easy to identify ${\bf t}_n(\lm)$ with the standard Ising
1741: transfer-matrix:
1742: \[ \fl
1743: {\bf t}_{\rm Ising}=\exp{\lk \sum_{k=1}^n\, K^*_2\, \sigma^x_k\rk}\:
1744: \exp{ \lk\sum_{k=1}^n\, K_1 \,\sigma^z_{k-1}\,\sigma^z_k\rk};\quad\;
1745: \tanh K^*_2=\frac{b}{a};\quad\; \tanh K_1=\frac{bc}{\kp}\,.
1746: \]
1747: 
1748: 
1749: 
1750: 
1751: 
1752: \section{Conclusion}\label{discus}
1753: 
1754: This paper is devoted to the solution of the eigenvalue and eigenvector problems
1755: for the finite-size inhomogenous periodic Baxter-Bazhanov-Stroganov
1756: quantum chain model. We use an approach which had been developed in full detail for
1757: the quantum Toda chain in \cite{KarLeb1,KarLeb2} and in \cite{KhLS} for
1758: the relativistic deformation of the Toda chain. This approach consists of two
1759: main steps: In order to find eigenvectors for the transfer matrix $A_n(\lm)+D_n(\lm)$
1760: first we find the eigenvectors of the off-diagonal operator $B_n(\lm)$ adapting
1761: the well known recurrent procedure described in \cite{KarLeb2}
1762: to our root-of-unity case. Then, using these
1763: eigenvectors, we construct the eigenvectors for the \BBS transfer matrix and show that the
1764: coefficients of the decompositions of one set of eigenvectors in terms of the other set
1765: factorizes into a product of single variable functions, each satisfying the Baxter
1766: type equation. We show that the condition for these equations to have nontrivial solution
1767: is equivalent to the functional relations for the transfer matrix eigenvalues in
1768: the \BBS or $\tau^{(2)}$ model. In case of $N=2$ the Baxter equation can be solved
1769: and as result we obtain the eigenstates of the transfer matrix of the generalized Ising model
1770: at the  free fermion point. We shortly give the relation of the $N=2$ \BBS model parameters
1771: to the standard Ising model parametrization.
1772: 
1773: 
1774: 
1775: 
1776: 
1777: \section*{Acknowledgements}
1778: This work was partially supported by the grant
1779: INTAS-OPEN-03-51-3350, the
1780: Heisenberg-Landau program and France--Ukrainian project ``Dnipro'' and the grant
1781: RFBR-05-01-086.
1782: 
1783: \appendix
1784: \setcounter{section}{1}
1785: \section*{Appendix: Amplitudes $\:r_{m,k}\:$ in the homogeneous case}
1786: 
1787: The determination of the amplitudes $r_{m,k}$ for the inhomogenous
1788: \BBS model had been reduced to solving Equation \r{Brec} with \r{relBm}
1789: of Section \ref{recinh}.
1790: Here we show that  in the {\it homogenous} case this task simplifies to
1791: solving just one quadratic equation, using a trigonometric parametrization.
1792: 
1793: In the homogeneous case we have
1794: \be \fl a_m=a,\;\;b_m=b,\;\;c_m=c,\;\;d_m=d,\;\;\kp_m=\kp,\;\;r_m=r,\;\;
1795: {\cal L}_m(\lm^N)={\cal L}(\lm^N)\label{hom}\ee and\\[-9mm]
1796: \be\label{calABCDhom}
1797: \left(\ba{cc}
1798: {\cal A}_m(\lm^N) &  {\cal B}_m(\lm^N)\\
1799: {\cal C}_m(\lm^N) &  {\cal D}_m(\lm^N)
1800: \ea\right)\;=\;\left({\cal L}(\lm^N)\right)^m\!.
1801: \ee
1802: Using the fact that a $2\times 2$ matrix $\:\bf M\:$ with eigenvalues
1803: $\mu_+$ and $\mu_-$ satisfies
1804: \[
1805: {\bf M}^m\;=\:\frac{\mu_+^m-\mu_-^m}{\mu_+-\mu_-}\;{\bf M}
1806: \:-\,\frac{\mu_+^m \mu_--\mu_-^m \mu_+}{\mu_+-\mu_-}\;{\bf 1}\,,
1807: \]
1808: from the matrix $\;{\cal L}(\lm^N)\;$ we obtain
1809: \bdm {\cal B}_m(\lm^N)\;=\;-\epsilon\, \lm^N r^N\,
1810: \frac{x_+^m-x_-^m}{x_+-x_-},
1811: \edm
1812: where $x_+(\lm^N)$ and $x_-(\lm^N)$ are the eigenvalues of $\:{\cal L}(\lm^N)$.
1813: These eigenvalues are the roots of the characteristic polynomial
1814: $\;x^2\,-\tau(\lm^N)\, x\, +
1815: \delta(\lm^N)\:=\,0\,$:
1816: %
1817: \be x_{\pm}\;=\;\frac{1}{2}(\,\tau\pm \sqrt{\tau^2\,-4\delta}\,),\label{xpm}  \ee
1818: where, see \r{detlN},
1819: \be\label{taulm}
1820: \tau(\lm^N)=\tr {\cal L}(\lm^N)=
1821: 1+\frac{b^N d^N}{\kp^N} - \epsilon\lm^N (\kp^N + a^N c^N), \ee
1822: \be\label{deltalm} \delta(\lm^N)= \det\, {\cal L}(\lm^N)\: =\:
1823: (b^N/\kp^N-\epsilon\lm^N\,a^N)\,(d^N-\epsilon\lm^N\,c^N\,\kp^N). \ee
1824: 
1825: Introducing the variable $\phi$ by $x_+/x_-=e^{{\rm i}\phi}$ we
1826: find that roots of ${\cal B}_m$ correspond to roots $\phi_{m,s}$
1827: of $e^{{\rm i}m\phi}=1$ (without $\phi=0$):
1828: \be
1829: \label{zerphi} \phi_{m,s}=2\pi s/m, \qquad s=1,2,\ldots,m-1.
1830: \ee
1831: Now we need to find the explicit relation between $\lm^N$ and
1832: $\:\phi$. We have
1833: %
1834: \be \tau+\sqrt{\tau^2-4\delta}\;=\;e^{{\rm i}\phi}\, (\tau-
1835: \sqrt{\tau^2-4\delta})\hs\mbox{or}\hs
1836: \tau^2=4 \delta \cos^2 \frac \phi 2\label{tdphi}.\ee
1837:  Taking into account (\ref{taulm}) and (\ref{deltalm}), we
1838: consider (\ref{tdphi}) as a quadratic equation for $\lm^N$:
1839: %
1840: \bea \fl \lefteqn{\lm^{2N} (a^{2N} c^{2N}+ \kp^{2N}- 2 a^N
1841: c^N\kp^N\cos\phi)\;+(b^{2N} d^{2N}+ \kp^{2N}- 2 b^N
1842: d^N\kp^N\cos\phi)/\kp^{2N}}\ny\\[2mm]
1843: \fl &-& 2\epsilon \lm^N \lk(a^N-b^N)(c^N-d^N)+\frac{a^N b^N c^N
1844: d^N}{\kp^N}+\kp^N - (a^N d^N+b^N c^N)\cos\phi \rk\;=\;0\,. \eea
1845: %
1846: The solution $\lm^N(\phi)$ of this equation describes the relation
1847: between the variables $\lm^N$ and $\phi$. Therefore we can
1848: translate the zeroes (\ref{zerphi}) of $\:{\cal B}_m(\lm^N(\phi))$
1849: in terms of variable $\phi$, to zeroes $\,\lm^N(\phi_{m,s}).\:$ From
1850:  (\ref{Brec}) we get
1851: \be r_{m,s}^N=\epsilon \lm^N(\phi_{m,s}),\qquad s=1,2,\ldots,m-1. \label{ire0m} \ee
1852: The value of $r_{m,0}^N$ can be found recursively from \r{rel0N} using \r{hom}:
1853: \be \label{irel0N}
1854: r_{m,0}^N= r_{m-1,0}^N\, a^N c^N+ r^N \kp^{N(m-1)}\,,\qquad
1855: r_{1,0}^N= r^N=a^N-b^N\,.\ee
1856: 
1857: 
1858: \section*{References}
1859: \bibliographystyle{amsplain}
1860: \begin{thebibliography}{**}
1861: 
1862: \bibitem{B_tau}
1863: Baxter R J 2004  \emph{Transfer matrix functional relations for
1864:  the generalized $\tau_2(t_q)$ model}, \JStP~ \textbf{117}~ 1--25
1865: %cond-mat/0409493;
1866: 
1867: \bibitem{BaxInv} Baxter R J 1989 \emph{Superintegrable chiral Potts model:
1868: Thermodynamic properties, an "Inverse" model, and a simple
1869: associated Hamiltonian}
1870:  ~\JStP~ {\bf 57} 1--39
1871: 
1872: \bibitem{BS}
1873: Bazhanov V V and Stroganov Yu G 1990 \emph{Chiral Potts model as a
1874: descendant of the six-vertex model}, \JStP~ \textbf{59}~ 799--817
1875: 
1876: \bibitem{BBP}
1877: Baxter R J, Bazhanov V V and Perk J H H 1990 \emph{Functional
1878: relations for the transfer matrices of the Chiral Potts model},
1879: \IJMP~ {\textbf B 4}~ 803--869
1880: 
1881: \bibitem{BaxtEV}
1882: Baxter R J 1990 \emph{Chiral Potts model: eigenvalues of the transfer matrix},
1883: \PL~ \textbf{A146} 110--114
1884: 
1885: \bibitem{BaxtOP}
1886: Baxter R J 2005 \emph{Derivation of the order parameter of the chiral Potts model},
1887: \PRL~  \textbf{94} 130602; {\tt arXiv:cond-mat/0501227}
1888: 
1889: \bibitem{Kore}
1890: Korepanov I G 1987 \emph{Hidden symmetries in the 6-vertex model},
1891: Chelyabinsk Polytechnical Institute, {\it archive VINITI} No. 1472-V87 (in
1892: Russian);~~ 1994 \emph{Hidden symmetries in the 6-vertex model of
1893: Statistical Physics}, {\it Zapiski Nauchn. Semin. POMI}~ \textbf{215}
1894: 163--177; arXiv:hep-th/9410066;~~ 2000 \emph{The method of
1895: vacuum vectors in the theory of Yang-Baxter equation} arXiv:nlin.SI/0010024
1896: 
1897: \bibitem{Skly1} Sklyanin E K 1990 \emph{Functional Bethe Ansatz} in
1898: \emph{Integrable and Superintegrable Systems}, ed. Kupershmidt B A,
1899: World Scientific, Singapore, 8--33
1900: 
1901: \bibitem{KarLeb1} Kharchev S and Lebedev D 1999 \emph{Integral representations for the
1902: eigenfunctions of a quantum periodic Toda chain}, {\it Lett. Math. Phys.}~
1903: {\textbf 50}~ 53--77
1904: 
1905: \bibitem{KarLeb2} Kharchev S and Lebedev D 2000 \emph{Eigenfunctions of
1906: $GL(N,\mathbb{R})$ Toda chain: The Mellin-Barnes representation},
1907: {\it JETP Lett.}~ {\textbf 71}~ 235 --238
1908: 
1909: \bibitem{KhLS} Kharchev S, Lebedev D, and Semenov-Tian-Shansky M 2002
1910: \emph{Unitary representations of $U_{q}(sl(2,R))$, the modular
1911: double, and the multiparticle $q$-deformed Toda chains},
1912: {\it Comm. Math. Phys.} \textbf{225}~  573--609
1913: 
1914: \bibitem{AMCP} Albertini G, McCoy B M M and Perk J J H 1898 \emph{Eigenvalue spectrum
1915: of the superintegrable chiral Potts model} {\it Adv. Stud. Pure Math.}~
1916: \textbf{19}~ 1--55
1917: 
1918: \bibitem{TarasovL} Tarasov V O 1990 \emph {Transfer matrix of the superintegrable
1919: chiral Potts model. Bethe ansatz spectrum}, \PL {\textbf A147}
1920: ~487--490
1921: 
1922: \bibitem{Tarasov}
1923: Tarasov V O 1992 \emph{Cyclic monodromy matrices for the R-matrix of
1924: the six-vertex model and the chiral Potts model with fixed spin
1925: boundary conditions},
1926: %, Proc.RIMS Research Project 91 on Infinite Analysis (1991) 1--13,
1927: \IJMP~ {\bf A7} Suppl. {\bf 1B}~ 963--975
1928: 
1929: \bibitem{BaxSkew} Baxter R J 1993 \emph{Chiral Potts model with skewed boundary
1930: conditions} \JStP \textbf{73}~  461--495
1931: 
1932: \bibitem{Bugrij} Bugrij A I, Iorgov N Z and Shadura V N  2005
1933: \emph{Alternative Method of Calculating the Eigenvalues of the
1934: Transfer Matrix of the $\tau_2$  Model for $N=2$} {\it JETP Lett.}~
1935: \textbf{82}~ 311 --315
1936: 
1937: \bibitem{Lisovy} Lisovyy O  2006
1938: \emph{ Transfer matrix eigenvectors of the Baxter-Bazhanov-Stroganov $\tau_2$-model for N=2
1939: } \JPA \textbf{39}~ 2265--2285; arXiv:nlin.SI/0512026
1940: 
1941: 
1942: \bibitem{Gu81} Gutzwiller M  1981 \emph{The quantum mechanical Toda lattice II},
1943:  {\it Ann. of Phys.}~ \textbf{133}~  304--331
1944: 
1945: \bibitem{Sk85} Sklyanin E 1985 \emph{The quantum Toda chain} in {\it Lectures Notes in Physics}
1946: \textbf{226}~ 196--233, Springer, New York
1947: 
1948: \bibitem{Skly2} Sklyanin E K 1992 \emph{Quantum Inverse Scattering Method. Selected Topics},
1949: in \emph{Quantum Group and Quantum Integrable Systems}, Nankai
1950: {\it Lectures in Math. Phys.}, ed. Mo-Lin Ge, World Scientific,
1951: Singapore,~ 63--97.
1952: 
1953: \bibitem{Iorgov} Iorgov N Z 2006 \emph{Eigenvectors of open Bazhanov--Stroganov quantum chain},
1954: SIGMA \textbf{2}, No.~19; arXiv:nlin.SI/0602010
1955: 
1956: \bibitem{KS82} Kulish P P and Sklyanin E K 1982 \emph{Lecture notes in Physics}
1957: \textbf{151} (Springer)~ 61
1958: 
1959: \bibitem{KR87} Kirillov A N and Reshetikhin N Yu 1987 \JPA~ \textbf{20}~ 1565
1960: 
1961: \bibitem{BR89} Bazhanov V V and Reshetikhin N Yu 1989 \emph {Critical RSOS models and
1962: conformal field theory}, \IJMP~ \textbf{A4}~ 115
1963: 
1964: \bibitem{N2003} Nepomechie R I 2003 \emph {Functional relations and Bethe ansatz for the XXZ
1965: chain}, \JStP~ {\textbf 111}~ 1363
1966: 
1967: \bibitem{McCR90} McCoy B M M and Roan S S, 1990 \emph{ Excitation spectrum and phase
1968: structure of the chiral Potts model}, \PL~ {\textbf A150}~ 347
1969: 
1970: \bibitem{IorgShad} Iorgov N and Shadura V 2005 \emph{Wave functions of the Toda chain
1971: with boundary interactions}, {\it Theor. Math. Phys.} \textbf{142}~ 289--305;  arXiv:nlin.SI/0411002
1972: 
1973: \bibitem{BB}
1974: Bazhanov V V and Baxter R J 1992 \emph{New solvable lattice models
1975: in three dimensions}, \JStP \textbf{69}~ 453--485
1976: 
1977: \bibitem{GPS} von Gehlen G, Pakuliak S and Sergeev S 2005 \emph{The Bazhanov-Stroganov model
1978: from 3D approach}. \JPA \textbf{38}~ 7269--7298
1979: 
1980: \bibitem{CPMBaxSol}
1981: Au-Yang H, Bai-Qi Jin and Perk J H H 2001 \emph{Baxter's solution
1982: for the free energy of the chiral Potts model}, \JStP~ \textbf{102}~
1983: 471--499
1984: 
1985: \end{thebibliography}
1986: 
1987: \end{document}
1988: