1: \section{Analysis} \label{DGsecana}
2: In this section we discuss two of the basic processes of analysis:
3: pull-back and push-forward, and how they affect asymptotic behavior of smooth
4: functions (as discussed in the previous section) and conormal
5: distributions, which we also introduce.
6:
7: What are pull-back and push-forward, and why are they important?
8: Pull-back is composition, push-forward is integration. They are important
9: since they may be used as building blocks for other operations.
10: This allows to carry out recurring ugly calculations (e.g.\ those
11: involving Fourier transform) once in the proof of theorems about pull-back
12: and push-forward, and then never look at them again.
13: Let us illustrate this in two simple but central examples:%
14: \footnote{
15: In the examples, $\R$ may be replaced by any manifold, equipped with a
16: fixed density.
17: For the moment we naively neglect the distinction between functions,
18: distributions and the respective densities; also, we neglect such
19: tedious matters as integrability.
20: }
21: \begin{examples}\label{DGexpfpb}
22: \
23: \begin{description}
24: \item [Applying an operator to a function]
25: Let $\pi_1,\pi_2:\R^2\to\R$ be the projections onto the first and second
26: coordinate. If $v$ is a function on $\R$ then its pull-back
27: $\pi_2^*v=v\circ\pi_2$ is the function $(x,y)\mapsto v(y)$ on $\R^2$.
28: If $u$ is a function on $\R^2$ then its push-forward $\pi_{1*}u$ is the
29: function
30: \begin{equation}\label{DGeqpf1}
31: \pi_{1*}u(x) = \int u(x,y) dy.
32: \end{equation}
33: If $A$ is an operator, acting on functions $v$ on $\R$, with
34: integral kernel $A(x,y)$ then
35: \begin{equation}\label{DGeqaction}
36: (Av)(x) = \int A(x,y)v(y)\, dy = \pi_{1*}(A\cdot\pi_2^*v).
37: \end{equation}
38: Though this may look like an exercise in
39: formal nonsense, it shows
40: that mapping properties of $A$ may be read off from the structure
41: (e.g.\ asymptotic type) of
42: the function (distribution) $A$, if one understands how such structure
43: is affected by pull-back and push-forward.%
44: \footnote{
45: Also, one needs to understand how structure is affected by
46: multiplication. This is trivial for nice functions, geometrically non-trivial
47: for functions with different asymptotic types, and analytically non-trivial
48: for distributions. See Subsection
49: \ref{DGsubsecdist} for the latter case.}
50:
51: \item[Composition of operators]
52: If $A,B$ are operators, acting on functions on $\R$, with
53: integral kernels $A(x,y),B(y,z)$, then
54: $C=A\circ B$ has integral kernel $C(x,z)=\int A(x,y)B(y,z)\,dy$, i.e.
55: \begin{equation}\label{DGeqcomp}
56: C=\pi_{2*}(\pi_3^*A\cdot\pi_1^*B)
57: \end{equation}
58: where $\pi_1,\pi_2,\pi_3:\R^3\to\R^2$ are the projections leaving out
59: the first, second and third variable, respectively.
60: Again, understanding how pull-back, push-forward and product
61: affect the structure
62: of distributions allows to predict, for example, whether a class of operators
63: with a given structure of its kernels is closed under composition.
64: \end{description}
65: \end{examples}
66:
67: Another place where one needs to understand the behavior of distributions
68: under push-forward is in the 'specializations' mentioned in
69: Footnote \ref{DGfnspec} in the Introduction, since many of them are obtained
70: from the full kernel of $Q$ by integration (i.e.\ push-forward).
71:
72: The maps used for pull-back and push-forward in the examples
73: are rather trivial projections,
74: so it's legitimate to ask: Why be so formal, why
75: not talk simply of 'integration in $y$' instead of 'push-forward by $\pi_1$'?
76: The answer is given by:
77: \begin{principle}
78: The push-forward of a complicated function by a simple map should be
79: analyzed by rewriting it as push-forward of a simple function by a complicated
80: map.
81: \end{principle}
82: The point is that the 'complication' of the map lies mainly in its global
83: geometry, so by a partition of unity the problem can be reduced to the sum of
84: relatively simple local problems. (In contrast, the 'complication' of the
85: function is local, typically.)
86: Melrose's Push-Forward Theorem gives the result of this analysis, for the
87: case of smooth functions.
88: In Subsection \ref{DGsubsecpf} we discuss all these matters, starting from
89: an example.
90: We also sketch the idea of a proof of the Push-Forward Theorem in the
91: special case that the target space is $\Rplus$.
92:
93: When integrating one needs measures. Therefore, push-forward
94: is best defined as acting on measures (or densities) rather than
95: functions. The
96: push-forward of a smooth density may be not smooth, and (what's equivalent)
97: the pull-back of a distribution is not always defined.
98: For the reader unfamiliar with these matters,
99: we give the precise definitions and a short discussion
100: of pull-back and push-forward, and how they act on smooth functions,
101: distributions and smooth and distributional densities, in the Appendix.%
102: \footnote{\label{DGfndens}%
103: The reader who prefers to neglect the distinction between
104: functions and densities is invited to do so, but will probably begin
105: to acknowledge their usefulness when making computations herself.
106: }
107:
108: In Subsection \ref{DGsubsecpb} we state Melrose's
109: Pull-Back Theorem, which tells how pull-back by a \Mb-map affects nice
110: functions. This is rather trivial in comparison to the Push-Forward Theorem.
111:
112:
113: Finally, in Subsection \ref{DGsubsecdist} we introduce conormal
114: distributions and discuss how pull-back,
115: push-forward and multiplication affect them. As an
116: illustration, we define pseudodifferential operators and
117: study their composition.
118:
119:
120: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
121: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Push-Forward Subsec.
122: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
123:
124: \subsection{Push-forward and asymptotic type}\label{DGsubsecpf}
125: We begin by analyzing a few examples of push-forward under the
126: projection $\R_+^2\to\R_+,(x,y)\mapsto x$.
127: In other words, we set
128: \begin{equation}\label{DGeqpft}
129: \utilde(x) = \int_0^\infty u(x,y)\,dy,\quad x>0.
130: \end{equation}
131: Assuming that $u$ is smooth in $(0,\infty)^2$ and $\supp u$ is bounded,
132: we ask how the behavior of $u$ near the boundary of $\R_+^2$ affects the
133: behavior of $\utilde$ near 0.
134: \begin{examples}\label{DGexpft}
135: \
136: \begin{enumerate}
137: \item
138: If $u$ is smooth up to the boundary then so is $\utilde$ (by
139: first-year analysis).
140: More generally (and just as easy),
141: $$ u \text{ nice with index family }(E,F)
142: \Rightarrow \utilde\text{ nice with index set }F$$
143: if the integral \eqref{DGeqpft} exists at all, i.e.\ if
144: \begin{equation}\label{DGeqint1}
145: \Re z>-1 \text{ for } (z,p)\in E.
146: \end{equation}
147: \item
148: If $u(x,y)=y^{-1}v(x/y,y)$ with $v$ smooth on $\R^2_+$ and compactly supported then
149: \begin{equation}\label{DGeqhyperbel}
150: \utilde(x)=\int_0^\infty v(\frac{x}y,y)\,\frac{dy}y
151: \sim \sum_{i=0}^\infty (a_ix^i + b_ix^i\log x)\quad\text{ as }x\to 0,
152: \end{equation}
153: i.e., $\utilde$ is nice, but not smooth (the index set is
154: $\N_0\times\{0,1\}$).%
155: \footnote{
156: Proof: Taylor expand $v(\xi,\eta)$ at $\xi=0$ (for each fixed $\eta$), then
157: Taylor expand each coefficient and the remainder at $\eta=0$ to obtain, for
158: any $N$,
159: \begin{equation}\label{DGeqexpansion}
160: v(\xi,\eta) = \sum_{\alpha=0}^{N-1} \xi^\alpha \eta^Na_\alpha(\eta)
161: + \sum_{\beta=0}^{N-1} \eta^\beta \xi^N b_\beta(\xi)
162: + \sum_{\alpha,\beta=0}^{N-1} c_{\alpha,\beta} \xi^\alpha \eta^\beta
163: + \xi^N \eta^N r(\xi,\eta)
164: \end{equation}
165: with $a_\alpha,b_\beta,r$ smooth up to the boundary. Assume $\supp v
166: \subset [0,C]^2$. Then in the integral \eqref{DGeqhyperbel}
167: one may replace $\int_0^\infty$ by $\int_{x/C}^1$. To obtain the asymptotics,
168: simply integrate \eqref{DGeqexpansion} term by term,
169: using the substitution $z=x/y$ in the second sum.
170:
171: Note that the $\log$-terms only come from the terms $\alpha=\beta$ in
172: the third sum.
173: }
174: More generally, if $v$ has index sets $(E,F)$ then $\utilde$ is
175: nice with index set%
176: \footnote{
177: Same proof, after the (non-trivial) analysis argument that our
178: definition \eqref{DGdefnice} of niceness
179: implies an expansion like \eqref{DGeqexpansion}. Alternatively, one may define
180: niceness by this expansion.
181: }
182: \footnote{\label{DGfndyy}%
183: Why did we write the integral \eqref{DGeqhyperbel} with
184: $dy/y$ instead of simply $dy$? Since then the result
185: \eqref{DGeqextunion} is beautifully symmetric!
186: Cf. '\Mb-densities' below.
187: }
188: \begin{equation} \label{DGeqextunion}
189: E\extunion F := E\cup F \cup \{(z,p'+p''+1):\, (z,p')\in E, \, (z,p'')\in F\}.
190: \end{equation}
191:
192: \item
193: For $u(x,y)=\sqrt{x^2+y^2}$ explicit integration shows (restricting
194: to $y\leq 1$ for integrability -- this does not affect the essential point)
195: \begin{equation}\label{DGeqlog}
196: \utilde(x)= \text{ (smooth near zero) } - \frac12 x^2\log x.
197: \end{equation}
198: Thus, $\utilde$ is nice, and again a logarithm appears.
199: \end{enumerate}
200: \end{examples}
201: The common feature of Examples 2 and 3 is that $u$ has asymptotic type
202: $\beta$, where $\beta$ is the blow-up of 0 in $\Rplus^2$.
203: \medskip
204:
205: {\em Claim: This already suffices to explain the similarity of the results
206: \eqref{DGeqhyperbel} and \eqref{DGeqlog}.}
207:
208: \proof
209: We first show this by a simple calculation and then explain how it
210: may be seen directly by 'looking at pictures'.
211:
212: {\em Calculation: } Consider any $u$ of type $\beta$, so that $w=\beta^*u$
213: is nice on $W=[\Rplus^2,0]$, and assume $w$ has no
214: logarithmic terms in its expansions.
215: We split up the integral
216: $$
217: \int_0^\infty\!\! u(x,y)\,dy = A+B
218: $$
219: 'smoothly near $y=x$'. That is, with any cut-off function $\phi\in
220: C_0^\infty(\Rplus)$ which equals one near 0, and with $\psi=1-\phi$, we set
221: \begin{align*}
222: A=\int_0^\infty\!\! u(x,y)\,\phi(y/x)\,dy &
223: = \int_0^\infty\!\! x u(x,x\eta_1)\,\phi(\eta_1)\,d\eta_1
224: = \int_0^\infty\!\! x w_1(x,\eta_1)\,\phi(\eta_1)\,d\eta_1\\
225: B=\int_0^\infty\!\! u(x,y)\,\psi(y/x)\,dy &
226: = \int_0^\infty\!\! w_2(x/y,y)\,\psi(y/x)y\,\frac{dy}y.
227: \\
228: \end{align*}
229: Here, $w_1$ is just $w$ expressed in projective coordinates near the point
230: $A$ in Figure \ref{DGfigpf}(a)
231: (i.e.\ $w_1(\xi_1,\eta_1)=u(\xi_1,\xi_1\eta_1)$ or $w_1 = \beta_1^*u$
232: with $\beta_1$ from (\ref{DGeqproj}c));
233: the integral $A$ is like Example 1 (with $u(\xi,\eta)=\xi w_1(\xi,\eta)
234: \phi(\eta)$).
235: Similarly, $w_2$ is just $w$ expressed in projective coordinates
236: near $B$ in Figure \ref{DGfigpf}(a) (i.e.\ $w_2(\xi_2,\eta_2)=u(\xi_2\eta_2,\eta_2)$
237: or $w_2=\beta_2^*u$); the integral $B$ is like Example 2
238: (with $v(\xi,\eta) = w_2(\xi,\eta)\psi(\xi^{-1})\eta$).
239:
240: Since by assumption $w_{1/2}$ have no $\log$'s in their expansions
241: and in Example 1 no logarithms are created, we conclude:
242: The log terms in Examples 2 and 3 are of the same nature.%
243: \footnote{
244: But we also see that for general $\beta$-singular $u$ infinitely many
245: $\log$ terms will appear. For $u(x,y)=\sqrt{x^2+y^2}$ only
246: one $\log$-term appears (see \eqref{DGeqlog}); this is due to the fact
247: that in $\sqrt{x^2+y^2}=y\sqrt{1+(x/y)^2}=x\xi_2^{-1}\sqrt{1+\xi_2^2}$
248: only one power of $x$ occurs.
249: Such fine points are lost under (regular) coordinate changes and therefore
250: invisible in the geometric setup of the Push-Forward Theorem.
251: }
252:
253: \begin{figure}[htbp]
254: \input{figpf.pstex_t}
255: \caption{Level lines for push-forward under $[\Rplus^2,(0,0)]\to\Rplus$}
256: \label{DGfigpf}
257: \end{figure}
258:
259: {\em Pictures: }
260: We now show how the same result can be 'seen' geometrically.
261: Since $\beta^*u=w$ and $\beta$ is a diffeomorphism in the interior, we have
262: $u=\beta_*w$, so
263: \begin{equation} \label{DGeqpfcomp}
264: \utilde = \pi_{1*}u = \pi_{1*}\beta_*w = f_*w
265: \end{equation}
266: with $f=\pi_1\circ\beta:W\to\Rplus$.
267: This says simply that $\utilde(x)$ equals the integral of $w$ over
268: the fiber $f^{-1}(x)$ for each $x$ (let's postpone the question of
269: measures for a moment). This is clear since the values of $w$ on
270: $f^{-1}(x)$ are precisely the values of $u$ on $\pi_1^{-1}(x)$, which are
271: integrated to obtain $\utilde(x)$.
272:
273: Some fibers of $f$ are shown in Figure \ref{DGfigpf}(a), some of $\pi_1$
274: in Figure \ref{DGfigpf}(c)
275: and some of $g(\xi,\eta)=\xi \eta$ in Figure \ref{DGfigpf}(b).
276: Pictorially, we see:%
277: \footnote{
278: This can be made precise by expressing $f$ in projective
279: local coordinates \eqref{DGeqproj}:
280: \begin{itemize}
281: \item On $W\setminus\lb$ ('near A') $f$ is expressed
282: as $f_1(\xi_1,\eta_1)=\xi_1$ (using $\beta_1$ in (\ref{DGeqproj}c)),
283: i.e.\ $f_1=\pi_1$,
284: \item on $W\setminus\rb$ ('near B') as $f_2(\xi_2,\eta_2)=\xi_2\eta_2$
285: (using $\beta_2$), i.e. $f_2=g$.
286: \end{itemize}
287: }
288: \begin{itemize}
289: \item Near A, Figure \ref{DGfigpf}(a) looks like Figure \ref{DGfigpf}(c),
290: \item near B, Figure \ref{DGfigpf}(a) looks like Figure \ref{DGfigpf}(b).
291: \end{itemize}
292:
293: Therefore, push-forward of $w$ by $f$ is the sum of push-forward
294: (of $w$ near $A$)
295: by $\pi_1$ and push-forward (of $w$ near $B$) by $g$, and this was
296: precisely the calculation above. This also explains why the cut-off
297: had to be chosen as a smooth function of $y/x$, see Figure \ref{DGfigpf}(a).
298:
299: In summary, we may say that the $\log$ terms in Examples 2 and 3 arise
300: from the fact that the fibers of $g$ and of $f=\pi_1\circ\beta$ approach the
301: corner as in Figure \ref{DGfigpf}(b) for $x\to 0$.
302: \qed
303: \medskip
304:
305: In Melrose's Push-Forward Theorem these considerations are
306: generalized to any \Mb-map $f:W\to Z$:
307: Under certain conditions on $f$,
308: it says that the push-forward of a nice density $\mu$
309: on $W$ is a nice density on $Z$, and computes the index sets of the
310: latter from the index sets of the former and the 'boundary geometry'
311: of $f$.
312:
313: The conditions on $f$ are best understood if we consider the special case
314: $Z=\Rplus$ first. Before we can state them, we need some definitions.
315:
316: From now on, we assume that all index sets satisfy
317: $(z,p)\in E\implies (z+1,p)\in E$, the condition ensuring coordinate
318: invariance of niceness, and that bhs's are embedded and connected
319: (see Definition~\ref{DGdefmwc} and Footnote~\ref{DGfnmwc}).
320:
321: \bigskip
322:
323: {\em Densities on manifolds with corners.}
324: If $W$ is a mwc, then a density on $W$ is, by definition, a density on the
325: interior $\Wint$ (concerning densities see the Appendix and
326: Footnote~\ref{DGfndens}). The notion of niceness carries over to densities
327: immediately, e.g.\ on $\Rplus^2$:
328: \begin{definition}
329: A density $\mu=u\, dxdy$ on $\Rplus^2$ is {\em nice} with index sets $E,F$
330: if $u$ is nice with index sets $E,F$.%
331: \footnote{
332: As usual, this should be checked for coordinate independence.
333: But only under coordinate changes $(x,y)\mapsto (\xtilde,\ytilde)$ for
334: which $\xtilde,\ytilde$ are still bdf's of the coordinate axes!
335: Cf.\ \ref{DGsubsubsecinv}.
336: }
337: \end{definition}
338:
339: The following is a slight variant in book-keeping, which makes lots of things
340: more transparent%
341: \footnote{\label{DGfnbdens}%
342: Examples:
343: \begin{enumerate}
344: \item
345: $\mu$ locally integrable $\Longleftrightarrow$ $\Re z>0$ whenever
346: $(z,p)\in E\cup F$
347: (rather than $-1$).
348: \item
349: The transformation under projective coordinates becomes especially simple:
350: Say $\xi_1=x,\eta_1=y/x$, then
351: \begin{equation}\label{DGeqbdenstrf}
352: \frac{dx}x\frac{dy}{y} = \frac{d\xi_1}{\xi_1}\frac{d\eta_1}{\eta_1}.
353: \end{equation}
354: \item See Footnote \ref{DGfndyy} after Example \ref{DGexpft}.2.
355: \end{enumerate}
356: }
357: (though it may seem artificial to the uninitiated):
358: \begin{definition} \label{DGdefbdens}
359: A {\em \Mb-density}
360: on $\Rplus^2$ is just a density, except that we write it as
361: $\mu=u(x,y) \frac{dx}x\frac{dy}y$ instead. When talking about smoothness or
362: the index family of $\mu$ then we mean smoothness or the index family
363: of $u$ in
364: such a representation.
365: \end{definition}
366:
367: Of course, a \Mb-density on $\Rplus\times\R$ is of the form
368: $u(x,y)\,\frac{dx}x dy$. That is, the $\frac{dx}x$ factor only occurs
369: in the variables $x$ defining some bhs. It is easy to see that the index
370: family of a \Mb-density is well-defined on any mwc.
371:
372: \bigskip
373:
374: {\em Boundary geometry of a \Mb-map $f:W\to \Rplus$.}
375: \begin{definition} \label{DGdefexp1}
376: Let $f:W\to\Rplus$ be a \Mb-map. For any bhs $G$ of $W$ define $e_f(G)$ to be
377: the order of vanishing of $f$ at $G$.
378: \end{definition}
379:
380: In other words,
381: in the local Definition \ref{DGdefbmap} with $w\in G$, we set
382: $e_f(G)=\alpha_{1j_0}$ if $f(w)=0$ and $x_{j_0}$ is a bdf for $G$, and
383: $e_f(G)=0$ if $f(w)\not=0$. This is clearly locally constant and therefore
384: constant on $G$ by connectedness, so $e_f(G)$ is well-defined.
385: Note that
386: \begin{equation} \label{DGequnion}
387: f^{-1}(0) = \bigcup \{G: e_f(G) > 0\}.
388: \end{equation}
389:
390: \begin{theorem}[Push-Forward Theorem, special case $Z=\Rplus$]
391: \label{DGthpft1}
392: Let $W$ be a manifold with corners and $f:W\to\Rplus$ a \Mb-map which is
393: a fibration over $(0,\infty)$.%
394: \footnote{
395: I.e.\ $f:f^{-1}((0,\infty))\to(0,\infty)$ is a fibration in the sense
396: of Footnote \ref{DGfnfibr} in the Appendix,
397: except that $L$ is allowed to be a mwc.}
398: Let $\Ecal$ be an index family for $W$.
399: Assume that $f,\Ecal$ satisfy
400: the integrability condition \eqref{DGeqint} below.
401:
402: If $\mu$ is a compactly supported \Mb-density on $W$, nice
403: with index family $\Ecal$, then $f_*\mu$ is a \Mb-density on $\Rplus$,
404: nice with index family $f_\#\Ecal$ (defined in \eqref{DGeqindexset} below).
405: \end{theorem}
406: The integrability condition is:
407: \begin{equation}\label{DGeqint}
408: \inf \Ecal(G)>0 \text{ whenever } e_f(G) = 0
409: \end{equation}
410: where for any index set $E$
411: \begin{equation}\label{DGeqinfdef}
412: \inf E := \inf \{\Re z: (z,p)\in E\}
413: \end{equation}
414: (which is actually a minimum).%
415: \footnote{
416: Geometrically, $e_f(G)=0$ means that $f>0$ on $G$, so
417: the fibers $f^{-1}(x)$, $x>0$,
418: will hit only these $G$, and actually transversally
419: as in Figure \ref{DGfigpf}(a) at the $x$-axis.
420: So \eqref{DGeqint} generalizes \eqref{DGeqint1}
421: and comes from the fact that
422: $\int_0^1 x^z \,\frac{dx}x$ exists iff $\Re z>0$. }
423: To define $f_\#\Ecal$, associate to every face $F$
424: (i.e.\ non-empty intersection of boundary hypersurfaces) of $W$
425: the index set
426: $$\tilde{\Ecal}(F) = \Extunion_G \left\{(\frac{z}{e_f(G)},p):\,(z,p)\in
427: \Ecal(G)\right\}$$
428: where the extended union (defined in \eqref{DGeqextunion}) is
429: over all bhs's $G$ containing $F$ and having $e_f(G)>0$.
430: Then, define
431: \begin{equation}\label{DGeqindexset}
432: f_\#\Ecal = \bigcup_F \tilde{\Ecal}(F).
433: \end{equation}
434:
435: \begin{remarks}
436: \
437: \begin{enumerate}
438: \item
439: $f$ needs to be a fibration in the interior to ensure that
440: $f_*\mu$ is smooth in the interior.
441: \item
442: The definition of $f_\#\Ecal$ given above is a little more precise
443: than the one given in \cite{DGMel:CCDMWC} (which may yield an index set that
444: is 'too big'). But the Push-Forward Theorem with this 'smaller' $f_\#\Ecal$
445: follows directly from Melrose's by introducing a suitable partition of unity.%
446: \footnote{
447: Clearly, in \eqref{DGeqindexset} it is enough to take the union over all
448: {\em minimal} faces (with respect to inclusion), for example the corners
449: $A$, $B$ in Figure \ref{DGfigpf}(a). Thus, any 'regime' on $W$ (see
450: \ref{DGsubsubsecreg}) contributes some asymptotic terms.}
451: \item
452: The {\em proof} was essentially done above: Localize
453: as in the discussion of Example \ref{DGexpft}.3, this reduces to the
454: cases of Examples \ref{DGexpft}.1/2 (modulo
455: replacing $x,y$ by powers $x^\nu,y^\mu$ with $\nu,\mu>0$
456: determined by the $e_f(G)$, and modulo straight-forward
457: generalization to higher dimensions.)
458: \item
459: See the article \cite{DGGriGru:SALPFT}
460: \inthisbook\ for a discussion of the relation of
461: the Push-Forward Theorem (with $Z=\Rplus$)
462: and the 'Singular Asymptotics Lemma' by Br\"uning and Seeley
463: (\cite{DGBruSee:RSA}).
464: \end{enumerate}
465: \end{remarks}
466:
467: \medskip
468:
469: {\em Push-Forward Theorem with general target space.}
470: Here, some additional assumptions on the map $f$ are needed.
471: Before we can state these, we need to look a little closer at the
472: geometry of \Mb-maps:
473: \medskip
474:
475: {\em Boundary geometry of \Mb-maps.}
476: By definition, $f:W\to Z$ is a \Mb-map iff
477: \begin{equation} \label{DGeqfH}
478: f_H:=\rho_H\circ f:W\to\Rplus
479: \end{equation}
480: is a \Mb-map for all bhs's $H$ of $Z$,
481: and bdf's $\rho_H$ of $H$.
482: So we can define:
483: \begin{definition} \label{DGdefexp2}
484: The {\em exponent matrix} of a \Mb-map $f:W\to Z$ is the set of integers
485: $$ e_f(G,H) = e_{f_H}(G),\quad G\text{ bhs of }W,\, H\text{ bhs of }Z.$$
486: \end{definition}
487: Thus, $e_f(G,H)\not=0$ iff $f(G)\subset H$, and in this case if $p\in W$
488: has small distance $\eps$ from $G$ and distance $\geq$ const $>0$ from all
489: other bhs's of $W$, then $f(p)$ has distance of order $\eps^{e_f(G,H)}$
490: from $H$ (say in Euclidean metric for any local coordinate systems based
491: at points of $G$ and $H$).
492:
493: Referring to the Definition \ref{DGdefbmap} of \Mb-maps,
494: we have $e_f(G,H)=\alpha_{ij}$
495: in \eqref{DGeqbmap} if $G=\{x_j=0\}$ and $H=\{x_i'=0\}$ locally.
496:
497: Recall that a {\em face} of a mwc $W$
498: is a non-empty intersection of boundary hypersurfaces,
499: or $W$ itself. Each face is a mwc.
500: A \Mb-map $f$ induces a map
501: $$\fbar:\text{faces of }W\to\text{ faces of }Z$$
502: characterized by
503: \begin{equation} \label{DGeqftilde}
504: x\in \interior{F} \implies f(x)\in \interior{(\fbar(F))}.
505: \end{equation}
506: Alternatively, $\fbar(F)=$ the intersection of the bhs's $H$ of $Z$ satisfying
507: $f(F)\subset H$.
508:
509: In summary, the 'combinatorics' of a \Mb-map $f$ can be described either
510: by giving the pairs $(G,H)$ with $f(G)\subset H$, or equivalently by
511: the map $\fbar$, or (a little more refined) by the matrix $e_f$.
512: \begin{definition}\label{DGdefbfibr}
513: A \Mb-map $f:W\to Z$ is a {\em \Mb-fibration} if for each face $F$ of $W$,
514: \begin{itemize}
515: \item[(a)]
516: $\codim \fbar(F)\leq \codim F$
517: (it is enough to require this of bhs's $F$), and
518: \item[(b)]
519: $f$ is a fibration $\interior{F}\to \interior{(\fbar(F))}$. %
520: \footnote{
521: Melrose's definition in \cite{DGMel:CCDMWC} looks different, but is
522: equivalent: As is easily seen, condition (a) is equivalent to his
523: '\Mb-normality' (we assume all \Mb-maps to be interior), and, assuming (a),
524: condition (b) is equivalent to his '\Mb-submersion' condition, at least for
525: proper maps (for which submersion $\iff$ fibration), which is all that
526: matters anyway.}
527: \end{itemize}
528: \end{definition}
529: For example, if $Z=\R_+$ then $f$ is a \Mb-fibration iff $f$ is a
530: fibration over $(0,\infty)$ (here (a) is empty). The polar-coordinate
531: map is not a \Mb-fibration (since $\ff$ gets mapped to a codimension two
532: face), nor is any other non-trivial blow-down map.
533: An important example of a \Mb-fibration where condition (a) is non-empty
534: is given by the projection from the 'triple \Mb-space' in \eqref{DGeqbfibrex}.
535:
536: For a \Mb-map $f:W\to Z$ and an index family $\Ecal$ on $W$, define
537: the index family $f_\#\Ecal$ on $Z$ by
538: \begin{equation}\label{DGeqindexfam}
539: f_\#\Ecal(H)= (f_H)_\#\Ecal,
540: \end{equation}
541: with the right hand side defined by \eqref{DGeqindexset} and
542: \eqref{DGeqfH}.
543: The integrability condition is:
544: \begin{equation} \label{DGeqint2}
545: \inf \Ecal(G)>0 \quad\text{whenever}\quad e_f(G,H)=0\ \;\;\forall\ H.
546: \end{equation}
547: (The latter condition means that $f(G)\not\subset\partial Y$.)
548: \begin{theorem}[Push-Forward Theorem]
549: \label{DGthpft}
550: Let $W,Z$ be manifolds with corners and $f:W\to Z$ a \Mb-fibration.
551: Let $\Ecal$ be an index family for $W$, and assume \eqref{DGeqint2}.
552:
553: If $\mu$ is any compactly supported \Mb-density $\mu$ on $W$, nice with
554: index family $\Ecal$, then
555: the push-forward $f_*\mu$ is a \Mb-density on $Z$, nice with
556: index family $f_\#\Ecal$ (defined in \eqref{DGeqindexfam}).
557: \end{theorem}
558: See \cite{DGMel:CCDMWC} for a proof.
559: \begin{remarks}\label{DGrempft}
560: \
561: \begin{enumerate}
562: \item
563: {\em Why the \Mb-fibration conditions are needed:}
564: For example, (a) is violated for the polar-coordinate map, and this map
565: does not preserve niceness (this was the reason for doing blow-ups
566: in the first place!).
567: Also, if
568: $f_*\mu$ is to be nice then the expansion coefficients in
569: the asymptotics at any face should be smooth in the interior of the face, so
570: one should require a fibration condition here, which explains (b).
571: (Thus, (a) ensures good behavior of $f_*\mu$
572: when approaching the boundary, while
573: (b) does so {\em in} the boundary and locally in the interior.)
574: \item
575: {\em On determining the asymptotic type of $g_*\mu$ from
576: the asymptotic type of $\mu$}, for a map $g:Z\to Z'$ between mwc's:
577:
578: For $Z'=\Rplus$ the answer is given essentially by Theorem \ref{DGthpft1}:
579: If $g:Z\to\Rplus$ is a fibration in the interior and $\mu$ has
580: type $\beta:W\to Z$ then applying the theorem
581: to $f=g\circ\beta$ shows that $g_*\mu$ is nice on $\Rplus$.
582: Compare \eqref{DGeqpfcomp} where $g=\pi_1$.
583: Clearly, this only works if $(\beta^{-1})_*$
584: maps nice densities on $Z$ to nice densities on $W$; for
585: blow-ups $\beta$ this is clearly true, see \eqref{DGeqbdenstrf}
586: in Footnote \ref{DGfnbdens}.
587:
588: For general $Z'$ the problem is: Given $g:Z\to Z'$ and a blow-up
589: $\beta:W\to Z$, find a blow-up $\beta':W'\to Z'$ such that densities
590: of type $\beta$ are pushed forward to densities of type $\beta'$.
591: By the Push-Forward Theorem, this would be satisfied if
592: $\gtilde=(\beta')^{-1}\circ g\circ\beta:W\to W'$
593: was a \Mb-fibration.
594: \begin{equation} \label{DGeqcd}
595: \begin{CD}
596: W @>\gtilde>>W'\\
597: @V\beta VV @VV\beta' V \\
598: Z @>g>> Z'
599: \end{CD}
600: \end{equation}
601: Note that even for $\gtilde$ to be a well-defined map
602: implies restrictions on $\beta'$. The problem
603: when this is possible, and how to find $\beta'$, seems to be difficult.
604: \item
605: The support condition on $\mu$ in Theorem \ref{DGthpft} merely excludes
606: problems of non-integrability at infinity. Clearly, it could be weakened
607: to: $f$ is proper on $\supp\mu$.
608: (We need this extension when we discuss \PDO s.)
609: \end{enumerate}
610: \end{remarks}
611:
612: \subsection{Pull-back and asymptotic type} \label{DGsubsecpb}
613: Though just as important as the Push-Forward Theorem, this is child's play
614: in comparison. The main point was already mentioned in Remark
615: \ref{DGrembmap}.1. Doing the book-keeping on the index sets easily yields:
616:
617: \begin{theorem}[Pull-Back Theorem] \label{DGthpbt}
618: Let $f:W\to Z$ be a \Mb-map.
619: Then, for any function $v$ on $Z$ which is nice with
620: index family $\Fcal$, the pull-back $f^*v$ is a nice function
621: on $W$ with index family $f^\#\Fcal$ defined by: If $G$ is a bhs of $W$ then
622: \begin{multline*}
623: f^\#\Fcal(G) = \Big\{(q+\sum_H e_f(G,H)z_H, \sum_H p_H):\,q\in\N_0
624: \text{ and}\\
625: \text{ for each bhs $H$ of $Z$:}
626: \begin{cases}
627: (z_H,p_H)\in \Fcal(H) &\text{if } e_f(G,H)\not=0,\\
628: (z_H,p_H)=(0,0) &\text{if } e_f(G,H)=0.
629: \end{cases}
630: \Big\}.
631: \end{multline*}
632: \end{theorem}
633:
634: \begin{remarks} \label{DGrempbt}
635: \
636: \begin{enumerate}
637: \item
638: {\em On determining the asymptotic type of $g^*v$ from the
639: asymptotic type of $v$.}
640: Here one is given $g$ and a blow-up $\beta'$ in diagram \eqref{DGeqcd},
641: and needs to find a blow-up $\beta$
642: such that $g^*v$ has type
643: $\beta$ whenever $v$ has type $\beta'$. By the Pull-Back Theorem,
644: this is satisfied if $\gtilde$ is
645: well-defined and a \Mb-map (and $g$ is surjective).
646: Note that $(\beta')^{-1}\circ g$ is usually only defined on the interior since
647: $\beta'$ is not a diffeomorphism on the boundary, so $W$ has to be chosen
648: 'big' enough so that $\gtilde$ may extend continuously from the interior
649: to all of $W$.
650: \item
651: {\em The triple \Mb-space.}
652: As example consider the case relevant for composition
653: in the \Mb-\PDO\ calculus:
654: $g=\pi_1:\Rplus^3\to\Rplus^2$ is the projection $\pi_1(x_1,x_2,x_3)=(x_2,x_3)$.
655: The solution is easy: If $\beta':W'\to\Rplus^2$ is any blow-up then let
656: $W=\Rplus\times W'$, $\beta=\id_{\Rplus}\times\beta':\Rplus\times W'\to
657: \Rplus\times\Rplus^2$. In the case relevant for us, where $W'=[\Rplus^2,0]$
658: is just the blow-up of zero, $W$ is the blow-up of the $x_1$-axis.
659:
660: However, in the composition problem $W$ and $\beta$ need to work for
661: several maps $g$ {\em simultaneously}, and this makes
662: the problem more interesting.
663: Let $\pi_i:\Rplus^3\to\Rplus^2$ be the projection that forgets the
664: $i$'th coordinate, for $i=1,2,3$.
665: \medskip
666: \begin{quote}
667: {\bf Problem:}
668: Find a blow-up $\beta:W\to\Rplus^3$ such that whenever $v$ has type
669: $\beta':[\Rplus^2,0]\to\Rplus^2$ then $\pi_i^*v$ has type $\beta$ for
670: $i=1,2,3$.
671: \end{quote}
672: \medskip
673:
674: In other words, $\pitilde_i=(\beta')^{-1}\circ\pi_i\circ\beta:W\to[\Rplus^2,0]$
675: must be a \Mb-map for $i=1,2,3$.
676: It is clear that at least all three coordinate axes must be blown up.
677: The most naive thing to try is to blow up one axis (say the $x_1$-axis)
678: and then (the preimages of)
679: the other two. However, it is easily seen
680: that $\pi_2$ and $\pi_3$ are still not well-defined on the resulting space.
681:
682:
683: But there is a beautiful solution which even preserves the symmetry:
684: First, blow up zero in $\Rplus^3$. Then, blow up the preimages of
685: the three coordinate axes (in any order, since they are separated now!).
686: the result is called triple \Mb-space $X^3_b$
687: and shown in Figure \ref{DGfigtriple}.
688: \begin{figure}[htbp]
689: \input{figtriple.pstex_t}
690: \caption{The triple \Mb-space (and projection $\pitilde_3$)}
691: \label{DGfigtriple}
692: \end{figure}
693:
694: Let us convince ourselves pictorially that the maps
695: \begin{equation} \label{DGeqbfibrex}
696: \pitilde_i:X_b^3\to X_b^2:= [\Rplus^2,(0,0)]
697: \end{equation}
698: are well-defined and \Mb-fibrations.
699: By symmetry, it is enough to consider $\pitilde_3$.
700: It is well-defined since the $x_3$-axis was blown up.
701: Denote the bhs's of $X_b^3$ by $\bface_1,\bface_2,\bface_3$ (the 'old'
702: bhs's from $\Rplus^3$), $\ff_1,\ff_2,\ff_3$ (the front faces of the axis
703: blow-ups), and $\fff$ (the front face of the point blow-up) as in
704: Figure \ref{DGfigtriple}.
705: The bhs's are mapped as follows:
706: \begin{align} \label{DGeqbhs}
707: \begin{split}
708: \ff_2, \bface_1& \mapsto \lb\\
709: \ff_1, \bface_2&\mapsto \rb\\
710: \fff,\ff_3&\mapsto\ff\\
711: \bface_3&\mapsto X^2_b,
712: \end{split}
713: \end{align}
714: and all these maps are onto. Also, $\interior{(X_b^3)}\to\interior{(X_b^2)}$.
715: Therefore, the preimage of each bhs is a union of bhs's, which almost
716: shows that $\pitilde_3$ is a \Mb-map (see Remark \ref{DGrembmap}.1; of course
717: one may check the full condition \eqref{DGeqbmap} by direct calculation).
718: Also, \eqref{DGeqbhs} defines the map $\pitildebar_3$
719: (see \ref{DGeqftilde}) on bhs's and
720: this determines $\pitildebar_3$ for all faces.
721: Since all faces on the right of \eqref{DGeqbhs} have codimension at most one,
722: condition (a) in the Definition \ref{DGdefbfibr} of a \Mb-fibration
723: is satisfied.
724:
725: Condition (b) is easily checked for each face: For example,
726: $\interior{\bface_1}\to\interior{\lb}$ is basically the same map
727: as $[\Rplus^2,(0,0)]\to\Rplus$ from \ref{DGfigpf}, and so is
728: $\interior{\fff}\to\interior{\ff}$ near the boundary. All maps from
729: codimension two faces are either diffeomorphisms or constant, so they
730: are fibrations trivially.
731: \end{enumerate}
732: \end{remarks}
733:
734: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
735: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Distributions subsec
736: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
737:
738: \subsection{Distributions} \label{DGsubsecdist}
739: So far, all singular behavior occurred at the boundary.
740: Now we turn to the description of singularities in the interior of
741: a mwc $Z$. This means talking about distributions%
742: \footnote{
743: There are also distributions whose singular support is
744: contained in the boundary. We will not discuss them here (although
745: they are not really more difficult).
746: See \Mnotes, for example.}.
747: In many situations only a very special class of distributions occurs,
748: the 'step 1 polyhomogeneous conormal' (here called 'conormal') ones%
749: \footnote{
750: This is not the most general kind of what's usually called
751: conormal distributions, but they are easy
752: to define and sufficient for many purposes.}.
753: They are smooth outside a
754: submanifold, and at the submanifold have a special explicitly
755: describable kind of singular behavior, which is in some sense similar
756: to the behavior of a nice function at the boundary.
757:
758: In the case of manifolds most of
759: this material is quite standard (see e.g.\ \cite{DGHor:ALPDOIII},
760: \cite{DGSim:PDO});
761: we will briefly recall
762: the definition, give some examples and state the push-forward and pull-back
763: theorems. As an illustration, we
764: use this to show that the set of (properly supported) classical
765: pseudodifferential operators on $\R$ is closed under composition.
766: The extension of the definition and basic properties of conormal
767: distributions to manifolds with corners is quite straight-forward if
768: the singular submanifold hits the boundary in a 'product-type' way.
769:
770: For lack of space we
771: do not treat the transformation of the principal symbol under pull-back
772: and push-forward. However,
773: this is important for the composition formula for pseudodifferential
774: operators (see the references above).
775: \subsubsection{Conormal distributions on manifolds}
776: \label{DGsubsubseccon}
777: \begin{definition} \label{DGdefcon}
778: Let $Z$ be a manifold and $Y\subset Z$ a submanifold. A distribution
779: $u\in \Dcal(Z)$ is {\em conormal} with respect to $Y$ if, for some $m\in\R$,
780: \begin{itemize}
781: \item
782: $u$ is smooth on $Z\setminus Y$, and
783: \item
784: in any local coordinate system $x:U\subset Z\to\R^n$ sending $Y\cap U$ to
785: $\R^k\times\{0\}^{n-k}\subset\R^n$ there is a representation
786: \begin{equation} \label{DGeqcon}
787: u(y,z) = \int_{\R^{n-k}} e^{iz\zeta} a(y,\zeta)\,d\zeta
788: \end{equation}
789: where $y=(x_1,\ldots,x_k), z=(x_{k+1},\ldots,x_n)$ and $a$ is a smooth
790: function on $(Y\cap U)\times \R^{n-k}$ with asymptotics
791: \begin{equation}\label{DGeqconasymp}
792: a(y,\zeta) \sim \sum_{j=0}^\infty a_{m-j}(y,\zeta),
793: \end{equation}
794: as $|\zeta|\to\infty$,
795: where $a_l$ is homogeneous of degree $l$ in $\zeta$, for each $l$.%
796: \footnote{
797: The meaning of the asymptotics is that, for any $N$, if $a^{(N)}$ is the
798: sum up to the term $a_{-N}$ then
799: $|a(y,\zeta)-a^{(N)}(y,\zeta)| \leq C|\zeta|^{-N-1}$, plus analogous
800: estimates for all derivatives in $y$ and $\zeta$.
801:
802: The {\em order} of $u$ is defined to be $m+(n-2k)/4$, if $a_m\not\equiv 0$.
803: }
804: \end{itemize}
805: \end{definition}
806: Note that \eqref{DGeqcon} is simply the inverse Fourier transform in $z$
807: (i.e.\ 'transversal' to $Y$), with smooth dependence
808: on the parameter $y\in Y$.
809: \begin{examples} \label{DGexcon}
810: \
811: \begin{enumerate}
812: \item
813: For $Z=\R$ and $Y=\{0\}$ the distributions $\delta$ and $p.v.\frac1x$
814: are conormal, and also all of their derivatives and anti-derivatives,
815: which include for example $x_+^\alpha$ for $\alpha\in \N_0$
816: (the function vanishing for $x\leq0$ and equal to $x^\alpha$ for $x>0$).
817: Any $(-1)$-homogeneous distribution is a linear combination of $\delta$
818: and $p.v.\frac1x$,
819: so a conormal distribution (with $m\in\Z$) may be thought
820: of as 'series' of such terms, of increasing regularity.%
821: \footnote{
822: More generally, one can define conormality without reference
823: to the Fourier transform: \eqref{DGeqcon}
824: and \eqref{DGeqconasymp} are equivalent to the existence of
825: distributions $u_s(y,\cdot)\in\Dcal'(\R^{n-k}) \cap
826: C^\infty(\R^{n-k}\setminus 0)$,
827: homogeneous
828: of degree $s$ and depending smoothly on the parameter $y$, such that
829: $$ u - \sum_{j=0}^N u_{j+n-k-m}\in C^\infty(\R^k,C^{N'}(\R^{n-k}))$$
830: (locally) for all $N$, where $N'=N-C$, with $C$ only depending on $n$.
831: This may look weaker than the definition
832: above, but is actually equivalent (exercise!).
833: }
834: \item
835: For $Z=\R^n$, $n>1$, and $Y=\{0\}$ there is much more freedom since
836: now the space of $l$-homogeneous distributions is infinite-dimensional for
837: each $l$. The simplest example is $\delta$ again.
838: \item
839: For $Z=\R^n\times\R^n$ (with coordinates $w,w'\in\R^n$) and
840: $Y=\{w=w'\}$ (the diagonal) the conormal distributions are
841: the integral kernels of classical pseudodifferential operators on $\R^n$
842: since \eqref{DGeqcon} precisely amounts to their 'usual' definition, in the
843: coordinates $y=w$, $z=w-w'$, see for example \cite{DGShu:PDOST}, Section 3.7.
844: (And similarly for \PDO s on any manifold.)
845: The order of the conormal distribution is the order of the operator
846: in the usual sense.
847: For example, the differential operator $P=\sum_\alpha a_\alpha(w)(\partial
848: /\partial w)^\alpha$ has kernel $P(w,w') = \sum_\alpha a_\alpha(w)
849: \delta^{(\alpha)}(w-w')$.
850: \end{enumerate}
851: \end{examples}
852: \begin{remarks}\label{DGremcondist}
853: \
854: \begin{itemize}
855: \item
856: It's not obvious, but the definition is actually independent of the chosen
857: coordinate system $x$, see \cite{DGHor:ALPDOIII}.
858: Of course, $a$ will depend on the choice of coordinates, but its leading term
859: $a_m$ is invariant if considered as section of the conormal bundle of $Y$.
860: It is called the {\em principal symbol} of $u$. One easily sees that it depends
861: only on the restriction of $u$ to arbitrarily small neighborhoods of $Y$.
862: \item
863: The definition carries over immediately to distribution densities or,
864: more generally, to distributions with values in any bundle over $Z$.
865: \end{itemize}
866: \end{remarks}
867:
868: We now consider push-forward and pull-back of distributions under a
869: smooth map $f:W\to Z$. The proofs of the following theorems are quite
870: easy, given the coordinate invariance of Definition \ref{DGdefcon}.
871: They can be found in \Mnotes.
872: The push-forward of any distribution density $\mu$ on $W$ is a distribution
873: density on $Z$ (supposing, as usual, that $f$ is proper on $\supp \mu$), see
874: the Appendix.
875: The question arises whether conormality of $\mu$ with respect to a
876: submanifold $X\subset W$ implies
877: conormality of $f_*\mu$.
878: The answer is no in general%
879: \footnote{
880: Example: If $f$ is bijective and smooth then
881: $\singsupp f_*\mu \supset f(\singsupp \mu)$.
882: The latter need not be (contained in) a submanifold even if
883: $\singsupp\mu$ is, therefore
884: $f_*\mu$ may be not conormal even if $\mu$ is.}; it is a very tricky problem
885: to determine precise conditions when it is true.
886: It depends essentially on the behavior of the fibers of
887: $f$ and their tangency to $X$. We only consider the simplest case of a
888: fibration whose fibers meet $X$ transversally in isolated points only,
889: which is enough for many purposes.
890:
891: \begin{theorem}[Push-forward of conormal distributions]
892: \label{DGthpfdist}
893: Let $f:W\to Z$ be a fibration between manifolds. Let $X$ be a submanifold
894: of $W$ such that for each $x\in X$, the tangent spaces to $X$ and to the
895: fiber $f^{-1}(f(x))$ through $x$ intersect only in zero.%
896: \footnote{
897: Equivalently, $d(f_{|X})$ is injective.}
898: Let $\mu$ be a distribution density on $W$, conormal with respect to $X$,
899: and assume $f$ is proper on $\supp\mu$.
900: \begin{enumerate}
901: \item[(a)]
902: If $f_{|X}$ is a diffeomorphism onto $Z$ then $f_*\mu$ is smooth.
903: \item[(b)]
904: Otherwise, $f(X)$ is a proper submanifold of $Z$
905: and $f_*\mu$ is conormal with respect to $f(X)$.
906: \end{enumerate}
907: \end{theorem}
908: Thus, the 'vertical' (in fiber direction) singularities get integrated out,
909: while the others remain, as in the simple example of
910: $f:\R^2\to\R$, $(x,y)\mapsto x$:\\
911: For $X=\{(x,0):x\in\R\}$ one has, for example,
912: $f_*(\delta(y)\,dxdy) = 1\cdot dx$ (case (a)), and for $X=\{(0,0)\}$
913: one has $f_*(\delta(x)\delta(y)\,dxdy) = \delta(x)\,dx$ (case (b)).
914: (Calculations done using \eqref{DGeqpf}.)
915: See Remark \ref{DGrempft}.3 concerning the support condition.
916:
917: For pull-back the situation is different: While
918: the pull-back for general distributions is only defined under fibrations
919: (and conormality is always preserved under a fibration),
920: a {\em weaker} condition
921: on $f$ already allows to pull back conormal distributions:
922: \begin{theorem}[Pull-back of conormal distributions]
923: \label{DGthpbdist}
924: Let $Y\subset Z$ be a submanifold, and assume that $f:W\to Z$
925: is transversal{$\,$}%
926: \footnote{
927: I.e.\ if $z=f(x)\in Y$ then $T_zZ$ is spanned by $T_zY$ and
928: $df(T_xW)$.
929: }
930: to $Y$.Then $f^{-1}(Y)$ is a submanifold of $W$, and if
931: $u$ is a distribution on $Z$, conormal with respect to $Y$, then $f^*u$
932: is a distribution on $W$ which is conormal with respect to $f^{-1}(Y)$.
933: \end{theorem}
934: We saw in Examples \ref{DGexpfpb} that we also need to {\em multiply}
935: distributions. A complete discussion of when this is possible would lead us
936: too far astray, so we'll just sketch the procedure which allows to define a
937: product in this context.
938:
939: For two functions $u_1,u_2$ on a manifold $Z$, we can translate
940: the trivial identity
941: $u_1(x)u_2(x) = u_1(x)u_2(y)_{|x=y}$ into geometric terms as
942: $$ u_1u_2 = i^* (u_1\times u_2) $$
943: where $i:Z\to Z\times Z,x\mapsto (x,x)$ is the diagonal inclusion
944: and $(u_1\times u_2)(x,y)=u_1(x)u_2(y)$ defines the direct product
945: of $u_1$ and $u_2$ as function on $Z\times Z$.
946:
947: When trying to generalize this to distributions $u_1,u_2$ on $Z$,
948: we first note that the direct product is well-defined as
949: a distribution on $Z\times Z$ (since $u_1$ and $u_2$
950: 'depend on different sets of variables' in $Z\times Z$).
951: The problem arises with the pull-back: $i$ is certainly not a fibration
952: (it's not even surjective), so one would hope to apply Theorem
953: \ref{DGthpbdist}. But this fails since usually $u_1\times u_2$ is not
954: conormal, even if $u_1$ and $u_2$ are conormal;
955: for example for $Z=\R$ and $u_1=u_2=p.v.\frac1x$ one gets
956: $(p.v. \frac1x)(p.v.\frac1y)$, which is not
957: conormal (since its singular support, the union
958: of both coordinate axes, is not a manifold).
959: \medskip
960:
961: The following theorem allows a way out (at least in some situations):
962: \begin{theorem}[Direct product of conormal distributions]
963: \label{DGthproddist}
964: Let $u_i$ be distributions on manifolds $Z_i$,
965: conormal with respect to submanifolds
966: $Y_i$, for $i=1,2$.
967: Then the direct product $u_1\times u_2$ can be written
968: $u_1\times u_2 = v+w$
969: where $v$ is conormal with respect to
970: $Y_1\times Y_2$ and $w$ has wave front set contained in any given conic
971: neighborhood of
972: $(Z_1\times N^*Y_2) \cup (N^*Y_1\times Z_2)$.
973: \end{theorem}
974: We will not define wave front sets here, see \cite{DGHor:ALPDOI}.
975: $N^*X_i$ denotes the conormal bundle.
976: The point is that in the applications we have in mind
977: (Examples \ref{DGexpfpb}) the product is integrated in the end,
978: and then the position of $WF(w)$ guarantees that the
979: term resulting from $w$ is smooth (by a generalization of Theorems
980: \ref{DGthpfdist}(a) and \ref{DGthpbdist}), so the singularities are determined
981: only by the conormal part $v$.
982:
983: \subsubsection{Composition of pseudodifferential operators}
984: \label{DGsubsubsecdistpdo}
985: Let us check how these results show that the composition of
986: two pseudodifferential operators on a manifold $X$
987: is a pseudodifferential operator%
988: \footnote{\label{DGfnprop}%
989: The support condition in Theorem \ref{DGthpfdist}
990: translates into the condition that at least one of the factors
991: is properly supported (i.e.\ the two projections $X^2\to X$ are proper on
992: the support of the integral kernel).
993: We will neglect this in the following discussion. See also Remark
994: \ref{DGremclasspdo}.3.}. For simplicity we take $X=\R$ although the general
995: case works precisely the same way.
996:
997: Thus, we are given distributions $A,B$ on $\R^2$, conormal with
998: respect to the diagonal (see Example \ref{DGexcon}.3).
999: To avoid confusion later on, we will
1000: write $A\in\Dcal'(X_1\times X_2)$, $B\in\Dcal'(X_2\times X_3)$
1001: although $X_1=X_2=X_3=\R$. The diagonals will be denoted
1002: $\Delta_A\subset X_1\times X_2$ and $\Delta_B\subset X_2\times X_3$.
1003:
1004: The composition of $A$ and $B$ has integral kernel given by
1005: \eqref{DGeqcomp}. Here, the product should be expanded, as explained
1006: above, as diagonal pull-back of the direct product. However, matters
1007: can be simplified slightly. The $\pi_1,\pi_3$ pull-backs can be omitted
1008: since clearly one also has
1009: \begin{equation} \label{DGeqCeq}
1010: C= \pi_{2*}(i^* (A\times B))
1011: \end{equation}
1012: where $i$ is the embedding
1013: $$ i: X_1\times X_2\times X_3 \to X_1\times X_2\times X_2\times X_3,
1014: \quad (x_1,x_2,x_3)\mapsto (x_1,x_2,x_2,x_3).$$
1015:
1016: Using Theorem \ref{DGthproddist} we write
1017: \begin{equation} \label{DGeqprodsum}
1018: A\times B= v+w,
1019: \end{equation}
1020: with $v$ conormal with respect to $\Delta_A\times\Delta_B$
1021: and $WF(w)$ close to $(X_1\times X_2\times N^*\Delta_B) \cup
1022: (N^*\Delta_A\times X_2\times X_3)$.
1023:
1024: We first analyze $\pi_{2*}(i^*v)$:
1025: $i$ is transversal to $\Delta_A\times\Delta_B$ since
1026: the image of $di$ is $\{(\alpha,\beta,\beta,\gamma)\}$ and the
1027: tangent space of $\Delta_A\times\Delta_B$ is $\{(\delta,\delta,\eps,\eps)\}$
1028: (all free variables in braces range over $\R$),
1029: and these two subspaces clearly span $\R^4$.
1030: Therefore, Theorem \ref{DGthpbdist} shows that $i^*v$ is conormal
1031: with respect to $\Delta':=i^{-1}(\Delta_A\times\Delta_B) = \{x_1=x_2=x_3\}$,
1032: the space diagonal.
1033: Finally, the tangent spaces to $\Delta'$ and the fiber of $\pi_2$
1034: are $\{(\alpha,\alpha,\alpha)\}$ and $\{(0,\beta,0)\}$, so they have
1035: zero intersection, and $\pi_2(\Delta')=\Delta_C$, the diagonal in $X_1\times
1036: X_3$. Therefore, Theorem \ref{DGthpfdist}(b) applies, so
1037: \begin{equation} \label{DGeqvcon}
1038: \pi_{2*}(i^*v) \text{ is conormal with respect to the diagonal.}
1039: \end{equation}
1040:
1041: Finally, we analyze $w$, using standard results on wave front sets. First,
1042: by Theorem 8.2.4 in \cite{DGHor:ALPDOI}, the pull-back $i^*w$ is defined
1043: as a distribution if $WF(w)\cap N^*({\rm Im}\, i)=\emptyset$, and then
1044: $WF(i^*w)\subset i^*(WF(w))$.
1045: Now $N^*({\rm Im}\, i)=\{(0,\alpha,-\alpha,0)\}$ at every point,
1046: and this has non-zero angle
1047: with the fiber of $N^*\Delta_A\times X_2\times X_3$, which is
1048: $\{(\alpha,-\alpha,0,0)\}$, and similarly with the fiber of
1049: $X_1\times X_2\times N^*\Delta_B$.
1050: Therefore, by choosing $WF(w)$ close enough to these latter sets,
1051: we may assume that
1052: $WF(w)\cap N^*({\rm Im}\, i)=\emptyset$. Also, $WF(i^*w)$ is contained
1053: in a small conic neighborhood of $i^*$ of these sets, i.e.\ of
1054: (fiberwise)
1055: $\{(\alpha,-\alpha,0)\}\cup\{(0,\alpha,-\alpha)\}$.
1056: By another standard theorem (see \cite{DGSim:PDO}, ex.\ 6.7.8)
1057: the push-forward
1058: $\pi_{2*}u$ of a distribution $u$ on $X_1\times X_2\times X_3$ is
1059: smooth unless $WF(u)$ hits the conormal space to the fiber of $\pi_2$.
1060: Since the latter is $\{(\alpha,0,\beta)\}$, this is clearly
1061: not the case for $u=i^*w$, so finally we obtain:
1062: $$ \pi_{2*}(i^*w)\text { is smooth.} $$
1063: This together with \eqref{DGeqvcon}, \eqref{DGeqCeq} and \eqref{DGeqprodsum}
1064: shows that $C$ is conormal with respect to the diagonal, i.e.\ the integral
1065: kernel of a pseudodifferential operator.
1066:
1067: \subsubsection{Conormal distributions on manifolds with corners}
1068: \label{DGsubsubsecconmwc}
1069: The definition of conormal distributions depends in an essential way on the
1070: fact that 'normal slices' to $Y$ in $Z$ look the same at every point of $Y$.
1071: Therefore, it would be problematic to try to define conormality for
1072: $Z=\Rplus^2$, $Y=\{(x,x):x\in\Rplus\}$: At zero, there is not even a
1073: reasonable candidate for a 'normal slice'!
1074:
1075: However, if $Z$ looks like $\R^{n-k}\times Y$ near $Y$ then Definition
1076: \ref{DGdefcon} makes sense literally even if $Y$ is a mwc, when we
1077: require the $y$-dependence to be smooth up to the boundary everywhere.
1078: For example, this is the case for $Z=\Rplus\times\R$, $Y=\Rplus\times\{0\}$ or
1079: $Z=[\Rplus^2,0]$, $Y=\Delta_b:=\{\theta=\pi/4\}$ (using polar coordinates on
1080: $Z$, see Figure \ref{DGfigpolar}(b)).
1081:
1082: Thus, we have defined {\em distributions on a mwc $Z$ which are conormal
1083: with respect to an interior p-submanifold, smoothly at the boundary}
1084: (cf.\ \ref{DGsubsubsecgenbu}.4; 'interior' means that
1085: $Y\not\subset\partial Z$).
1086: One may actually allow nice (rather than smooth) behavior at the
1087: boundary, with respect to a given index family. This gives
1088: {\em nice conormal distributions.} The precise definitions are
1089: quite straight-forward, we leave them as exercise to the reader.
1090: (See \Mnotes.)
1091:
1092: The {\em push-forward} for nice conormal distributions may be analyzed
1093: by a combination of Theorems \ref{DGthpft} and \ref{DGthpfdist}. Since
1094: the point of Theorem \ref{DGthpft} was to allow maps $f$ more general than
1095: fibrations, the assumption on $f$ should be: $f$ is a \Mb-fibration, and
1096: a fibration (of mwc's) in some neighborhood of $X$, satisfying the additional
1097: transversality condition in Theorem \ref{DGthpfdist}. Then the result
1098: of the push-forward is nice conormal again. The proof is straightforward by
1099: use of a partition of unity, and is left as an exercise.
1100:
1101: Similarly {\em pull-back} of nice conormal distributions is easy by
1102: combining Theorems \ref{DGthpbt} and \ref{DGthpbdist}. $f$ needs to
1103: be a \Mb-map transversal to $Y$, then the pull-back by $f$ of a nice
1104: conormal function is nice conormal with respect to $f^{-1}(Y)$.
1105:
1106: \endinput
1107: