cond-mat0701605/nrg.tex
1: \documentclass[prl,aps,twocolumn,amsfonts,showpacs,floatfix]{revtex4}
2: 
3: 
4: \usepackage{epsfig}
5: \usepackage{psfrag}
6: \usepackage{color}
7: \usepackage{graphicx}
8: \usepackage{subfigure}
9: 
10: \begin{document}
11: 
12: %\twocolumn[\hsize\textwidth\columnwidth\hsize\csname@twocolumnfalse%
13: %\endcsname
14: 
15: 
16: \title{A Numerical Renormalization Group for Continuum One-Dimensional Systems}
17: \author{Robert M. Konik and Yury Adamov}
18: \affiliation{CMPMS Department, Brookhaven National Laboratory, Upton, NY 11973}
19: 
20: \date{\today}
21: 
22: \begin{abstract}
23: We present a renormalization group (RG) procedure which works naturally on 
24: a wide class of interacting one-dimension models based on perturbed (possibly strongly) continuum conformal and integrable models.
25: This procedure integrates Kenneth Wilson's numerical
26: renormalization group with Al. B. Zamolodchikov's truncated conformal spectrum
27: approach.
28: Key to the method is that such theories provide a set of completely understood eigenstates
29: for which matrix elements can be exactly computed.  In this procedure the RG
30: flow of physical observables can be studied both numerically and analytically.
31: To demonstrate the approach, we study the spectrum of a pair of coupled quantum Ising
32: chains and correlation functions in a single quantum Ising chain in the presence of a magnetic field.
33: \end{abstract}
34: \pacs{}
35: \maketitle
36: 
37: The numerical renormalization group (NRG), as developed by Kenneth Wilson \cite{wilson},
38: is a tremendously successful technique for the study of generic quantum impurity problems,
39: systems where interactions are confined to a single point.
40: But the NRG as such is not directly generalizable to systems where interactions are present in the bulk.
41: The natural generalization of the NRG in real space
42: treats boundary conditions between RG blocks inadequately, leading to qualitatively inaccurate
43: results.  To overcome this difficulty, Steven White
44: developed the density matrix renormalization group (DMRG) \cite{white}.  This tool is now ubiquitous
45: in the study of low dimensional strongly correlated {\it lattice} models and
46: can access both static and dynamic quantities \cite{white1}.
47: 
48: In this letter we offer a distinct generalization realizing a renormalization group
49: procedure for a wide range of strongly interacting {\it continuum} one-dimensional systems.
50: It can treat any model which is representable as a conformal or integrable field theory
51: (CFT/IFT),
52: with Hamiltonian, $H_0$,
53: plus a relevant \cite{note} perturbation (of arbitrary strength), $H_{\rm pert}$.
54: Beyond this there is no real constraint on $H_0$ or $H_{\rm pert}$.
55: In particular, the full theory, $H_0 + H_{\rm pert}$, need not be integrable or conformal.
56: Thus the technique can handle a standard array of models of perturbed
57: Luttinger liquids or Mott insulators.   It also capable of treating disordered
58: systems, either by envisioning $H_{\rm pert}$ as a random
59: field or, equally well, considering a non-unitary supersymmetric CFT, $H_0$ arising from
60: disorder averaging a system with quenched disorder \cite{efetov}.   This technique
61: also allows the study of coupled CFT/IFTs, allowing the 
62: study of systems between one and two dimensions.  In all cases, the low energy
63: spectrum and correlation functions of the model are computable.
64: 
65: Our starting point is the truncated spectrum approach (TSA) pioneered by Al. B. Zamolodchikov.
66: The TSA was developed to treat perturbations of simple conformal field theories.
67: While straightforward in conception, it has an advantage over other numerical techniques in that
68: it analytically embeds strongly correlated physics at the start, dramatically
69: lessening the computational burden.  In one of the TSA's first applications,
70: Al. B. Zamolodchikov studied a critical
71: Ising chain in a magnetic field \cite{zamo}, a continuum version
72: of the lattice model
73: \begin{equation}\label{ei}
74: H^{Ising}_0 \!=\! -J\sum_i(\sigma^z_i\sigma^z_{i+1}\!+\!\sigma^x_i); ~H_{\rm pert}\!=\!-h\sum_i\sigma^z_i,
75: \end{equation}
76: where $\sigma^{a}_i$ are the standard Pauli matrices and $i$ indexes the sites of the lattice.
77: The continuum model, itself integrable, has a complicated
78: spectrum of eight fundamental excitations.  The TSA was able to produce
79: the gaps of the first five excitations within 2\% of the analytic, infinite volume values by
80: diagonalizing a mere 39x39 matrix.  To see how remarkable this is, consider that in a 
81: computationally equivalent exact diagonalization of the lattice model,
82: one would be limited to studying a five site chain.
83: 
84: \begin{figure}
85: \includegraphics[height=1.3in,width=3.in,angle=0]{tsa.eps}
86: \caption{A schematic of the finite sized system, both in real space
87: and in terms of energy levels, analyzed in the TSA procedure.}
88: \end{figure}
89: 
90: The TSA begins by taking the model to be studied, $H= H_0+H_{\rm pert}$, and placing
91: it on a finite ring of circumference, R.  Doing so makes the spectrum
92: discrete (see Fig. 1).   We nonetheless expect to be able to obtain {\it infinite} volume
93: results provided we work in a regime where $R\Delta \gg 1$ with $\Delta$ a characteristic
94: energy scale of the system.  In the discrete system, the spectrum can then be ordered in
95: energy, $|1\rangle, |2\rangle, \ldots$.  
96: Non-perturbative information is input in the next step of the TSA where
97: the matrix elements,
98: $H_{{\rm pert}ij} = \langle i| H_{\rm pert}|j\rangle$, are computed {\it exactly}.  It is important to
99: stress this is always a practical possibility.  If $H_0$ is a CFT, the attendant
100: Virasoro algebra (or, equally good, some more involved algebraic structure
101: such as a current or a W-algebra) permits the computation of such matrix elements
102: straightforwardly.  If $H_0$ is instead an IFT, such matrix elements
103: are available through the form-factor bootstrap programme \cite{ffprog}.  In an IFT, the matrix
104: elements which we will want to focus on involve states, $|i\rangle$, with few excitations
105: and, as such, are readily computable.
106: 
107: With the matrix elements in hand, one can then express the full Hamiltonian as
108: a matrix.  The penultimate step in the procedure is to truncate the spectrum
109: at some energy, $E_{\rm trunc}$, making the matrix finite.  This matrix is then numerically diagonalized from which
110: the spectrum and correlations functions can be extracted.  
111: When $H_0$ represents
112: a theory with a relatively simple set of eigenstates, this procedure, even with
113: a crude truncation of states, works remarkable well in extracting the spectrum.  However when the starting
114: point Hamiltonian, $H_0$, is more complicated (say a CFT based on a $SU(2)_k$ current algebra such as would be encountered
115: in the study of spin chains), 
116: or one is interested in computing
117: correlation functions in the full theory, $H_0+H_{\rm pert}$, the simple truncation scheme
118: ceases to produce accurate results at reasonable numerical cost. 
119: For example, errors in an excitation gap, $\Delta$, introduced by the truncation of states
120: behave as power laws, $\delta \Delta \sim E_{\rm trunc}^{-\alpha}$ ($\alpha \sim 1$).
121: Thus increasing $E_c$ does not dramatically reduce $\delta Q$ while
122: at the same time greatly increasing the cost of the exact diagonalization
123: routine which should scale similarly to the partition of integers, i.e. 
124: as $e^{\beta\sqrt{E_{\rm trunc}}}/E_{\rm trunc}$.  It is the aim of this letter
125: to outline an RG technique offering a dramatic improvement on this truncation scheme.
126: 
127: \begin{figure}[tbh]
128: \includegraphics[height=1.1in,width=3.3in,angle=0]{nrgproc.eps}
129: \caption{An outline of the NRG algorithm}
130: \end{figure}
131: 
132: Our framework hews closely to the original Wilsonian conception of the NRG.  In developing
133: the NRG for the Kondo model, Wilson transformed the original Kondo Hamiltonian 
134: using the ``Kondo basis'' to a lattice model of an impurity situated at the end of a
135: infinite half line with sites far from the end characterized by rapidly diminishing
136: matrix elements.  We, in a sense, start in this position.  The ordering in energy of
137: the states provided by $H_0$ is in direct analogy to the half-line on which the impurity
138: lives in Wilson's Kondo work.  The next step in Wilson's NRG is an iterative numerical
139: procedure by which at each step a finite lattice is expanded by one site, the model diagonalized, and
140: high energy eigenstates thrown away.  
141: It is this iterative procedure that we mimic.  
142: \vskip .15in
143: \begin{figure}[tbh]
144: \includegraphics[height=2.2in,width=2.9in,angle=0]{rgfita.eps}
145: \caption{Plots showing the behavior of $\Delta_2$ as a function
146: of the truncation energy, $E_{\rm trunc}$ (for $R=5\tilde\lambda^{-1}$).}
147: \end{figure}
148: \vskip -.1in
149: Let us denote the initial basis offered by $H_0$ as $\{|i\rangle\}^\infty_{i=1}$.
150: We begin by keeping a certain number, say $N+\Delta N$, of 
151: the lowest energy states,$\{|i\rangle\}^{N+\Delta N}_{i=1}$ (in blue in Fig. 2).   
152: We diagonalize the problem so extracting an initial spectrum and set
153: of eigenvectors, $\{|\tilde i\rangle\}^{N+\Delta N}_{i=1}$ (in red in Fig 2).  
154: We then toss away a certain number, $\Delta N$, of the eigenvectors corresponding to
155: the highest energy, i.e. $\{|\tilde i\rangle\}^{N+\Delta N}_{i=N+1}$ .  A new basis is then formed, consisting
156: of the remaining eigenvectors together with
157: the first $\Delta N$ states of $\{|i\rangle\}^\infty_{i=1}$
158: that we had previously ignored, 
159: i.e. $\{|\tilde i\rangle\}^{N}_{i=1} \cup \{|i\rangle\}^{N+2\Delta N}_{i=N+\Delta N+1}$, and the procedure
160: is repeated.  We present the technique schematically in Fig. 2.
161: Convergence of the procedure in the Kondo problem is promoted by the small matrix elements
162: involving sites far from the impurity.  While matrix elements in the procedure
163: just described grow progressively smaller under the NRG (scaling as $1/E_{trunc}$), 
164: here numerical convergence is not necessarily the goal.  
165: Rather we aim to merely bring the quantity into a regime where its flow is governed by a simple flow equation.
166:  
167: The algorithm just described implements a Wilsonian RG in reverse.   It does so at all loop orders
168: and so the RG flow it describes is exact.  As the flow proceeds it, however, evolves closer and closer
169: to a flow described by a one-loop equation.  Analytically, the equation is nearly trivial as it is given
170: solely in terms of the anomalous dimension, $\alpha_Q$, of the flowing quantity, $Q$ (whether
171: it be an energy eigenvalue or a matrix element).
172: More specifically,
173: \begin{equation}\label{eii}
174: \frac{d\Delta Q}{d\ln E_{\rm trunc}} = -g(\alpha_Q)\Delta Q,
175: \end{equation}
176: where $\Delta Q = Q(E_{\rm trunc})-Q(E_{\rm trunc}=\infty)$ describes the deviation of the
177: quantity as a function of the truncation energy from its 'true' value (i.e. the value where
178: the cutoff in energy is taken to infinity).  The function $g(\alpha_Q)$ can be determined exactly using
179: high energy perturbation theory, well-controlled
180: provided $H_{\rm pert}$ is relevant.  For example, if $Q$ is some energy 
181: eigenvalue, then $g(\alpha_Q)=\alpha_Q =1$.  
182: 
183: 
184: The virtue of Eqn.(\ref{eii}) is that it allows us to run the RG in two steps.  We first implement the NRG
185: as described above until we reach a truncation energy placing us safely in the one-loop regime.
186: We then continue the RG by merely integrating the above equation allowing us to fully eliminate the
187: effects of truncation.
188: 
189: %in the case when $Q$ is an energy eigenvalue for then $\alpha=1$ uniformly
190: %-- we never need worry about a non-trivial wavefunction renormalization of $Q$
191: %altering $\alpha$ under the flow.  In the case of matrix elements of operators ${\cal O}$, 
192: %$\alpha$, will, in general, evolve under the flow.  But in a number
193: %of instances, $\alpha$ will not flow: i) ${\cal O}$ is an element of the stress-energy
194: %tensor or ii) $H_{\rm pert} + H_0$ is a gapped theory with $H_{\rm pert}$ a relevant perturbation
195: %of a CFT $H_0$.  In the latter instance the anomalous dimensions of operators do not flow and are
196: %all equivalent to their unperturbed counterparts \cite{zamo1}.  Outside of such cases, $\alpha$,
197: %while not known {\it a priori} can instead be directly determined from fitting Eqn.(\ref{eii}) to the NRG results.
198: 
199: We now consider two examples using this RG procedure, one where we compute the spectrum of a model
200: and one where we analyze correlation functions.  Both examples are chosen so that a straight
201: application of the TSA leads to poor results.
202: 
203: \noindent {\bf Spectrum:} In the first example, we consider a pair of quantum critical Ising chains
204: coupled together:
205: \begin{equation}\label{eiii}
206: H_0 = H^{Ising~1}_0 + H^{Ising~2}_0; ~~~ H_{\rm pert}=-\lambda\int dx \sigma^z_{1}\sigma^z_{2}.
207: \end{equation}
208: This model is known to be integrable and to have a spectrum equivalent to the sine-Gordon
209: model at $\beta^2=\pi$ \cite{llm}, that is, a spectrum with a pair of solitons with gap, $\Delta_{s1}=\Delta_{s2}$, together with
210: a set of six bound states with gaps, $\Delta_k = 2\Delta_s\sin(\pi k /14)$, $k=1,\cdots,6$.
211: By comparing conformal perturbation theory with a thermodynamic Bethe ansatz analysis,
212: $\Delta_{s1,2}$ can be expressed in terms of the coupling constant, $\lambda$:
213: $\Delta_{s1,2} = 11.2205920 \tilde\lambda$ with $\tilde\lambda = \lambda^{4/7}/(2\pi)^{3/7}$ \cite{fateev}.
214: 
215: The underlying finite volume Hilbert space of $H_0$ is considerably more complicated than that of single
216: Ising chain.  In a single chain there are four potential sectors of the Hilbert space \cite{zamo}: 
217: a sector, $I$, composed of even numbers of half-integer
218: fermionic modes acting over a unique vacuum, $|I\rangle$; a sector, $F$, composed
219: of odd numbers of half-integer modes over $|I\rangle$; and finally two sectors, $\sigma/\mu$,
220: composed of even numbers of both right and left moving integer fermionic modes over degenerate 
221: vacuua, $|\sigma\rangle/|\mu\rangle$.  The sectors $\sigma$ and $\mu$ are connected
222: by applying a product of an odd number of even mode fermionic operators.
223: Under periodic boundary conditions, however, the Hilbert
224: space of a single chain is reduced to two sectors, $I$ and $\sigma$.  In the two chain case, this is no longer true.
225: Not only do we have sectors of the form $I\otimes I$, $I\otimes \sigma$, $\sigma \otimes I$,
226: and $\sigma\otimes\sigma$ (such tensor products arising naturally from considering
227: two chains), but $F\otimes F$, $F\otimes\mu$, $\mu\otimes F$, and $\mu\otimes\mu$.  Unlike
228: a single chain, all possible sectors
229: are consistent with periodic boundary conditions.
230: The Hilbert space that results is thus much larger and applying the TSA with a simple truncation scheme
231: leads to poor results for the spectrum.  Computing the spectrum of this model is thus an ideal testing ground for our proposed
232: RG procedure.
233: 
234: \begin{table}
235: \vskip .2in
236: \begin{center}
237: \begin{tabular}{|l|l|l|l|l|}
238: \hline
239: exc. & Exact & TSA (10) & NRG & RG Improved \\
240: \hline
241: $\Delta_{s1}$ & 11.2206 & 11.92/12.67 & 11.32/11.54  & 11.17(2)/11.15(5)\\
242: \hline
243: $\Delta_{s2}$ & 11.2206 & 11.92/12.66 & 11.32/11.54 & 11.17(2)/11.16(6)\\
244: \hline
245: $\Delta_1$ & 4.9936 & 5.29/5.61 & 5.03/5.12 & 4.97(1)/4.97(2)\\
246: \hline
247: $\Delta_2$ & 9.7369 & 10.69/11.55 & 9.89/10.24 & 9.70(3)/9.7(1)\\
248: \hline
249: $\Delta_3$ & 13.9918 & 15.58/16.65 & 14.33/14.84 & 14.02(5)/14.20(5)\\
250: \hline
251: $\Delta_4$ & 17.5452 & 19.672/20.923 & 18.69/18.03 & 17.6(1)/17.7(1)\\
252: \hline
253: $\Delta_5$ & 20.2188 & 23.64/24.64 & 20.80/21.62 & 20.2(2)/20.5(2)\\
254: \hline
255: $\Delta_6$ & 21.8785 & 23.65/25.28 & 22.39/23.08 & 21.8(1)/21.8(2)\\
256: \hline
257: \end{tabular}
258: \caption{The excitation energies for two coupled Ising chains at values of $R=4/5\tilde\lambda^{-1}$.
259: (in units of $\tilde\lambda=\lambda^{4/7}/(2\pi)^{3/7}$).}
260: \end{center}
261: \vskip -.1in
262: \end{table}
263: 
264: In Figure 3 we outline the procedure by which we extract the values of the spectrum of the two
265: coupled Ising chains focusing for specificity on the second bound state, $\Delta_2$,
266: in the spectrum.  We first show the results of a straight TSA analysis (red squares) as a
267: function of increasing truncation energies (given in units of $1/R$).  While the gap, $\Delta_2$, is
268: converging towards its infinite volume value, $9.7368648\ldots\tilde\lambda$, it is doing so only slowly and
269: at exponentially increasing numerical cost.  We have performed the straight TSA analysis
270: up to level 15 (i.e. keeping states with energies less than $15/R$) where the Hilbert space contains $\approx 3500$
271: states.  At this point, the TSA produces a result deviating by $\approx 20\%$ from the exact value.
272: We also plot the value of $\Delta_2$ as 
273: given by the NRG algorithm as it iterates through states of ever higher energy.
274: Here we have run the algorithm so as to take into account states up to level 25 (in total $\approx 150000$/sector).
275: We see, reassuringly, that where the TSA results exists, the NRG algorithm produces matching
276: results (at a fraction of the numerical cost).  The NRG algorithm ends up producing a value of $\Delta_2$
277: with a $5\%$ error.
278: Finally we plot 
279: the result of fitting Eqn.(\ref{eii}) to the NRG results between level 20 and 25 (where we believe
280: a one-loop RG equation describes the NRG flow).  Extrapolating the fit to $E_{\rm trunc}=\infty$
281: gives  $\Delta_2=9.70\tilde\lambda$, a value deviating from the exact result by $0.5\%$.
282: 
283: In Table 1 we present the results of our RG analysis on the complete spectrum of the two coupled chains.  
284: In the first column we provide the exact value of spectrum as determined by integrability
285: and the TBA analysis of Ref. \cite{fateev}.  In the second column we give the values of the spectrum
286: at two different system sizes ($R=4/5\tilde\lambda^{-1}$)
287: computed using a straight TSA analysis truncating at level 10 (i.e. keeping $\approx 600$ states
288: in each of the relevant sectors).  
289: We see that the disagreement  with the exact result ranges up to 20\%.
290: In the third column we give the results coming from applying the NRG algorithm (again iterating
291: until we have reached level 25).
292: We see a marked improvement over the TSA analysis, but still we obtain results with errors 
293: ranging up to 5\%. In the final column, we give the results for the spectrum arrived
294: at by fitting the one-loop RG equation (Eqn.(\ref{eii})) to the NRG data.  We see that our errors are
295: now less than 1\% .
296: \vskip .1in
297: \begin{figure}[tbh]
298: \includegraphics[height=2.4in,width=3.in,angle=0]{me.eps}
299: \caption{A plot of the matrix element, $\langle 0| \sigma(0) |A_1(p)A_1(-p)\rangle$
300: as a function of energy, $\omega = 2(p^2+m_1^2)^{1/2}$.}
301: \end{figure}
302: 
303: \noindent {\bf Correlation Functions:}
304: We now turn to the computation of correlation functions using the above described RG methodology.  For simplicity
305: we consider only $T=0$ response functions although a generalization to finite temperature multi-point 
306: functions is readily realizable.
307: At zero temperature, the imaginary piece of a retarded correlation function, $G_{\rm ret}(x,t) = 
308: \langle {\cal O}(x,t){\cal O}(0)\rangle_{\rm ret}$,
309: has a spectral decomposition, $S_{\cal OO}(x,\omega) = -\frac{1}{\pi}{\rm Im}G_{\rm ret}(x,\omega>0)$ equal to \cite{review}
310: \begin{eqnarray}\label{eiv}
311: S_{\cal OO}(x,\omega) = \sum_{n,s_n}e^{iP_{s_n}x}|\langle 0|{\cal O}(0,0)|n;s_n\rangle|^2
312: \delta(\omega-E_{s_n}),
313: \end{eqnarray}
314: where $|n;s_n\rangle$ is an eigenstate of the system with energy/momentum $E_{s_n}/P_{s_n}$ 
315: built out of $n$ fundamental single
316: particle excitations carrying internal quantum numbers $s_n$.
317: Thus the computation of any response function is equivalent to the computation of a number
318: of matrix elements of the form $\langle 0|{\cal O}(x,0)|n;s_n\rangle$.  Ostensibly 
319: to compute the response function fully one would need to compute an infinite number of such
320: matrix elements.  In practice if one is interested in the response function at low energies
321: only a small finite number of such matrix elements need be computed \cite{review}.
322: We will illustrate the computation under the RG of one such non-trivial matrix element
323: for a critical Ising chain in a magnetic field (i.e. Eqn.(\ref{ei})). 
324: While the spectrum of this theory rapidly converges upon the increase of $E_{\rm trunc}$,
325: the matrix elements are less well behaved.  And unlike the two-chain case, analytical
326: results are available for comparison \cite{muss}. Thus this computation is a good test
327: of our RG methodology.
328: 
329: We specifically compute the two excitation contribution 
330: $f_2(\omega) = \langle0|\sigma(0)|A_1(p)A_1(-p)|0\rangle$ with
331: $p = (\omega^2-\Delta_{1I}^2)$, to
332: the spin-spin correlator, $S_{\sigma\sigma}(p=0,\omega>0)$.  (Here $\Delta_{1I}$ is
333: the gap of the lowest lying single particle excitation, $A_1$, in an Ising chain in a magnetic field.)
334: This contribution takes
335: the form
336: \begin{equation}
337: \delta S_{\sigma\sigma}(\omega) = \pi^{-1}\omega(\omega^2-4\Delta_{1I}^2)^{1/2}\Theta(\omega-2\Delta_{1I})
338: |f_2(\omega)|^2.
339: \end{equation}
340: We are able to compute the necessary {\it infinite} volume matrix element over a {\it continuous} range of energies
341: by studying a single matrix element in {\it finite} volume where the spectrum is {\it discrete}.  We do so by
342: continuously varying the system size, $R$.  Under such variations, the energy $\omega = 2(p^2+m_1^2)^{1/2}$ 
343: of the state, $|A_1(p)A_1(-p)\rangle$, changes continuously due to the quantization condition of the momentum, $p$,
344: (i.e. $p = 2\pi n /R + \delta(p,-p)$ where $\delta(p,-p)$ is a two-body scattering phase).  
345: 
346: In Figure 4 by varying $R$ we parametrically plot the results of our computations of $f_2(p)$ vs its exact value \cite{muss}.
347: A straight application of the TSA (with a level
348: 10 truncation) produces acceptable results at higher energies but does poorly at energies around threshold, $2\Delta_1$.
349: At larger values of $R$ (and so smaller energies), the TSA breaks down.  The TSA curve
350: in this region is then double valued.  Computing the same matrix
351: element with the NRG algorithm leads to a considerable improvement but at the lowest energies
352: a deviation from the exact result remains (see inset to Fig. 4).  
353: RG improving the computation of $f_2(p)$ largely removes this discrepancy even at energies next
354: to threshold.  (In applying Eqn. (2) to $Q=f_2(p)$, perturbation theory yields, $g(\alpha_{f_2})= 2(1-1/8)$ where $1/8$
355: is the anomalous dimension of the spin operator, $\sigma$).
356: 
357: In conclusion we have presented an RG scheme by which a large number of one-dimensional
358: continuum models can be studied with quantitative accuracy.  With this methodology, both
359: the spectrum and spectral functions of a model can be determined.
360: 
361: RMK and YA acknowledge support from the US DOE
362: (DE-AC02-98 CH 10886) together with useful discussions with A. Tsvelik.
363: 
364: \begin{references}
365: 
366: \bibitem{wilson} K. Wilson, Rev. Mod. Phys. {\bf 47}, 773 (1975).
367: \bibitem{white} S.R. White, Phys. Rev. Lett. {\bf 69}, 2863 (1992).
368: \bibitem{white1} T.D. K\"uhner and S.R. White, Phys. Rev. B {\bf 60}, 335 (1999).
369: \bibitem{note} Further studies need to be done on the efficacy of the method for marginal perturbations.
370: \bibitem{zamo} V. P. Yurov and Al. B. Zamolodchikov, Int. J. Mod. Phys A {\bf 6}, 4557 (1991).
371: \bibitem{efetov} K. Efetov, Adv. Phys. {\bf 32}, 53 (1983).
372: \bibitem{fateev} V. A. Fateev, Phys. Lett. B {\bf 324} 45 (1994).
373: \bibitem{zamo1} A. B. Zamolodchikov, Adv. Studies in Pure Math. {\bf 19}, 641 (1989).
374: \bibitem{muss} G. Delfino and G. Mussardo, Nucl. Phys. B {\bf 455}, 724 (1995).
375: \bibitem{llm} A. Leclair, A. Ludwig, and G. Mussardo, Nucl. Phys. B {\bf 512}, 523 (1998).
376: \bibitem{ffprog} F. Smirnov, ``Form Factors in Completely Integrable Models of Quantum Field Theory'',
377: World Scientific, Singapore (1992).
378: \bibitem{review} F. Essler and R. M. Konik in ``From Fields to Strings: Circumnavigating
379: Theoretical Physics'', ed. by M. Shifman, A. Vainshtein, and J. Wheater, World Scientific,
380: Singapore (2005).
381: \end{references}
382: 
383: 
384: \end{document}
385: