forked from ernestyalumni/Gravite
-
Notifications
You must be signed in to change notification settings - Fork 29
/
Copy pathlecture11.tex
243 lines (206 loc) · 15 KB
/
lecture11.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
\section{Symmetry}
\begin{framed}
This lecture is about symmetry but we will pick a number of elementary techniques in differential geometry that we will need in Einstein's theory. We shall motivate these techniques by appealing to the feeling that the round sphere $(S^2, \mathcal{O}, \mathcal{A},g^{\text{round}})$ has rotational symmetry, while the potato $(S^2, \mathcal{O},\mathcal{A}, g^{\text{potato}})$ does not.
So far we have considered symmetry by having inner product first, and then demanding that w.r.t. that inner product we classify linear maps $A$ acting on vectors $X$ and $Y$ such that inner product of $AX$ and $AY$ results in inner product $XY$.
Here we talk about an altogether different idea. Firstly, since the distinction between the two is entirely contained in $g$, we are talking about the rotational symmetry of $g^{\text{round}}$. Secondly, while an inner product is on one tangent space, there are many different tangent spaces with different inner products. $g$ talks about the distribution of these inner products over the sphere, and that distribution in some sense is rotationally invariant or not.
Therefore, the question is: How to describe the symmetries of a metric? This is important because nobody has solved Einstein's Equations without assuming some sort of additional assumptions such as symmetry of the solution.
\end{framed}
\subsection{Push-forward map}
\begin{definition}
Let $M$ and $N$ be smooth manifolds with tangent bundles $TM$ and $TN$ respectively. Let $\phi : M \to N$ be a smooth map. Then, the \textbf{push-forward map} of $\phi$ is the map
\begin{align}
\phi_{\ast} : & TM \to TN \nonumber \\
& X \mapsto \phi_\ast(X) \nonumber \\
\label{eq_defPushForwardMap}\text{where } & \phi_\ast(X) f := X (f \after \phi) && (\forall \, f \in C^\infty(N))
\end{align}
\end{definition}
\begin{SCfigure}[5][h]
\label{fig:L11_pushForwardMap}
\centering
\begin{tikzpicture}
\matrix (m) [matrix of math nodes, row sep=4em, column sep=6em, minimum width=2em]
{
TM & TN & \\
M & N & \mathbb{R} \\
};
\path[->]
(m-1-1) edge node [auto] {$\phi_\ast$} (m-1-2)
edge node [auto] {$\pi_{TM}$} (m-2-1)
(m-1-2) edge node [auto] {$\pi_{TN}$} (m-2-2)
(m-2-2) edge node [auto] {$f$} (m-2-3)
(m-2-1) edge node [auto] {$\phi$} (m-2-2)
edge [bend right=30] node [auto] {$f \after \phi$} (m-2-3);
\end{tikzpicture}
\caption{\textbf{Push-forward map}: $\phi_\ast$ takes a vector $X \in T_pM$ in the tangent space at the point $p \in M$ to the vector $\phi_\ast(X) \in T_qN$ in the tangent space at the point $\phi(p) = q \in N$, such that the action of $\phi_\ast(X)$ on any smooth function $f \in C^\infty(N)$ results in the same value as the action of $X$ on the function $(f \after \phi)$.}
\end{SCfigure}
\textbf{Note}: If we take an entire fibre at the point $p \in M$, applying $\phi_\ast$ on it remains within the fibre at the point $\phi(p) \in N$. That is \\
\[
\phi_\ast(T_pM) \subseteq T_{\phi(p)}N
\]
\textit{Mnemonic: ``vectors are pushed forward'' across tangent bundles in a manner dictated by the underlying map.}
\textbf{Components of push-forward $\phi_\ast$ w.r.t charts $(U,x) \in \mathcal{A}_M$ and $(V,y) \in \mathcal{A}_N$}: Let $p \in U$ and $\phi(p) \in V$. Since $\cibasis{x^i}_p$ is a vector, we have $\phi_\ast(\cibasis{x^i}_p)$ as a vector in N. Then we can select a component of this vector by using $dy^a$ as follows: \\
\begin{equation} \label{eq_pushForwardComponents}
\underbrace{dy^a\left(\phi_\ast\left(\left(\cibasis{x^i}\right)_p\right)\right)}_{:= \phi^a_{\ast i}} = \phi_\ast\left(\left(\cibasis{x^i}\right)_p\right)y^a = \left(\cibasis{x^i}\right)_p (y^a \after \phi) = \left(\cibasis{x^i}\right)_p (y \after \phi)^a := \left(\cibasis[\hat{\phi}^a]{x^i}\right)_p
\end{equation}
\begin{SCfigure}[5][h]
\centering
\begin{tikzpicture}
\matrix (m) [matrix of math nodes, row sep=4em, column sep=6em, minimum width=2em]
{
M \supseteq U & V \subseteq N \\
\underbrace{x(U)}_{\subseteq \mathbb{R}^{\text{dim }M}} & \underbrace{y(V)}_{\subseteq \mathbb{R}^{\text{dim }N}}\\
};
\path[->]
(m-1-1) edge node [auto] {$\phi$} (m-1-2)
edge node [auto] {$x$} (m-2-1)
edge node [sloped, anchor=center, above] {$y \after \phi =: \hat{\phi}$} (m-2-2)
(m-1-2) edge node [auto] {$y$} (m-2-2)
(m-2-1) edge node [below] {$y \after \phi \after x^{-1}$} (m-2-2);
\end{tikzpicture}
\caption{Components of push-forward map w.r.t charts $(U,x) \in \mathcal{A}_M$ and $(V,y) \in \mathcal{A}_N$.}
\end{SCfigure}
\begin{theorem}
If $\gamma : \mathbb{R} \to M$ is a curve in $M$, then $\phi \after \gamma : \mathbb{R} \to N$ is a curve in $N$. Then, $\phi_\ast$ pushes the tangent to a curve $\gamma$ (velocity) to the tangent to the curve $(\phi \after \gamma)$, i.e.,
\begin{equation}
\phi_\ast\left(v_{\gamma, p}\right) = v_{(\phi \after \gamma), \phi(p)}
\end{equation}
\end{theorem}
\begin{proof}
Let $p = \gamma(\lambda_0)$. Then $\forall \, f \in C^\infty(N)$,
\begin{align*}
\phi_\ast\left(v_{\gamma, p}\right) f & = v_{\gamma, p} (f \after \phi) && (\text{by Eq.~\ref{eq_defPushForwardMap}}) \\
& = ((f \after \phi) \after \gamma)^\prime (\lambda_0) && (\text{by Eq.~\ref{eq_velocity}}) \\
& = (f \after (\phi \after \gamma))^\prime (\lambda_0) && (\text{by associativity of composition}) \\
& = v_{(\phi \after \gamma), \phi(\gamma(\lambda_0))} f && (\text{by Eq.~\ref{eq_velocity}}) \\
& = v_{(\phi \after \gamma), \phi(p)} f && (\gamma(\lambda_0) = p) \\
\implies \phi_\ast\left(v_{\gamma, p}\right) & = v_{(\phi \after \gamma), \phi(p)}
\end{align*}
\end{proof}
\subsection{Pull-back map}
\begin{definition}
Let $M$ and $N$ be smooth manifolds with cotangent bundles $T^\ast M$ and $T^\ast N$ respectively. Let $\phi : M \to N$ be a smooth map. Then, the \textbf{pull-back map} of $\phi$ is the map
\begin{align}
\phi^{\ast} : & T^\ast N \to T\ast M \nonumber \\
& \omega \mapsto \phi^\ast(\omega) \nonumber \\
\label{eq_defPullBackMap}\text{where } & \phi^\ast(\omega)(X) := \omega (\phi_\ast(X)) && (\forall \, X \in T_pM)
\end{align}
\end{definition}
\textbf{Components of pull-back $\phi^\ast$ w.r.t charts $(U,x) \in \mathcal{A}_M$ and $(V,y) \in \mathcal{A}_N$}: Let $p \in U$ and $\phi(p) \in V$. Since $dy^a$ is a covector, we have $\phi^\ast(dy^a_p)$ as a covector in N. Then we can select a component of this covector by using $\cibasis{x^i}$ as follows: \\
\begin{align*}
\phi^{\ast a} & := \phi^\ast\left(\left(dy^a\right)_{\phi(p)}\right) \left(\left(\cibasis{x^i}\right)_p\right) \\
& = \left(dy^a\right)_{\phi(p)} \phi_\ast \left(\left(\cibasis{x^i}\right)_p\right) && (\text{by Eq.~\ref{eq_defPullBackMap}}) \\
& = \left(\cibasis[\hat{\phi}^a]{x^i}\right)_p = \phi^a_{\ast i} && (\text{by Eq.~\ref{eq_pushForwardComponents}})
\end{align*}
Thus, the components of the push-forward and pull-back maps are exactly the same.
\begin{align*}
\left(\phi_\ast\left(X\right)\right)^a & = \phi^a_{\ast i} X^i \\
\left(\phi^\ast\left(\omega\right)\right)_i & = \phi^a_{\ast i} \omega_a
\end{align*}
Remember, $a = (1, \dotsc, \text{dim} N)$ and $i = (1, \dotsc, \text{dim} M)$.
Claim: $\phi^\ast(df) = d(f \after \phi)$.
\textit{Mnemonic: ``covectors are pulled back'' across tangent bundles in a manner dictated by the underlying map.}
\textbf{Important application}:
\begin{definition}
Let $M$ and $N$ be smooth manifolds. Let $\phi : M \injmapto N$ be an injective map. If we know a metric $g$ on $N$, then the \textbf{induced metric} $g_M$ in $M$ is defined using the push-forward map $\phi_\ast$ as follows:
\begin{equation}\label{eq_inducedMetric}
\boxed{g_M(X,Y) := g\left(\phi_\ast(X),\phi_\ast(Y)\right)} \quad \quad (\forall \, X, Y \in T_pM)
\end{equation}
\end{definition}
In terms of components,
\begin{equation}
\left(\left(g_M\right)_{ij}\right)_p = \left(g_{ab}\right)_{\phi(p)} \left(\cibasis[\hat{\phi}^a]{x^i}\right)_{\phi(p)} \left(\cibasis[\hat{\phi}^b]{x^j}\right)_{\phi(p)}
\end{equation}
\textbf{Example}: $N = (\mathbb{R}^3, \stdtop, \mathcal{A})$ and $M = (S^2, \mathcal{O}, \mathcal{A})$, then we can have several injective maps $\phi : S^2 \injmapto \mathbb{R}^3$. For example, $S^2$ could live in $\mathbb{R}^3$ either as a potato or a round sphere. However, suppose $\mathbb{R}^3$ is equipped with the Euclidean metric $g_E$ ...(TODO complete this example)
\subsection{Flow of a complete vector field}
\begin{definition}
Let $X$ be a vector field on a smooth manifold $(M,\mathcal{O},\mathcal{A})$. A curve $\gamma : I \subseteq \mathbb{R} \to M$ is called an \textbf{integral curve of $X$} if
\[
v_{\gamma,\gamma(\lambda)} = X_{\gamma(\lambda)}
\]
\end{definition}
\begin{definition} A vector field $X$ is \textbf{complete} if all integral curves have $I = \mathbb{R}$ (i.e. domain is all of $\mathbb{R}$).
\end{definition}
\begin{theorem}
Compactly supported smooth vector field is complete.
\end{theorem}
\begin{definition} The \textbf{flow of a complete vector field $X$} on a manifold $M$ is a 1-parameter family
\begin{align*}
h^X : & \mathbb{R} \times M \to M \\
& (\lambda, p) \mapsto \gamma_p(\lambda)
\end{align*}
where $\gamma_p : \mathbb{R} \to M$ is the integral curve of $X$ with $\gamma(0) = p$.
\end{definition}
Then for fixed $\lambda \in \mathbb{R}$, $h_{\lambda}^X : M \to M$ is smooth.
\textbf{Picture}: If $S$ is a set of points in $M$, then $h^X_{\lambda}(S)$ can be seen as the new position of these points under the flow $h^X$ after the passage of $\lambda$ units of parameter. In general, $h^X_{\lambda}(S) \neq S (\text{ if } X \neq 0)$.
\subsection{Lie subalgebras of the Lie algebra $(\Gamma(TM), [\cdot, \cdot])$ of vector fields}
We know that $\Gamma(TM) = \lbrace \text{ set of all vector fields } \rbrace$, which can be seen as a $C^{\infty}(M)$-module, or as a $\mathbb{R}$-vector space.
$(\Gamma(TM), [\cdot, \cdot])$ with $[X,Y]$ defined by its action on a function $f$ by $[X,Y] f := X(Yf) - Y(Xf)$ is a Lie algebra since $X,Y \in \Gamma(TM) \implies [X,Y] \in \Gamma(TM)$ and the following properties are satisfied:
\begin{enumerate}[(i)]
\item \textbf{Anticommutativity}: $[X,Y] = -[Y,X]$
\item \textbf{Linearity}: $[\lambda X + Z, Y] = \lambda [X,Y] + [Z,Y]$ where $\lambda \in \mathbb{R}$
\item \textbf{Jacobi identity}: $[X,[Y,Z]] + [Z,[X,Y]] + [Y,[Z,X]] = 0$
\end{enumerate}
Let $X_1, \dotsc, X_s$ be $s$ (many) vector fields on $M$, such that
\[
\forall i,j \in \lbrace 1,\dotsc,s \rbrace \quad [X_i,X_j] = \underbrace{C^k_{ij} X_k}_{\text{linear combination of }X_k s} \quad \quad \text{where } C^k_{ij} \in \mathbb{R}
\]
$C^k_{ij}$ are called \textbf{structure constants}.
Let $\text{span}_{\mathbb{R}} \lbrace X_1,\dotsc,X_s \rbrace := \lbrace \text{all linear combinations of }X_k \rbrace$. Then $\left(\text{span}_{\mathbb{R}} \lbrace X_1,\dotsc,X_s \rbrace, [\cdot, \cdot]\right)$ is a Lie subalgebra of $(\Gamma(TM), [\cdot, \cdot])$.
\textbf{Example}: In $S^2$, assume that the vector fields $X_1,X_2,X_3$ satisfy $[X_1,X_2] = X_3$, $[X_2,X_3] = X_1$ and $[X_3,X_1] = X_2$. Then $\left(\text{span}_{\mathbb{R}} \lbrace X_1,X_2,X_3 \rbrace, [\cdot, \cdot]\right)$ ($= SO(3)$) is a Lie subalgebra. An instance of vector fields satisfying these conditions is (with $X_i, \theta, \phi$ all taken at a point $p$, and $x^1 = \theta, x^2 = \phi$)
\begin{align*}
& X_1 = - \sin\phi \cibasis{\theta} - \cot\theta \cos\phi \cibasis{\phi} \\
& X_2 = \cos\phi \cibasis{\theta} - \cot\theta \cos\phi \cibasis{\phi} \\
& X_3 = \cibasis{\phi}
\end{align*}
Note that the above is defined on a merely smooth manifold without any additional structure like metric.
\subsection{Symmetry}
\begin{definition}
A finite-dimensional Lie subalgebra $(L,[\cdot,\cdot])$ is said to be a \textbf{symmetry} of a metric tensor field $g$ if
\[
\boxed{g \left( \left(h^X_\lambda\right)_\ast (A), \left(h^X_\lambda\right)_\ast (B) \right) = g(A,B)} \quad\quad (\forall X \text{ (complete vector field) } \in L, \quad \forall \lambda \in \mathbb{R}, \quad \forall A, B \in T_pM)
\]
\end{definition}
In another formulation (using pullback), $\boxed{\left(h^X_\lambda\right)^\ast g = g}$. The pullback of $\phi : M \to M$ on $g$, itself, is defined as follows:
\[
(\phi^\ast g)(A, B) := g(\phi_\ast(A), \phi_\ast(B))
\]
%\textbf{Example}:
\subsection{Lie derivative}
It can be shown that the following expression is precisely the Lie derivative of $g$ w.r.t a vector field
\begin{equation}
\boxed{\mathcal{L}_X g := \lim_{\lambda \to 0} \frac{\left( h_\lambda^X \right)^\ast g - g}{\lambda}}
\end{equation}
Clearly, $L$ is a symmetry of $g$, iff $\mathcal{L}_X g = 0$.
\begin{definition}
The \textbf{Lie derivative} $\mathcal{L}$ on a smooth manifold $(M, \mathcal{O}, \mathcal{A})$ a pair of a vector field $X$ and a $(p,q)$-tensor field $T$ to a $(p,q)$-tensor field such that
\begin{enumerate}[(i)]
\item $\mathcal{L}_X f = Xf \quad \forall f \in C^{\infty}M$
\item $\mathcal{L}_X Y = [X, Y] \quad\quad \text{ where }X, Y \text{ are vector fields}$
\begin{framed}
This condition sucks in information about the vector field $X$. It is not $C^{\infty}$-linear in the lower index. If it were, the derivative would be independent of values of $X$ at nearby points to the point where derivative is evaluated. This is an important difference between the covariant derivative $\nabla$ and the Lie derivative.
\end{framed}
\item $\mathcal{L}_X (T + S) = \mathcal{L}_X T + \mathcal{L}_X S \quad \text{ where }T, S \text{ are } (p,q) \text{-tensors}$
\item \textbf{Leibnitz rule: } $\mathcal{L}_X T(\omega_1,\dotsc,\omega_p,Y_1,\dotsc,Y_q) = (\mathcal{L}_X T)(\omega_1,\dotsc,\omega_p,Y_1,\dotsc,Y_q) \\
+ T(\mathcal{L}_X \omega_1,\dotsc,\omega_p,Y_1,\dotsc,Y_q) + \dotsb + T(\omega_1,\dotsc,\mathcal{L}_X \omega_p,Y_1,\dotsc,Y_q) \\
+ T(\omega_1,\dotsc,\omega_p,\mathcal{L}_X Y_1,\dotsc,Y_q) + \dotsb + T(\omega_1,\dotsc,\omega_p,Y_1,\dotsc,\mathcal{L}_X Y_q) \quad \text{ where }T \text{ is a }(p,q)\text{-tensor}$
\begin{framed}
Note that for a $(p,q)$-tensor $T$ and a $(r,s)$-tensor $S$, since: \\
$(T \otimes S) (\omega_{(1)}, \dotsc, \omega_{(p+r)}, Y_{(1)}, \dotsc, Y_{(q+s)}) = \\ T(\omega_{(1)}, \dotsc, \omega_{(p)}, Y_{(1)}, \dotsc, Y_{(q)} ) \cdot S( \omega_{(p+1)}, \dotsc, \omega_{(p+r)} , Y_{(q+1)}, \dotsc, Y_{(q+s)})$, \\
Leibnitz rule implies $\mathcal{L}_X (T \otimes S) = (\mathcal{L}_X T) \otimes S + T \otimes (\mathcal{L}_X S)$.
\end{framed}
\item $\mathcal{L}_{X+Y} T = \mathcal{L}_X T + \mathcal{L}_Y T$
\end{enumerate}
\end{definition}
Observe that, in chart $(U,x)$
\[
\left(\mathcal{L}_X Y\right)^i = X^m \cibasis{x^m}(Y^i) - \underbrace{\cibasis{x^s} (X^i)}_{\text{requires knowing } X \text{ around the point} } Y^s
\]
However, for covariant derivative
\[
\left(\nabla_X Y\right)^i = X^m \cibasis{x^m}(Y^i) + \ccf{i}{sm} X^m Y^s
\]
In general, for a $(1,1)$-tensor $T$
\[
\left(\mathcal{L}_X T \right)\indices{^i_j} = X^m \cibasis{x^m}(T\indices{^i_j}) \underbrace{- \cibasis[X^i]{x^s} T\indices{^x_j}}_{('-' \text{ for lower index})} \underbrace{+ \cibasis[X^s]{x^j} T\indices{^i_s}}_{('+' \text{ for upper index})}
\]
\textbf{Application}: As above, it is easy to calculate components of Lie derivative of metric g, $\mathcal{L}_X g$. Thus, by checking whether the derivative equals $0$ or not, it can be determined whether a metric features a symmetry.