cond-mat0411405/bt.tex
1: %\documentstyle[aps,preprint,twocolumn]{revtex}
2: %\documentclass[aps,preprint]{revtex}
3: \documentclass[aps,draft]{revtex4}
4: 
5: \input epsf
6: %\usepackage[active]{srcltx}
7: 
8: %\tightenlines
9: \begin{document}
10: 
11: \title{
12: Fermionic Quasiparticle Representation of Tomonaga-Luttinger Hamiltonian
13: }
14: \author{A.V. Rozhkov}
15: 
16: \affiliation{
17: Institute for Theoretical and Applied Electrodynamics JIHT RAS,
18: Moscow, ul. Izhorskaya 13/19, 127412, Russian Federation
19: }
20: %\maketitle
21: 
22: \begin{abstract}
23: We find a unitary operator which asymptotically diagonalizes the
24: Tomonaga-Luttinger hamiltonian of one-dimensional spinless electrons. The
25: operator performs a Bogoliubov
26: rotation in the space of electron-hole pairs. If bare interaction of the
27: physical electrons is sufficiently small this transformation
28: maps the original Tomonaga-Luttinger system on a system of free fermionic
29: quasiparticles. Our representation is useful when the electron dispersion
30: deviates from linear form. For such situation we obtain non-perturbative
31: results for the electron gas free energy and the density-density
32: propagator.
33: \end{abstract}
34: 
35: \maketitle
36: \hfill
37: %\draft
38: 
39: 
40: 
41: \section{Introduction}
42: Bosonization is a standard approach to the problem of interacting 
43: electrons in one dimension (1D) \cite{haldane,boson}. Bosonization maps the
44: low-energy spectrum of the Tomonaga-Luttinger (TL) model on the spectrum of
45: free bosons.
46: 
47: In this paper we discuss a new method of treating such system.
48: We explicitly construct a unitary operator $U$ diagonalizing TL
49: hamiltonian.
50: Our unitary transformation maps the TL model on a system of free fermionic
51: quasiparticles. The description of the TL model in terms of these
52: quasiparticles
53: has several advantages over bosonization, as we will see below.
54: 
55: The diagonalizing procedure is closely related to the bosonization. The
56: operator $U$ can be thought of as a Bogoliubov transformation in the space
57: of particle-hole pair excitations. Alternatively, one can describe the action
58: of $U$ as a sequence of bosonization, Bogoliubov transformation which
59: diagonalizes the bosonic hamiltonian and further refermionization (fig.1).
60: 
61: When the dispersion of the physical electrons is linear both the 
62: quasiparticles
63: and the bosons offer equally good description of the TL model. If the
64: non-linear terms are substantial the free boson representation breaks down.
65: The fermionic quasiparticles show more resilience toward deviations from
66: the linear dispersion. They remain free as long as the bare interaction
67: constant is sufficiently small.
68: 
69: This latter property of the quasiparticle representation allows for
70: non-perturbative calculation of the free energy and the density-density
71: correlation function for the TL model with the non-linear dispersion. We
72: believe that these two results are new.
73: 
74: When the bare electron dispersion is linear the quasiparticle
75: representation could be used to determine the single-electron Green's
76: function. It coincides with the Green's function
77: obtained by other methods.
78: 
79: The existence of the quasiparticles does not contradict to the fact that
80: the
81: single-electron Green's function of TL model has no pole. The
82: quasiparticles
83: in TL model has zero overlap with the physical electrons: $\sqrt{Z} = 0$.
84: Therefore, it is convenient to think about TL model as $Z=0$ Fermi liquid
85: \cite{carmelo}.
86: 
87: The paper is organized as follows. We diagonalize the TL hamiltonian in
88: Sec. II. In Sec. III we offer an intuitive explanation to the method. In
89: this section all
90: technical complications are disregarded in order to create an integral
91: view. The density-density propagator is derived in Sec. IV. The
92: single-particle Green's function is calculated in Sec. V. Sec. VI is
93: reserved for discussion. Certain technical details can be found in Appendices.
94: 
95: 
96: 
97: 
98: \section{Diagonalization of TL hamiltonian}
99: 
100: The TL model hamiltonian is given by:
101: \begin{eqnarray}
102: H = H_{\rm kin} + H_{\rm int},\label{H}\\
103: H_{\rm kin} = i v_{\rm F}\int_{-L/2}^{L/2} dx 
104: \left( \colon\psi^\dagger_{{\rm L}}
105: \nabla\psi^{\vphantom{\dagger}}_{{\rm L}}\colon - \colon\psi^\dagger_{{\rm R}}
106: \nabla\psi^{\vphantom{\dagger}}_{{\rm R}}\colon \right) ,\label{H_kin}\\
107: H_{\rm int} = \int dx dx' \hat g(x-x') \rho_{\rm L}(x) \rho_{\rm R}(x'),
108: \label{int}\\
109: \rho_{\rm L,R} = \colon\psi_{\rm L,R}^\dagger \psi_{\rm
110: L,R}^{\vphantom{\dagger}}\colon.
111: \end{eqnarray}
112: The chirality label `L' corresponds to left-moving electrons, the label `R'
113: corresponds to right-moving electrons. The interaction of the electrons of
114: the same chirality is ignored since up to irrelevant operators such
115: interaction simply renormalizes the value of the Fermi velocity $v_{\rm F}$.
116: The symbol $:\ldots :$ denotes normal ordering of the field operators
117: $\psi$. A brief discussion of the normal ordering procedure we use in this
118: paper is given in Appendix A.
119: 
120: It is assumed that the cut-off of (\ref{H}) is infinite. To remove
121: ultraviolet divergences of the theory without cut-off we
122: replace usual zero-range interaction ($\hat g(x)=g_0\delta(x)$) by 
123: interaction acting over a finite range. Specifically, $\hat g(x)=g_0
124: \delta_\Lambda(x)$ where $\delta_\Lambda (x)$ is a broadened
125: version of the delta-function: its Fourier transform $\delta_\Lambda(q)$ is
126: such that $\delta_\Lambda(q)=1$ for $|q|$ smaller than some quantity
127: $\Lambda$, and $\delta_\Lambda (q)$ vanishes quickly for $|q|>\Lambda$.
128: The parameter $\Lambda$ thus defined regularizes ultraviolet divergences of
129: our theory.
130: 
131: It is easy to demonstrate (see, for example, \cite{haldane}) that the
132: following commutation relations are obeyed:
133: \begin{eqnarray}
134: \left[ \rho_{pq}, \rho_{p'-q'} \right] = \delta_{pp'} \delta_{qq'} p n_q
135: \label{rho_comm}\\
136: \left[\rho_{pq}^{\vphantom{\dagger}}, \psi^\dagger_{p'}(x) \right] =
137: \delta_{pp'} {\rm e}^{-iqx} \psi^\dagger_{p'}(x), \label{psi_comm}
138: \end{eqnarray}
139: where $p=+1$ for left-moving electrons, $p=-1$ for right-moving electrons
140: and
141: \begin{equation}
142: \rho_{pq} = \int dx {\rm e}^{-iqx} \rho_p (x),\quad n_q = \frac{Lq}{2\pi}.
143: \end{equation}
144: Using this commutation relations we will show that the unitary operator
145: \begin{eqnarray}
146: U={\rm e}^{\Omega},\\
147: {\Omega} = \sum_{q \ne 0 } \sum_p \alpha_q
148: \frac{p}{n_q} \rho_{p-q} \rho_{-pq}\label{Omega}
149: \end{eqnarray}
150: diagonalizes the hamiltonian (\ref{H}) except for the zero mode part.
151: Fortunately, the zero modes 
152: \begin{eqnarray}
153: N_{p} = \rho_{{p}q} \Big|_{q=0}
154: \end{eqnarray}
155: are decoupled from other degrees of freedom.
156: Also, their contribution to the low-energy spectrum is ${\cal O} (1/L)$.
157: 
158: Since $\rho$'s are quadratic in $\psi$
159: the above operator $U$ is quartic in fermionic operators $\psi$. In
160: general, it is impossible to work with such a non-linear
161: object. In our situation, however, the simplicity of commutation rules
162: (\ref{rho_comm})  and (\ref{psi_comm}) allows us to diagonalize the
163: hamiltonian with the help of $U$.
164: 
165: In order to transform the interaction term (\ref{int}) with the operator
166: $U$ it is enough to observe that the action of $U$ on the density operator
167: $\rho_{pq}$, $q \ne 0$, is a Bogoliubov rotation:
168: \begin{eqnarray}
169: U \rho_{pq} U^\dagger = u(\alpha_q) \rho_{pq} + v(\alpha_q) \rho_{-pq},
170: \label{BT}\\
171: u(\alpha_q) = u_q = \cosh(\alpha_q),\ v(\alpha_q) = v_q =
172: \sinh(\alpha_q),\\
173: u_q^2 - v_q^2 = 1.
174: \end{eqnarray}
175: This result is a simple consequence of the commutation relation
176: (\ref{rho_comm}). This identity can be derived by a variety of methods. For
177: example, one can differentiate its left-hand side with respect to $\alpha_q$ for
178: both $p={\rm L}$ and $p={\rm R}$ and solve the resultant differential
179: equation system.
180: 
181: The easiest way to transform the kinetic energy term is to
182: notice that the kinetic energy density can be expressed as a product of two
183: density operators. The derivation goes as follows. First, we normal order
184: the product of two density operators:
185: \begin{eqnarray}
186: \rho_p (x) \rho_p (y)&=&
187: \colon \psi_p^\dagger(x) \psi_p^{\vphantom{\dagger}}(x)
188: \psi_p^\dagger(y) \psi_p^{\vphantom{\dagger}}(y)\colon + 
189: s_p(x-y)\colon\psi_p^{\vphantom{\dagger}}(x) \psi_p^\dagger(y)\colon +
190: \label{rr}\\
191: &&s_p(x-y) \colon\psi_p^{\dagger}(x) \psi_p^{\vphantom{\dagger}}(y)\colon + 
192: b_p(x-y),
193: \nonumber \\
194: s_p(x)&=&\frac{p}{2\pi i\left(x-ip0\right)},\\
195: b_p(x)& = &\left(s_p(x)\right)^2,
196: \end{eqnarray}
197: As it is explained in Appendix A the normal ordering is used here to isolate
198: explicitly singular terms of the field operator products. Now we expand the
199: above identity into Laurent series with respect to powers of $(x-y)$:
200: \begin{eqnarray}
201: \rho_p (x) \rho_p (y)&=&b_p(x-y)
202: +\frac{ip}{2\pi} \colon\psi_p^{\vphantom{\dagger}}(x)
203: \nabla\psi_p^\dagger(x)\colon +
204: \frac{ip}{2\pi} \colon\psi_p^{\dagger}(x)
205: \nabla \psi_p^{\vphantom{\dagger}}(x)\colon 
206: + (\text{irrelevant\ operators}). \label{Laurent}
207: \end{eqnarray}
208: In this expansion an irrelevant
209: operator can be recognized by a factor of $(x-y)^n$ where $n>0$. For
210: example, the first term in (\ref{rr}) is an irrelevant operator:
211: \begin{equation}
212: \colon\psi_p^\dagger(x) \psi_p^{\vphantom{\dagger}}(x)
213: \psi_p^\dagger(y) \psi_p^{\vphantom{\dagger}}(y)\colon \ \approx
214: \left(x-y\right)^2\colon\psi_p^\dagger(x) \psi_p^{\vphantom{\dagger}}(x)
215: \nabla \psi_p^\dagger(x) \nabla\psi_p^{\vphantom{\dagger}}(x)\colon.
216: \label{irrelevant}
217: \end{equation}
218: Indeed, by simple power counting one can verify that its scaling dimension is
219: equal to $d=4<2$. Sending $y \rightarrow x$ in eq. (\ref{Laurent}), we
220: establish the identity:
221: \begin{equation}
222: \frac{ip}{2\pi} \left( \colon\psi_p^\dagger(x)
223: (\nabla\psi_p^{\vphantom{\dagger}}(x))\colon  -
224: \colon(\nabla\psi_p^\dagger(x)) \psi_p^{\vphantom{\dagger}}(x)\colon\right)
225: =\lim_{y\rightarrow x} \left\{ \rho_p(x) \rho_p(y) -
226: b_p(x-y)\right\}.
227: \label{rhorho}
228: \end{equation}
229: Thus, it is permissible to write the hamiltonian (\ref{H}) in the form:
230: \begin{eqnarray}
231: H= \frac{\pi v_{\rm F}}{L} \sum_{pq} \left(\rho_{pq} \rho_{p-q} + 
232: \frac{g_q}{2\pi v_{\rm F}} \rho_{pq} \rho_{-p-q} \right), \label{Hr}
233: \end{eqnarray}
234: where $g_q$ is the Fourier transform of $\hat g(x)$. 
235: For convenience we explicitly show the zero mode part of the
236: above expression:
237: \begin{eqnarray}
238: H= \frac{\pi v_{\rm F}}{L} \sum_{q \ne 0} \sum_p \left(\rho_{pq} \rho_{p-q} + 
239: \frac{g_q}{2\pi v_{\rm F}} \rho_{pq} \rho_{-p-q} \right) +
240: \frac{ \pi v_{\rm F}}{L} \left( N_{\rm L}^2 + N_{\rm R}^2 \right) +
241: \frac{g_0}{L} N_{\rm L} N_{\rm R}.
242: \end{eqnarray}
243: Such splitting is useful since our transformation $U$ does not act on the
244: zero modes.
245: We now apply $U$ and choose parameters $\alpha_q$ in such a way that the
246: term $\rho_{\rm L} \rho_{\rm R}$ vanishes:
247: \begin{eqnarray} 
248: \tanh 2\alpha_q = -\frac{g_q}{2\pi v_{\rm F}},\\
249: u^2_q = \frac{1}{2} \left( 1 + \frac{1}{\sqrt{ 1 -
250: \left({g_q}/{2\pi v_{\rm F}}\right)^2}} \right),\quad
251: v^2_q = \frac{1}{2} \left( - 1 + \frac{1}{\sqrt{ 1 -
252: \left({g_q}/{2\pi v_{\rm F}}\right)^2}} \right), \quad v_q g_q < 0.
253: \label{uv}
254: \end{eqnarray}
255: The transformed hamiltonian is:
256: \begin{eqnarray}
257: U H U^\dagger&=&\frac{\pi v_{\rm F}}{L} \sum_{q \ne 0} \sum_p
258: \left(u_q^2 + v_q^2 +
259: \frac{g_q}{\pi v_{\rm F}} u_q v_q \right) \rho_{pq} \rho_{p-q} + 
260: \frac{ \pi v_{\rm F}}{L} \left( N_{\rm L}^2 + N_{\rm R}^2 \right) +
261: \frac{g_0}{L} N_{\rm L} N_{\rm R} \\
262: &=&\frac{\pi v_{\rm F}}{L} \sum_{pq} 
263: \left(u_q^2 + v_q^2 +
264: \frac{g_q}{\pi v_{\rm F}} u_q v_q \right) \rho_{pq} \rho_{p-q} + 
265: \frac{ \pi \left(v_{\rm F} - \tilde v_{\rm F} \right)}{L}
266: \left( N_{\rm L}^2 + N_{\rm R}^2 \right) +
267: \frac{g_0}{L} N_{\rm L} N_{\rm R}, \\
268: \tilde v_{\rm F}&=& v_{\rm F}
269: \sqrt{1-\left(\frac{g_0}{2\pi v_{\rm F}}\right)^2}.\label{tilde_v}
270: \end{eqnarray}
271: From now on we will ignore the zero mode contribution to the hamiltonian:
272: the above formulas clearly show that it is small ($\sim 1/L$) for a
273: macroscopic system. Ultimately, we have:
274: \begin{eqnarray}
275: U H U^\dagger&=&\frac{\pi v_{\rm F}}{L} \sum_{pq } 
276: \left(u_q^2 + v_q^2 +
277: \frac{g_q}{\pi v_{\rm F}} u_q v_q \right) \rho_{pq} \rho_{p-q}.
278: \label{Hfin}
279: \end{eqnarray}
280: This expression can be re-written in terms of fields $\psi$. Let us show how
281: this is done in a technically trivial case of
282: $\Lambda \rightarrow \infty$ (which is equivalent to $\hat g (x) =
283: g_0 \delta(x)$). When $\Lambda$ is infinite the functions $u_q$, $v_q$ and
284: $g_q$ are all constants independent of $q$. Their values are given by
285: (\ref{uv}) with $g_q \equiv g_0$. Therefore, hamiltonian (\ref{Hfin}) is
286: equal to
287: \begin{eqnarray}
288: U H U^\dagger&=&\frac{\pi \tilde v_{\rm F}}{L} \sum_{pq } 
289: \rho_{pq} \rho_{p-q},
290: \end{eqnarray}
291: where $\tilde v_{\rm F}$ is given by (\ref{tilde_v}). Finally, inverting
292: (\ref{rhorho}), we obtain
293: \begin{eqnarray}
294: U H U^\dagger = i \tilde v_{\rm F} \int dx \sum_p p \colon 
295: \psi_p^\dagger \nabla \psi_p^{\vphantom{\dagger}}\colon + {\rm const}.
296: \end{eqnarray}
297: These transformations, apart from minor differences, are equivalent to
298: the refermionization as it is done in \cite{vD-S}.
299: 
300: In certain situations, however, it is convenient to have finite $\Lambda$.
301: Then a subtler reasoning is required. It is possible to prove that:
302: \begin{eqnarray}
303: U H U^\dagger = i\tilde v_{\rm F} \int dx \sum_p p \colon \psi_p^\dagger
304: \nabla\psi_p^{\vphantom{\dagger}}\colon + ({\rm irrelevant\ operators}).
305: \label{Hqp}
306: \end{eqnarray}
307: Such proof is rather technical. It can be found in Appendix B.
308: 
309: Let us agree that operators with the tilde are transformations of
310: corresponding operators without the tilde, that is $\tilde O=
311: U^\dagger O U$. Then the hamiltonian can be expressed in terms of operators
312: $\tilde\psi$ as follows:
313: \begin{eqnarray}
314: H = i\tilde v_{\rm F} \int dx \sum_p p \colon \tilde \psi_p^\dagger
315: \nabla\tilde\psi_p^{\vphantom{\dagger}}\colon + ({\rm irrelevant\ operators}).
316: \label{H0}
317: \end{eqnarray}
318: The above equation is the desired mapping of TL model on the model of 
319: quasiparticles $\tilde\psi$ whose interactions are small irrelevant
320: operators. Due to the irrelevance of the interaction the low-energy
321: properties of $\tilde\psi$ are those of free fermions.
322: 
323: Using the developed framework it is possible to discuss the effect of
324: non-linear dispersion on the spectrum and correlations of (\ref{H}). We add
325: an extra term to the TL hamiltonian:
326: \begin{eqnarray}
327: H' = H + H_{\rm nl}, \\ 
328: H_{\rm nl} = \sum_p \int dx dx' \hat h(x-x') \nabla\psi^\dagger_p(x)
329: \nabla \psi_p^{\vphantom{\dagger}} (x'), \label{H2}
330: \end{eqnarray}
331: The subscript `nl' stands for `non-linear' dispersion. Function $h_q$
332: which is the Fourier transform of $\hat h(x)$ has these properties:
333: \begin{eqnarray}
334: h_{q} = v'_{\rm F} < v_{\rm F}/\Lambda\ {\rm for}\ |q|<\Lambda,
335: \label{q=0}\\
336: h_{q} < v_{\rm F}/|q| \ {\rm for}\ |q| > \Lambda.
337: \end{eqnarray}
338: The second of these two conditions guarantees that for
339: $v'_{\rm F} < v_{\rm F}/\Lambda$ the modified kinetic energy
340: $pv_{\rm F}q + h_q q^2$ has the same sign as the original kinetic energy 
341: $pv_{\rm F}q$. That is, $H_{\rm nl}$ does not induce an instability of the
342: ground state by creating spurious Fermi point.
343: 
344: Condition (\ref{q=0}) implies that for small $|q|<\Lambda$ it is permissible
345: to use
346: \begin{eqnarray}
347: H_{\rm nl} = v'_{\rm F} \sum_p \int dx \nabla \psi^\dagger_p (x) \nabla
348: \psi^{\vphantom{\dagger}}_p(x)
349: \end{eqnarray}
350: instead of (\ref{H2}). The quantity $2v'_{{\rm F}}$ is the Fermi velocity
351: derivative with respect to the momentum.
352: 
353: We need to express $H_{\rm nl}$ in terms of $\tilde\psi$. To do so it is
354: convenient to rewrite $H_{\rm nl}$ with the help of density operators $\rho_p$
355: rather than field operators $\psi_p$. This trick was already used in
356: (\ref{rhorho}). It is easy to establish that
357: \begin{eqnarray}
358: v'_{\rm F} \sum_p \colon (\nabla \psi^\dagger_p (x))
359: (\nabla \psi^{\vphantom{\dagger}}_p (x)) \colon - \frac{1}{6} \nabla^2
360: \rho_p(x) = \label{Hnl_aux}\\
361: \frac{2\pi v'_{\rm F}}{3} \sum_p \lim_{y \rightarrow x} \left\{ i p \rho_p (x)
362: \left[ \colon \psi^\dagger_p (y) (\nabla \psi^{\vphantom{\dagger}}_p (y))
363: \colon -
364: \colon (\nabla \psi^\dagger_p (y)) \psi^{\vphantom{\dagger}}_p (y) \colon
365: \right] - 4\pi b_p(x-y) \rho_p (y) \right\}. \nonumber
366: \end{eqnarray}
367: According to (\ref{rhorho}) the expression in the square brackets
368: proportional to the product of two density operators. Thus:
369: \begin{eqnarray}
370: {\cal H}_{\rm nl} (x) = 
371: v'_{\rm F} \sum_p \colon (\nabla \psi^\dagger_p (x))
372: (\nabla \psi^{\vphantom{\dagger}}_p (x)) \colon - \frac{1}{6} \nabla^2
373: \rho_p(x) = \label{rho3}\\
374: \frac{4\pi^2v'_{\rm F}}{3}
375: \sum_p \lim_{y \rightarrow x} \lim_{z \rightarrow y} \left\{
376: \rho_p (x) \left[ \rho_p (z) \rho_p(y) - b_p (z-y) \right] -
377: 2b_p (x-y) \rho_p(y) \right\}. \nonumber
378: \end{eqnarray}
379: Now we have to substitute $\rho_{pq} = u \tilde \rho_{pq} + v \tilde
380: \rho_{-pq}$ for $q \ne 0$ in
381: the above expression. The resultant third order polynomial of $\tilde \rho$'s
382: must be rewritten in terms of $\tilde \psi$ with the help of (\ref{rr})
383: and (\ref{rho3}). The final expression for $H_{\rm nl}$ is:
384: \begin{eqnarray}
385: H_{\rm nl} = \sum_p \int dx \left\{
386: \tilde v'_{\rm F}\colon (\nabla \tilde \psi^\dagger_p)
387: (\nabla \tilde \psi^{\vphantom{\dagger}}_p ) \colon +
388: i p \tilde g' \tilde \rho_{-p} \left( \colon \tilde \psi^\dagger_p 
389: (\nabla \tilde \psi^{\vphantom{\dagger}}_p) \colon -
390: \colon (\nabla \tilde \psi^\dagger_p) \tilde \psi^{\vphantom{\dagger}}_p
391: \colon \right) + \tilde \mu \tilde \rho_p\right\}. \label{nlq}
392: \end{eqnarray}
393: The details of the derivation together with the exact formulas for
394: coefficients $\tilde v'_{\rm F}$, $\tilde g'$ and $\tilde \mu$ can be found
395: in Appendix C. Here we quote only the expressions which are valid if the
396: interaction parameter $g_0$ is small. Let us define $\alpha_0$ as
397: $\alpha_0 = \alpha_{q=0}$. When interaction is small
398: ($\alpha_0 \approx g_0/4\pi v_{\rm F} \ll 1$) the coefficients are:
399: \begin{eqnarray}
400: \tilde v'_{\rm F} - v'_{\rm F} \approx v'_{\rm F} \alpha^2_0,\\
401: \tilde g' \approx 2\pi v'_{\rm F}\alpha_0
402:  = \frac{g_0 v'_{\rm F}}{2 v_{\rm F}},\\
403: \tilde \mu \approx \gamma \Lambda^2 v'_{\rm F} \alpha_0^2,
404: \end{eqnarray}
405: where $\gamma$ is a non-universal constant of order unity.
406: 
407: The total hamiltonian $H'$ is the sum of $H$, eq. (\ref{H0}), and
408: $H_{\rm nl}$.  Due to $H_{\rm nl}$ the quasiparticle dispersion becomes
409: non-linear (the first term of (\ref{nlq})). Also, $H_{\rm nl}$ introduces
410: additional interactions between the quasiparticles. The operators
411: corresponding to these interaction (the second term in (\ref{nlq})) are
412: ${\cal O}(v'_{\rm F} g_0)$ and irrelevant. Therefore, they can be
413: neglected provided that 
414: \begin{eqnarray} 
415: \tilde g'\Lambda / v_{\rm F} \ll 1 \Leftrightarrow
416: v'_{{\rm F}} g_0 \Lambda \ll v_{\rm F}^2. \label{free}
417: \end{eqnarray}
418: The quasiparticles remain free even if the bare
419: dispersion is not linear as long as the bare interaction is sufficiently
420: small. Note, that the free boson representation for TL model is much
421: less tolerant of $H_{\rm nl}$. This term is equivalent to cubic interaction
422: between bosons (see (\ref{rho3}) and Ref.\cite{haldane}). For the bosons to
423: remain free a stricter condition $v'_{{\rm F}} \Lambda \ll v_{\rm F}$ has
424: to be satisfied.
425: 
426: Equations (\ref{H0}) and (\ref{nlq}) allow us to describe the thermodynamics
427: of the electronic liquid with the generic dispersion. For example, the free
428: energy equals to
429: \begin{equation}
430: F = T\sum_{pk} \log \left( 1 + {\rm e}^{-\tilde \varepsilon_p(k)/T} \right)
431: =\frac{LT}{\pi} \int \frac{d \tilde \varepsilon}
432: {\sqrt{ \tilde v^2 _{\rm F} + 4 \tilde v'_{\rm F} \tilde \varepsilon}}
433: \log \left( 1 + {\rm e}^{-\tilde \varepsilon/T} \right), \label{free_en}
434: \end{equation}
435: where $\tilde \varepsilon_p(k) = p \tilde v _{\rm F} k + \tilde v'_{\rm F} 
436: k^2$ is the dispersion of the quasiparticles. This result is
437: non-perturbative in $v'_{\rm F} \Lambda /v _{\rm F}$. 
438: Corrections to the free energy due to the neglected quasiparticle
439: interaction are ${\cal O} ((v'_{\rm F} g_0)^2)$. Thus, our result for $F$
440: is accurate provided that (\ref{free}) holds true. 
441: 
442: The specific heat at constant chemical potential can be derived from
443: (\ref{free_en}) in the limit of low temperature
444: ($T \ll \tilde v_{\rm F}^2/ \tilde v'_{\rm F}$):
445: \begin{eqnarray}
446: C(T) = \frac{\pi}{3 \tilde v_{\rm F}} T + \frac{14 \pi^3
447: (\tilde v'_{\rm F})^2}{5 \tilde v_{\rm F}^5} T^3
448: + {\cal O} \left(({\tilde v'_{\rm F} T}/{\tilde v_{\rm F}^2})^5 \right)
449: +  {\cal O} (g_0^4) + {\cal O} ((\tilde g')^2). \label{C(T)}
450: \end{eqnarray}
451: We observe that non-zero dispersion curvature generates $T^3$
452: contribution to the specific heat. (Note: irrelevant operators neglected
453: along the way also contribute to the specific heat. However, their
454: contribution is ${\cal O}(g_0^4)$ and ${\cal O} ((\tilde g')^2)$.
455: For small $g_0$ such corrections can be disregarded.)
456: 
457: The same result for the specific heat could be obtained with the help of
458: bosonization technique. There one must expand the free energy in orders of
459: $H_{\rm nl}$. These calculations are done in Appendix D. They provide an
460: important consistency check for the proposed approach.
461: 
462: \section{Non-rigorous overview of the method}
463: 
464: We have finished the introduction of our method and ready to apply it to
465: Green's function calculations. At this point we would like to make a break
466: between two rather technical parts and explain why the method works on
467: intuitive level. 
468: 
469: The central
470: issue of this discussion could be loosely formulated as follows: why
471: the strongly interacting bosons could be mapped on the weakly interacting
472: fermions, what kind of `cancellation' of interactions takes place? To answer
473: this question we direct our attention to most essential aspects of the
474: approach. All technical complications will be disregarded: we will ignore
475: zero modes, normal ordering and assume that $\Lambda = \infty$. This makes
476: the following presentation more transparent.
477: 
478: Let us start with the bosonized form of the TL hamiltonian, familiar to
479: many researchers \cite{haldane}:
480: \begin{eqnarray}
481: H'&=& H_{\rm kin} + H_{\rm int} + H_{\rm nl}, \label{Hbos}\\
482: H_{\rm kin} + H_{\rm int}&=& \int dx \frac{\tilde v_{\rm F}}{2} \left\{ 
483: {\cal K} \left( \nabla \Theta \right)^2 +
484: {\cal K}^{-1} \left( \nabla \Phi \right)^2 \right\}, \\
485: H_{\rm nl} &=& \int dx
486: \frac{2 \pi^2 v'_{\rm F}}{3\sqrt{2}}
487: \left\{ \left( \nabla \Theta + \nabla \Phi \right)^3
488: + \left( \nabla \Theta - \nabla \Phi \right)^3 \right\}, \label{Hnl_bos}
489: \end{eqnarray}
490: where boson fields $\Theta$ and $\Phi$ are connected to the density
491: operators in the usual way:
492: \begin{eqnarray}
493: \rho_{p} = \frac{1}{\sqrt{2}} \left( \nabla \Phi + p \nabla \Theta \right).
494: \end{eqnarray}
495: Tomonaga-Luttinger liquid parameter ${\cal K}$ is defined below. For the
496: purpose of this Section it is enough to remember that for small $g_0$
497: parameter ${\cal K}$ is close to unity:
498: \begin{eqnarray}
499: 1 - {\cal K} = {\cal O} (g_0).
500: \end{eqnarray}
501: As we see from (\ref{Hnl_bos}), the boson interaction is proportional to
502: $v'_{\rm F}$. Now we perform the Bogoliubov transformation:
503: \begin{eqnarray}
504: \tilde \Theta = {\cal K}^{1/2} \Theta,\\
505: \tilde \Phi = {\cal K}^{-1/2} \Phi.
506: \end{eqnarray}
507: Our hamiltonian becomes:
508: \begin{eqnarray}
509: H'&=&\int dx \frac{\tilde v_{\rm F}}{2} \left\{ 
510: \left( \nabla \tilde \Theta \right)^2 +
511: \left( \nabla \tilde \Phi \right)^2 \right\} \\
512: &+& \frac{2 \pi^2 v'_{\rm F}}{3\sqrt{2}}
513: \left\{ \left( {\cal K}^{-1/2} \nabla \tilde \Theta +
514: {\cal K}^{1/2} \nabla \tilde \Phi \right)^3 +
515: \left( {\cal K}^{-1/2} \nabla \tilde \Theta -
516: {\cal K}^{1/2}\nabla \tilde \Phi \right)^3 \right\}. \nonumber
517: \end{eqnarray}
518: We rewrite it as follows:
519: \begin{eqnarray}
520: H'&=&H_{\rm free} + \Delta H,\\
521: H_{\rm free}&=&\int dx \frac{\tilde v_{\rm F}}{2} \left\{ 
522: \left( \nabla \tilde \Theta \right)^2 +
523: \left( \nabla \tilde \Phi \right)^2 \right\} 
524: + \frac{2 \pi^2 v'_{\rm F}}{3\sqrt{2}} \left\{ \left( \nabla \tilde \Theta +
525: \nabla \tilde \Phi \right)^3 +
526: \left( \nabla \tilde \Theta -
527: \nabla \tilde \Phi \right)^3 \right\} ,\label{Hfree}\\
528: \Delta H&=& \int dx \frac{2 \pi^2 v'_{\rm F}}{3\sqrt{2}} \left\{
529: \left( {\cal K}^{-1/2} \nabla \tilde \Theta + {\cal K}^{1/2}
530: \nabla \tilde \Phi \right)^3 +
531: \left( {\cal K}^{-1/2} \nabla \tilde \Theta - {\cal K}^{1/2}
532: \nabla \tilde \Phi \right)^3 \right.\\
533: &-& \left. \left( \nabla \tilde \Theta + \nabla \tilde \Phi \right)^3 -
534: \left( \nabla \tilde \Theta - \nabla \tilde \Phi \right)^3 \right\}.
535: \nonumber
536: \end{eqnarray}
537: We split the above hamiltonian into two part for a reason. The first part
538: $H_{\rm free}$ becomes a free quasiparticle hamiltonian upon
539: refermionization. Indeed, the first term in the curly brackets of
540: (\ref{Hfree}) 
541: refermionizes to become the kinetic energy of the fermions with linear
542: dispersion, the second term of (\ref{Hfree}) becomes $q^2$-correction
543: to the linear dispersion.
544: 
545: Refermionized hamiltonian $\Delta H$ contains all the quasiparticle
546: interaction. It also contains terms quadratic in $\tilde \psi$
547: which renormalize the value
548: of $v'_F$. The important point, however, is that $\Delta H$ is small if
549: $g_0$ is small:
550: \begin{eqnarray}
551: \Delta H = {\cal O} ((1 - {\cal K})v'_{\rm F}) = {\cal O} (g_0 v'_{\rm F})
552: \end{eqnarray}
553: This can be established by observing that $\Delta H$ vanishes when $g_0 =0$
554: $\Leftrightarrow$ ${\cal K} = 1$. Note also, that $\Delta H$ is small if
555: (\ref{free}) is satisfied.
556: Power counting shows that $\Delta H$ is irrelevant. Smallness and
557: irrelevance of the quasiparticle interaction implies that the
558: quasiparticles could be viewed as weakly interacting. 
559: 
560: We could look at our method from a different prospective. The hamiltonian
561: $H' = H_{\rm kin} + H_{\rm int} + H_{\rm nl}$ in addition to a quadratic
562: (in $\psi$)
563: part $H_{\rm kin}$ has small marginal operator $H_{\rm int}$. Plus, it has
564: an irrelevant operator $H_{\rm nl}$ which we want to account for
565: non-perturbatively. Since perturbation
566: theory in the marginal operator diverges, we must either sum certain
567: diagrams to all orders or, as it is done in the theory of
568: superconductivity, perform a Bogoliubov transformation which kills the
569: undesirable marginal operator. This is what our transformation $U$ does:
570: $U (H_{\rm kin} + H_{\rm int}) U^\dagger$ is quadratic in fermionic fields.
571: 
572: Operator $H_{\rm nl}$ is quadratic in $\psi$. In general, however, operator
573: $U H_{\rm nl} U^\dagger$ does not have to be quadratic. Thus, after the
574: removal of the marginal operator by the Bogoliubov transformation, new
575: interactions between the
576: quasiparticles are generated. Fortunately, they are irrelevant and small.
577: The smallness of the generated interactions becomes obvious if we observe
578: that for small $g_0$
579: transformation $U$ is close to the identity transformation: $U = 1 + {\cal
580: O}(g_0)$. Thus, the quasiparticle interaction induced by the action of
581: $U$ on $H_{\rm nl}$ could be treated perturbatively. This is the core of
582: the approach we proposed in the previous Section.
583: 
584: \section{Density-density propagator}
585: In this section we will calculate density-density Green's function.
586: The derivation of the correlation function for the total density operator
587: $\rho=\rho_{\rm L} + \rho_{\rm R}$ is quite simple if we note that $\rho$
588: is proportional to the quasiparticle density operator:
589: $\rho_q = (u_q + v_q)\tilde\rho_q$. Using this identity one obtains for
590: small $|q|$:
591: \begin{eqnarray}
592: {\cal D}_q (\tau) = - \frac{1}{L}
593: \left< T_\tau \left\{\rho_q(\tau) \rho_{-q}(0)\right\} \right> = 
594: {\cal K} \left(\tilde {\cal D}_{{\rm L}q}(\tau) +
595: \tilde {\cal D}_{{\rm R}q}(\tau)\right),\label{D}\\
596: {\cal K} = \left.\left(u_q + v_q\right)^2\right|_{q=0} = \sqrt{\frac{2\pi
597: v_{\rm F} - g_0}{2\pi v_{\rm F} + g_0}}.
598: \end{eqnarray}
599: The chiral Green's function $\tilde {\cal D}_{pq}(\tau) = -
600: \left< T_\tau \left\{ \tilde\rho_{pq}(\tau) \tilde \rho_{p-q}(0)
601: \right\}\right>/L$ will be calculated below to the zeroth order in the
602: quasiparticle interaction constant $\tilde g'$. Unlike expansion in orders
603: of $g_0$ the perturbative expansion in orders of $\tilde g'$ is a
604: well-defined procedure: as we explained, the quasiparticle-quasiparticle
605: interaction is irrelevant. With accuracy ${\cal O}((\tilde g')^0)$ the
606: propagator $\tilde {\cal D}_{pq}$ can be expressed as a convolution of two
607: single-quasiparticle propagators:
608: \begin{eqnarray}
609: \tilde {\cal D}_{pq}(\tau) = - \frac{1}{L}
610: \sum_k \left< T_\tau \left\{\tilde\psi^\dagger_{p(k+q)}
611: (\tau) \tilde\psi^{\vphantom{\dagger}}_{p(k+q)} (0) \right\} \right>
612: \left< T_\tau \left\{\tilde\psi^{\vphantom{\dagger}}_{pk} (\tau)
613: \tilde\psi^\dagger_{pk} (0) \right\} \right> ,\\
614: \tilde {\cal D}_{pq\omega} = \sum_k \frac{n(\tilde\varepsilon_p(k)) - 
615: n(\tilde\varepsilon_p(k+q))}{i\omega + \tilde
616: \varepsilon_p(k) - \tilde \varepsilon_p(k+q)}.\label{Dp}
617: \end{eqnarray}
618: In eq. (\ref{Dp}) $n(\varepsilon)$ is Fermi distribution function. Thus, at
619: $T=0$:
620: \begin{eqnarray}
621: \tilde {\cal D}_{pq\omega} = \frac{p}{4\pi \tilde v'_{\rm F}q} \log \left\{
622: \frac{i \omega - p \tilde v_{\rm F} q + \tilde v'_{\rm F} q^2}
623: {i \omega - p \tilde v_{\rm F} q - \tilde v'_{\rm F} q^2} \right\},\\
624: {\cal D}_{q\omega} =  \frac{{\cal K}} {4\pi \tilde v'_{\rm F} q}
625: \log \left\{ \frac{ \omega^2 + ( \tilde v_{\rm F} q -
626: \tilde v'_{\rm F} q^2)^2}{ \omega^2 + (\tilde v_{\rm F} q +
627: \tilde v'_{\rm F} q^2)^2 }\right\}.\label{Df}
628: \end{eqnarray}
629: This result has the accuracy of ${\cal O} ((v'_{\rm F} g_0)^2)$, as
630: explained above. The
631: omitted corrections are due to the quasiparticle interaction $\tilde g'$.
632: At the same time, the above expression for ${\cal D}$ is accurate to all
633: orders in the dispersion curvature $\tilde v'_{\rm F}$.
634: 
635: The retarded propagator $D_{q\omega}$ is obtained by the analytical
636: continuation:
637: \begin{eqnarray}
638: D_{q\omega} = \frac{{\cal K}} {4\pi \tilde v'_{\rm F} q}
639: \log \left\{ \frac{ ( \tilde v_{\rm F} q -
640: \tilde v'_{\rm F} q^2)^2 - (\omega + i0)^2}{ (\tilde v_{\rm F} q +
641: \tilde v'_{\rm F} q^2)^2 - (\omega +i0)^2}\right\}. \label{Dret}
642: \end{eqnarray}
643: For vanishing $\tilde v'_{\rm F}$ we can write for the Green's
644: function the following expansion:
645: \begin{eqnarray}
646: { D}_{q\omega} =  \frac{\tilde v_{\rm F} {\cal K} q^2}{\pi
647: ((\omega + i0)^2 - \tilde v_{\rm F}^2 q^2)} +
648: \frac{( \tilde v'_{\rm F})^2 {\cal K} q^5}{6\pi}
649: \left(\frac{1}{(\omega - \tilde v_{\rm F} q + i0)^3} -
650: \frac{1}{(\omega + \tilde v_{\rm F} q + i0)^3}\right) +
651: {\cal O} \left((v'_{\rm F})^4 \right) +
652: {\cal O} \left((v'_{\rm F}g_0)^2 \right).\label{D_exp}
653: \end{eqnarray}
654: The first term coincides with the well-known bosonization
655: result \cite{boson}. The second term could be also found within the
656: bosonization approach: one has to develop a perturbation theory expansion
657: in powers of $\tilde v'_{\rm F}$.
658: How this could be done is shown in Appendix E. These calculations serve as
659: yet another consistency check for our method.
660: 
661: One can make another interesting observation when examining (\ref{D_exp}).
662: By looking at this expansion it is impossible to guess that the propagator
663: (\ref{Dret})
664: has a branch-cut, not a pole. To determine the correct complex structure
665: the non-perturbative in $v'_{\rm F}$ calculations are required.
666: 
667: Such complex
668: structure of the propagator indicates that there is no coherently propagating
669: bosonic mode \cite{boson}. Instead, we have a continuum of quasiparticle -
670: quasihole pair excitations. It is possible to visualize this continuum 
671: by calculating the spectral function $B_{q\omega}$:
672: \begin{eqnarray}
673: B_{q\omega} = - 2 {\rm Im\/} D_{q\omega} =
674: \frac{{\cal K}} {2 \tilde v'_{\rm F} q} \left\{ 
675: \vartheta \left( \omega^2 - (\tilde v_{\rm F} q - \tilde v'_{\rm F} q^2)^2
676: \right) -
677: \vartheta \left( \omega^2 - (\tilde v_{\rm F} q + \tilde v'_{\rm F} q^2)^2
678: \right)
679: \right\} {\rm sgn\ } \omega.
680: \end{eqnarray}
681: (Note: unlike single-fermion spectral function $B_{q\omega}$ does not have to
682: be positive.) Every point of $(q,\omega)$ plane where $B_{q\omega} \ne 0$ 
683: corresponds to a quasiparticle - quasihole excitation with total energy
684: $|\omega|$ and total momentum $q$. The set of these points is shown on fig.2.
685: 
686: A previous attempt to account for non-zero $v'_{\rm F}$ has been made in
687: \cite{haldane,kopietz}. However, the bosonic representation used in the
688: latter references is not very convenient for such a task. To illustrate the
689: nature of the problem we consider a case of free electrons ($g_0=0$) with
690: strongly non-linear dispersion $v'_{\rm F} \Lambda \sim v_{\rm F}$:
691: \begin{eqnarray}
692: H_{\rm free} = H_{\rm kin} + H_{\rm nl}. \label{free_f}
693: \end{eqnarray}
694: The quasiparticle representation is trivial: $U = 1$ and $\psi = \tilde
695: \psi$. The spectrum of (\ref{free_f}) is 
696: \begin{eqnarray}
697: \varepsilon_k = v_{\rm F} |k| + v'_{\rm F} k^2. \label{E}
698: \end{eqnarray}
699: Density-density correlation function is given by (\ref{Df}). In the
700: bosonization framework operator 
701: $H_{\rm nl}$ corresponds to cubic interaction between bosons with the
702: dimensionless coupling constant $v'_{\rm F}\Lambda/v_{\rm F}$ of order unity
703: (see (\ref{Hnl_bos})). To compute either spectrum or Green's functions one
704: must resort to the perturbation theory whose accuracy, however, is not
705: obvious due to lack of small parameter.
706: 
707: If in addition to $v'_{\rm F}$ we have $g_0 \ne 0$ then the Bogoliubov
708: rotation of the TL bosons is required.
709: When the Bogoliubov transformation acts on $H_{\rm nl}$ it generates extra
710: interaction terms of the form $\tilde \rho_p^{\vphantom{2}}
711: \tilde \rho_{-p}^2$. The
712: coupling constant for this kind of interaction is of the order of
713: $v'_{\rm F} g_0$. Thus, in the bosonic representation one has to deal with
714: two kinds of interaction terms and two coupling constants one of which is
715: of the order of unity.
716: 
717: As we have seen in Section III, if we add weak
718: electron-electron interaction $H_{\rm int}$ (eq.(\ref{int})) to
719: $H_{\rm free}$ (eq.(\ref{free_f})) the quasiparticle representation
720: evolves continuously with $g_0$: $U=1 + {\cal O} (g_0)$,
721: $\tilde H' = H_{\rm free} + {\cal O}(g_0)$. Non-zero $g_0$ generates
722: interaction between the quasiparticles. Yet, this interaction remains
723: small (if (\ref{free}) is fulfilled) and irrelevant.
724: Consequently, the Green's function
725: is given by (\ref{Df}) with renormalized $v_{\rm F}$ and $v'_{\rm F}$, the
726: spectrum of the interacting hamiltonian has the form (\ref{E}). In
727: short, the advantage of the quasiparticle representation steams from
728: its ability to account for $v'_{\rm F}$ non-perturbatively.
729: 
730: \section{Single-electron Green's function}
731: Now we will show how to calculate the single-electron Green's function.
732: At present we can find the Green's function for the case of $v'_{\rm F} =
733: 0$ only. Although, this Green's function has been evaluated before by a
734: number of approaches we would like to include this derivation to demonstrate
735: different aspects of the method.
736: 
737: In order to calculate the single-electron Green's function it is necessary
738: to know how to express the electron field operator $\psi^\dagger$ in terms
739: of the quasiparticle operators $\tilde\psi^\dagger$ and $\tilde\rho$. The
740: following derivation answers this question. We begin by introducing unitary
741: operator $\exp(\varphi_p(x))$ where
742: \begin{equation}
743: \varphi_p(x) = \sum_{q \ne 0} \frac{\beta_{pq}}{n_q} {\rm e}^{iqx}\rho_{pq},
744: \quad \beta_{pq} - {\rm real}\ {\rm coefficients}. 
745: \end{equation}
746: Let us calculate the commutator of this operator with $\rho_{pq}$:
747: \begin{equation}
748: \left[ {\rm e}^{\varphi_p(x)}, \rho_{pq} \right] =
749: \left( {\rm e}^{\varphi_p(x)} \rho_{pq} {\rm e}^{-\varphi_p(x)} -
750: \rho_{pq} \right) {\rm e}^{\varphi_p(x)}.
751: \end{equation}
752: The first term in brackets can be calculated easily:
753: \begin{equation}
754: {\rm e}^{\varphi_p(x)} \rho_{pq}{\rm e}^{-\varphi_p(x)} = \rho_{pq} + \left[
755: \varphi_p (x), \rho_{pq} \right] = \rho_{pq} + \beta_{p-q} p {\rm e}^{-iqx}.
756: \end{equation}
757: Thus, the formula for the commutator is:
758: \begin{equation}
759: \left[ {\rm e}^{\varphi_p}, \rho_{pq} \right] = p\beta_{p-q} {\rm e}^{-iqx}
760: {\rm e}^{\varphi_p}.
761: \end{equation}
762: If we choose $\beta_{pq}=p$ the product $\psi^\dagger_p {\rm e}^{\varphi_p}$
763: commute with $\rho_{p'q}$ for any $p'$ and $q \ne 0$. Therefore, it commutes
764: with $\Omega$, eq.(\ref{Omega}). This means that such product is invariant
765: under the action of $U$. The action of $U$ on the field operator is given by:
766: \begin{eqnarray}
767: \tilde \psi_p^\dagger = U^\dagger \psi_p^\dagger U= \psi_p^\dagger
768: {\rm e}^{\varphi_p} U^\dagger {\rm e}^{-\varphi_p} U.
769: %= \psi_p^\dagger {\cal F}_p^\dagger,\\
770: %{\cal F}_p^\dagger(x) = \exp \left( -p \sum_q \frac{1}{n_q} {\rm e}^{iqx}
771: %\left( \left(u_q - 1\right) \rho_{pq} + v_q \rho_{-pq} \right) \right),\\
772: %\left[ \psi_p^\dagger (x) , {\cal F}_p^\dagger (x) \right] = 0.
773: \end{eqnarray}
774: The latter formula is easy to invert:
775: \begin{eqnarray}
776: \psi_p^\dagger(x) = \tilde \psi_p^\dagger(x) {\rm e}^{\tilde \varphi_p(x)}
777: U {\rm e}^{-\tilde \varphi_p(x)} U^\dagger
778:  = \tilde \psi_p^\dagger (x) \tilde {\cal F}_p^\dagger(x),\\
779: \tilde {\cal F}_p^\dagger(x) = \exp \left( -p \sum_{q \ne 0} \frac{1}{n_q}
780: {\rm e}^{iqx}
781: \left( w_q \tilde \rho_{pq} + v_q \tilde \rho_{-pq}
782: \right) \right),\\
783: \left[\tilde \psi_p^\dagger (x),
784: \tilde {\cal F}_p^\dagger (x) \right] = 0,\\
785: w_q = u_q - 1.
786: \end{eqnarray}
787: It gives us the desired equation for $\psi^\dagger$ in terms of $\tilde
788: \psi^\dagger$ and $\tilde \rho$.
789: 
790: Let us calculate the correlation function:
791: \begin{eqnarray}
792: \left< \psi_p^\dagger (x,\tau) \psi_p^{\vphantom{\dagger}} (0,0) \right> =
793: \left< \tilde\psi_p^\dagger (x,\tau) \tilde {\cal F}_p^\dagger (x,\tau)
794: \tilde {\cal F}_p^{\vphantom{\dagger}} (0,0)
795: \tilde \psi_p^{\vphantom{\dagger}} (0,0) \right>. \label{corr}
796: \end{eqnarray}
797: The simplest way to evaluate this expression is to transform it in the
798: following manner. The density operators with $pn_q<0 (pn_q>0)$ must be shifted
799: to the left (right) end, the quasiparticle field operators stay in the
800: middle. The reason
801: for such choice is that $\tilde \rho_{pq}\left|0\right>=0$ for $pn_q>0$ and 
802: $\left<0\right|\tilde \rho_{pq}=0$ for $pn_q<0$ where $\left| 0 \right>$ is
803: the ground state of (\ref{H}). The details of this
804: derivation are given in Appendix F. The result is
805: \begin{eqnarray}
806: \left< \psi_p^\dagger (x,\tau) \psi_p^{\vphantom{\dagger}} (0,0) \right>
807: =\left< \tilde \psi_p^\dagger (x,\tau)
808: \tilde \psi_p^{\vphantom{\dagger}} (0,0) \right> {\exp}\left\{  - \sum_q
809: (1-{\rm e}^{iqx- \tilde v_ {\rm F} |q|\tau})\frac{v_q^2}{\left|n_q\right|}
810: \right\}
811: = \frac{1}{2\pi(ipx- \tilde v_ {\rm F} \tau)}
812: \left( \frac{a^2}{x^2 + \tilde v_{\rm F}^2 \tau^2}\right)^\theta,\label{G}
813: \end{eqnarray}
814: where we used the notation $\theta=2v_q^2|_{q=0}^{\vphantom{2}} =
815: ({\cal K} + 1/{\cal K} - 2)/2$.  The same formula is derived using
816: bosonization \cite{haldane,boson}.
817: 
818: It is well-known fact that the above Green's function does not have pole. This
819: means that the quasiparticle state has zero overlap with the physical electron
820: state.
821: 
822: 
823: \section{Discussion}
824: 
825: In this paper we solve the TL model with the help of the unitary
826: transformation. The transformation maps the original hamiltonian on the
827: hamiltonian of weakly interacting quasiparticles.
828: 
829: Our approach easily incorporates deviation of the electron
830: dispersion from the linear form. As long as the bare interaction constant is
831: sufficiently small two new non-perturbative results (eq. (\ref{free_en}) and
832: eq. (\ref{Df})) can be derived. The derived results accounts for the
833: curvature parameter $\tilde v'_{\rm F}$ to all orders; they contain
834: errors of order $(\tilde g')^2 \sim (v'_{\rm F} g_0)^2$ due to neglected
835: interaction among the quasiparticles.
836: 
837: The ability of our method to account for $\tilde v'_{\rm F}$ to all orders
838: allows us to resolve the complex structure of the density-density
839: propagator.
840: 
841: In principle, our diagonalization technique is not the only way to derive
842: the quasiparticle representation. It is possible (fig.1) to bosonize (\ref{H})
843: then perform the Bogoliubov transformation and then fermionize the diagonal
844: bosonic hamiltonian \cite{luther-emery,kivelson}. Alternatively,
845: one can construct non-fermionic excitations using appropriate exponents of
846: bosonic operators \cite{ng}. In our case use of bosonization looks like an
847: unnecessary detour.
848: 
849: To conclude, we propose a unitary transformation which maps TL model on
850: a model of free fermions. Such approach reproduces or generalizes the 
851: TL correlation functions calculated using bosonization.
852: 
853: \section{Acknowledgments}
854: The author is grateful to D.P. Arovas, A. Castro Neto, F. Guinea, A.J.
855: Millis for useful discussions. 
856: 
857: The author would like to thank the referees who
858: suggested writing Appendices D and E. 
859: 
860: Support of the ``Dynasty" foundation of
861: Dmitri Zimin is gratefully acknowledged.
862: 
863: \section{Appendix A}
864: In order to differentiate products of field operators one must perform 
865: normal ordering of the product under consideration. The subject of this
866: Appendix is the definition of the normal ordering procedure we use in this
867: paper.
868: 
869: Most commonly a normal ordered product of two or more field operators is
870: defined with the reference to some non-interacting ground state
871: $\left|0\right>$. For example, for product of two operators such definition
872: reads:
873: \begin{eqnarray}
874: \colon\psi^\dagger (x) \psi (x')\colon = \psi^\dagger (x) \psi (x') -
875: \left<0\right|\psi^\dagger (x) \psi (x') \left|0\right>. \label{Nord}
876: \end{eqnarray}
877: The ground state is not specified in this equation. Thus, we actually have
878: infinite number of normal ordering procedures, each corresponding to a
879: particular state $\left|0\right>$. Usually, the problem at hand dictates
880: the choice of this state.
881: 
882: Normal ordered product (\ref{Nord}) possesses two properties: (i) it is
883: well-defined and analytical at $x=x'$; (ii) its ground state expectation
884: value is zero.
885: 
886: Since the ground state of the Tomonaga-Luttinger hamiltonian cannot be
887: approximated by a ground state of non-interacting physical fermions $\psi$
888: TL ground state cannot be used in (\ref{Nord}). Consequently, it is
889: necessary to explain what state we use in our definition of the normal
890: order.
891: 
892: In the bosonization framework it is customary \cite{haldane,vD-S} to normal
893: order with respect to the the ground state of the hamiltonian:
894: \begin{eqnarray}
895: H_0 = \sum_k v_{\rm F} pkc^\dagger_k c^{\vphantom{\dagger}} _k. \label{Hk0}
896: \end{eqnarray}
897: We accept this definition in our work as well.
898: Thus, for a product of two field operators we have:
899: \begin{eqnarray}
900: \colon\psi^\dagger_p (x) \psi^{\vphantom{\dagger}}_p (x')\colon =
901: \psi^\dagger_p (x) \psi_p^{\vphantom{\dagger}}  (x') - s_p(x-x'),
902: \label{Nord'}\\
903: s_p(x) = \frac{p}{2\pi i (x -ip0)},
904: \end{eqnarray}
905: where $s_p$ is equal to $\left<0 \right| 
906: \psi^\dagger_p (x) \psi^{\vphantom{\dagger}}_p (x') \left|0 \right>$.
907: In this equation the pole of 
908: $\psi^\dagger_p (x) \psi^{\vphantom{\dagger}}_p (x')$ at $x=x'$ is
909: explicitly shown. The normal ordered product is well-behaved at $x=x'$ and
910: it satisfies the property (i), formulated above.
911: 
912: It is not a miracle that property (i) remains intact despite the fact that the
913: ground states of (\ref{H}) and (\ref{Hk0}) are drastically different. Note that
914: (i) is `ultraviolet' in its nature -- it refers to short distance
915: ($|x-x'| \ll 1/\Lambda$), high energy ($|k| \gg \Lambda$) behavior
916: of the
917: operator product. Since interaction (\ref{int}) is limited in $q$-space
918: high-energy structure of the TL ground state is the same as that of
919: (\ref{Hk0}). This guarantees that (i) is satisfied.
920: 
921: Yet, the TL ground state expectation value of (\ref{Nord'}) for generic
922: values of $x$ and $x'$ is not zero. That is, property (ii) is violated.
923: Fortunately, we never need it in this paper.
924: 
925: The normal ordered product of four field operators is defined by the
926: equation:
927: \begin{eqnarray}
928: \psi_p^\dagger(x) \psi_p^{\vphantom{\dagger}}(x')
929: \psi_p^\dagger(y) \psi_p^{\vphantom{\dagger}}(y')&=&
930: \colon \psi_p^\dagger(x) \psi_p^{\vphantom{\dagger}}(x')
931: \psi_p^\dagger(y) \psi_p^{\vphantom{\dagger}}(y')\colon + 
932: s_p(x-y')\colon\psi_p^{\vphantom{\dagger}}(x') \psi_p^\dagger(y)\colon +\\
933: &&s_p(x'-y) \colon\psi_p^{\dagger}(x) \psi_p^{\vphantom{\dagger}}(y')\colon + 
934: s_p(x-y')s_p(x'-y).
935: \end{eqnarray}
936: As above, the normal ordered product can be differentiated everywhere.
937: 
938: For completeness, let us define the normal ordering of the density 
939: operators $\rho_p(x)$:
940: \begin{eqnarray}
941: \colon \rho_p (x) \rho_p (y) \colon&=&\rho_p (x) \rho_p (y) -
942: b_p (x-y),\label{Nrho2}\\
943: \colon \rho_p (x) \rho_p (y) \rho_p (z) \colon&=&\rho_p (x) \rho_p (y)
944: \rho_p (z) - b_p (x-y) \rho_p (z) - b_p (x-z) \rho_p (y) -\label{Nrho3}\\
945: &&b_p (y-z) \rho_p (x),\nonumber\\
946: b_p(x)&=&(s_p(x))^2.\label{bp}
947: \end{eqnarray}
948: We will use these definitions in Appendix C. 
949: 
950: The above rules can be
951: generalized for $T>0$: to normal order a product of density operators at
952: finite temperature one has to replace function $b_p(x)$ by function:
953: \begin{eqnarray}
954: b_{pT} (x)&=& \langle \tilde \rho_p(x) \tilde \rho_p(0) \rangle =
955: -\frac{T^2}{4 \tilde v_{\rm F}^2\sinh^2 (\pi T (x/ \tilde v_{\rm F} -ip0))}.
956: \label{bpT}
957: \end{eqnarray}
958: It is obvious that $b_{pT} = b_p$ at $T=0$.
959: In Appendix D we will need the finite $T$ definition.
960: 
961: 
962: \section{Appendix B}
963: In this Appendix we provide the derivation of (\ref{Hqp}) starting from the
964: formula (\ref{Hfin}). To express the
965: hamiltonian (\ref{Hfin}) in terms of $\psi$ rather than $\rho$ we introduce
966: a function $\Delta_q$:
967: \begin{eqnarray}
968: \Delta_q = u_q^2 + v_q^2 + \frac{g_q}{\pi v_{\rm F}} u_q v_q - 1 
969: \end{eqnarray}
970: and its Fourier transform $\hat \Delta(x)$:
971: \begin{eqnarray}
972: \hat \Delta(x) = \sum_q \left(u_q^2 + v_q^2 + \frac{g_q}{\pi v_{\rm F}}
973: u_q v_q - 1 \right) {\rm e}^{iqx}.
974: \end{eqnarray}
975: We defined $g_q$ such that $g_q \rightarrow 0$ when $|q|\rightarrow\infty$.
976: Thus, $v_q \rightarrow 0$, $u_q \rightarrow 1$ for $|q| \gg \Lambda$.
977: Therefore, the function $\Delta_q$
978: vanishes for large $|q|$. This implies that $\hat \Delta(x)$ vanishes for
979: $|x| \gg 1/\Lambda$. We may write:
980: \begin{eqnarray}
981: U H U^\dagger = 
982: \frac{\pi v_{\rm F}}{L} \sum_{pq} \rho_{pq} \rho_{p-q} +
983: \frac{\pi v_{\rm F}}{L} \sum_{pq} \Delta_q \rho_{pq} \rho_{p-q} =
984: \label{HDelta} \\
985: i v_{\rm F} \int dx \sum_p p :\psi_p^\dagger
986: \nabla\psi_p^{\vphantom{\dagger}}: +
987: \frac{\pi v_{\rm F}}{L} \sum_p
988: \int dx dx' \hat\Delta(x-x') \rho_p (x) \rho_p(x').
989: \nonumber
990: \end{eqnarray}
991: The first term in the above equation was obtained by inversion of
992: (\ref{rhorho}). The second term must be normal ordered first:
993: \begin{eqnarray}
994: \frac{\pi v_{\rm F}}{L} \sum_p \int dx dx' \hat\Delta(x-x')
995: \rho_p (x) \rho_p(x') =
996: \frac{\pi v_{\rm F}}{L} \sum_p \int dx dx' \hat\Delta(x-x') 
997: \left\{ b_p(x-x') + \colon \rho_p(x) \rho_p(x') \colon \right\}.
998: \end{eqnarray}
999: The additive constant proportional to $\int dx dx' \hat \Delta b_p$ will be
1000: disregarded. The normal ordered product of
1001: two $\rho$'s can be expanded into Taylor series in powers of $(x-x')$:
1002: \begin{eqnarray}
1003: \sum_p \int dx dx' \hat\Delta(x-x') \colon \rho_p (x) \rho_p(x')\colon &=&
1004: \sum_p \left[ \int dx' \hat\Delta(x') \right] \int dx \colon
1005: \rho_p^2 (x) \colon \label{aux}\\
1006: &+&\frac{1}{2}\left[ \int dx' x'^2 \hat\Delta(x') \right]
1007: \int dx \colon \rho_p (x) \nabla^2 \rho_p(x) \colon + \ldots,\nonumber
1008: \end{eqnarray}
1009: where ellipsis stand for higher order terms of the Taylor series. Since
1010: $\hat \Delta (x)$ vanishes for large $|x|$ the integrals of $\hat \Delta$
1011: are well-defined. We did not show the term proportional to
1012: $\colon \rho_p \nabla \rho_p \colon$ since it is total derivative
1013: $\nabla \colon \rho_p^2\colon$, thus, it vanishes upon integration.
1014: 
1015: We notice that the first term of expansion (\ref{aux}) is marginal. All
1016: other terms
1017: are irrelevant. Indeed, power counting shows that the scaling dimension of
1018: the operator $\colon \rho_p \nabla^2 \rho_p \colon$ is $4 > 2$. The scaling
1019: dimensions of the omitted terms are even higher. Thus, it is permissible
1020: to write:
1021: \begin{eqnarray}
1022: \frac{\pi v_{\rm F}}{L} \sum_p \int dx dx' \hat\Delta(x-x')
1023: \rho_p (x) \rho_p(x') = {\pi v_{\rm F}} \Delta_{0} \sum_p
1024: \int dx \colon \rho_p^2(x) \colon + (\text{irrelevant\ operators}),\\
1025: \Delta_0 = \Delta_q\big|_{q=0}.
1026: \end{eqnarray}
1027: Consequently, transforming $\colon \rho_p^2 \colon$ into the kinetic energy
1028: operator with the help of (\ref{rhorho}) and substituting the result into
1029: (\ref{HDelta}), we get the desired equation:
1030: \begin{eqnarray}
1031: U H U^\dagger & =&
1032: \int dx \sum_p ip\tilde v_{\rm F} : \psi_p^\dagger
1033: \nabla\psi_p^{\vphantom{\dagger}}: + ({\rm irrelevant\ operators}), \\
1034: \tilde v_{\rm F}&=&v_{\rm F} ( 1 + \Delta_{0} ) = 
1035: v_{\rm F}\left.\left(u_q^2 + v_q^2 + \frac{g_q}{\pi v_{\rm F}} u_q v_q
1036: \right)\right|_{q=0} = v_{\rm F}
1037: \sqrt{1-\left(\frac{g_0}{2\pi v_{\rm F}}\right)^2}.
1038: \end{eqnarray} 
1039: Note, that the omitted operators, in addition to being irrelevant, are also
1040: small ($\sim \Delta_0 \sim g_0^2/v_{\rm F}^2$) if interaction is small
1041: $g_0 \ll v_ {\rm F}$.
1042: 
1043: \section{Appendix C}
1044: 
1045: In this Appendix we show how $H_{\rm nl}$ can be expressed in terms of the
1046: quasiparticle field operators $\tilde \psi^\dagger$, $\tilde \psi$.
1047: We start with expression for the local density ${\cal H}_{\rm nl}$:
1048: \begin{eqnarray}
1049: {\cal H}_{\rm nl} (x)&=&v'_{\rm F} \sum_p \colon (\nabla \psi_p^\dagger(x))
1050: (\nabla \psi_p(x)) \colon - \frac{1}{6}\nabla^2 \rho_p (x) = \label{Hnl1}\\
1051: &&\frac{4\pi^2v'_{\rm F}}{3} \sum_p \lim_{z \rightarrow x}
1052: \lim_{y \rightarrow x} \left\{
1053: \rho_p (z) \left[ \rho_p (x) \rho_p(y) - b_p (x-y) \right] -
1054: 2b_p (z-y) \rho_p(y) \right\}. \nonumber
1055: \end{eqnarray}
1056: The last line can be abbreviated if one use the definition of the normal
1057: ordering for the density operators $\rho_p (x)$, eq. (\ref{Nrho3}).
1058: With this notation we can rewrite the formula for ${\cal H}_{\rm nl}$:
1059: \begin{eqnarray}
1060: {\cal H}_{\rm nl} = \frac{4\pi^2 v'_{\rm F}}{3} \sum_p \colon \rho_p^3 \colon.
1061: \label{Hnl2}
1062: \end{eqnarray}
1063: To express the density operator $\rho_p (x)$ in terms of the quasiparticle
1064: density operators $\tilde \rho_p (x)$ we must remember that:
1065: \begin{eqnarray}
1066: \rho_{pq} = u_q \tilde \rho_{pq} + v_q \tilde \rho_{-pq}\text{\ for\ }
1067: q \ne 0,\\
1068: N_p = \tilde N_p \text{\ for\ zero\ modes.}
1069: \end{eqnarray}
1070: Therefore, the density operator in co-ordinate space is equal to:
1071: \begin{eqnarray}
1072: \rho_p (x) = \delta \tilde \rho_{p0} +
1073:  \tilde \rho_p^u (x) + \tilde \rho_{-p}^v (x),\\
1074: \delta \tilde \rho_{p0} = \left.
1075: \left( (1 - u_q) \tilde N_p - v_q \tilde N_{-p} \right)\right|_{q=0},
1076: \end{eqnarray}
1077: where $\tilde \rho_p^{u}$ is the convolution of the density operator
1078: $\tilde \rho_p (x)$ with $\hat u(x)$:
1079: \begin{eqnarray}
1080: \tilde \rho_p^u (x) = (\hat u*\tilde \rho_p )(x) =
1081: \int dx' \hat u(x-x') \tilde \rho_p (x'),\\
1082: \hat u(x) = \int \frac{dq}{2\pi} u_q {\rm e}^{iqx}.
1083: \end{eqnarray}
1084: Likewise, $\tilde \rho_{-p}^v = (\hat v*\tilde \rho_{-p}) (x)$, where
1085: function $\hat v(x)$ is defined in the same manner as $\hat u(x)$. It is
1086: tempting to equate ${\cal H}_{\rm nl}$ and
1087: \begin{eqnarray}
1088: \frac{4\pi^2 v'_{\rm F}}{3} \sum_p \colon \left( \delta \tilde \rho_{p0} 
1089: + \tilde \rho_p^u + \tilde \rho_{-p}^v \right)^3 \colon.
1090: \end{eqnarray}
1091: This, however, is not exactly accurate since the normal ordering and
1092: Bogoliubov transformation do not `commute' with each other. Such phenomena
1093: becomes obvious if we were to think in terms of Bose creation and
1094: annihilation operators. Normal ordering places all creation operators to the
1095: left of all annihilation operators. Bogoliubov transformation mixes creation
1096: and annihilation operators, thus, it spoils normal ordering. 
1097: %(Warning: the above passage
1098: %must not create an illusion that our manipulations in this
1099: %Appendix depend on existence of the TL bosons. We want to remind that our
1100: %normal ordering of the density operators is merely a convenient notation
1101: %which allows us to write a compact formula like (\ref{Hnl2}) instead of
1102: %(\ref{Hnl1}). However, the result of the Appendix could be reproduced without
1103: %any problem if all expressions $\colon \rho^3 \colon$ are replaced by
1104: %expressions similar to (\ref{Hnl1}).)
1105: 
1106: To rewrite ${\cal H}_{\rm nl}$ in terms of $\tilde \rho^{u,v}$ correctly let
1107: us first examine the expression in square brackets in eq. (\ref{Hnl1}):
1108: \begin{eqnarray}
1109: \rho_p (x) \rho_p (y) - b_p (x-y) =
1110: \colon \rho_p (x) \rho_p (y) \colon =
1111: \left( \delta \tilde \rho_{p0} + \tilde \rho_p^u (x) + \tilde \rho_{-p}^v (x)
1112: \right)  \left( \delta \tilde \rho_{p0} + \tilde \rho_p^u (y) +
1113: \tilde \rho_{-p}^v (y) \right) - b_p (x-y) = \\
1114: \left\{ \colon \tilde \rho_p^u (x) \tilde \rho_{p}^u (y) \colon +
1115: (\hat u * b_p *\hat u) (x-y) - b_p (x-y) \right\} +
1116: \left\{ \colon \tilde \rho_{-p}^v (x) \tilde \rho_{-p}^v (y) \colon +
1117: (\hat v * b_{-p} * \hat v) (x-y) \right\} + \nonumber\\
1118: \left\{\tilde \rho_p^u (x) \tilde \rho_{-p}^v (y) +
1119: \tilde \rho_{-p}^v (x) \tilde \rho_{p}^u (y) \right\} +
1120: \delta \tilde \rho_{p0} \left\{ \tilde \rho_p^u (x) + \tilde \rho_{-p}^v (x)
1121: + \tilde \rho_p^u (y) + \tilde \rho_{-p}^v (y) \right\} +
1122: \delta \tilde \rho_{p0}^2 , \nonumber
1123: \end{eqnarray}
1124: where we applied the following transformations:
1125: \begin{eqnarray}
1126: \tilde \rho^u_p (x) \tilde \rho^u_p (y) = \int dx' dy' u(x-x') u(y-y') \tilde 
1127: \rho_p (x') \tilde \rho_p (y')\\
1128:  = \int dx' dy' u(x-x') u(y-y') \left(
1129: \colon \tilde \rho_p (x') \tilde \rho_p (y') \colon + b_p (x'-y') \right)
1130: \nonumber\\
1131: = \colon \tilde \rho_p (x) \tilde \rho_p (y) \colon + (\hat u*b_p*\hat u).
1132: \nonumber
1133: \end{eqnarray}
1134: Similar transformations could be done for $\tilde \rho^v(x) \tilde \rho^v(y)$.
1135: Since $u_q^2 - v_q^2 = 1$ the identity:
1136: \begin{eqnarray}
1137: \hat u*f*\hat u = f + (\hat v*f*\hat v)
1138: \end{eqnarray}
1139: holds true. Therefore:
1140: \begin{eqnarray}
1141: \colon \rho_p (x) \rho_p (y) \colon = 
1142: \colon \left( \tilde \rho_p^u (x) + \tilde \rho_{-p}^v (x) \right)
1143: \left( \tilde \rho_p^u (y) + \tilde \rho_{-p}^v (y) \right) \colon
1144: + \left( \hat v * (b_p + b_{-p}) * \hat v \right)(x-y) + \\
1145: \delta \tilde \rho_{p0} \left\{ \tilde \rho_p^u (x) + \tilde \rho_{-p}^v (x)
1146: + \tilde \rho_p^u (y) + \tilde \rho_{-p}^v (y) \right\} +
1147: \delta \tilde \rho_{p0}^2 , \nonumber
1148: \end{eqnarray}
1149: The extra terms in this formula are non-singular functions of $(x-y)$. In
1150: particular:
1151: \begin{eqnarray}
1152: \left( \hat v * (b_p + b_{-p}) * \hat v \right) (x-y)=
1153: \int \frac{dq}{(2\pi)^2} {\rm e}^{iq(x-y)} |q| v_q^2.
1154: \end{eqnarray}
1155: Observe that the normal ordered product of $\rho$'s differs from the normal
1156: ordered product of $\tilde \rho$'s by a non-singular operator.
1157: 
1158: Generalizing the above calculations for the product of three $\rho$'s 
1159: we obtain the following expression for ${\cal H}_{\rm nl}$:
1160: \begin{eqnarray}
1161: {\cal H}_{\rm nl} = \frac{4\pi^2v'_{\rm F}}{3}
1162: \sum_p \colon (\tilde \rho^u_p)^3
1163: \colon + \colon (\tilde \rho^v_{-p})^3 \colon +
1164: 3 \tilde \rho^u_p \colon ( \tilde \rho_{-p}^v )^2 \colon +
1165: 3 \tilde \rho^v_{-p} \colon ( \tilde \rho_{p}^u )^2 \colon +
1166: \tilde c\left(\delta \tilde \rho_{p0} + \tilde \rho^u_p + \tilde \rho_{-p}^v
1167: \right) + (\text{z.m.}), \label{Hnl3}\\
1168: \tilde c = \frac{3}{4\pi^2} \int dq |q| v_q^2,
1169: \end{eqnarray}
1170: where `z.m.' stand for zero modes terms which vanish in the
1171: thermodynamic limit.
1172: 
1173: For small momenta the above formula can be simplified. What must be done,
1174: in its substance, amounts to inversion of equations (\ref{Hnl1}) and
1175: (\ref{rhorho}).
1176: To illustrate this statement, consider the first and the second terms of 
1177: (\ref{Hnl3}). Limiting ourselves to the case $\Lambda = \infty$ we can
1178: neglect the $q$-dependence of $u_q$ and $v_q$. Then,
1179: $\tilde \rho_p^u = u \tilde \rho_p$, $\tilde \rho_p^v = v \tilde \rho_p$
1180: and it is possible to write:
1181: \begin{eqnarray}
1182: \frac{4\pi^2v'_{\rm F}}{3}
1183: \sum_p \colon (\tilde \rho^u_p)^3 \colon + \colon (\tilde \rho^v_p)^3 \colon
1184: = \frac{4\pi^2v'_{\rm F}}{3} (u^3 + v^3) \sum_p \colon \tilde \rho_p^3
1185: \colon =  v_{\rm F}'  (u^3 + v^3)
1186: \left( \colon \nabla \tilde \psi_p^\dagger \nabla \tilde
1187: \psi_p^ {\vphantom{\dagger}} \colon - \frac{1}{6} \nabla^2 \tilde \rho_p
1188: \right). \label{finite_Lambda}
1189: \end{eqnarray}
1190: When handling the third and the fourth terms of (\ref{Hnl3}) we are to act in
1191: the similar fashion: the product $\colon \tilde \rho_p^2 \colon$ should be
1192: rewritten in terms of $\tilde \psi$ with the help of (\ref{rhorho}). These
1193: two terms give the quasiparticle interaction.
1194: 
1195: For a finite value of $\Lambda$ our task becomes somewhat more complicated.
1196: In such a situation one can generalize the procedure of Appendix B.  Let us
1197: briefly discuss the core of this generalization. As an example, consider
1198: the expression:
1199: \begin{eqnarray}
1200: \colon (\tilde \rho_p^v)^3 \colon  = \int dx' dx'' dx''' v(x-x') v(x-x'')
1201: v(x-x''')
1202: \colon \tilde \rho_p (x')\tilde \rho_p (x'')\tilde \rho_p (x''') \colon.
1203: \end{eqnarray}
1204: It is easy to show with the help of Taylor expansion (see Appendix B) that
1205: the normal ordered product 
1206: $\colon \tilde \rho_p (x')\tilde \rho_p (x'')\tilde \rho_p (x''') \colon $
1207: is equal
1208: to $\colon \tilde \rho_p^3 (x)\colon + (\text{more\ irrelevant\ operators})$,
1209: where `more irrelevant operators' stands for operators whose scaling
1210: dimension is bigger than 3.
1211: Therefore, we obtain the following equation:
1212: \begin{eqnarray}
1213: \colon (\tilde \rho_p^v)^3 \colon = a\colon \tilde \rho_p^3
1214: \colon + (\text{more\ irrelevant\ operators}),
1215: \end{eqnarray}
1216: where coefficient $a$ is equal to $v_q^3 |_{q=0}$. The expression 
1217: $\colon \tilde \rho_p^3 \colon$ has to be transformed further as in eq.
1218: (\ref{finite_Lambda}). The remaining terms of (\ref{Hnl3})
1219: are transformed similarly.
1220: Therefore, one establishes:
1221: \begin{eqnarray}
1222: {\cal H}_{\rm nl}&=&
1223: \sum_p \left\{ \tilde v'_{\rm F} \colon (\nabla \tilde \psi_p^\dagger)
1224: (\nabla \tilde \psi_p) \colon - \frac{\tilde v'_{\rm F}}{6}\nabla^2 \tilde
1225: \rho_p
1226: + i p \tilde g' \tilde \rho_{-p} \left( \colon \tilde \psi^\dagger_p 
1227: (\nabla \tilde \psi^{\vphantom{\dagger}}_p) \colon -
1228: \colon (\nabla \tilde \psi^\dagger_p) \tilde \psi^{\vphantom{\dagger}}_p
1229: \colon \right)\right\} +  \label{Hnl} \\
1230: &&\tilde \mu ( \tilde \rho_{\rm L} +
1231: \tilde \rho_{\rm R} ) +
1232: (\text{more\ irrelevant\ operators}), \nonumber\\
1233: \tilde v'_{\rm F}&=&v'_{\rm F} \left. \left( u_q^3 + v_q^3 \right)
1234: \right|_{q=0},\\
1235: \tilde g'&=&2\pi v'_{\rm F} \left.\left( u_q^2 v_q^{\vphantom{2}} +
1236: u_q^{\vphantom{2}} v_q^2 \right) \right|_{q=0},\\
1237: \tilde \mu&=& v'_{\rm F} \left(\int dq |q| v_q^2 \right).
1238: \end{eqnarray}
1239: The values of $u_q$ and $v_q$ are given by (\ref{uv}). Finally, we find for
1240: the hamiltonian:
1241: \begin{eqnarray}
1242: H_{\rm nl} = \sum_p \int dx \left\{
1243: \tilde v'_{\rm F}\colon (\nabla \tilde \psi^\dagger_p)
1244: (\nabla \tilde \psi^{\vphantom{\dagger}}_p ) \colon +
1245: i p \tilde g' \tilde \rho_{-p} \left( \colon \tilde \psi^\dagger_p 
1246: (\nabla \tilde \psi^{\vphantom{\dagger}}_p) \colon -
1247: \colon (\nabla \tilde \psi^\dagger_p) \tilde \psi^{\vphantom{\dagger}}_p
1248: \colon \right) + \tilde \mu \tilde \rho_p\right\} + (\text{more\
1249: irrelevant\ operators}).
1250: \end{eqnarray}
1251: We omitted the full derivative $\nabla^2 \rho$ from the hamiltonian since it
1252: contributes on the system boundary only.
1253: 
1254: We don't want to prove here that the quasiparticle interaction produces no
1255: ultraviolet divergence of the perturbation theory. Instead, this issue is
1256: examined at the end of Appendix F. This is because the formalism developed
1257: there is particularly convenient for discussion of ultraviolet properties of
1258: the quasiparticle interaction.
1259: 
1260: Finally, let us briefly explain, why the quasiparticle interaction
1261: generated by the action of $U$ on $H_{\rm nl}$ is small. This explanation
1262: is equivalent to the one given in Sec.III. Initially, our $H_{\rm
1263: nl}$ could be expressed as a cube of $\rho_p(x)$, eq. (\ref{Hnl2}).
1264: (Although, superficially, this expression appears to be sixth order in
1265: $\psi_p$, a subtler analysis shows that it is quadratic in $\psi_p$, eq.
1266: (\ref{Hnl1}).) The
1267: Bogoliubov transformation $U$ converts (\ref{Hnl2}) into (\ref{Hnl3}).
1268: Unlike (\ref{Hnl2}), eq. (\ref{Hnl3}) is indeed a sixth order polynomial in
1269: $\psi_p$. However, for small $g_0$ operator $U$ is close to the identity
1270: transformation: $U= 1 + {\cal O}(g_0)$. Therefore,
1271: \begin{eqnarray}
1272: v'_{\rm F} \colon \rho^3_p \colon = v'_{\rm F}\colon \tilde \rho^3_p \colon
1273: + {\cal O}(g_0v'_{\rm F}).
1274: \end{eqnarray}
1275: The first term on the right is proportional to $\nabla \tilde
1276: \psi_p^\dagger \nabla \tilde \psi_p^{\vphantom{\dagger}}$, eq. (\ref{Hnl1}).
1277: All generated terms, such as quasiparticle interactions and corrections to
1278: $v'_{\rm F}$, are ${\cal O}(g_0v'_{\rm F})$.
1279:                                                                                 
1280: \section{Appendix D}
1281: In this Appendix we show how the result for the specific heat,
1282: eq.(\ref{C(T)}), can be obtained within the bosonization framework. More
1283: specifically, ${\cal O}((v'_{\rm F})^2)$ correction to the specific heat
1284: will be calculated with the help of the perturbation theory for the free
1285: boson hamiltonian:
1286: \begin{eqnarray}
1287: H = \frac{\pi \tilde v_{\rm F}}{2} \sum_{pq} \tilde \rho_{pq} \tilde
1288: \rho_{p-q}.\label{Hbos0}
1289: \end{eqnarray}
1290: We will see that it is the same as 
1291: ${\cal O}((v'_{\rm F})^2)$ term in eq.(\ref{C(T)}). 
1292: 
1293: Our first step is to rewrite the non-linear dispersion hamiltonian $H_{\rm
1294: nl}$ in a form particularly suitable for $T>0$ calculations: since in this
1295: Appendix we work at finite temperature it is natural to express $H_{\rm nl}$
1296: with the help of the finite temperature generalization of (\ref{Nrho2})
1297: and (\ref{Nrho3}). As it was explained, such generalization is achieved by
1298: placing $b_{pT}$, eq. (\ref{bpT}), instead of $b_p$, eq.(\ref{bp}).
1299: 
1300: Using the new rules of the normal ordering we could prove that at
1301: $T\ge 0$ the following is true:
1302: \begin{eqnarray}
1303: \tilde v'_{\rm F} \sum_p \colon (\nabla \tilde 
1304: \psi_p^\dagger(x)) (\nabla \tilde \psi_p(x)) \colon - \frac{1}{6}
1305: \nabla^2 \tilde \rho_p (x) = 
1306: \frac{4\pi^2 \tilde v'_{\rm F}}{3} \sum_p \colon \tilde \rho_p^3 \colon +
1307: \frac{T^2}{4 \tilde v_{\rm F}^2}\tilde \rho_p,\label{HnlT_dens}
1308: \end{eqnarray}
1309: where $\colon\ldots\colon$ denotes here the finite temperature normal
1310: order.
1311: The proof of this equation is simple. Start from eq.(\ref{rho3}):
1312: \begin{eqnarray}
1313: \tilde v'_{\rm F} \sum_p \colon (\nabla \tilde 
1314: \psi_p^\dagger(x)) (\nabla \tilde \psi_p(x)) \colon - \frac{1}{6}
1315: \nabla^2 \tilde \rho_p (x) = \\
1316: \frac{4\pi^2 \tilde v'_{\rm F}}{3}
1317: \sum_p \lim_{y \rightarrow x} \lim_{z \rightarrow y} \left\{
1318: \tilde \rho_p (x) \left[ \tilde \rho_p (z) \tilde \rho_p(y) -
1319: b_p (z-y) \right] - 2b_p (x-y) \tilde \rho_p(y) \right\} = \nonumber\\
1320: \frac{4\pi^2 \tilde v'_{\rm F}}{3}
1321: \sum_p \lim_{y \rightarrow x} \lim_{z \rightarrow y} \left\{
1322: \tilde \rho_p (x) \left[ \tilde \rho_p (z) \tilde \rho_p(y) - b_{pT} (z-y)
1323: \right] - 2b_{pT} (x-y) \tilde \rho_p(y) \right\} + \nonumber\\
1324:  3 \tilde \rho_p(x)
1325: \lim_{y \rightarrow x} \left\{ b_{pT} (x-y) - b_p (x-y) \right\} .
1326: \nonumber
1327: \end{eqnarray}
1328: The last limit is 
1329: \begin{eqnarray}
1330: \lim_{y \rightarrow x} \left\{ b_{pT} (x-y) - b_p (x-y) \right\} =
1331: \frac{T^2}{12 \tilde v_{\rm F}^2}.
1332: \end{eqnarray}
1333: Thus, bosonized $H_{\rm nl}$ equals to
1334: \begin{eqnarray}
1335: H_{\rm nl} = \frac{4\pi^2 \tilde v'_{\rm F}}{3} \sum_p \int dx
1336: \left( \colon \tilde \rho_p^3 \colon + \frac{T^2}{4 \tilde v_{\rm F}^2}
1337: \tilde \rho_p \right) + {\cal O}(g_0 v'_{\rm F}). \label{HnlT}
1338: \end{eqnarray}
1339: This formula does not imply that $H_{\rm nl}$ depends on the temperature.
1340: In this Appendix the symbol $\colon \ldots \colon$ denotes 
1341: the finite $T$
1342: generalization of the normal ordering, consequently, the expression
1343: $\colon \tilde \rho^3 \colon$ varies with temperature. The $T^2$ term
1344: compensates the temperature dependence of the normal ordered product
1345: $\colon \tilde \rho^3 \colon$ so that $H_{\rm nl}$ is temperature
1346: independent as it must be. We remind the reader that
1347: such a way of writing is a matter of convenience.
1348: 
1349: In the above expression we ignore the quasiparticle interaction and the
1350: correction to the chemical potential since we are interested in ${\cal
1351: O}((v'_{\rm F})^2)$ corrections only while the neglected terms contribute
1352: to the thermodynamics at ${\cal O} ((v'_{\rm F} g_0)^2)$.
1353: 
1354: The correction to the free energy due to $H_{\rm nl}$ is given by the usual
1355: perturbation theory formula:
1356: \begin{eqnarray}
1357: \delta F = - \frac{1}{2} \int d\tau \langle H_{\rm nl} (\tau)
1358: H_{\rm nl} (0) \rangle = - (\tilde v'_{\rm F})^{2} L \int d\tau dx \left(
1359: \frac{16\pi^4}{9} \langle \colon \tilde \rho^3_{\rm L}(x,\tau) \colon
1360: \colon \tilde \rho^3_{\rm L}(0,0) \colon \rangle + \frac{\pi^4 T^4}{9
1361: \tilde v_{\rm F}^4} 
1362: \langle \tilde \rho_{\rm L} (x,\tau) \tilde \rho_{\rm L}(0,0) \rangle \right).
1363: \end{eqnarray}
1364: Using Wick's theorem we find:
1365: \begin{eqnarray}
1366: \langle \colon \tilde \rho^3_{\rm L}(x,\tau) \colon
1367: \colon \tilde \rho^3_{\rm L}(0,0) \colon \rangle  =
1368: 6 \langle \tilde \rho_{\rm L} (x,\tau) \tilde \rho_{\rm L} (0,0)
1369: \rangle^3,\\
1370: \langle \tilde \rho_{\rm L} (x,\tau) \tilde \rho_{\rm L} (0,0) \rangle = -
1371: \frac{T^2}{4 \tilde  v_{\rm F}^2 \sinh^2(\pi T (x/\tilde v_{\rm F} - i\tau))}.
1372: \end{eqnarray}
1373: Therefore:
1374: \begin{eqnarray}
1375: \delta F = ( \tilde v'_{\rm F})^2 L \int d\tau dx \left(
1376: \frac{\pi^4 T^6} {6 \tilde v_{\rm F}^6}
1377: \frac{1}{\sinh^6 (\pi T (x/ \tilde v_{\rm F}
1378: - i\tau))} + \frac{ \pi^4 T^6}{36 \tilde v_{\rm F}^6}
1379: \frac{1}{\sinh^2 (\pi T (x/ \tilde v_{\rm F} - i\tau))} \right).
1380: \end{eqnarray}
1381: It is easy to calculate the integral 
1382: \begin{eqnarray}
1383: \int_{-\infty}^{+\infty}
1384: \frac{dx}{ \sinh^2(\pi T (x/ \tilde v_{\rm F} - i\tau))} =
1385: - \frac{2 \tilde v_{\rm F}}{\pi T}
1386: \end{eqnarray}
1387: for $0 < \tau < \beta$.
1388: To find the integral of $1/\sinh^6$ we use a trick. One can check directly
1389: that:
1390: \begin{eqnarray}
1391: \frac{1}{\sinh^{6} r} = \frac{8}{15}
1392: \frac{1}{\sinh^2 r} + \frac{d^2 I(r)}{dr^2},\\
1393: I(x) = \frac{1}{20}
1394: \frac{1}{\sinh^4 r} -
1395: \frac{2}{15} \frac{1}{\sinh^2 r}.
1396: \end{eqnarray}
1397: The integral of the full derivative is zero and we establish
1398: for $0 < \tau < \beta$
1399: \begin{eqnarray}
1400: \int_{-\infty}^{+\infty}
1401: \frac{dx}{\sinh^6 (\pi T (x/ \tilde v_{\rm F} - i\tau))} =
1402: \frac{8}{15} \int_{-\infty}^{+\infty}
1403: \frac{dx}{\sinh^2 (\pi T (x/ \tilde v_{\rm F} - i\tau))} =
1404: - \frac{16 \tilde v_{\rm F}}{15 \pi T}.
1405: \end{eqnarray}
1406: Therefore, one finds the following corrections to the free energy, the 
1407: entropy and the specific heat:
1408: \begin{eqnarray} 
1409: \delta F/L = - \frac{7\pi^3( \tilde v'_{\rm F})^2 T^4}
1410: {30 \tilde v_{\rm F}^5},\\
1411: \delta S/L = - \frac{\partial (\delta F/L)}{\partial T} = 
1412: \frac{14\pi^3( \tilde v'_{\rm F})^2 T^3}{15 \tilde v_{\rm F}^5},\\
1413: \delta C = T \frac{\partial (\delta S/L)}{\partial T} = 
1414: \frac{14\pi^3( \tilde v'_{\rm F})^2 T^3}{5 \tilde v_{\rm F}^5}.
1415: \end{eqnarray}
1416: The last expression coincides with the second term of eq.(\ref{C(T)}) as it
1417: should be.
1418: 
1419: \section{Appendix E}
1420: In this Appendix we calculate ${\cal O}((v'_{\rm F})^2)$ correction
1421: to the zero-temperature boson propagator. Such correction must coincide
1422: with the second term of (\ref{D_exp}). 
1423: 
1424: The correction in question to the chiral Matsubara propagator comes from
1425: $H_{\rm nl}$, eq.(\ref{HnlT}), with $T=0$. The appropriate Feynman
1426: diagram is shown on fig.3. As one can see from eq.(\ref{D})
1427: the corresponding expression is:
1428: \begin{eqnarray}
1429: \delta {\cal D}_p (x,\tau)= - \frac{{\cal K}}{2}
1430: \left(\frac{4\pi^2 \tilde v'_{\rm F}}{3}\right)^2
1431: \int d\tau' d\tau'' dx' dx'' \langle T_\tau \left\{ \tilde \rho_p (0,0)
1432: \colon \tilde \rho_p^3 (x',\tau') \colon
1433: \colon \tilde \rho_p^3 (x'',\tau'') \colon \tilde \rho_p(x,\tau) \right\}
1434: \rangle_{\rm con- \atop nected} + {\cal O}((v'_{\rm F} g_0)^2).
1435: \end{eqnarray}
1436: The time-ordered average with respect to hamiltonian (\ref{Hbos0}) 
1437: is equal to:
1438: \begin{eqnarray}
1439: \int d\tau' d\tau'' dx' dx'' 
1440: \langle T_\tau \left\{ \tilde \rho_p (0,0)
1441: \colon \tilde \rho_p^3 (x',\tau') \colon
1442: \colon \tilde \rho_p^3 (x'',\tau'') \colon \tilde \rho_p(x,\tau) \right\}
1443: \rangle_{\rm con- \atop nected} = \label{loop} \\
1444: 3\times 3 \times 2 \times 2 
1445: \int d\tau' d\tau'' dx' dx'' \tilde {\cal D}_p^0(\tau', x')
1446: \left( \tilde {\cal D}_p^0 (\tau'' - \tau', x'' - x') \right)^2 
1447: \tilde {\cal D}_p^0 (\tau-\tau'',x-x'') ,\nonumber
1448: \end{eqnarray}
1449: where $\tilde {\cal D}_p^0(x,\tau)$ is the chiral free boson Green's
1450: function
1451: $
1452: \tilde {\cal D}_p^0 (x,\tau) = - (1/L)
1453: \langle T_\tau \left\{ \tilde \rho_p (0,0)
1454: \tilde \rho_p (x,\tau) \right\}\rangle 
1455: $ for hamiltonian (\ref{Hbos0}).
1456: 
1457: The numerical factor on the right-hand side of (\ref{loop})
1458: corresponds to 36 possible ways
1459: of contracting the density operators into a connected diagram presented on
1460: fig.3. First, the external operator $\tilde \rho_p (0,0)$ could contract
1461: with either of two vertexes; the operator $\tilde \rho_p(x,\tau)$ contracts
1462: with the remaining vertex. This gives a factor of two. Second, a given
1463: external
1464: operator could contract with either of three density operators in a vertex.
1465: This gives a factor of three for one external operator and another factor
1466: of three for the second external operator. Finally, after contracting with
1467: the external operators, every vertex has two uncontracted density
1468: operators. For them there are two ways of contraction. Thus, we have yet
1469: another factor of two.
1470: 
1471: In the Fourier space we can write:
1472: \begin{eqnarray}
1473: {\cal D}_{pq\omega} \approx {\cal K} \tilde {\cal D}_{pq\omega}^0 +
1474: 32 \pi^4 {\cal K} ( \tilde v'_{\rm F})^2
1475: \left( \tilde {\cal D}_{pq\omega}^0 \right)^2 \tilde \Pi_{pq\omega},\\
1476: \tilde {\cal D}_{pq\omega}^0 = \frac{1}{2\pi}
1477: \frac{pq}{i\omega - p \tilde v_{\rm F} q},\\
1478: \tilde \Pi_{pq\omega} = -T\sum_{\Omega} \int \frac{dk}{2\pi}
1479: \tilde {\cal D}_{pk\Omega}^0 \tilde {\cal D}_{p(q-k)(\omega-\Omega)}^0
1480: = \frac{1}{48 \pi^3} \frac{p q^3}{i\omega - p \tilde v_{\rm F} q}.
1481: \end{eqnarray}
1482: Therefore, for correction to the full propagator we write:
1483: \begin{eqnarray}
1484: \delta {\cal D}_{q\omega} = \sum_p \delta {\cal D}_{pq\omega} =
1485: \frac{{\cal K}( \tilde v'_{\rm F})^2 q^5}{6\pi} \left(
1486: \frac{1}{(i\omega - \tilde v_{\rm F} q)^3}
1487: - \frac{1}{(i\omega + \tilde v_{\rm F} q)^3} \right).
1488: \end{eqnarray}
1489: After analytical continuation we recover the ${\cal O}((v'_{\rm F})^2)$
1490: term of (\ref{D_exp}).
1491: 
1492: 
1493: \section{Appendix F}
1494: In this Appendix we provide a detailed derivation of the single-electron
1495: Green's function for TL model, equation (\ref{G}). The formalism developed for
1496: the Green's function is also a convenient tool to prove that the quasiparticle
1497: perturbation theory has no ultraviolet divergences.
1498: 
1499: We start with equation (\ref{corr}). First, the product $\tilde
1500: {\cal F}^\dagger_p \tilde {\cal F}^{\vphantom{\dagger}}_p$ has to be 
1501: normal-ordered:
1502: \begin{eqnarray}
1503: \tilde {\cal F}^\dagger_p (x,\tau)
1504: \tilde {\cal F}^{\vphantom{\dagger}}_p(0,0)&=&
1505: {\rm e}^{{\bf W}_p (x + ipy,0)} {\rm e}^{-{\bf W}_p^\dagger (x + ipy,0)} 
1506: {\rm e}^{{\bf V}_p (x - ipy,0)} {\rm e}^{-{\bf V}_p^\dagger (x - ipy,0)} \times
1507: \label{FF} \\
1508: %&&{\rm e}^{p\sum_q
1509: %\frac{1}{n_q} \left(1-{\rm e}^{iqx - pqy}\right)
1510: %w_q \tilde \rho_{pq} \vartheta(-pq)} 
1511: %{\rm e}^{p\sum_q
1512: %\frac{1}{n_q} \left(1-{\rm e}^{iqx - pqy}\right)
1513: %w_q \tilde \rho_{pq} \vartheta(pq) } \times 
1514: %\nonumber
1515: %\\
1516: %&&{\rm e}^{p\sum_q
1517: %\frac{1}{n_q} \left(1-{\rm e}^{iqx + pqy}\right)
1518: %v_q \tilde \rho_{-pq}\vartheta(pq) }
1519: %{\rm e}^{p\sum_q
1520: %\frac{1}{n_q} \left(1-{\rm e}^{iqx + pqy}\right)
1521: %v_q \tilde \rho_{-pq} \vartheta(-pq) } \times \nonumber\\
1522: &&{\rm e}^{-p\sum_q
1523: \frac{1}{n_q} \left(1-{\rm e}^{iqx - pqy}\right) w_q^2 \vartheta(pq) }
1524: {\rm e}^{p\sum_q \frac{1}{n_q} \left(1-{\rm e}^{iqx + pqy}\right)
1525: v_q^2  \vartheta(-pq)}, \nonumber\\
1526: &&{\bf W}_p(x,x') =
1527: {p\sum_q \frac{1}{n_q} \left({\rm e}^{iqx'}-{\rm e}^{iqx  }\right)
1528: w_q \tilde \rho_{pq} \vartheta(-pq)} ,\\
1529: &&{\bf V}_{p}(x,x')=
1530: {p\sum_q \frac{1}{n_q} \left({\rm e}^{iqx'}
1531: - {\rm e}^{iqx }\right)
1532: v_q \tilde \rho_{-pq}\vartheta(pq) }.
1533: \end{eqnarray}
1534: Here $y=\tilde v_{\rm F} \tau$.
1535: This equation was derived with the help of commutation relations
1536: (\ref{rho_comm}). In order to calculate $\tilde \rho(\tau)$ we used the
1537: expression (\ref{Hr}).
1538: 
1539: Second, we transform the whole expression 
1540: $\tilde \psi^\dagger_p (x,\tau) \tilde {\cal F}^\dagger_p (x,\tau)
1541:  \tilde {\cal F}^{\vphantom{\dagger}}_p (0,0)
1542: \tilde \psi^{\vphantom{\dagger}}_p(0,0)$. We shift $\exp({\bf W}_p)$
1543: to the left past the
1544: quasiparticle field operator $\tilde \psi^\dagger_p$. The exponent
1545: $\exp(-{\bf W}^\dagger_p)$
1546: is shifted to the very right. To perform these shifts the
1547: commutation rule (\ref{psi_comm}) has to be used.
1548: \begin{eqnarray}
1549: &&\tilde \psi^\dagger_p (x,\tau) \tilde {\cal F}^\dagger_p (x,\tau)
1550:  \tilde {\cal F}^{\vphantom{\dagger}}_p (0,0)
1551: \tilde \psi^{\vphantom{\dagger}}_p(0,0) =  
1552: \left\{{\rm e}^{{\bf W}_p(x + ipy,0)}
1553: \left( 
1554: \tilde \psi^\dagger_p(x,\tau) \tilde \psi^{\vphantom{\dagger}}_p (0,0)
1555: \right) {\rm e}^{-{\bf W}_p^\dagger (x+ipy,0)}\right\}
1556: \times \label{field}\\
1557: &&\left({\rm e}^{{\bf V}_p(x - ipy,0)}
1558: {\rm e}^{-{\bf V}_p^\dagger(x - ipy,0)} \right)
1559: {\rm e}^{-p\sum_q \frac{1}{n_q} \left(1-{\rm e}^{iqx - pqy}\right)
1560: \left(w_q^2 + 2w_q\right)\vartheta(pq)}
1561: {\rm e}^{p\sum_q \frac{1}{n_q} \left(1-{\rm e}^{iqx + pqy}\right)
1562: v_q^2  \vartheta(-pq)}. \nonumber
1563: \end{eqnarray}
1564: After taking the expectation value of this expression and recalling that
1565: $w_q^2 + 2 w_q = v_q^2$ we obtain the desired expression for the correlation
1566: function.
1567: 
1568: The above formula sets a convenient background for a rather general
1569: discussion of the ultraviolet properties of the quasiparticle-quasiparticle
1570: interactions. Imagine that we add the following term to the TL hamiltonian:
1571: \begin{eqnarray}
1572: H_{\rm extra} = \sum_p \int dx dx' \hat Z(x-x') \psi^\dagger_p (x)
1573: \psi^{\vphantom{\dagger}}_p (x').
1574: \end{eqnarray}
1575: Operator $H_{\rm nl}$ is a particular case of $H_{\rm extra}$ with $\hat Z
1576: = - \hat h''$.
1577: 
1578: With the help of (\ref{field}) it is easy to rewrite $H_{\rm extra}$ in terms
1579: of the quasiparticle operators $\tilde \psi$ and $\tilde \rho$:
1580: \begin{eqnarray}
1581: H_{\rm extra} = \sum_p \int dx dx' \hat Z(x-x') \zeta(x-x') \left\{
1582: {\rm e}^{{\bf W}_p(x,x')} \left( \tilde \psi^\dagger_p (x)
1583: \tilde \psi^{\vphantom{\dagger}}_p (x') \right)
1584: {\rm e}^{-{\bf W}_p^\dagger (x,x')} \right\}
1585: \left( {\rm e}^{{\bf V}_{-p}(x,x')}
1586: {\rm e}^{-{\bf V}_{-p}^\dagger(x,x')}\right),\label{Hextra}\\
1587: \zeta(x) =
1588: {\rm e}^{\sum_q \frac{1}{|n_q|} \left(1-{\rm e}^{iqx  }\right) v_q^2 }.
1589: \end{eqnarray}
1590: Function $\zeta(x)$ is well-defined for $|x| \ll 1/\Lambda$ and vanishes
1591: algebraically for large $|x|$. Operator (\ref{Hextra}) is normal-ordered,
1592: therefore, it is safe to expand the exponentials:
1593: \begin{eqnarray}
1594: H_{\rm extra} = \sum_{n, m, n', m'} H_{nmn'm'} =
1595:  \sum_{n,m,n',m'} \int dx dx' \frac{\hat Z \zeta}{n!m!n'!m'!}
1596: {\bf W}_p^n \tilde \psi^\dagger_p \tilde \psi^{\vphantom{\dagger}}_p
1597: (-{\bf W}_p^\dagger)^m {\bf V}_p^{n'}(-{\bf V}_p^\dagger)^{m'}
1598: \end{eqnarray}
1599: Since ${\bf W} = {\cal O}(g_0^2)$ and ${\bf V} = {\cal O}(g_0)$ such
1600: expansion is controlled by the smallness of $g_0$. 
1601: We want to argue that any interaction term $H_{nmn'm'}$,
1602: $n + m + n' + m' > 0$ does not lead to ultraviolet divergence of perturbation
1603: theory. Consider, for example, the second order correction to the ground
1604: state energy:
1605: \begin{eqnarray}
1606: \delta E_{nn'} =  - \sum_N \sum_{p_1\ldots p_{N}  \atop k_1\ldots k_{N}  }
1607: \frac{ \left| \left< {p_1 \ldots p_N \atop k_1 \ldots k_N} \Big| H_{n0n'0}
1608: \Big|0\right> \right|^2}{\sum_i \tilde \varepsilon (p_i) + 
1609: \tilde \varepsilon (k_i)} \delta(\sum_i p_i + k_i). \label{2-corr}
1610: \end{eqnarray}
1611: Summation in the above formula goes over $N\leq n+n'+1$ quasiparticle
1612: momenta $p_i$ and over $N$ quasihole momenta $k_i$. We include $m=m'=0$
1613: terms only because other terms annihilate the ground state. Thus,
1614: $m,m'\ne 0$ terms do not contribute to the second order ground state
1615: correction. They do, however, contribute to higher order corrections and to
1616: corrections to different propagators.
1617: 
1618: If the matrix element does not vanish at large momenta the expression
1619: (\ref{2-corr}) suffers from the ultraviolet divergence. We will show that the
1620: matrix element goes to zero if at least one $|p_i|$ or $|k_i|$ exceeds
1621: $2(n+n')\Lambda$.
1622: 
1623: Let us examine definitions of ${\bf W}$ and ${\bf V}$. Since $v_q$ and $w_q$
1624: vanishes for $|q| > \Lambda$ every individual operator ${\bf W,V}$ changes the
1625: total momentum by no more than $\Lambda$. Thus, momentum change produced by
1626: $({\bf W}^n {\bf V}^{n'})$ is no more than $(n+n') \Lambda$. We will use this
1627: information in our next step.
1628: 
1629: First step in evaluating the matrix element from (\ref{2-corr}) is creation
1630: of a quasiparticle-quasihole pair by applying $\tilde\psi^\dagger \tilde\psi$
1631: product to $\left| 0 \right>$. Since the total momentum of all excitations is
1632: zero the momentum of this pair cancels the momentum of
1633: $({\bf W}^n {\bf V}^{n'})$. Thus, the pair momentum magnitude is limited by
1634: $(n+n')\Lambda$. Two elementary excitations, quasiparticle and quasihole, of
1635: which the pair is composed have their momenta bound by $(n + n')\Lambda$.
1636: 
1637: As we explained above, by acting on the ground state the product
1638: $\tilde \psi^\dagger \tilde \psi$ creates an excited state with a
1639: quasiparticle and a quasihole. On this state the operator 
1640: $({\bf W}^n {\bf V}^{n'})$ acts creating yet another excited state. Every
1641: individual operator ${\bf V}$ or ${\bf W}$ when acting on a state with finite
1642: number of the elementary excitations can either (i) create another
1643: quasiparticle-quasihole pair or (ii) replace an elementary excitation with
1644: another one of the same chirality and charge but different momentum.
1645: 
1646: In case (i) both newly created elementary excitations have their momenta
1647: bound by $\Lambda$. This is because single ${\bf W}$ or ${\bf V}$ operator
1648: cannot make quasiparticle-quasihole excitation with total momentum higher than
1649: $\Lambda$.
1650: 
1651: In case (ii) an individual operator ${\bf W}$ or ${\bf V}$ can change
1652: momentum of an elementary excitation by no more that $\Lambda$. As above, this
1653: is a consequence of $v_q$ and $w_q$ vanishing at $|q| > \Lambda$. Therefore,
1654: the product $({\bf W}^n {\bf V}^{n'})$ cannot change momentum of an elementary
1655: excitation by more than $(n+n')\Lambda$. Consequently, momentum of an
1656: elementary excitation is bound by $2(n+n')\Lambda$.
1657: 
1658: This means that in (\ref{2-corr}) summation is effectively performed over a
1659: sphere $|k_i| < 2 (n + n') \Lambda$, $|p_i| < 2 (n + n') \Lambda$. Thus, this
1660: expression has no ultraviolet divergence.
1661: 
1662: The above argument can be easily generalized to other perturbation theory
1663: formulas to prove that the perturbation theory for the quasiparticles has no
1664: ultraviolet divergences.
1665: 
1666: The absence of the ultraviolet divergences, together with irrelevance of the
1667: quasiparticle - quasiparticle interaction, implies that the quasiparticle
1668: perturbation theory is well-defined.
1669: 
1670: \begin{thebibliography}{99}
1671: 
1672: \bibitem{haldane} F.D.M. Haldane, {\sl J. Phys. C: Solid State Phys.} {\bf
1673: 14}, 2585 (1981).
1674: 
1675: \bibitem{boson} A.O. Gogolin, A.A. Nersesyan, A.M. Tsvelik, {\sl
1676: Bosonization and Strongly Correlated Systems}, Cambridge University Press
1677: (1998).
1678: 
1679: \bibitem{carmelo} J.M.P. Carmelo, A.H. Castro Neto, {\sl Phys. Rev. B} {\bf
1680: 54}, 11230 (1996).
1681: 
1682: \bibitem{vD-S} Jan von Delft, Herbert Schoeller, {\sl Annalen der Physik}
1683: {\bf 4}, 225 (1998); preprint {\sl cond-mat/9805275}
1684: 
1685: \bibitem{kopietz} Peter Kopietz, {\sl Bosonization of Interacting Fermions
1686: in Arbitrary Dimensions}, Springer (1997).
1687: 
1688: \bibitem{luther-emery} A. Luther and V.J. Emery, {\sl Phys. Rev. Lett.}
1689: {\bf 33}, 589 (1974).
1690: 
1691: \bibitem{kivelson} V.J. Emery, S. Kivelson, {\sl Phys. Rev. B} {\bf 46}
1692: 10812 (1992)
1693: 
1694: \bibitem{ng} Tai-Kai Ng, unpublished, preprint {\sl cond-mat/0302604}.
1695: 
1696: \bibitem{jaxodraw} The diagram is drawn
1697: with JaxoDraw, see D. Binosi and L. Theussl, preprint {\sl hep-ph/0309015}.
1698: 
1699: \end{thebibliography}
1700: \newpage
1701: 
1702: \begin{figure} [!t]
1703: \centering
1704: \leavevmode
1705: %\epsfxsize=8cm
1706: \epsfysize=6cm
1707: %\epsfbox[18 144 592 718] {fig1.eps}
1708: \epsfbox{fig1.eps}
1709: \caption[]
1710: {\label{fig1} 
1711: Commutative diagram explains the relation between transformation $U$,
1712: and the bosonization. `BT' stands for the Bogoliubov transformation of
1713: bosons.
1714: }
1715: \end{figure}
1716: \begin{figure} [!t]
1717: \centering
1718: \leavevmode
1719: %\epsfxsize=8cm
1720: \epsfysize=5cm
1721: \epsfbox{boson.eps}
1722: %\epsfbox[18 -100 592 718] {boson.eps}
1723: \caption[]
1724: {\label{fig2} 
1725: When $v'_{\rm F} = 0$ the spectral density of $D_{q\omega}$ is delta-function
1726: centered at $\omega = \tilde v_{\rm F} q$ line (dash line on the figure).
1727: This is the dispersion curve of the TL bosons. For $v'_{\rm F} \ne 0$ the
1728: spectral
1729: density is non-zero within the whole shaded area. This area represents the
1730: continuum of the quasiparticle-quasihole excitations. The TL bosons acquire
1731: finite life-time in such a situation.
1732: }
1733: \end{figure}
1734: \begin{figure} [!b]
1735: \centering
1736: \leavevmode
1737: %\epsfxsize=12cm
1738: \epsfysize=4cm
1739: %\epsfbox[18 144 592 718] {3bos.eps}
1740: \epsfbox{3bos.eps}
1741: \caption[]
1742: {\label{fig3} 
1743: Correction to $\tilde {\cal D}^0_{pq}$ due to the
1744: dispersion curvature. Wavy lines correspond to $\tilde {\cal D}^0$, the
1745: vertexes are proportional to $\tilde v'_{\rm F}$ \cite{jaxodraw}.
1746: }
1747: \end{figure}
1748: 
1749: 
1750: \end{document}
1751: 
1752: 
1753: 
1754: