generated from pbelmans/latex-template
-
Notifications
You must be signed in to change notification settings - Fork 0
Commit
This commit does not belong to any branch on this repository, and may belong to a fork outside of the repository.
- Loading branch information
1 parent
40be2cf
commit 92a60ef
Showing
4 changed files
with
146 additions
and
5 deletions.
There are no files selected for viewing
This file contains bidirectional Unicode text that may be interpreted or compiled differently than what appears below. To review, open the file in an editor that reveals hidden Unicode characters.
Learn more about bidirectional Unicode characters
This file contains bidirectional Unicode text that may be interpreted or compiled differently than what appears below. To review, open the file in an editor that reveals hidden Unicode characters.
Learn more about bidirectional Unicode characters
120 changes: 120 additions & 0 deletions
120
V5B1-Advanced-Topics-in-Complex-Analysis/Lecture 17.tex
This file contains bidirectional Unicode text that may be interpreted or compiled differently than what appears below. To review, open the file in an editor that reveals hidden Unicode characters.
Learn more about bidirectional Unicode characters
Original file line number | Diff line number | Diff line change |
---|---|---|
@@ -0,0 +1,120 @@ | ||
\section{Lecture 17 -- 5th December 2024}\label{sec: lecture 17} | ||
We prove the Bochner-Martinelli integral formula. | ||
\begin{theorem}[Bochner-Martinelli Integral Formula]\label{thm: Bochner-Martinelli integral} | ||
Let $G\subseteq\CC^{n}$ be a bounded domain with smooth boundary. If $f$ is a continuously differentiable function on the boundary $\partial G$ then | ||
$$f(\tau)=\int_{\partial G}f(w)\beta(w,\tau)-\int_{G}\overline{\partial}f(w)\wedge \beta(w,\tau)$$ | ||
for all $\tau\in G$.\marginpar{Theorem 1.2} | ||
\end{theorem} | ||
\begin{proof} | ||
For fixed $\tau\in G$ and $r>0$ sufficiently small that $\overline{B_{r}(\tau)}\subsetneq G$. We have by Stokes' theorem and \Cref{lem: Bochner-Martinelli computation} | ||
\begin{align*} | ||
\int_{\partial G}f(w)\beta(w,\tau)-\int_{\partial B_{r}(\tau)}f(w)B(w,z)&=\int_{G\setminus\overline{B_{r}(\tau)}}\dform(f(w)\beta(w,\tau))\\ | ||
&=\int_{G\setminus\overline{B_{r}(\tau)}}\overline{\partial}f(w)\wedge \beta(w,\tau). | ||
\end{align*} | ||
Note further that | ||
$$\lim_{r\to 0}\int_{G\setminus\overline{B_{r}(\tau)}}\overline{\partial}f(w)\wedge \beta(w,\tau)=\int_{G}\overline{\partial}f(w)\wedge \beta(w,\tau)$$ | ||
so we can rearrange the equations above to observe | ||
$$\int_{\partial B_{r}(\tau)}f(w)\beta(w,\tau)=f(\tau)\int_{\partial B_{r}(\tau)}\beta(w,\tau)-\int_{\partial B_{r}(\tau)}(f(w)-f(\tau))\beta(w,\tau).$$ | ||
To prove the result, we show $\int_{\partial B_{r}(\tau)}\beta(w,\tau)=1$ and that the second summand of the expression above vanishes. | ||
|
||
For the first claim, we apply Stokes' theorem once more, and recall that the volume of $B_{r}(\tau)$ is $\frac{\pi^{n}}{n!}r^{2n}$ and compute | ||
\footnotesize | ||
\begin{align*} | ||
\int_{\partial B_{r}(\tau)}\beta(w,\tau) &= \int_{\partial B_{r}(\tau)}\frac{(n-1)!}{(2\pi i)^{n}}\sum_{j=1}^{n}\frac{\overline{w_{j}}-\overline{z_{j}}}{\Vert w-\tau\Vert^{2n}}\dform\overline{w_{1}}\wedge\dform w_{1}\wedge\dots\wedge\dform w_{j-1}\wedge\dform w_{j}\wedge \dform\overline{w_{j+1}}\wedge\dots\wedge\dform\overline{w_{n}}\wedge\dform w_{n} \\ | ||
&= \frac{(n-1)!}{r^{2n}(2\pi i)^{n}}\int_{\partial B_{r}(\tau)}\sum_{j=1}^{n}(\overline{w_{j}}-\overline{z_{j}})\dform\overline{w_{1}}\wedge\dform w_{1}\wedge\dots\wedge\dform w_{j-1}\wedge\dform w_{j}\wedge \dform\overline{w_{j+1}}\wedge\dots\wedge\dform\overline{w_{n}}\wedge\dform w_{n} \\ | ||
&=\frac{(n-1)!}{r^{2n}(2\pi i)^{n}}\int_{B_{r}(\tau)}\dform\left(\sum_{j=1}^{n}(\overline{w_{j}}-\overline{z_{j}})\dform\overline{w_{1}}\wedge\dform w_{1}\wedge\dots\wedge\dform w_{j-1}\wedge\dform w_{j}\wedge \dform\overline{w_{j+1}}\wedge\dots\wedge\dform\overline{w_{n}}\wedge\dform w_{n}\right) \\ | ||
&= \frac{(n-1)!}{r^{2n}(2\pi i)^{n}}\int_{B_{r}(\tau)}n\cdot\dform\overline{w_{1}}\wedge\dform w_{1}\wedge\dots\wedge\dform w_{j}\wedge \dform\overline{w_{j+1}} \\ | ||
&= \frac{(n-1)!}{r^{2n}(2\pi i)^{n}}\int_{B_{r}(\tau)}n(2i)^{n}\dform V \\ | ||
&= 1 | ||
\end{align*} | ||
\large | ||
For the second claim, the computation above yields | ||
\footnotesize | ||
\begin{align*} | ||
&\frac{(n-1)!}{r^{2n}(2\pi i)^{n}}\left(\int_{B_{r}(\tau)}n\cdot\dform\overline{w_{1}}\wedge\dform w_{1}\wedge\dots\wedge\dform w_{n}\wedge \dform\overline{w_{n}}\right)\\ | ||
&\hspace{1cm}+\frac{(n-1)!}{r^{2n}(2\pi i)^{n}}\int_{B_{r}(\tau)}\sum_{j=1}\partial_{\overline{w_{j}}}f(w)(\overline{w_{j}}-\overline{\tau_{j}})\cdot\dform\overline{w_{1}}\wedge\dform w_{1}\wedge\dots\wedge\dform w_{n}\wedge \dform\overline{w_{n}}. | ||
\end{align*} | ||
\large | ||
By boundedness of $U$, $|f(w)-f(\tau)|\leq C\Vert w-\tau\Vert$ and $|\partial_{\overline{w_{j}}}f(w)(\overline{w_{j}}-\overline{z_{j}})|\leq C\Vert w-z\Vert$ for some large constant $C$. So for all $w\in\partial B_{r}(\tau)$, the quantities $|f(w)-f(\tau)|,|\partial_{\overline{w_{j}}}f(w)(\overline{w_{j}}-\overline{z_{j}})|$ are bounded above by $Cr$ so | ||
$$\left|\int_{\partial B_{r}(\tau)}(f(w)-f(\tau))\beta(w,\tau)\right|\leq\frac{(n-1)!}{r^{2n}(2\pi)^{n}}\left(2\int_{B_{r}(\tau)}n\cdot 2^{n}Cr\dform V\right)=2Cr$$ | ||
which vanishes as $r\to0$ proving the claim. | ||
\end{proof} | ||
\begin{remark} | ||
Note that the Bochner-Martinelli kernel is not a holomorphic form. | ||
\end{remark} | ||
The Bochner-Martinelli integral formula allows us to prove Hartogs' so-called ``Kugelsatz,'' allowing us to extend holomorphic functions over compact domains. For this, we will require the following lemmata, only one of which we prove. | ||
\begin{lemma}\label{lem: Kugelsatz lemma 1} | ||
Let $G\subseteq\CC^{n}$ be a domain and $\beta(w,z)$ the Bochner-Martinelli kernel on $G$ and $f$ continuously differentiable on $\partial G$. Then $\overline{\partial}_{z}\beta(w,z)=-\overline{\partial}_{w}\beta_{1}(w,z)$ where\marginpar{Lemma 1.5} | ||
$$\beta_{1}(w,z)=(n-1)\beta(w,z)\wedge(\overline{\partial}_{w}\beta)^{n-2}\wedge\overline{\partial}_{z}\beta\cdot\left(\frac{1}{2\pi i}\right)^{n}\frac{1}{\Vert w-z\Vert^{2n}}.$$ | ||
\end{lemma} | ||
\begin{proof} | ||
See \cite[\S 4, Prop. 4.9]{Range}. | ||
\end{proof} | ||
\begin{lemma}\label{lem: Kugelsatz lemma 2} | ||
Let $G\subseteq\CC^{n}$ and $f$ holomorphic on $\partial G$. Then | ||
$$F(z)=\int_{\partial G}f(w)B(w,z)$$ | ||
is holomorphic in $z$ for $z\notin\partial G$.\marginpar{Lemma 1.6} | ||
\end{lemma} | ||
\begin{proof} | ||
We compute | ||
\begin{align*} | ||
\overline{\partial}_{z}F(z) &= \int_{\partial G}f(w)\overline{\partial}_{z}B(w,z) \\ | ||
&= -\int_{\partial G}f(w)\overline{\partial}_{w}\beta_{1}(w,z) \\ | ||
&= -\int_{\partial G}\overline{\partial}_{w}(f(w)\beta_{1}(w,z)) \\ | ||
&= -\int_{\partial G}\dform_{w}(f(w)\beta_{1}(w,z)) \\ | ||
&= 0 | ||
\end{align*} | ||
as desired. | ||
\end{proof} | ||
In turn: | ||
\begin{theorem}[Hartogs' Kugelsatz]\label{thm: Kugelsatz} | ||
Let $U\subseteq\CC^{n}$ be an open, $K\subseteq U$ compact, and $U\setminus K$ connected. If $n\geq 2$ and $f$ is holomorphic on $U\setminus K$ then $f$ extends holomorphically to $U$. | ||
\end{theorem} | ||
\begin{proof} | ||
Let $G\subsetneq U$ be a domain and $G_{0}$ a subdomain of $G$ fully containing $K$. By hypothesis $f$ is holomorphic in $G\setminus\overline{G_{0}}$ and by the Bochner-Martinelli integral formula \Cref{thm: Bochner-Martinelli integral}, we have | ||
$$f(z)=\int_{\partial G}f(w)\beta(w,z)-\int_{\partial G_{0}}f(w)\beta(w,z).$$ | ||
Denoting the integrals $F_{1}(z),F_{2}(z)$, respectively, $F_{1}$ is holomorphic for all $z\notin\partial G$ by \Cref{lem: Kugelsatz lemma 2} and $F_{2}$ holomorphic for all $z\notin G_{0}$. In particular, $F_{1}$ is also holomorphic on all of $G$ and thus holomorphic in $K$ and $F_{2}(z)\to0$ as $|z|\to\infty$. | ||
|
||
Since $n\geq 2$, there exist hyperplanes that do not meet $G_{0}$ on which $F_{2}$ is zero, showing $F_{2}$ is zero by the identity theorem \Cref{thm: multivariate identity} and $F_{1}$ is a holomorphic extension of $f$ since it coincides with $f$ on $G\setminus\overline{G_{0}}$ once again using the identity theorem. | ||
\end{proof} | ||
|
||
We now return to a consideration of the multivariate Cauchy-Riemann equations, in particular considering their role in producing holomorphic functions corresponding to fixed differential form data. | ||
|
||
The motivating problem is the existence of holomorphic functions satisfying Cousin-I distributions. | ||
\begin{definition}[Cousin-I Distribution]\label{def: Cousin-I distribution} | ||
Let $U\subseteq\CC^{n}$ be open. A Cousin-I distribution on $U$ is given by $\{(U_{i},f_{i})\}_{i\in I}$ where $\{U_{i}\}_{i\in I}$ form an open cover of $U$, $f_{i}\in\Kcal(U_{i})$ holomorphic such that $f_{i}-f_{j}$ is holomorphic on $U_{ij}=U_{i}\cap U_{j}$. | ||
\end{definition} | ||
A solution to the Cousin-I problem is a meromorphic function compatible with the data of a Cousin-I distribution in the following sense. | ||
\begin{definition}[Solution to Cousin-I Problem] | ||
Let $U\subseteq\CC^{n}$ be open and $\{(U_{i},f_{i})\}_{i\in I}$ be a Cousin-I distribution on $U$. A solution to the Cousin-I problem is a meromorphic function $f\in\Kcal(U)$ such that $f|_{U_{i}}-f_{i}$ is holomorphic for all $i\in I$. | ||
\end{definition} | ||
Solutions to problems of this type are constructed by finidng a smooth solution and ``altering'' it to a holomorphic solution by solving a $\overline{\partial}$-equation. | ||
|
||
We discuss a sequence of reductions for this problem, first recalling the following notions from the theory of smooth manifolds. | ||
\begin{definition}[Partition of Unity]\label{def: partition of unity} | ||
Let $U\subseteq\RR^{n}$ be open and $\{U_{i}\}_{i\in I}$ a locally finite open cover of $U$ such that each $U_{i}$ is relatively compact in $U$. A partition of unity $\{\psi_{i}\}_{i\in I}$ subordinate to the cover $\{U_{i}\}_{i\in I}$ is the data of smooth functions $\psi_{i}:U_{i}\to \RR$ such that $\mathrm{supp}(\psi_{i})\subseteq U_{i}$ and $\sum_{i\in I}\psi_{i}(x)=1$ for all $x\in U$. | ||
\end{definition} | ||
Partitions of unity exist for open sets of $\CC^{n}$. | ||
\begin{proposition}\label{prop: partitions of unity exist} | ||
Let $U\subseteq\RR^{n}$ be open and $\{U_{i}\}_{i\in I}$ a locally finite open cover of $U$ such that each $U_{i}$ is relatively compact in $U$. There exists a partition of unity $\{\psi_{i}\}_{i\in I}$ subordinate to the cover $\{U_{i}\}_{i\in I}$. | ||
\end{proposition} | ||
A proof of \Cref{prop: partitions of unity exist} can be found in \cite[\S 2]{LeeSM}. Note in particular that holomorphic phenomena are in particular smooth which justifies their use in what follows. | ||
|
||
\begin{proposition}\label{prop: equivalent Cousin-I distributions} | ||
Let $U\subseteq\CC^{n}$ be open and $\{(U_{i},f_{i})\}_{i\in I},\{(V_{j},g_{j})\}_{j\in J}$ be Cousin-I distributions such that their union is a Cousin-I distribution. If $f\in\Kcal(U)$ is a solution to the Cousin-I distribution $\{(U_{i},f_{i})\}_{i\in I}$ then $f$ is a solution to the Cousin-I distribution $\{(V_{j},g_{j})\}_{j\in J}$ as well. | ||
\end{proposition} | ||
\begin{proof} | ||
Let $f\in\Kcal(U)$ be a solution to the Cousin-I problem for the distribution $\{(U_{i},f_{i})\}_{i\in I}$ and $\{\psi_{j}\}_{j\in J}$ be a partition of unity subordinate to the cover $\{V_{j}\}_{j\in J}$. We want to show that $f|_{V_{j}}-g_{j}\in\Ocal(V_{j})$ for all $j\in J$. For this, we note that $V_{j}$ admits a cover by $\{U_{i}\cap V_{j}\}_{i\in I}$ on which $f_{i}|_{U_{i}\cap V_{j}}-g_{j}|_{U_{i}\cap V_{j}}$ is holomorphic so $f_{i}\psi_{j}-g_{j}$ is holomorphic on all $V_{j}$ and similarly $f-f_{i}|_{U_{i}\cap V_{j}}$ is holomorphic for all $U_{i}\cap V_{j}$ and all $i\in I$ with fixed $j\in J$ which extends holomorphically to $f|_{V_{j}}-f_{i}\psi_{j}$ on $V_{j}$ But their sum is $f|_{V_{j}}-g_{j}$ and holomorphic as it is the finite sum of holomorphic functions by local finiteness of the cover. | ||
\end{proof} | ||
This motivates the following definition. | ||
\begin{definition}[Equivalent Cousin-I Distributions] | ||
Let $U\subseteq\CC^{n}$ be open and $\{(U_{i},f_{i})\}_{i\in I},\{(V_{j},g_{j})\}_{j\in J}$ be Cousin-I distributions. $\{(U_{i},f_{i})\}_{i\in I},\{(V_{j},g_{j})\}_{j\in J}$ are equivalent Cousin-I distributions if their union is a Cousin-I distribution. | ||
\end{definition} | ||
|
||
This reduces solving the Cousin-I problem to the following. | ||
\begin{proposition}\label{prop: Cousin-I is CR01} | ||
Let $U\subseteq\CC^{n}$ be open and $f\in C^{\infty}(U)$, $f=\sum_{j=1}^{n}f_{j}\dform\overline{z_{j}}$, and there exists $u$ a smooth function such that $\overline{\partial}u=f$ and $\overline{\partial}f=0$. Then the Cousin-I problem admits a solution. | ||
\end{proposition} | ||
\begin{proof} | ||
Let $\{(U_{i},f_{i})\}_{i\in I}$ be a Cousin-I distribution as above and set $g_{i}=\sum_{j\in I}\phi_{j}f_{ji}$ which is defined on $U_{i}$. Note that on $U_{ik}$ the difference $g_{i}-g_{k}$ is given by $f_{ik}$ as expected and that $\overline{\partial}g_{i}-\overline{\partial}g_{j}=0$. We can define function $f$ as in the hypothesis of the theorem by taking $f|_{U_{i}}=\overline{\partial}g_{i}$ which statisfies the hypothesis by inspection and admits a solution $u$ to the Cousin-I problem. | ||
\end{proof} |
This file contains bidirectional Unicode text that may be interpreted or compiled differently than what appears below. To review, open the file in an editor that reveals hidden Unicode characters.
Learn more about bidirectional Unicode characters