math0507445/hh13.tex
1: %Paper homogeneous - 29-07-02 
2: % Submitted on 29-07-02  
3: %Typeset with LaTeX format   
4: 
5: % Revision C&U 24-07-03
6: %Format for Proceedings   
7: % Shortened after mail of D.Larson, 3-8-03
8: 
9: %Preamble   
10: 
11: \documentclass{proc-l}   
12:    
13: \usepackage{amssymb,amscd,enumerate}   
14:    
15: % theorems, corollaries, lemmas, propositions   
16: % in the most emphatic (plain) style   
17: % definitions in the less emphatic (definitions) style   
18: % notations in the least emphatic (remark) style   
19: %\input{ma\os_l}   
20: 
21: %\textwidth 5.7in
22: %\evensidemargin .35in
23: %\oddsidemargin .35in
24: %\textheight 7.6in
25:    
26: \theoremstyle{plain}   
27: \newtheorem{theorem}{Theorem}   
28: \newtheorem{lemma}{Lemma}   
29: \newtheorem{proposition}{Proposition}   
30: \newtheorem{corollary}{Corollary}   
31:    
32: \theoremstyle{definition}   
33: \newtheorem{definition}{Definition}   
34:    
35: \theoremstyle{remark}   
36: \newtheorem*{remark}{Remark}   
37: \newtheorem*{note}{Note}   
38: \newtheorem*{notation}{Notation}   
39: \newtheorem*{conjecture}{Conjecture}   
40: \newtheorem*{observation}{Observation}   
41:    
42:    
43: \newcommand{\Lblock}[9]{\left[ \begin{array}{c|c|c}   
44: #1 & #2 & #3 \\   
45: %\hdotsfor{1} & \hdotsfor{1} & \hdotsfor{1} \\   
46: \hline   
47: #4 & #5 & #6 \\   
48: %\hdotsfor{1} & \hdotsfor{1} & \hdotsfor{1} \\   
49: \hline   
50: #7 & #8 & #9 \\   
51: \end{array} \right] } 
52: \newcommand{\homog}[2]{\hbox{\ensuremath{\mathcal{H}(2,#1,#2)}}}   
53: \newcommand{\dd}{\hbox{\ensuremath{\mathcal{D}_{1/2}}}}   
54: \newcommand{\HH}{\hbox{\ensuremath{\mathcal{H}}}}   
55: \newcommand{\BB}{\hbox{\ensuremath{\mathcal{B}}}}   
56: \newcommand{\sfi}{\hbox{\ensuremath{\mathcal{S}(\varphi)}}}   
57: \newcommand{\Sfi}{\hbox{\ensuremath{\mathcal{S}(\varphi)}}}   
58: \newcommand{\I}[2]{I^{#1}_{#2}}   
59: \newcommand{\N}{\hbox{\ensuremath{\mathbb{N}}}}   
60: \newcommand{\Z}{\hbox{\ensuremath{\mathbb{Z}}}}   
61: \newcommand{\Q}{\hbox{\ensuremath{\mathbb{Q}}}}   
62: \newcommand{\R}{\hbox{\ensuremath{\mathbb{R}}}}   
63: \newcommand{\C}{\hbox{\ensuremath{\mathbb{C}}}}   
64: %\newcommand{\nprec}{\hbox{\prec \!\!\!\!\! / \ }}   
65: \newcommand{\comment}[1]{}   
66:    
67: \DeclareMathOperator{\kker}{Ker} \DeclareMathOperator{\supp}{Supp}   
68:    
69: \numberwithin{equation}{section}   
70:    
71: \begin{document}   
72: % 2 authors   
73:    
74: %%%%%%%%%%%%%%%%%%%%%%%% PREAMBULO   
75:    
76: %%%%%%%%%%%%%A continuacion va el titulo   
77:    
78: \title
79: {Local Bases for Refinable Spaces}
80:    
81: % Author Info   
82:    
83: \author[C.Cabrelli]
84: {Carlos Cabrelli} 
85: \address{
86: Depto.~de   
87: Matem\'atica \\ FCEyN\\ Univ.~de Buenos Aires\\ Cdad.~Univ., Pab.~I\\ 1428 Capital   
88: Federal\\ ARGENTINA\\ and CONICET, Argentina}   
89: \email[%
90: %Carlos~A.~Cabrelli]
91: ]{cabrelli@dm.uba.ar, sheinek@dm.uba.ar, umolter@dm.uba.ar} 
92:    
93: \author[S.B.Heineken]{Sigrid B.~Heineken}   
94: %\email[Sigrid~B.~Heineken]{sheinek@dm.uba.ar}   
95:    
96: \author[U.M.Molter]{Ursula~M.~Molter}   
97: %\email[Ursula~M.~Molter]{umolter@dm.uba.ar}   
98:    
99: \thanks{The research of   
100: the authors is partially supported by Grants:   
101: CONICET, PIP456/98, and UBACyT X058 and X108. The authors also acknowledge partial support  from the Guggenheim Foundation and the Fulbright Commission during the period in which part of this research was performed.}   
102:    
103: %%%%%%%%%%%%%%%%%%% Aqui van los keywords y la subject class   
104:    
105: \keywords{Homogeneous functions, shift-invariant spaces, accuracy, refinable functions}  
106: \subjclass{Primary:39A10, 42C40, 41A15}   
107:    
108: \date{\today}   
109:    
110: %%%%%%%%%%%%% Abstract   
111:    
112: \begin{abstract}   
113: %texto del abstract aqui   
114: %(a   
115: %finite square matrix with entries from the mask of the   
116: %generator). 
117: We provide a new   
118: representation of a refinable shift invariant space with a compactly
119: supported generator, in terms of    
120: functions with a special property of homogeneity. In particular
121: these functions include all the homogeneous polynomials
122: that are reproducible by the generator, what links this representation
123: to the accuracy of the space. 
124: We completely characterize the class of homogeneous functions in
125: the space and show that they can reproduce the generator. 
126: As a result we conclude that the homogeneous functions can
127: be constructed from the vectors associated to
128: the spectrum of the scale matrix (a   
129: finite square matrix with entries from the mask of the   
130: generator).  Furthermore, we prove
131: that the kernel of the transition operator has
132: the same dimension than the kernel of this finite matrix.
133: This relation provides an easy test for the linear independence
134: of the integer translates of the generator.
135: This could be potentially   
136: useful in applications to approximation theory, wavelet theory and   
137: sampling.    
138: \end{abstract}   
139:    
140: \maketitle   
141:    
142: %%%%%%%%%%%%%%%%%%%%%%%%%% FIN PREAMBULO   
143:    
144: %%%%%%%%%%%%%% A partir de aqui va el paper   
145: \section{Introduction}   
146:    
147:    
148: A function  $\varphi: \R \longrightarrow \C$ is called {\em refinable} if    
149: it satisfies the equation:   
150: \begin{equation}\label{dilation-equation}   
151: \varphi(x) = \sum_{k=0}^N c_k \varphi(2x - k),    
152: \end{equation}   
153: for some complex scalars $c_0,...,c_N$.   
154: The scalars ${c_k}$ are the {\em mask} of the refinable function.   
155: We consider the case in which $\varphi$ is compactly supported.   
156: Define the {\em Shift Invariant Space} (SIS) generated by $\varphi$   
157: as:    
158: $$ {\mathcal S}(\varphi) = \{f: \R \longrightarrow \C: f(x) =   
159: \sum_{k\in \Z} y_k \varphi(x+k), y_k \in \C \}.   
160: $$   
161: A refinable SIS is a SIS with a refinable generator.   
162: Refinable SIS and refinable generators  have been studied extensively, since   
163: they are very important in Approximation Theory and Wavelet Theory.   
164:    
165: Many properties of  $\varphi$  can be obtained imposing conditions   
166:  on the mask.
167: One fundamental question is when the space $\sfi$ contains polynomials   
168:  and of which degree. The {\em accuracy} of $\varphi$ is the maximum integer $n$ such that   
169: all the polynomials of degree less or equal than $n-1$ are   
170: contained in $\sfi $.   
171:    
172: The accuracy is related to the approximation order of   
173: $\sfi$, (\cite{Jia95}, \cite{deB90} and references therein), 
174: and with the zero moments 
175:  and the  smoothness of the associated wavelet when $\varphi$ generates   
176:  a Multiresolution Analysis \cite{Mey92}.   
177: There are many well known equivalent conditions for accuracy.   
178: The one that interests us here is the following, 
179: \cite{Dau92}, \cite{CHM98}:   
180: \begin{proposition} \label{accuracy}  
181: Let $\varphi$ be a compactly supported function satisfying   
182: \eqref{dilation-equation}. Then 
183: $\varphi$ has accuracy $n$ if and only if  
184: $\{1,2^{-1},...,2^{-(n-1))}\}$ are   
185:  eigenvalues of the $(N+1)\times (N+1)$ matrix $T$ defined by  
186: $T = \{c_{2i-j}\}_{i,j=0,...,N}$, (the {\em scale matrix})  and   
187: there exist polynomials $p_0,...,p_{n-1}$ with degree$(p_i)=i$   
188: such that the vector $v_i= \{p_i(k)\}_{k=0,...,N}$     
189: is a left eigenvector of $T$ corresponding to the eigenvalue   
190: $2^{-i}$.    
191: \end{proposition}   
192: Here, and always throughout the paper, we assume $c_t = 0$, if $t \not= 0,\dots, N.$ 
193: One interesting property is that if $\varphi$ has accuracy $n$,   
194: then for $s=0,1,...,n-1$ it is true that   
195: %\begin{equation}   
196: $x^s = \sum_{k\in\Z} p_s(k) \varphi(x-k),   
197: $%\end{equation}   
198: where $p_s$ is the polynomial that provides the eigenvector   
199: for the eigenvalue $2^{-s}$.   
200: So, if the polynomial $x^s$ is in $\Sfi$, then the left eigenvector of $T$, 
201: corresponding to the eigenvalue $2^{-s}$, provides the coefficients needed
202: to write $x^s$ as a linear combination of the translates of $\varphi$.
203: 
204: A {\em local basis} for $\Sfi$ is a set of functions in $\Sfi$ whose restriction
205: to the $[0,1]$ form a basis for the space of
206: all the functions in $\Sfi$, restricted to the $[0,1]$.
207: 
208: When $\varphi$ is the B-spline of order $m$ (so $N=m$), then all
209: polynomials of degree less or equal than $m-1$ are in $\Sfi$. 
210: Moreover, the set $\{1, x, x^2, \dots, x^{m-1}\}$ is a local basis
211: for $\Sfi$, and the spectrum of $T$ consists exactly of
212: $\{1, 2^{-1}, \dots, 2^{-(m-1)}\}$.
213: 
214: Now, if $\varphi$ is not a B-spline then $T$ could have some
215: eigenvalue $\lambda$ different from a power of $1/2$.
216: If powers of $1/2$ are associated to homogeneous polynomials, which
217: functions in $\Sfi$ are associated to an arbitrary eigenvalue $\lambda$? 
218: Will the functions, associated to all the eigenvalues, provide also
219: local bases of $\Sfi$, or equivalently, will these functions reproduce
220: the generator $\varphi$? 
221: %If $\varphi$ has accuracy $p < N$, then no
222: %polynomials of degree bigger or equal than $p$ will be in $\Sfi$.
223: %Will the extra eigenvalues of $T$ tell us something about the
224: %order of approximation of $\Sfi$?
225: 
226: In the case that $\lambda$ is a simple eigenvalue, Blu and Unser
227: \cite{BU02} and later Zhou \cite{Zho02} showed 
228: that $\lambda$ is associated to what they call a $2$-scale $\lambda$-homogeneous
229: function, that is a function in the SIS that satisfies the relation
230: $h(x) = \lambda h(2x)$.
231: 
232: However, to obtain a complete representation of the space it is
233: necessary to consider the whole spectrum of $T$. 
234: This motivates the study of the spectral properties of $T$ for a {\em general}
235: refinable $\varphi$.
236: This is achieved in this paper: we are able to completely characterize the
237: SIS in terms of functions associated to the spectrum of $T$.
238: We prove that these functions provide a local basis of $\Sfi$ (c.f. Theorem~\ref{main}).
239: The advantage of this local basis is that it contains all possible
240: homogeneous polynomials in the space, and those functions in the
241: basis which are not polynomials, still preserve some
242: kind of homogeneity. Furthermore this basis can be easily
243: obtained from the spectrum of the finite matrix $T$. We are also able to 
244: prove that this matrix is necessarily invertible if the translates of $\varphi$
245: are linearly independent.
246: \begin{definition} \label{defi-hom}
247: Let $\lambda \in \C,  \lambda \not= 0$ and $r\geq 1$ an integer.   
248: A function $h$ is $(2,\lambda,r)$-{\em homogeneous} if it satisfies   
249: the following equation:   
250: \begin{equation} \label{homog-eq}  
251: \sum_{k=0}^r {r \choose k}   (- \lambda)^{r-k} h(2^{-k} x) = 0   
252: \;\; \textrm{a.e.}.   
253: \end{equation}  
254: $r$ is called the order of homogeneity,   
255: and $\lambda$ the degree.
256: \end{definition}
257: %For $r = 1$ these functions are called {\em two-scale homogeneous} and they satisfy
258: %$h(x) = \frac{1}{\lambda} h(\frac{x}{2}).$
259: 
260: If $\HH \subset \Sfi$ is the span of all the
261: $(2,\lambda,r)$-homogeneous functions in $\Sfi$, for any $\lambda \in \C$ and
262: any positive integer $r$, we show that under the hypothesis of linear independence
263: of the translates of $\varphi$, $\text{dim} (\HH) = N+1$, and that there is a
264: basis of $\HH$, corresponding to the spectrum of $T$. 
265: More precisely, given a basis $\BB = \{v_0, \dots, v_{N}\}$ of $\C^{N+1}$ that
266: yields the Jordan form of $T$ we associate to each vector $v \in \BB$ a unique
267: $(2,\lambda,r)$-homogeneous function in $S(\varphi)$, where $\lambda$ and $r$
268: satisfy   $v(T-\lambda I)^r = 0$.
269: 
270: The first $N$ of these functions are a local basis of functions 
271: in $S(\varphi)$ restricted to $[0,1]$.   
272: This allows to reconstruct   
273: the generator $\varphi$ from the homogeneous functions,
274: and gives a new representation for the functions in $S(\varphi)$.   
275: 
276: Furthermore, we show that to 
277: each non-zero vector in the kernel of $T$, there corresponds a non-trivial linear 
278: combination of the integer translates of $\varphi$ yielding the zero function. 
279:    
280: \section{Notation}  \label{notation} 
281: Let $\varphi: \R \longrightarrow \C$ be a function supported in $[0,N]$ 
282: satisfying \eqref{dilation-equation}.  
283:    
284: We will often use an infinite column vector associated to $\varphi$,   
285: namely    
286: \begin{equation}   
287: \left(\phi(x)\right)^t = \left[ \dots,\ \varphi(x-1),\ \varphi(x), \  
288: \varphi(x+1),\ \dots\ \right]. \label{vfi}   
289: \end{equation}   
290: Let $\ell(\Z)$ be the space of all the sequences defined in $\Z$.   
291: We say that the integer translates of $\varphi$ are {\em globally linearly independent}, 
292: or {\em linearly independent} if  
293: $\sum_{k\in \Z} \alpha_k \varphi(\cdot - k) \equiv 0 \quad \Longrightarrow \quad \alpha_k = 0\   
294: \forall k$, 
295: for any sequence $\alpha \in \ell(\Z)$.  
296:  
297: The {\em subdivision operator} associated to the mask $c_k$ is   
298: the operator   
299: \begin{equation} 
300: \label{subdiv}   
301:  S_c :\ell(\Z)\rightarrow \ell(\Z)   
302: \quad \text{defined by} \quad   
303: S_c(\alpha)_j = \sum_{i \in \Z} \alpha_i c_{2i-j}.   
304: \end{equation} 
305: 
306: \begin{note} The subdivision operator is sometimes defined in a different but equivalent  
307: way as $\tilde{S}_c(\alpha)_j = \sum_{i\in \Z} \alpha_i c_{j-2i}$. If $h: \ell(\Z)
308: \rightarrow \ell(\Z)$ is the 
309: operator $h(\alpha)_k = \alpha_{-k}$, then $h\tilde{S}_c h = S_c$ and
310: therefore  $S_c$ and $\tilde{S}_c$ 
311: share most of the properties. 
312: For a nice account of properties of the subdivision operator
313: see \cite{BJ02}. \end{note} 
314:    
315: If $L = L_{\varphi}$ is the double infinite matrix   
316: $L = \left[c_{2i-j}\right]_{i,j \in \Z}$,   
317: then the refinement equation can be written as   
318: $\phi(x) = L \phi(2x)$.   
319:  
320: Using the matrix $L$, the subdivision operator \eqref{subdiv} can be recast as: 
321: $S_c \alpha = \alpha L,\ \alpha \in \ell(\Z),$ 
322: where $\alpha$ on the right hand side of the equation is thought as an infinite row vector. 
323: Note that the scaling matrix $T$ defined in Proposition~\ref{accuracy} 
324: is a finite submatrix of $L$. 
325: We will consider in our analysis the matrices $M, T_0, T_1$, that are  
326: submatrices of $T$ and are defined as: 
327: $M=[c_{2i-j}]_{i,j=1,...,N-1},\ T_0=[c_{2i-j}]_{i,j=0,...,N-1},\ 
328: T_1=[c_{2i-j}]_{i,j=1,...,N}$. That is, 
329: \begin{equation} \label{Tone}   
330: T = %\begin{bmatrix} c_0 & 0 & \hdotsfor{3} & 0 \\   
331:      %               c_2 & c_1 &  c_0 & 0  & \hdotsfor{1} & 0\\   
332:      %               \vdots & c_3 & c_2 & c_1 & \hdots & \vdots \\   
333:      %               0 & \vdots & \vdots & \vdots & \vdots & c_{N-2}\\   
334:      %               0 & \hdotsfor{3} &  0 & c_{N}   
335: %\end{bmatrix} =    
336: \left[ \begin{array}{c|c|c}   
337: c_0 & 0 & 0 \\   
338: \hline    
339: \vdots & M & \vdots\\   
340: \hline    
341: 0 & 0 & c_N    
342: \end{array} \right] .   
343: \end{equation}   
344:  Note that $c_0$ and $c_N$ must be non-zero, 
345: since $\supp(\varphi) = [0,N]$. 
346:  
347:    
348: Now, if $Y \in \ell(\Z)$,   
349: define $Y^0$ and $Y^M$ as the restriction of $Y$ to   
350: the indexes $\{0, \dots, N\}$,   
351: and   
352: $\{1, \dots, N-1\}$, respectively, i.e.,    
353: $$Y^0 = (Y_0, \dots, Y_{N}), \quad Y^M = (Y_1, \dots, Y_{N-1}). $$
354: \begin{note}   
355: Throughout this paper, $(L - \lambda I)$ is considered as an operator on   
356: $\ell(\Z)$, defined by left-multiplication, (i.e. $ Y \longmapsto   
357: Y(L-\lambda I)$, where $Y$ is a double infinite row vector).   
358: $I$ is the identity operator acting on $\ell(\Z)$. By an abuse of   
359: notation, we will use   
360: the notation $I$ for all identity operators, without   
361: distinguishing the space they are acting on. 
362: %Note also that properties of the matrix $L$ 
363: %translate directly into properties of the subdivision operator $S_c$. 
364: \end{note}   
365:    
366: \section{The point spectrum of $L$}   
367:   
368: The following proposition, will show, how the spectral properties of $L$    
369: are related to those of $T$.  The case $r=1$ has been studied earlier by \cite{CHM00}, 
370: \cite{JRZ98}, \cite{Zho00,Zho02}.  
371: \begin{proposition} \label{I}   
372: Let $\lambda \in \C.$   
373: \begin{enumerate}   
374: \item \label{p1}   
375: Let $Y \in \ell(\Z)$ and $r \in \N$, $r \geq 1$. 
376: If $Y \in \kker(L - \lambda I)^r$, then $   
377: Y^0 \in \kker(T - \lambda I)^r.$    
378: Moreover, if    
379: $\lambda \not= 0$, $Y \not= 0$ and $Y \in \kker(L - \lambda I)^r$, then   
380: $Y^0 \not= 0$.   
381: \item If $v \in \kker(T - \lambda I)^r$ and $\lambda \not= 0$,   
382: then there exists an extension $Y_v \in \ell(\Z)$ of $v$,   
383: (i.e. $Y_v^0 = v$) such that $Y_v \in \kker(L - \lambda I)^r$. \label{p2}   
384: \end{enumerate}   
385: \end{proposition}   
386: \begin{proof}   
387: The matrix $L$ (and therefore $L-\lambda I$) can be decomposed in blocks as   
388: \begin{equation} \label{L-blocks}   
389: L = \Lblock{R}{0}{0}{P}{T}{Q}{0}{0}{S},\quad 
390: (L - \lambda I) = \Lblock{R - \lambda I}{0}{0}{P}{T - \lambda I}{Q}%   
391: {0}{0}{S -\lambda I},   
392: \end{equation}   
393: where we decompose $\Z$ as   
394: $\Z =  A^{-}  \cup  A^0    \cup  A^{+},$    
395: with   
396: $A^{-} = \Z  \cap (-\infty,-1]$, $A^0 = \Z  \cap  [0,N]$ and   
397: $A^{+} = \Z  \cap [N+1, +\infty)$,   
398: and   
399: \begin{equation*}   
400: R = L|_{A^-\times A^-} \quad P = L|_{A^0\times A^-} \quad T =   
401: L|_{A^0\times A^0} \quad Q = L|_{A^0\times A^+} \quad S =   
402: L|_{A^+\times A^+}.   
403: \end{equation*}   
404: This block form of the matrix, is closed under multiplication. So if $r \geq 1,   
405: r \in \N$   
406: \begin{equation} \label{PQ}   
407: L^r = \Lblock{R^r}{0}{0}{P_r}{T^r}{Q_r}{0}{0}{S^r}, \quad \text{and} \quad   
408: (L - \lambda I)^r = \Lblock{(R - \lambda I)^r}{0}{0}%   
409: {P_r^{\lambda}}{(T - \lambda I)^r}{Q_r^{\lambda}}%   
410: {0}{0}{(S -\lambda I)^r} ,   
411: \end{equation}   
412: where   
413: $P_r = \sum_{k=0}^{r-1} T^k P R^{r-k-1}$ and $
414: Q_r = \sum_{k=0}^{r-1} T^k Q S^{r-k-1},$   
415: and $P_r^{\lambda}$ and $Q_r^{\lambda}$ are analogous, with   
416: the obvious changes.    
417:    
418: Note that the matrix $S$ is upper triangular,    
419: with diagonal $(0, 0, 0, \dots)$, and hence    
420: $(S - \lambda I)^r$ is upper triangular, with diagonal   
421: $((-\lambda)^r, (-\lambda)^r, (-\lambda^r), \dots)$.   
422:    
423: Analogously, we observe that $R$ is lower triangular with zeroes in the    
424: main diagonal, so  $(R -\lambda I)^r$ is   
425: lower triangular with diagonal    
426: $((-\lambda)^r,(-\lambda)^r,(-\lambda)^r,\dots)$.   
427:    
428:    
429: If $Y = (Y^-,Y^0,Y^+)$, then   
430: $Y(L - \lambda I)^r$ can be written as
431: \begin{equation}    
432: (Y^-(R - \lambda I)^r + Y^0 P_r^{\lambda},    
433: Y^0(T - \lambda I)^r,   
434: Y^0 Q_r^{\lambda} + Y^+(S - \lambda I)^r). 
435: \label{ydescomp}   
436: \end{equation}   
437: So if $Y \in \kker(L - \lambda I)^r$, then $Y^0 \in \kker(T - \lambda I)^r$.   
438:    
439: We now want to show that if $Y \in \kker(L - \lambda I)^r$,    
440: $\lambda \not= 0$, $Y \not= 0$, then $Y^0 \not= 0$.    
441:    
442: For this, let $k_0 \in \Z$    
443: be such that $Y_{k_0} \not= 0$. If $0 \leq k_0 \leq N$, we are   
444: done. Assume that $k_0 > N$. Then, since   
445: $Y(L-\lambda I)^r = 0$, in particular, the $k_0$ element of this   
446: product is $0$.    
447: But since $\lambda \not=0$, $(S - \lambda I)^r$ is   
448: upper triangular with $(-\lambda)^r$ in the diagonal,    
449: therefore the only nonzero elements of   
450:  column $k_0$ of $(L - \lambda I)^r$ are between $0$ and $k_0$.   
451: Hence there has to be a $k_1, 0 \leq k_1 < k_0$ such that   
452: $Y_{k_1} \not= 0$. Again, if $0 \leq k_1 \leq N$ we are   
453: done, otherwise we repeat the argument until $k_j$ is in the    
454: desired interval.   
455: If $k_0 < 0$, the argument works in the same way, reversing   
456: the role of $(S-\lambda I)^r$ and $(R - \lambda I)^r$.   
457:    
458: For the proof of part~\ref{p2}, assume that $v \in \C^{N+1}$,    
459: $v \in \kker(T - \lambda I)^r$.    
460: We want to find an infinite  vector $Y \in \ell (\Z)$, such that    
461: $Y^0 = v$ and $Y \in \kker(L - \lambda I)^r$.   
462: From equation \eqref{ydescomp} we know that if $Y \in \ell (\Z)$,   
463: \begin{equation*}   
464: \left[Y (L - \lambda I)^r \right]^+  =  
465:  Y^0 Q_r^{\lambda} + Y^+(S-\lambda I)^r,\quad   
466: \left[Y (L - \lambda I)^r \right]^-  =  Y^0 P_r^{\lambda} + Y^-(R-\lambda I)^r.   
467: \end{equation*}   
468: Therefore, if $ Y \in \kker(L - \lambda I)^r$, and $Y^0 = v$, then $Y^+$ and $Y^-$ have 
469: to satisfy   
470: \begin{equation}\label{4*}   
471: Y^+(S-\lambda I)^r = -v Q_r^{\lambda}  
472: \quad \text{and} \quad 
473: Y^-(R-\lambda I)^r = -v P_r^{\lambda} .   
474: \end{equation}   
475: Using again that $(S - \lambda I)^r$ and $(R - \lambda I)^r$, are triangular, if   
476: $\lambda \not= 0$, there are unique solutions for $Y^+$ and $Y^-$ and they can be   
477: obtained recursively.   
478: \end{proof}   
479: The last proposition tells us that the elements of the spectrum of   
480: $T$ are intimately related to those of $L$. But by the special form   
481: of $T$ (see equation \eqref{Tone}), we can actually use the $(N-1) \times
482: (N-1)$ matrix $M$ to obtain the spectrum of $T$, as the following proposition
483: shows:
484: \begin{proposition} \label{IIbis}   
485: Let $\lambda \not=0 \in \C$.    
486: \begin{enumerate}   
487: \item Let $v^0 = (v_0,\dots, v_N) \in \C^{N+1}$ and $r \in \N$, $r \geq 1$.    
488: Then,   
489: if $v^0 \in \kker(T - \lambda I)^r$ then $v^M = (v_1, \dots, v_{N-1})   
490:  \in \kker(M - \lambda I)^r$.    
491: Moreover, if $\lambda \not= c_0$, $\lambda \not= c_N$,   
492: $v^0 \in \kker(T - \lambda I)^r$, and   
493: $v^0 \not= 0$, then   
494: $v^M \not= 0$.   
495: \item if $v^M = (v_1,\dots,v_{N-1}) \in \kker(M - \lambda I)^r$ and    
496: $\lambda \not= c_0$ and $\lambda \not= c_N$,   
497: then there exists an extension $v^0 \in \C^{N+1}$ of $v$,   
498: such that $v^0 \in \kker(T - \lambda I)^r$.   
499: \end{enumerate}   
500: \end{proposition}   
501: \begin{proof}   
502: The proof is immediate by noting the special block-form of $T$. 
503: \end{proof}   
504: 
505: \section{The kernel of $L$}
506:    
507: The case $\lambda = 0$ could not be handled with the   
508: methods of Proposition~\ref{I},   
509: since the matrices $R$ and $S$ in \eqref{L-blocks} have zeros in   
510: the main diagonal. Instead,    
511:  we need some results from   
512: the theory of difference equations which we present below \cite{Hen62}.   
513:    
514: \subsection{Difference Equations}   
515:    
516: Consider the linear difference equation with constant coefficients of order $r$   
517: \begin{equation}   
518: u_0y_n + u_1 y_{n+1} + \dots + u_r y_{n+r} = 0  \quad y = \{y_n\}_{n \in \Z},   
519: \label{diff}   
520: \end{equation}   
521: where $u_k \in \C, u_0 \not=0, u_r \not= 0$ with characteristic polynomial   
522: %\begin{equation*}   
523: $P(x) = \sum_{k=0}^r u_{k} x^{k}.   
524: $%\end{equation*}
525:    
526: A {\em solution} to the equation~\eqref{diff} is  a sequence $Y$ in $\ell(\Z)$, that 
527: satisfies \eqref{diff}   for all $k \in \Z$. A vector $y=(y_0, \dots, y_m)$ with 
528: $m \geq r+1$ is a {\em finite solution} of \eqref{diff}, if it satisfies 
529: \eqref{diff} for $n=0$ to $n=m-r-1$. 
530:  
531: The {\em space of solutions}  $S \subset \ell(\Z)$, has dimension $r$, and a basis   
532: of this space (the fundamental basis) can be written in the following way:   
533: 
534: Let $h \geq 1$ be an integer, $d_1,\dots,d_h$ arbitrary non-zero complex numbers 
535: with $d_i \not= d_j$ if $i \not= j$. Let $r_1,\dots,r_h$ be positive integers. To each pair 
536: $(d_i,r_i)$, $i=1,\dots,h$ we will associate a sequence $a_i = \{a_{ik}\}_{k\in \Z}$ 
537: defined as follows: 
538: %\begin{equation*}   
539: $\text{Set} \quad r = r_1+\dots+r_h \quad \text{and}\quad r_0 = 0.   
540: $\ Let $0 \leq i \leq r-1$ and $s = s(i), j = j(i)$ be the unique integers that   
541: satisfy   
542: $%\begin{equation*}   
543: r_0+\dots+r_{s-1} \leq i < r_1+\dots+r_s,\ j(i) = i - \sum_{k=0}^{s(i)-1}r_k$.   
544: %\end{equation*}   
545: \begin{equation}\label{aik}   
546: \text{Define} \quad a_{ik} =    
547: \begin{cases}   
548: \text{sg}(k)\frac{|k|!}{(|k|-j(i))!}\, d^k_{s(i)} & \text{for}\ |k|  \geq j(i)\\   
549: 0 & |k| < j(i).   
550: \end{cases} \quad i=0,\dots,r-1, \quad k \in \Z,
551: \end{equation}   
552: where $\text{sg}(k)$ is the sign of $k$.
553: 
554: So, if $P$ is the characteristic polynomial associated to~\eqref{diff}, 
555: consider the pairs $\{(d_i,r_i): \text{where}\ d_i\ \- \text{is a root of $P$ and $r_i$  
556: its multiplicity} \}$. 
557: The sequences $\{a_{ik}\}_{k \in \Z}$, $i = 0,\dots,r-1$ form a basis   
558: of $S$, the subspace of $\ell(\Z)$ of the solutions to \eqref{diff}.   
559:  
560: It is also known from the theory of difference equations, that every solution is 
561: determined unequivocally by any $r$ consecutive elements of it. Hence, if $y$ is a solution 
562: such that $r$ consecutive elements are $0$, then $y$ is the zero solution. 
563:  
564: We will now associate to the pairs $\{(d_i,r_i): i = 1,\dots,h\}$ the $r\times r$ 
565: matrix $A = [a_{ij}]_{i,j = 0, \dots, r-1}$. Then (cf. Henrici \cite{Hen62}, pg.~214)  
566: \begin{equation} \label{deter}   
567: \det(A) = \prod_{1 \leq l < s \leq h} (d_l - d_s)^{r_l + r_s}   
568: \prod_{i=1}^h (r_i - 1)!! ,   
569: \end{equation}   
570: where $0!! = 1$ and $k!! = k!(k-1)!\dots 1!$.   
571: Since $d_i \not= d_j$ for $i\not=j$, $\det(A) \not=0$ and $A$ is invertible.  
572:  
573: Let us now 
574: consider a system of $k$ linear difference equations with constant coefficients of order $r$.   
575: \begin{equation} \label{sys-diff}   
576: u_{i0}y_n + \dots + u_{ir} y_{n+r} = 0 
577: , \quad i = 1, \dots k, \quad n \in \Z,   
578: \end{equation}   
579: and let $P_i$ be the characteristic polynomial of equation $i$, 
580: $P_i(x) = \sum_{j=0}^r u_{ij}x^j$. Define $
581: \mathcal{D} = \cup_{i=1}^k \{d : P_i(d) = 0\} = \{d_1,\dots,d_s\}$, 
582: and for each $d \in \mathcal{D}$ define 
583: $r_d = max\{r_i: r_i \text{ is the multiplicity of $d$ in $P_i$}\}$. 
584: Note that $r_d \geq 1\ \forall d \in \mathcal{D}$. We then have the pairs 
585: $(d_i,r_{d_i}) = (d_i,r_i)$. Define the index of the system 
586: to be  $t = \sum_{d\in \mathcal{D}} r_d \leq kr$. 
587: Let $\ell$ be the degree of the maximum common divisor $p$, of 
588: $\{P_a,\dots,P_k\}$. Hence, $P_i(x) = p(x) \tilde{P}_i(x)$, with degree $\tilde{P}_i = r-\ell$. 
589: (Note that $\ell$ could be $0$). 
590: With the above notation, we have the following proposition:
591: \begin{proposition}\label{pdiff}   
592: The space $S_k$ of solutions to the system~\eqref{sys-diff} 
593: has dimension $\ell$, where $\ell$ is the degree of the maximum common divisor of 
594: the characteristic polynomials. Moreover, if $t$ is the index of the system~\eqref{%
595: sys-diff}, and
596: $z$ is a vector of length $t$ that satisfies~\eqref{sys-diff}, then it can be 
597: extended to a sequence $y_{z} = \{y_j\}_{j\in \Z}$ solution  of \eqref{sys-diff} and such that 
598: $y_j = z_j, \,\, j=1,\dots,t$. 
599: \end{proposition}   
600: \begin{proof}   
601: Let $p$ be the maximum common divisor of $P_1,\dots,P_k$, and let $\ell = \deg(p)$. 
602: It is clear, that $\dim(S_k) \geq \ell$.  
603:  
604: For the other inequality,  
605: consider the $t \times t$ matrix $A = [a_{ij}]_{i,j = 0, \dots, t-1}$,   
606: with $a_{ij}$ defined in \eqref{aik} for the pairs 
607: $\{(d_i,r_i)\}$ defined above and $t$ being the index of the 
608: system~\eqref{sys-diff}.  
609: Since $d_i \not= d_j$ for $i\not=j$, $\det(A) \not=0$ by  
610: \eqref{deter} and $A$ is invertible.  
611:  
612: Assume now that $y \in S_k$, then $y$ is a solution to all $k$ difference equations, 
613: hence there exist 
614: $\alpha^1, \dots, \alpha^k$ vectors of length $r$, such that 
615: \begin{equation} 
616: A_i \alpha^i = \left[
617: y_0, \dots, y_{t-1}\right]^t  
618: \quad 1\leq i \leq k, 
619: \end{equation} 
620: where $A^i$ is an $t \times r$ matrix whose columns are a fundamental system for 
621: equation $i$. Note that $A^i$ is a sub-matrix of $A$, whose columns correspond 
622: to some columns $\{i_1,\dots,i_r\}$ of $A$. 
623:  
624: Let now $\tilde{\alpha}^i$ be vectors of length $t$, such that $\tilde{\alpha}^i_h = 0$ 
625: whenever $h \not\in \{i_1,\dots,i_r\}$ and $\tilde{\alpha}^i_{i_s} = \alpha_s$, 
626: $s = 1, \dots, r$. Then we have for $i, j = 1,\dots,k$ 
627: \begin{equation} 
628: A \tilde{\alpha}^i = A_i \alpha^i = \left[ 
629: y_0, \dots, y_{t-1}\right]^t = A_j \alpha^j =  A \tilde{\alpha}^j, 
630: \end{equation} 
631: and hence  
632: $A (\tilde{\alpha}^i - \tilde{\alpha}^j) = 0$, for all $i \not= j$ 
633: and therefore, by the invertibility of $A$, $\tilde{\alpha}^i = \tilde{\alpha}^j$, for all $i \not= j$. Therefore the only non-zero elements of $\alpha_i$ can be those corresponding 
634: to the columns associated to the roots of $p$. Hence $y$ is a linear combination 
635: of $\ell$ columns, and therefore $\dim(S_k) \leq \ell$. 
636: \end{proof}   
637:  
638: By noting that for the previous proof, we only used the 
639: first $t$ coordinates of the infinite sequences, we have the following immediate Corollary.  
640: \begin{corollary} 
641: \label{pp2} If $z$ is a vector of length $t$ that satisfies~\eqref{sys-diff}, then it can be 
642: extended to a sequence $y_{z} = \{y_j\}_{j\in \Z}$ solution  of \eqref{sys-diff} and such that 
643: $y_j = z_j, \,\, j=1,\dots,t$. 
644: \end{corollary}   
645: \subsection{The $\kker(L)$}   
646:    
647: We can now return to our double infinite matrix $L$ and look at the special case $\lambda = 0$.    
648: As it turns   
649: out, the kernel of $L$ is characterized by the vectors in the   
650: kernel of $M$. Since $c_0$ and $c_N$ are non-zero, the matrices   
651: $T$ and $M$ have kernels of the same dimension.   
652: Moreover, we have the following Proposition: 
653: \begin{proposition} \label{II} 
654: Consider the polynomials $p_e$ and $p_o$  
655: of degree $q = \frac{N-1}{2}$ (we assume $N$ to be odd)   
656: $ p_e(x) = c_0 + c_2 x + \dots + c_{2q} x^q, \ 
657: p_o(x) = c_1 + c_3 x + \dots + c_{2q+1} x^q.$ 
658: Then $\dim(\kker(L)) = \dim(\kker(M)) = \text{\rm degree}(p)$, 
659: where $p$ is the  
660: maximum common divisor of the polynomials $p_e$ and $p_o$. 
661: In particular, if $\dim(\kker(M)) >0$, $p_e$ and $p_o$ 
662: have a common root. Furthermore 
663: \begin{enumerate} 
664: \item \label{one} 
665: For every $Y \in \kker(L)$, $Y \not= 0$, we have $Y^M \not=0$ 
666: and $Y^M \in \kker(M)$. 
667: \item \label{two} 
668: Conversely, for each $v \in \kker(M), v \not= 0$, we have 
669: $Y_v \not= 0$ and $Y_v \in \kker(L)$.   
670: \end{enumerate} 
671: \end{proposition}   
672: \begin{proof}   
673: Let us observe first, that $Y \in \ell(\Z)$ is in the Kernel of $L$, if and 
674: only if $Y$ satisfies the system of difference equations: 
675: \begin{equation} \label{eod} 
676: \left\{ 
677: \begin{array}{ll} 
678: c_0v_n + c_2 v_{n+1} + \dots + c_{2q} v_{n+q} & = 0  \\   
679: c_1v_n + c_3 v_{n+1} + \dots + c_{2q+1} v_{n+q} & = 0. 
680: \end{array} 
681: \right. 
682: \end{equation} 
683: Therefore, by Proposition~\ref{pdiff}, $\kker(L)$ is the subspace generated 
684: by the fundamental solutions associated to the roots of $p$, the maximum 
685: common divisor of $p_o$ and $p_e$. This shows that $\dim(\kker(L)) = \text{degree}(p)$. 
686:  
687: On the other side, if $Y \in \kker(L)$, since $(YL)^M = Y^M M$ we conclude 
688: that $Y^M \in \kker(M)$, and if $Y^M$ is the zero vector, then the solution 
689: $Y$ of \eqref{eod} has $N-1$ consecutive zeros, so $Y = 0$. 
690: Hence, if $Y \not= 0$, then $Y^M \not= 0$, which proves~(\ref{one}). 
691:     
692: To see that if $v=(v_1,\dots,v_{N-1})$ satisfies $vM = 0$, then $v$ can be extended, just 
693: note that the   
694: sequence $v_1, \dots, v_{N-1}$ must satisfy the difference equations system  
695: of order $q = \frac{N-1}{2}$ (we assumed $N$ to be odd) given by \eqref{eod}. 
696: Since the index $t$ of the system~\ref{eod} satisfies $t \leq 2q = N-1$,  
697: and $v$ is a non-trivial common solution of length $N-1$, by Corollary~\ref{pp2},  
698: this solution can   
699: be extended in such a way that the extension satisfies both difference equations. 
700: This proves~(\ref{two}). 
701:  
702: From~(\ref{one})~and~(\ref{two}) it is immediate that $\dim(\kker(L)) = \dim(\kker(M))$.   
703: \end{proof}   
704: 
705: \begin{note} The fact that $\dim(\kker(M)) >0$ implies that $p_e$ and $p_o$ 
706: have a common root was proved under some minor technical conditions by Meyer \cite{Mey91}. Related results can also be found in \cite{JW93}.
707: \end{note}
708: 
709: \subsection{Invertibility of $L$}   
710:    
711: Propositions~\ref{I} and \ref{IIbis}, relate the spectral properties   
712: of the matrix $M$ to the ones of the operator $L$. The next proposition   
713: shows a necessary condition for the independence of the integer translates of   
714: the function $\varphi$, in terms of the matrix $M$.   
715: \begin{proposition} \label{prop6}   
716: With the above notation, consider the following properties   
717: \begin{enumerate}   
718: \item $\{\varphi(\cdot - k)\}_{k\in \Z}$ are globally linearly independent, \label{iI}   
719: \item The operator $L:\ell(\Z) \longrightarrow \ell(\Z),   
720: Y \longmapsto YL$ is one-to-one, \label{iII}   
721: \item The matrix $M$ is invertible. \label{iIII}   
722: \end{enumerate}   
723: Then (\ref{iII} $\Longleftrightarrow$ \ref{iIII}) and (\ref{iI}    
724: $\Longrightarrow$ \ref{iII}).   
725: \end{proposition}   
726: \begin{proof}   
727: {\bf (\ref{iI} $\Longrightarrow$ \ref{iII})}   
728: Assume $YL = 0$. Define $F(x) = Y \phi(x)$. Then we have   
729: $F(x) = Y\phi(x) = YL\phi(2x) = 0.$
730: Now, $Y\phi(x) = 0  \Longrightarrow Y = 0$, therefore $\kker(L) = \{0\}$.   
731:    
732: \noindent   
733: {\bf (\ref{iII} $\Longleftrightarrow$ \ref{iIII})}   
734: is a consequence of Proposition~\ref{II}.    
735: \end{proof}   
736:    
737: \noindent   
738: {\bf Note:} We do not know if either   
739:  (\ref{iII}) or   
740: (\ref{iIII}) implies (\ref{iI}).    
741:    
742: \section{Homogeneous functions}   
743:    
744: Assume now that $Y \in \kker(L-\lambda I)^r$, and define the function $
745: h \in \Sfi$ as
746: $h(x) = Y\phi(x)$. So $h$ is $(2, \lambda,r)$ homogeneous (c.f.~\eqref{homog-eq}),
747: since it satisfies:   
748: \begin{align*}   
749: 0 &= Y (L - \lambda I)^r\phi(x) =    
750: Y \left(\sum_{k=0}^r \begin{pmatrix}r\\k   
751: \end{pmatrix} (-\lambda)^{k} L^{r-k}\right) \phi(x) \\   
752:  &= Y \left(\sum_{k=0}^r \begin{pmatrix}r\\k\end{pmatrix}   
753:  (-\lambda)^{k} \phi(\frac{x}{2^{r-k}})\right)   
754:  = \sum_{k=0}^r \begin{pmatrix}r\\k\end{pmatrix}   
755:  (-\lambda)^{k} h(\frac{x}{2^{r-k}}). 
756: \end{align*}   
757: We will denote by   
758: \homog{\lambda}{r}, the space of all $(2,\lambda,r)$ homogeneous   
759: functions.   
760: \begin{remark}[1]   As pointed out by the referee, a $(2,\lambda,r)$-homogeneous function, is a particular case of a poly-scale refinable distribution. The concept of poly-scale refinable distribution is a generalization of refinability. See for instance \cite{DD02, Sun05}.
761: \end{remark}
762: \begin{remark}[2]   
763: Note that if    
764: $h \in \homog{\lambda}{r}$  then   
765: $h \in \homog{\lambda}{s}$ for every $s \geq r$.  
766: Therefore the ``order of homogeneity'' will be defined by    
767: $\min\{s: h \in \homog{\lambda}{s}\}$.   
768: \end{remark}    
769: \begin{remark}[3] If $h$ is homogeneous (of any order) and $\lambda \not= 1$, then $h(0) = 0$.    
770: The values of any homogeneous function of order $r$   
771: in $(0,+\infty)$, are completely determined   
772: by its values on any interval of the type    
773: $\left[\frac{1}{2^{k+r}}, \frac{1}{2^k}\right)$,   
774: $k \in \Z$.    
775: (Analogously, the values on $(-\infty,0)$, are obtained from the values   
776: in any interval of the type   
777: $\left(-\frac{1}{2^{k}}, -\frac{1}{2^{k+r}}\right]$).   
778: \end{remark}   
779: \begin{remark}[4]   
780: In the case of order of homogeneity $1$, (e.g. $r=1$), $h$ is a $2$-scale   
781: homogeneous function as described  in \cite{Zho02}.   
782: \end{remark}   
783: \begin{proposition}   
784: Assume $\{\varphi(\cdot-k)\}$ are linearly independent. Let   
785: $\phi$ be as in \eqref{vfi}. If $g_1, \dots, g_n \in \sfi$,   
786: $g_i = Y^i \phi$, then   
787: $\{g_1, \dots, g_n\}$ are linearly independent functions   
788: if and only if 
789: $ \{Y^1, \dots, Y^n\}$ are linearly independent in $\ell(\Z)$.   
790: \end{proposition}   
791: \begin{proof}   
792: We observe that   
793: $\sum_{i=1}^n \alpha_i g_i = \sum_{i=1}^n \alpha_i\left(Y^i \phi\right) =   
794: \left(\sum_{i=1}^n \alpha_i Y^i \right)\phi.$    
795: This equation, together with the linear independence of the translates   
796: of $\varphi$, tells us that   
797: $\sum_i \alpha_i g_i \equiv 0$ if and only if    
798: $\left(\sum_i \alpha_i Y^i \right)= 0$, which proves the desired result.   
799: \end{proof}   
800: \begin{theorem}\label{tone}   
801: Assume that $\{\varphi(\cdot-k)\}$ are linearly independent. If $h = Y\phi$,   
802: ($ h \in \sfi$), and $h \in \homog{\lambda}{r}$   
803: then $v_h = Y^0 \in \kker(T-\lambda I)^r$. Reciprocally, if   
804: $v \in \kker(T-\lambda I)^r$, then the function $h = Y_v \phi$ is   
805: in \homog{\lambda}{r}. (Here $Y_v$ is the unique extension   
806: of $v$ to a vector in $\kker(L - \lambda I)^r$ by Prop.~\ref{I}.)   
807: \end{theorem}   
808: \begin{proof}   
809: For the first claim, note that   
810: \begin{equation*}   
811: 0  = \sum_{k=0}^r \begin{pmatrix}r\\k\end{pmatrix} (-\lambda)^{r-k} h(2^{-k} x)   
812:  = \sum_{k=0}^r \begin{pmatrix}r\\k\end{pmatrix} (-\lambda)^{r-k}%   
813: YL^k \phi(x)\\   
814:  = Y \left(L - \lambda I\right)^r \phi(x).   
815: \end{equation*}   
816: Then, $ Y \left(L - \lambda I\right)^r=0$ and by 
817:  Proposition~\ref{I}, $v_h \in \kker(T- \lambda I)^r$.   
818:    
819: For the converse first observe that if $v=0$ the result is trivial. Assume 
820: $v\not= 0$ and $v \in \kker(T- \lambda I)^r$, then by   
821: Proposition~\ref{prop6} $\lambda \not= 0$. Hence (by Prop.~\ref{I}) there is a unique extension 
822:  $Y_v \in \kker(L- \lambda I)^r$, so   
823: $h = Y_v \phi$ is in \homog{\lambda}{r}.   
824: \end{proof}    
825:    
826: \subsection{Local basis of homogeneous functions}   
827: \begin{theorem}   
828: Let $\Lambda$ be the set of eigenvalues of $T$, and    
829: let $\mathcal{B} = \{v_0,\dots,v_N\}$  be a basis of    
830: $\C^{N+1}$ that gives the Jordan form of $T$.    
831: Let $
832: \mathcal{H} = \bigoplus_{\lambda \in \Lambda}   
833:  \mathcal{H}_{\lambda} \subset \sfi,   
834: $ where $\mathcal{H}_{\lambda} (\varphi) = \{h \in \sfi : h \in \homog{\lambda}{k},\   
835: \text{for some}\ k \geq 1 \}, \lambda \in \Lambda$.   
836: Then we have that $\dim(\mathcal{H})= N+1$.   
837: \end{theorem}
838: \begin{remark}   
839: Note that we can choose both $v_0=(1,0,\dots,0)$ and $v_N=(0,\dots,0,1)$ to be in   
840: the basis $\mathcal{B}$, corresponding to the eigenvalues $c_0$ and   
841: $c_N$ respectively.   
842: \end{remark}   
843: \begin{proof}   
844: If $v_i \in \mathcal{B}, (0 \leq i \leq N)$,    
845: then $v_i \in \kker(T - \lambda I)^k$ and   
846: $v_i \not\in \kker(T - \lambda I)^{k-1}$, for some $\lambda \in \Lambda$,   
847: and $k \geq 1$. So to each $v_i \in \mathcal{B}$, we can associate   
848: a unique pair $(\lambda,k)$. Let us denote such $v_i = v(\lambda,k)$.   
849: (Note that by the previous observation, $v_0 = v(c_0,1)$   
850: and $v_N = v(c_N,1)$).   
851:    
852: After Theorem~\ref{tone}, we can associate to each $v(\lambda,k)$   
853: a function $h_{v(\lambda,k)}$ in $\homog{\lambda}{k}\cap \sfi$.   
854: Furthermore, the functions $\{h_{v(\lambda,k)}\}_{v \in \mathcal{B}}$,   
855: are linearly independent.   
856: For this, observe that since the vectors in $\mathcal{B}$ are   
857: linearly independent, its extensions $\{Y_v\}$ are   
858: linearly independent in $\ell(\Z)$, and therefore   
859: the functions $\{h_{v(\lambda,k)}\}_{v \in \mathcal{B}}$   
860: are linearly independent.   
861:    
862: One can see that if a finite number of functions are homogeneous   
863: for the same $\lambda$, then a linear combination of them   
864: is also homogeneous for the same $\lambda$. More precisely,   
865: $\sum_{i=0}^n \alpha_i h_i(\lambda,k_i) = h(\lambda, k)$,  
866: where  $k = \max_i(k_i)$.
867:   
868: Hence, if $p_T$, the characteristic polynomial of $T$, is factorized as:   
869: $p_T(x) = \prod_{\lambda \in \Lambda} (x - \lambda)^{r_\lambda}$, 
870: then $\dim(\mathcal{H}_{\lambda}) = r_{\lambda}$ and a basis of   
871: $\mathcal{H}_{\lambda}$ is the set of $(2,\lambda,k)$-homogeneous   
872: functions associated to the vectors $v \in \mathcal{B}$,   
873: such that $v = v(\lambda,k)$, for some $k \geq 1$.   
874: \end{proof}   
875: 
876: \begin{theorem} \label{main}
877: Assume that $\varphi$ satisfies \eqref{dilation-equation} and has
878: linearly independent integer translates.
879: Let $\mathcal{B} = \{v_0, \dots, v_N\}$ be as before, a Jordan   
880: basis for $T$, and let $B$ be the $(N+1)\times (N+1)$ matrix that has the   
881: vectors $v_i$ as rows. 
882: Let   
883: \[ \phi^0(x) = \left[\varphi(x), \varphi(x+1), \dots, 
884: \varphi(x+N)\right]^t \quad \text{and}\quad   
885: h(x) = \left[h_0(x), h_1(x), \dots,    
886: h_N(x)\right]^t,\]   
887: where $h_i$ is the homogeneous function associated to the vector $v_i$. 
888: Then we have
889: \begin{enumerate}[\upshape (i)]
890: \item   $h(x) = B\phi^0(x)$   $x \in [-1,1]$. \label{i}
891: \item   \label{ii}
892: $ \displaystyle
893: \phi^0(x) = T \phi^0(2x) \quad \text{and} 
894: \quad h(x) = B T B^{-1} h(2x) \quad x \in [-1/2,1/2] 
895: $   
896: where $BTB^{-1}$ is in Jordan form.  
897: \item \label{iii} There exists a local basis of $\sfi$ consisting of homogeneous functions. 
898: Moreover, if $\varphi$ has accuracy $n$, this basis can be chosen to contain the
899: polynomials $\{1, x, \cdots, x^{n-1}\}$.
900: \end{enumerate}
901: \end{theorem}
902: 
903: \begin{remark}
904:   Note that \eqref{ii} is a statement about the refinability of both $\phi^0$ and $h$, where
905:   the scaling matrix of $h$ is the Jordan form of the scaling matrix of $\phi^0$.
906: \end{remark}
907: 
908: \begin{proof} 
909: Since the support of $\varphi$ is $[0,N]$, $\varphi(x+k) = 0$ if 
910: $k \not\in \{0, 1, \dots, N\}$ for $x \in [-1,1]$. Then 
911: $$
912: h_i(x) = Y^i\Phi(x) = v_i \Phi^0(x) \ x \in [-1,1].
913: $$
914: So we have, 
915: \begin{equation} \label{*}
916: h(x) = B\phi^0(x) \ x \in [-1,1].
917: \end{equation}
918: This shows that for every interval $I \subset [-1,1]$ 
919: the functions $\{\varphi(x), \varphi(x + 1), \cdots, \varphi(x +N)\}$ span the same space than the functions $\{h_0, \cdots, h_N\}$ when restricted to $I$.
920: 
921: Since $\{\varphi(x), \varphi(x + 1), \cdots, \varphi(x +N-1)\}$ is a local
922: basis of \sfi, if $v_N$ has been chosen to be $v_N = (0,\cdots,0,1)$, using equation \eqref{*} $\{h_0, \cdots, h_{N-1}\}$ are a local basis for \sfi. Moreover, if $\varphi$ has accuracy $n \leq N$, 
923: then we can choose $v_0, \cdots, v_{n-1}$ to be the eigenvectors associated to the 
924: eigenvalues $\{1, \cdots, 2^{-n+1}\}$, and hence $\{h_0, \cdots, h_{N-1}\} = 
925: \{1, x, \cdots, x^{n-1}, h_n, \cdots, h_{N-1}\}$.
926: 
927: This proves \eqref{i} and \eqref{iii}.
928: 
929: For \eqref{ii}, using again that the support of $\varphi$ is $[0,N]$,   it is easily seen that
930: if $ x \in [-\frac{1}{2},\frac{1}{2}]$ then  $\phi(x) = T \phi(2x).$  Then we have
931: $ B\Phi^0(x) = BTB^{-1}B\Phi^0(2x) \quad x \in [-\frac{1}{2},\frac{1}{2}].
932: $
933: \end{proof}
934: 
935: \subsection{Generalizations} 
936: 
937: Throughout the paper ``function'' meant ``measurable function''. However, with the obvious modifications, all results hold for the case that $\varphi$ is a generalized function or
938: distribution.
939: 
940: It is interesting to consider the generalization of the results to arbitrary dilations $M \geq 1$ in $\R$, and in higher dimensions with arbitrary dilation matrices. It can be shown that most of the results are still true in these cases \cite{CHnM05}.   
941: 
942: \section{Examples for $N=3$} 
943:  
944: %\subsection{B-spline} 
945: \noindent
946: {\bf B-spline} 
947: The simplest case of refinable functions are the {\em B-splines}. They are the 
948: (normalized) convolutions of the characteristic function of $[0,1]$ with itself.  
949: In particular the B-spline of order $3$ is the refinable function that 
950: satisfies:
951: \begin{equation} 
952: b(x) = \frac{1}{4} b(2x) + \frac{3}{4}b(2x-1) + \frac{3}{4} b(2x-2) 
953: + \frac{1}{4} b(2x-3). 
954: \end{equation} 
955: The B-splines are those functions, for which the accuracy is maximum and so coincides with the 
956: dimension of the matrix $T_0$, so in this case, the eigenvalues of $T_0$ are $1$ (for the 
957: constant functions), $\frac{1}{2}$ (for the linear functions), 
958: and $\frac{1}{4}$ (for the quadratic functions). 
959:  
960: %\subsection{Daubechies D$_4$} 
961: \noindent
962: {\bf Daubechies D$_4$} 
963: Daubechies wavelets, are those refinable functions of $N$ coefficients, that are 
964: orthogonal and provide the highest order of accuracy possible. (Note that the  
965: splines do not form an orthonormal basis). 
966: D$_4$ satisfies:
967: \begin{equation*} 
968: D_4(x) = \frac{1+\sqrt{3}}{4} D_4(2x) + \frac{3+\sqrt{3}}{4}D_4(2x-1) + 
969: \frac{3-\sqrt{3}}{4} D_4(2x-2) + \frac{1-\sqrt{3}}{4} D_4(2x-3). 
970: \end{equation*} 
971: D$_4$ has accuracy 2 (it reproduces the constant and the linear functions). 
972: In this case the matrix $T_0$ has eigenvalues $1$, $\frac{1}{2}$ and 
973: $c_0 = \frac{1+\sqrt{3}}{4}$. So a basis for $\text{span}\{D_4(x),%
974:  D_4(x-1), D_4(x-2)\}_{x \in [0,1]}$ is also given by 
975: $\text{span}\{1, x, h_{c_0}(x)\}_{x \in [0,1]}$ where $h_{c_0}$ is the homogeneous function associated to $c_0$. 
976:  
977: %\subsection{$(\lambda,1)$-Homogeneous functions are not enough} 
978: \noindent
979: {\bf $(\lambda,1)$-Homogeneous functions are not enough} 
980: In the two previous examples, we could always obtain a basis of  
981: $\text{span}\{f(x), f(x-1), f(x-2)\}_{x \in [0,1]}$ just by using 1-homogeneous functions. The following example is to illustrate, that even in the simple case of only 4 coefficients, 
982: it may be necessary to use homogeneous functions of order bigger than 1. 
983: Consider the function: 
984: \begin{equation} 
985: f(x) = \frac{1}{3} f(2x) + \frac{2}{3}f(2x-1) + \frac{2}{3} f(2x-2) 
986: + \frac{1}{3} f(2x-3). 
987: \end{equation} 
988: It can be shown that $f$ has accuracy 1, and the eigenvalues of $T$ are $\{1,\frac{1}{3}\}$. 
989: So in this case, $\text{span}\{f(x), f(x-1), f(x-2)\}_{x \in [0,1]} = 
990: \text{span}\{1, h_{\{1/3,1\}}(x), h_{\{1/3,2}\}(x)\}_{x \in [0,1]}$, where $h_{\{1/3,1\}}$ is a 1-homogeneous function corresponding to the eigenvalue $1/3$, and $h_{\{1/3,2\}}$ is a 2-homogeneous function corresponding to the eigenvalue $1/3$. 
991:  
992:  \section{Acknowledgments} We wish to thank the anonymous referee for many helpful suggestions.
993:  
994: %%%%%%%%%%%%%%%%%%% AQUI VA LA BIBLIOGRAFIA   
995:    
996: \bibliographystyle{amsalpha}   
997: % supone que vas a usar un archivo con bibliografia   
998: % que se llama cyu.bib   
999: \bibliography{cyu}   
1000:    
1001: %%%%%%%%%%%%%%%%%%% Termina el paper   
1002: \end{document}   
1003:    
1004:    
1005:    
1006: