math0205052/t-w.tex
1: \def\versiondate{25 Sep.\ 2002}
2: \input math.macros
3: \input Ref.macros
4: 
5: %\proofmodetrue
6: %\leftsectionheadtrue
7: \checkdefinedreferencetrue
8: %\continuousnumberingtrue
9: \continuousfigurenumberingtrue
10: \theoremcountingtrue
11: \sectionnumberstrue
12: %\figuresectionnumberstrue
13: \forwardreferencetrue
14: %\lefteqnumberstrue
15: %\tocgenerationtrue
16: \citationgenerationtrue
17: \nobracketcittrue
18: \hyperstrue
19: \initialeqmacro
20: 
21: \bibsty{myapalike}
22: \input\jobname.lbl 
23: 
24: \def\T{{\Bbb T}}
25: \def\GM{{\ss GM}}  %% geometric mean
26: \def\ip#1{(\changecomma #1)}
27: \def\bigip#1{\bigl(\bigchangecomma #1\bigr)}
28: \def\biggip#1{\biggl(\bigchangecomma #1\biggr)}
29: \def\leftip#1{\left(\leftchangecomma #1\right)}  
30: \def\changecomma#1,{#1,\,}
31: \def\bigchangecomma#1,{#1,\;}
32: \def\leftchangecomma#1,{#1,\ }
33: \def\constant#1{{\bf #1}}
34: \def\trig{{\ss Poly}}  %% trig poly
35: \def\onp{{\bf p}}  % the orthonormal polynomials for a measure
36: \def\tilonp{\widetilde{{\bf p}}}  % modified orthonormal polynomials 
37: \def\qonp{{\bf q}}  % further modified orthonormal polynomials 
38: \def\bfz{{\bf 0}}
39: \def\bfo{{\bf 1}}
40: \def\hl{{\ss HL}}  % Helson-Lowdenslager space
41: \def\supp{{\rm supp}\,}
42: \def\F{{\cal F}}
43: \def\G{{\cal G}}
44: 
45: 
46: 
47: \ifproofmode \relax \else\head{To appear in {\it Geom. Funct. Anal.}}
48: {Version of \versiondate}\fi 
49: \vglue20pt
50: 
51: \title{Szeg\H{o} Limit Theorems}
52: 
53: \author{Russell Lyons}
54: 
55: 
56: 
57: 
58: 
59: 
60: \abstract{%
61: The first Szeg\H{o} limit theorem has been extended by
62: Bump-Diaconis and Tracy-Widom to limits of other minors of Toeplitz
63: matrices.
64: We use a more geometric method to
65: extend their results still further. Namely, we allow more general measures and
66: more general determinants.
67: We also give a new extension to higher dimensions, which extends a theorem
68: of \refbauthor{HelLow}.
69: }
70: 
71: \bottomII{Primary 
72: 47B35. %Toeplitz operators, Hankel operators, Wiener-Hopf operator
73: Secondary
74: 11C20. %Matrices, determinants
75: }
76: {Toeplitz determinants, minors, spectral factors.}
77: {Research partially supported by 
78: NSF grant DMS-0103897.}
79: 
80: 
81: \bsection{Introduction}{s.intro}
82: 
83: Let $\lambda$ denote Lebesgue measure on the unit-length circle $\T := \R/\Z$.
84: For a finite positive measure $\mu$ on $\T$, its Fourier coefficients are
85: $\widehat\mu(n) := \int_{\T} e^{-2\pi i n t}\,d\mu(t)$.
86: For $n \ge 0$, define 
87: $$
88: D_n(\mu) := \det[\widehat\mu(k-j)]_{0 \le j, k \le n}
89: \,.
90: $$
91: The Szeg\H{o} limit theorems determine the asymptotics of $D_n(\mu)$.
92: The first Szeg\H{o} limit theorem (\ref b.GZ/, p.~44), which is the one of
93: concern here, determines $\lim_{n \to\infty} D_{n+1}(\mu)/D_n(\mu)$.
94: To state this beautiful result of Szeg\H{o}'s in its extended form due to
95: Kolmogorov and Kre\u{\i}n, define,
96: for $f \ge 0$ and $\log^+ f \in L^1(\lambda)$, %or $\log^- f \in L^1(\lambda)$,
97: the {\bf geometric mean} of $f$ by 
98: $$
99: \GM(f) := \exp \int_{\T} \log f \,d\lambda
100: \,.
101: $$
102: According to the arithmetic mean-geometric mean inequality, if $0 \le f \in
103: L^1(\lambda)$, then $\log^+ f \in L^1(\lambda)$ and $0 \le \GM(f) \le \int f
104: \,d\lambda$.
105: 
106: \procl t.first \procname{Szeg\H{o} Limit Theorem}
107: Let $\mu$ be a finite positive measure on $\T$ with infinite support.
108: Let $f:= [d\mu/d\lambda]$ be the Radon-Nikod\'ym derivative of the
109: absolutely continuous part of $\mu$.
110: Then 
111: $$
112: \lim_{n \to\infty} D_{n+1}(\mu)/D_n(\mu) = \GM(f)
113: \,.
114: \label e.basic
115: $$
116: \endprocl
117: 
118: This result has been extended in various ways, three of which we consider
119: here.
120: The first two extensions were proved by \ref b.BD/ and \ref b.TW/, while
121: the third was proved by \ref b.HelLow/.
122: Our theorems extend theirs further yet.
123: These extensions concern quotients of determinants where, instead of the
124: determinant in the numerator having one more particular row and column than
125: the determinant in the denominator, as in \ref e.basic/, the numerator
126: contains finitely many more rows and columns arising from inner
127: products among arbitrary vectors.
128: In addition, we shall consider the case where $\mu$ is complex, as did
129: \refbauthor{BD} and \refbauthor{TW}.
130: Theorems \briefref t.nonneg/ and \briefref t.nonneg-d/ of this paper are
131: used in \ref b.LyonsSteif:dyn/ for studying entropy and phase transitions
132: in stationary determinantal processes. 
133: 
134: The approach of \refbauthor{BD} relies on certain identities for symmetric
135: functions and representations of the symmetric group, while \refbauthor{TW}
136: proceed via factorization and analysis of Toeplitz operators.
137: Our approach is different, although it bears some similarities to that of
138: \refbauthor{TW}. Namely, we go to the level of vectors in Hilbert space
139: and analyze projections on various subspaces. This more geometric
140: method appears to be more flexible
141: and leads to a more transparent proof and formulation of the results.
142: Although we use row operations that leave determinants unchanged, one could
143: instead use multivectors and continue a geometric approach. However, to
144: maintain greater accessibility, we have omitted exterior algebra.
145: 
146: In a brief \ref s.notn/, we recall some general notation and facts from
147: complex analysis.
148: This allows us to state and prove our first result in \ref s.positive/, which
149: concerns the case $\mu \ge 0$.
150: We turn to the case of complex $\mu$ in \ref s.cplx/.
151: This requires a lemma about convergence of non-orthogonal linear
152: projections in Hilbert space, whose proof is relegated to the appendix of
153: the paper.
154: \ref b.HelLow/ proved an analogue of \ref t.first/ for higher dimensions.
155: In \ref s.high/, we extend their theorem by proving analogues of
156: our results for higher
157: dimensions; both the real and complex case are treated there.
158: Actually, \refbauthor{HelLow} stated their result not as a limit of
159: quotients of determinants, but as an extremum problem, as in \ref b.GZ/.
160: 
161: 
162: 
163: 
164: 
165: \bsection{General Notation and Hardy Spaces}{s.notn}
166: 
167: 
168: Write $e_n(t) := e^{2\pi i n t}$ and for $f \in L^1(\lambda)$, write $\widehat
169: f(n) := \int_{\T} f \overline{e_n} \,d\lambda$.
170: For $p \ge 1$, let $H^p(\T)$ denote the Hardy space of those $f \in
171: L^p(\lambda)$ with $\widehat f(n) = 0$ for all $n < 0$.
172: Write $P_{H^2}$ for the orthogonal projection from $L^2(\lambda) \to
173: H^2(\T)$.
174: For all $f \in H^1(\T)$, we have $\GM(|f|) > 0$
175: (see \ref b.Rudin:RCA/, Theorem 17.17, p.~344).
176: For the converse, for any function
177: $f \ge 0$ with $\log f \in L^1(\lambda)$, define
178: $$
179: \Phi_f(z) := \exp {1 \over 2} \int_{\T} {e_1(t) + z
180: \over e_1(t) - z} \log f(t) \,d\lambda(t)
181: \label e.defPhi_f
182: $$
183: for $|z| < 1$.
184: The {\bf outer} function
185: $$
186: \varphi_f(t) := \lim_{r \uparrow 1} \Phi_f(r e_1(t))
187: \label e.defvarphi_f
188: $$
189: exists for $\lambda$-a.e.\ $t \in \T$ and satisfies $|\varphi_f|^2 = f$
190: $\lambda$-a.e.
191: Also, $\varphi_f \in H^p(\T)$ iff $f \in L^{p/2}(\lambda)$.
192: If $f \in L^2(\lambda)$, then the
193: limit in \ref e.defvarphi_f/ also holds in $L^{1}(\lambda)$.
194: See \ref b.Rudin:RCA/, Theorem 17.11, p.~340, and Theorem 17.16, p.~343.
195: Cauchy's integral formula shows that if $f \in L^2(\lambda)$, then
196: $$
197: \widehat {\varphi_f}(0) = \Phi_f(0) = \sqrt{\GM(f)}
198: \,.
199: \label e.PhiGM
200: $$
201: %Clearly, $\Phi_f$ has no zeroes in the unit disc.
202: %
203: %if $f \in L^1(\lambda)$ and $\GM(f) > 0$, then $\Phi_f \in H^2(\T)$ and
204: %Also, $\Phi_f(r e_1(\cbuldot)) \to \varphi_f$ in $L^2(\T)$.
205: %if $\GM(f) > 0$,
206: By the factorization Theorem 17.17, p.~344, of \ref b.Rudin:RCA/,
207: if $g \in H^p(\T)$ and $\varphi$ is an outer function such that 
208: $g \varphi \in L^q(\lambda)$, then $g \varphi \in H^q(\T)$.
209: Define $\Phi_f := \varphi_f := \constant 0$ if $\log f \notin L^1(\lambda)$.
210: 
211: 
212: \bsection {Positive Measures}{s.positive}
213: 
214: Note that the determinant $D_n(\mu)$
215: is the same when $j, k$ take values in any index set
216: of $n+1$ consecutive integers.
217: We shall, in fact, use the index set $ \{ -1, -2, \ldots, -n \}$ for
218: $D_{n-1}(\mu)$.
219: 
220: In this section, we extend \ref t.first/ to more general determinants
221: as follows.
222: 
223: \procl t.nonneg
224: Let $\mu$ be a finite positive measure on $\T$ with infinite support.
225: Let $f:= [d\mu/d\lambda]$ be the Radon-Nikod\'ym derivative of the
226: absolutely continuous part of $\mu$.
227: Given any functions $f_0, \ldots, f_{r}, g_0, \ldots, g_r \in L^2(\mu)$, let
228: $F_j := P_{H^2}(f_j \overline {\varphi_f})$ and $G_j := P_{H^2}(g_j \overline
229: {\varphi_f})$.
230: Define 
231: $$
232: p_j := \cases{
233: f_j &if $0 \le j \le r$,\cr
234: e_j &if $-1 \ge j \ge -n$\cr
235: }
236: $$
237: and
238: $$
239: q_j := \cases{
240: g_j &if $0 \le j \le r$,\cr
241: e_j &if $-1 \ge j \ge -n$.\cr
242: }
243: $$
244: %$$
245: %a_{j, k}^{(n)} := \cases{
246: %\int f_j \overline {f_k} \,d\mu &if $0 \le j, k \le r$,\cr
247: %\noalign{\vskip5pt}
248: %\widehat\mu(k-j) &if $-1 \ge j, k \ge -n$,\cr
249: %\noalign{\vskip5pt}
250: %\widehat{f_j\mu}(k) &if $0 \le j \le r$ and $-1 \ge k \ge -n$,\cr
251: %\noalign{\vskip5pt}
252: %\overline{\widehat{f_k\mu}(j)} &if $-1 \ge j \ge -n$ and $0 \le k \le
253: %r$.\cr
254: %}
255: %$$
256: We have 
257: $$
258: \lim_{n \to\infty}
259: %\det[a_{j, k}^{(n)}]/\det[\widehat\mu(k-j)]_{-1 \ge j, k \ge -n}
260: D_{n-1}(\mu)^{-1}
261: \det\left[\int p_j \overline{q_k} \,d\mu\right]_{-n \le j, k \le r}
262: =
263: \det\left[ \int F_j \overline{G_k} \,d\lambda\right]_{0 \le j, k \le r}
264: \,.
265: $$
266: \endprocl
267: 
268: 
269: \procl r.defined
270: If $\GM(f) = 0$, then $\varphi_f = \constant 0$, so
271: $F_j = G_j = \constant 0$. Otherwise, $f > 0$ $\lambda$-a.e., so that
272: $\lambda \ll f\lambda$ and there is no ambiguity about the equivalence
273: class of $f_j$ or $g_j$ with respect to $\lambda$.
274: Also, $\int |f_j \overline{\varphi_f}|^2 \,d\lambda = \int |f_j|^2 f \,d\lambda
275: \le \int |f_j|^2 \,d\mu < \infty$, so that $f_j \overline{\varphi_f} \in
276: L^2(\lambda)$ and $F_j$ is well defined. Likewise $G_j$ is well defined.
277: \endprocl
278: 
279: \procl r.nonsing
280: Since $\mu$ has infinite support, $[\widehat\mu(k-j)]_{-1 \ge j, k \ge -n}$
281: is non-singular. Indeed, if it were singular, then since it is a Gram
282: matrix $[\ip{e_j, e_k}_\mu]$, where the subscript $\mu$ indicates that
283: the inner product is taken in $L^2(\mu)$, it would follow
284: that the vectors $e_{-1}, \ldots, e_{-n}$ would be linearly dependent,
285: i.e., there would be scalars $a_j$ such that $\sum_j a_j e_j = 0$
286: $\mu$-a.e. This would imply that $\mu$ would have support contained in the
287: zero set of this trigonometric polynomial, i.e., $\mu$ would have support
288: of cardinality at most $n-1$. 
289: \endprocl
290: 
291: 
292: \procl r.bdtw
293: The case considered by \ref b.BD/ and \ref b.TW/ is that where all
294: functions $f_j$ and $g_j$ are of the form $e_n$ for various $n \ge 0$.
295: These give minors of the Toeplitz matrix other than merely $D_n(\mu)$.
296: In this case, when $f_j := e_j$ and $g_k := e_k$,
297: the limiting matrix entries $\int F_j \overline{G_k} \,d\lambda$
298: become 
299: $$
300: \int P_{H^2}(e_j \overline{\varphi_f}) \overline{P_{H^2}(e_k
301: \overline{\varphi_f})}\,d\lambda 
302: =
303: \sum_{l=0}^{\min(j, k)} \overline{\widehat\varphi_f(j-l)}
304: \widehat\varphi_f(k-l)
305: \,.
306: \label e.twform
307: $$
308: \refbauthor{BD} gave a different formula than \ref e.twform/; \refbauthor{TW}
309: gave the same formula as ours.
310: Both sets of authors assumed that $\mu$ was absolutely continuous.
311: In addition, \refbauthor{BD} assumed that $f = e^g$ for some $g$ satisfying
312: %belongs to the Wiener algebra and the Besov space $B_2^{1/2}$ 
313: $\sum_{n \in \Z} \big( |\widehat g(n)| + |n \widehat g(n)|^2 \big) < \infty$,
314: while \refbauthor{TW} assumed that $f$ was bounded above and bounded away
315: from 0.
316: On the other hand, \refbauthor{BD} showed the strong Szeg\H{o} limit
317: theorem, which gives finer asymptotics.
318: \endprocl
319: 
320: 
321: \procl r.gren-ros
322: The special case $r:=0$, $f_0 := g_0 := e_j$ for any fixed $j > 0$ and
323: $\mu$ is absolutely continuous is due
324: to Kolmogorov and Wiener (see \ref b.GZ/, Section 10.9).
325: \endprocl
326: 
327: 
328: \procl r.noneed
329: In case one of $F_j$ or $G_k$ is easier to calculate than the other, one
330: could use instead of $\int F_j \overline{G_k} \,d\lambda$ either of the
331: equivalent expressions
332: $\int F_j \overline{g_k} \varphi_f \,d\lambda$ or
333: $\int f_j \overline{G_k \varphi_f} \,d\lambda$.
334: \endprocl
335: 
336: \proofof t.nonneg
337: For the ease of the reader, we treat first the case $r=0$, $f_0 = g_0 =
338: \constant 1$, when \ref t.nonneg/ becomes the Szeg\H{o} limit theorem.
339: Since $\widehat \mu(k-j) = \ip{e_j, e_k}_\mu$, we have that 
340: $$
341: %\det[\widehat\mu(k-j)]_{0 \ge j, k \ge -n}/
342: %\det[\widehat\mu(k-j)]_{-1 \ge j, k \ge -n}
343: D_n(\mu)/D_{n-1}(\mu)
344: =
345: \|P_n e_0\|_\mu^2
346: \,,
347: $$
348: where $P_n$ is the orthogonal projection onto $\{e_{-1}, \ldots,
349: e_{-n}\}^\perp$ in $L^2(\mu)$.
350: (This is sometimes called ``Gram's formula".)
351: This quotient therefore tends (monotonically)
352: to $\|P_\infty e_0\|_\mu^2$, where $P_\infty$ is the
353: orthogonal projection of $L^2(\mu)$ onto 
354: $$
355: H_\infty := \{ g \in L^2(\mu) \st \all {n < 0} \ip{g, e_n}_\mu = 0 \}
356: \,.
357: $$
358: Now $g \in H_\infty$ iff $g \in L^2(\mu)$ and $g \mu$ is an analytic measure.
359: By the F.\ and M.\ Riesz theorem (\ref b.Rudin:RCA/, Theorem 17.13,
360: p.~341), it follows that $g \in H_\infty$ iff $g = 0$ a.e.\ with respect to
361: the singular part of $\mu$, $g \in L^2(f)$ and $g f \in H^1(\T)$.
362: In particular, we may from now on disregard the singular part of $\mu$.
363: That is, $P_\infty e_0$ is the same as the orthogonal projection of $e_0$ in
364: $L^2(f)$ onto $H_\infty := \{ g \in L^2(f) \st \all {n < 0} \ip{g, e_n}_f = 0
365: \}$ and its norm in $L^2(\mu)$ is the same as its norm in $L^2(f)$.
366: Write $h_0 := f \cdot P_\infty e_0$
367: and $\varphi := \varphi_f$.
368: 
369: If $\GM(f) > 0$, then
370: $h_0/f \in L^2(f)$, $h_0 \in H^1(\T)$ and $h_0/\varphi \in H^2(\T)$.
371: Also, for all $g \in H_\infty$, we have 
372: $$
373: \bigip{h_0/f, g}_\mu = \ip{e_0, g}_\mu
374: \,.
375: $$
376: For all $m \ge 0$, we have $e_m/\overline\varphi \in H_\infty$, whence
377: $$
378: \bigip{h_0/f, e_m/\overline\varphi}_\mu = \ip{e_0, e_m/\overline\varphi}_\mu
379: \,,
380: $$
381: or in other words, 
382: $
383: \widehat{(h_0/\varphi)}(m)
384: =
385: \widehat{\overline\varphi}(m)
386: $.
387: Since $\widehat{\overline\varphi}(m) = \sqrt{\GM(f)} \delta_{0, m}$ for $m \ge
388: 0$, we obtain that
389: $$
390: \widehat{(h_0/\varphi)}(m)
391: =
392: \sqrt{\GM(f)} \delta_{0, m}
393: \,.
394: \label e.FC
395: $$
396: for $m \ge 0$. 
397: Since $h_0/\varphi \in H^2(\T)$, \ref e.FC/ holds for all $m$.
398: That is, $h_0 = \sqrt{\GM(f)} \varphi$.
399: Therefore, 
400: $$
401: \|P_\infty e_0\|_\mu^2 
402: =
403: \|h_0/f\|_\mu^2
404: =
405: \int {|h_0|^2 \over f} \,d\lambda
406: =
407: \GM(f)
408: \,.
409: $$
410: 
411: If $\GM(f) = 0$, then for all $g \in H_\infty$, 
412: $$
413: \GM(|g f|)^2
414: =
415: \GM(|g f|^2)
416: =
417: \GM(|g|^2 f) \GM(f) 
418: =
419: 0
420: $$
421: since $\GM(|g|^2 f) \le \int |g|^2 f \,d\lambda < \infty$ as $g \in L^2(f)$.
422: Since $g f \in H^1(\T)$, this means that $g f = \constant 0$ as noted in \ref
423: s.notn/. In other words, $g = 0$ $\mu$-a.e.
424: Therefore $H_\infty = 0$ and so the limit is 0.
425: 
426: We have thus proved the Szeg\H{o} formula.
427: This proof also shows immediately that the linear span of
428: $ \{ e_n \st n \ge 0 \}$ is dense in $L^2(\mu)$ iff $\GM(f) = 0$, a theorem
429: of Kolmogorov and Kre\u\i n.
430: (More precisely, as written, this proof decides the density of the linear span
431: of $ \{ e_n \st n \le -1 \}$ by deciding whether its orthocomplement is 0, but
432: this is equivalent.)
433: 
434: Now we continue with the general case.
435: Consider $j \ge 0$.
436: Since 
437: $$
438: P_n f_j = f_j - \sum_{-1 \ge i \ge -n} a_i e_i
439: $$
440: for some constants $a_i$, row operations can be used to change the $j$th
441: row from its initial value $[\ip{f_j, q_k}_\mu]_{-n \le k \le r}$ to $[\ip{
442: P_n f_j, q_k}_\mu]_{-n \le k \le r}$ without changing the determinant.
443: Since $P_n$ is an orthogonal projection, we have $\ip{P_n f_j, q_k}_\mu =
444: \ip{P_n f_j, P_n q_k}_\mu$.
445: If we change all rows $j \ge 0$ in this manner, we obtain a block diagonal
446: matrix, which shows that
447: $$
448: D_{n-1}(\mu)^{-1}
449: \det\left[\ip{p_j, q_k}_\mu\right]_{-n \le j, k \le r}
450: =
451: \det[\ip{P_n f_j, P_n g_k}_\mu]_{0 \le j, k \le r}
452: \,.
453: $$
454: Thus, the limit is
455: $$
456: \det[\ip{P_\infty f_j, P_\infty g_k}_\mu]_{0 \le j, k \le r}
457: \,.
458: $$
459: As before, if $\GM(f) = 0$, then $H_\infty = 0$ and the limit is 0. Otherwise, 
460: the reasoning that led to \ref e.FC/ now leads to 
461: $$
462: [f (P_\infty f_j)/\varphi]^{\widehat{\quad}}(m) = \widehat{f_j \overline\varphi}(m)
463: $$
464: for all $m \ge 0$, whence $f (P_\infty f_j)/\varphi = F_j$.
465: Likewise, $f (P_\infty g_k)/\varphi = G_k$.
466: This gives the formula since 
467: $$
468: \ip{P_\infty f_j, P_\infty g_k}_\mu 
469: =
470: \int P_\infty f_j \cdot \overline{P_\infty g_k} \,d\mu
471: =
472: \int P_\infty f_j \cdot \overline{P_\infty g_k} \cdot f\,d\lambda
473: =
474: \int F_j \overline{G_k} \,d\lambda
475: \,.
476: \Qed
477: $$
478: 
479: \procl r.genfn
480: A bivariate generating function for the matrix entries of \ref e.twform/ is 
481: $$
482: \sum_{j, k \ge 0} z^j \overline\zeta^k 
483: \sum_{l=0}^{\min(j, k)} \overline{\widehat\varphi_f(j-l)}
484: \widehat\varphi_f(k-l)
485: =
486: {\Phi_f(z) \overline{\Phi_f(\zeta)} \over 1 - \overline\zeta z}
487: \,.
488: $$
489: \endprocl
490: 
491: 
492: \bsection{Complex Measures}{s.cplx}
493: 
494: We now consider the case of absolutely continuous complex measures, $\mu$.
495: The proof of the main result in this section,
496: \ref t.cplx/, could be modified so as to allow a positive
497: singular part to $\mu$ and to include all of \ref t.nonneg/.
498: However, the proof would become less elegant.
499: 
500: Let $\trig_n$ denote the linear span of $ \{ e_0, e_1, \ldots, e_n \}$.
501: Given a pair of functions $\varphi, \psi \in L^2(\lambda)$, consider the
502: condition 
503: $$
504: \texists {\epsilon > 0} \texists {n_0}
505: \all {n \ge n_0} \all {S \in \trig_n} \texists {T \in \trig_n \setminus \{
506: 0 \}}  \quad
507: \ip{\varphi S, \psi T}_\lambda
508: \ge
509: \epsilon \|\varphi S\|_\lambda \|\psi T\|_\lambda
510: \,.
511: \label e.good
512: $$
513: Of course, this holds if $\varphi = \psi$, since we may then take $\epsilon
514: := 1$ and $T := S$.
515: Some readers may prefer the following restatement of \ref e.good/. Given
516: two subspaces $H_1$ and $K_1$ of a Hilbert space $H$, define 
517: $$
518: \epsilon(H_1, K_1; H) := \epsilon(H_1, K_1) := \inf_{{x \in H_1 \atop \|x\|
519: = 1}} \sup_{{y \in K_1 \atop \|y\|=1}} |(x, y)|
520: \,.
521: $$
522: The condition $\epsilon(H_1, K_1) > 0$ is weaker than $H_1 = K_1$ and
523: stronger than $H_1 \cap K_1^\perp = 0$.
524: Our condition \ref e.good/ is equivalent to
525: $$
526: \liminf_{n \to\infty} \epsilon\big(\varphi \cdot \trig_n, \psi \cdot \trig_n;
527: L^2(\lambda)\big) > 0
528: \,.
529: $$
530: 
531: Condition \ref e.good/ will be used via the following criterion.
532: We write $H_n \uparrow H_\infty$ to mean that $H_n \subseteq H_{n+1}$ for all
533: $n$ and $\bigcup H_n$ is dense in $H_\infty$.
534: 
535: \procl l.projection
536: Suppose that $H$ is a Hilbert space, $H_n, K_n$ are non-zero
537: closed subspaces for $1 \le n \le \infty$
538: with $H = H_n + K_n^\perp$ and
539: $H_n \cap K_n^\perp = 0$ for all $1 \le n \le \infty$.
540: Suppose that $H_n \uparrow H_\infty$ and $K_n \uparrow K_\infty$.
541: Let $T_n : H \to K_n^\perp$ be the linear projection along $H_n$ ($1 \le n \le
542: \infty$).
543: Then $T_n \to T_\infty$ in the strong operator topology iff
544: $$
545: \liminf_{n \to\infty} \epsilon(H_n, K_n) > 0
546: \,.
547: \label e.gengood
548: $$
549: \endprocl
550: 
551: This lemma should be known, but we could not locate a reference.
552: Thus, we include its proof in an appendix.
553: Note that when $H_n = K_n$, which will correspond to the case $\varphi =
554: \psi$ in our application, it is trivial that $T_n \to T_\infty$ in the
555: strong operator topology.
556: 
557: As we have noted already, $\epsilon(H_n, K_n) > 0$
558: implies that $H_n \cap K_n^\perp = 0$.
559: If $\dim H_n = \dim K_n <\infty$, as will be the case in our application of
560: \ref e.gengood/, this in turn implies that $H = H_n + K_n^\perp$.
561: 
562: \procl t.cplx
563: Suppose that $\mu = \psi\overline{\varphi }\lambda$ for some pair of outer
564: functions $\varphi, \psi \in H^2(\T)$ that satisfies condition \ref e.good/.
565: Given any functions $f_0, \ldots, f_{r}, g_0, \ldots, g_r \in L^2(|\varphi|^2
566: + |\psi|^2)$, let $F_j := P_{H^2}(f_j \overline \varphi)$ and $G_j :=
567: P_{H^2}(g_j \overline \psi)$.
568: Define 
569: $$
570: p_j := \cases{
571: f_j &if $0 \le j \le r$,\cr
572: e_j &if $-1 \ge j \ge -n$\cr
573: }
574: $$
575: and
576: $$
577: q_j := \cases{
578: g_j &if $0 \le j \le r$,\cr
579: e_j &if $-1 \ge j \ge -n$.\cr
580: }
581: $$
582: We have 
583: $$
584: \lim_{n \to\infty}
585: D_{n-1}(\mu)^{-1}
586: \det\left[\int p_j \overline{q_k} \,d\mu\right]_{-n \le j, k \le r}
587: =
588: \det\left[ \int F_j \overline{G_k} \,d\lambda\right]_{0 \le j, k \le r}
589: \,.
590: $$
591: \endprocl
592: 
593: Note that $L^2(|\varphi|^2 + |\psi|^2) \subseteq L^2(|\mu|)$ by the
594: Cauchy-Schwarz inequality.
595: 
596: 
597: \proof
598: Let $H_n(\varphi) := e_1 \varphi \trig_n$.
599: By virtue of \ref e.good/, we have for $n \ge n_0$,
600: $$
601: \overline{H_n(\varphi)} \cap \overline{H_n(\psi)}^\perp = 0
602: \,,
603: $$
604: and so
605: $$
606: L^2(\T) = \overline{H_n(\varphi)} + \overline{H_n(\psi)}^\perp
607: \,.
608: $$
609: A consequence of Beurling's theorem (\ref b.Rudin:RCA/, Theorem 17.23,
610: p.~350) is that 
611: $$
612: H_n(\varphi) \uparrow H^2_0(\T) := e_1 H^2(\T)
613: \,.
614: \label e.beurling
615: $$
616: Thus the projection along $\overline{H_n(\varphi)}$ to
617: $\overline{H_n(\psi)}^\perp$ tends to the orthogonal projection
618: $P_{\overline{H^2_0(\T)}^\perp} = P_{H^2}$.
619: 
620: Now $\int p_j \overline{q_k} \,d\mu = \ip{\overline\varphi p_j, \overline\psi 
621: q_k}_\lambda$.
622: Let $F^{(n)}_j$ be the projection of $\overline\varphi f_k$ along
623: $\overline{H_n(\varphi)}$ to $\overline{H_n(\psi)}^\perp$.
624: Row operations show that for $n \ge n_0$, 
625: $$
626: D_{n-1}(\mu)^{-1}
627: \det[\int p_j \overline{q_k} \,d\mu]_{-n \le j, k \le r}
628: =
629: \det[\ip{F^{(n)}_j, \overline\psi g_k}_\lambda]_{0 \le j, k \le r}
630: \,.
631: $$
632: Because of our assumption \ref e.good/ and \ref l.projection/, the limit is
633: $
634: \det[\ip{F_j, \overline\psi g_k}_\lambda]_{0 \le j, k \le r}
635: $,
636: which is the same as
637: $
638: \det[\ip{F_j, G_k}_\lambda]_{0 \le j, k \le r}
639: $.
640: \Qed
641: 
642: \procl r.beurling
643: The limit \ref e.beurling/ is often used to prove Beurling's theorem and it
644: has a simple direct proof: If $g \in H^2_0(\T)$ and $g \perp H_n(\varphi)$
645: for all $n \ge 0$, then $\widehat{g \overline\varphi}(k) = 0$ for all $k \ge
646: 1$, i.e., $g \overline\varphi \in \overline{H^1(\T)}$.
647: Dividing by $\overline\varphi$, we get that $g \in \overline{H^2(\T)} =
648: (H^2_0(\T))^\perp$, so that $g = \constant 0$.
649: \endprocl
650: 
651: %\procl r.bdtw
652: The case considered by \ref b.BD/ and \ref b.TW/ is that where all
653: functions $f_j$ and $g_j$ are of the form $e_n$ for various $n \ge 0$.
654: In addition, \refbauthor{BD} assumed that $f = e^g$ for some $g$ satisfying
655: %belongs to the Wiener algebra and the Besov space $B_2^{1/2}$
656: $\sum_{n \in \Z} \big( |\widehat g(n)| + |n \widehat g(n)|^2 \big) < \infty$, 
657: while \refbauthor{TW} assumed that
658: $\varphi$ and $\psi$ are bounded above and that the Toeplitz matrix
659: corresponding to $\psi\overline\varphi$ has uniformly invertible finite
660: sections.
661: 
662: The assumption of \refbauthor{BD} implies that of \refbauthor{TW}.
663: Indeed, write $g = g_1 + g_2$, where $g_1 := \sum_{n \ge 0} \widehat
664: g(n) e_n$ and $g_2 := \sum_{n < 0} \widehat g(n) e_n$.
665: Set $f_1 := e^{g_1 + \overline{g_1}}$ and
666: $f_2 := e^{g_2 + \overline{g_2}}$.
667: Then $f = \psi \overline\varphi$ with $\psi := e^{g_1} = \varphi_{f_1}$
668: and $\varphi := e^{\overline{g_2}} = \varphi_{f_2}$, so that $\psi$ and
669: $\varphi$ are bounded outer functions.
670: Furthermore, since
671: $g$ is continuous, Kre\u\i{}n's Theorem (\ref
672: b.BS:large/, Theorem 1.15, p.~18) in combination with a theorem of Gohberg
673: and Feldman (\ref b.BS:large/, Theorem 2.11, p.~39) shows that the Toeplitz
674: matrix of $f$ has uniformly invertible finite sections.
675: 
676: Our theorem covers that of \refbauthor{TW} since boundedness of $\psi$ and
677: uniform invertibility of finite
678: sections implies uniform boundedness of the projections along
679: $\overline{H_n(\varphi)}$ to $\overline{H_n(\psi)}^\perp$, as we see by
680: simply writing the equations: If $g \in L^2(\lambda)$ is written as $g = u + v$
681: with $u \in \overline{H_n(\varphi)}$ and $v \in \overline{H_n(\psi)}^\perp$,
682: then write $\overline u = \sum_{k=1}^n a_k e_k \varphi$. The coefficients
683: $a_k$ are determined by the requirement that $g - u \perp
684: \overline{H_n(\psi)}$, i.e., by the equations 
685: $$
686: \all {k \in [1, n]} \quad
687: \widehat{\psi g}(-k)
688: =
689: \sum_{j=1}^n {a_j} \widehat{\psi\overline\varphi}(j-k)
690: \,.
691: $$
692: Now observe that $\left[\sum_{k=1}^n |\widehat{\psi g}(-k)|^2 \right]^{1/2}
693: \le \|\psi g\|_\lambda \le \|\psi\|_\infty \|g\|_\lambda$.
694: %\endprocl
695: 
696: We next give some additional cases when \ref e.good/ holds.
697: We begin with a reformulation of \ref e.good/, for which we are grateful to
698: Doron Lubinsky.
699: Let $w := |\psi|^2$ and $\sigma := \varphi/\psi$.
700: Then 
701: $$
702: \ip{\varphi S, \psi T}_\lambda
703: =
704: \int \sigma S \overline T w\,d\lambda
705: \,,
706: $$
707: so that 
708: $$
709: \sup_{T \in \trig_n} 
710: {|\ip{\varphi S, \psi T}_\lambda| \over \|\psi T\|_\lambda}
711: =
712: \sup_{T \in \trig_n} 
713: {|\ip{\sigma S, T}_w| \over \|T\|_w}
714: =
715: \|P_{\trig_n}(\sigma S)\|_w
716: \,,
717: $$
718: where the orthogonal projection onto $\trig_n$ takes place in $L^2(w)$.
719: Note that $\sigma \in L^2(w)$ since $\varphi \in L^2(\lambda)$.
720: Let $\onp_n$ be the standard orthogonal polynomials for the weight $w$, i.e.,
721: $\onp_n \in \trig_n$ with positive leading coefficient and $\ip{\onp_m,
722: \onp_n}_w = \delta_{m, n}$.
723: If we write 
724: $$
725: \sigma S = \sum_{k \ge 0} a_k \onp_k
726: \,,
727: \label e.defa_k
728: $$
729: then $\|\sigma S\|^2_w =
730: \sum_{k \ge 0} |a_k|^2$ and
731: $\|P_{\trig_n}(\sigma S)\|_w^2 = \sum_{k=0}^n |a_k|^2$.
732: Thus, \ref e.good/ is equivalent to
733: $$
734: \texists {\epsilon > 0} \texists {n_0}
735: \all {n \ge n_0} \all {S \in \trig_n} 
736: \sum_{k=0}^n |a_k|^2
737: \ge
738: \epsilon 
739: \sum_{k=0}^\infty |a_k|^2
740: \,,
741: \label e.goodonp
742: $$
743: where $a_k$ are defined by \ref e.defa_k/.
744: 
745: \procl p.poly
746: If $\varphi, \psi \in L^2(\T)$ and $\sigma := \varphi/\psi$
747: is an analytic polynomial that has 
748: no zeroes in the closed unit disc, then \ref e.good/ holds.
749: \endprocl
750: 
751: Of course, this means that \ref t.cplx/ applies to the pair $\varphi, \psi$
752: if, in addition, they are outer functions.
753: 
754: \proof 
755: We use the notation above and show that \ref e.goodonp/ holds.
756: Let the zeroes of $\sigma$ be $z_1, \ldots, z_m$, all outside the
757: closed unit disc, and suppose first that each zero is simple.
758: Let $S \in \trig_n$.
759: Since $(\sigma S)(z_j) = 0$, we have 
760: $$
761: \sum_{l=1}^m a_{n+l} \onp_{n+l}(z_j)
762: =
763: -\sum_{k=0}^n a_{k} \onp_k(z_j)
764: $$
765: for $1 \le j \le m$ in the notation of \ref e.defa_k/.
766: Write these equations as 
767: $$
768: \sum_{l=1}^m a_{n+l} \tilonp_{n+l}(z_j)/z_j^{n+1}
769: =
770: -\sum_{k=0}^n a_{k} \tilonp_k(z_j)/z_j^{n+1}
771: =: \zeta_j
772: \,,
773: \label e.system
774: $$
775: where 
776: $$
777: \tilonp_k(z) :=
778: \onp_k(z) \left(\overline{\Phi_w(\overline z^{-1})}\right)^{-1}
779: \,.
780: $$
781: Recall Szeg\H{o}'s asymptotics 
782: $$
783: \hbox{for all } \rho > 1,\quad \lim_{k \to\infty} z^{-k} \tilonp_k(z)
784: =
785: 1 \hbox{ uniformly for } |z| > \rho
786: \label e.asymp
787: $$
788: (see \ref b.GZ/, p.~51).
789: Because $|z_j| > 1$, it follows that 
790: $$
791: C := \sup_{j, n} \sum_{k=0}^n |\tilonp_k(z_j)/z_j^{n+1}|^2 <\infty
792: \,.
793: $$ 
794: By the Cauchy-Schwarz inequality, we obtain that
795: $$
796: |\zeta_j|^2 \le C \sum_{k=0}^n |a_k|^2
797: \,.
798: $$
799: Let $M_n := [\tilonp_{n+l}(z_j)/z_j^{n+1}]_{1 \le j, l \le m}$ be the matrix
800: of coefficients in the system of equations \ref e.system/ for $a_{n+l}$,
801: considered as variables.
802: If $M_n$ is not singular, then 
803: $$
804: [a_{n+l}]_{1 \le l \le m} = M_n^{-1} [\zeta_j]_{1 \le j \le m}
805: \,.
806: $$
807: Now by \ref e.asymp/, the matrix $M_n$ tends to the Vandermonde
808: matrix determined by $z_1, \ldots, z_m$.
809: Hence, for all large $n$, we have not only that $M_n$ is nonsingular, but
810: also that its inverse has $\ell^2$-norm bounded by some constant $D$
811: that depends only on $\sigma$.
812: Therefore, for all large $n$, we have 
813: $$
814: \sum_{l=0}^m |a_{n+l}|^2
815: \le
816: D^2 \sum_{j=1}^m |\zeta_j|^2
817: \le
818: C D^2 m \sum_{k=0}^n |a_k|^2
819: \,.
820: $$
821: This clearly implies \ref e.goodonp/ for $\epsilon := 1/(1+C D^2 m)$
822: and finishes the proof for the case of simple zeroes.
823: 
824: Now suppose that the zero $z_j$ of $\sigma$ has multiplicity $r_j$, so that
825: $s := \sum_{j=1}^m r_j$ is the degree of $\sigma$.
826: Multiply \ref e.asymp/ by $z^p$ and
827: take the $r$th derivative (\ref b.Rudin:RCA/, Theorem 10.28, p.~214)
828: to obtain that
829: $$
830: \lim_{k \to\infty} 
831: \qonp_{k, p}^{(r)}(z) 
832: =
833: z^{p-r} \prod_{t=0}^{r-1} (p-t)
834: \quad\hbox{ for } |z| > 1
835: \,,
836: \label e.asymp-der
837: $$
838: where
839: $$
840: \qonp_{k, p}^{(r)}(z) :=
841: \left({d \over dz}\right)^r \left( z^{-k+p} \tilonp_k(z) \right)
842: \,.
843: $$
844: Since $z \mapsto z^{-n-1} \sigma(z) S(z)
845: \left(\overline{\Phi_w(\overline z^{-1})}\right)^{-1}$
846: has a zero at $z_j$ of order at least $r_j$, it follows that
847: $$
848: \sum_{l=1}^s a_{n+l} \qonp^{(r)}_{n+l, l-1}(z_j)
849: =
850: -\sum_{k=0}^n a_{k} \qonp^{(r)}_{k, l-1}(z_j)
851: $$
852: for $0 \le r < r_j$ and $1 \le j \le m$.
853: We may now follow the same reasoning as for the case of simple zeroes, but
854: instead of finding a coefficient matrix tending to a Vandermonde matrix,
855: we find instead a limit matrix that has $r_j$ columns corresponding to each
856: $z_j$, namely, for each $r = 0, 1, \ldots, r_j - 1$, it has the column
857: $$
858: \left[z_j^{l-1-r} \prod_{t=0}^{r-1} (l-1-t) \right]_{1 \le l \le s}
859: \,.
860: $$
861: Thus, by reasoning analogous to before, it suffices to establish that these
862: columns form a nonsingular matrix.
863: To do this, we show that there is no nontrivial linear relation among the
864: rows.
865: Indeed, if $b_1, \ldots, b_s$ are constants such that for $1 \le j \le m$
866: and $0 \le r < r_j$, 
867: $$
868: \sum_{l=1}^s b_l z_j^{l-1-r} \prod_{t=0}^{r-1} (l-1-t) = 0
869: \,,
870: $$
871: then the polynomial $\sum_{l=1}^s b_l z^{l-1}$ has a zero at $z_j$ of order
872: at least $r_j$ for each $1 \le j \le m$. But since this polynomial has
873: degree at most $s-1$, this implies that the polynomial is identically zero,
874: i.e., all $b_l = 0$, as desired.
875: \Qed
876: 
877: 
878: 
879: 
880: It would be interesting to have a good characterization of those $\varphi,
881: \psi$ such that \ref e.good/ holds.
882: 
883: 
884: 
885: 
886: 
887: 
888: 
889: 
890: \bsection{Higher Dimensions}{s.high}
891: 
892: The technology we use to replace the theory of Hardy spaces for higher
893: dimensions was provided by \ref b.HelLow/, who proved the extension of \ref
894: t.first/ to higher dimensions. We review the relevant definitions and facts
895: from their theory before giving our theorems, which extend Theorems \briefref
896: t.nonneg/ and \briefref t.cplx/. Fix a positive integer
897: $d$ and let $\lambda$ be Lebesgue measure on $\T^d := \R^d/\Z^d$.
898: 
899: For $k \in \Z^d$ and $x \in \T^d$, let $e_k(x) := e^{2\pi i k \cdot x}$.
900: For $f \in L^2(\T^d)$, write $\widehat f(k) := (f, e_k)_\lambda := \int_{\T^d}
901: f \overline{e_k} \,d\lambda$.
902: Let $S \subset \Z^d$ have the properties $S \cup (-S) = \Z^d \setminus \{ \bfz
903: \}$, $S \cap (-S) = \emptyset$, and $S+S \subset S$. The associated {\bf
904: ordering} of $\Z^d$ is that where $k \prec l$ iff $l - k \in S$.
905: For example, we could have $(k_1, k_2, \ldots, k_d) \prec \bfz$
906: if $k_i < 0$ when $i$ is the first index such that $k_i \ne 0$, which we
907: call the {\bf lexicographic ordering}.
908: The replacement for the Hardy spaces $H^p(\T)$ ($p \ge 1$) are the {\bf
909: Helson-Lowdenslager spaces}
910: $$
911: \hl^p := \hl^p(\T^d, S) := \Big \{ \varphi \in L^p(\T^d) \st \supp \widehat
912: \varphi \subset S \cup \{ \bfz \} \Big \}
913: \,.
914: $$
915: Let $P_{\hl^2} : L^2(\T^d) \to \hl^2$ be the orthogonal projection $\sum_{k
916: \in \Z^d} a_k e_k \mapsto \sum_{k \in S \cup \{ \bfz \}} a_k e_k$.
917: For $0 \le f \in L^1(\lambda)$ and any set $R \subseteq \Z^d$,
918: let $[R]$ be the linear span of $ \{ e_k \st k \in R\}$ and $[R]_f$ be
919: its closure in $L^2(f)$.
920: In place of outer functions, we use {\bf spectral factors},
921: which are the functions $\varphi \in \hl^2$ with the properties 
922: %\supp \widehat \varphi &\subset S \cup \{ \bfz \}
923: %\,,
924: %\label e.support
925: %\cr
926: $$
927: \widehat\varphi(\bfz) > 0
928: \label e.at0
929: $$
930: and
931: $$
932: 1/\varphi \in [S \cup \{ \bfz \}]_{|\varphi|^2}
933: \,.
934: \label e.inverse
935: $$
936: \ref b.HelLow/ show that for $0 \le f \in L^1(\T^d)$, the condition $\GM(f) >
937: 0$ is equivalent to the existence of a spectral factor $\varphi$ such that
938: $|\varphi|^2 = f$. (More precisely, they prove $\GM(f) > 0$ iff $\exists
939: \varphi \in \hl^2$ satisfying %\ref e.support/ and
940: \ref e.at0/. Their proof
941: shows that in this case, $\varphi$ can be chosen so that also \ref e.inverse/
942: holds.)
943: 
944: \procl l.divide
945: Suppose that $\varphi$ satisfies \ref e.inverse/, $h \in \hl^1$, and
946: $h/\varphi \in L^2(\T^d)$. Then $h/\varphi \in \hl^2$.
947: \endprocl
948: 
949: \proof
950: Let $f := |\varphi|^2$. By \ref e.inverse/, there exist trigonometric
951: polynomials $p_n$ with $\supp \widehat {p_n} \subset S \cup \{ \bfz \}$ and
952: such that $p_n \to 1/\varphi$ in $L^2(f)$.
953: Since $h/\varphi \in L^2(\T^d)$, we have $h/f \in L^2(f)$.
954: Thus, 
955: \begineqalno
956: \widehat{h/\varphi}(k)
957: &=
958: \int {h \over \varphi} \overline{e_k} \,d\lambda
959: =
960: \int {h \overline{e_k}\over f} {1\over \varphi} f \,d\lambda
961: =
962: \ip{h\overline{e_k}/f, 1/\overline\varphi}_f
963: =
964: \lim_{n \to\infty} \ip{h\overline{e_k}/f,  \overline{p_n}}_f
965: \cr&=
966: \lim_{n \to\infty} \int h p_n \overline{e_k}\,d\lambda
967: =
968: \lim_{n \to\infty} \widehat{h p_n}(k)
969: \,.
970: \cr
971: \endeqalno 
972: Since $h p_n \in \hl^1$, we have $\widehat{h p_n}(k) = 0$ for $k \notin S$, 
973: whence $\widehat{h/\varphi}(k) = 0$ for $k \notin S$.
974: That is, $h/\varphi \in \hl^2$.
975: \Qed
976: 
977: For $A \subseteq \Z^d$, let $(A)$ denote the set of corresponding complex
978: exponentials $ \{ e_k \st k \in A \}$.
979: For any two finite ordered sets of functions $\F, \G \subset L^2(\mu)$
980: of the same cardinality, let 
981: $$
982: \bigip{\F, \G}_\mu :=
983: \det\big[\ip{p, q}_\mu\big]_{p \in \F , q \in \G }
984: \,.
985: $$
986: Here, the ordering of the sets $\F, \G$ is used to order the rows and columns
987: of the matrix whose determinant appears in this equation.
988: Also, we write $\ip{p, q}_\mu := \int p \overline q \,d\mu$ even if $\mu$
989: is a complex measure.
990: Write $\F \Cup \G$ for the set $\F \cup \G$ ordered by concatenating
991: $\G$ after $\F$.
992: 
993: We are now ready to state and prove our extension of \ref t.nonneg/.
994: 
995: \procl t.nonneg-d
996: Let $w : \T^d \to \CO{0,\infty}$ be measurable with
997: $\GM(w) > 0$. Let $\varphi$ be a spectral factor for $w$.
998: Given any two finite ordered sets of functions $\F, \G \subset L^2(w)$
999: of the same cardinality, let $\F' := \Seq{ P_{\hl^2}(f \overline \varphi) \st
1000: f \in \F }$ and define $\G'$ likewise.
1001: Let $S_n \subset -S$ be finite ordered sets increasing to $-S$.
1002: We have 
1003: $$
1004: \lim_{n \to\infty}
1005: {\bigip{\F \Cup (S_n) , \G \Cup (S_n)}_w
1006: \over
1007: \bigip{(S_n), (S_n)}_w}
1008: =
1009: \bigip{\F', \G'}_\lambda
1010: \,.
1011: \label e.past
1012: $$
1013: \endprocl
1014: 
1015: The case $\F = \G = \Seq{\constant 1}$ is the theorem of \ref b.HelLow/.
1016: Actually, there are special considerations in that case that allow
1017: \refbauthor{HelLow} to make the same conclusion when $w = [d\mu/d\lambda]$
1018: and the matrix entries are given by inner products in $L^2(\mu)$.
1019: 
1020: \proof
1021: Let $P_n$ be the orthogonal projection onto $(S_n)^\perp$ in $L^2(w)$.
1022: Also, let $P_\infty$ be the
1023: orthogonal projection of $L^2(w)$ onto 
1024: $$
1025: H_\infty := (-S)^\perp
1026: =
1027: \{ g \in L^2(w) \st \all {n \in -S} \ip{g, e_n}_w = 0 \}
1028: \,.
1029: $$
1030: Define $\F'_n := \Seq{P_n f \st f \in \F}$ and likewise for $\G'_n$,
1031: $\F'_\infty$, and $\G'_\infty$.
1032: By row and column operations, we have 
1033: $$
1034: {\bigip{\F \Cup (S_n) , \G \Cup (S_n)}_w
1035: \over
1036: \bigip{(S_n), (S_n)}_w}
1037: =
1038: \bigip{\F_n', \G_n'}_w
1039: \,.
1040: $$
1041: This therefore tends to
1042: $\bigip{\F_\infty', \G_\infty'}_w$ (indeed, we have entry-wise convergence
1043: of the corresponding matrices).
1044: Now $g \in H_\infty$ iff $g \in L^2(w)$ and $\supp \widehat{g w} \subseteq
1045: S \cup \{ \bfz \}$.
1046: Thus, 
1047: $$
1048: H_\infty 
1049: =
1050: \{ g \in L^2(w) \st g w \in \hl^1 \}
1051: \,.
1052: $$
1053: 
1054: Let $f \in L^2(w)$ and
1055: write $h := w \cdot P_\infty f$.
1056: Since $h/w \in H_\infty$, we have $h/w \in L^2(w)$ and $h \in \hl^1$.
1057: {}From the first of these relations, we see that $h/\varphi \in L^2(\T^d)$.
1058: Also, 
1059: $$
1060: \all {g \in H_\infty}\quad  \bigip{h/w, g}_w = \ip{f, g}_w
1061: \,.
1062: \label e.whatyouget
1063: $$
1064: For all $m \in S \cup \{ \bfz \}$, we have $e_m \varphi \in \hl^2 \subset
1065: \hl^1$, so that $e_m /\overline \varphi \in H_\infty$.
1066: Therefore, \ref e.whatyouget/ implies that
1067: $$
1068: \bigip{h/w, e_m/\overline\varphi}_w = \ip{f, e_m/\overline\varphi}_w
1069: \,,
1070: $$
1071: or in other words, 
1072: $
1073: \widehat{(h/\varphi)}(m)
1074: =
1075: \widehat{f\overline\varphi}(m)
1076: $ for all $m \in S \cup \{ \bfz \}$.
1077: Since $h/\varphi \in \hl^2$ by \ref l.divide/, it follows that
1078: $h/\varphi = P_{\hl^2}(f \overline\varphi)$.
1079: Thus, we have proved that for all $f$, we have 
1080: $$
1081: P_\infty f = {P_{\hl^2}(f \overline\varphi) \over \overline\varphi}
1082: \,.
1083: \label e.projform
1084: $$
1085: 
1086: This gives \ref e.past/ since 
1087: $$
1088: \ip{P_\infty f, P_\infty g}_w 
1089: =
1090: \int P_\infty f \cdot \overline{P_\infty g} \cdot w\,d\lambda
1091: =
1092: \int P_{\hl^2}(f \overline\varphi) \overline{P_{\hl^2}(g \overline\varphi)}
1093: \,d\lambda
1094: \,.
1095: \Qed
1096: $$
1097: 
1098: Unlike in \ref t.nonneg/, the limit \ref e.past/ is not necessarily 0 when
1099: $\GM(w) = 0$. For example, let $S$ be the lexicographic ordering
1100: on $\Z^2$ and $w(x_1, x_2)$ be a function that depends only on $x_2$ with
1101: $\GM(w) = 0$.  If $\F := \G := \{ e_{(1, 0)} \}$, then the left-hand side of
1102: \ref e.past/ is equal to $\widehat w(\bfz)$, which need not equal 0.
1103: 
1104: However, if the order is archimedean, such as if $S = \{ k \in \Z^d \st k
1105: \cdot x > 0 \}$, where $x \in \R^d$ has at least two coordinates whose
1106: quotient is irrational, then the limit \ref e.past/ is 0 when $\GM(w) = 0$,
1107: as we now show.
1108: (All archimedean orders arise in this way; in fact, for a characterization of
1109: all orders, see \ref b.Teh/, \ref b.Zaiceva/, or \ref b.Trevisan/.)
1110: 
1111: \procl p.GM0
1112: Let $w : \T^d \to \CO{0,\infty}$ be measurable with $\GM(w) = 0$. 
1113: Suppose that the order induced by $S$ is archimedean.
1114: Given any two finite ordered sets of functions $\F, \G \subset L^2(\mu)$
1115: of the same cardinality,
1116: and any finite ordered sets $S_n \subset -S$ increasing to $-S$,
1117: we have 
1118: $$
1119: \lim_{n \to\infty}
1120: {\bigip{\F \Cup (S_n) , \G \Cup (S_n)}_w
1121: \over
1122: \bigip{(S_n), (S_n)}_w}
1123: =
1124: 0
1125: \,.
1126: $$
1127: \endprocl
1128: 
1129: \proof
1130: It suffices to establish that $H_\infty = 0$ in the notation of the proof of
1131: \ref t.nonneg-d/.
1132: Now for all $g \in H_\infty$, 
1133: $$
1134: \GM(|g w|)^2
1135: =
1136: \GM(|g w|^2)
1137: =
1138: \GM(|g|^2 w) \GM(w) 
1139: =
1140: 0
1141: $$
1142: since $\GM(|g|^2 w) \le \int |g|^2 w \,d\lambda < \infty$ as $g \in L^2(w)$.
1143: Since $g w \in \hl^1$, this means that $g w = \constant 0$ by the main result
1144: of \ref b.arens/. 
1145: Thus, $H_\infty = 0$.
1146: \Qed
1147: 
1148: \procl r.uniq
1149: Suppose that $\GM(w) > 0$ and that $\varphi$ is a spectral factor for $w$.
1150: By \ref e.projform/, we have $w P_\infty \bfo = \varphi
1151: P_{\hl^2}(\overline\varphi) = \overline{\widehat\varphi(\bfz)} \varphi$, so
1152: that $\varphi$ is uniquely determined by $w$.
1153: It is essentially by this formula that \ref b.HelLow/ proved the existence
1154: of a spectral factor.
1155: This method goes back to \ref b.Szego:outer/.
1156: \endprocl
1157: 
1158: The extension of \ref t.cplx/ is relatively straightforward:
1159: 
1160: \procl t.cplx-d
1161: Let $S_n \subset -S$ be finite ordered sets increasing to $-S$.
1162: Let $\mu = \psi\overline\varphi\lambda$ for some pair of spectral factors
1163: $\varphi, \psi$ that satisfy the condition
1164: $$
1165: \liminf_{n \to\infty} \epsilon\big(\varphi \cdot [-S_n], \psi
1166: \cdot [-S_n]; L^2(\lambda)\big) > 0
1167: \,.
1168: \label e.good-d
1169: $$
1170: Given any two finite ordered sets of functions
1171: $\F, \G \subset L^2(|\varphi|^2 + |\psi|^2)$
1172: of the same cardinality, let $\F' := \Seq{ P_{\hl^2}(f \overline \varphi) \st
1173: f \in \F }$ and $\G' := \Seq{ P_{\hl^2}(g \overline \psi) \st
1174: g \in \G }$.
1175: We have 
1176: $$
1177: \lim_{n \to\infty}
1178: {\bigip{\F \Cup (S_n) , \G \Cup (S_n)}_\mu
1179: \over
1180: \bigip{(S_n), (S_n)}_\mu}
1181: =
1182: \bigip{\F', \G'}_\lambda
1183: \,.
1184: \label e.past
1185: $$
1186: \endprocl
1187: 
1188: \proof
1189: Let $H_n(\varphi) := \{ \varphi e_k \st k \in -S_n \}$.
1190: By virtue of \ref e.good-d/, we have for $n \ge n_0$,
1191: $$
1192: \overline{H_n(\varphi)} \cap \overline{H_n(\psi)}^\perp = 0
1193: \,,
1194: $$
1195: and so
1196: $$
1197: L^2(\T^d) = \overline{H_n(\varphi)} + \overline{H_n(\psi)}^\perp
1198: \,.
1199: $$
1200: We claim that
1201: $$
1202: H_n(\varphi) \uparrow \hl^2_0 := \left(\overline{\hl^2}\right)^\perp
1203: \,.
1204: \label e.beurling-d
1205: $$
1206: Indeed, if $g \in \hl^2_0$ and $g \perp H_n(\varphi)$
1207: for all $n \ge 0$, then $\widehat{g \overline\varphi}(k) = 0$ for all $k \in
1208: S$, i.e., $g \overline\varphi \in \overline{\hl^1}$.
1209: By \ref l.divide/, we may divide by $\overline\varphi$ to obtain that $g \in
1210: \overline{\hl^2} = (\hl^2_0)^\perp$, so that $g = \constant 0$.
1211: This proves \ref e.beurling-d/.
1212: 
1213: Thus the projection along $\overline{H_n(\varphi)}$ to
1214: $\overline{H_n(\psi)}^\perp$ tends to the orthogonal projection
1215: $P_{\overline{\hl^2_0}^\perp} = P_{\hl^2}$.
1216: 
1217: Now $\ip{f, g}_\mu = \ip{\overline\varphi f, \overline\psi g}_\lambda$ for any
1218: $f, g \in L^2(|\varphi|^2+|\psi|^2)$.
1219: Let $\F_n$ be the image of $\F$ under
1220: the projection along
1221: $\overline{H_n(\varphi)}$ to $\overline{H_n(\psi)}^\perp$.
1222: Let $\G'' := \{ \overline\psi g \st g \in \G \}$.
1223: Row operations show that for $n \ge n_0$, 
1224: $$
1225: {\bigip{\F \Cup (S_n) , \G \Cup (S_n)}_\mu
1226: \over
1227: \bigip{(S_n), (S_n)}_\mu}
1228: =
1229: \bigip{\F_n, \G''}_\lambda
1230: \,.
1231: $$
1232: Because of our assumption \ref e.good-d/ and \ref l.projection/, the limit is
1233: $
1234: \bigip{\F', \G''}_\lambda
1235: $,
1236: which is the same as
1237: $
1238: \bigip{\F', \G'}_\lambda
1239: $.
1240: \Qed
1241: 
1242: 
1243: \bsection{Appendix}{s.app}
1244: 
1245: In order to prove \ref l.projection/, we first demonstrate the following
1246: lemma.
1247: 
1248: 
1249: \procl l.proj1
1250: Suppose that $H$ is a Hilbert space, $H_1$ and $K_1$ are non-zero
1251: closed subspaces, $H = H_1 + K_1^\perp$, and $H_1 \cap K_1^\perp = 0$.
1252: Let $T : H \to K_1^\perp$ be the linear projection along $H_1$
1253: and 
1254: $$
1255: \epsilon := \epsilon(H_1, K_1)
1256: \,.
1257: $$
1258: Then 
1259: $$
1260: \|T\| \le {1 \over 2} + \sqrt{{1 \over 2(1 - \sqrt{1 - \epsilon^2})} + {1
1261: \over 4}}
1262: \label e.bddT
1263: $$
1264: and
1265: $$
1266: \epsilon \ge {1 \over \sqrt{1 + \|T\|^2}}
1267: \,.
1268: \label e.bddepsi
1269: $$
1270: \endprocl
1271: 
1272: \proof
1273: Let $v \in H$ with $\|v\| = 1$.
1274: Write $w := T v \in K_1^\perp$ and $u := v - w \in H_1$.
1275: Choose $y \in K_1$ such that $\ip{u, y} \ge \epsilon \|u\| \|y\|$ and
1276: $\|y\| = \epsilon \|u\|$.
1277: We have 
1278: \begineqalno
1279: 1 
1280: &=
1281: \|v\|^2
1282: =
1283: \|u + w\|^2
1284: =
1285: \|u\|^2 + \|w\|^2 + 2 \Re \ip{u, w}
1286: \cr&=
1287: \|u\|^2 + \|w\|^2 + 2 \Re \ip{u - y, w}
1288: \ge
1289: \|u\|^2 + \|w\|^2 - 2 \|u - y\| \| w\|
1290: \cr&=
1291: \|u\|^2 + \|w\|^2 - 2 \| w\| \left[ \|u\|^2 + \|y\|^2 - 2 \Re\ip{u, y}
1292: \right]^{1/2}
1293: \cr&\ge
1294: \|u\|^2 + \|w\|^2 - 2 \| w\| \left[ \|u\|^2 + \|y\|^2 - 2 \epsilon \|u \|
1295: \|y \| \right]^{1/2}
1296: \cr&=
1297: \|u\|^2 + \|w\|^2 - 2 \| w\| \|u\| \sqrt{1- \epsilon^2}
1298: \cr&=
1299: \left(\|u\| - \|w\|\right)^2 + 2 \| w\| \|u\| (1 - \sqrt{1- \epsilon^2})
1300: \cr&\ge
1301: 2 \| w\| \|u\| (1 - \sqrt{1- \epsilon^2})
1302: =
1303: 2 \| w\| \|v - w\| (1 - \sqrt{1- \epsilon^2})
1304: \cr&\ge
1305: 2 \| w\| (\|w\| - 1) (1 - \sqrt{1- \epsilon^2})
1306: \,.
1307: \endeqalno
1308: Simple algebra shows that this inequality implies 
1309: $$
1310: \|T v\| = \|w\|
1311: \le {1 \over 2} + \sqrt{{1 \over 2(1 - \sqrt{1 - \epsilon^2})} + {1
1312: \over 4}}
1313: \,.
1314: $$
1315: Since $\|v\| = 1$, this is equivalent to \ref e.bddT/.
1316: 
1317: To prove \ref e.bddepsi/, note that by definition of $\epsilon$, we have
1318: that
1319: $$
1320: \texists {x \in H_1 \setminus \{ 0 \}} \all {y \in
1321: K_1} \quad |\ip{x, y}| \le \epsilon \|x\| \|y\|
1322: \,.
1323: \label e.genbad
1324: $$
1325: Choose such an $x$ with $\|x\|= 1$ and set $y := P_{K_1} x$, the orthogonal
1326: projection of $x$ onto $K_1$.
1327: Then $y \ne 0$ because $H_1 \cap K_1^\perp = 0$.
1328: Write $z := x - y \in K_1^\perp$.
1329: Since $y = x - z$ with $x \in H_1$ and $z \in K_1^\perp$, it follows that
1330: $T y = -z$.
1331: By \ref e.genbad/ applied to $y$, we have 
1332: \begineqalno
1333: \|z\|^2
1334: &=
1335: \|x - y\|^2
1336: =
1337: 1 + \|y\|^2 - 2 \Re \ip{x, y}
1338: \cr&\ge
1339: 1 + \|y\|^2 - 2 \epsilon \|y\|
1340: \label e.interm
1341: \cr&=
1342: 1 + (\|y\| - \epsilon)^2 - \epsilon^2
1343: \cr&\ge
1344: 1 - \epsilon^2
1345: \,.
1346: \endeqalno
1347: Furthermore, $\|z\|^2 = 1 - \|y\|^2$, whence comparison to \ref e.interm/
1348: shows that $\|y\| \le \epsilon$.
1349: Therefore, 
1350: $$
1351: \|T\| 
1352: \ge
1353: \|T y\|/\|y\|
1354: =
1355: \|z\|/\|y\|
1356: \ge
1357: \sqrt{1 - \epsilon^2}/\epsilon
1358: \,,
1359: $$
1360: which is equivalent to \ref e.bddepsi/.
1361: \Qed
1362: 
1363: 
1364: 
1365: \proofof l.projection
1366: For each $n$ and any $v \in H_n$, we have $T_m v = 0 = T_\infty v$ for all
1367: $m \ge n$.
1368: Also, for each $v \in K_\infty^\perp$, we have $T_m v = v = T_\infty v$ for
1369: all $m$.
1370: Therefore, $T_n v \to T_\infty v$ for all $v$ belonging to the dense
1371: set $\bigcup_n H_n + K_\infty^\perp$.
1372: It follows by continuity and the principle of uniform boundedness that $T_n
1373: \to T_\infty$ in the strong operator topology iff 
1374: $$
1375: \sup_n \|T_n\| <\infty
1376: \,.
1377: \label e.unifbdd
1378: $$
1379: If \ref e.gengood/ holds, then \ref e.unifbdd/ is a consequence of \ref
1380: e.bddT/, while if \ref e.unifbdd/ holds, then \ref e.gengood/ is a
1381: consequence of \ref e.bddepsi/.
1382: \Qed
1383: 
1384: 
1385: 
1386: 
1387: \medbreak
1388: \noindent {\bf Acknowledgements.}\enspace
1389: I am grateful to Doron Lubinsky for allowing me to include \ref e.goodonp/
1390: and to Jeff Geronimo for a reference.
1391: 
1392: 
1393: 
1394: \bibfile{\jobname}
1395: \def\noop#1{\relax}
1396: \input \jobname.bbl
1397: 
1398: \filbreak
1399: \begingroup
1400: \eightpoint\sc
1401: \parindent=0pt\baselineskip=10pt
1402: 
1403: Department of Mathematics,
1404: Indiana University,
1405: Bloomington, IN 47405-5701
1406: \emailwww{rdlyons@indiana.edu}
1407: {http://php.indiana.edu/\string~rdlyons/}
1408: 
1409: and
1410: 
1411: School of Mathematics,
1412: Georgia Institute of Technology,
1413: Atlanta, GA 30332-0160
1414: \email{rdlyons@math.gatech.edu}
1415: 
1416: \endgroup
1417: 
1418: 
1419: \bye
1420: 
1421: 
1422: 
1423: