cond-mat0310569/erg.tex
1: %\documentclass[prb,twocolumn,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}% Physical Review B
2: \documentclass[prb,twocolumn,showpacs,amsmath,amssymb]{revtex4}
3: \usepackage{amsfonts}
4: 
5: \usepackage{graphicx}% Include figure files
6: \usepackage{dcolumn}% Align table columns on decimal point
7: \usepackage{bm}% bold math
8: \newcommand{\be}{\begin{equation}}
9: \newcommand{\ee}{\end{equation}}
10: \newcommand{\lbl}{\label}
11: \newcommand{\ptl}{\partial}
12: 
13: \def\mm#1{\mathrm{#1}}
14: \def\tm#1{\textrm{#1}}
15: \def\eq#1{(\ref{#1})}
16: \def\la{\langle}
17: \def\ra{\rangle}
18: 
19: \newcommand{\beqn}{\begin{eqnarray}}
20: \newcommand{\eeqn}{\end{eqnarray}}
21: \newcommand{\nn}{\nonumber}
22: \newcommand{\ai}{\emph{ab initio }}
23: \newcommand{\dr}{$\rho$ }
24: 
25: \def\eq#1{(\ref{#1})}
26: \def\ket#1{\vert{#1}\rangle}
27: \def\bra#1{\langle{#1}\vert}
28: \def\br#1{\langle{#1}}
29: \def\ord#1{{\cal O}\left({#1}\right)}
30: \def\hf{\frac{1}{2}}
31: 
32: %\nofiles
33: 
34: \begin{document}
35: 
36: %\preprint{APS/123-QED}
37: 
38: \title{Projection method for rapid ab initio calculations of metals }
39: % Force line breaks with \\
40: \author{Abdelouaheb Kenoufi$^a$}\email{kenoufi@lpt1.u-strasbg.fr}
41: \author{Janos Polonyi$^{a,b}$}\email{polonyi@lpt1.u-strasbg.fr}
42: \affiliation{$^a$Laboratoire de Physique Th\'eorique, Universit\'e Louis
43: Pasteur\\
44: 3-5 rue de l'Universit\'e F-67084 Strasbourg Cedex, France}
45: \affiliation{$^b$Department of Atomic Physics, Lorand E\"otv\"os University,
46: Budapest, Hungary}
47: 
48: 
49: 
50: \date{\today}% It is always \today, today,
51:              %  but any date may be explicitly specified
52: 
53: \begin{abstract}
54: An improvement of the Energy Renormalization Group method is proposed for
55: systems with small gap, based on the projection methods developed by
56: H. Feshbach. It is tested for the ground state energy of the
57: one-dimensional tight-binding model.
58: \end{abstract}
59: 
60: \pacs{31.15.Ar,71.15.Dx}
61: %\keywords{Suggested keywords}%Use showkeys class option if keyword
62:                               %display desired
63: \maketitle
64: 
65: \section{Introduction}
66: The key point in the construction of fast, $\ord{N}$ numerical methods is
67: the gap in the excitation spectrum above the ground state which
68: renders the interactions short ranged. Systems without gap
69: display infrared singular, long range interactions which slow down
70: the convergence of the numerical algorithms. It seems natural to seek
71: different strategies to deal with the short and the long range quantum
72: fluctuations. In particular, one may use rapid numerical methods
73: for the short range fluctuations and treat the more difficult
74: long range sector with slower, more sophisticated method. Such
75: a mixed numerical algorithm is discussed in this paper.
76: 
77: The strategy of the renormalization group\cite{Wilson_1975} is a natural
78: candidate for the construction of such an algorithm. In fact, the
79: renormalization group is a systematic method to successively eliminate certain
80: degrees of freedom or fluctuation modes in such a manner that their impact on
81: the dynamics is accumulated in the effective theory which is constructed
82: for the remaining degrees of freedom. The algorithm proposed in this paper
83: consists of two steps. First, a rapid numerical method is applied for the 
84: elimination of the short range fluctuations. What is left is a dynamical problem
85: of the long range fluctuations described by an effective Hamiltonian.
86: This problem is dealt with in the second step of the algorithm by the 
87: diagonalization of the effective Hamiltonian.
88: 
89: The question of central importance for such a mixed method is the
90: relation between the total dimension of the Hilbert space and the
91: dimensionality of the linear space of the effective theory
92: where the exact diagonalization is performed. Let us denote by
93: $\cal E$ an intrinsic energy scale of the system and introduce
94: $N_>$ and $N_<$ as the number of modes with energy superior or
95: inferior of $\cal E$. One may call $N_>$ and $N_<$
96: ultraviolet and infrared cut-off. Our algorithm will be
97: $\ord{N_>^2}$ but will slow down as $N_<$ is
98: increased. Since $N_<$ growths with the size of the system
99: in the absence of gap and remains finite when the gap is present
100: our algorithm might be useful for systems with weak gap
101: or truly gapless models of finite size. The numerical efficiency
102: compared to other methods will be judged by the prefactor
103: of $N_>$ in the required computer time so long the system size or the
104: gap is kept fixed. We believe that this prefactor will be rather small
105: because the modes treated in this step are short ranged.
106: 
107: The traditional renormalization group method\cite{Wilson_1975}
108: consists of the repeated application of a three step procedure.
109: The first step is the blocking, the elimination of certain
110: variables from the system. This is usually achieved by
111: the lowering of the ultraviolet cut-off, the highest energy the
112: fluctuations may reach in the system. The second step is the
113: construction of the effective theory for the remaining modes.
114: Finally, in the third step which gave the name of the procedure,
115: one performs a rescaling of the energy or other scales of the effective
116: theory in order to restore the original ultraviolet cut-off.
117: This last step is not always necessary.
118: 
119: There have already been a proposal in the literature for a partial implementation
120: of this idea, the so called energy space renormalization
121: group\cite{BHG_1997_3,BHG_1998}, realizing the blocking and the
122: rescaling steps. In order to render this scheme systematical
123: one can not be content with the naive elimination of the unwanted
124: modes, the restriction of the Hamiltonian into a subspace, but should
125: realize the second step as well, the accumulation of the effects of the
126: excluded directions within the subspace retained. For this purpose we
127: use a projection method developed in nuclear physics\cite{HF_1962,Rau_1996,Muller_Rau_1996}.
128: 
129: In sec. \ref{dmf}, we expand the density matrix formalism, which is the
130: foundement of \ai algorithms.
131: The locality principle and its use in linear-scaling methods are presented in
132: sec. \ref{np}.
133: An exemple of such algorithm, the so called Fermi Operator Expansion is
134: presented in section \ref{lsa}. In the last section
135: we develop a Numerical Renormalization Group method in Hilbert Space around
136: the Fermi level and propose an improvement inspired by projection method.
137: 
138: 
139: 
140: \section{Density matrix formalism}\lbl{dmf}
141: In the framework of Density Functional Theory (DFT)
142: \cite{HK_1964,K_1965,Cap_2003}, particularly in the Kohn-Sham
143: scheme \cite{K_1965}, rapid \ai calculation methods
144: allowing linear scaling or near-linear scaling computation time have been
145: developed recently \cite{G_1999,WJ_2002}. Most of the rapid \ai algorithm
146: is based on the one-particle reduced density matrix \dr which is
147: assumed to be a projector on the subspace spanned by the
148: low-lying occupied states according to the \emph{auf bau} principle :
149: \beqn
150: \rho &=&
151: \sum\limits_{i}f_{\infty,\mu}(\epsilon_{i})\ket{\psi_{i}}\bra{\psi_{i}}\nn \\
152: &=&\sum\limits_{i=1}^{N/2}\ket{\psi_{i}}\bra{\psi_{i}}
153: \eeqn
154: where $N$ denotes the number of electrons and
155: $(\epsilon_{i},\ket{\psi_{i}})$ is an eigenfunction of the
156: Kohn-Sham Hamiltonian $H$ and $f_{\beta}$ is defined as the
157: Fermi-Dirac distribution function at the inverse temperature $\beta$,
158: \be
159: f_{\beta,\mu}(\epsilon)=\frac{1}{1+e^{\beta(\epsilon-\mu)}}.
160: \ee
161: The form $H=\sum\limits_{i}\epsilon_{i}\ket{\psi_{i}}\bra{\psi_{i}}$
162: allows us to write the density matrix as
163: \beqn\lbl{ff}
164: \rho &=& f_{\infty,\epsilon_{F}}(H)\nonumber\\
165: &=&\Theta(\epsilon_{F}-H)
166: \eeqn
167: $\Theta$ being the Heaviside function.
168: The average energy and the particle number can be written as
169: \be
170: E_{BS}=\sum\limits_{i,j}\rho_{ij}H_{ji}
171: \ee
172: and
173: \be
174: N=2\sum\limits_{i,j}\rho_{ij}S_{ji}
175: \ee
176: where the density matrix is given in a localized orbital basis
177: \be\lbl{rang}
178: \rho=\sum\limits_{ij}\rho_{ij}\ket{\phi_i}\bra{\phi_j}
179: \ee
180: and $H_{ij} = \bra{\phi_{i}}H\ket{\phi_{j}}$,
181: $S_{ij}=\br{\phi_{i}}\ket{\phi_{j}}$.
182: 
183: 
184: 
185: \section{Principle of "nearsightedness"}\lbl{np}
186: This principle states \cite{K_1959,JC_1964,K_1993} that the matrix
187: elements of the one-electron density matrix are negligible beyond
188: the distance $ca$ where $a$ is the lattice spacing,
189: \be\lbl{dmd2}
190: \vert i-j \vert>c \Rightarrow \rho_{ij}\simeq 0,
191: \ee
192: giving
193: \be\lbl{enp2}
194: E_{BS}\simeq2\sum\limits_{i}\sum\limits_{max(0,i-c)<j<min(N,i+c)}\rho_{ij}H_{ji}.
195: \ee
196: 
197: The decay of the density matrix $\rho$ in real-space depends on the
198: material. For systems with gap the
199: decay is exponential \cite{K_1959,JC_1964,K_1993,K_1995,IB}
200: \be
201: \rho(\vec{r},\vec{r}')=e^{-\alpha\vert\vec{r}-\vec{r}'\vert}
202: \ee
203: where $\alpha\sim\sqrt{\Delta\epsilon_{gap}}$ for the tight-binding limit and
204: $\alpha\sim a_{lattice}\cdot\Delta\epsilon_{gap}$ for the weak-binding limit.
205: For systems with no gap the decay of \dr is algebraic at zero
206: temperature\cite{IB,GI_1998,MYS}
207: \be
208: \rho(\vec{r},\vec{r}')=k_{F}\frac{\cos(k_{F}\vert\vec{r}-\vec{r}'\vert)}{\vert\vec{r}-\vec{r}'\vert^{2}}
209: \ee
210: Such an algebraic decay reflects the presence of long range correlations
211: and prevents linear-scaling in the numerical calculations.
212: The electron states tend to be more localized for disordered systems
213: and the matrix elements of the density matrix decay faster with
214: the distance\cite{Anderson,K_1996}.
215: 
216: 
217: 
218: \section{Fermi Operator Expansion}\lbl{lsa}
219: The polynomial expansion of \dr in Chebychev polynomials, the so called
220: Fermi Operator Expansion\cite{G_1994,G_1995,G_1998} is an important ingredient
221: of rapid algorithms.
222: 
223: Chebychev polynomials are defined by the recursion formula
224: \beqn\lbl{recursion}
225: T_{0}(x)&=&1\nn\\
226: T_{1}(x)&=&x \\
227: T_{n+2}(x)&=&2xT_{n+1}(x)-T_{n}(x)\nn
228: \eeqn
229: for $-1\le x\le1$. It is easy to see that actually
230: $T_{n}(x)=\cos(n\arccos x)$.
231: We use the functional form \eq{ff} in order to fit \dr
232: with a Chebychev polynomials up to order $p$,
233: \be
234: \rho_\beta=f_{\beta,\mu}(H)\simeq\sum\limits_{i=0}^pa_i(\beta_s,\mu_s)T_i(H_s)
235: \ee
236: where $H_s$ is the dimensionless Hamiltonian scaled and shifted into
237: the interval $[-1,1]$,
238: \beqn
239: H_{s}&=&\frac{H-\bar{E}}{\Delta E}\nn\\
240: \bar{E}&=&\hf[\min(spec(H))+\max(spec(H))]\\
241: \Delta E&=&\hf[\max(spec(H))-\min(spec(H))]\nn
242: \eeqn
243: and $\beta_{s}=\beta\Delta E$, $\mu_{s}=(\mu-\bar{E})/\Delta E$.
244: The smallest and largest eigenvalue of $H$ can be computed by using the
245: Lanczos method which scales linearly with the size of the matrix.
246: The projection coefficients
247: \be
248: a_n(\beta_s,\mu_s)=\frac{2-\delta_{n0}}{\pi}\int_{0}^{\pi}\cos (n
249: \theta)\frac{1}{1+e^{\beta_{s}(\cos \theta-\mu_{s})}}d\theta
250: \ee
251: will be computed numerically by means of Fast Fourier Transform (FFT).
252: The particle number conservation fixes the value of the Fermi energy level
253: $\epsilon_F$ found by solving Eq. \eq{rang}.
254: 
255: The accuracy of Fermi Operator Expansion can be estimates by recalling that
256: one truncates the Chebychev polynoms $T_n$ of Eq. \eq{recursion}
257: in such a manner that only the matrix elements $(T_{n}(H))_{ij}$ with
258: $\vert i-j\vert\le c$ are retained. The computation time will be of
259: order $pc^2N=O(N)$. It can be shown \cite{BHG_1997_1,BHG_1997_2}
260: that the order of the Chebychev expansion should be
261: \be
262: p\simeq\frac{2}{3}(d-1)\beta_s=\frac{2}{3}(d-1)\beta\Delta E
263: \ee
264: for the accuracy $10^{-d}$ of the expansion coefficients $\{a_{i}\}$.
265: If the system has an HOMO-LUMO gap $\Delta E_{gap}$ and
266: \be
267: \beta\ge\frac{2\log_{10}d}{\Delta E_{gap}}
268: \ee
269: then
270: \be
271: p\ge\frac{4(d-1)\Delta E\log_{10}d}{3\Delta E_{gap}}.
272: \ee
273: Since the range of correlations in the density matrix is bounded,
274: \be
275: range(\rho)\le p~c\simeq \frac{2}{3}c(d-1)\beta\Delta E
276: \ee
277: the correlations grows with the inverse temperature for gapless systems
278: and the linear-scaling methods are rendered inapplicable.
279: 
280: 
281: 
282: 
283: 
284: 
285: 
286: \section{Energy Renormalization Group}\lbl{erg}
287: We present now a renormalization group method in the energy space in order
288: to treat systems with small gap. In the original version of this
289: method\cite{BHG_1997_3,BHG_1998} one starts with a series of
290: inverse temperatures $\beta_n\to\infty$ and the corresponding
291: density matrices $\rho_n$ which tend to be concentrated around the
292: Fermi level as $n\to\infty$. This alone would not represent any
293: improvement as far as the numerical difficulties of obtaining
294: the density matrices are concerned. But the density matrices are
295: constructed in decreasing subspaces ${\cal H}_n\supset{\cal H}_{n+1}$
296: where ${\cal H}_{n+1}$ is span by the eigenvectors of $\rho_n$ with
297: large eigenvalues.
298: 
299: This algorithm is modified in order to implement the blocking
300: in energy space. First a common chemical potential is introduced
301: for each temperature which is adjusted at the end of the
302: computation to dial the desired particle number. This modification
303: is needed to clear the way for the blocking. The Hamiltonians
304: were simply truncated in the original algorithm as their subspaces
305: were restricted. In order to retain the dynamics of the excluded
306: dimensions we employ a method developed in Nuclear Physics\cite{HF_1962}
307: which yields an exact, $\ord{N^2}$ algorithm.
308: 
309: 
310: 
311: 
312: \subsection{Blocking in the Hilbert pace}
313: A geometric series of inverse temperatures $\beta_n=q^n\beta_0$
314: is introduced for $q>1$ together with the corresponding
315: density matrices $\rho_{n,\mu}=f_{\beta_n,\mu}(H)$.
316: The zero temperature expectation value of an observable $A$ is written as
317: a telescopic series
318: \be \lbl{somme}
319: \la A\ra=\mm{Tr(\rho_{\infty,\mu}A)}
320: =\sum\limits_n\mm{Tr(\Delta_{n,\mu}A)}.
321: \ee
322: where
323: \be
324: \Delta_{n,\mu}=\rho_{n,\mu}-\rho_{n-1,\mu}\nn,~~~
325: \Delta_{0,\mu}=\rho_{0,\mu}.
326: \ee
327: Each term in this equation corresponds to a more restricted energy subspaces
328: centered at the Fermi energy level as $n$ is increased. The localization
329: in the energy leads to delocalized states in real space in the
330: absence of disorder.
331: The ground state is approached by the telescopic summation
332: by zooming onto the Fermi level and the corresponding density matrix
333: projects on more and more extended states.
334: 
335: The order of the Chebychev expansion $p$ is chosen to be
336: independent of $n$ and the coefficients obtained by FFT are
337: \be
338: a'_m(\beta_n,\mu)=\la\Delta_{n,\mu},T_m\ra
339: \ee
340: where
341: \be
342: \Delta_{n,\mu}=f_{\beta_n,\mu}(H_n)-f_{\beta_{n-1},\mu}(H_n).
343: \ee
344: 
345: \begin{figure}[!h] \begin{center}
346: \includegraphics[angle=-90,width=8cm]{erg1.ps}
347: \caption{\lbl{fig:figure1}Spectral representation of $\rho_n$ and
348: $\Delta_n$.}
349: \end{center} \end{figure}
350: 
351: 
352: 
353: 
354: \subsection{Fixed-point}
355: The convergence of the telescopic series can be expressed as the existence
356: of a fixed point of the blocking in the energy space for energy dependent
357: operators. In fact, let us suppose that a continuous operator $A$ is
358: commuting with the Hamiltonian $H$ and can be diagonalized in a basis of
359: eigenvectors of $H$. We can then express its expectation value by means
360: of $\rho$ as
361: \beqn
362: \mm{Tr}(\Delta_{n+1,\mu}A)&=&\int d\epsilon\ A(\epsilon)
363: \Delta_{n+1,\mu}(\epsilon)\nn \\
364: &=&\frac{\beta_n}{\beta_{n+1}}\int d\epsilon\ A
365: \left(\epsilon\frac{\beta_n}{\beta_{n+1}}\right)\Delta_{n,\mu}(\epsilon)\\
366: &=&\frac{\beta_n}{\beta_{n+1}}\mm{Tr}\left(\Delta_{n,\mu}
367: A\left(\cdot\frac{\beta_n}{\beta_{n+1}}\right)\right)\nonumber
368: \eeqn
369: This expression allows us to rescale the operator $A$ around the Fermi-level
370: by the factor $\beta_{n+1}/\beta_n$ and to keep $\Delta_{n,\mu}$ unchanged.
371: Since $A$ is a continuous operator the iteration of this step obviously
372: leads to a fixed-point,
373: \be
374: \mm{Tr}(\Delta_{n+1,\mu}A)-\mm{Tr}(\Delta_{n,\mu}A)\to0
375: \ee
376: as $n\to\infty$.
377: 
378: \subsection{Projection}
379: The identification of the subspaces proceeds by the construction of
380: the projectors $P_n:\mathcal{H}_n\to\mathcal{H}_{n+1}$.
381: We introduce first the following pseudo-projectors constructed by means
382: of the Chebychev expansion
383: \be
384: G_n=\frac{\ptl\rho_{n,\mu}}{\ptl\mu}=\beta_n\rho_{n,\mu}(1-\rho_{n,\mu})
385: \ee
386: Once the series $\{G_n\}$ is found another set of matrices $\{C_n\}$
387: is formed. The columns of $C_n$ are basis vectors of $\mathcal{H}_{n+1}$
388: by means of a heuristic version of the singular value decomposition with
389: column pivoting\cite{BHG_1997_3,BHG_1998,GVL_1996}. Since the
390: dynamics of the excluded dimensions is retained in our case this choice
391: is a less sensitive step of the algorithm then in the original
392: version and influences the sparsity of the resulting density matrices only.
393: As the next step, the overlap matrices $S_n=C_n^*C_n$ are constructed.
394: Finally, the projectors are given as $P_{i}=C_{i}S_{i}^{-1}C_{i}^{*}$.
395: $S_i^{-1}$ can actually be obtained as $S_{i}^{-1/2}=\lim_{k\to\infty}A_k$
396: by the help of the algorithm\cite{Lar}
397: \beqn\lbl{larin_1}
398: A_k&=&\frac{1}{2}(3A_k-A_kB_kA_k)\nn\\
399: B_k&=&\frac{1}{2}(3B_k-B_kA_kB_k)
400: \eeqn
401: with $A_0=-\sqrt\alpha\cdot\openone$, $B_0=-\sqrt\alpha \cdot S_{i}$
402: and $\alpha=1/\max\limits_{j,k}|(S_{i})_{jk}|$.
403: The projected Hamiltonian is of the form
404: \be
405: H_{n+1}^{ERG}=S_n^{-\hf}C_n^*H_nC_nS_n^{-\hf}
406: \ee
407: 
408: \begin{figure}[!h]
409: \begin{center}
410: \includegraphics[angle=-90,width=8cm]{erg2.ps}
411: \caption{\lbl{fig:figure2}Spectral representation of $G_{i}$ and
412: $\Delta_{i}$.}
413: \end{center}
414: \end{figure}
415: 
416: 
417: 
418: \begin{table*}
419: \begin{ruledtabular}
420: \caption{\lbl{table1} Relative errors of energy computations for different size of systems with $\beta_{0}=5$, $q=10$ and with a Chebychev
421: expansion $p=10$}
422: 
423: \begin{tabular}{cccccccc}
424: $N$&$E_{Exact}$&$E_{ERG}$&$E_{HF}$&$Error_{ERG}$&$Error_{HF}$&$Error_{ERG}-Error_{HF}$\\
425: \hline
426: 256& 93.39 &95.44& 91.58& 2.20 &1.94& 0.26 \\
427: 384& 139.90& 142.95 &139.09& 2.18 &0.58& 1.60 \\
428: 512& 186.41& 190.47& 186.61& 2.18 &0.10 &2.07  \\
429: 640 &232.93& 237.98& 234.12& 2.17 &0.51 &1.66  \\
430: 768 &279.44& 285.50& 281.63& 2.17 &0.79 &1.38  \\
431: 896 &325.95& 333.01& 329.15& 2.17 &0.98 &1.19  \\
432: 1024& 372.47& 380.52& 376.66& 2.16 &1.13 &1.04  \\
433: 1152 &418.98& 428.04& 424.17& 2.16 &1.24 &0.92  \\
434: 1280 &465.49& 475.55& 471.69& 2.16 &1.33 &0.83  \\
435: 1408& 512.00& 523.06& 519.20& 2.16 &1.41 &0.75  \\
436: 1536 &558.52& 570.58& 566.72& 2.16 &1.47 &0.69  \\
437: 1664& 605.03& 618.09& 614.23& 2.16 &1.52 &0.64  \\
438: 1792& 651.54& 665.60& 661.74& 2.16 &1.57 &0.59  \\
439: 1920 &698.05& 713.12& 709.26& 2.16 &1.60 &0.55  \\
440: 2048 &744.57& 760.63& 756.77& 2.16& 1.64& 0.52 \\
441: \end{tabular}
442: \end{ruledtabular}
443: \end{table*}
444: 
445: 
446: Up to now we have excluded certain directions of the Hilbert
447: space which are supposed to be less important from the point
448: of view of the ground state dynamics. In order to perform
449: the analogue of the Kadanoff-Wilson blocking we have to construct
450: an effective Hamiltonian\cite{HF_1962,Rau_1996} $H_{n+1}$ in
451: the restricted space with the same dynamics around the Fermi level as
452: those of $H_n$,
453: \beqn\label{fesbach}
454: H_{n+1}&=&S_n^{-\hf}C_n^*\left(H_n+H_nQ_n\frac{1}{\mu-Q_nH_nQ_n}Q_nH_n\right)\nn\\
455: &&\times C_nS_n^{-\hf}\\
456: &=&H_{n+1}^{ERG}+S_n^{-\hf}C_n^*H_nQ_n\frac{1}{\mu-H_nQ_nH_n}Q_nH_nC_nS_n^{-\hf}\nn
457: \eeqn
458: where $Q_n=1-P_n$. The exclusion of directions from the Hilbert space
459: renders the finding of the projection of the eigenvectors of the original
460: Hamiltonian into the restricted space a nonlinear problem. This
461: complication appears as a nonlinear dependence of the eigenvector
462: equation in the restricted space on the eigenvalue. The energy eigenvalue
463: was replaced by the Fermi level, $\mu$, in the 'self-energy',
464: the second term on right hand side of Eq. \eq{fesbach}.
465: The inverse in the right hand side can be obtained
466: by the well-known \emph{Schultz's} or \emph{Hotelling's}
467: method\cite{NR,Householder,PanReif} as $(\mu-H_n)^{-1}=\lim_{j\to\infty}X_j$
468: where
469: \be\lbl{pr_1}
470: X_j=X_{j-1}[2\openone-(\mu-H_n)X_{j-1}]
471: \ee
472: with the initial-guess \cite{PanReif}
473: $X_0=(\mu-H_n)^*/\sum\limits_{j,k}(\mu-H_n)_{jk}^2$.
474: 
475: The calculation ends when the dimension of the
476: subspace is sufficiently small for explicit diagonalization.
477: 
478: One can introduce approximations which render the method $\ord{N}$.
479: One possibility is the note that $X_j$ of Eq. \eq{pr_1} converges
480: quadratically and the order of 30 iterations, a value independent
481: of the system size was always sufficient in our numerical test. Another
482: possibility is based on the adjustment of the chemical potential at the
483: end of the computation. This circumstance allows us to make the replacement
484: \be\label{approx}
485: \frac{1}{\mu-Q_nH_nQ_n}\to\frac{1}{\mu}
486: \ee
487: in Eq. \eq{fesbach} where $\mu$ will include the 'average'
488: of $Q_nH_nQ_n$ within ${\cal H}_n$. Such a simplification is
489: more acceptable for large $n$ where $\mm{dim}{\cal H}_n$ is
490: not too large and the evolution is slow.
491: 
492: 
493: 
494: 
495: 
496: \subsection{Numerical test}
497: We considered a lattice of $2N$ sites in one dimension with nearest
498: neighbor interaction described by the Hamiltonian\cite{Atkins_Friedman}
499: \be\label{hamnd}
500: H=2\sum\limits_ia_i^+a_i-\sum\limits_{<i,j>}a_i^+a_j
501: \ee
502: at half filling. Being the simplest model for the conducting band electrons the matrix elements 
503: of the density matrix, computed in the appendix for half-filling, show
504: metal-like decrease with the distance.\\ 
505: Table \ref{table1} shows the results of energy calculations with the
506: algorithm of Eq. \eq{fesbach} for different sizes.
507: It has been reported\cite{BHG_1997_3,BHG_1998} that the CPU time of the
508: ERG method scales as $N\ln^2N$. The computation of Eq. \eq{fesbach}
509: which was done by applying the approximation \eq{approx}
510: does not change this result since it contains matrix multiplications only. 
511: 
512: \section{Conclusion}
513: A new application of the renormalization group method is presented in
514: this work. This method is designed to retain the dynamics of modes excluded
515: from the computation and was developed for the path integral.
516: But it is an ideal tool to improve systematically the truncations
517: of the Hilbert space committed in the operator formalism, too.
518: As an example the improvement of the Energy Renormalization Group
519: was presented. Here the Kadanoff-Wilson blocking is performed
520: in energy space and the effects of the directions of the Hilbert
521: space lost by the truncation is retained. Therefore the dimension
522: of the linear space is reduced but the physics which can be described
523: by states within the reduced space remained the same. As long as the ground
524: state and the low lying excitations are kept in the linear spaces
525: constructed in this sequence the salient features of the model
526: can be described in a systematical and more economical manner.
527: 
528: The elimination of dimensions makes the eigenvalue equation
529: nonlinear in the eigenvalues, an effect which is well known in
530: many-body theory. In fact, say the self energy of a particle
531: receives a complicated, energy dependent contribution from
532: 'virtual', particle-number changing processes which leave
533: from and return to the one-particle sector of the Fock space.
534: We employed a widely
535: used approximation which becomes exact for the ground state
536: and the low lying excitations, the replacement of the
537: energy eigenvalue by the Fermi level in the self-energy.
538: The computational need of the resulting method is $\ord{N^2_>}$
539: with a prefactor which growth with the volume. Nevertheless
540: we find this result remarkable since systems with small gap can
541: safely be treated by exact diagonalization in a low dimensional
542: subspace.
543: 
544: We employed a further simplification of the effective Hamiltonian
545: in our numerical test. We replaced the part of the Hamiltonian
546: which belongs to the eliminated directions and appears in the self-energy
547: by a 'mean-operator' which is proportional to the identity. This
548: approximation is supposed to become exact for the ground state and
549: the low lying excitations of a Fermi-liquid. The method
550: is $\ord{N_>\ln^2N_>}$ when this simplification is used.
551: 
552: Our method was tested numerically in the case of the one dimensional
553: tight binding model. The ground state energy improved and a
554: reduction of its {\em error} by 25\% was found compared to the
555: original algorithm for $N=2048$.
556: 
557: The main question left open by the present work is the dependence
558: of the computational requirement on $N_<$, the physical size of the system,
559: and the explorations of alternative approximations which ultimately
560: speed up the algorithm in this respect.
561: 
562: 
563: 
564: 
565: 
566: \begin{acknowledgments}
567: We thank Jean Richert, Jacques Harthong, Xavier Blanc and Claude Demangeat for useful and
568: interesting discussions.
569: \end{acknowledgments}
570: 
571: 
572: \appendix
573: 
574: \section{One-particle density matrix}
575: This Appendix contains some details of the computation of the density
576: matrix for the tight binding model of Eq. \eq{hamnd} at half filling.
577: 
578: The matrix of eigenvectors corresponding to the $N$ lowest eigenvalues of $H$
579: is given by
580: \be
581: C_{\mu\nu}=\frac{1}{\sqrt{N+\frac{1}{2}}}\sin\bigg(\frac{\pi \mu \nu}{2N+1}\bigg).
582: \ee
583: for $1\le\mu\le2N$, and $1\le\nu\le N$. The reduced density matrix
584: \be
585: \rho=CC^{*}
586: \ee
587: is a projector with the diagonal matrix elements
588: \be
589: \rho_{\mu\mu}=\hf.
590: \ee
591: If $\mu$ and $\nu$ have same parities, i.e $\mu-\nu$ ia even then
592: $\rho_{\mu\nu}=0$. For $\nu=\mu+2k+1$
593: \be
594: \rho_{\mu\nu}=\frac{1}{4N+2}\left[\frac{(-1)^{k}}{\sin\frac{\pi}{2}\frac{2k+1}{2N+1}}
595: -\frac{(-1)^{\mu+k}}{\sin\frac{\pi}{2}\frac{\mu+k+\frac{1}{2}}{2N+1}}
596: \right].
597: \ee
598: 
599: In order to find rate of decrease of $\rho_{\mu\nu}$ we consider the
600: limit $N\to\infty$ but keep $\nu-\mu=2k$ fixed,
601: \be
602: \rho_{\mu,\mu+2k}\approx\frac{(-1)^{k}}{2k\pi}.
603: \ee
604: 
605: 
606: 
607: 
608: \begin{thebibliography}{99}
609: \bibitem{Wilson_1975}{K. G. Wilson, Rev. Mod. Phys. {\bf 47}, 773 (1975)}
610: \bibitem{BHG_1997_3}{R. Baer and M. Head-Gordon, J. Chem. Phys. {\bf 109}, 10159 (1998)}
611: \bibitem{BHG_1998}{R. Baer and M. Head-Gordon, Phys. Rev. B {\bf 58}, 15296 (1998)}
612: \bibitem{HF_1962}{H. Feshbach, Annals of Physics, {\bf 19}, 287-313 (1962)}
613: \bibitem{Rau_1996}{J. Rau, cond-mat/9607198, (1996)}
614: \bibitem{Muller_Rau_1996}{J. Muller, J. Rau, Phys. Lett. B 386 (1996) 274}
615: %\bibitem{Shankar_1994}{R. Shankar, Rev. Mod. Phys. {\bf 66}, 129 (1994)}
616: \bibitem{HK_1964}{P. Hohenberg and W. Kohn, Phys. Rev. B {\bf 136}, 864 (1964)}
617: \bibitem{K_1965}{W. Kohn and L. J. Sham, Phys. Rev. B {\bf 140}, 1133 (1965)}
618: \bibitem{Cap_2003}{K. Capelle, cond-mat/0211443, (2003)}
619: \bibitem{G_1999}{S. Goedecker, Rev. Mod. Phys. {\bf 71}, 1085 (1999)}
620: \bibitem{WJ_2002}{S. Y. Wu, C. S. Jayanthi, Phys. Rep. {\bf 358}, 1 (2002)}
621: \bibitem{K_1993}{W. Kohn, Chem. Phys. Lett. {\bf 208}, 167 (1993)}
622: \bibitem{JC_1964}{J. des Cloizeaux, Phys. Rev. {\bf 135}, (1964);
623: ibid (1964) A698}
624: \bibitem{K_1959}{W. Kohn, Phys. Rev. {\bf 115}, 809 (1959)}
625: \bibitem{K_1995}{W. Kohn, Int. J. Quart. Chem. {\bf 56}, 229 (1995)}
626: \bibitem{IB}{S. Ismail-Beigi, T. Arias, Phys. Rev. Lett. {\bf 82}, 2127 (1999)}
627: \bibitem{GI_1998}{S. Goedecker and O. Ivanov, Solid State Commun. {\bf 105}, 665 (1998)}
628: \bibitem{MYS}{S. Ismail-Beigi, T. Arias, Phys. Rev. Lett. {\bf 82}, 2127 (1999)}
629: \bibitem{Anderson}{W. Kohn, Phys. Rev. Lett. {\bf 21} (1968) 13}
630: \bibitem{K_1996}{W. Kohn, Phys. Rev. Lett. {\bf 76} (1996) 3168}
631: \bibitem{G_1994}{S. Goedecker and L. Colombo, Phys. Rev. Lett. {\bf 73} (1994) 122}
632: \bibitem{G_1998}{S. Goedecker, J. Comp. Phys. {\bf 118}, 261 (1998)}
633: \bibitem{G_1995}{S. Goedecker and M. Teter, Phys. Rev. B {\bf 51}, 9455 (1995)}
634: 
635: %\bibitem{VKS_1996}{A. Voter, J. Kress, and R. Silver,
636: %Phys. Rev. B {\bf 53}, 12733 (1996)}
637: %\bibitem{WN_1982}{S. R. White and R. M. Noack, Phys. Rev. Lett. {\bf 68} (1982) 3487}
638: %\bibitem{White_1993}{S. R. White, Phys. Rev. B {\bf 48}, 10345 (1993)}
639: \bibitem{BHG_1997_1}{R. Baer and M. Head-Gordon, Phys. Rev. Lett. {\bf 79} (1997) 3962}
640: \bibitem{BHG_1997_2}{R. Baer and M. Head-Gordon, J. Chem. Phys. {\bf 107}, 10003 (1997)}
641: \bibitem{GVL_1996}{G. H. Golub and C. F. Van Loan, \emph{Matrix computations},
642: 3rd ed., (John Hopkins University Press, Baltimore, 1996), p. 591}
643: \bibitem{Lar}{V. B. Larin, Technichskaya Kibernetika {\bf 1}, 210 (1992)}
644: \bibitem{NR}{W. H. Press, S. A. Teukolsky, W. T. Vetterling, and B. P. Flannery,
645: \em{Numerical Recipes in C++}, 2nd ed., 2001, p. 60}
646: \bibitem{Householder}{A. S. Householder, \emph{The theory of matrices in numerical analysis},
647: Dover Publications, Inc. New-York, 1975, p. 94}
648: \bibitem{PanReif} V. Pan, and J. Reif, \em{Proceedings of the Seventeenth
649: Annual ACM Symposium on Theory of Computing}, 1985
650: \bibitem{Atkins_Friedman}{P.W Atkins and R.S Friedman, \emph{Molecular
651: Quantum Mechanics}, 3rd ed., (Oxford University Press, 1997), p. 267}
652: \end{thebibliography}
653: 
654: \bibliography{apssamp}% Produces the bibliography via BibTeX.
655: 
656: \end{document}
657: