Skip to content

Commit

Permalink
Further fixes from Martijn and new version
Browse files Browse the repository at this point in the history
Signed-off-by: Marcello Seri <[email protected]>
  • Loading branch information
mseri committed Nov 6, 2023
1 parent 626dbea commit 1a6ab8b
Show file tree
Hide file tree
Showing 6 changed files with 52 additions and 25 deletions.
2 changes: 1 addition & 1 deletion 3b-liegroups.tex
Original file line number Diff line number Diff line change
Expand Up @@ -39,7 +39,7 @@ \section{Lie groups}
\end{example}

\begin{exercise}
\item Prove that any Lie group is orientable.
Prove that any Lie group is orientable.
\end{exercise}

\begin{definition}
Expand Down
16 changes: 10 additions & 6 deletions 6-differentiaforms.tex
Original file line number Diff line number Diff line change
Expand Up @@ -413,6 +413,9 @@ \section{Exterior derivative}
& = \sum_{i<j} \left(\frac{\partial \omega_j}{\partial x^i} - \frac{\partial \omega_i}{\partial x^j} \right) dx^i\wedge dx^j,
\end{align}
consistently with our previous definitions.

It is worth observing here that $d(df) = 0$ since $\frac{\partial^2 f}{\partial x^i \partial x^j} = \frac{\partial^2 f}{\partial x^j \partial x^i}$.
We will revisit this fact in Lemma~\ref{lem:ext_deriv}.
\end{example}

\begin{definition}
Expand Down Expand Up @@ -479,7 +482,7 @@ \section{Exterior derivative}
\end{equation}
\end{exercise}

\begin{lemma}
\begin{lemma}\label{lem:ext_deriv}
The exterior derivative satisfies the following properties.
For all $\omega,\omega_1,\omega_2\in\Omega^k(M)$, $\nu\in\Omega^h(M)$ and $f\in C^\infty(M)$,
\begin{enumerate}[(i)]
Expand Down Expand Up @@ -548,14 +551,14 @@ \section{Exterior derivative}
\begin{equation}
d\eta = \left(\frac{\partial u}{\partial z}-\frac{\partial v}{\partial y} + \frac{\partial w}{\partial x}\right) dx\wedge dy \wedge dz.
\end{equation}
If you compare the results we obtained above with the gradient ($\nabla$), divergence ($\nabla\cdot$) and curl ($\nabla\times$) from multivariable analysis, you would likely notice that the components of the $2$-form $\d\omega$ are exactly the components of the curl of the vector field with components $(P, Q, R)$.
If you compare the results we obtained above with the gradient ($\nabla$), divergence ($\nabla\cdot$) and curl ($\nabla\times$) from multivariable analysis, you would likely notice that the components of the $2$-form $d\omega$ are exactly the components of the curl of the vector field with components $(P, Q, R)$.
Similarly, the formula for the divergence will look very close to the formula for $d\eta$.
What is going on?

The standard euclidean metric on $\R^n$ is the metric associated to the metric tensor\footnote{Cf. Definition~\ref{def:metric}.} $g_{ij} = \delta_{ij}$.
The euclidean metric on $\R^n$ is the metric associated to the metric tensor\footnote{Cf. Definition~\ref{def:metric}.} $g_{ij} = \delta_{ij}$.
We can use the musical isomorphisms\footnote{Cf. Example~\ref{ex:musicaliso}.} to identify vector fields and $1$-forms, obtaining for the components with respect to cartesian coordinates that $v_i = v^i$.

Moreover, the interior multiplication yields another map $\beta: \fX(\R^3)\to\Omega^2(\R^3)$ defined by $\beta(X) = \iota_X (dx\wedge dy\wedge dz)$, which is linear over $C^\infty(\R^3)$ (why?) and, thus, corresponds to a smooth bundle homomorphism from $T\R^3$ to $\Lambda^2\R^3$ (why?).
Moreover, the interior multiplication yields another map $\beta: \fX(\R^3)\to\Omega^2(\R^3)$ defined by $\beta(X) = \iota_X (dx\wedge dy\wedge dz)$, which is linear over $C^\infty(\R^3)$ (why?) and, thus, corresponds to a smooth bundle isomorphism from $T\R^3$ to $\Lambda^2\R^3$ (why?).

In a similar fashion, we can also define a smooth bundle isomorphism $\bigstar: C^\infty(\R^3) \to \Omega^3(\R^3)$ via
\begin{equation}
Expand All @@ -579,7 +582,7 @@ \section{Exterior derivative}

The interest and need to generalize the operations of vector calculus in $\R^3$ to higher dimensional spaces have been one of the drives to develop the theory of differential forms.

This is a more general fact related to 3 dimensional manifolds with a Riemannian metric $M$. Then the maps defined above are still defined in the same exact way\footnote{There is a more general operator $\star : \Omega^k(M) \to \Omega^{n-k}(M)$ appearing here. This is called \emph{Hodge star} and is defined as the operator such that $\tau \wedge \star \sigma = g^\dagger(\tau,\sigma) \omega$, where $\omega := \star 1$ is the Riemannian volume and $g^\dagger$ is the inner product induced by $g$ on covector fields via the musical isomorphism between tangent and cotangent bundle. See \cite[Exercise 16-18]{book:lee} or \cite[Chapters 6.2--6.5]{book:abrahammarsdenratiu}.} up to using the correct isomorphism induced by $g$ and one obtains the following commutative diagram.
This is a more general fact related to 3 dimensional manifolds with a Riemannian metric $M$. Then the maps defined above are still defined in the same exact way\footnote{There is a more general operator $\star : \Omega^k(M) \to \Omega^{n-k}(M)$ appearing here. This is called \emph{Hodge star} and is defined as the operator such that $\tau \wedge \star \sigma = g^\dagger(\tau,\sigma) \omega$, where $\omega := \star 1$ is the Riemannian volume and $g^\dagger$ is the inner product induced by $g$ on covector fields via the musical isomorphism between tangent and cotangent bundle (cf. Example~\ref{ex:musicaliso}). See \cite[Exercise 16-18]{book:lee} or \cite[Chapters 6.2--6.5]{book:abrahammarsdenratiu}.} up to using the correct isomorphism induced by $g$ and one obtains the following commutative diagram.
\begin{equation}
% https://q.uiver.app/?q=WzAsNSxbMSwxLCJcXE9tZWdhXjEoTSkiXSxbMCwwLCJcXG1hdGhjYWx7WH0oTSkiXSxbMSwyLCJcXE9tZWdhXjAoTSkiXSxbMiwyLCJcXE9tZWdhXjMoTSkiXSxbMiwxLCJcXE9tZWdhXjIoTSkiXSxbMSwwLCJcXGZsYXQiLDAseyJvZmZzZXQiOi0xfV0sWzAsMSwiXFxzaGFycCIsMCx7Im9mZnNldCI6LTF9XSxbMCw0LCJkIiwwLHsib2Zmc2V0IjotMX1dLFszLDIsIlxcc3RhciIsMCx7InN0eWxlIjp7InRhaWwiOnsibmFtZSI6ImFycm93aGVhZCJ9fX1dLFs0LDMsImQiXSxbMiwwLCJkIiwwLHsib2Zmc2V0IjoxfV0sWzAsNCwiXFxzdGFyIiwyLHsib2Zmc2V0IjoxLCJzdHlsZSI6eyJ0YWlsIjp7Im5hbWUiOiJhcnJvd2hlYWQifX19XV0=
\begin{tikzcd}
Expand All @@ -602,7 +605,7 @@ \section{Exterior derivative}
\item $\mathrm{curl} := \sharp \star d\; \flat : \fX(M) \to \fX(M)$;
\end{itemize}
where symbols' juxtaposition means their composition.
Since $\sharp$ and $\flat$ are defined in terms of $g$, it becaomes clear that all those operators are tightly related to the metric.
Since $\sharp$ and $\flat$ are defined in terms of $g$, it becomes clear that all those operators are tightly related to the metric.
There could be a lot more to say, but that will be more suite for a course in Riemannian geometry.
A comprehensive free resource on the subject is \cite{book:derivations}.
\end{example}
Expand Down Expand Up @@ -646,6 +649,7 @@ \section{Exterior derivative}
\end{equation}
\end{enumerate}
\end{exercise}
If you want to know more about symplectic manifolds and the role of differential geometry in classical mechanics, have a look at one of \cite{book:abrahammarsdenratiu, book:arnold, book:knauf, lectures:seri:hm}.

\section{Lie derivative}

Expand Down
10 changes: 6 additions & 4 deletions 6b-cohomology.tex
Original file line number Diff line number Diff line change
Expand Up @@ -66,11 +66,11 @@ \section{Poincar\'e lemma}
\begin{theorem}\label{thm:deRham-invariance}
Let $M$ be a smooth manifold and $[0,1]\times M$ the product manifold with boundary $(\{0\}\times M) \cup (\{1\}\times M) \cup ((0,1)\times\partial M)$.
Let $i_t : M \hookrightarrow [0,1]\times M$ be the injection $i_t(p) := (t,p)$ and $\pi : [0,1]\times M \to M$ the projection onto $M$.
Then, there is a map
Then, there is a linear map
\begin{equation}
K : \Omega^\ell([0,1]\times M)\to \Omega^{\ell-1}(M)
\end{equation}
such that for every differential $\ell$-form $\omega\in\Omega^\ell([0,1]\times M)$ one has
such that for every differential $\ell$-form $\omega\in\Omega^\ell([0,1]\times M)$ one has\footnote{If you took an algebraic topology course, you may recognise such map $K$ as a cochain homotopy.}
\begin{equation}
K(d\omega) + d(K(\omega)) = i^*_1(\omega) - i^*_0(\omega)
\end{equation}
Expand Down Expand Up @@ -137,7 +137,7 @@ \section{Poincar\'e lemma}

To conclude the proof, take a closed $\ell$-form on $[0,1]\times M$, then
\begin{equation}
i_1^*([w]) - i_0^*([w]) = [K(d\omega) + dK(\omega)] = [dK(\omega)] =0,
i_1^*([\omega]) - i_0^*([\omega]) = [K(d\omega) + dK(\omega)] = [dK(\omega)] =0,
\end{equation}
completing the proof.
\end{proof}
Expand Down Expand Up @@ -214,7 +214,9 @@ \section{Poincar\'e lemma}
Let $F:M\to N$ and $G:N\to M$ be continuous maps such that $F\circ G$ and $G\circ F$ are homotopic to the identity maps.
By the Whitney Approximation Theorem~\ref{thm:WhitneyApproxCont} we can approximate $F$ and $G$ by smooth maps that we keep denoting with the same symbols.
By the previous theorem, then, $(F\circ G)^*$ and $(G\circ F)^*$ coincide with the maps induced by the identity.
Since $\id^*$ is clearly the identity, we see that $F^*$ is an inverse to $G^*$, which concludes the proof.
Since $\id^*$ is clearly the identity, we see that $F^*$ is an inverse to $G^*$.
Hence $F^*$ is an isomorphism between the corresponding de Rham cohomology groups,
concluding the proof.
\end{proof}

We are almost there.
Expand Down
35 changes: 24 additions & 11 deletions 7-integration.tex
Original file line number Diff line number Diff line change
@@ -1,4 +1,4 @@
We finally have all the main ingredients to generalize our line integral detour and discuss integration of $n$-forms over $n$-dimensional manifolds.
We finally have all the main ingredients to generalise our line integral detour and discuss integration of $n$-forms over $n$-dimensional manifolds.

\newthought{We know from calculus one}, or our line integral examples, that the direction in which we traverse the interval, or a curve, can actually make a difference.
As it turns out, the sign of the integral of a differential $n$-form is only fixed after choosing an orientation of the manifold.
Expand Down Expand Up @@ -48,7 +48,7 @@ \section{Orientation on vector spaces}
In fact, the orientation is completely characterized by the action of $n$-forms on the bases, as the following lemma shows.

\begin{lemma}\label{lemma:orient}
Let $V$ be a $n$-dimensional vector space and let $ \omega\in\Lambda^n(V)$ be nowhere vanishing.
Let $V$ be a $n$-dimensional vector space and let $\omega\in\Lambda^n(V)$ be nonzero.
Then, all bases $\{v_1, \ldots, v_n\}$ for which $\omega(v_1,\ldots, v_n) > 0$ give the same\footnote{Not necessarily the positive orientation!} orientation for $V$.
\end{lemma}
% \begin{proof}
Expand All @@ -63,10 +63,10 @@ \section{Orientation on vector spaces}
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.
\end{exercise}

This allows to define an eqivalence relation between orientations in terms of the nonzero elements in $\Lambda^n(V)$. We call these nonzero elements \emph{volume elements}.
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}.

Two volume elements $\omega_1, \omega_2$ are equivalent if there exists $c > 0$ such that $\omega_1 = c\, \omega_2$.
Then, the previous exercise implies that the classes of equivalence $[\omega]$ of volume elments on $V$ uniquely determine orientations on $V$.
Then, the previous exercise implies that the classes of equivalence $[\omega]$ of volume elements on $V$ uniquely determine orientations on $V$.

\begin{remark}
Of course, if $V$ is a vector space, then an orientation on $V$ canonically determines an orientation on the dual space $V^*$ by declaring that the basis dual to a positive basis is itself positive.
Expand All @@ -76,7 +76,7 @@ \section{Orientation on vector spaces}

\section{Orientation on manifolds}
Tangent and cotangent spaces are vector spaces, so what we discussed in the previous section directly apply (as usual).
Moreover, we can bundle up over our manifold into a vector bundle.
Moreover, we can extend our point of view to the various tensor bundles over smooth manifold that we studied so far.
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 $TM$.

\begin{definition}
Expand Down Expand Up @@ -113,11 +113,11 @@ \section{Orientation on manifolds}
Locally, for some chart with coordinates $(x^i)$ on $M$, we have $\omega_p = w(p)\, dx^{i_1}\wedge\cdots\wedge dx^{i_n}$ and $\eta_p = \eta(p)\, dx^{i_1}\wedge\cdots\wedge dx^{i_n}$, where both coefficients are in $C^\infty(M)$.
Since $\omega(p) \neq 0$ for all $p\in M$, $f(p) = \eta(p)/\omega(p)$ is a smooth function.

Conversely, if $\omega$ is a basis for $\Omega^n$, then it must be nonvanishing for all $p\in M$ since each fiber is one-dimensional.
Conversely, if $\omega$ is a basis\footnote{Here we mean a basis for $\Omega^n(M)$ as a $C^\infty(M)$-module, not as a vector space.} for $\Omega^n(M)$, then it must be nonvanishing for all $p\in M$ since each fiber is one-dimensional.
\medskip

\textbf{Part II: 1. $\Leftrightarrow$ 3.}
Assume $M$ is orientable witha volume form $\omega$.
Assume $M$ is orientable with a volume form $\omega$.
By eventually restricting the domains, let the atlas $\cA = \{(U_i, \varphi_i)\}$ be such that $\varphi_i(U_i) \subset \R^n$ is connected.
Denote $\omega_0 = dx^1 \wedge\cdots\wedge dx^n$ the standard volume form on $\R^n$, then by the previous part of the proof, $(\varphi_i)_*\omega = f_i\,\omega_0$ for some smooth function $f_i \neq 0$.
Without loss of generality\footnote{Is it clear why?} we can assume $f_i > 0$.
Expand Down Expand Up @@ -164,7 +164,7 @@ \section{Orientation on manifolds}

\begin{definition}
A manifold $M$ with an oriented atlas is called \emph{oriented manifold}.
If an orientation exists, we call \emph{orientation} the equivalence class of atlases with the same orientation.
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.
Otherwise we say that the manifold is \emph{non-orientable}.
\end{definition}

Expand All @@ -175,7 +175,19 @@ \section{Orientation on manifolds}
If $(U,\varphi)$ is a chart with local coordinates $(x^i)$ such that, in the coordinate representation, the volume form $\omega = \omega(x) dx^1\wedge\cdots\wedge dx^n$ with $\omega(x) > 0$, then we say that the chart $\varphi$ is \emph{positively oriented} with respect to $\omega$, otherwise we say that it is \emph{negatively oriented}.
Similarly, if $M$ is connected and the charts for the oriented manifold are positively (resp. negatively) oriented, we say that the manifold is positively (resp. negatively) oriented.
\end{definition}
%

In fact, we can go one step further and define what does it mean for a diffeomorphism to preserve or reverse orientation.

\begin{definition}
Let $(M, \omega)$ and $(N, \eta)$ be pairs of an oriented smooth manifold and its volume form.
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 $\omega$.
We call it \emph{orientation reversing} if $F^*\eta$ has the opposite orientation as $\omega$.
\end{definition}

\begin{exercise}
Show that a diffeomorphism $F: M \to N$ is \emph{orientation preserving} if and only if for all $p\in M$, $dF_p : T_p M \to T_{F(p)}$ is an orientation preserving as a linear map, that is, if $\det(dF_p)>0$.
\end{exercise}

% \begin{theorem}
% Let $M$ be a $n$-dimensional smooth manifold.
% A nowhere-vanishing $n$-form $\omega\in\Omega^n(M)$ uniquely determines an orientation.
Expand Down Expand Up @@ -227,6 +239,7 @@ \section{Orientation on manifolds}
and $\frac{2}{1-p^2}>0$.
If we perform the same computation on $U_2$, however, we obtain $(\varphi_2)_*(X) = -\frac{2}{1+p^2}\frac{\partial}{\partial x}\Big|_{\varphi_2(p)}$, with the negative coefficient $-\frac{2}{1+p^2} < 0$, corresponding to the opposite orientation on $U_2$.
Of course, in this case, not all is lost: by choosing $\widetilde\varphi_2(p) = \varphi_2(-p^1, p^2)$ we obtain $(\widetilde\varphi_2)_*(X) = \frac{2}{1+p^2} \frac{\partial}{\partial x}\Big|_{\widetilde\varphi_2(p)}$ with the positive coefficient $\frac{2}{1+p^2} > 0$, which shows that $X_p$ defines an orientation on the whole $\bS^1$.
Note again that here the orientation is given as a positive basis for $T_p \bS^1$ at all points, not as a volume form.
\end{example}

\begin{exercise}
Expand Down Expand Up @@ -569,7 +582,7 @@ \section{Integrals on manifolds}
\end{equation}
\end{proof}

This corollary has deep consequences in classical mechanics, which I am going to teach in the master and you are welcome to attend!
This corollary has deep consequences in classical mechanics, which we teach in the master and you are welcome to attend! If you are curious you can have a look at \cite[Chapter 3]{lectures:seri:hm} and, in particular, at Liouville theorem (Theorem 3.30).

\begin{remark}
The integral defined in this section can be extended rather immediately to measurable functions.
Expand All @@ -588,7 +601,7 @@ \section{Integrals on manifolds}
\section{Stokes' Theorem}

Stokes' theorem states that if $\omega$ is an $(n-1)$-form on an orientable $n$-manifold $M$, then the integral of $d\omega$ over $M$ equals the integral of $\omega$ over $\partial M$, generalising our observations for the line integral.
This is a beautiful and very important results, with deep consequences. The most immediate ones are the classical theorems of Gauss, Green and Stokes, which are just a special cases of this result.
This is a beautiful and very important result, with deep consequences. The most immediate ones are the classical theorems of Gauss, Green and Stokes, which are just a special cases of this result.

We are going to state the theorem, discuss some of its consequences and then give its proof.

Expand Down
10 changes: 9 additions & 1 deletion aom.bib
Original file line number Diff line number Diff line change
Expand Up @@ -208,7 +208,15 @@ @misc{lectures:merry
title = {Lecture notes on Differential Geometry},
year = 2019,
note = {Unpublished lecture notes},
url = {https://www.merry.io/courses/differential-geometry/}
url = {https://web.archive.org/web/20231106145652/https://www2.math.ethz.ch/will-merry/files/Merry%20-%20Differential%20Geometry%20%282019%29.pdf}
}

@misc{lectures:seri:hm,
author = {Seri, Marcello},
title = {Hamiltonian Mechanics},
year = {2023},
note = {Unpublished lecture notes, version 1.11},
url = {https://github.com/mseri/hammech20/releases/tag/1.11}
}

@misc{lectures:teufel,
Expand Down
4 changes: 2 additions & 2 deletions aom.tex
Original file line number Diff line number Diff line change
Expand Up @@ -214,7 +214,7 @@
\setlength{\parskip}{\baselineskip}
Copyright \copyright\ \the\year\ \thanklessauthor

\par Version 1.5 -- \today
\par Version 1.6 -- \today

\vfill
\small{\doclicenseThis}
Expand Down Expand Up @@ -270,7 +270,7 @@ \chapter*{Introduction}
Finally, a colleague mentioned~\cite{book:lang}. I don't have experience with this book but from a brief look it seems to follow a similar path as these lecture notes, so it might provide yet an alternative reference after all.

The idea for the cut that I want to give to this course was inspired by the online \href{https://www.video.uni-erlangen.de/course/id/242}{Lectures on the Geometric Anatomy of Theoretical Physics} by Frederic Schuller, by the lecture notes of Stefan Teufel's Classical Mechanics course~\cite{lectures:teufel} (in German), by the classical mechanics book by Arnold~\cite{book:arnold} and by the Analysis of Manifolds chapter in~\cite{book:thirring}.
In some sense I would like this course to provide the introduction to geometric analysis that I wish was there when I prepared my \href{https://www.mseri.me/lecture-notes-hamiltonian-mechanics/}{first edition} of the Hamiltonian mechanics course.
In some sense I would like this course to provide the introduction to geometric analysis that I wish was there when I prepared my \href{https://www.mseri.me/lecture-notes-hamiltonian-mechanics/}{first edition} of the Hamiltonian mechanics course (see also my lecture notes for that course \cite{lectures:seri:hm}).
In addition to the reference above, these lecture notes have found deep inspiration from~\cite{lectures:merry,lectures:hitchin} (all freely downloadable from the respective authors' websites), and from the book~\cite{book:abrahammarsdenratiu}.

I am extremely grateful to Martijn Kluitenberg for his careful reading of the notes and his invaluable suggestions, comments and corrections, and to Bram Brongers\footnote{You can also have a look at \href{https://fse.studenttheses.ub.rug.nl/25344/}{his bachelor thesis} to learn more about some interesting advanced topics in differential geometry.} for his comments, corrections and the appendices that he contributed to these notes.
Expand Down

0 comments on commit 1a6ab8b

Please sign in to comment.