You signed in with another tab or window. Reload to refresh your session.You signed out in another tab or window. Reload to refresh your session.You switched accounts on another tab or window. Reload to refresh your session.Dismiss alert
\newthought{Since to speak of continuity we need topological spaces}, it may be a good idea to remind you what they are and set some notation.
56
56
I will be very brief: if you need a more extensive reminder, you can refer to Appendix A of either~\cite{book:tu} or~\cite{book:lee}.
57
57
58
-
\begin{definition}
58
+
\begin{definition}\idxdef{Topological space}
59
59
Let $X$ be some set and $\cT$ a set of subsets of $X$.
60
60
A pair $(X, \cT)$ is a \emph{topological space}\footnote{In such case the elements $O\in\cT$ of $\cT$ are all subsets of $X$ called \emph{open} subsets and $\cT$ is a \emph{topology} on $X$.} if
A topological space $(X, \cT)$ is \emph{Hausdorff} if every two distinct points admit disjoint open neighbourhoods. That is, for every pair $x\neq y$ of points in $X$, there exist open subsets $U_x, U_y\in\cT$ such that $x\in U_x$, $y\in U_y$ and $U_x \cap U_y = \emptyset$.
Hausdorff spaces are still rather general: in particular, any metric space with the metric topology\footnote{Recall that in a metric space $X$ the \emph{metric topology} is defined in the following way: a set $U\subset X$ is called open if for any $x\in U$ there exists $\epsilon>0$ such that $U$ fully contains the ball of radius $\epsilon$ around $x$.} is Hausdorff.
98
98
99
-
\begin{definition}
99
+
\begin{definition}\idxdef{Second countable}
100
100
A topological space $(X, \cT)$ is \emph{second countable} if there exists a countable set $\cB\subset\cT$ such that any open set can be written as a union of sets in $\cB$.
101
101
In such case, $\cB$ is called a (countable) basis for the topology $\cT$.
A topological space\sidenote[][-1.5em]{From now on, if we say that $X$ is a topological space we are implying that there is a topology $\cT$ defined on $X$.} $M$ is a \emph{topological manifold} of dimension $n$, or topological $n$-manifold, if it has the following properties:
111
111
\marginnote[-0.5em]{Note that the finite dimensionality is a somewhat artificial restriction: manifolds can be infinitely dimensional~\cite{book:lang:infinite}. For example, the space of continuous functions between manifolds is a so-called infinite-dimensional Banach manifold.\vspace{1em}}
Reusing the notation of the definition above, we call \emph{(coordinate) chart} the pair $(U, \varphi)$ of a \emph{coordinate neighbourhood}\footnote{Or \emph{coordinate open set}} $U$ and an associated \emph{coordinate map}\footnote{Or \emph{coordinate system}.} $\varphi: U\to V$ onto an open subset $V=\varphi(U)\subseteq\R^n$ of $\R^n$.
121
121
Furthermore, we say that a chart is \emph{centred at $p\in U$} if $\varphi(p) = 0$.
Before entering into the details of new definitions, let's recall what will be the most important tools throughout the rest of the course.
186
186
187
-
\begin{definition}
187
+
\begin{definition}\idxdef{Partition of unity}
188
188
A map $f: U \to V$ between open sets $U\subset\R^n$ and $V\subset\R^m$ is in $C^r(U,V)$ or \emph{of class $C^r$}, if it is continuously differentiable $r$-times.
189
189
It is called a $C^r$-\emph{diffeomorphism}\footnote{With this definition a homeomorphism is a $C^0$-diffeomorphism} if it is bijective and of class $C^r$ with inverse of class $C^r$.
190
190
We say that $f$ is \emph{smooth}, or of class $C^\infty$, if it is of class $C^r$ for every $r \geq1$.
Let $U\subseteq\R^n$ and $V\subseteq\R^k$ be open sets and $f: U \to\R^k$, $g: V\to\R^m$ two continuously differentiable functions such that $f(U)\subseteq V$.
A \emph{differentiable structure}, or more precisely a \emph{smooth structure}, on a topological manifold is an equivalence class of smooth atlases.
262
262
\end{definition}
263
263
264
264
\begin{remark}
265
265
The union of all atlases in a differentiable structure is the \emph{unique} \emph{maximal} atlas in the equivalence class.\footnote{There is a one-to-one correspondence between differentiable structures and maximal differentiable atlases \cite[Proposition 1.17]{book:lee}: for convenience and to lighten the notation, from now on, we will always regard a differentiable structure as a differentiable maximal atlas without further comments.}
A \emph{smooth manifold} of dimension $n$ is a pair $(M, \cA)$ of a topological $n$-manifold $M$ and a smooth structure $\cA$ on $M$.
270
270
\marginnote[1em]{There are no preferred coordinate charts on a manifold: all coordinate systems compatible with the differentiable structure are on equal footing.}
271
271
\end{definition}
@@ -883,7 +883,7 @@ \section{Partitions of unity}\label{sec:partition_of_unity}
883
883
884
884
We are not there yet. To extend this result to our needs will need a new tool, which will be useful throughout the course and in many courses to come.
885
885
886
-
\begin{definition}
886
+
\begin{definition}\idxdef{Partition of unity}
887
887
Let $M$ be a smooth manifold. A \emph{partition of unity} is a collection $\{\rho_\alpha\mid\alpha\in A\}$ of functions $\rho_\alpha:M\to\R$ such that
888
888
\begin{enumerate}[(i)]
889
889
\item$0\leq\rho_\alpha\leq1$ for all $p\in M$ and $\alpha\in A$;
@@ -899,7 +899,7 @@ \section{Partitions of unity}\label{sec:partition_of_unity}
899
899
900
900
Throughout the course we will be mostly interested in partitions of unity $\{\rho_\alpha\mid\alpha\in A\}$ which are \emph{subordinate} to an open cover $\{U_\alpha\mid\alpha\in A\}$, that is, such that $\supp_\alpha(\rho_\alpha) \subset U_\alpha$ for each $\alpha\in A$.
901
901
902
-
\begin{theorem}\label{thm:partitionof1}
902
+
\begin{theorem}\label{thm:partitionof1}\idxthm{Partition of unity (existence)}
903
903
\marginnote[1.5em]{We are going to omit the proof of this theorem, for its details you can refer to~\cite[Proposition 13.6]{book:tu} or~\cite[Theorem 2.23]{book:lee}.}
904
904
Let $M$ be a smooth manifold. For any open cover $\{U_\alpha\mid\alpha\in A\}$ of $M$, there exists a partition of unity $\{\rho_\alpha\mid\alpha\in A\}$ subordinate to $\{U_\alpha\mid\alpha\in A\}$.
905
905
\end{theorem}
@@ -972,7 +972,7 @@ \section{Manifolds with boundary}\label{sec:mbnd}
972
972
Explicitly state the definitions above in the case of manifolds with boundary.
973
973
\end{exercise}
974
974
975
-
\begin{definition}\label{def:diffmanifoldwb}
975
+
\begin{definition}\label{def:diffmanifoldwb}\idxdef{Manifold with boundary}
976
976
\marginnote{Remember that the differentiable structure is an equivalence class of smooth atlases.}
977
977
A \emph{smooth manifold with boundary} of dimension $n$ is a pair $(M, \cA)$ of a topological $n$-manifold with boundary $M$ and a smooth differentiable structure $\cA = \{(U_\alpha, \varphi_\alpha) \mid\alpha\in A\}$ on $M$.
Copy file name to clipboardExpand all lines: 10-integration.tex
+12-12Lines changed: 12 additions & 12 deletions
Original file line number
Diff line number
Diff line change
@@ -11,7 +11,7 @@
11
11
\section{Orientation on vector spaces}
12
12
Let's proceed step by step by first revisiting some results from multivariable analysis. If you need a reference to review this material, you can refer to \cite[Chapter 6.2]{book:abrahammarsdenratiu} or \cite[Chapters 21.1-21.2]{book:tu}.
13
13
14
-
\begin{definition}
14
+
\begin{definition}\idxdef{Oriented manifold}
15
15
Let $V$ be a one-dimensional vector space. Then $V\setminus\{0\}$ has two components.
16
16
An \emph{orientation} of $V$ is a choice of one of these components, which one then labels as ``positive'' and ``negative''.
17
17
A \emph{positive basis} of $V$ then is a choice of any non-zero vector belonging to the positive component, while a \emph{negative basis} of $V$ is a choice of any non-zero vector belonging to the negative component.
@@ -25,7 +25,7 @@ \section{Orientation on vector spaces}
25
25
If $V$ be a $n$-dimensional vector space, we know by Proposition~\ref{prop:dimLkV}, that $\Lambda^n(V)$ is a one-dimensional vector space.
26
26
Moreover, if $\{e_1,\ldots,e_n\}$ is a basis for $V$, then $e^1\wedge\cdots\wedge e^n$ is a basis for $\Lambda^n(V)$.
27
27
28
-
\begin{definition}
28
+
\begin{definition}\idxdef{Orientation of a vector space}\idxdef{Positive orientation}
29
29
Let $V$ be a $n$-dimensional vector space.
30
30
An \emph{orientation} on $V$ is a choice of orientation\footnote{That is, we are talking about a representative from an equivalence class.} on the one-dimensional vector space $\Lambda^n(V)$.
31
31
Therefore there are exactly two orientations: we say that a basis $\{e_1,\ldots,e_n\}$ of $V$ is \emph{positive} (or positively oriented) if $e^1\wedge\cdots\wedge e^n$ is a positive basis of $\Lambda^n(V)$ and \emph{negative} (or negatively oriented) otherwise.
@@ -63,7 +63,7 @@ \section{Orientation on vector spaces}
63
63
Let $V$ be a $n$-dimensional vector space, prove that two nonzero $n$-forms on $V$ determine the same orientation if and only if each is a positive multiple of the other.
64
64
\end{exercise}
65
65
66
-
This allows to define an equivalence relation between orientations in terms of the nonzero elements in $\Lambda^n(V)$. We call these nonzero elements \emph{volume elements}.
66
+
This allows to define an equivalence relation between orientations in terms of the nonzero elements in $\Lambda^n(V)$. We call these nonzero elements \emph{volume elements}\idxdef{Volume element} on $V$.
67
67
68
68
Two volume elements $\omega_1, \omega_2$ are equivalent if there exists $c > 0$ such that $\omega_1 = c\,\omega_2$.
69
69
Then, the previous exercise implies that the classes of equivalence $[\omega]$ of volume elements on $V$ uniquely determine orientations on $V$.
@@ -79,9 +79,9 @@ \section{Orientation on manifolds}
79
79
Moreover, we can extend our point of view to the various tensor bundles over smooth manifold that we studied so far.
80
80
Differential $n$-forms, then, seem a reasonable concept to define a notion of orientation for a manifold, at least if we think about their pointwise meaning of assigning an orientation to each fiber of $TM$.
81
81
82
-
\begin{definition}
83
-
A \emph{volume form}\footnote{Sometimes also called \emph{orientation form}.} on a $n$-dimensional smooth manifold $M$ is a $n$-form $\omega\in\Omega^n(M)$ such that $\omega(p) \neq0$ for all $p\in M$.
84
-
We say that $M$ is \emph{orientable} if there exists a volume form on $M$.
A \emph{volume form}\footnote{Sometimes also called \emph{orientation form}.}\idxdef{Volume form} on a $n$-dimensional smooth manifold $M$ is a $n$-form $\omega\in\Omega^n(M)$ such that $\omega(p) \neq0$ for all $p\in M$.
84
+
We say that $M$ is \emph{orientable}\idxdef{Orientable manifold} if there exists a volume form on $M$.
85
85
\end{definition}
86
86
87
87
\begin{example}
@@ -162,7 +162,7 @@ \section{Orientation on manifolds}
162
162
\end{proof}
163
163
164
164
165
-
\begin{definition}
165
+
\begin{definition}\idxdef{Oriented manifold}\idxdef{Orientation of a manifold}
166
166
A manifold $M$ with an oriented atlas is called \emph{oriented manifold}.
167
167
If an orientation exists, we call \emph{orientation} the equivalence class of atlases with the same orientation, i.e., the family of atlases whose union is still an oriented atlas.
168
168
Otherwise we say that the manifold is \emph{non-orientable}.
@@ -178,7 +178,7 @@ \section{Orientation on manifolds}
178
178
179
179
In fact, we can go one step further and define what does it mean for a diffeomorphism to preserve or reverse orientation.
Let $(M, \omega)$ and $(N, \eta)$ be pairs of an oriented smooth manifold and its volume form.
183
183
A diffeomorphism $F: M \to N$ is called \emph{orientation preserving} if $F^* \eta$ is a volume form on $M$ with the same orientation as $\omega$.
184
184
We call it \emph{orientation reversing} if $F^*\eta$ has the opposite orientation as $\omega$.
@@ -374,7 +374,7 @@ \section{Integrals on manifolds}
374
374
375
375
% TODO: properly explain how the change of variable in Euclidean space requires the absolute value of the Jacobian determinant while the transformation of a k-form only involves the determinant itself, and thus in our definition we need to take into account the orientation and compensate with a minus sign if necessary.
376
376
377
-
\begin{definition}\label{def:intnform:chart}
377
+
\begin{definition}\label{def:intnform:chart}\idxdef{Integral of n-forms supported on a single chart}
378
378
Let $M$ be a smooth $n$-manifold and $(U,\varphi)$ be a chart from an oriented atlas of $M$ with coordinates $(x^i)$.
379
379
If $\omega\in\Omega^n(M)$ is a $n$-form, $n > 0$, with compact support in $U$, we define the integral of $\omega$ as\sidenote[][-1em]{Recall that for a diffeomorphism $\phi$, $\phi_* = (\phi^{-1})^*$.}
380
380
\begin{equation}
@@ -418,7 +418,7 @@ \section{Integrals on manifolds}
418
418
419
419
To be able to integrate charts which are not supported in the domain of a single chart, we now need the help of a partition of unity.
420
420
421
-
\begin{definition}
421
+
\begin{definition}\idxdef{Integral of n-forms on manifolds}
422
422
Let $M$ be a smooth oriented manifold and $\cA = \{(U_i,\varphi_i)\}$ a positively oriented atlas.
423
423
If $\omega\in\Omega^n(M)$ has compact support, then the \emph{integral of $\omega$} is defined as
424
424
\begin{equation}\label{eq:intnform}
@@ -449,7 +449,7 @@ \section{Integrals on manifolds}
449
449
\end{proof}
450
450
451
451
Which immediately implies the following nice result.
452
-
\begin{theorem}[Global change of variables]\label{thm:gcv}
452
+
\begin{theorem}[Global change of variables]\label{thm:gcv}\idxthm{Global change of variables}
453
453
Suppose $M$ and $N$ are oriented $n$-manifolds and $F:M\to N$ is an orientation preserving diffeomorphism.
454
454
If $\omega\in\Omega^n(N)$ has compact support, then $F^*\omega$ has compact support and the following holds
455
455
\begin{equation}
@@ -608,7 +608,7 @@ \section{Stokes' Theorem}
608
608
609
609
We are going to state the theorem, discuss some of its consequences and then give its proof.
0 commit comments