Hausdorff space
A topological space $X$ is called a Hausdorff space if for all $a,b\in X$ such that $a\neq b$,
there are disjoint open sets $A,B\subseteq X$ such that $a\in A$ and $b\in B$.
Proposition.
Let $X$ be a Hausdorff space. Then every subspace of $X$ is Hausdorff.
(show proof)
Proof.
Let $U$ be a subspace of $X$.
Let $p,q\in U$.
Then there exist disjoint open subsets $P,Q$ of $X$ such that $p\in P$ and $q\in Q$.
Note that $U\cap P$ and $Q\cap P$ are disjoint open subsets of $U$ such that $p\in U\cap P$ and $q\in U\cap Q$.
Thus $U$ is Hausdorff.
$\blacksquare$
Proposition.
Let $X$ be a Hausdorff space. If a sequence $(p_i)$ of $X$ converges to both $p$ and $q$, then $p=q$.
(show proof)
Proof.
Suppose for contradiction that $p\neq q$.
Since $X$ is Hausdorff, there exists disjoint open subsets $A$ and $B$ such that $p\in A$ and $q\in B$.
But by convergence, there exists $n\in N$ such that $p_n\in A$ and $p_n\in B$, implying $A\cap B\neq\emptyset$, a contradiction.
Thus $p=q$.
$\blacksquare$
Proposition.
Let $X$ be a Hausdorff space. Then every compact subset of $X$ is closed.
(show proof)
Proof.
Let $U\subseteq X$ be compact.
If $U$ is empty, then it is trivially closed.
Now suppose $U$ is non-empty.
Let $p\in X\setminus U$.
Then for all $q\in U$, there exist disjoint open sets $U_q,V_q$ such that $q\in U_q$ and $p\in V_q$.
Note that $\{U_q:q\in U\}$ is an open cover of $U$ and thus has a finite subcover $\mathcal C$, and for some finite $S\subseteq U$, $\mathcal C=\{U_q:q\in S\}$.
Since $U$ is non-empty, $S$ is non-empty.
Then $\{V_q:q\in S\}$ is a non-empty finite set of open sets, thus $\cap\{V_q:q\in S\}$ is open.
Note that $p\in\cap\{V_q:q\in S\}\subseteq X\setminus U$, implying $X\setminus U$ is a neighborhood of $p$.
Since $X\setminus U$ is a neighborhood of every point in it, it is open.
Thus $U$ is closed.
$\blacksquare$
Proposition.
Let $X$ be a Hausdorff space. Then $X$ is locally compact if and only if it has a precompact basis.
(show proof)
Proof.
Suppose $X$ is locally compact.
Let $p\in X$, then there exists a compact neighborhood $U_p$ of $p$, thus there exists an open set $S_p$ such that $p\in S_p\subseteq U_p$.
Since $X$ is Hausdorff and $U_p$ is compact, $U_p$ is closed, implying $\overline{U_p}=U_p$.
Let $\tau$ be the topology of $X$, and let $\mathcal B$ denote $\{S_p\cap U:p\in X,U\in\tau\}$.
Note that given $p\in X$ and $U\in\tau$, $S_p\cap U$ is open and $\overline{S_p\cap U}\subseteq\overline{U_p}=U_p$, implying $\overline{S_p\cap U}$ is compact.
Then every element of $\mathcal B$ is open and precompact.
Let $S$ be an open subset of $X$.
Then $S=\cup\{S_p\cap S:p\in X\}$ and $\{S_p\cap S:p\in X\}\subseteq\mathcal B$.
Thus $\mathcal B$ is a precompact basis of $X$.
Suppose $X$ has a precompact basis $\mathcal B$.
Then $X=\cup\mathcal B$.
Let $p\in X$, then for some $U\in\mathcal B$, $p\in U$.
Since $U$ is precompact, $\overline U$ is compact.
Thus $\overline U$ is a compact neighborhood of $p$.
Hence $X$ is locally compact.
$\blacksquare$
Countability
Let $X$ be a topological space.
If for all $x\in X$, there exists a sequence $(N_i)$ of neighborhoods of $x$, such that
every neighborhood $N$ of $x$ is a superset of $N_i$ for some $i\in N$, then $X$ is said to be first-countable.
If there exists a sequence $(U_i)$ of open sets of $X$,
such that every open set $U$ of $X$ is the union of $\{U_i:i\in S\}$ for some $S\subseteq N$,
then $X$ is said to be second-countable.
Proposition.
Second-countability implies first-countability.
(show proof)
Proof.
Suppose $X$ is second-countable and let $(U_i)$ be a sequence of open sets as required by second-countability.
Let $x\in X$, define $N_i=U_i$ if $x\in U_i$ and $N_i=X$ otherwise,
then $(N_i)$ is a sequence of neighborhoods of $x$.
Let $V$ be a neighborhood of $x$, then for some open set $S$, we have $x\in S\subseteq V$.
Note that there exists $M\subseteq N$ such that $\cup\{U_i:i\in M\}=S$,
thus for some $i\in M$, $x\in U_i$, implying $N_i=U_i\subseteq S\subseteq V$.
We have shown that $X$ is first-countable.
$\blacksquare$
Proposition.
Let $X$ be a first-countable space. Then every subspace of $X$ is first-countable.
(show proof)
Proof.
Let $U$ be a subspace of $X$.
Let $x\in U$.
Then there exists a sequence $(N_i)$ of neighborhoods of $x$, with respect to $X$, such that
every neighborhood $N$ of $x$, with respect to $X$, is a superset of $N_i$ for some $i\in N$.
Then $(U\cap N_i)$ is a sequence of neighborhoods of $x$, with respect to $U$.
Let $N$ be a neighborhood of $x$, with respect to $U$.
Then $N\cup(X\setminus U)$ is a neighborhood of $x$, with respect to $X$,
and thus is a superset of $N_i$ for some $i\in N$.
Then $N$ is a superset of $U\cap N_i$.
We have shown that $U$ is first-countable.
$\blacksquare$
Proposition.
Let $X$ be a first-countable space, let $S\subseteq X$, and let $p\in X$.
Then $p\in\overline S$ if and only if there exists a sequence of $S$ that converges to $p$ with respect to $X$.
(show proof)
Proof.
Let $(N_i)$ be a sequence of neighborhoods of $p$ as required by first-countability.
Recursively define $M_0=N_0$ and $M_{i+1}=M_i\cap N_{i+1}$, then $(M_i)$ is a sequence of neighborhoods of $p$ satisfying the first-countability requirement,
and $M_{i+1}\subseteq M_i$.
Suppose $p\in\overline S$.
If $p\in S$, then a sequence $(p_i)$ such that $p_i=p$ trivially converges to $p$.
Now suppose $p$ is a limit point of $S$.
Then for each $i$, there exists $p_i\in S$ such that $p_i\in M_i\setminus\{p\}$.
And $(p_i)$ forms a sequence of $S$.
Let $U$ be a neighborhood of $p$, then for some $n\in N$, $M_n\subseteq U$,
thus for all $k\ge n$, $M_k\subseteq U$, implying $p_k\in U$.
Thus $(p_i)$ converges to $p$.
Suppose there exists a sequence $(p_i)$ of $S$ that converges to $p$.
If $p\in S$, then trivially, $p\in\overline S$.
Now suppose $p\notin S$.
Let $U$ be a neighborhood of $p$, then for some $k\in N$, $p_k\in U$.
Since $p_k\in S$, $p_k\neq p$.
We have shown that $p$ is a limit point of $S$, and thus $p\in\overline S$.
$\blacksquare$
Proposition.
Let $X$ be a second-countable space. Then every subspace of $X$ is second-countable.
(show proof)
Proof.
Let $U$ be a subspace of $X$.
Let $(U_i)$ be a sequence of open sets of $X$ as required by second-countability.
Then $(U\cap U_i)$ is a sequence of open sets of $U$.
Let $V$ be an open set of $U$, then for some open set $S$ of $X$, $V=S\cap U$.
Then for some $M\subseteq N$, $S=\cup_{i\in M}\{U_i\}$.
Then $V=\cup_{i\in M}\{U\cap U_i\}$.
We have shown that $U$ is second-countable.
$\blacksquare$
Proposition.
Let $X$ be a second-countable space, then every open cover of $X$ has a countable subcover.
(show proof)
Proof.
Let $(U_i)$ be a sequence of open sets as required by second-countability.
Let $\mathcal C$ be an open cover of $X$.
Define $M$ to be the set of natural numbers $k$ such that there exists $C\in\mathcal C$ with $U_k\subseteq C$.
This defines a function $f:M\to\mathcal C$ such that for all $k\in M$, $U_k\subseteq f(k)$.
Let $p\in X$, then for some $C\in\mathcal C$, $p\in C$.
Since $C$ is open, for some $L\subseteq N$, $C=\cup\{U_i:i\in L\}$.
Then for some $k\in L$, $p\in U_k\subseteq C$, implying $k\in M$.
And we have $p\in U_k\subseteq f(k)$.
Thus $\{f(k):k\in M\}$ covers $X$, which is a countable subcover of $\mathcal C$.
$\blacksquare$
Locally Euclidean
Given $n\in N$, a topological space $X$ is said to be locally Euclidean of dimension $n$
if for all $x\in X$, there exists an open set of $X$ containing $x$ that is homeomorphic to an open set of $R^n$.
Locally half-Euclidean
Define $H^n=\{x\in R^n|x_n\ge0\}$, $\Int H^n=\{x\in R^n|x_n\gt0\}$, $\partial H^n=\{x\in R^n|x_n=0\}$ for $n\gt0$,
and $H^0=R^0$, $\Int H^0=R^0$, $\partial H^0=\emptyset$.
Given $n\in N$, a topological space $X$ is said to be locally half-Euclidean of dimension $n$
if for all $x\in X$, there exists an open set of $X$ containing $x$ that is homeomorphic to an open set of $R^n$ or $H^n$.
Manifold
Given $n\in N$, a topological space $M$ is said to be an $n$-manifold if it is Hausdorff,
second-countable, and locally Euclidean of dimension $n$.
Manifold with boundary
Given $n\in N$, a topological space $M$ is said to be an $n$-manifold if it is Hausdorff,
second-countable, and locally half-Euclidean of dimension $n$.
Proposition.
For all $n\in N$, $R^n$ is an $n$-manifold and $H^n$ is an $n$-manifold with boundary.
(show proof)
Proof.
$R^0$ is trivially a $0$-manifold.
Let $n\gt0$, then
$R^n$ is clearly a Hausdorff space and locally Euclidean of dimension $n$.
The only thing to show is second-countability.
Let $X=\{B_r(p):r\in Q,p\in Q^n\}$, then $\abs{X}=\omega$, and there exists a tuple $(X_i)$ as a bijection from $\omega$ to $X$.
Let $S$ be an open set of $R^n$ and let $p\in S$, then there exists $\delta\gt0$ such that $B_\delta(p)\subseteq S$.
Note that there exists $q\in Q^n$ such that $0\lt d(p,q)\lt\frac{\delta}{3}$.
Let $s\in Q$ such that $\frac{\delta}{3}\lt s\lt\frac{2\delta}{3}$, then $p\in B_s(q)\subseteq S$.
Therefore, we can define a map $f:S\to X$ such that $p\in f(p)\subseteq S$ for all $p\in S$.
And $S=\cup\{f(p):p\in S\}$. Note that $\{f(p):p\in S\}=\{X_i:i\in U\}$ for some $U\subseteq N$.
Thus $\cup\{X_i:i\in U\}=S$.
We have shown that $R^n$ is an $n$-manifold.
Then $H^n$, as a subspace of $R^n$, is Hausdorff and second-countable.
Trivially, it is also locally half-Euclidean of dimension $n$.
Thus $H^n$ is an $n$-manifold with boundary.
$\blacksquare$
Coordinate chart
Suppose $M$ is an $n$-manifold (with boundary), a coordinate chart on $M$ is a homeomorphism $\varphi:U\to V$ where $U$ is an open subset of $M$ and
$V$ is an open subset of $R^n$ (or $H^n$).
A coordinate chart may also be called a chart, and a chart $\varphi$ from $U$ may be denoted $(\varphi,U)$.
Proposition.
Suppose $M$ is an $n$-manifold (with boundary). The restriction of a coordinate chart $\varphi:U\to V$ on $M$ to an open subset $S$ of $U$ is a coordinate chart $\varphi|_S:S\to\varphi(S)$ on $M$.
(show proof)
Proof.
Let $T=\varphi(S)$. Then $\varphi|_S:S\to T$ is bijective, and $\varphi^{-1}(T)=S$.
Since $\varphi$ is continuous, $\varphi|_S:S\to T$ is continuous.
Since $\varphi^{-1}$ is continuous, $\varphi^{-1}|_T:T\to S$ is continuous.
Since $\varphi^{-1}|_T=\varphi|_S^{-1}$, $\varphi|_S^{-1}$ is continuous.
Thus $\varphi|_S:S\to T$ is a homeomorphism.
Since $U$ is an open subset of $M$ and $S$ is an open subset of $U$, $S$ is an open subset of $M$.
Since $\varphi^{-1}$ is continuous, the preimage of $\varphi^{-1}$ on $S$, which is $T$, is an open subset of $V$.
Since $V$ is an open subset of $R^n$ (or $H^n$), $T$ is an open subset of $R^n$ (or $H^n$).
Hence $\varphi|_S:S\to T$ is a coordinate chart on $M$.
$\blacksquare$
Transition map
Let $M$ be an $n$-manifold (with boundary), and let $\varphi_1,\varphi_2$ be two charts on $M$ with domains $U_1,U_2$ respectively,
then $U_1\cap U_2$ is an open subset of both $U_1$ and $U_2$, thus $\varphi_1|_{U_1\cap U_2}$ and $\varphi_2|_{U_1\cap U_2}$ are coordinate charts on $M$.
We call $\varphi_2|_{U_1\cap U_2}\circ\varphi_1|_{U_1\cap U_2}^{-1}$, also denoted $\varphi_2\circ\varphi_1^{-1}$,
the transition map from $\varphi_1$ to $\varphi_2$, which is clearly homeomorphic.
If in addition, it is diffeomorphic, then $\varphi_1$ and $\varphi_2$ are said to be smoothly compatible, or just compatible.
Proposition.
Let $M$ be a $n$-manifold (with boundary), and let $(\varphi_1,U),(\varphi_2,V)$ be compatible charts on $M$, then the restriction of $\varphi_1$ on any open subset of $U$ is compatible with $\varphi_2$.
(show proof)
Proof.
Let $W$ be an open subset of $U$. We have shown that $\varphi_1|_W$ is indeed a chart on $M$.
Since $W\cap V\subseteq U\cap V$, $\varphi_1(W\cap V)\subseteq\varphi_1(U\cap V)$ and $\varphi_2(W\cap V)\subseteq\varphi_2(U\cap V)$.
Clearly, the transition map from $\varphi_1|_W$ to $\varphi_2$ and its inverse are just restrictions of the transition map from $\varphi_1$ to $\varphi_2$ and its inverse,
and they are thus smooth. Hence $\varphi_1|_W$ and $\varphi_2$ are compatible.
$\blacksquare$
Atlas
Let $M$ be an $n$-manifold (with boundary), an atlas for $M$ is a collection $\mathcal A$ of charts whose domains cover $M$.
If any two charts in $\mathcal A$ are smoothly compatible, then $\mathcal A$ is said to be a smooth atlas of $M$.
If no proper extension of $\mathcal A$ is a smooth atlas, then $\mathcal A$ is said to be maximal.
A maximal smooth atlas on $M$ is called a smooth structure on $M$.
A chart in a smooth structure is called a smooth chart.
Smooth manifold
A smooth $n$-manifold (with boundary) is an $n$-manifold (with boundary) with a smooth structure.
We may use $(M,\mathcal A)$ to denote an $n$-manifold (with boundary) $M$ with a smooth structure $\mathcal A$.
Proposition.
Let $M$ be a smooth manifold (with boundary), then the restriction of a smooth chart $(\varphi,U)$ to an open subset of $U$ is again a smooth chart.
(show proof)
Proof.
Let $\mathcal A$ denote the smooth structure of $M$ and let $(\varphi,U)\in\mathcal A$.
Since $(\varphi,U)$ is a chart, the restriction of $\varphi$ to an open subset $S$ of $U$, denoted $(\varphi,S)$, is also a chart.
Let $(\psi,V)\in\mathcal A$, then $(\varphi,U)$ and $(\psi,V)$ are compatible, thus $(\varphi,S)$ and $(\psi,V)$ are compatible.
Suppose for contradiction that $(\varphi,S)\notin\mathcal A$, then $\mathcal A\cup\{(\varphi,S)\}$ is a proper extension of $\mathcal A$ that is also a smooth atlas,
implying $\mathcal A$ is not maximal, a contradiction.
Thus $(\varphi,S)\in\mathcal A$.
$\blacksquare$
Open submanifold
Let $M$ be a smooth $n$-manifold (with boundary), let $U$ be an open subset of $M$,
then the topological subspace $U$ of $M$ together with the set of restrictions of smooth charts of $M$ on $U$
forms a smooth $n$-manifold (with boundary)
(show proof),
Proof.
The topological subspace $U$ is clearly Hausdorff and second-countable.
Let $p\in U$, then there exists a chart $(\varphi,V)$ on $M$ such that $p\in V$,
and $(\varphi,U\cap V)$ is a chart on $M$.
Note that $U\cap V$ is an open subset of $U$, thus $(\varphi,U\cap V)$ is also a chart on $U$.
We have shown that the topological subspace $U$ is an $n$-manifold (with boundary).
Let $\mathcal A$ denote the smooth structure of $M$ and let $\mathcal A'$ denote the set of restrictions of smooth charts of $M$ on $U$.
For the same reason as above, every function in $\mathcal A'$ is a chart on $U$.
Clearly, $\mathcal A'$ covers $U$ and is thus an atlas of $U$.
Let $\varphi_1',\varphi_2'\in\mathcal A'$, then for some $\varphi_1,\varphi_2\in\mathcal A$, $\varphi_1',\varphi_2'$ are restrictions of $\varphi_1,\varphi_2$ on $U$.
Since $\varphi_1$ and $\varphi_2$ are compatible, $\varphi_1'$ and $\varphi_2'$ are also compatible.
Thus $\mathcal A'$ is a smooth atlas of $U$.
Suppose some chart $\psi$ of $U$ is compatible with every chart in $\mathcal A'$ but not in $\mathcal A'$.
Let $\varphi\in\mathcal A$, then the restriction $\varphi'$ of $\varphi$ on $U$ is in $\mathcal A'$, and thus compatible with $\psi$.
Then $\varphi$ is also compatible with $\psi$.
Thus $\mathcal A\cup\{\psi\}$ is a proper extension of $\mathcal A$ that is a smooth atlas of $M$, a contradiction.
We have shown that $\mathcal A'$ is maximal.
$\blacksquare$
called an open submanifold of $M$.
Proposition.
Let $M$ be a smooth $n$-manifold (with boundary), let $U$ be an open subset of $M$, and let $V$ be an open subset of $U$.
Then the open submanifold $V$ of $M$ is equivalent to the open submanifold $V$ of the open submanifold $U$ of $M$.
(show proof)
Proof.
Let $V_a$ denote the former and let $V_b$ denote the latter.
A smooth manifold is a set with a topology and a smooth structure.
Clearly, $V_a$ and $V_b$ share the same underlying set and the same topology.
Let $\mathcal A$ denote the smooth structure of $M$.
Then the smooth structure of $V_a$, denoted $\mathcal A_a$, is $\{\varphi|_V:\varphi\in\mathcal A\}$.
And the smooth structure of $V_b$, denoted $\mathcal A_b$, is $\{\varphi|_V:\varphi\in\mathcal A_U\}$, where $\mathcal A_U$ denotes $\{\varphi|_U:\varphi\in\mathcal A\}$.
Suppose $\psi\in\mathcal A_a$, then $\varphi|_V=\psi$ for some $\varphi\in\mathcal A$.
Note that $\varphi|_U\in\mathcal A_U$, thus $\psi=\varphi|_V=(\varphi|_U)|_V\in\mathcal A_b$.
Now suppose $\psi\in\mathcal A_b$, then for some $\phi\in\mathcal A_U$, $\phi|_V=\psi$, and for some $\varphi\in\mathcal A$, $\varphi|_U=\phi$.
Thus $\psi=\phi|_V=(\varphi|_U)|_V=\varphi|_V$, implying $\psi\in\mathcal A_a$.
We have shown that $V_a$ and $V_b$ also share the same smooth structure.
$\blacksquare$
Proposition.
Let $M$ be an $n$-manifold (with boundary) with a smooth atlas $\mathcal A$.
Let $\overline{\mathcal A}$ be the collection of charts of $M$ smoothly compatible with every chart in $\mathcal A$.
Then $\overline{\mathcal A}$ is a maximal smooth atlas of $M$, which is said to be the smooth structure of $M$ determined by $\mathcal A$.
(show proof)
Proof.
Since $\mathcal A\subseteq\overline{\mathcal A}$, $\overline{\mathcal A}$ is an atlas of $M$.
Let $\varphi_1:U_1\to V_1,\varphi_2:U_2\to V_2$ be charts in $\overline{\mathcal A}$.
Let $x\in\varphi_1(U_1\cap U_2)$. Since $\mathcal A$ is an atlas of $M$,
there exists $\varphi_3:U_3\to V_3$ in $\mathcal A$ such that $\varphi_1^{-1}(x)\in U_3$.
Since $\varphi_1$ and $\varphi_3$ are compatible, $\varphi_3\circ\varphi_1^{-1}$ is diffeomorphic.
Since $\varphi_2$ and $\varphi_3$ are compatible, $\varphi_2\circ\varphi_3^{-1}$ is diffeomorphic.
Since $\varphi_1^{-1}(x)\in U_1\cap U_2\cap U_3$, $x\in\varphi_1(U_1\cap U_2\cap U_3)$, which is open.
Note that $(\varphi_2\circ\varphi_3^{-1})\circ(\varphi_3\circ\varphi_1^{-1})$, when restricted from $\varphi_1(U_1\cap U_2\cap U_3)$ to $\varphi_2(U_1\cap U_2\cap U_3)$, is a diffeomorphism.
Let $u\in\varphi_1(U_1\cap U_2\cap U_3)$, then $(\varphi_2\circ\varphi_3^{-1})\circ(\varphi_3\circ\varphi_1^{-1})(u)=\varphi_2\circ\varphi_1^{-1}(u)$.
If $\varphi_1(U_1\cap U_2\cap U_3)$ is an open subset of $H^n$,
then we need an extra step to find a smooth map on an open ball centered at $x$
that agrees with $\varphi_2\circ\varphi_1^{-1}$ on its domain if $x_n\gt0$ or on its intersection with $H^n$ if $x_n=0$.
We have shown that $\varphi_2\circ\varphi_1^{-1}$ is smooth.
By a symmetric argument, $\varphi_1\circ\varphi_2^{-1}$ is also smooth.
Thus $\varphi_2\circ\varphi_1^{-1}$ is diffeomorphic, implying $\varphi_1$ and $\varphi_2$ are compatible.
Hence $\overline{\mathcal A}$ is a smooth atlas of $M$.
Now suppose for contradiction that there exists a proper extension $\overline{\mathcal A}'$ of $\overline{\mathcal A}$ that is a smooth atlas.
Then some chart $\varphi$ is in $\overline{\mathcal A}'$ but not in $\overline{\mathcal A}$.
And $\varphi$ is compatible with every chart in $\overline{\mathcal A}'$, and thus every chart in $\mathcal A$.
But then $\varphi$ should be in $\overline{\mathcal A}$, a contradiction.
Hence $\overline{\mathcal A}$ is maximal.
$\blacksquare$
Standard Euclidean space
The identity map on $R^n$ alone forms a smooth atlas, which determines a smooth structure on $R^n$.
$R^n$ together with this smooth structure is called standard $R^n$.
Notation.
Given $x\in R^{n+1}$, we use $(x_1,\ldots,\hat{x_i},\ldots,x_{n+1})$ to denote a point in $R^n$ constructed by eliminating the $i$-th coordinate of $x$.
Standard sphere
Let $n\in N$, $r\in R^+$ and $p\in R^{n+1}$. We define the $n$-sphere with radius $r$ centered at $p$, denoted $S(n,r,p)$, to be $$\{u\in R^{n+1}|d(u,p)=r\}$$
With the subspace topology inherited from $R^{n+1}$, $S(n,r,p)$ is clearly Hausdorff and second-countable.
For $i\in\{1,\ldots,n+1\}$, define $U_i^+=\{u\in S(n,r,p)|u_i\gt p_i\}$ and $U_i^-=\{u\in S(n,r,p)|u_i\lt p_i\}$,
then these sets are open subsets of $S(n,r,p)$, and for all $u\in S(n,r,p)$, $u\in U_i^+$ or $u\in U_i^-$ for some $i\in\{1,\ldots,n+1\}$.
For $i\in\{1,\ldots,n+1\}$, let $\hat p_i$ denote $(p_1,\ldots,\hat{p_i},\ldots,p_{n+1})$,
define $f_i^+:U_i^+\to B(n,r,\hat p_i)$ such that $f_i^+(x)=(x_1,\ldots,\hat{x_i},\ldots,x_{n+1})$ for all $x\in U_i^+$
and $f_i^-:U_i^-\to B(n,r,\hat p_i)$ such that $f_i^-(x)=(x_1,\ldots,\hat{x_i},\ldots,x_{n+1})$ for all $x\in U_i^-$,
where $B(n,r,\hat p_i)$ is the open ball of $R^n$ with radius $r$ centered at $\hat p_i$.
Then these functions are homeomorphisms, thus rendering $S(n,r,p)$ an $n$-manifold, and they form a smooth atlas $\mathcal A$ on $S(n,r,p)$
(show proof).
Proof.
For $i\in\{1,\ldots,n+1\}$, define $g_i^+:B(n,r,\hat p_i)\to U_i^+$ such that $g_i^+(y)=(y_1,\ldots,y_{i-1},p_i+\sqrt{r^2-d(y,\hat p_i)^2},y_i,\ldots,y_n)$ for all $y\in B(n,r,\hat p_i)$
and $g_i^-:B(n,r,\hat p_i)\to U_i^-$ such that $g_i^-(y)=(y_1,\ldots,y_{i-1},p_i-\sqrt{r^2-d(y,\hat p_i)^2},y_i,\ldots,y_n)$ for all $y\in B(n,r,\hat p_i)$.
Then for each $i$, for all $x\in U_i^+$, $g_i^+(f_i^+(x))=x$, for all $x\in U_i^-$, $g_i^-(f_i^-(x))=x$, and for all $y\in B(n,r,\hat p_i)$, $f_i^+(g_i^+(y))=y$ and $f_i^-(g_i^-(y))=y$.
Thus $f_i^+$ and $f_i^-$ are bijective, and $g_i^+$ and $g_i^-$ are their inverses respectively. Since these functions are trivially continuous, they are homeomorphisms.
Let $i,j\in\{1,\ldots,n+1\}$.
- If $i=j$, then $f_j^+\circ g_i^+$ is an identity map, which is smooth, and the domain of $f_j^-\circ g_i^+$ is empty, thus it is smooth.
- If $i\neq j$, then the component functions of $f_j^+\circ g_i^+$ are either a projection function or $p_i+\sqrt{r^2-d(y,\hat p_i)^2}$, thus $f_j^+\circ g_i^+$ is smooth, and similarly, $f_j^-\circ g_i^+$ is also smooth.
By a symmetric argument, transition maps from $g_i^-$ to $f_j^+$ or $f_j^-$ are smooth.
$\blacksquare$
The smooth $n$-manifold formed by the $n$-manifold $S(n,r,p)$ together with the smooth structure determined by $\mathcal A$ is called the standard $n$-sphere with radius $r$ centered at $p$.
If $r=1$ and $p$ is the origin, then we also call it the
standard unit sphere of dimension $n$, denoted $S^n$.
Definition.
Let $M$ be a smooth $n$-manifold with boundary.
A smooth chart $(\varphi,U)$ on $M$ to an open subset of $R^n$ is called an
interior chart.
A smooth chart $(\varphi,U)$ on $M$ to an open subset of $H^n$ such that $\varphi(U)\cap\partial H^n\neq\emptyset$ is called a
boundary chart.
Let $p\in M$, then $p$ is called
- an interior point if $p\in U$ for some interior chart $(\varphi,U)$ of $M$;
- a boundary point if $p\in U$ for some boundary chart $(\varphi,U)$ of $M$ such that $\varphi(p)\in\partial H^n$.
Proposition.
Let $M$ be a smooth $n$-manifold with boundary, then every smooth chart of $M$ is either an interior chart or a boundary chart but not both.
(show proof)
Proof.
Let $(\varphi,U)$ be a smooth chart of $M$, then $\varphi(U)$ is an open subset of either $R^n$ or $H^n$.
Suppose $\varphi(U)$ is both an open subset of $R^n$ and an open subset of $H^n$ such that $\varphi(U)\cap\partial H^n\neq\emptyset$,
then there exists $p\in\varphi(U)$ such that $p\in\partial H^n$. But then no open ball centered at $p$ is contained by $\varphi(U)$,
contradicting that $\varphi(U)$ is an open subset of $R^n$. Hence $\varphi$ cannot be both an interior chart and a boundary chart.
If $\varphi(U)$ is an open subset of $R^n$, then it is an interior chart.
If $\varphi(U)$ is an open subset of $H^n$, then
- if $\varphi(U)\cap\partial H^n=\emptyset$, then $\varphi(U)$ is also an open subset of $R^n$, thus it is an interior chart;
- if $\varphi(U)\cap\partial H^n\neq\emptyset$, then it is a boundary chart.
$\blacksquare$
Notation.
Let $M$ be a smooth $n$-manifold with boundary, then the set of interior points of $M$ is denoted $\Int M$,
and the set of boundary points of $M$ is denoted $\partial M$.
Lemma.
Let $n\neq0$, let $U\subseteq R^n$ be open, and let $f:U\to R^n$ be a smooth function such that $Df(x)$ is invertible for all $x\in U$,
then $f(U)$ is open.
(show proof)
Proof.
Let $y\in f(U)$, then there exists $x\in U$ such that $f(x)=y$.
By inverse function theorem, there exists an open subset $V$ of $U$ containing $x$ such that $f(V)$ is open.
Since $y=f(x)\in f(V)$, there exists an open ball $B$ centered at $y$ such that $B\subseteq f(V)\subseteq f(U)$.
Therefore $f(U)$ is open.
$\blacksquare$
Lemma.
Let $n\neq0$, let $U$ be an open subset of $R^n$ and $V$ be an open subset of $H^n$ such that $V\cap\partial H^n\neq\emptyset$,
then $U$ and $V$ are not diffeomorphic.
(show proof)
Proof.
Suppose for contradiction that $U$ and $V$ are diffeomorphic, then there exists a diffeomorphism $f:U\to V$.
Since $V\cap\partial H^n\neq\emptyset$, there exists $y\in V\cap\partial H^n$.
Let $x$ denote $f^{-1}(y)$.
Then there exists an open subset $W$ of $R^n$ containing $y$ and a smooth function $g:W\to R^n$ that agrees with $f^{-1}$ on $W\cap V$.
And we can find an open ball $B$ centered at $x$ such that $B\subseteq U$ and $f(B)\subseteq W\cap V$.
Then $g\circ f|_B$ is the identity function on $B$.
By chain rule, for every $u\in B$, $I_n=D(g\circ f|_B)(u)=Dg(f|_B(u))D(f|_B)(u)$, implying $D(f|_B)(u)$ is invertible.
Therefore, $f(B)$ is an open subset of $R^n$ by the above lemma.
Since $y=f(x)\in f(B)\subseteq V\subseteq H^n$,
there exists an open ball centered at $y$ that is a subset of $H^n$.
But we also have $y\in\partial H^n$, a contradiction.
$\blacksquare$
Proposition.
Every point of a smooth manifold with boundary is either an interior point or a boundary point, but not both.
(show proof)
Proof.
Let $M$ be a smooth $n$-manifold with boundary.
If $n=0$, then every point of $M$ is an interior point and not a boundary point.
Now suppose $n\neq0$.
Let $p\in M$. Then some smooth chart $(\varphi,U)$ covers $p$.
- If $\varphi(U)$ is an open subset of $R^n$, then $p$ is an interior point.
- If $\varphi(U)$ is an open subset of $H^n$:
- if $\varphi(p)\in\partial H^n$, then $p$ is a boundary point,
- if $\varphi(p)\in\Int H^n$, let $B$ be an open ball centered at $\varphi(p)$ covered by the codomain of $\varphi$,
then the preimage $A$ of $B$ is an open subset of $M$, and $\varphi|_A:A\to B$ is a smooth chart
that covers $p$ with the codomain $B$ being an open subset of $R^n$, hence $p$ is an interior point.
We have shown that $p$ is either an interior point or a boundary point.
Suppose for contradiction that $p\in M$ is both an interior point and a boundary point.
Then there exist a smooth chart $(\varphi_1,U_1)$ covering $p$ such that $\varphi_1(U_1)$ is an open subset of $R^n$
and a smooth chart $(\varphi_2,U_2)$ covering $p$ such that $\varphi_2(U_2)$ is an open subset of $H^n$ and $\varphi_2(p)\in\partial H^n$.
Since $\varphi_1$ and $\varphi_2$ are compatible, $\varphi_2\circ\varphi_1^{-1}:\varphi_1(U_1\cap U_2)\to\varphi_2(U_1\cap U_2)$ is a diffeomorphism.
Since $U_1\cap U_2$ is an open subset of both $U_1$ and $U_2$,
$\varphi_1(U_1\cap U_2)$ is an open subset of $\varphi_1(U_1)$ and $\varphi_2(U_1\cap U_2)$ is an open subset of $\varphi_2(U_2)$.
Thus $\varphi_1(U_1\cap U_2)$ is an open subset of $R^n$ and $\varphi_2(U_1\cap U_2)$ is an open subset of $H^n$.
Since $p\in U_1\cap U_2$, $\varphi_2(p)\in\varphi_2(U_1\cap U_2)$, and since we also have $\varphi_2(p)\in\partial H^n$,
we have $\varphi_2(U_1\cap U_2)\cap\partial H^n\neq\emptyset$.
Then by the above lemma, $\varphi_1(U_1\cap U_2)$ and $\varphi_2(U_1\cap U_2)$ are not diffeomorphic, a contradiction.
$\blacksquare$
$0$-dimensional smooth manifold with boundary
Let $M$ be a $0$-dimensional smooth manifold with boundary. Then $M$ is countable and has no boundary point.
Let $p\in M$, then there exists a unique smooth chart $(\varphi,U)$ of $M$ such that $p\in U$.
And we have $U=\{p\}$ and $\varphi(U)=R^0$.
(show proof)
Proof.
Let $p\in M$, then there exists a smooth chart $(\varphi,U)$ of $M$ such that $p\in U$.
Since $\varphi(p)\in\varphi(U)$, $\varphi(U)$ is non-empty and equals $R^0$, implying $p$ is an interior point.
By injectivity of $\varphi$, $U=\{p\}$.
Since any smooth chart $(\varphi,U)$ of $M$ such that $p\in U$ would satisfy these conditions, such chart is unique.
Since $M$ is second-countable, there exists a countable basis $B$ of $M$.
Let $p\in M$, then $\{p\}$ is open, thus there exists a subset $S$ of $B$ such that $\{p\}=\cup S$.
For all $s\in S$, $s\subseteq\{p\}$, thus $s$ is either $\emptyset$ or $\{p\}$.
Since not all $s\in S$ is empty, some $s\in S$ is $\{p\}$, thus $\{p\}\in B$.
Thus $$\abs{M}=\abs{\{\{p\}:p\in M\}}\le\abs{B}\le\omega$$
$\blacksquare$
Standard half-Euclidean space
Since $H^n$ is a subspace of $R^n$, it is trivially an $n$-manifold with boundary.
Together with the smooth structure determined by the smooth atlas $\{I_{H^n}\}$,
$H^n$ is a smooth $n$-manifold with boundary, called standard $H^n$.
Clearly, the interior of $H^n$ is $\Int H^n$ and the boundary of $H^n$ is $\partial H^n$.
Proposition.
A smooth $n$-manifold is a smooth $n$-manifold with boundary.
(show proof)
Proof.
Let $M$ be a smooth $n$-manifold, then it is an $n$-manifold with boundary.
The smooth structure $\mathcal A$ of $M$ is clearly a smooth atlas of $M$ as a manifold with boundary.
Now suppose for contradiction that we have a proper extension $\mathcal A'$ of $\mathcal A$ that is a smooth atlas of $M$ as a manifold with boundary.
If some chart $(\varphi,U)$ in $\mathcal A'$ has $\varphi(U)$ being an open subset of $H^n$ such that $\varphi(U)\cap\partial H^n\neq\emptyset$,
then there exists $p\in U$ such that $\varphi(p)\in\partial H^n$, implying $n\neq0$.
Note that for some chart $(\psi,V)$ in $\mathcal A$ and thus in $\mathcal A'$, $p\in V$ and $\psi(V)$ is an open subset of $R^n$.
Then $\psi\circ\varphi^{-1}$ is a diffeomorphism, a contradiction to a lemma above.
Thus every chart $(\varphi,U)$ in $\mathcal A'$ has $\varphi(U)$ being an open subset of $R^n$,
implying $\mathcal A'$ is a smooth atlas of $M$ as a manifold without boundary, contradicting to $\mathcal A$ being maximal.
We have shown that $\mathcal A$ is a smooth structure of $M$ as a manifold with boundary.
Hence $M$ is a smooth $n$-manifold with boundary.
$\blacksquare$
Note.
Due to the above proposition, from this point on, if we want to discuss both smooth manifolds and smooth manifolds with boundary,
we will only discuss smooth manifolds with boundary.
Proposition.
Let $M$ be a smooth $n$-manifold with boundary.
- $\Int M$ is an open submanifold of $M$ without boundary.
- $\partial M$ is a smooth $(n-1)$-manifold if $n\neq0$, where every boundary chart of $M$ restricts to a smooth chart of $\partial M$ (with the last coordinate removed).
- $M$ is a smooth $n$-manifold if and only if $\partial M=\emptyset$.
(show proof)
Proof.
Let $p\in\Int M$, then for some interior chart $(\varphi,U_p)$, $p\in U_p$.
For every $u\in U_p$, we can find an open ball $B_u\subseteq\varphi(U_p)$ centered at $\varphi(u)$,
and $(\varphi,\varphi^{-1}(B_u))$ is an interior chart and $u\in\varphi^{-1}(B_u)$, implying $u\in\Int M$.
Thus $U_p\subseteq\Int M$.
Hence $\Int M=\bigcup_{p\in\Int M}U_p$ is an open subset of $M$.
Then $\Int M$ is an open submanifold with boundary of $M$.
Clearly, every point of $\Int M$ is locally Euclidean, thus $\Int M$ is an $n$-manifold.
Suppose for contradiction that some smooth chart $(\varphi,U)$ of $\Int M$ is a boundary chart.
Then for some $p\in U$, $\varphi(p)\in\partial H^n$, thus $p$ is a boundary point of $\Int M$.
But $p$ is an interior point of $M$ and thus an interior point of $\Int M$ as shown above, a contradiction.
Thus every smooth chart of $\Int M$ is an interior chart, and the smooth structure of $\Int M$ as a manifold with boundary is a smooth structure of $\Int M$ as a manifold.
Hence $\Int M$ is a smooth $n$-manifold, without boundary.
Suppose $n\neq0$.
$\partial M$ is clearly Hausdorff and second-countable.
Let $p\in\partial M$, then there exists a boundary chart $(\varphi,U)$ such that $p\in U$ and $\varphi(p)\in\partial H^n$.
Let $S$ denote $\varphi(U)\cap\partial H^n$ and define $f:R^n\to R^{n-1}$ by $f(x_1,\ldots,x_n)=(x_1,\ldots,x_{n-1})$,
then $f|_S:S\to f(S)$ is a diffeomorphism and thus a homeomorphism.
Let $y\in f(S)$, then $(f|_S)^{-1}(y)=(y_1,\dots,y_{n-1},0)\in S$, which we will denote as $x$. Note that for some open subset $V$ of $R^n$, $V\cap H^n=\varphi(U)$.
And we have $x\in V$. Then for some $r\gt0$, $B_r(x)\subseteq V$.
Let $y'\in B_r(y)$, and denote $(f|_S)^{-1}(y')=(y'_1,\dots,y'_{n-1},0)$ as $x'$, then $x'\in B_r(x)$, thus $x'\in V\cap\partial H^n$, implying $x'\in S$.
Hence $y'=(f|_S)(x')\in f(S)$.
We have shown that $f(S)$ is an open subset of $R^{n-1}$.
Note that $\varphi^{-1}|_S:S\to\varphi^{-1}(S)$ is homeomorphic and $\varphi(p)\in S$, thus $p\in\varphi^{-1}(S)$.
Let $q\in U\cap\partial M$. If $\varphi(q)\in\Int H^n$, then $q$ is an interior point, a contradiction.
Thus $\varphi(q)\in S$ and $q\in\varphi^{-1}(S)$.
Let $q\in\varphi^{-1}(S)$, then $q\in U$ and $\varphi(q)\in\partial H^n$, implying $q\in U\cap\partial M$.
We have shown that $\varphi^{-1}(S)=U\cap\partial M$, which is an open subset of $\partial M$.
Hence $f|_S\circ\varphi|_{U\cap\partial M}:U\cap\partial M\to f(S)$ is a homeomorphism from an open subset of $\partial M$ containing $p$ to an open subset of $R^{n-1}$.
Therefore, $\partial M$ is an $(n-1)$-manifold.
Similarly, we can see that for every boundary chart $(\varphi,U)$ of $M$, $f\circ\varphi|_{U\cap\partial M}$ (with codomain being the range) is a chart of $\partial M$,
and these charts form a smooth atlas, which generates a smooth structure of $\partial M$.
If $M$ is a smooth $n$-manifold, then every point of $M$ is an interior point, thus $\partial M=\emptyset$.
If $\partial M=\emptyset$, then $M=\Int M$, which is a smooth $n$-manifold.
$\blacksquare$
Smooth function on manifolds
Let $M$ be a smooth $n$-manifolds with boundary.
A function $f:M\to R^m$ is said to be smooth if for all $p\in M$,
there exists a smooth chart $(\varphi,U)$ where $p\in U$
such that $f\circ\varphi^{-1}:\varphi(U)\to R^m$ is smooth.
In particular, the set of smooth real-valued functions defined on $M$ is denoted $C^\infty(M)$.
Lemma.
Let $M$ be a smooth $n$-manifolds with boundary.
If $f:M\to R^m$ and $g:M\to R^k$ are smooth, then
for all $p\in M$, there exists a smooth chart $(\varphi,U)$ where $p\in U$
such that both $f\circ\varphi^{-1}$ and $g\circ\varphi^{-1}$ are smooth.
(show proof)
Proof.
Let $p\in M$, then
there exists a smooth chart $(\varphi_f,U_f)$ where $p\in U_f$
such that $f\circ\varphi_f^{-1}$ is smooth, and
there exists a smooth chart $(\varphi_g,U_g)$ where $p\in U_g$
such that $g\circ\varphi_g^{-1}$ is smooth.
Note that $U_f\cap U_g$ is an open subset of both $U_f$ and $U_g$ and $p\in U_f\cap U_g$.
Thus $(\varphi_f,U_f\cap U_g)$ and $(\varphi_g,U_f\cap U_g)$ are smooth charts of $M$,
implying $\varphi_g\circ\varphi_f^{-1}$ is smooth.
Hence $g\circ\varphi_f^{-1}=(g\circ\varphi_g^{-1})\circ(\varphi_g\circ\varphi_f^{-1})$ is smooth.
$\blacksquare$
Proposition.
Let $M$ be a smooth $n$-manifolds with boundary.
- If $f:M\to R^m$ and $g:M\to R^m$ are smooth, then $f+g$ is smooth.
- If $f:M\to R^m$ and $c:M\to R$ are smooth, then $cf$ is smooth.
(show proof)
Proof.
Let $p\in M$, then there exists a smooth chart $(\varphi,U)$ where $p\in U$
such that both $f\circ\varphi^{-1}$ and $g\circ\varphi^{-1}$ are smooth.
Thus $(f+g)\circ\varphi^{-1}=(f\circ\varphi^{-1})+(g\circ\varphi^{-1})$ is smooth.
Hence $f+g$ is smooth.
Let $p\in M$, then there exists a smooth chart $(\varphi,U)$ where $p\in U$
such that both $f\circ\varphi^{-1}$ and $c\circ\varphi^{-1}$ are smooth.
Thus $(cf)\circ\varphi^{-1}=(c\circ\varphi^{-1})(f\circ\varphi^{-1})$ is smooth.
Hence $cf$ is smooth.
$\blacksquare$
Note.
The above proposition implies that given a smooth $n$-manifold with boundary $M$:
- for all $f,g\in C^\infty(M)$, $f+g,fg\in C^\infty(M)$;
- the set of smooth functions from $M$ to $R^m$ is a vector space; in particular, $C^\infty(M)$ is a vector space.
Smooth map between manifolds
Let $M,N$ be smooth manifolds with boundary.
A map $f:M\to N$ is said to be smooth if for all $p\in M$,
there exist a smooth chart $(\varphi_M,U)$ where $p\in U$ and
a smooth chart $(\varphi_N,V)$ such that $f(U)\subseteq V$
and $\varphi_N\circ f\circ\varphi_M^{-1}:\varphi_M(U)\to\varphi_N(V)$ is smooth.
Given smooth $f:M\to N$ and any charts $(\varphi_M,U)$ of $M$ and $(\varphi_N,V)$ of $N$, $\varphi_N\circ f\circ\varphi_M^{-1}:\varphi_M(U\cap f^{-1}(V))\to\varphi_N(V)$
is called a coordinate representation of $f$.
Diffeomorphism between manifolds
Let $M,N$ be smooth manifolds with boundary.
A smooth bijective map $f:M\to N$ that has a smooth inverse is called a diffeomorphism.
If a diffeomorphism between $M$ and $N$ exists, they are called diffeomorphic.
Proposition.
A smooth map between smooth manifolds with boundary is continuous.
(show proof)
Proof.
Let $M,N$ be smooth manifolds with boundary.
Let $f:M\to N$ be a smooth map.
Let $B$ be an open subset of $N$ and let $A$ be the preimage of $f$ on $B$.
Let $p\in A$, then there exist a smooth chart $(\varphi_M,U)$ where $p\in U$ and
a smooth chart $(\varphi_N,V)$ where $f(p)\in V$ such that $f(U)\subseteq V$
and $\varphi_N\circ f\circ\varphi_M^{-1}:\varphi_M(U)\to\varphi_N(V)$ is smooth, and thus continuous.
Then $f|_U=\varphi_N^{-1}\circ(\varphi_N\circ f\circ\varphi_M^{-1})\circ\varphi_M$,
which is continuous.
Note that $V\cap B$ is an open subset of $N$, and $U\cap A$ is the preimage of $f|_U$ on $V\cap B$ and is thus open with respect to $U$ and thus $M$.
We have shown that for all $p\in A$, there exists an open subset $U_p$ of $M$ such that $p\in U_p\subseteq A$.
Note that $\cup\{U_p:p\in A\}=A$, hence $A$ is open.
And we have shown that $f$ is continuous.
$\blacksquare$
Localness of smoothness
Let $M,N$ be smooth manifolds with boundary and $f:M\to N$.
- If $f$ is smooth, then its restriction on any open submanifold of $M$ is smooth.
- If for each $p\in M$, there exists an open submanifold $U$ of $M$ where $p\in U$ such that $f|_U$ is smooth, then $f$ is smooth.
(show proof)
Proof.
Suppose $f$ is smooth and $U$ is an open submanifold of $M$.
Let $p\in U$, then there exist a smooth chart $(\varphi_M,V)$ where $p\in V$ and
a smooth chart $(\varphi_N,W)$ where $f(p)\in W$
such that $f(V)\subseteq W$
and $\varphi_N\circ f\circ\varphi_M^{-1}:\varphi_M(V)\to\varphi_N(W)$ is smooth.
Note that $(\varphi_M,U\cap V)$ is a smooth chart on $U$,
and $\varphi_N\circ f\circ\varphi_M^{-1}:\varphi_M(U\cap V)\to\varphi_N(W)$ is a restriction
of $\varphi_N\circ f\circ\varphi_M^{-1}:\varphi_M(V)\to\varphi_N(W)$ and is thus smooth.
Thus the restriction of $f$ on $U$ is smooth.
Suppose for each $p\in M$, there exists an open submanifold $U$ of $M$ where $p\in U$ such that $f|_U$ is smooth.
Let $p\in M$, then there exists an open submanifold $U$ of $M$ where $p\in U$ such that
there exist a smooth chart $(\varphi_U,V)$ of $U$ where $p\in V$ and
a smooth chart $(\varphi_N,W)$ where $f(p)\in W$
such that $f(V)\subseteq W$
and $\varphi_N\circ f\circ\varphi_U^{-1}:\varphi_U(V)\to\varphi_N(W)$ is smooth.
Note that $(\varphi_U,V)$ is also a smooth chart of $M$.
We have shown that $f$ is smooth.
$\blacksquare$
Note.
By localness of smoothness, it is trivial that $f:M\to N$ is smooth if and only if $f|_U$ is smooth for every smooth chart $(\varphi,U)$ of $M$.
Proposition.
Let $M,N$ be smooth manifolds with boundary,
then every coordinate representation of a smooth map $f:M\to N$ is smooth.
(show proof)
Proof.
Let $(\psi_M,A)$ and $(\psi_N,B)$ be charts of $M$ and $N$.
Let $\hat f$ denote $\psi_N\circ f\circ\psi_M^{-1}:\psi_M(A\cap f^{-1}(B))\to\psi_N(B)$.
Let $q\in\psi_M(A\cap f^{-1}(B))$ and let $p$ denote $\psi_M^{-1}(q)$, then $p\in A\cap f^{-1}(B)$.
Thus there exist a smooth chart $(\varphi_M,U)$ where $p\in U$ and
a smooth chart $(\varphi_N,V)$ where $f(p)\in V$
such that $f(U)\subseteq V$
and $\varphi_N\circ f\circ\varphi_M^{-1}:\varphi_M(U)\to\varphi_N(V)$ is smooth.
Since $(\varphi_M,U)$ and $(\psi_M,A\cap f^{-1}(B))$ are compatible, $\varphi_M\circ\psi_M^{-1}$ is smooth.
Thus $\varphi_N\circ f\circ\psi_M^{-1}:\psi_M(U\cap A\cap f^{-1}(B))\to\varphi_N(V\cap B)$ is smooth.
Since $(\varphi_N,V)$ and $(\psi_N,B)$ are compatible, $\psi_N\circ\varphi_N^{-1}$ is smooth.
Thus $\psi_N\circ f\circ\psi_M^{-1}:\psi_M(U\cap A\cap f^{-1}(B))\to\psi_N(V\cap B)$ is smooth.
Since $p\in U\cap A\cap f^{-1}(B)$, $q=\psi_M(p)\in\psi_M(U\cap A\cap f^{-1}(B))$.
Note that $\psi_M(U\cap A\cap f^{-1}(B))$ is open.
We have shown that $\hat f$ is smooth.
$\blacksquare$
Note.
By the above proposition, it is trivial that
the set of smooth functions from a smooth $n$-manifolds with boundary $M$ to $R^m$ is the set of smooth maps from $M$ to standard $R^m$, as a smooth $m$-manifold.
Thus the notation $C^\infty(M)$ may refer to either without ambiguity.
Proposition.
Let $M,N,P$ be smooth manifolds with boundary.
If $f:M\to N$ and $g:N\to P$ are smooth, then $g\circ f:M\to P$ is smooth.
(show proof)
Proof.
Let $p\in M$, then $f(p)\in N$.
Thus there exist a smooth chart $(\varphi_N,V)$ where $f(p)\in V$ and
a smooth chart $(\varphi_P,W)$ where $g(f(p))\in W$
such that $g(V)\subseteq W$
and $\varphi_P\circ g\circ\varphi_N^{-1}:\varphi_N(V)\to\varphi_P(W)$ is smooth.
Let $U$ denote $f^{-1}(V)$, then $U$ is open and $p\in U$.
Let $(\varphi_M,S)$ be a smooth chart of $M$ where $p\in S$.
Then $(\varphi_M,U\cap S)$ is a smooth chart of $M$ and $p\in U\cap S$.
And $\varphi_N\circ f\circ\varphi_M^{-1}:\varphi_M(U\cap S)\to\varphi_N(V)$ is smooth.
Thus $\varphi_P\circ(g\circ f)\circ\varphi_M^{-1}:\varphi_M(U\cap S)\to\varphi_P(W)$ is smooth.
Note that $(g\circ f)(U\cap S)\subseteq W$.
We have shown that $g\circ f:M\to P$ is smooth.
$\blacksquare$
Proposition.
Let $M,N,P$ be smooth manifolds with boundary.
If $f:M\to N$ and $g:N\to P$ are diffeomorphic, then $g\circ f:M\to P$ is diffeomorphic.
(show proof)
Proof.
This follows directly from the above proposition.
$\blacksquare$
Proposition.
Let $M,N$ be smooth $n$-manifolds with boundary, let $U$ be an open submanifold of $M$, and let $f:M\to N$ be a diffeomorphism,
then $f|_U:U\to f(U)$ is a diffeomorphism.
(show proof)
Proof.
$f|_U:U\to f(U)$ is trivially bijective.
Since $f$ is a diffeomorphism, it is a homeomorphism.
Thus $f(U)$ is an open subset of $N$, and thus defines an open submanifold of $N$.
Let $p\in U$, then for some smooth chart $(\varphi,A)$ of $M$, $p\in A$, and for some smooth chart $(\psi,B)$ of $N$, $f(p)\in B$.
Note that $(\psi,B\cap f(U))$ is a smooth chart of $f(U)$, as an open submanifold of $N$, and $f(p)\in B\cap f(U)$.
Also note that $(\varphi,A\cap f^{-1}(B\cap f(U)))$ is a smooth chart of $U$, as an open submanifold of $M$, and $p\in A\cap f^{-1}(B\cap f(U))$.
Since $f(A\cap f^{-1}(B\cap f(U)))\subseteq B\cap f(U)$, $\psi\circ f\circ\varphi^{-1}:\varphi(A\cap f^{-1}(B\cap f(U)))\to\psi(B\cap f(U))$
is a coordinate representation of $f$ and is thus smooth.
We have shown that $f|_U$ is smooth.
Note that $(f|_U)^{-1}=f^{-1}|_{f(U)}$, which is smooth for a symmetric argument.
Hence $f|_U:U\to f(U)$ is a diffeomorphism.
$\blacksquare$
Proposition.
Let $M,N$ be smooth $n$-manifolds with boundary, and let $f:M\to N$ be a diffeomorphism,
then $f(\Int M)=\Int N$, $f(\partial M)=\partial N$, and $f|_{\Int M}:\Int M\to\Int N$ is a diffeomorphism.
(show proof)
Proof.
Suppose for contradiction that $p\in\Int M$ and $f(p)\in\partial N$.
Then there exist a smooth chart $(\varphi_M,U)$ of $M$ where $p\in U$ and $\varphi_M(U)$ is an open subset of $R^n$, and
a smooth chart $(\varphi_N,V)$ of $N$ where $f(p)\in V$ and $\varphi_N(V)$ is an open subset of $H^n$,
such that $f(U)=V$ and $\varphi_N(f(p))\in\partial H^n$.
Since coordinate representations of smooth maps are smooth, $\varphi_N\circ f\circ\varphi_M^{-1}:\varphi_M(U)\to\varphi_N(V)$ is a diffeomorphism.
Since $f(p)\in V$, $\varphi_N(f(p))\in\varphi_N(V)$. Thus $\varphi_N(V)\cap\partial H^n\neq\emptyset$.
Since $\partial H^n\neq\emptyset$, $n\neq0$.
Then by a lemma above, $\varphi_M(U)$ and $\varphi_N(V)$ are not diffeomorphic, a contradiction.
Hence if $p\in\Int M$, then $f(p)\in\Int N$.
By a symmetric argument, if $q\in\Int N$, then $f^{-1}(q)\in\Int M$.
If $y\in f(\Int M)$, then for some $x\in\Int M$, $f(x)=y$, thus $y\in\Int N$.
If $y\in\Int N$, then $f^{-1}(y)\in\Int M$ and $f(f^{-1}(y))=y$, thus $y\in f(\Int M)$.
Hence $f(\Int M)=\Int N$.
If $y\in f(\partial M)$, then for some $x\in\partial M$, $f(x)=y$, then $y\notin\Int N$, thus $y\in\partial N$.
If $y\in\partial N$, then $f^{-1}(y)\notin\Int M$, thus $f^{-1}(y)\in\partial M$ and $f(f^{-1}(y))=y$, implying $y\in f(\partial M)$.
Hence $f(\partial M)=\partial N$.
Since $\Int M$ is an open submanifold of $M$, $f|_{\Int M}:\Int M\to\Int N$ is a diffeomorphism.
$\blacksquare$
Proposition.
Let $M$ be a smooth $m$-manifold with boundary and $N$ be a smooth $n$-manifold with boundary such that
$M$ and $N$ are non-empty and diffeomorphic, then $m=n$.
(show proof)
Proof.
Suppose $M$ and $N$ are smooth manifolds without boundary.
Since $M$ is non-empty, there exists $p\in M$.
Since $M$ and $N$ are diffeomorphic, there exists a diffeomorphism $f:M\to N$.
Then there exist a smooth chart $(\varphi_M,U)$ of $M$ where $p\in U$ and
a smooth chart $(\varphi_N,V)$ of $N$ where $f(p)\in V$ such that $f(U)=V$.
Since coordinate representations of smooth maps are smooth, $\varphi_N\circ f\circ\varphi_M^{-1}:\varphi_M(U)\to\varphi_N(V)$ is a diffeomorphism.
Since $\varphi_M(U)$ is a non-empty open subset of $R^m$ and $\varphi_N(V)$ is a non-empty open subset of $R^n$, $m=n$.
Suppose $M$ and $N$ are smooth manifolds with boundary.
Since $M$ and $N$ are diffeomorphic, there exists a diffeomorphism $f:M\to N$.
Then $f|_{\Int M}:\Int M\to\Int N$ is a diffeomorphism.
Thus $\Int M$ is a smooth $m$-manifold, $\Int N$ is a smooth $n$-manifold, and $\Int M$ and $\Int N$ are diffeomorphic.
Since $M$ is non-empty, there exists $p\in M$.
If $p\in\Int M$, then $\Int M$ is non-empty.
If $p\in\partial M$, then for some boundary chart $(\varphi,U)$, $p\in U$ and $\varphi(p)\in\partial H^n$.
Note that for some open subset $V$ of $R^n$, $V\cap H^n=\varphi(U)$.
Since $\varphi(p)\in V$, for some $r\gt0$, $B_r(\varphi(p))\subseteq V$.
Let $q$ denote $\varphi(p)+\frac{r}{2}\vb e_n$, then $B_{\frac{r}{2}}(q)\subseteq V\cap H^n=\varphi(U)$.
Hence $(\varphi,\varphi^{-1}(B_{\frac{r}{2}}(q)))$ is an interior chart of $M$.
Since $\varphi^{-1}(q)\in\varphi^{-1}(B_{\frac{r}{2}}(q))$, $\varphi^{-1}(q)\in\Int M$, implying $\Int M$ is non-empty.
By a symmetric argument, since $N$ is non-empty, $\Int N$ is non-empty.
Therefore, $m=n$.
$\blacksquare$
Local diffeomorphism
Let $M,N$ be smooth $n$-manifolds with boundary.
A map $F:M\to N$ is called a local diffeomorphism if for all $p\in M$,
there exists an open subset $U$ of $M$
such that $p\in U$, $F(U)$ is open, and $F|_U:U\to F(U)$ is a diffeomorphism.
Clearly, a diffeomorphism is a local diffeomorphism, and a local diffeomorphism is smooth.
Coordinate ball
Given an open ball $B$ of $R^n$ centered on $\partial H^n$, we call $B\cap H^n$ an open half ball of $H^n$.
Let $M$ be a manifold with boundary and $B\subseteq M$.
If there exists a chart $(\varphi,B)$ of $M$ such that $\varphi(B)$ is an open (half) ball, then $B$ is called a coordinate (half) ball.
If in addition, $M$ is smooth and $(\varphi,B)$ is a smooth chart of $M$,
then $B$ is called a smooth (half) ball.
If in addition, there exists a smooth (half) ball $B'$ of $M$, where $\overline B\subseteq B'$, and a smooth chart $(\varphi,B')$ of $M$
such that for some $0\lt r\lt r'$, $\varphi(B)=B_r(0)(\cap H^n)$, $\varphi(\overline B)=\overline{B_r(0)(\cap H^n)}$, and $\varphi(B')=B_{r'}(0)(\cap H^n)$,
then $B$ is called a regular (half) ball.
Lemma.
Let $M$ be a manifold with boundary and let $(\varphi,B)$ be a chart of $M$ such that $\varphi(B)=B_r(0)(\cap H^n)$ for some $r\gt0$,
then for all $s$ with $0\lt s\lt r$, $\varphi^{-1}(\overline{B_s(0)(\cap H^n)})=\overline{\varphi^{-1}(B_s(0)(\cap H^n))}$.
(show proof)
Proof.
Let $p\in\varphi^{-1}(\overline{B_s(0)(\cap H^n)})$. Then $\varphi(p)\in\overline{B_s(0)(\cap H^n)}$.
Thus there exists a sequence $(q_i)$ of $B_s(0)(\cap H^n)$ that converges to $\varphi(p)$.
And we can define a sequence $(p_i)$ of $\varphi^{-1}(B_s(0)(\cap H^n))$ such that $p_i=\varphi^{-1}(q_i)$.
Let $U$ be a neighborhood of $p$, then it contains an open subset $V$ of $M$ such that $p\in V$.
Note that $\varphi(V\cap B)$ is an open subset of $B_r(0)(\cap H^n)$ and $\varphi(p)\in\varphi(V\cap B)$,
thus for some $t\gt0$, $B_t(\varphi(p))(\cap H^n)\subseteq\varphi(V\cap B)$.
Since $(q_i)$ converges to $\varphi(p)$, for some $n\in N$, for all $k\ge n$, $q_k\in B_t(\varphi(p))(\cap H^n)$,
implying $p_k=\varphi^{-1}(q_k)\in V\cap B\subseteq U$.
We have shown that $(p_i)$ converges to $p$.
Hence $p\in\overline{\varphi^{-1}(B_s(0)(\cap H^n))}$.
Let $p\in\overline{\varphi^{-1}(B_s(0)(\cap H^n))}$.
Then there exists a sequence $(p_i)$ of $\varphi^{-1}(B_s(0)(\cap H^n))$ that converges to $p$.
And we can define a sequence $(q_i)$ of $B_s(0)(\cap H^n)$ such that $q_i=\varphi(p_i)$.
Since $(q_i)$ is bounded, it has a convergent subsequence $(q'_j)$.
Let $q$ denote the limit of $(q'_j)$, then $q\in\overline{B_s(0)(\cap H^n)}$ and $\varphi^{-1}(q)\in B$.
Then we can define a sequence $(p'_j)$ of $\varphi^{-1}(B_s(0)(\cap H^n))$ such that $p'_j=\varphi^{-1}(q'_j)$,
which is a subsequence of $(p_i)$ and thus converges to $p$.
By the same reasoning as the last paragraph, $(p'_j)$ converges to $\varphi^{-1}(q)$.
Thus $p=\varphi^{-1}(q)\in\varphi^{-1}(\overline{B_s(0)(\cap H^n)})$.
$\blacksquare$
Proposition.
Every smooth manifold (with boundary) has a countable basis of regular balls (or regular half balls).
(show proof)
Proof.
Let $M$ be a smooth $n$-manifold with boundary.
Then the domains of smooth charts of $M$ forms an open cover of $M$, which has countable subcover $\mathcal C$.
Let $U\in\mathcal C$. Then there exists a smooth chart $(\varphi_U,U)$ of $M$.
Let $\mathcal B_U$ denote the set of open balls $B_r(q)$ such that $r$ is rational, $q$ has rational coordinates, and $B_{2r}(q)\subseteq\varphi_U(U)$.
And let $\mathcal B'_U$ denote
- $\emptyset$ if $(\varphi_U,U)$ is an interior chart;
- the set of open half balls $B_r(q)\cap H^n$ such that $r$ is rational, $q$ has rational coordinates, and $B_{2r}(q)\cap H^n\subseteq\varphi_U(U)$, if $(\varphi_U,U)$ is a boundary chart.
Define $\mathcal A_U$ to be $\{\varphi_U^{-1}(B):B\in\mathcal B_U\}$ and $\mathcal A'_U$ to be $\{\varphi_U^{-1}(B):B\in\mathcal B'_U\}$.
Then $\bigcup_{U\in\mathcal C}\p{\mathcal A_U\cup\mathcal A'_U}$
is a countable set of subsets of $M$, which we will denote as $\mathcal A$.
Let $A\in\mathcal A$.
Then for some $U\in\mathcal C$, either $A\in\mathcal A_U$ or $A\in\mathcal A'_U$.
-
If $A\in\mathcal A_U$, then for some $B\in\mathcal B_U$, $A=\varphi_U^{-1}(B)$.
Note that $B=B_r(q)$ for some rational $r$ and $q$ such that $B_{2r}(q)\subseteq\varphi_U(U)$.
We define $A'$ to be $\varphi_U^{-1}(B_{2r}(q))$.
Regardless of whether $(\varphi_U,U)$ is an interior chart or a boundary chart,
$B_r(q)$ and $B_{2r}(q)$ are open subsets of $\varphi_U(U)$,
thus $A$ and $A'$ are open subsets of $U$ and thus $M$.
Define $h:R^n\to R^n$ such that $h(x)=x-q$,
then $(h\circ\varphi_U,A')$ is a smooth chart,
and we have $(h\circ\varphi_U)(A)=B_r(0)$ and $(h\circ\varphi_U)(A')=B_{2r}(0)$.
Note that $$\overline A=\overline{(h\circ\varphi_U)^{-1}(B_r(0))}=(h\circ\varphi_U)^{-1}(\overline{B_r(0)})\subseteq(h\circ\varphi_U)^{-1}(B_{2r}(0))=A'$$
And we have $(h\circ\varphi_U)(\overline A)=\overline{B_r(0)}$.
We have shown that $A$ is a regular ball.
-
If $A\in\mathcal A'_U$, then for some $B\in\mathcal B'_U$, $A=\varphi_U^{-1}(B)$.
Since $\mathcal B'_U$ is non-empty, $(\varphi_U,U)$ is a boundary chart,
and $B=B_r(q)\cap H^n$ for some rational $r$ and $q$ such that $B_{2r}(q)\cap H^n\subseteq\varphi_U(U)$.
We define $A'$ to be $\varphi_U^{-1}(B_{2r}(q)\cap H^n)$.
Since $(\varphi_U,U)$ is a boundary chart,
$B_r(q)\cap H^n$ and $B_{2r}(q)\cap H^n$ are open subsets of $\varphi_U(U)$,
thus $A$ and $A'$ are open subsets of $U$ and thus $M$.
Define $h:R^n\to R^n$ such that $h(x)=x-q$,
then $(h\circ\varphi_U,A')$ is a smooth chart,
and we have $(h\circ\varphi_U)(A)=B_r(0)\cap H^n$ and $(h\circ\varphi_U)(A')=B_{2r}(0)\cap H^n$.
Note that $$\overline A=\overline{(h\circ\varphi_U)^{-1}(B_r(0)\cap H^n)}=(h\circ\varphi_U)^{-1}(\overline{B_r(0)\cap H^n})\subseteq(h\circ\varphi_U)^{-1}(B_{2r}(0)\cap H^n)=A'$$
And we have $(h\circ\varphi_U)(\overline A)=\overline{B_r(0)\cap H^n}$.
We have shown that $A$ is a regular half ball.
Thus every element of $\mathcal A$ is a regular ball or a regular half ball.
Let $S$ be an open subset of $M$.
Then $S=\cup\{S\cap U:U\in\mathcal C\}$.
Let $U\in\mathcal C$, then $\varphi_U(S\cap U)$ is an open subset of $\varphi_U(U)$.
- Suppose $(\varphi_U,U)$ is an interior chart.
Then $\varphi_U(S\cap U)$ is an open subset of $R^n$.
Let $p\in\varphi_U(S\cap U)$, then for some $r\gt0$, $B_r(p)\subseteq\varphi_U(S\cap U)$.
Note that there exists $q\in B_{\frac{r}{4}}(p)$ such that $q$ has rational coordinates,
and there exists $s\in(\frac{r}{2},\frac{3r}{4})$ such that $s$ is rational.
Then $B_{\frac{s}{2}}(q)$ is an open ball containing $p$ such that
$\frac{s}{2}$ is rational, $q$ has rational coordinates, and $B_s(q)\subseteq B_r(p)\subseteq\varphi_U(S\cap U)\subseteq\varphi_U(U)$.
Thus $p\in\cup\{B\in\mathcal B_U|B\subseteq\varphi_U(S\cap U)\}$.
The other direction is trivial.
Hence $\varphi_U(S\cap U)=\cup\{B\in\mathcal B_U|B\subseteq\varphi_U(S\cap U)\}$.
Let $\mathcal B_{U,S}$ denote $\{B\in\mathcal B_U|B\subseteq\varphi_U(S\cap U)\}$,
then $S\cap U=\cup\{\varphi_U^{-1}(B):B\in\mathcal B_{U,S}\}$.
Note that $\{\varphi_U^{-1}(B):B\in\mathcal B_{U,S}\}\subseteq\mathcal A_U\subseteq\mathcal A_U\cup\mathcal A'_U\subseteq\mathcal A$.
Thus $S\cap U$ is the union of a subset of $\mathcal A$.
- Suppose $(\varphi_U,U)$ is a boundary chart.
Then $n\neq0$ and $\varphi_U(S\cap U)$ is an open subset of $H^n$.
Let $p\in\varphi_U(S\cap U)$.
- If $p\in\Int H^n$, then for some $r\gt0$, $B_r(p)\subseteq\varphi_U(S\cap U)$.
By the same reasoning as above, $p\in\cup\{B\in\mathcal B_U|B\subseteq\varphi_U(S\cap U)\}$
- If $p\in\partial H^n$, then for some $r\gt0$, $B_r(p)\cap H^n\subseteq\varphi_U(S\cap U)$.
Note that there exists $q\in B_{\frac{r}{4}}(p)$ such that $q$ has rational coordinates and $q\in\partial H^n$,
and there exists $s\in(\frac{r}{2},\frac{3r}{4})$ such that $s$ is rational.
Then $B_{\frac{s}{2}}(q)\cap H^n$ is an open half ball containing $p$ such that
$\frac{s}{2}$ is rational, $q$ has rational coordinates, and $B_s(q)\cap H^n\subseteq B_r(p)\cap H^n\subseteq\varphi_U(S\cap U)\subseteq\varphi_U(U)$.
Thus $p\in\cup\{B\in\mathcal B'_U|B\subseteq\varphi_U(S\cap U)\}$.
In either case, we have $p\in\cup\{B\in\mathcal B_U\cup\mathcal B'_U|B\subseteq\varphi_U(S\cap U)\}$.
The other direction is trivial.
Hence $\varphi_U(S\cap U)=\cup\{B\in\mathcal B_U\cup\mathcal B'_U|B\subseteq\varphi_U(S\cap U)\}$.
Let $\mathcal B_{U,S}$ denote $\{B\in\mathcal B_U|B\subseteq\varphi_U(S\cap U)\}$
and $\mathcal B'_{U,S}$ denote $\{B\in\mathcal B'_U|B\subseteq\varphi_U(S\cap U)\}$,
then $S\cap U=\cup\{\varphi_U^{-1}(B):B\in\mathcal B_{U,S}\cup\mathcal B'_{U,S}\}$.
Note that $\{\varphi_U^{-1}(B):B\in\mathcal B_{U,S}\cup\mathcal B'_{U,S}\}\subseteq\mathcal A_U\cup\mathcal A'_U\subseteq\mathcal A$.
Thus $S\cap U$ is the union of a subset of $\mathcal A$.
Since each $S\cap U$ is the union of a subset of $\mathcal A$, $S$ is the union of a subset of $\mathcal A$.
Thus $\mathcal A$ is a basis of $M$.
Note that if $M$ is a smooth manifold without boundary, then every smooth chart is an interior chart, thus
every element of $\mathcal A$ is a regular ball.
$\blacksquare$
Paracompactness
Let $X$ be a topological space.
A set $\mathcal F$ of subsets of $X$ is said to be locally finite
if for all $p\in X$, there exists a neighborhood $B$ of $p$ such that $\{S\in\mathcal F|S\cap B\neq\emptyset\}$ is finite.
Given covers $\mathcal C,\mathcal D$ of $X$, if for all $V\in\mathcal D$, there exists $U\in\mathcal C$
such that $V\subseteq U$, then $\mathcal D$ is said to be a refinement of $\mathcal C$.
If every open cover of $X$ has a locally finite open refinement, then $X$ is said to be paracompact.
Proposition.
Let $M$ be a smooth manifold with boundary and let $\mathcal C$ be a locally finite set of subsets of $M$.
- $\{\overline U:U\in\mathcal C\}$ is locally finite.
- $\overline{\bigcup\mathcal C}=\bigcup_{U\in\mathcal C}\overline U$.
(show proof)
Proof.
-
Let $p\in M$ then there exists an open set $B$ containing $p$ such that $\{S\in\mathcal C|S\cap B\neq\emptyset\}$, denoted $\mathcal D$, is finite.
Let $\mathcal C'$ denote $\mathcal C\setminus\mathcal D$,
then for all $S\in\mathcal C'$, we have $S\cap B=\emptyset$, thus $B\subseteq M\setminus S$, implying $B\subseteq\text{Ext} S$, and thus $\overline S\cap B=\emptyset$.
Hence $\{S\in\mathcal C|\overline S\cap B\neq\emptyset\}\subseteq\mathcal D$, which is finite.
Note that $S\mapsto\overline S$ defines a surjection from $\{S\in\mathcal C|\overline S\cap B\neq\emptyset\}$ to $\{S\in\{\overline U:U\in\mathcal C\}|S\cap B\neq\emptyset\}$.
Therefore, $\{\overline U:U\in\mathcal C\}$ is locally finite.
-
Let $p\in\overline{\cup\mathcal C}$.
Then $p$ has a neighborhood $B_p$ such that $\{S\in\mathcal C|S\cap B_p\neq\emptyset\}$, denoted $\mathcal D$, is finite,
and for every neighborhood $B$ of $p$, $\cup\mathcal C\cap(B\setminus\{p\})\neq\emptyset$.
Suppose for contradiction that $p\notin\cup_{U\in\mathcal C}\overline U$,
then for all $U\in\mathcal D$, $U\in\mathcal C$, then $p\notin\overline U$, thus $p\in\text{Ext} U$.
Then either $\mathcal D$ is empty or $p\in\bigcap_{U\in\mathcal D}\text{Ext} U$, which is open.
If $\mathcal D$ is empty, then $\cup\mathcal C\cap(B_p\setminus\{p\})=\emptyset$, a contradiction.
Otherwise, $B_p\cap\bigcap_{U\in\mathcal D}\text{Ext} U$, denoted $B'_p$, is a neighborhood of $p$,
and $\cup\mathcal C\cap(B'_p\setminus\{p\})=\emptyset$, a contradiction.
Thus $p\in\cup_{U\in\mathcal C}\overline U$.
Let $p\in\cup_{U\in\mathcal C}\overline U$.
Then for some $U\in\mathcal C$, $p\in\overline U\subseteq\overline{\cup\mathcal C}$.
$\blacksquare$
Proposition.
Every smooth manifold with boundary is locally compact.
(show proof)
Proof.
Let $M$ be a smooth manifold with boundary, then it has a countable basis $\mathcal B$ of regular balls or regular half balls.
Note that given an open ball $B$, $\overline B$ is closed and bounded, and thus compact in $R^n$.
Similarly, given an open half ball $B$, $\overline B$ is compact in $R^n$, and thus compact in $H^n$.
Let $p\in M$, then there exists $B\in\mathcal B$ such that $p\in B$, and $\overline B$ is a neighborhood of $p$.
If $B$ is a regular ball, then there exists $B'\subseteq M$ such that
$\overline B\subseteq B'$, and there exists a smooth chart $(\varphi,B')$ of $M$
such that for some $0\lt r\lt r'$, $\varphi(B)=B_r(0)$, $\varphi(\overline B)=\overline{B_r(0)}$, and $\varphi(B')=B_{r'}(0)$.
Note that $\overline{B_r(0)}$ is compact in $\varphi(B')$.
Since $\varphi^{-1}$ is continuous, $\overline B=\varphi^{-1}(\overline{B_r(0)})$ is compact in $B'$, and thus compact in $M$.
If $B$ is a regular half ball, then there exists $B'\subseteq M$ such that
$\overline B\subseteq B'$, and there exists a smooth chart $(\varphi,B')$ of $M$
such that for some $0\lt r\lt r'$, $\varphi(B)=B_r(0)\cap H^n$, $\varphi(\overline B)=\overline{B_r(0)\cap H^n}$, and $\varphi(B')=B_{r'}(0)\cap H^n$.
Note that $\overline{B_r(0)\cap H^n}$ is compact in $\varphi(B')$.
Since $\varphi^{-1}$ is continuous, $\overline B=\varphi^{-1}(\overline{B_r(0)\cap H^n})$ is compact in $B'$, and thus compact in $M$.
We have shown that there exists a compact neighborhood of $p$.
Hence $M$ is locally compact.
$\blacksquare$
Lemma.
Every second-countable, locally compact Hausdorff space $X$ has a sequence $(K_i)$ of compact subsets
such that $X=\cup_i K_i$ and $K_i\subseteq\Int K_{i+1}$ for all $i\in N$.
(show proof)
Proof.
Let $X$ be such a space. Since $X$ is Hausdorff and locally compact, it has a precompact basis $\mathcal B$.
Since $X$ is second-countable and $\mathcal B$ is an open cover of $X$,
it has a countable subcover $\mathcal C$, which can be indexed into a sequence $(C_i)$ (extend with $\emptyset$ if $\mathcal C$ is finite) of open precompact sets that covers $X$.
Suppose $A$ is a compact set and $n\in N$, then $(C_i)$ covers $A$ and thus has a finite subcover,
implying for some $k\in N$ such that $k\ge n+1$, $A\subseteq\cup_{j=0}^kC_j\subseteq\cup_{j=0}^k\overline{C_j}$.
Note that $\cup_{j=0}^k\overline{C_j}$ is compact, $n+1\in N$, and $A\subseteq\cup_{j=0}^kC_j\subseteq\Int\cup_{j=0}^k\overline{C_j}$.
This defines a function $(A,n)\mapsto(\cup_{j=0}^k\overline{C_j},n+1)$ where $k\ge n+1$ from and onto pairs formed by a compact subsets of $X$ and a natural number.
By recursion, we can define a sequence of pairs $(K_i,n_i)$
such that $(K_0,n_0)=(\overline{C_0},0)$ and $(K_{i+1},n_{i+1})=(\cup_{j=0}^k\overline{C_j},n_i+1)$ where $k\ge n_i+1$,
and from which we can extract a sequence of compact sets $(K_i)$.
Note that $n_i=i$ and $C_0\subseteq K_0$.
Let $i\in N^+$,
then $K_i=\cup_{j=0}^k\overline{C_j}$ where $k\ge n_{i-1}+1=i$,
thus $C_i\subseteq K_i$.
Hence $X\subseteq\cup_i C_i\subseteq\cup_i K_i\subseteq X$.
Note that $K_i\subseteq\Int K_{i+1}$ by the definition of the recursion function.
$\blacksquare$
Proposition.
Every second-countable, locally compact Hausdorff space is paracompact.
In particular, given such a space $X$, an open cover $\mathcal C$ of $X$, and a basis $\mathcal B$ of $X$,
there exists a locally finite open refinement of $\mathcal C$ that is a countable subset of $\mathcal B$.
(show proof)
Proof.
Define $(K_i)$ as in the last lemma, then $X=\cup_i K_i$ and $K_i\subseteq\Int K_{i+1}$ for all $i\in N$.
Let $K_{-1}=K_{-2}=\emptyset$ and
define $V_j=K_j\setminus\Int K_{j-1}$ and $W_j=\Int K_{j+1}\setminus K_{j-2}$ for $j\in N$.
Since $X$ is Hausdorff, compact sets are closed, thus each $V_j$ is closed and each $W_j$ is open.
Then each $V_j$ is a closed subset of a compact set, and is thus compact.
Note that $V_j\subseteq W_j$ for $j\in N$.
Let $p\in X$, then there exists $j\in N$ such that $p\in K_j$.
Thus there exists a least $k\in N$ such that $p\in K_k$, implying $p\in V_k$.
Then we have $X=\cup_{j\in N}V_j$.
Let $j\in N$, then for all $p\in V_j$, there exists $C_p\in\mathcal C$ such that $p\in C_p$.
Note that $p\in C_p\cap W_j$, which is open, thus there exists $B_p\in\mathcal B$ such that $p\in B_p\subseteq C_p\cap W_j$.
Then $\{B_p:p\in V_j\}$ is an open cover of $V_j$ and thus has a finite subcover $\mathcal B_j$, which is a subset of $\mathcal B$.
Then $\cup_{j\in N}\mathcal B_j$, denoted $\mathcal B'$, is an open refinement of $\mathcal C$ that is a countable subset of $\mathcal B$.
Note that for all $B\in\mathcal B'$, $B\subseteq W_j$ for some $j\in N$.
Also note that $W_j\cap W_{j+k+3}=\emptyset$ for all $j,k\in N$.
Let $p\in X$, then for some $B\in\mathcal B'$, $p\in B$, which is a neighborhood of $p$, and $B\subseteq W_j$ for some $j\in N$.
Then $\{S\in\mathcal B'|B\cap S\neq\emptyset\}\subseteq\cup\{\mathcal B_k:k\in N,k\lt j+3,j\lt k+3\}$, which is finite.
Hence $\mathcal B'$ is locally finite.
$\blacksquare$
Support
The support of a function $f:X\to R^m$, where $X$ is a topological space, is defined to be
$$\overline{\{p\in X|f(p)\neq0\}}$$
which is denoted $\text{supp}f$.
If $\text{supp}f\subseteq U$ for some subset $U$ of $X$, then we say $f$ is supported in $U$.
Partition of unity
Let $M$ be a topological space, let $A$ be a set serving as indexes, and let $(U_a)_{a\in A}$ be an open cover of $M$.
Then a
partition of unity with respect to $(U_a)_{a\in A}$ is an indexed set $(\psi_a)_{a\in A}$ of continuous real-valued functions defined on $M$ such that:
- For all $p\in M$ and $a\in A$, $0\le\psi_a(p)\le 1$.
- For all $a\in A$, $\psi_a$ is supported in $U_a$.
- $\{\text{supp}\psi_a:a\in A\}$ is locally finite.
- For all $p\in M$, $\sum_{a\in A}\psi_a(p)=1$.
Note that the summation $\sum_{a\in A}\psi_a(p)$ is a short-hand of $\sum_{a\in\{a\in A|\psi_a(p)\neq0\}}\psi_a(p)$,
which is well-defined if $\{\text{supp}\psi_a:a\in A\}$ is locally finite and thus $\{a\in A|\psi_a(p)\neq0\}$ is finite.
We will apply similar short-hands from here on.
If in addition, $M$ is a smooth manifold with boundary, and $\psi_a$ is smooth for all $a\in A$, then $(\psi_a)_{a\in A}$ is called a
smooth partition of unity.
Lemma.
The function
$$
f(t) =
\begin{cases}
e^{-\frac{1}{t}} & t\gt0\\
0 & t\le0
\end{cases}
$$
is smooth.
(show proof)
Proof.
Note that by limit of composite functions, $$\lim_{t\to 0^+}e^{-\frac{1}{t}}=0$$
Let $k\in N$ and suppose $\lim_{t\to 0^+}e^{-\frac{1}{t}}t^{-k}=0$, then
$$
\lim_{t\to 0^+}e^{-\frac{1}{t}}t^{-(k+1)}
=\lim_{t\to 0^+}\frac{t^{-(k+1)}}{e^{\frac{1}{t}}}
=\lim_{t\to 0^+}\frac{-(k+1)t^{-(k+2)}}{e^{\frac{1}{t}}(-t^{-2})}
=\lim_{t\to 0^+}\frac{(k+1)t^{-k}}{e^{\frac{1}{t}}}
=0
$$
By induction, for all $k\in N$, $\lim_{t\to 0^+}e^{-\frac{1}{t}}t^{-k}=0$.
Let $k\in N$ and suppose
$$
f^{(k)}(t) =
\begin{cases}
e^{-\frac{1}{t}}\sum_{j=0}^{2k}c_jt^{-j} & t\gt0\\
0 & t\le0
\end{cases}
$$
for some $c_0,\ldots,c_{2k}\in R$.
Note that for each $j$, with domain restricted to $R^+$, $$\p{t^{-j}e^{-\frac{1}{t}}}'=(-j)t^{-(j+1)}e^{-\frac{1}{t}}+t^{-(j+2)}e^{-\frac{1}{t}}$$
Thus $f^{(k+1)}$ takes the form $e^{-\frac{1}{t}}\sum_{j=0}^{2(k+1)}d_jt^{-j}$ for some $d_0,\ldots,d_{2(k+1)}\in R$ when $t\gt0$.
Note that $$\lim_{h\to 0^+}\frac{f^{(k)}(h)-f^{(k)}(0)}{h}=\lim_{h\to 0^+}e^{-\frac{1}{h}}\sum_{j=0}^{2k}c_jh^{-(j+1)}=0$$
Thus $f^{(k+1)}(0)=0$, which concludes the inductive step.
By induction, for all $k\in N$, $f^{(k)}$ exists on $R$.
Hence $f$ is smooth.
$\blacksquare$
Lemma.
Given $a,b\in R$ such that $a\lt b$, there exists a smooth function $h:R\to R$ such that
$$
h(t)
\begin{cases}
=1 & t\le a\\
\in(0,1) & t\in(a,b)\\
=0 & t\ge b
\end{cases}
$$
(show proof)
Proof.
Let $f$ be the function in the lemma above, then the function
$$\frac{f(b-t)}{f(b-t)+f(t-a)}$$
clearly satisfies the conditions.
$\blacksquare$
Lemma.
Given $p\in R^n$ and $r_1,r_2\in R^+$ such that $r_1\lt r_2$, there exists a smooth function $H:R^n\to R$ such that
$$
H(x)
\begin{cases}
=1 & d(x,p)\le r_1\\
\in(0,1) & d(x,p)\in(r_1,r_2)\\
=0 & d(x,p)\ge r_2
\end{cases}
$$
(show proof)
Proof.
Let $H(x)=h(d(x,p))$, where $h$ is the function obtained in the lemma above with $a$ being $r_1$ and $b$ being $r_2$.
Note that $d(x,p)=\norm{x-p}$ and the norm function is smooth on $R^n\setminus{0}$,
thus $H(x)$ is smooth on $R^n\setminus\{p\}$.
Since $H(x)$ is constant on $B_{r_1}(p)$, it is also smooth at $p$.
$\blacksquare$
Existence of smooth partition of unity
Let $M$ be a smooth manifold with boundary and let $(U_a)_{a\in A}$ be an open cover of $M$,
then there exists a smooth partition of unity with respect to $(U_a)_{a\in A}$.
(show proof)
Proof.
Let $a\in A$, then $U_a$ is an open submanifold of $M$, and thus has a basis $\mathcal B_a$ of regular balls or regular half balls.
Let $\mathcal B$ denote $\cup\{\mathcal B_a:a\in A\}$, then $\mathcal B$ is a basis of $M$.
Since $M$ is second-countable, locally compact, and Hausdorff,
there exists a locally finite open refinement of $\{U_a:a\in A\}$ that is a countable subset of $\mathcal B$,
which we will index into a bijection $(B_i)_{i\in k}$ where $k$ is countable.
Let $i\in k$, then for some $a\in A$, $B_i\in\mathcal B_a$, implying $B_i$ either a regular ball or a regular half ball of $U_a$.
Then for some smooth ball $B'_i$ of $U_a$ where $\overline{B_i}\subseteq B'_i$, there exists a smooth chart $(\varphi_i,B'_i)$ of $U_a$
such that for some $0\lt r_i\lt r'_i$, $\varphi_i(B_i)=B_{r_i}(0)(\cap H^n)$, $\varphi_i(\overline{B_i})=\overline{B_{r_i}(0)(\cap H^n)}$, and $\varphi_i(B'_i)=B_{r'_i}(0)(\cap H^n)$.
Let $H_i:R^n\to R$ be a smooth function such that
$$
H_i(x)
\begin{cases}
=1 & d(x,0)\le\frac{r_i}{2}\\
\in(0,1) & d(x,0)\in(\frac{r_i}{2},r_i)\\
=0 & d(x,0)\ge r_i
\end{cases}
$$
Define $f_i:M\to R$ by
$$
f_i(x)=
\begin{cases}
(H_i\circ\varphi_i)(x) & x\in B'_i\\
0 & x\notin B'_i
\end{cases}
$$
Then
$$
f_i(x)
\begin{cases}
\gt0 & x\in B_i\\
=0 & x\notin B_i
\end{cases}
$$
implying $\text{supp}f_i=\overline{B_i}$, and $f_i$ is clearly smooth.
Define $f:M\to R$ by $$f(p)=\sum_{i\in k}f_i(p)$$
Let $p\in M$, since $\{B_i:i\in k\}$ is locally finite and $B_i$ is injective, there exists an open subset $S$ of $M$ containing $p$
such that for some finite $j\subseteq k$, for all $i\in k$, $B_i\cap S\neq\emptyset$ if and only if $i\in j$.
Thus for all $x\in S$, $f(x)=\sum_{i\in j}f_i(x)$.
Note that for each $i\in j$, $f_i$ is smooth on the open submanifold $S$.
Thus $f|_S$ is smooth.
We have shown that $f$ is smooth.
Let $p\in M$, then for some $i\in k$, $p\in B_i$, thus $f(p)\ge f_i(p)\gt 0$.
Hence $f(M)\subseteq R^+$, implying $\frac{1}{f}$ is smooth, thus $$g_i=\frac{f_i}{f}$$ is smooth for each $i$.
Clearly, let $p\in M$, then $\sum_{i\in k}g_i(p)=1$ and for each $i\in k$, $0\le g_i(p)\le 1$.
Let $i\in k$, then for some $a_i\in A$, $B'_i\subseteq U_{a_i}$.
Now for each $a\in A$, let $j_a$ denote $\{i\in k|a_i=a\}$ and define $\psi_a:M\to R$ by
$$\psi_a(p)=\sum_{i\in j_a}g_i(p)$$
Then $\psi_a$ is smooth and $0\le\psi_a(p)\le 1$ for all $p\in M$ and $a\in A$.
Let $a\in A$, then $\{B_i:i\in j_a\}$ is locally finite, thus
$$\text{supp}\psi_a=\overline{\bigcup_{i\in j_a}B_i}=\bigcup_{i\in j_a}\overline{B_i}\subseteq U_a$$
Let $p\in M$, we showed that there exists an open subset $S$ of $M$ containing $p$
such that for some finite $j\subseteq k$, for all $i\in k$, $B_i\cap S\neq\emptyset$ if and only if $i\in j$.
If $i\in k\setminus j$, then $B_i\cap S=\emptyset$, thus $\overline{B_i}\cap S=\emptyset$.
Hence for all $i\in k$, if $\overline{B_i}\cap S\neq\emptyset$, then $i\in j$.
Let $j'$ denote $\{i\in j|\overline{B_i}\cap S\neq\emptyset\}$,
then $a_i$ defines a surjection from $j'$ to $\{a\in A|\text{supp}\psi_a\cap S\neq\emptyset\}$, which is thus finite.
We have shown that $\{\text{supp}\psi_a:a\in A\}$ is locally finite.
Let $p\in M$, then $\{i\in k|g_i(p)\neq0\}$, denoted $l$, is finite.
Let $A'$ denote $\{a_i:i\in l\}$, then $A'$ is finite, and
$$\sum_{a\in A}\psi_a(p)=\sum_{a\in A}\sum_{i\in j_a}g_i(p)=\sum_{a\in A'}\sum_{i\in j_a\cap l}g_i(p)=\sum_{i\in l}g_i(p)=1$$
Therefore, $(\psi_a)_{a\in A}$ is a smooth partition of unity with respect to $(U_a)_{a\in A}$.
$\blacksquare$
Bump function
Let $M$ be a smooth manifold with boundary and let $U,V\subseteq M$ such that $V$ is open and $\overline U\subseteq V$,
then a bump function for $U$ supported in $V$ is a continuous function $f:M\to R$ such that
$0\le f(p)\le 1$ for all $p\in M$, $f(p)=1$ for all $p\in U$, and $f$ is supported in $V$.
Proposition.
Let $M$ be a smooth manifold with boundary and let $U,V\subseteq M$ such that $V$ is open and $\overline U\subseteq V$,
then there exists a smooth bump function for $U$ supported in $V$.
(show proof)
Proof.
Define $S_0=\text{Ext}U$ and $S_1=V$, then $(S_a)_{a\in\{0,1\}}$ is an open cover of $M$.
Thus there exists a smooth partition of unity $(\psi_a)_{a\in\{0,1\}}$ with respect to $(S_a)_{a\in\{0,1\}}$.
Note that $\psi_1$ is a smooth real-valued function defined on $M$,
$0\le\psi_1(p)\le 1$ for all $p\in M$, and $\psi_1$ is supported in $V$.
Since $\psi_0$ is supported in $\text{Ext}U$,
for all $p\in U$, $\psi_0(p)=0$, thus $\psi_1(p)=\sum_{a\in\{0,1\}}\psi_a(p)=1$.
We have shown that $\psi_1$ is a smooth bump function for $U$ supported in $V$.
$\blacksquare$
Generalized smoothness on manifolds
Let $M,N$ be smooth manifolds with boundary, $U\subseteq M$, and $f:U\to N$.
If for every $p\in U$, there exists an open submanifold $U_p$ of $M$ containing $p$
and a smooth function $f_p:U_p\to N$ that agrees with $f$ on $U\cap U_p$,
then $f$ is said to be smooth.
Generalized smoothness trivially implies continuity.
Note that this definition of smoothness is equivalent to the original definition of smoothness if $U$ is open and thus an open submanifold of $M$.
Also note that this definition of smoothness is equivalent to generalized smoothness on Euclidean spaces
if $M,N$ are standard Euclidean spaces.
Proposition.
Let $M,N$ be smooth manifolds with boundary, let $U\subseteq M$, and let $f:U\to N$ be smooth.
Then for all $V\subseteq U$, $f|_V$ is smooth.
(show proof)
Proof.
Let $p\in V$, then $p\in U$, thus there exists an open submanifold $U_p$ of $M$ containing $p$
and a smooth function $f_p:U_p\to N$ that agrees with $f$ on $U\cap U_p$.
Let $x\in V\cap U_p$, then $f_p(x)=f(x)=f|_V(x)$.
Thus $f_p$ agrees with $f|_V$ on $V\cap U_p$.
Hence $f|_V$ is smooth.
$\blacksquare$
Proposition.
Let $M,N,P$ be smooth manifolds with boundary, let $U\subseteq M$ and $V\subseteq N$,
let $f:U\to N$ and $g:V\to P$ be smooth such that $f(U)\subseteq V$,
then $g\circ f$ is smooth.
(show proof)
Proof.
Let $p\in U$, then there exists an open submanifold $V_p$ of $N$ containing $f(p)$
and a smooth function $g_p:V_p\to P$ that agrees with $g$ on $V\cap V_p$,
and there exists an open submanifold $U_p$ of $M$ containing $p$
and a smooth function $f_p:U_p\to N$ that agrees with $f$ on $U\cap U_p$.
Note that $f_p^{-1}(V_p)$ is open, thus $f_p|_{f_p^{-1}(V_p)}$ is smooth, and $f_p(p)=f(p)\in V_p$.
Now we have an open submanifold $f_p^{-1}(V_p)$ of $M$ containing $p$
and a smooth function $g_p\circ f_p|_{f_p^{-1}(V_p)}$ such that for all $x\in U\cap f_p^{-1}(V_p)$,
$(g_p\circ f_p|_{f_p^{-1}(V_p)})(x)=g_p(f_p(x))=g(f(x))=(g\circ f)(x)$.
We have shown that $g\circ f$ is smooth.
$\blacksquare$
Generalized diffeomorphism on manifolds
Let $M,N$ be smooth manifolds with boundary, $U\subseteq M$, and $V\subseteq N$.
A bijective function $F:U\to V$ such that $F:U\to N$ and $F^{-1}:V\to M$ are smooth is called a diffeomorphism.
Trivially, a diffeomorphism $F:U\to V$ is also a homeomorphism.
If $U,V$ are open, this definition of diffeomorphism is equivalent to that with $U,V$ taken as open submanifolds.
If $M,N$ are Euclidean spaces, then this definition of diffeomorphism is equivalent to that of Euclidean spaces.
Proposition.
Let $M,N,P$ be smooth manifolds with boundary. Let $U\subseteq M$, $V\subseteq N$, and $W\subseteq P$.
let $f:U\to V$ and $g:V\to W$ be diffeomorphisms, then $g\circ f:U\to W$ is a diffeomorphism.
(show proof)
Proof.
Trivial.
$\blacksquare$
Proposition.
Let $M$ be a smooth manifold with boundary, let $U$ be a closed subset of $M$,
let $V$ be an open subset of $M$ such that $U\subseteq V$, and let $f:U\to R^m$ be smooth.
Then there exists a smooth function $F:M\to R^m$ that agrees with $f$ on $U$ and is supported in $V$.
(show proof)
Proof.
Let $p\in U$, then there exists an open submanifold $U_p$ of $M$ such that $p\in U_p\subseteq V$
and a smooth function $f_p:U_p\to R^m$ that agrees with $f$ on $U\cap U_p$.
Define $U'=U\cup\{U\}$ and $U_U=M\setminus U$,
then $(U_p)_{p\in U'}$ is an open cover of $M$.
Let $(\psi_p)_{p\in U'}$ be the corresponding smooth partition of unity.
For each $p\in U$, $\text{supp}\psi_p\subseteq U_p$.
Extend $f_p$ so that $f_p(x)=0$ for all $x\in M\setminus U_p$, then $\psi_pf_p:M\to R^m$ is smooth,
and $\text{supp}\psi_pf_p\subseteq\text{supp}\psi_p$.
Thus $\{\text{supp}\psi_pf_p:p\in U\}$ is locally finite.
Hence we can define $F:M\to R^m$ by
$$F(x)=\sum_{p\in U}(\psi_pf_p)(x)$$
Let $p\in M$, then there exists some open subset $B$ of $M$ containing $p$
and some finite $S\subseteq U$ such that $F|_B=\sum_{p\in S}(\psi_pf_p)|_B$, which is smooth.
Thus $F$ is smooth.
Note that $\text{supp}\psi_U\subseteq M\setminus U$,
thus for all $x\in U$,
$$F(x)
=\sum_{\{p\in U|(\psi_pf_p)(x)\neq0\}}\psi_p(x)f_p(x)
=\p{\sum_{\{p\in U|(\psi_pf_p)(x)\neq0\}}\psi_p(x)}f(x)
=\p{\sum_{\{p\in U|\psi_p(x)\neq0\}}\psi_p(x)}f(x)
=\p{\sum_{\{p\in U'|\psi_p(x)\neq0\}}\psi_p(x)}f(x)
=f(x)
$$
Let $x\in M$ such that $F(x)\neq0$.
Then there exists $p\in U$ such that $(\psi_pf_p)(x)\neq0$, implying $\psi_p(x)\neq0$, and thus $x\in\text{supp}\psi_p\subseteq\bigcup_{p\in U}\text{supp}\psi_p$.
Note that $\{\text{supp}\psi_p:p\in U\}$ is locally finite.
Hence $$\text{supp}F\subseteq\overline{\bigcup_{p\in U}\text{supp}\psi_p}=\bigcup_{p\in U}\overline{\text{supp}\psi_p}=\bigcup_{p\in U}\text{supp}\psi_p\subseteq V$$
We have shown that $F$ is the desired function.
$\blacksquare$
Lemma.
Let $M$ be a smooth manifold with boundary, let $U$ be an open subset of $M$, and let $p\in U$,
then there exists a closed neighborhood of $p$ contained in $U$.
(show proof)
Proof.
Suppose $p$ is an interior (boundary) point, then there exists an interior (a boundary) chart $(\varphi,S)$ such that $p\in S\subseteq U$.
Since $\varphi(S)$ is open, we can find an open (half) ball $B$ centered at $\varphi(p)$ such that $\overline B\subseteq\varphi(S)$.
Note that $\overline B$ is a compact subset of $\varphi(S)$.
Then $\varphi^{-1}(\overline B)$ is a compact subset of $S$ and thus $M$, implying it is closed.
Note that $p\in\varphi^{-1}(B)\subseteq\varphi^{-1}(\overline B)\subseteq S\subseteq U$.
Hence $\varphi^{-1}(\overline B)$ is a closed neighborhood of $p$ contained in $U$.
$\blacksquare$
Local extension of smooth function
Let $M$ be a smooth manifold with boundary, let $p\in M$, and let $U$ be a neighborhood of $p$.
Then there exists a neighborhood $U^*\subseteq U$ of $p$, such that for all smooth $f:U\to R^m$,
there exists a smooth $\overline f:M\to R^m$ that agrees with $f$ on $U^*$
(show proof).
Proof.
Since $U$ is a neighborhood of $p$, there exists an open subset $B$ of $M$ such that $p\in B\subseteq U$.
By the lemma above, there exists a closed neighborhood $U^*$ of $p$ contained in $B$, and thus contained in $U$.
Let $f:U\to R^m$ be smooth, then $f|_{U^*}$ is smooth.
Thus there exists a smooth function $\overline f:M\to R^m$ that agrees with $f|_{U^*}$ on $U^*$, and hence agrees with $f$ on $U^*$.
$\blacksquare$
We call $\overline f$ a local extension of $f$.