← Introduction → Lorentz Spaces and Relativity
What should you be acquainted with? 1. Linear Algebra in particular: inner product spaces, also called spaces with a Euclidean product over the real numbers. We refer to Introduction to Linear Algebra by Serge Lang or Linear Algebra Problem Book by Paul Halmos or Sage for Linear Algebra.

Inner Products

Linear Algebra Refreshed

This is not intended to be a basic course in linear algebra!

Dual and bi-dual space

We will only consider finite dimensional vector-spaces over the reals. The dual of $E$ will be denoted by $E^*$, i.e. $E^*$ is the vector-space of all linear mappings (or linear functionals) $x^*:E\rar\R$. If $e_1,\ldots,e_n$ is a basis for $E$, then we define its dual basis
$e_j^*\in E^*$ by $e_j^*(e_k)\colon=\d_{jk}$. For any $x\in E$ the numbers $e_j^*(x)$ are just the coefficents of $e_j$ in the decomposition of $x$, i.e. $$ x=\sum_j e_j^*(x)e_j~. $$ Now for all $x\in E$ the mapping $x^*\mapsto x^*(x)$ is a linear functional on $E^*$ and thus defines an element of the bi-dual $E^{**}$. $E$ and $E^{**}$ are in fact canonically isomorphic, i.e. there is an isomorphism $J:E\rar E^{**}$, which is defined without refering to any basis; indeed, define \begin{equation}\label{lareq1}\tag{LAR1} J:E\rar E^{**},\quad J(x)(x^*)\colon=x^*(x)~. \end{equation} Of course, any two finite dimensional vector-spaces of the same dimension are isomorphic, but in general we need to know bases of both spaces to define such an isomorphism. In our case $J$ is defined without refering to any basis. It is this isomorphism $J$ that allows to identify a finite dimensional vector-space with its bi-dual: if $x^{**}:E^*\rar\R$ is linear, then there is a unique $x\in E$, such that for all $x^*\in E^*$: $x^{**}(x^*)=x^*(x)$. In other words: every linear functional on $E^*$ is the evaluation at some vector $x\in E$. The benefit of this observation is that whenever a property $P$ for a finite dimensional vector-space $E$ implies a property $Q$ for its dual $E^*$, then property $P$ for $E^*$ implies property $Q$ for $E$.
Verify that $Je_1,\ldots,Je_n$ is the dual basis to $e_1^*,\ldots,e_n^*$, i.e. $e_j^{**}=Je_j$, and that for all $x^*\in E^*$: $$ x^*=\sum_j x^*(e_j)e_j^*~. $$
Let $E$ be the space of all real polynomials on $\R$ of degree less than $n$. Suppose $x_1,\ldots,x_n\in\R$ are pairwise distinct. Show that the linear functionals $x_j^*(p)\colon=p(x_j)$ form a basis of $E^*$.
Given a basis $e_1,\ldots,e_n$ and its dual basis $e_1^*,\ldots,e_n^*$, we define the matrix $(a_{jk})$ of $u\in\Hom(E)$ - i.e. $u:E\rar E$ is linear - with respect to this basis by \begin{equation}\label{lareq2}\tag{LAR2} a_{jk}\colon=e_j^*(u(e_k)) \quad\mbox{i.e.}\quad u(e_k)=\sum_j a_{jk}e_j~. \end{equation} Of course, we could consider matrices of linear mappings with different domain and range and different bases for these spaces. However, we will hardly use this level of generality and thus stick to the case of linear maps $u:E\rar E$, i.e. endomorphisms, and a single basis for $E$. Anyway, here is the general definition:
Given a linear map $u:E\rar F$, bases $e_1,\ldots,e_m$ for $E$ and $f_1,\ldots,f_n$ for $F$ with dual basis $f_1^*,\ldots,f_n^*\in F^*$; the $n\times m$-matrix $A$ of $u$ with respect to these bases is given by $A\colon=(f_j^*(u(e_k))_{j,k=1}^{n,m}\in\Ma(n,m,\R)$.
Now suppose $(a_{jk})$ is the matrix of an operator $u\in\Hom(E)$ with respect to a basis $e_1,\ldots,e_n$ and $x\in E$ is an arbitary vector. Given the components $x_j\colon=e_j^*(x)$ of $x$ with respect to the basis $e_1,\ldots,e_n$, how can we compute the components $y_1,\ldots,y_n$ of $u(x)$ with respect to the basis $e_1,\ldots,e_n$? By definition these numbers are given by $$ y_j\colon=e_j^*(u(x)) =e_j^*\Big(u\Big(\sum_k x_ke_k\Big)\Big) =e_j^*\Big(\sum_k x_ku(e_k)\Big) =\sum_k x_ka_{jk} =\sum_k a_{jk}x_k $$ and this is just the $j$-th row of an $n\times1$ matrix obtained by multiplying the matrix $(a_{jk})$ from the right with the $n\times 1$-matrix $(x_1,\ldots,x_n)^t$, i.e. $$ \left(\begin{array}{c} y_1\\ \vdots\\ y_n \end{array}\right) =\left(\begin{array}{c} a_{11}&\cdots&a_{1n}\\ \vdots&\ddots&\vdots\\ a_{n1}&\cdots&a_{nn} \end{array}\right) \left(\begin{array}{c} x_1\\ \vdots\\ x_n \end{array}\right) $$
Let $A,B$ be the matrices of $u,v\in\Hom(E)$ with respect to a basis $e_1,\ldots,e_n$ for $E$. The $AB$ is the matrix of the composition $uv$ with respect to $e_1,\ldots,e_n$. If $u$ is an isomorphism then the matrix of $u^{-1}$ with respect to $e_1,\ldots,e_n$ is $A^{-1}$.
Suppose $e_1,\ldots,e_n$ and $b_1,\ldots,b_n$ are two bases for $E$, then the matrix $U=(u_{jk})$ of the linear map $u:e_j\mapsto b_j$ with respect to the basis $e_1,\ldots,e_n$ is given by $u_{jk}=e_j^*(b_k)$ and its inverse $U^{-1}=(u^{jk})$ is given by $u^{jk}=b_j^*(e_k)$.
By \eqref{lareq2}: $u_{jk}=e_j^*(u(e_k))=e_j^*(b_k)$ and $$ \sum_l b_j^*(e_l)e_l^*(b_k) =b_j^*\Big(\sum_l e_l^*(b_k)e_l\Big) =b_j^*(b_k) =\d_{jk}~. $$ Analogously we get $\sum_l e_j^*(b_l)b_l^*(e_k)=\d_{jk}$.
Let $A\colon=(a_{jk})$ be the matrix of $v\in\Hom(E)$ with respect to the basis $e_1,\ldots,e_n$ and let $U$ be the matrix of exam. Then $B=(b_{jk})\colon=U^{-1}AU$ is the matrix of $v\in\Hom(E)$ with respect to the basis $b_1,\ldots,b_n$. Suggested solution.
Suppose $f,x_1^*,\ldots,x_n^*\in E^*$ such that $\bigcap\ker x_j^*\sbe\ker f$. Prove that $f$ is a linear combination of $x_1^*,\ldots,x_n^*$. Suggested solution.

Rank-Nullity Theorem

Finally let us note the well known Rank-Nullity Theorem: Suppose $u\in\Hom(E,F)$ - i.e. $u:E\rar F$ is linear - and $\dim E < \infty$, then \begin{equation}\label{lareq5}\tag{LAR5} \dim\im u+\dim\ker u=\dim E~. \end{equation} It's an obvious consequence of the Isomorphism Theorem, which states that $\im u$ is isomorphic to the quotient space $E/\ker u$ and hence $$ \dim\im u=\dim(E/\ker u)=\dim E-\dim\ker u~. $$ A unique isomorphism $\wh u:E/\ker u\rar\im u$ is given by requiring $\wh u\pi=u$, where $\pi:E\rar E/\ker u$ denotes the quotient map.
An application of the Rank-Nullity Theorem is the frequently utilized fact that two subspace $F_1$ and $F_2$ of $E$ which satisfy $\dim F_1+\dim F_2 > \dim E$ have a non-trivial intersection, more precisely: $$ \dim(F_1\cap F_2)\geq\dim F_1+\dim F_2-\dim E $$ and equality holds if and only if $F_1+F_2=E$. To see this define $u:F_1\times F_2\rar E$ by $(x,y)\mapsto x+y$. By the Rank-Nullity Theorem we have: $$ \dim\ker u=\dim(F_1\times F_2)-\dim\im u \geq\dim F_1+\dim F_2-\dim E $$ with equality iff $u$ is onto, i.e. iff $F_1+F_2=E$. Obviously $\ker u=\{(x,-x):x\in F_1\cap F_2\}$ is isomorphic to $F_1\cap F_2$, i.e. $\dim\ker u=\dim(F_1\cap F_2)$.

Entire functions of an endomorphism

If $f(x)=a_0+a_1x+\cdots a_nx^n$ is a polynomial of degree $n$ and $u\in\Hom(E)$, then it's pretty clear how to define $f(u)$: $$ f(u)\colon=a_0u^0+a_1u+\cdots a_nu^n, $$ where $u^0$ is the identity, denoted by $1$ or $id$. The following definition is intended to extend the notion of a function $f(u)$ of an operator $u\in\Hom(E)$ from polynomials $f$ to entire functions $f$.
For any $u\in\Hom(E)$ the exponential $e^u\in\Hom(E)$ is defined by: $$ \exp(u)=e^u\colon=\sum_{k=0}^\infty\frac{u^k}{k!}~. $$ More generally, let $f:\R\rar\R$ be an entire function, i.e. the Taylor-series $\sum a_kx^k$ of $f$ at e.g. $0$ converges absolutely in all points $x\in\R$ to $f(x)$. Then we define for any $u\in\Hom(E)$: $$ f(u)\colon=\sum_{k=0}^\infty a_ku^k~. $$
Here we actually need some minor facts from functional analysis: $E$ with any norm is complete and thus $\Hom(E)$ with the operator norm is also complete; since the sum is absolutely convergent, the operator $f(u)$ is well defined. We won't discuss these facts, for we won't need it elsewhere. Also, by the Cayley-Hamilton Theorem $$ f(u)=a_{n-1}u^{n-1}+\cdots+a_1u+a_0u^0, $$ where $a_{n-1},\ldots,a_0$ are real numbers. Usually we don't write $a_0u^0$ but simply $a_0$, because it should be clear that we don't mean the real number $a_0$ but this number times the identity.
Obviuously, if $A$ is the matrix of $u\in\Hom(E)$ with respect to some basis, then the exponential matrix $$ e^A\colon=\sum_{k=0}^\infty\frac{A^k}{k!} $$ is the matrix of the endomorphism $e^u$ with respect to the same basis.
Let $u\in\Hom(\R^2)$ be the linear map with matrix (with respect to some basis) $$ \left(\begin{array}{cc} 0&-1\\ 1&0 \end{array}\right) $$ 1. Verify that for any $t\in\R$ the matrix of $e^{tu}$ (with respect to the same basis) is given by $$ \left(\begin{array}{cc} \cos t&-\sin t\\ \sin t&\cos t \end{array}\right) $$ 2. Find functions $a_0(t)$ and $a_1(t)$ such that $e^{tu}=a_0(t)+a_1(t)u$.
Let $u\in\Hom(\R^2)$ be the linear map with matrix (with respect to some basis) $$ \left(\begin{array}{cc} 0&1\\ 1&0 \end{array}\right) $$ 1. Verify that for any $t\in\R$ the matrix of $e^{tu}$ (with respect to the same basis) is given by $$ \left(\begin{array}{cc} \cosh t&\sinh t\\ \sinh t&\cosh t \end{array}\right) $$ 2. Find functions $a_0(t)$ and $a_1(t)$ such that $e^{tu}=a_0(t)+a_1(t)u$.

The adjoint of linear mappings

Let $E,F$ be vector-spaces. The adjoint or dual of a linear mapping $u:E\rar F$ is the linear mapping $u^*:F^*\rar E^*$ defined by \begin{equation}\label{lareq3}\tag{LAR3} \forall x\in E,\,\forall y^*\in F^*:\qquad u^*(y^*)(x)=y^*(u(x))~. \end{equation} Sometimes $u^*(y^*)$ is called the pull-back of $y^*$, because $u$ maps $E$ into $F$ and $u^*$ pulls back the functional $y^*$ on $F$ to the functional $u^*(y^*)$ on $E$. Suppose $e_1,\ldots,e_n$ is a basis of $E$ with dual basis $e_1^*,\ldots,e_n^*$, then $e_j^{**}\colon=J(e_j)$ is the dual basis to $e_j^*$ (cf. exam) and thus for any linear mapping $u:E\rar E$: $$ e_j^{**}(u^*(e_k^*)) =J(e_j)(u^*(e_k^*)) =u^*(e_k^*)(e_j) =e_k^*(u(e_j)), $$ i.e. the matrix of $u^*$ with respect to the dual basis $e_1^*,\ldots,e_n^*$ is the transposed matrix of the matrix of $u$ with respect to the basis $e_1,\ldots,e_n$.
Verify that for all $u\in\Hom(E)$: $Ju=u^{**}J$ - or in diagram form:
diagram1
Thus identifying $E$ with $E^{**}$ (via $J$), i.e. we consider $Jx$ and $x$ to be the same, we may simply write: $u^{**}=u$.
Suggested solution.
If $E$ and $F$ are finite dimensional vector-spaces, then a linear mapping $u:E\rar F$ is onto if and only if $u^*$ is one-one and since $u^{**}=u$, $u$ is one-one if and only if $u^*$ is onto.
Suppose $u^*(y^*)=0$ for some $y^*\in F^*\sm\{0\}$, then for all $x\in E$: $0=u^*(y^*)(x)=y^*(u(x))$, i.e. $u$ cannot be onto. Conversely if $u$ is not onto, then there is some $y^*\in F^*\sm\{0\}$, such that for all $x\in E$: $0=y^*(u(x))=u^*(y^*)(x)$, i.e. $u^*(y^*)=0$.
A linear system of equations $Ax=y$ is solvable for every $y$ iff the linear system $A^tx=0$ only admits the trivial solution $x=0$.
If $u,v\in\Hom(E)$, then the following holds:
  1. The mapping $u\mapsto u^*$ is a linear mapping from $\Hom(E)$ into $\Hom(E^*)$.
  2. $(uv)^*=v^*u^*$, $u^{-1*}=u^{*-1}$ and $(e^u)^*=e^{u^*}$.
  3. If $u$ and $v$ commute, i.e. $uv=vu$, then: $e^{u+v}=e^u e^v$ and in particular $1=e^u e^{-u}$, i.e. $(e^u)^{-1}=e^{-u}$.
$\proof$ 2. $(uv)^*(x^*)(x)=x^*(uv(x))=u^*(x^*)(v(x))=v^*(u^*(x^*))(x)=v^*u^*(x^*)(x)$. In particular $1=1^*=(uu^{-1})^*=u^{-1*}u^*$ and analogously $1=u^*u^{-1*}$. Since for all $n\in\N_0$: $u^{*n}=u^{n*}$ it follows (by continuity of $*$): $(e^u)^*=e^{u^*}$. 3. If $uv=vu$, then the binomial formula holds, i.e.: $$ (u+v)^n=\sum_{k=0}^n{n\choose k}u^kv^{n-k} $$ and therefore by rearranging terms: $$ e^ue^v =\sum_{k,j=0}\frac{u^kv^j}{k!j!} =\sum_{n=0}\sum_{k=0}^n\frac{u^kv^{n-k}}{k!(n-k)!} =\sum_{n=0}\frac1{n!}\sum_{k=0}^n{n\choose k}u^kv^{n-k} =e^{u+v} $$ $\eofproof$

The trace

Suppose $u:E\rar E$ is linear and $e_1,\ldots,e_n$ an arbitrary basis with dual basis $e_1^*,\ldots,e_n^*$. then the trace
of $u$ is defined by \begin{equation}\label{lareq4}\tag{LAR4} \tr(u)\colon=\sum_je_j^*(u(e_j))~. \end{equation} Well, that's just a number, but what makes it important is the fact, that it doesn't depend on the basis: indeed, let $b_1,\ldots,b_n$ be another basis; using the facts that the matrices $(e_j^*(b_k))$ and $(b_j^*(e_k))$ are inverse (cf. exam) we get \begin{eqnarray*} \sum_jb_j^*(u(b_j)) &=&\sum_{j,k}b_j^*(u(e_k^*(b_j)e_k)) =\sum_{j,k}e_k^*(b_j)b_j^*(u(e_k))\\ &=&\sum_{j,k,l}e_k^*(b_j)b_j^*(e_l)e_l^*(u(e_k)) =\sum_{k,l}\d_{kl}e_l^*(u(e_k)) =\sum_ke_k^*(u(e_k))~. \end{eqnarray*} Thus the trace of $u\in\Hom(E)$ is just the sum of all diagonal elements of the matrix $A=(a_{jk})$ of $u$ with respect to any basis of $E$. In particular, if $u$ is diagonalizable with real eigen values $\l_1,\ldots,\l_n$, then $\tr u=\sum\l_j$; actually this is true for all $u\in\Hom(E)$ (cf. exam).
Prove that for all $u,v\in\Hom(E)$: $\tr(uv)=\tr(vu)$ and $\tr u^*=\tr u$. Suggested solution.
Beware! in general $\tr(uvw)\neq\tr(uwv)$.
Let $E$ be the space of all real polynomials of degree less than $n$ and $u$ the mapping which sends a polynomial $p$ to its formal derivative $p^\prime$. Compute the trace of $u$.
For $x\in E$ and $x^*\in E^*$ the mapping $x^*\otimes x:y\mapsto x^*(y)x$ is a linear mapping into $E$ and its trace is $x^*(x)$. Since any linear operator $u\in\Hom(E)$ is a linear combination of these type of operators (why?), the trace is uniquely defined by requiring that it must be a linear functional on $\Hom(E)$ such that for all $x\in E$ and all $x^*\in E^*$: $\tr(x^*\otimes x)=x^*(x)$. Anyhow, the trace of an endomorphism $u$ is defined algebraically, but is there also some geometric meaning of the trace?

Inner Product Spaces

Cf. students version in german.

Bi-linear forms

What Euclidean spaces are to classical physics, Lorentz spaces are to relativistic physics. The algebraic concept which comprises both is the concept of an inner product space: A bi-linear mapping (or bi-linear form) $B:E\times E\rar\R$ is called symmetric, if for all $x,y\in E$: $B(x,y)=B(y,x)$; $B$ is said to be non-singular, if for any $x\in E\sm\{0\}$ there is a vector $y\in E$ such that $B(x,y)\neq0$.
Suppose $B:E\times E\rar\R$ is bi-linear. $B$ is non-singular if and only if for some basis $e_1,\ldots,e_n$ for $E$: $$ \forall j=1\ldots,n:\quad B(x,e_j)=B(y,e_j) \quad\Rar\quad x=y~. $$
If $B$ is non-singular and for all $j$: $B(x,e_j)=B(y,e_j)$, then $B(x-y,e_j)=0$ for all $j$ and thus for all $z\in E$: $B(x-y,z)=0$, i.e. $x-y=0$. Conversely, if $B(x,z)=0$ for all $z\in E$, then $B(x,e_j)=0$ for all $j$, i.e. $B(x,e_j)=B(0,e_j)$ and thus: $x=0$.
The polarization formula $$ 4B(x,y)=B(x+y,x+y)-B(x-y,x-y) $$ evidently shows that any symmetric bi-linear form $B:E\times E\rar\R$ is uniquely determined by the quadratic form $x\mapsto B(x,x)$.
A quadratic form $q$ on $\R^n$ is just a homogeneous polynomial of degree two, i.e. for $x=(x_1,\ldots,x_n)\in\R^n$: $$ q(x)=\sum_{j,k=1}^na_{jk}x_jx_k, \quad a_{jk}\in\R~. $$ By the polarization formula there is a unique symmetric bi-linear form $B:\R^n\times\R^n\rar\R$ such that $B(x,x)=q(x)$ and this form is given by $$ B(x,y)=\tfrac12\sum_{j,k=1}^n(a_{jk}+a_{kj})x_jy_k~. $$
An inner product $\la.,.\ra$ on $E$ is a real valued, bi-linear, symmetric and non-singular mapping on $E\times E$. The inner product is said to be Euclidean if for all $x\in E\sm\{0\}$: $\la x,x\ra > 0$.
Here is an inifinite dimensional example:
Suppose $\mu$ is a signed measure on some measurable space $(\O,\F)$. Then $$ \la f,g\ra\colon=\int_\O fg\,d\mu $$ is an inner product on $E\colon=L_2(\O,\F,|\mu|)$.
If $\la.,.\ra$ is an inner product on $E$, then $x\mapsto\la x,x\ra$ is the associated quadratic form and $$ \Vert x\Vert\colon=|\la x,x\ra|^{1/2} $$ the corresponding norm of $x$ - though it's a norm in the usual sense in the Euclidean case only! This is why we will rarely use this notion for the whole space. The set $S(E)\colon=\{x\in E:\la x,x\ra=\pm 1\}$ is called the unit sphere of $E$. For $n\in\N$ and $\nu=0,\ldots,n$ we denote by $\R_\nu^n$ the vector-space $\R^n$ furnished with the inner product \begin{equation}\label{ipseq1}\tag{IPS1} \la x,y\ra\colon=-\sum_{1\leq j\leq\nu} x_jy_j+\sum_{j>\nu} x_jy_j~. \end{equation}
Show that $\R_\nu^n$ is indeed an inner product space and for $\nu=1,\ldots,n-1$ the quadratic form $x\mapsto\la x,x\ra$ takes on any real value. For $\nu=0$ these values are always non-negative and for $\nu=n$ these values are always non-positive.
The unit sphere of $\R_\nu^n$ will also be denoted by $S_\nu^{n-1}$, i.e.: $$ S_\nu^{n-1} =\Big\{x=(x_1,\ldots,x_n)\in\R^n:-\sum_{j=1}^\nu x_j^2+\sum_{j=\nu+1}^n x_j^2=\pm1\Big\},\quad S^{n-1}\colon=S_0^{n-1}~. $$
If $E$ is an $n$-dimensional real vector space, then $(u,v)\mapsto\tr(uv)$ is an inner product on $\Hom(E)$.
Take any basis $e_1,\ldots,e_n$ of $E$ and denote by $(a_{jk})$ and $(b_{jk})$ the corresponding matrices of $u\in\Hom(E)$ and $v\in\Hom(E)$, respectively. Then $$ \tr(uv)=\sum_{j,k}a_{jk}b_{kj}~. $$ Given $(a_{jk})$ we put $b_{jk}\colon=a_{kj}$ and for this choice we get $\tr(uv)=\sum a_{jk}^2$, which doesn't vanish provided $u\neq0$.

Lagrange-Sylvester Theorem

Two vectors $x,y$ in an inner product space $(E,\la.,.\ra)$ are said to be orthogonal
(or perpendicular) to each other, if $\la x,y\ra=0$. For $A\sbe E$ let $\lhull{A}$ be the subspace generated by $A$. The set $A^\perp$ is the set of all vectors orthogonal to all vectors in $A$, i.e. \begin{equation}\label{ipseq2}\tag{IPS2} A^\perp\colon=\{x\in E:\forall y\in A:\la x,y\ra=0\}~. \end{equation} Then $A^\perp$ is a subspace and $A^\perp=\lhull{A}^\perp$. For $x\in A^\perp$ we will write: $x\perp A$ and if $A=\{y\}$: $x\perp y$.
Let $e_0=(1,0)$ and $e_1=(0,1)$ be the canonical basis of $\R_1^2$. For $x=ae_0+be_1\in\R_1^2$ the space $x^\perp$ is the subspace generated by the vector $be_0+ae_1$. Thus if $a=b$, then $x$ is orthogonal to itself - that's a new feature, that does not occure in the Euclidean case!
Let $(E,\la.,.\ra)$ be an inner product space. 1. Verify the parallelogram identity $$ \la x+y,x+y\ra+\la x-y,x-y\ra=2\la x,x\ra+2\la y,y\ra~. $$ 2. If $\la x,x\ra=\la y,y\ra$, then $x-y$ and $x+y$ are orthogonal.
Our first and only result is about particular bases in inner product spaces; these bases are called orthonormal bases and generalize the concept of orthonormal bases in Euclidean spaces. In linear algebra the subsequent theorem is known under Sylvester's law of inertia or Lagrange method for quadratic forms.
Let $E$ be an $n$-dimensional vector-space and $\la.,.\ra$ an inner product on $E$. Then there exists a basis $e_1,\ldots,e_n$ for $E$ and a uniquely determined number $\nu\in\N_0$, which does not depend on the basis, such that for all $j,k=1,\ldots,n$: $$ \la e_j,e_k\ra=\e_j\d_{jk} \quad\mbox{where}\quad \forall j\leq\nu:\e_j=-1 \quad\mbox{and}\quad \forall j>\nu:\e_j=+1~. $$ $\nu$ is said to be the index of the inner product.
$\proof$ Let $(.|.)$ be an arbitrary Euclidean product on $E$ and define a linear mapping $A:E\rar E$ by $(Ax|y)=\la x,y\ra$. Then $A:(E,(.|.))\rar(E,(.|.))$ is self-adjoint - because $\la.,.\ra$ is symmetric - and invertible - because $\la.,.\ra$ is non-singular. Let $E^-$ and $E^+$, respectively, be the subspaces generated by all eigen vectors with stricly negative eigen values and strictly positive eigen values, respectively (cf. spectral theorem). These subspaces are orthogonal with respect to $(.|.)$; but they are also orthogonal with respect to $\la.,.\ra$: choose a basis $x_1,\ldots$ for $E^+$ such that $Ax_j=a_jx_j$ for some $a_j > 0$; then for all $x=\sum\l_jx_j\in E^+$ and all $y\in E^-$: $$ \la x,y\ra =(Ax|y) =(\sum\l_jAx_j|y) =\sum a_j\l_j(x_j|y)=0~. $$ If $x_1,\ldots,x_n$ is an orthonormal basis of the Euclidean space $(E,(.|.))$ and $Ax_j=a_jx_j$ for some $a_j\in\R\sm\{0\}$, then $e_j\colon=x_j/\sqrt{|a_j|}$ has the desired properties.
Putting $\nu\colon=\dim(E^-)$ we get a number which by definition depends on the Euclidean product $(.|.)$ choosen initially. To bypass this dependency we define a subspace $N$ as follows: Let $N$ be a subspace of $E$ of maximal dimension, such that for all $x\in N\sm\{0\}$: $\la x,x\ra < 0$ and put $\nu\colon=\dim(N)$. For $x\in E^-\sm\{0\}$ we have: $\la x,x\ra < 0$ and therefore: $\nu\geq\dim(E^-)$. Suppose that $\nu > \dim(E^-)$, then there is a non trivial vector $x\in N\cap E^+$, which is impossible and thus $\dim N=\dim E^-$. $\eofproof$
Such a basis $e_1,\ldots,e_n$ is called an orthonormal basis for $E$; in this case we have for all $x\in E$ the expansion: \begin{equation}\label{ipseq3}\tag{IPS3} x=\sum_{j=1}^n\e_j\la x,e_j\ra e_j \quad\mbox{where}\quad \e_j\colon=\la e_j,e_j\ra=\pm1 \end{equation} as can be easily verified by employing exam. Moreover, if $b_1,\ldots,b_n$ is another orthonormal basis, then $\nu=|\{j:\la b_j,b_j\ra=-1\}$. If $\nu=0$, then $\la.,.\ra$ is a Euclidean product and $(E,\la.,.\ra)$ is a Euclidean space. If $\nu=1$, then $(E,\la.,.\ra)$ is called a Lorentz space and $\la.,.\ra$ is said to be a Lorentz product.
The sum of Euclidean products is again a Euclidean product, but the sum of two Lorentz products need not be an inner product.
Obviously, the canonical basis $e_1,\ldots,e_n$ - i.e. $e_j\colon=(0,\ldots,0,1,0,\ldots,0)$ where $1$ is in the $j$-th slot - of $\R^n$ is an orthonormal basis for all spaces $\R_\nu^n$. The following example in particular demonstrates that given a basis $e_1,\ldots,e_n$, we can choose a Euclidean product $(.|.)$ such that $e_1,\ldots,e_n$ is orthonormal with respect to this Euclidean product.
Suppose $E$ is an $n$-dimensional vector-space and $e_1,\ldots,e_n$ a basis. Define for any $\nu\in\{0,\ldots,n\}$: $$ \la x,y\ra \colon=-\sum_{j\leq\nu}e_j^*(x)e_j^*(y) +\sum_{j>\nu}e_j^*(x)e_j^*(y)~. $$ Then $\la.,.\ra$ is an inner product with index $\nu$ and $e_1,\ldots,e_n$ is an orthonormal basis with respect to this inner product.
Given an arbitary inner product, how to find the index of this product and how can we find an orthonormal basis? Suppose $e_1,\ldots,e_n$ is an arbitrary basis for $E$ and $a_{jk}\colon=\la e_j,e_k\ra$ - the matrix $(a_{jk})$ is called the Gramian of the basis $e_1,\ldots,e_n$ (cf. subsection); by declaring $e_1,\ldots,e_n$ to be orthonormal with respect to $(.|.)$ (cf. exam), the proof of the theorem shows that $(a_{jk})$ is the matrix of the operator $A$ with respect to the basis $e_1,\ldots,e_n$ and the index $\nu$ is just the dimension of the sum of the eigen spaces for all strictly negative eigen values of $A$. An orthonormal basis is formed by the suitably normalized eigen vectors of $A$.
Verify that $(x_1,x_2),(y_1,y_2)\mapsto x_1y_2+x_2y_1$ defines a Lorentz product on $\R^2$. Find an orthonormal basis.
Let $e_1,e_2$ be the canonical basis of $\R^2$. Then the matrix $(a_{jk})$ is given by $a_{11}=a_{22}=0$ and $a_{12}=a_{21}=1$. The operator (or matrix) $A$ has eigen values $-1,1$ - thus it's a Lorentz product - with eigen vectors $b_1=e_1-e_2$ and $b_2=e_1+e_2$. Since $\la b_1,b_1\ra=-2$ and $\la b_2,b_2\ra=2$ the vectors $b_1/\sqrt2$ and $b_2/\sqrt2$ form an orthonormal basis.
Verify that $(x_1,x_2),(y_1,y_2)\mapsto x_1y_2+x_2y_1-x_3y_3$ defines an inner product on $\R^3$. Find its index and determine an orthonormal basis.
The map $A\mapsto\det(A)$ is a quadratic form on $\Ma(2,\R)$, the space of all real $2\times2$ matrices. Verify that the associated symmetric bi-linear form $g:\Ma(2,\R)\times\Ma(2,\R)\rar\R$, $g(A,A)=\det(A)$ is an inner product and compute its index. Suggested solution.
The space $E$ of all complex hermitian $2\times2$ matrices $A$ is a real vector-space of dimension $4$. Verify that $A\mapsto-\det(A)$ is a quadratic form and the associated symmetric bi-linear form $g:E\times E\rar\R$, $g(A,A)=-\det(A)$ is a Lorentz product. Cf. exam.
Verify that the inner product space $\Ma(2,\R)$, $\la A,B\ra\colon=\tr(AB)$, is a Lorentz space, i.e. the inner product has index $1$.
Verify that $\la A,B\ra\colon=\tr(AB^t)$ is a Euclidean product on $\Ma(n,\R)$.
Compute the index of the inner product space $\Ma(n,\R)$, $\la A,B\ra\colon=\tr(AB)$. Suggested solution

The musical isomorphisms

Let $B$ be a bi-linear form on an $n$-dimensional vector-space $E$. For $x\in E$ let $x^\flat\in E^*$ be the linear functional defined by $$ \forall y\in E:\quad x^\flat(y)\colon=B(x,y)~. $$ We first verify that

$x\mapsto x^\flat$ is an isomorphism from $E$ onto $E^*$ if and only if $B$ is non-singular.

Indeed, $x^\flat=0$ implies for all $y\in E$: $B(x,y)=0$; since $B$ is non-singular we infer that: $x=0$, i.e. $x\mapsto x^\flat$ one-one and since both $E$ and $E^*$ have the same dimension $x\mapsto x^\flat$ is an isomorphism. Conversely if $x\mapsto x^\flat$ is an isomorphism, then $B$ is non-singular, for suppose $B(x,y)=0$ for all $y\in E$, then $x^\flat(y)=0$ for all $y\in E$, i.e. $x^\flat=0$ and thus $x=0$.
Let $B:E\times E\rar\R$ be bi-linear and let $e_1,\ldots,e_n$ be any basis for $E$. Then the $n\times n$-matrix $(a_{jk})$, where $a_{jk}=B(e_j,e_k)$ is symmetric, if and only if $B$ is symmetric; it is invertible if and only if $B$ is non-singular.
$\proof$ 1. If $B$ is symmetric, then $(a_{jk})$ is obviously symmetric. Conversely the symmetry of $(a_{jk})$ implies by bi-linearity of $B$ that for all $x,y\in E$: $B(x,y)=B(y,x)$.
2. $B$ is non-singular if and only iff $x\mapsto x^\flat$ is an isomorphism, but the matrix of this map with respect to the bases $e_1,\ldots,e_n$ and $e_1^*,\ldots,e_n^*$ is given by $$ e_j^{**}(e_k^\flat) =J(e_j)(e_k^\flat) =e_k^\flat(e_j) =B(e_k,e_j)~. $$ Thus $B$ is non-singular iff $(a_{kj})$ is invertible, which in turn is equivalent to invertibility of its transpose $(a_{jk})$ (cf.
lemma). $\eofproof$
Show that $\o:\R^2\times\R^2\rar\R$, $(x_1,x_2),(y_1,y_2)\mapsto x_1y_2-x_2y_1$ is a non-singular bi-linear form, which is alternating (or anti-symmetric), i.e. $\o(y,x)=-\o(x,y)$.
If $B$ is non-singular then the inverse of $x\mapsto x^\flat$ will be denoted by $x^*\mapsto x^{*\sharp}$; thus we have: $$ \forall y\in E:\quad x^*(y)=B(x^{*\sharp},y)~. $$ These isomorphisms are called the musical isomorphisms.
Compute $e_j^{*\sharp}$ for $j=1,2$ for the bi-linear form $\o$ in the previous example.
For any linear functional $f:\Hom(E)\rar\R$ there is exactly one $u\in\Hom(E)$ such that for all $v\in\Hom(E)$: $f(v)=\tr(uv)$.
Now suppose $B=\la.,.\ra$ is an inner product then $$ \forall y\in E:\quad x^\flat(y)=\la x,y\ra=\la y,x\ra~. $$ Now let $\nu$ denote the index. For any orthonormal basis $e_1,\ldots,e_n$ we get: \begin{equation}\label{ipseq4}\tag{IPS4} \forall j:\quad e_j^{*\sharp}=\e_je_j \quad\mbox{where}\quad \e_j\colon=\la e_j,e_j\ra=\pm1~. \end{equation} More generally, if $b_j$ is an arbitrary basis with dual basis $b_j^*$, then there is some non-singular matrix $(g^{jk})$ such that for all $j$: \begin{equation}\label{ipseq5}\tag{IPS5} b_j^{*\sharp}=\sum_l g^{lj}b_l~. \end{equation} Moreover, we have $$ \d_{jk} =b_j^*(b_k) =\la b_j^{*\sharp},b_k\ra =\sum_l g^{lj}\la b_l,b_k\ra =\sum_l \la b_k,b_l\ra g^{lj}~. $$ Hence the matrix $(g^{jk})$ is the inverse of the matrix $g_{jk}\colon=\la b_j,b_k\ra$. The symmetric matrix $$ (\la b_j,b_k\ra)_{j,k=1}^n $$ is called the Gramian matrix of the basis $b_1,\ldots,b_n$ - notice, the Gramian depends on the inner product!

Convention: Whenever $(a_{jk})$ is an invertible matrix we denote its inverse by $(a^{jk})$.

Suppose $(E,\la.,.\ra)$ is an inner product space and $e_1,\ldots,e_n$ is an arbitrary basis with Gramian matrix $g_{jk}\colon=\la e_j,e_k\ra$, then \begin{equation}\label{ipseq6}\tag{ISP6} \tr u=\sum_{j,k=1}^n g^{jk}\la u(e_j),e_k\ra \end{equation} and in particular for an orthonormal basis $e_1,\ldots,e_n$ satisfying $\la e_j,e_j\ra=\e_j=\pm1$: $$ \tr u=\sum_{j=1}^n\e_j\la u(e_j),e_j\ra~. $$
By \eqref{ipseq5} we have by symmetry of $(g^{jk})$: $$ \tr u =\sum_je_j^*(u(e_j)) =\sum_j\la e_j^{*\sharp},u(e_j)\ra =\sum_{j,k}\la g^{kj}e_k,u(e_j)\ra =\sum_{j,k}g^{jk}\la u(e_j),e_k\ra~. $$ Usually the relation \eqref{ipseq6} is computationally a bit laborious:
$b_1=(1,2)$ and $b_2=(2,1)$ form a basis for $\R^2$. Compute the trace of the linear mapping $u$ given by $u(b_1)=b_1-b_2$, $u(b_2)=b_1+b_2$. Suggested solutions.
Suppose $E$ is an inner product space. For $u\in\Hom(E)$ compute the trace of the linear mapping $v\rar uv$ on the inner product space $\Hom(E)$ with inner product $(v,w)\mapsto\tr(v^*w)$.

Adjoint mappings on inner product spaces

If $E$ carries an inner product structure, then the musical isomorphisms identify $E$ and its dual $E^*$. Thus also the adjoint of a linear map $u:E\rar F$ may be viewed as a linear map $u^*:F\rar E$ - yes! we use in general the same symbol for this map. For clearity let us temporarily write $u^d$ for the dual map, i.e. $u^d\in\Hom(E^*)$.We just define $$ u^*(y^{*\sharp})\colon=u^d(y^*)^\sharp $$
sharp
Another way to state this: the mapping $u^*:F\rar E$ satisfies \begin{equation}\label{ipseq7}\tag{IPS7} \forall x\in E,y\in F:\qquad\la u^*(y),x\ra=\la y,u(x)\ra, \end{equation} Using this notation we get for the right hand side: $y^\flat(u(x))=u^d(y^\flat)(x)$ and the left hand side equals $u^*(y)^\flat(x)$, i.e. \eqref{ipseq7} is equivalent to: for all $y\in E$: $u^d(y^\flat)=u^*(y)^\flat$ or for all $y^*\in E^*$: $u^d(y^*)^\sharp=u^*(y^{*\sharp})$, which is just the definition of $u^*$.
For $u,v\in\Hom(E)$ and $\l\in\R$ we have: $1^*=1$, $(u+v)^*=u^*+v^*$, $(\l u)^*=\l u^*$, $(uv)^*=v^*u^*$, $u^{-1*}=u^{*-1}$ and $(e^u)^*=e^{u^*}$. More generally: for every entire function $f:\R\rar\R$: $f(u)^*=f(u^*)$.
Given an inner product space $(E,\la.,.\ra)$, an operator $u\in\Hom(E)$ is said to be self-adjoint (skew-symmetric), if $u^*=u$ ($u^*=-u$). $u$ is said to be a linear isometry if $u^*=u^{-1}$.
If $u\in\Hom(E)$ is self-adjoint, then $(x,y)\mapsto\la u(x),y\ra$ is a symmetric bi-linear form; it's an inner product iff $u$ is invertible.
If $u\in\Hom(E)$ is skew-symmetric, then for all $x\in E$: $\la u(x),x\ra=0$ and in particular $\tr u=0$.
$u\in\Hom(E)$ is an isometry iff for all $x,y\in E$: $\la u(x),u(y)\ra=\la x,y\ra$.
Use the polarization formula to prove that an isomorphism $u\in\Hom(E)$ is an isometry iff for all $x\in E$: $\la u(x),u(x)\ra=\la x,x\ra$.
Given the matrix $A=(a_{jk})$ of $u\in\Hom(E)$ with respect to an orthonormal basis $e_1,\ldots,e_n$ we want to determine the matrix $A^*\colon=(a_{jk}^*)$ of the adjoint $u^*\in\Hom(E)$ with respect to this basis; by definition we have \begin{eqnarray*} a_{jk}^* &=&e_j^*(u^*(e_k)) =\la e_j^{*\sharp},u^*(e_k)\ra =\e_j\la e_j,u^*(e_k)\ra\\ &=&\e_j\la u(e_j),e_k\ra =\e_j\sum_l a_{lj}\la e_l,e_k\ra =\e_j\e_k a_{kj}~. \end{eqnarray*} Putting $D\colon=diag\{\e_1,\ldots,\e_n\}$ we conclude that the matrix $A^*$ of $u^*$ with respect to the orthonormal basis $e_1,\ldots,e_n$ is given by: \begin{equation}\label{ipseq8}\tag{IPS8} A^*=DA^tD~. \end{equation} Remember from linear algebra that multiplying a matrix $A$ from the left (right) by a diagonal matrix $D$ multiplies the $j$-th row (column) of $A$ with $\e_j$.
$A=(a_{jk})$ is the matrix of a self-adjoint operator on $\R_1^{n+1}$ (with respect to the canonical basis $e_0,e_1,\ldots,e_n$), if and only if for all $k\geq1$: $a_{0k}=-a_{k0}$ and for all $j,k\geq1$: $a_{jk}=a_{kj}$.
Describe all skew-symmetric linear operators $u:\R_1^2\rar\R_1^2$ and $u:\R_1^3\rar\R_1^3$.
Let $(E,\la.,.\ra)$ be an inner product space. Determine for all $a,b\in E$ the adjoint of the linear map $u:x\mapsto\la x,a\ra b$.
For all $x,y\in E$ we have $$ \la u(x),y\ra =\la x,a\ra\la b,y\ra =\la x,\la y,b\ra a\ra, \quad\mbox{i.e.}\quad u^*(x)=\la x,b\ra a~. $$
If $A$ is a subset of the inner product space $(E,\la.,.\ra)$ and $u\in\Hom(E)$, then$$ u(A)^\perp =(u^*)^{-1}(A^\perp) \colon=\{y\in E: u^*(y)\in A^\perp\} =\colon[u^*\in A^\perp]~. $$ In particular, if $u$ is an isometry, then $u(A)^\perp=u(A^\perp)$.
$y\in E$ is in $u(A)^\perp$ iff for all $x\in A$: $\la u(x),y\ra=0$, which holds if and only if for all $x\in A$: $\la x,u^*(y)\ra=0$, i.e. $u^*(y)\in A^\perp$.
If $(E,\la.,.\ra)$ is an inner product space of index $\nu$, then $(u,v)\mapsto\tr(u^*v)$ is an inner product on $\Hom(E)$ and its index is $2\nu(n-\nu)$. If $E$ is Euclidean, then $(u,v)\mapsto\tr(u^*v)$ is Euclidean and the norm $\Vert u\Vert\colon=\tr(u^*u)^{1/2}$ is called the Hilbert-Schmidt norm (or Frobenius norm) and is usually denoted by $\Vert u\Vert_{HS}$.
Suppose $e_1,\ldots,e_n$ is an orthonormal basis with $\la e_j,e_j\ra=\e_j=\pm1$ and $(u_{jk})$ and $(v_{jk})$ the matrices of $u$ and $v$, respectively. $$ \tr(u^*v) =\sum_l\e_l\la u^*ve_l,e_l\ra =\sum_l\e_l\la ve_l,ue_l\ra =\sum_{l,m,k}\e_l\la v_{ml}e_m,u_{kl}e_k\ra =\sum_{l,m}\e_l\e_mv_{ml}u_{ml}~. $$ It follows easily that the mappings $E^{jk}x\colon=e_k^*(x)e_j$ form an orthonormal basis: the matrix of $E^{jk}$ with respect to the basis $e_1,\ldots,e_n$ has exactly one non-vanishing entry: the entry in its $j$-th row and $k$-th column is $1$. Moreover $\tr(E^{jk*}E^{jk})=\e_j\e_k$, which is $-1$ for exactly $2\nu(n-\nu)$ pairs of indices $j,k$.
For any real $2\times2$ matrix $X\in\Ma(2,\R)$ satisfying $\det X=1$ the mapping $u:\Ma(2,\R)\rar\Ma(2,\R)$, $u(A)\colon=XA$ is an isometry, when $\Ma(2,\R)$ carries the inner product $g$ from exam.
If $(E,\la.,.\ra)$ is an inner product space then $\tr(u^*v)=\tr(uv^*)$.

Orthogonal Decompositions

In a Euclidean space $E$ we have for every subspace $F$ an orthogonal decomposition $E=F\oplus F^\perp$. Alas, this is no more true in arbitrary inner product spaces. To hold the subspace $F$ must meet a certain condition.

Non-degenerate subspaces

Let $\la.,.\ra$ be an inner product on $E$. For any subspace $F$ of $E$ we have $\dim F+\dim F^\perp=n=\dim E$ and $F^{\perp\perp}=F$.
$\proof$ Let $e_1,\ldots,e_k$ be a basis for $F$, $b_1,\ldots,b_k$ the canonical basis for $\R^k$ and define $u:E\rar\R^k$ by: $$ u(x)\colon=\sum_{j=1}^k\la x,e_j\ra b_j~. $$ Then we have by \eqref{ipseq7} for the adjoint $u^*:\R^{k}\rar E$: $$ \forall x\in E:\quad \la u^*(b_j),x\ra =\la b_j,u(x)\ra =\la x,e_j\ra, $$ i.e. $u^*(b_j)=e_j$. Hence $u^*$ is one-one and by exam $u$ must be onto. Now $F^\perp=\ker u$, indeed $x\in\ker u$ iff for all $j$: $\la e_j,x\ra=0$, i.e. iff for all $y\in F$: $\la y,x\ra=0$. We therefore conclude by the Rank-Nullity Theorem \eqref{lareq5} that: $$ \dim F^\perp=\dim\ker u=\dim E-\dim\im u=n-k~. $$ 2. We clearly have $F\sbe F^{\perp\perp}$ and by 1. $\dim F^{\perp\perp}=\dim F$. $\eofproof$
This result immediately implies that $E=F\oplus F^\perp$, if and only if $F\cap F^\perp=\{0\}$. However, in general we do not have $F\cap F^\perp=\{0\}$: for example, let $F$ be the subspace of $\R_1^2$ generated by $(1,1)$, then $F^\perp=F$.
Suppose $x\neq0$ and $x\in F\cap F^\perp$, which means that for all $y\in F$: $y\perp x$, but this simply states that the inner product $\la.,.\ra$ restricted to $F\times F$ is singular. Therefore $F\cap F^\perp=\{0\}$ iff $\la.,.\ra|F\times F$ is non-singular.
Let $\la.,.\ra$ be an inner product on $E$. A subspace $F$ of $E$ is called non-degenerated, if $\la.,.\ra|F\times F$ is non-singular, i.e. if and only if $(F,\la.,.\ra)$ is an inner product space as well.
Let $\la.,.\ra$ be an inner product on $E$. If $F$ is a non-degenerated subspace of $E$, then $E=F\oplus F^\perp$. Moreover $F^\perp$ is non-degenerated.
$\proof$ We only need to prove the last statement. $F^\perp$ is not degenerated iff $\{0\}=F^{\perp\perp}\cap F^\perp$ and since $F^{\perp\perp}=F$, $F^\perp$ is non-degenerated iff $F$ is non-degenerated. $\eofproof$
If a subspace $F$ of an inner product space $(E,\la.,.\ra)$ is non-degenerated, then for all bases $b_1,\ldots,b_m$ for $F$ the Gramian $(\la b_j,b_k\ra)\in\Ma(m,\R)$ is non-singular. If for some basis $b_1,\ldots,b_m$ for $F$ the Gramian $(\la b_j,b_k\ra)\in\Ma(m,\R)$ is non-singular, then $F$ is non-degenerated.
If a subspace $F$ of $E$ contains a vector $x\neq0$ orthogonal to itself, then $F$ need not be degenerated. If $\dim F=1$ then $F$ is degenerated iff for some $x\in F\sm\{0\}$: $\la x,x\ra=0$.
Suppose $(E,\la.,.\ra)$ is an inner product space and $u\in\Hom(E)$, such that $\im u$ is non-degenerated. Then there is an orthogonal decomposition: $E=\im u\oplus\ker u^*$. Moreover $\im u$ is non-degenerated iff $\ker u^*$ is non-degenerated. Suggested solution.
Verify that the subspace $F$ of $\R_1^3$ generated by $(1,1,0)$ and $(1,1,1)$ is degenerated.
The subspace $S$ of all self-adjoint operators $u\in\Hom(E)$ is a non-degenerated subspace of $\Hom(E)$ with the inner product $(u,v)\mapsto\tr(u^*v)$? Cf. exam.
Let $A$ be the subspace of all skew-symmetric operators. Since $u=(u+u^*)/2+(u-u^*)/2$ and $u+u^*\in S$, $u-u^*\in A$, we have $\Hom(E)=S+A$ and since $S\cap A=\{0\}$: $\Hom(E)=S\oplus A$. Moreover, for all $u\in S$ and all $v\in A$: $$ \tr(u^*v)=\tr(uv)=\tr((uv)^*)=\tr(v^*u^*)=-\tr(vu^*)=-\tr(u^*v) $$ i.e. $\tr(u^*v)=0$. Hence $A\sbe S^\perp$ and since $\dim S^\perp=\dim\Hom(E)-\dim S=\dim A$, we conclude that $A=S^\perp$ and $\Hom(E)=S\oplus S^\perp$, i.e. $S$ is non-degenerated.
Given the inner product $g$ on $\Ma(2,\R)$ (cf. exam). Show that $\{A\in\Ma(2,\R):\tr A=0\}$ is a non-degenertated subspace and compute its index.

Orthogonal projection and orthonormalization

A projection $P$ in a vector-space $E$ is a linear map $P:E\rar E$ such that $P^2=P$. Putting $F\colon=\im P$ and $G\colon=\im(1-P)$ we say that $P$ is a projection onto $F$ with kernel $G$: $$ x\in\ker P\, \Lrar\, Px=0\, \Lrar\, (1-P)x=x\, \Rar\, x\in G\, \Rar\, Px=P(1-P)y=0 \Rar x\in\ker P, $$ i.e. $\ker P=\im(1-P)$. Since for all $x\in E$: $x=Px+(1-P)x$ and $x\in F\cap G$ implies $x=Px=P(1-P)y=0$, we have $E=F\oplus G$ and for all $x\in F$ and all $y\in G$: $Px=x$ and $Py=0$. Conversely if we have a decomposition $E=F\oplus G$, then there is a unique projection $P$ onto $F$ with kernel $G$: just put for all $x\in F$ and all $y\in G$: $Px=x$ and $Py=0$.
If $P\in\Hom(E)$ is a projection onto $F$ with kernel $G$, then $\tr P=\dim F$.
If $P\in\Hom(E)$ is a projection onto $F$ with kernel $G$, then $P^*\in\Hom(E^*)$ is a projection onto $\{x^*\in E^*: x^*|G=0\}$ with kernel $\{x^*\in E^*: x^*|F=0\}$. Suggested solution.
time cone
The picture above describes the projections of a vector $x\in\R_1^2$ on the subspaces generated by $b_0$ (with kernel $\lhull{b_1}$) and $b_1$ (with kernel $\lhull{b_0}$), respectively.
In an arbitary vector-space $E$ for a subspace $F$ there is no 'canonical' candidate for a complementary space $G$, i.e. a subspace $G$ of $E$ such that $E=F\oplus G$. However in an inner product space there is such a space in case $F$ is non-degenerated: the space $F^\perp$.
Let $(E,\la.,.\ra)$ be an inner product space, $F$ a non-degenerated subspace of $E$, then there is a unique projection $\Prn_F\in\Hom(E)$ onto $F$ with kernel $F^\perp$, i.e. $\Prn_F|F=id$ and $\Prn_F|F^\perp=0$. Moreover $\Prn_F$ is self-adjoint and $\Prn_{F^\perp}=1-\Prn_F$. $\Prn_F$ is called the orthogonal projection on $F$.
$\proof$ Since $\Prn_F|F=id_F$ and $\Prn_F|F^\perp=0$, $\Prn_F$ is unique for $F$ is non-degenerated and thus $E=F\oplus F^\perp$. $\Prn_F$ is a projection: for every $x\in E$ we have a unique decomposition $x=y+z$, $y\in F$, $z\in F^\perp$ and thus $\Prn_F(x)=y=\Prn_F(y)=\Prn_F^2(x)$. Finally, for $x=a+b$, $y=c+d$ such that $a,c\in F$ and $b,d\in F^\perp$ we get: $$ \la\Prn_Fx,y\ra-\la x,\Prn_Fy\ra =\la a,y\ra-\la x,c\ra =\la a,c\ra-\la a,c\ra =0, $$ i.e. $\Prn_F$ is self-adjoint. Finally $(1-\Prn_F)|F=0$ and $(1-\Prn_F)|F^\perp=id_{F^\perp}$, i.e. $1-\Prn_F=\Prn_{F^\perp}$. $\eofproof$
The mapping $P:u\mapsto(u+u^*)/2$ is the orthogonal projection onto the subspace $S$ of all self-adjoint operators $u\in\Hom(E)$. Cf. exam.
If $u\in S$, then $u^*=u$ and thus $Pu=u$. If $u\in S^\perp$, then by exam: $u^*=-u$ and therefore $Pu=0$.
If $F$ is degenerated, then there is no map $P:E\rar F$ such that $P|F=id$ and $P|F^\perp=0$.
Suppose $(E,\la.,.\ra)$ is an inner product space and $P\in\Hom(E)$ is a projection onto a non-degenerate subspace $F$, i.e. $P^2=P$ and $\im P=F$. Then $P$ is the orthogonal projection onto $F$, if and only if $P^*=P$. Suggested solution.
If $b_1,\ldots,b_n$ is an arbitrary basis for $F$, then $\Prn_F$ is given by $$ \Prn_F(x)=\sum_{j,k=1}^n g^{jk}\la x,b_j\ra b_k~. $$ where $(g^{jk})$ is the inverse of the Gramian matrix $g_{jk}\colon=\la b_j,b_k\ra$. In particular if $e_1,\ldots,e_n$ is an orthonormal basis for $F$ with $\e_j=\la e_j,e_j\ra=\pm1$, then $$ \Prn_F(x)=\sum_{j=1}^n\e_j\la x,e_j\ra e_j~. $$
Indeed, define a linear mapping $P\in\Hom(E)$ by $Px\colon=\sum_{j,k=1}^n g^{jk}\la x,b_j\ra b_k$. In order to verify that $P$ is the orthogonal projection onto $F$, we need to establish $Pb_l=b_l$ and for $x\in F^\perp$: $Px=0$. For all $l\leq n$: $$ Pb_l =\sum_{j,k}g^{jk}\la b_l,b_j\ra b_k =\sum_{j,k}g^{jk}g_{lj}b_k =\sum_{k}\d_{kl}b_k =b_l~. $$ For $x\in F^\perp$ we have by definition of $F^\perp$ for all $j$: $\la x,b_j\ra=0$ and thus $Px=0$.
Suppose $F=\lhull{(1,1,1),(0,1,1)}$. Compute the orthogonal projection onto $F$ in $\R_1^3$.
Suppose $F$ is a non-degenerated subspace of $F$. Then the linear map $$ R_F:x\mapsto2\Prn_F(x)-x=x-2\Prn_{F^\perp}(x) $$ is called the reflection about the subspace $F$. Verify that for all $x\in F$: $R_F(x)=x$, for all $x\in F^\perp$: $R_F(x)=-x$ and $R_F^*=R_F$.
In a two dimensional Euclidean space every linear isometry is the composition of two reflections at most.
In an $n$-dimensional Euclidean space every linear isometry is the composition of $n$ reflections at most. This is a special case of the so called Dieudonné-Cartan Theorem. Hint: Utilize the result in dimension $n=2$ and apply the spectral theorem for orthogonal operators (in real spaces) in the general case.

Orthogonal projections and critical points

In the Euclidean case the orthogonal projection $\Prn_F(x)$ is the unique point in $F$ which minimizes the functional $f(y)\colon=\norm{x-y}^2=\la x-y,x-y\ra$, i.e. $$ \norm{x-\Prn_F(x)}^2=\inf\{\norm{x-y}^2:y\in F\}~. $$ In the general case of an inner product there is no minimizer even if $F$ is non-degenerated, so let's look for critical points of the functional $f(y)=\la y-x,y-x\ra$ subjected to $y\in F$.
For any inner product $\la.,.\ra$ on $E$ the derivative of $f:E\rar\R$, $f(y)=\la y-x,y-x\ra$ at $y\in E$ in the direction of $u\in E$ is given by $df(y)u=2\la y-x,u\ra$.
$\proof$ Putting $z=y-x$ the directional derivative is given by definition by $$ df(y)u \colon=\lim_{t\to0}\frac{f(y+tu)-f(y)}{t} =\lim_{t\to0}\frac{\la z+tu,z+tu\ra-\la z,z\ra}{t} =2\la z,u\ra =2\la y-x,u\ra~. $$ $\eofproof$
Suppose $e_1,\ldots,e_k$ is an orthonormal basis for the non-degenerated subspace $F^\perp$; employing Lagrange multipliers we have to find critical points of the functional $$ y\mapsto\la y-x,y-x\ra-2\sum_j\l_j\la y,e_j\ra,\quad y\in E~. $$ By
lemma this gives us the equations: $$ \forall u\in E:\, \la y-x,u\ra=\sum_j\l_j\la u,e_j\ra, \quad\mbox{and}\quad \forall j:\, \la y,e_j\ra=0~. $$ Thus $y-x=\sum\l_je_j\in F^\perp$ and $y\in F$; since $E=F\oplus F^\perp$ we have $$ (x-y)+y=x=\Prn_{F^\perp}(x)+\Prn_{F}(x) $$ and thus: $y=\Prn_F(x)$ and $x-y=\Prn_{F^\perp}(x)$.
Suppose $F$ is a non-degenerated subspace of the inner product space $E$. Then $\Prn_F(x)$ is the unique critical point of the functional $f:F\rar\R$, $$ f(y)=\la y-x,y-x\ra~. $$
This is a general pattern: minimizers in the Euclidean case will often turn into critical points for inner product spaces.

Gram-Schmidt algorithm

For the Gram-Schmidt orthonormalization procedure we start with an orthonormal set of vectors $e_1,\ldots,e_m$, put $F=\lhull{e_1,\ldots,e_m}$ and take any vector $x$ not in $F$. The vector $$ z\colon=x-\Prn_F(x) =x-\sum_{j=1}^m\e_j\la x,e_j\ra e_j, \quad \e_j\colon=\la e_j,e_j\ra=\pm1, $$ is by definition orthogonal to $F$. Now in the non-Euclidean case $\la z,z\ra$ may vanish - that's the catch of the algorithm! If this does not occur, then the vectors $e_1,\ldots,e_m,z/|\la z,z\ra|^{1/2}$ form an orthonormal basis for the subspace generated by $F$ and $x$. In order to find an orthonormal basis for $E$ starting with some basis vectors $b_1,\ldots,b_n$ we normalize $b_1$ and then proceed as in the Euclidean case. Provided none of the obtained vectors $z$ is orthogonal to itself, the algorithm terminates successfully!
Suppose $b_1,\ldots,b_n$ is a basis for $E$ such that none of the spaces $E_k\colon=\lhull{b_1,\ldots,b_k}$, $k=1,\ldots,n$, is degenerated. Then the Gram-Schmidt algorithm terminates successfully. Suggested solution.
The Gram-Schmidt orthonormalization procedure can not be applied to the basis $(1,1),(1,-1)$ in $\R_1^2$.
Apply the Gram-Schmidt orthonormalization procedure to the basis $(1,1,1),(0,1,1),(1,1,0)$ in $\R_1^3$.
Write a computer program that determines a set of orthonormalized vectors in $\R_\nu^n$ given $m$ linear independent vectors $x_1,\ldots,x_m$, $m\leq n$. Suggested solution

Diagonalization

In the Euclidean setting the spectral theorem for self-adjoint operators $u$ states that $u$ is diagonalizable and there exists an orthonormal basis of eigen vectors. This no longer holds in case of an arbitrary inner product space!
$e_0=(1,1,0)$, $e_1=(0,0,1)$, $e_2=(-1,0,0)$ is a basis of $\R_1^3$; define $u\in\Hom(\R_1^3)$ by $u(e_0)=e_0$, $u(e_1)=e_1+e_0$ and $u(e_2)=e_2+e_1$. Then $u$ is self-adjoint but not diagonalizable.
Since $e_0,e_1\perp e_0$, $e_2\perp e_1$ and $\la e_0,e_2\ra=\la e_1,e_1\ra=\la e_2,e_2\ra=1$, we get \begin{eqnarray*} \la u(e_0),e_1\ra=\la e_0,e_1\ra=0&=&\la e_1+e_0,e_0\ra=\la u(e_1),e_0\ra\\ \la u(e_0),e_2\ra=\la e_0,e_2\ra=1&=&\la e_2+e_1,e_2\ra=\la u(e_2),e_0\ra\\ \la u(e_1),e_2\ra=\la e_1+e_0,e_2\ra=1&=&\la e_2+e_1,e_1\ra=\la u(e_2),e_1\ra~. \end{eqnarray*} Thus $u$ is self-adjoint and the matrix of $u$ with respect to the basis $e_0,e_1,e_2$ is $$ \left(\begin{array}{ccc} 1&1&0\\ 0&1&1\\ 0&0&1 \end{array}\right) $$ which is a Jordan-matrix. Hence $u$ is not diagonalizable. Even in case a self-adjoint operator is diagonalizable its eigen values need not be real:
The linear map $u(e_0)=e_0-e_1$, $u(e_1)=e_0+e_1$ on $\R_1^2$ is self-adjoint and its eigen values are not real.
Suppose $E$ is an inner product space and $u:E\rar E$ is self-adjoint.
  1. If $F$ is an invariant subspace, i.e. $u(F)\sbe F$, the so is $F^\perp$.
  2. If $x$ and $y$ are eigen vectors with different real eigen values, then $x\perp y$.
Beware! If $E$ is a real vector-space, $u\in\Hom(E)$ and $\l\in\C$ is a root of the characteristic polynomial $c_u$, the corresponding eigen vector need not be in $E$, it's a vector in the complexification $\C\otimes E$ of $E$.

Volume Forms and Orientation

Volume forms

Suppose $E$ is an $n$-dimensional vector-space; an (alternating) $n$-form $\o$ on $E$ is an $n$-linear alternating map $\o:E^n\rar\R$, i.e. it is linear in all of its $n$ components and for all permutations $\pi\in S_n$ and all $x_1,\ldots,x_n\in E$ we have: $$ \o(x_{\pi(1)},\ldots,x_{\pi(n)})=\sign(\pi)\o(x_1,\ldots,x_n)~. $$ Objects of this type will be explored in greater generality in section! The simplest non-trivial case is an alternating $2$-form on a $2$-dimensional space $E$: Let $e_1,e_2$ be any basis for $E$ with dual basis $e_1^*,e_2^*$, then $$ \o:(x,y)\mapsto e_1^*(x)e_2^*(y)-e_2^*(x)e_1^*(y) $$ is an alternating $2$-form and $\o(e_1,e_2)=1$.
1. Interchanging any two components of an alternating $n$-form $\o$ changes the sign, i.e. $$ \o(x_1,\ldots,x_j,\ldots,x_k,\ldots,x_n) =-\o(x_1,\ldots,x_k,\ldots,x_j,\ldots,x_n)~. $$ Thus for every alternating $n$-form $\o$ on the $n$-dimensional vector-space $E$ we have: $\o(x_1,\ldots,x_n)=0$ if for some $j\neq k$: $x_j=x_k$.
If $x_1,\ldots,x_n\in E$ are linearly dependent, then for every alternating $n$-form $\o$ on the $n$-dimensional vector-space $E$: $\o(x_1,\ldots,x_n)=0$.
Assume w.l.o.g. that $x_n=\l_1x_1+\cdots\l_{n-1}x_{n-1}$, then by linearity of $\o$ in the last slot and exam: $$ \o(x_1,\ldots,x_n)=\sum_j\l_j\o(x_1,\ldots,x_{n-1},x_j)=0~. $$
Suppose for $j=1,\ldots,n$ we have $e_1,\ldots,e_n\in E$ and $x_j\colon=\sum_{k=j}^n a_{kj}e_k\in E$ - or in matrix notation: $$ (x_1,x_2,x_3,\ldots,x_n) =(e_1,e_2,e_3,\ldots,e_n) \cdot \left(\begin{array}{cccc} a_{11}&0&0&\cdots&0\\ a_{21}&a_{22}&0&\cdots&0\\ a_{31}&a_{32}&a_{33}&\cdots&0\\ \vdots&\vdots&\vdots&\ddots&\vdots\\ a_{n1}&a_{n2}&a_{n3}&\ldots&a_{nn} \end{array}\right)~. $$ Then for every alternating $n$-form $\o$ on the $n$-dimensional vector-space $E$: $\o(x_1,\ldots,x_n)=a_{11}\ldots a_{nn}\o(e_1,\ldots,e_n)$. Suggested solution.
If $e_1,\ldots,e_n$ is a basis for $E$ with dual basis $e_1^*,\ldots,e_n^*$, then for any alternating $n$-form $\o$ and all $x_1,\ldots,x_n\in E$: $$ \o(x_1,\ldots,x_n) =\sum_{\pi\in S_n}\sign(\pi)\Big(\prod_{k=1}^n e_{\pi(k)}^*(x_k)\Big)\,\o(e_1,\ldots,e_n) =\det(e_j^*(x_k))\,\o(e_1,\ldots,e_n), $$ where $\det(a_{jk})$ denotes the determinant of the matrix $(a_{jk})$. Thus the value $\o(e_1,\ldots,e_n)$ uniquely determines $\o$. In other words, the space of all $n$-forms on an $n$-dimensional space $E$ is one dimensional.
$\proof$ For every $j$ we have $x_j=\sum_k a_{kj}e_k$, $a_{kj}=e_k^*(x_j)$, and thus by multi-linearity and the previous exams: \begin{eqnarray*} \o(x_1,\ldots,x_n) &=&\sum_{1\leq k_1,\ldots,k_n\leq n}\o(a_{k_11}e_{k_1},\ldots,a_{k_nn}e_{k_n})\\ &=&\sum_{\pi\in S_n}\o(a_{\pi(1)1}e_{\pi(1)},\ldots,a_{\pi(n)n}e_{\pi(n)})\\ &=&\sum_{\pi\in S_n}\sign(\pi)a_{\pi(1)1}\cdots a_{\pi(n)n}\o(e_1,\ldots,e_n)~; \end{eqnarray*} $\eofproof$
Hence every basis $e_1,\ldots,e_n$ of $E$ determines a unique alternating $n$-form $\vol{}:E^n\rar\R$ such that $\vol{}(e_1,\ldots,e_n)=1$ - it's called the volume form
associated with the basis $e_1,\ldots,e_n$ and it's given by \begin{equation}\label{vfoeq1}\tag{VFO1} \vol{}(x_1,\ldots,x_n) =\sum_{\pi\in S_n}\sign(\pi)\Big(\prod_{k=1}^n e_{\pi(k)}^*(x_k)\Big) =\det(e_j^*(x_k))~. \end{equation} As a by-product the formula for the determinant of a matrix pops up quite naturally. Now suppose $\o$ is another alternating $n$-form, then $\o$ is uniquely determined by the value $c\colon=\o(e_1,\ldots,e_n)$ and $\o=c\vol{}$. Therefore we have
An $n$-form is either a volume-form or trivial i.e. zero and it is a volume-form iff for some and hence for all bases $x_1,\ldots,x_n$ of $E$: $\o(x_1,\ldots,x_n)\neq0$.
If $\o$ is a volume form on the $n$-dimensional vector-space $E$ and $\o(x_1,\ldots,x_n)=0$, then the vectors $x_1,\ldots,x_n$ are linearly dependent.
Suppose $x_1,\ldots,x_n$ are linearly independent, then it's a basis and thus $\o(x_1,\ldots,x_n)\neq0$.

The determinant of a linear mapping

Given a volume-form $\vol{}$ on a real vector-space $E$ of dimension $n$ and a linear map $u\in\Hom(E)$ we define $$ \forall x_1,\ldots,x_n\in E:\quad \eta(x_1,\ldots,x_n) \colon=\vol{}(u(x_1),\ldots,u(x_n))~. $$ Then $\eta$ is another alternating $n$-form - it's the pull-back of the form $\vol{}$ (cf.
subsetion). By lemma there is some constant $\det(u)\in\R$ such that $\eta=\det(u)\vol{}$, i.e. \begin{equation}\label{vfoeq2}\tag{VFO2} \forall x_1,\ldots,x_n\in E:\quad \vol{}(u(x_1),\ldots,u(x_n)) =\det(u)\vol{}(x_1,\ldots,x_n)~. \end{equation} The constant $\det(u)$ is called the determinant of $u$. This definition is actually independent of the volume form, for suppose $\o$ is another volume form, then for some constant $c\neq0$: $\o=c\vol{}$ and thus $$ \o(u(x_1),\ldots,u(x_n)) =c\vol{}(u(x_1),\ldots,u(x_n)) =c\det(u)\vol{}(x_1,\ldots,x_n) =\det(u)\o(x_1,\ldots,x_n)~. $$ If $e_1,\ldots,e_n$ is a basis such that $\vol{}(e_1,\ldots,e_n)=1$, then by \eqref{vfoeq1} $$ \det(u)\colon=\vol{}(u(e_1),\ldots,u(e_n))=\det(e_j^*(u(e_k))) $$ where $\det$ on the right is the determinant of the matrix $(e_j^*(u(e_k)))$, i.e. the matrix of $u$ with respect to the basis $e_1,\ldots,e_n$.
Suppose $e_1,\ldots,e_n$ and $b_1,\ldots,b_n$ are bases for $E$. Then for all $u\in\Hom(E)$: $$ \det(b_j^*(u(b_k)))=\det(e_j^*(u(e_k))) $$ i.e. the determinant of the matrix of $u$ with respect to a basis does not depend on the basis. Suggested solution.
Finally we verify another important property of the map $\det:\Hom(E)\rar\R$: it's multiplicative, i.e. $$ \forall u,v\in\Hom(E):\quad\det(uv)=\det(u)\det(v)~. $$ This property is almost obvious! Indeed, let $u,v:E\rar E$ be linear mappings, then \eqref{vfoeq2} implies for $x_j\colon=v(e_j)$ - we still assume $\vol{}(e_1,\ldots,e_n)=1$: $$ \det(uv) =\vol{}(uv(e_1),\ldots,uv(e_n)) =\det(u)\vol{}(v(e_1),\ldots,v(e_n)) =\det(u)\det(v)~. $$ By proposition or exam $\det u=0$ is equivalent to the fact that for any basis $e_1,\ldots,e_n$ of $E$ the vectors $u(e_1),\ldots,u(e_n)$ are linearly dependent, i.e. $u$ is not one-one and since $E$ is of finite dimension this tantamounts to $u$ is not an isomorphism. Take for example the characteristic polynomial $c_u$ of $u\in\Hom(E)$, it's defined by $$ c_u(t)\colon=\det(u-t) $$ and it vanishes iff $u-t$ is not one-one, i.e. iff $t$ is an eigen value of $u$. \begin{eqnarray*} c_u(t) &=&\det(u-t) =(-1)^n\det(t-u)\\ &=&(-1)^n(t^n-a_1(u)t^{n-1}+-\cdots+(-1)^na_n(u))\\ &=&(-1)^n(t-\l_1)\cdots(t-\l_n), \end{eqnarray*} where $\l_1,\ldots,\l_n\in\C$ are the roots of the characteristic polynomial, multiplicities accounted. Thus we get: $$ a_n(u)=\det u=\l_1\cdots\l_n~. $$ Moreover, since for all isomorphisms $v\in\Hom(E)$: $$ \det(u-t) =\det(v(u-t)v^{-1}) =\det(v^{-1}uv-t), $$ we conclude that for all $j=1,\ldots,n$: $$ a_j(v^{-1}uv)=a_j(u)~. $$
If $E=F\oplus G$ and $u\in\Hom(E)$ satisfies: $u(F)\sbe F$ and $u(G)\sbe G$, then $\det u=\det(u|F)\det(u|G)$. Suggested solution.
If $E$ has a basis such that the matrix $A$ of $u\in\Hom(E)$ is triangular (lower or upper, this doesn't matter), then $\det u$ is the product of all diagonal entries in $A$.
The essential properties of the determinant in a nutshell:
  1. The determinant of a linear mapping $u\in\Hom(E)$ does not depend on the volume form and by Vieta's formulas it's the product of all possibly complex eigen values of $u$, i.e. the product of all $n\colon=\dim E$ zeros of its characteristic polynomial.
  2. $u\in\Hom(E)$ is an isomorphism if and only if $\det(u)\neq0$ or: $\vol{}(x_1,\ldots,x_n)\neq0$ if and only if the vectors $x_1,\ldots,x_n$ are linearly independent.
  3. $\det:\Hom(E)\rar\R$ is multiplicative, i.e. for all $u,v\in\Hom(E)$: $\det(uv)=\det(u)\det(v)$ and $\det(u^{-1})=1/\det u$ if $u$ is an isomorphism.
  4. If $e_1,\ldots,e_n$ is a basis for $E$ with dual basis $e_1^*,\ldots,e_n^*$, then the volume form on $E$ associated with the basis $e_1,\ldots,e_n$ is given by $$ \o:(x_1,\ldots,x_n)\mapsto\det(e_j^*(x_k)), $$ where $\det$ on the right hand side is the determinante of an $n$ by $n$ matrix. Putting for $u\in\Hom(E)$: $A=(a_{jk})=(e_j^*(u(e_k)))$ we have: $$ \det u =\det(e_j^*(u(e_k))) =\det A \colon=\sum_{\pi\in S_n}\sign(\pi)a_{\pi(1)1}\cdots a_{\pi(n)n}~. $$
One of the few applications of the last formula is the following:
Verify that for any $n\times n$ matrix $A$ we have: $\det A^t=\det A$. Suggested solution.
If $c_u(t)=\det(u-t)=(-1)^n(t^n-a_1(u)t^{n-1}+-\cdots+(-1)^na_n(u))$ is the characteristic polynomial for $u\in\Hom(E)$, then $a_1(u)=\tr u$. Thus $\tr u$ is the sum of all $n$ possibly complex roots of the characteristic polynomial.
Verify that for all $u\in\Hom(E)$ and all isomorphisms $v\in\Hom(E)$: $$ \lim_{t\to0}\frac{\det(1+tu)-1}t=\tr u \quad\mbox{and}\quad \lim_{t\to0}\frac{\det(v+tu)-\det(v)}t=\det(v)\tr(v^{-1}u)~. $$ For arbitrary matrices $A,B$ we have $$ \lim_{t\to0}\frac{\det(A+tB)-\det(A)}t=\tr(A^{ad}B) $$ where $A^{ad}$ denotes the adjugate matrix of $A$.

Orientation

Every volume form $\vol{}$ on an $n$-dimensional vector-space determines an orientation of a basis $e_1,\ldots,e_n$ as follows: $e_1,\ldots,e_n$ is said to be positive (oriented) if $\vol{}(e_1,\ldots,e_n) > 0$, otherwise it's said to be negatively oriented. Accordingly an isomorphism $u\in\Hom(E)$ is said to be orientation preserving if for some basis $e_1,\ldots,e_n$ for $E$ the basis $u(e_1),\ldots,u(e_n)$ has the same orientation as the basis $e_1,\ldots,e_n$ and by the definition of the determinant this is equivalent to $\det u > 0$, which in turn means that for every basis $e_1,\ldots,e_n$ for $E$ the basis $u(e_1),\ldots,u(e_n)$ has the same orientation as the basis $e_1,\ldots,e_n$.
Conversely, every basis $e_1,\ldots,e_n$ determines a unique volume form $\vol{}$ satisfying $\vol{}(e_1,\ldots,e_n)=+1$. Hence we could also say that every basis determines an orientation!
If the basis $e_1,\ldots,e_n$ is positive with respect to the volume form $\vol{}$, then for any permutation $\pi\in S_n$ the basis $e_{\pi(1)},\ldots,e_{\pi(n)}$ is positive iff $\pi$ is even, and it's negative iff $\pi$ is odd.
Finally the so called matrix determinant lemma:
Let $A\in\Ma(m+n,\R)$ be the block matrix $$ A=\left( \begin{array}{cc} A_{11}&A_{12}\\ A_{21}&A_{22} \end{array} \right)~. $$ If $A_{11}\in\Ma(m,\R)$ is invertible then (suggested solution): $$ \det A=\det(A_{11})\det(A_{22}-A_{21}A_{11}^{-1}A_{12})~. $$
Special cases: 1. If $m=n$ and $A_{21}A_{11}=A_{11}A_{21}$, then $$ \det A=\det(A_{11}A_{22}-A_{21}A_{12})~. $$ 2. If $A_{12}=0$ or $A_{21}=0$, then $\det A=\det A_{11}\det A_{22}$. This generalizes exam.

Volume forms and subspaces

Assume we are given a volume form $\vol{}$ on $E$ and functional $x^*\in E^*\sm\{0\}$. Is there some sort of restriction of $\vol{}$ to the subspace $F\colon=[x^*=0]$. Plain restriction is not possible, because any $n$-form restricted to a subspace of dimension less than $n$ vanishes. So we have to come up with a different construction: Choose $x_1\in E$ such that $x^*(x_1)=1$ and put $$ \forall x_2,\ldots,x_n \vol F(x_2,\ldots,x_n)\colon=\vol{}(x_1,x_2,\ldots,x_n)~. $$ This is obviously a volume form on $F$. Moreover, it doesn't depend on $x_1$; for if $x^*(x)=1$, then $x_1-x\in F$ and thus $$ \vol{}(x_1,x_2,\ldots,x_n)-\vol{}(x,x_2,\ldots,x_n) =\vol{}(x_1-x,x_2,\ldots,x_n) =0~. $$ $\vol F$ is commonly denoted by $x_1\contract\vol{}$ and it is called the contraction of $x^*$ and $\vol F$. By induction we may extend this construction to subspaces $F$ of arbitrary codimension. We won't do that, we rather turn to the converse construction: given a subspace $F=[x^*=0]$ of codimension $1$ and a volume form $\vol F$ on $F$ can we find a volume form $\vol{}$ on $E$ such that $\vol F(x_2,\ldots,x_n)\colon=\vol{}(x_1,x_2,\ldots,x_n)$ for some $x_1\in[x^*=1]$? Yes, this can be achieved and the model is Laplace's expansion of determinants: First we extend $\vol F$ to an $(n-1)$-form on $E$ by choosing $x_1\in[x^*=1]$ and putting $$ \forall x_2,\ldots,x_n\in E:\quad \vol F(x_2,\ldots,x_n)\colon=\vol F(x_2-x^*(x_2)x_1,\ldots,x_n-x^*(x_n)x_1)~. $$ Finally we put $$ \forall x_1,\ldots,x_n\in E:\quad \vol{}(x_1,\ldots,x_n) \colon=\sum_{l=1}^n(-1)^{l-1}x^*(x_j)\vol F(x_1,\ldots,x_{l-1},x_{l+1},\ldots,x_n)~. $$ This is indeed an $n$-form on $E$ and for $x_2,\ldots,x_n\in F$ and $x^*(x_1)=1$ we get: $\vol{}(x_1,\ldots,x_n)=\vol F(x_2,\ldots,x_n)$. Usually this volume form on $E$ is denoted by $x^*\wedge\vol F$ and called the exterior or wedge product of $x^*$ and $\vol F$, cf. exterior product.

Volume forms on inner product spaces

Let $(E,\la.,.\ra)$ be an inner product space, $e_1,\ldots,e_n$ an orthonormal basis and $\e_j=\la e_j,e_j\ra$. Then the matrix $(a_{jk})$ of $u$ with respect to the basis $e_1,\ldots,e_n$ is by \eqref{ipseq4}: $$ a_{jk} =e_j^*(u(e_k)) =\la e_j^{*\sharp},u(e_k)\ra =\e_j\la e_j,u(e_k)\ra $$ and thus: \begin{equation}\label{vfoeq3}\tag{VFO3} \det(u)=\e\det(\la e_j,u(e_k)\ra) \end{equation} where $\e=\e_1\cdots\e_n=\det D$ and $D=diag\{\e_1,\ldots,\e_n\}$.
Let $(E,\la.,.\ra)$ be an $n$-dimensional inner product space and $x\in E$, $\la x,x\ra=\e=\pm1$. The linear mapping $R:u\mapsto u-2\e x\la x,u\ra$ is the reflection about the subspace orthogonal to $x$. Compute $\det R$ and $\tr R$.
Choose a basis $f_1,\ldots,f_{n-1}$ for the non-degenerated space $F\colon=x^\perp$. Then $f_1,\ldots,f_{n-1},x$ is a basis for $E$ and $R(f_j)=f_j$ and $R(x)=-x$. Hence $\det R=-1$ and $\tr R=n-2$.
For $u\in\Hom(E)$ we have: $\det(u^*)=\det(u)$. 2. If $u$ is a linear isometry, then $\det(u)=\pm1$. 3. If $u$ is a skew-symmetric isomorphism, then $\dim E$ is even -so there is no skew-symmetric isomorphism on a space $E$ of odd dimension.
$\proof$ By subsection we know that the matrix of $u^*$ with respect to the dual basis $e_1^*,\ldots,e_n^*$ is the transposed matrix $A^t$ of the matrix $A$ of $u$ with respect to the basis $e_1,\ldots,e_n$. Since $\det(A)=\det(A^t)$ (cf exam), we conclude that $$ \det(u^*)=\det A^t=\det A=\det(u)~. $$ 2. If $u$ is a linear isometry, then $u^{-1}=u^*$ and thus by 1. and multiplicativity of $\det$: $$ \det(u)^2=\det(u)\det(u^*)=\det(uu^*)=\det(1)=1~. $$ 3. This follows from 1. and $\det(-u)=(-1)^n\det u$. $\eofproof$
Suppose $e_1,\ldots,e_n$ is an orthonormal basis for $E$ and $\vol{}$ the associated volume form, i.e. $\vol{}(e_1,\ldots,e_n)=1$. Then for any orthonormal basis $b_1,\ldots,b_n$ for $E$: $\vol{}(b_1,\ldots,b_n)=\pm1$. We will say that $\vol{}$ is a volume form associated to an ONB, if $\vol{}(e_1,\ldots,e_n)=\pm1$ for some ONB $e_1,\ldots,e_n$ and hence for all ONBs.
For $a,b\in\R$ the mapping $T:u\mapsto au+bu^*$ is a linear mapping on $\Hom(E)$. Cf. exam. Compute the trace and the determinant of this mapping.
Since $au+bu^*=(a+b)(u+u^*)/2+(a-b)(u-u^*)/2$, the mapping $T$ is given by $(a+b)P+(a-b)Q$, where $P$ and $Q$ denote the orthogonal projections onto the subspaces $S$ and $A=S^\perp$ respectively. Since $\dim S=n(n+1)/2$ and $\dim A=n(n-1)/2$ we get $\tr P=\dim S$, $\tr Q=\dim A$, i.e. $$ \tr T=\tfrac12n(n+1)(a+b)+\tfrac12n(n-1)(a-b) $$ and by exam: $$ \det T=(a+b)^{n(n+1)/2}(a-b)^{n(n-1)/2}~. $$
Suppose $E$ is an inner product space, $e_1,\ldots,e_n$ an orthonormal basis for $E$. The matrix $G\colon=(\la u(e_j),u(e_k)\ra)_{j,k=1}^n$ is called the Gramian of $u$. Verify that $\det(u)^2=|\det G|$. If $\vol{}$ is a volume form on $E$ associated to an ONB, then $$ |\det G|=\vol{}(u(e_1),\ldots,u(e_n))^2~. $$
By \eqref{vfoeq3}: $\det G=\det(\la u^*u(e_j),e_k\ra)=\e\det(u^*u)=\e\det(u)^2$, where $\e=\e_1\cdots\e_n$ and $\e_j\colon=\la e_j,e_j\ra=\pm1$. Finally, by definition of the determinant: $$ \det u\vol{}(e_1,\ldots,e_n)=\vol{}(u(e_1),\ldots,u(e_n))~. $$
Let $(E,\la.,.\ra)$ be an inner product space of dimension $n$. Prove that for all $u\in\Hom(E)$ and all $x_1,\ldots,x_n\in E$: $$ \det u\det(\la x_j,x_k\ra)=\det(\la u(x_j),x_k\ra)~. $$
Let $e_1,\ldots,e_n$ be an orthonormal basis, $\e_j=\la e_j,e_j\ra$ and define $v\in\Hom(E)$ by $v(e_j)\colon=x_j$. Then for $\e\colon=\e_1\cdots\e_n$ by \eqref{vfoeq3}: $$ \det(\la v^*uve_j,e_k\ra)=\e\det(v^*uv)=\e\det(v)^2\det u~. $$ On the other hand again by \eqref{vfoeq3}: $$ \det(\la x_j,x_k\ra)=\det(\la v^*ve_j,e_k\ra)=\e\det(v)^2~. $$
If $u$ is a linear orientation preserving isometry on an inner product space of odd dimension, then $1$ is an eigen value of $u$. Suggested solution.
If $u$ is a linear orientation reversing isometry on an inner product space of odd dimension, then $-1$ is an eigen value of $u$.
Given a unit vector $N$ in an inner product space $(E,\la.,.\ra)$ and a volume form $\vol{}$ on $F$ the form $N\contract\vol{}$ defines a volume form on $F$. Conversely every volume form $\vol F$ on $F$ gives rise to a volume form on $E$: $$ \vol E(x_1,\ldots,x_n) \colon=\sum_{j=1}^n(-1)^{j-1}\la x_j,N\ra\vol F(x_1,\ldots,x_{l-1},x_{l+1},\ldots,x_n), $$ which satisfies for all $x_2,\ldots,x_n\in F$: $N\contract\vol E(x_2,\ldots,x_n)=\vol F(x_2,\ldots,x_n)$. Cf. subsection.
Suppose that for some ONB $e_1,\ldots,e_n$ for $E$: $\vol{}(e_1,\ldots,e_n)=1$. If $f_2,\ldots,f_n$ is an ONB for $F$ such that $N,f_2,\ldots,f_n$ is a positive ONB for $E$, then $f_2,\ldots,f_n$ is a positive ONB for $F$ with respect to $\vol F$, i.e. $\vol F(f_2,\ldots,f_n)=1$. Conversely, if $f_2,\ldots,f_n$ is a positive ONB for $F$ with respect to $\vol F$, then $$ \vol E(N,f_2,\ldots,f_n)=\la N,N\ra~. $$ In any case if $\vol{}$ is a volume form on $E$ associated to an ONB, then $N\contract\vol{}$ is a volume form on $F$ associated to an ONB. If $\vol F$ is a volume form on $F$ associated to an ONB, then $\vol E$ is a volume form on $E$ associated to an ONB.
We close with some additional results, which may be useful elsewhere but which we won't refer to.
Suppose $u\in\Hom(E)$, then $\det e^u=e^{\tr u}$.
Prove Liouville's formula in case $u$ is diagonalizable.
If $E$ is an $n$-dimensional Euclidean space then for all $u\in\Hom(E)$ and any orthonormal basis $e_1,\ldots,e_n$: $$ |\det u|\leq\prod_j\norm{u(e_j)}~. $$
Here is a proof based on the geometric-arithmetic mean inequality.
Equality holds in Hadamard's inequality iff $u(e_1),\ldots,u(e_n)$ are pairwise orthogonal. Suggested solution.
If $E$ is an $n$-dimensional Euclidean space then for all $u\in\Hom(E)$ (suggested solution): $$ |\det u|=\inf\Big\{\prod_j\norm{u(e_j)}: e_1,\ldots,e_n\mbox{ is an ONB}\Big\}~. $$
Let $x_1,\ldots,x_n$ be (linearly independent) vectors in the $n$-dimensional inner product space $E$ such that none of the spaces $E_{n-j}\colon=\lhull{x_1,\ldots,x_j}^\perp$ is degenerated. Choose an orthonormal basis $b_1,\ldots,b_n$ for $E$ such that $b_1=x_1/\Vert x_1\Vert$, $b_{j+1}\in E_{n-j}$. Finally, put $\e_j\colon=\la b_j,b_j\ra$. Then for every volume form $\vol{}$: $$ \vol{}(x_1,\ldots,x_n)=\prod_j\e_j\la x_j,b_j\ra\,\vol{}(b_1,\ldots,b_n)~. $$
Let $P_{n-j}$ be the orthogonal projection onto $E_{n-j}$. By exam and exam we conclude that \begin{eqnarray*} \vol{}(x_1,\ldots,x_n) &=&\vol{}(x_1,P_{n-1}(x_2),x_3,\ldots,x_n)\\ &=&\vol{}(x_1,P_{n-1}(x_2),P_{n-2}(x_3),\ldots,x_n)\\ &=&\cdots =\vol{}(x_1,P_{n-1}(x_2),P_{n-2}(x_3),\ldots,P_1(x_n))~. \end{eqnarray*} Since the matrix $(\e_j\la b_j,P_{n-k+1}x_k\ra)$ is lower triangular, the determinant of this matrix is just the product of its diagonal entries (cf. exam or exam). Finally orthogonal projections are self-adjoint and $b_j\in\lhull{x_1,\ldots,x_{j-1}}^\perp=E_{n-j+1}$.
Hadamard's inequality will follow: in the context of proposition put $x_j=u(e_j)$, then by Cauchy-Schwarz: $$ |\det u| =\prod_j|\la u(e_j),b_j\ra| \leq\prod_j\Vert u(e_j)\Vert $$ and equality holds iff $u(e_j)=\l_jb_j$, i.e. $u(e_1),\ldots,u(e_n)$ are pairwise orthogonal.
In a Euclidean space $E$ we may use the Gramian to determine the distance of a point $x\in E$ to a subspace $F$ generated by linearly independet vectors $x_1,\ldots,x_n$: $$ d(x,F)=\sqrt{\frac{G(x,x_1,\ldots,x_n)}{G(x_1,\ldots,x_n)}}, $$ where $G(x_1,\ldots,x_n)\colon=\det(\la x_j,x_k\ra)$. The particular cases $\dim E=3$ and $\dim F=1$ and $\dim F=2$, respectively, are probably familiar from elemetary Euclidean geometry: $$ d(x,F)=\frac{\Vert x\times x_1\Vert}{\Vert x_1\Vert} \quad\mbox{and}\quad d(x,F)=\frac{|\la x,x_1\times x_2\ra|}{\Vert x_1\times x_2\Vert} \quad\mbox{respectively.} $$
W.l.o.g. we may assume that $E=\lhull{x,F}$. Choose a volume form $\vol{}$ on $E$ associated to an ONB, i.e. $\vol{}(e_0,\ldots,e_n)=\pm 1$ for any ONB $e_0,\ldots,e_n$ for $E$. Let $N$ be a unit vector normal to $F$ and denote by $P_F$ the orthogonal projection on $F$, then $x-P_F(x)N=\la x,N\ra N=\pm d(x,F)N$ and thus $$ \vol{}(x,x_1,\ldots,x_n) =\vol{}(x-P_F(x),x_1,\ldots,x_n) =\vol{}(\pm d(x,F)N,x_1,\ldots,x_n) =\pm d(x,F)\vol F(x_1,\ldots,x_n) $$ where $\vol F\colon=N\contract\vol{}$ is a volume form on $F$ associated to an ONB for $F$ by exam. By exam: $G(x,x_1,\ldots,x_n)=\vol{}(x,x_1,\ldots,x_n)^2$ and $G(x_1,\ldots,x_n)=\vol F(x_1,\ldots,x_n)^2$.
Compute the distance of $x=(1,1,1,1)$ to the subspace generated by $x_1=(1,-1,0,0)$ and $x_2=(0,0,-1,1)$ in $\R^4$.

Cramer's rule from a geometric point of view

We finish this section with a geometric interpretation of Cramer's rule: If we choose a volume $\vol{}$ on a Euclidean space $E$ such that for a given orthonormal basis $e_1,\ldots,e_n$ for $E$: $\vol{}(e_1,\ldots,e_n)=1$, then the "Lebesgue" measure $\l$ on $E$ is the unique translation invariant Radon measure satisfying $\l(W)=1$, where $W$ denotes the cube $W\colon=\{\sum t_je_j:\,0\leq t_j\leq1\}$. Moreover for any set of vectors $x_1,\ldots.x_n$: $$ \l\Big(\sum_{j=1}^{n} t_jx_j:\,0\leq t_j\leq1\Big) =|\vol{}(x_1,\ldots,x_n)| =\sqrt{G(x_1,\ldots,x_n)} $$ and for any set of vectors $x_1,\ldots,x_n,x_{n+1}$ (cf.
exam): $$ \l\Big(\sum_{j=1}^{n+1}t_jx_j:\,0\leq t_j\leq1,\sum t_j=1\Big) =\frac1{n!}|\vol{}(x_2-x_1,\ldots,x_{n+1}-x_1)|~. $$
Compute the Lebesgue measure of the set $$ \Big\{\sum_{j=1}^{n} t_jx_j:\,0\leq t_j\leq1\Big\} $$ in Euclidean space $\R^n$, where $x_j=\sum_{k=1}^je_k$ and $e_1,\ldots,e_n$ denotes the canonical ONB.
Suppose we are given a simplex $S\colon=\convex{x_j:j\leq n+1}$ in an $n$-dimensional Euclidean space $E$. For any volume form $\vol{}$ on $E$ we put $$ V\colon=\frac1{n!}|\vol{}(x_2-x_1,\ldots,x_{n+1}-x_1)| \quad\mbox{and}\quad V_j\colon=\frac1{n!}|\vol{}(x_1-x,\ldots,x_{j-1}-x,x_{j+1}-x,\ldots,x_{n+1}-x)|, $$ i.e. $V$ is the Lebesgue measure of the simplex $S$ and $V_j$ is the Lebesgue measure of the simplex $S_j\colon=\convex{x,x_1,\ldots,x_{j-1},x_{j+1},\ldots,x_n}$. Then we have for all $x\in S$: $$ x=\sum_{j=1}^{n+1}\tfrac{V_j}Vx_j~. $$
cramer
W.l.o.g. we may assume that $E=\R^n$ with the canonical Euclidean product. The point $x$ is in $S$ iff there are non-negative numbers $t_j$, $j\leq n+1$, such that $\sum_jt_jx_j=x$ and $\sum_jt_j=1$. This means that the coefficients $t_j\in[0,1]$ satisfy the system of linear equations $$ \sum t_j=1 \quad\mbox{and}\quad \sum_{j=1}^{n+1}t_j(x_j-x)=0~. $$ By Cramer's rule we have: $t_j=D_j/D=|D_j|/|D|$, where $$ D =\det\left(\begin{array}{ccc} 1&\ldots&1\\ x_1-x&\ldots&x_{n+1}-x \end{array}\right) $$ and $D_j$ denotes the determinant of the matrix obtained from the previous matrix by replacing the $j$-th column with the column $(1,0,\ldots,0)$; by Laplace expansion we get: $$ D_j =(-1)^{j-1}\det(x_1-x,\ldots,x_{j-1}-x,x_{j+1}-x,\ldots,x_{n+1}-x) =(-1)^{j-1}\vol{}(x_1-x,\ldots,x_{j-1}-x,x_{j+1}-x,\ldots,x_{n+1}-x)| $$ and again by Laplace expansion: $D=\sum_{j=1}^{n+1}D_j$. Since $V=\sum V_j$, we also have $|D|=\sum_{j=1}^{n+1}|D_j|$ which shows in addition that all the numbers $D,D_1,\ldots,D_{n+1}$ have the same sign!
Prove that for any simplex $S$ in $\R^n$ the prism $S\times[0,1]$ is the union of $n+1$ simplices with pairwise disjoint interior, all of equal volume $\l(S)/(n+1)$. Suggested solution. In particular the cube $W=[0,1]^n$ is the union of $n!$ simplices with pairwise disjoint interior, all of which have the same volume $\l(W)/n!$.

Local Manifolds

Let us briefly discuss models physicists employ for the classical 'space' or the relativistic 'spacetime'.

Tangent spaces of open subsets of $\R^n$

Suppose we are given an open (and connected) subset $M$ of $\R^n$ - we'll call it a local manifold. The tangent space $T_xM$ of $M$ at $x$ is an $n$-dimensional real vector space disjoint from any other tangent space $T_yM$ for $y\neq x$. Thus the most obvious way to define $T_xM$ is $$ T_xM\colon=\{(x,v):v\in\R^n\}, $$ where $(x,v)+(x,u)\colon=(x,u+v)$ and $\l(x,v)\colon=(x,\l v)$. For $(x,v)$ we usually write $v_x$. The elements of $T_xM$ are called tangent vectors
of $M$ at the point $x$. In particular we put $$ E_x^{x_1}\colon=(x,1,0,\ldots,0), \ldots, E_x^{x_n}\colon=(x,0,\ldots,0,1)~. $$ These vectors form a basis for $T_xM$ and we will mostly denote them by $E_x^1,\ldots,E_x^n$. In some cases it's more convenient to denote the components (also called coordinates) of a point $x\in M$ by $r,\theta,t,\ldots$ and not by $x_1,x_2,\ldots$, thus $x=(r,\theta,t,\ldots)$. In these cases the corresponding basis vectors of $T_xM$ will be denoted by $E_x^r,E_x^\theta,E_x^t,\ldots$.
The set $TM\colon=\bigcup\{T_xM:x\in M\}=M\times\R^n$ is called the tangent space of $M$ and the mapping $\pi_M:TM\rar M$, $v_x\mapsto x$ the projection onto $M$.
A smooth mapping $X:M\rar TM$ is said to be a vector field, if for all $x\in M$: $\pi_M(X_x)=x$ (i.e. $X_x\in T_xM$). Thus \begin{equation}\label{diffeq1} X_x =\sum_j\z_j(x)\,E_x^j =(x,\z_1(x),\ldots,\z_n(x)) \end{equation} where $\z_j:M\rar\R$ are smooth functions.
Finally the vector-space $\G(M)$ denotes the space of all vector fields on $M$, i.e. $(X+Y)_x\colon=X_x+Y_x$ and $(\l X)_x\colon=\l X_x$. This space is also a so-called $C^\infty(M)$-modul, which simply means that for any smooth function $f:M\rar\R$ (i.e. $f\in C^\infty(M)$) and any vector field $X$ there is another vector field $fX$ defined by $(fX)_x\colon=f(x)X_x$. The 'simplest' vector fields are coordinate fields: let $x_1,\ldots,x_n$ be the coordinates of a point $x\in M$, then the mapping $x\mapsto E_x^{x_j}$ is said to be the $j$-th coordinate field. In general the notation for coordinate fields depends on the notation for the coordinates: if we denote the coordinates by $x_1,\ldots,x_n$, then the corresponding coordinate fields will be denoted by $E^{x_1},\ldots,E^{x_n}$ or simpler by $E^1,\ldots,E^n$ in case there is no misunderstanding. If we denote the coordinates by $x,y,t,\ldots$ or $r,\theta,\ldots$, then the associated coordinate fields will be denoted by $E^x,E^y,E^t\ldots$ or $E^r,E^\theta,\ldots$.
Suppose $c:I\rar M$, $t\mapsto c(t)$, is a curve with components $c_j(t)\colon=x^j(c(t))$. The tangent vector at $c(t)$ $$ c^\prime(t)\colon =\sum_j c_j^\prime(t)\,E_{c(t)}^j\in T_{c(t)}M $$ is said to be the derivative of the curve $c$ at the point $c(t)$. $c$ is said to be smooth if all functions $c_j$, $j=1,\ldots,n$, are smooth.
A curve $c:I\rar M$ is called an trajectory
or an integral curve of the vector field $X=\sum_j\z_j\,E^j$ through the point $x\in M$, if $c(0)=x$ and for all $t\in I$: $c^\prime(t)=X_{c(t)}$. This happens if and only if the components $c_j$ of the curve satisfy the following ordinary system of differential equations: $$ \forall j=1,\ldots,n:\qquad c_j^\prime(t)=\z_j(c_1(t),\ldots,c_n(t))~. $$ Suppose $N$ is another open subset of $\R^m$, say, $F:M\rar N$ smooth, $x\in M$ and $y\colon=F(x)\in N$. Let $E_x^{x_j}$ and $E_y^{y_k}$ denote the coordinate fields of $T_xM$ and $T_yN$, respectively. Then we define a linear map $T_xF:T_xM\rar T_yN$ by \begin{equation}\label{diffeq3} T_xF(E_x^{x_k})\colon=\sum_j\pa_{x_k}F_j(x)\,E_y^{y_j} \end{equation} This map is called the tangent map of $F$ at $x$ and its characterized by the fact that the matrix of $T_xF$ with respect to the bases $E_x^j$ and $E_y^k$ is just the Jacobian $DF(x)$ of $F$ at $x$. Thus we have the chain rule: \begin{equation}\label{diffeq4} T_xG\circ F=T_{F(x)}G\circ T_xF \end{equation} and in particular for a smooth curve $c:I\rar M$: $(F\circ c)^\prime(t)=T_{c(t)}F(c^\prime(t))$. The mapping $TF:TM\rar TN$ is usually denoted by $F_*$, i.e. $$ F_*(E_x^{x_j})\colon=T_xF(E_x^{x_j})~. $$ Hence we also write for $(F\circ c)^\prime(t)=T_{c(t)}F(c^\prime(t))$: $$ (F\circ c)^\prime(t)=F_*(c^\prime(t))~. $$ This $*$-notation is utilized in case there is no ambiguity about the point $x$ at which the tangent map is taken!
tangent map

Local Pseudo-Riemannian manifolds

Suppose each tangent space $T_xM$ of $M$ is furnished with an inner product $\la.,.\ra_x$. This amount to the assertion that for each $x\in M$ the matrix $$ (g_{jk}(x))\colon=(\la E^j,E^k\ra_x) =(\la E_x^j,E_x^k\ra_x) =(\la E_x^j,E_x^k\ra) $$ is non-singular and symmetric. If for all $j,k$ the functions $x\mapsto g_{jk}(x)$ are smooth, then $\la.,.\ra$ (or the matrix valued function $x\mapsto(g_{jk}(x))$) is called a Pseudo-Riemannian metric on $M$ and the pair $(M,\la.,.\ra)$ is called a local Pseudo-Riemannian manifold. As the map that sends each symmetric matrix to its list of eigenvalues is continuous, the index of the inner products $\la.,.\ra_x$ is constant on connected components. If all inner products $\la.,.\ra_x$ are Euclidean, $\la.,.\ra$ is called a Riemannian metric and $(M,\la.,.\ra)$ is called a local Riemannian manifold. If all inner products $\la.,.\ra_x$ are Lorentz products, $\la.,.\ra$ is called a Lorentz metric and $(M,\la.,.\ra)$ is called a local Lorentz manifold. The standard model of classical physics is $M\colon=\R^n$ furnished with the Riemannian metric $g_{jk}(x)=\d_{jk}$ - this is called the cannonical Euclidean metric on $M$, usually denoted by $\can$, i.e. $$ \forall x\in M\,\forall j,k:\quad \can(E_x^j,E_x^k)=\d_{jk}~. $$ The standard model of special relativity is $M\colon=\R^{n+1}$ furnished with the Lorentz metric $g_{jk}(x)=\e_j\d_{jk}$, where $\e_0=-1$ and $\e_1=\cdots=\e_n=+1$ - the space $M$ endowed with this metric is called the $(n+1)$-dimensional Minkowski space - we will use the symbol $\R_1^{n+1}$ both for the $(n+1)$-dimensional Minkowski space, which is a manifold and the vector space $\R^{n+1}$ endowed with the canonical Lorentz product. In both the Euclidean and the Minkowski space the coordinate fields $E_x^1,\ldots,E_x^n$ and $E_x^0,\ldots,E_x^n$, respectively, form an orthonormal basis at any point $x\in M$.

The existance of coordinates such that the corresponding coordinate fields form an orthonormal basis at each point essentially characterizes these spaces: the Euclidean space and the Minkowski space; the first is the space of classical physics and the latter is the spacetime of special relativity.

The length of a piecewise smooth curve $c:[0,1]\rar M$ in a local Riemannian manifold $(M,\la.,.\ra)$ is defined by $$ L(c)\colon=\int_0^1\sqrt{\la c^\prime(s),c^\prime(s)\ra_{c(s)}}\,ds $$ In classical physics the length of a curve is of minor importance, it's just a number among others. That's completely different in relativity: suppose $c:[0,1]\rar M$ is a piecewise smooth curve in a local Lorentz manifold $(M,\la.,.\ra)$ such that for all $s$: $\la c^\prime(s),c^\prime(s)\ra < 0$ - physicists will say it's a world line
, then $$ T(c)\colon=\int_0^1\sqrt{-\la c^\prime(s),c^\prime(s)\ra_{c(s)}}\,ds $$ is called the proper time of the curve - usually we will drop the subscript $c(s)$ in $\la c^\prime(s),c^\prime(s)\ra_{c(s)}$!
The curve $c(s)=(s,R\cos(\o s),R\sin(\o s))$ is a world line in three dimensional Minkowski space $M$ iff $|R\o| < 1$. In this case the proper time of $c:[0,2\pi/\o]\rar M$ is $2\pi\sqrt{1-R^2\o^2}/\o$.
We have $c^\prime(s)=E_{c(s)}^0-R\o\sin(\o s)E_{c(s)}^1+R\o\cos(\o s)E_{c(s)}^2$ and thus $$ -\la c^\prime(s),c^\prime(s)\ra =1-R^2\o^2\sin^2(\o s)-R^2\o^2\cos^2(\o s) =1-R^2\o^2~. $$ A world line $c:[0,T]\rar M$ is said to be an observer - sometimes called a material particle - if $$ \forall s\in(0,T):\quad \la c^\prime(s),c^\prime(s)\ra=-1 $$ This holds if and only if for all $\t\in[0,T]$ the proper time of the curve $c:[0,\t]\rar M$ equals $\t$, i.e. $c$ is parametrized by it's proper time! In exam the curve $$ \t\mapsto(\t\b,R\cos(\o\t\b),R\sin(\o\t\b)) \quad\mbox{where}\quad \b\colon=1/\sqrt{1-R^2\o^2} $$ is an observer in three dimensional Minkowski space $M$. We will say that it's an observer moving with constant speed $R\o$ (or constant angular velocity $\o$) on a circle of radius $R$. Beware, on a local Lorentz manifold there is no notion of distance, cf. exam. Hence, strictly speaking, a circle of radius $R$ is not defined. The time measured by this observer for a full rotation is $T\colon=2\pi\sqrt{1-R^2\o^2}/\o$ - i.e. The time elapsed for this observer from event $c(0)=(0,R,0)$ through event $c(T)=(2\pi/\o,R,0)$ is $T$. In general it needs some reparametrization to turn a world line into an observer.
The time elapsed for the observer $k(s)=(s,R,0)$ from event $k(0)=c(0)$ through event $c(T)=k(2\pi/\o)$ is $2\pi/\o$, which is larger than $T$, the time elapsed for the observer $c$.
$(c(t)\colon=(t,at^2/2)$, $t\in[0,1]$, is a world line in two dimensional Minkowski space iff $|a| < 1$. Show that the proper time is given by $$ \t(t)=\frac{\arcsin(at)+at\sqrt{1-a^2t^2}}{2a}, $$ i.e. if $\t\mapsto t(\t)$ denotes the inverse of $t\mapsto\t(t)$, then the time elapsed for the observer $k(\t)\colon=c(t(\t))$ from event $(0,0)$ through event $(t,c(t))$ is $\t(t)$.
In the following chapter we will see (cf. exam) that the world line of the previous example describes an observer accelerating constantly in a fixed direction.
Let $c:[a,b]\rar M$ be a smooth curve in the Lorentz manifold $M$ such that for all $t\in[a,b]$: $c^\prime(t)$ is space-like. Then $$ L(c)\colon=\int_a^b\Vert c^\prime(t)\Vert\,dt $$ is called the Lorentz-length of the curve. Verify that the lenght of the curve $c(t)=(\cosh(t),\sinh(t))$, $t\in[a,b]$ in two dimensional Minkowski space is given by $b-a$ - i.e. the curve is parametrized by arc-length!

Submanifolds of local manifolds

For our purposes it will suffice to define a submanifold $M$ of a local maifold $U(\sbe\R^{n+k})$ as level set of a smooth function $F:U\rar\R^k$ - i.e. $M=[F=0]$, such that for all $x\in M$ the linear mapping $T_xF:T_xU\rar T_{F(x)}\R^k$ is onto - such mappings are called submersions. Computationally this comes down to showing that the Jacobi matrix $DF(x)$ has rank $k$ at every point $x\in M$. The tangent space $T_xM$ of $M$ at $x$ is then defined by $$ T_xM\colon=\ker(T_xF:T_xU\rar T_{F(x)}\R^k) $$ By the Rank-Nullity Theorem \eqref{lareq5} this implies that $T_xM$ is an $n$-dimensional subspace of $T_xU$. Sometimes it might be more convenient to define $M$ as level set of $k$ real valued functions - the components $F_1,\ldots,F_k$ of $F$. If $(U,\la.,.\ra)$ is Riemannian, then any submanifold is Riemannian, but this no longer holds for Lorentz manifolds - a submanifold of a local Lorentz manifold $(U,\la.,.\ra)$ need not be a Pseudo-Riemannian at all. Only if all subspaces $T_xM$ are not degenerated, $M$ is Pseudo-Riemannian: For example the submanifold $$ H^n\colon=\{(x_0,\ldots,x_n):-x_0^2+x_1^2+\cdots+x_n^2=-1, x_0 > 0\} $$ of Minkowski space $\R_1^{n+1}$ with the induced Lorentz-metric is a Riemannian manifold. Indeed, the tangent space at a point $x=(x_0,x_1,\ldots,x_n)$ is the subspace orthogonal to $N_x\colon=x_0E_x^0+\cdots+x_nE_x^n$, i.e. $$ T_xM=\Big\{\sum v_jE_x^j:-x_0v_0+x_1v_1+\cdots+x_nv_n=0\Big\}~. $$ Since $N_x$ is time-like it's orthogonal complement must be space-like (by lemma). Hence $H^n$ is a Riemannian manifold. $H^n$ is a model of the $n$-dimensional hyperbolic space of constant curvature $-1$; its geodesics are obtained in just the same way geodesics are obtained on the Euclidean sphere $S^n$, i.e. by intersecting the submanifold with any $2$-dimensional plane in $\R^{n+1}$ containing the origin.
$\R\times S^1\colon=\{(x_0,x_1,x_2)\in\R_1^3: x_1^2+x_2^2=1\}$ is a Lorentz submanifold of $\R_1^3$. This is another example of a Lorentz manifold which admits orthonormal coordinate fields globally.
$\R\times S^{n-1}\colon=\{(x_0,x_1,\ldots,x_n)\in\R_1^{n+1}: x_1^2+\cdots+x_n^2=1\}$ is a Lorentz submanifold of $\R_1^{n+1}$.
$S^{n}\colon=\{(x_0,x_1,\ldots,x_n)\in\R_1^{n+1}: x_0^2+\cdots+x_n^2=1\}$ is not a Pseudo-Riemannian submanifold of $\R_1^{n+1}$.
If $c:[0,1]\rar U$ is a curve in $U$ such that for all $t\in[0,1]$: $c(t)\in M\colon=[F=0]$, then for all $t$: $c^\prime(t)\in T_{c(t)}M$. Conversely, if $c(0)\in M$ and for all $t\in[0,1]$: $c^\prime(t)\in T_{c(t)}M$, then for all $t\in[0,1]$: $c(t)\in M$.
This follows from the chain rule: since $F(c(t))=0$, we get $T_{c(t)}F(c^\prime(t))=0$, i.e. $c^\prime(t)\in\ker(T_{c(t)}F:T_{c(t)}U\rar T_{F(c(t))})=M_{c(t)}$. Conversely, if for all $t$: $c^\prime(t)\in T_{c(t)}M$, then $F\circ c$ is constant and since $c(0)\in M$ we conclude that $F(c(t))=0$, i.e. $c(t)\in M$.
If $(M,\la.,.\ra)$ is a local Riemannian manifold, then $\R\times M$ endowed with the metric $g(E^0,E^0)=-1$, $g(E^0,E^j)=0$ for $j=1,\ldots,n$ and $g(E^j,E^k)=\la E^j,E^k)$ is a local Lorentz manifold.
There are also submanifolds of e.g. $\R^3$ without any Lorentz metric: the sphere $S^2$ cannot be endowed with a Lorentz metric - the topological obstruction is essentially known as the hairy ball theorem. For the same reason there is no Lorentz metric on $S^{2n}(\sbe\R^{2n+1})$. A vector field $X$ on a submanifold $M$ of $U$ is defined to be a vector field in a neigbourhood of $M$ in $U$ (usually we don't care about that neighbourhood) such that $$ \forall x\in M:\quad X_x\in T_xM~. $$
On $S^3(\sbe\R^4)$ there is a Lorentz metric.
Let $N\colon=x_1E^1+x_2E^2+x_3E^3+x_4E^4$ be the unit normal field, i.e. for all $x\in S^3$ and all $X_x\in T_xS^3$: $\can(X_x,N_x)=0$ and $\can(X_x,X_x)=1$, then $$ Z\colon=-x_3E^1-x_4E^2+x_1E^3+x_2E^4 $$ is a unit vector field on $S^3$. Putting $$ g_x(X,Y)\colon=\can_x(X,Y)-2\can_x(X,Z)\can_x(Y,Z) $$ we get a Lorentz metric on $S^3$ (cf. exam) such that for all $x\in S^3$: $g_x(Z,Z)=-1$.
On $S^{2n-1}(\sbe\R^{2n})$ there is a Lorentz metric.
Suppose $F:N\rar M$ is a smooth mapping such that for all $y\in N$: $T_yF$ is injective - such mappings are called immersions. For any Pseudo-Riemannian metric $\la.,.\ra$ on $M$ we define for any pair of vector fields $X,Y$ on $N$: $$ g(X,Y)_x\colon=\la T_xF(X_x),T_xF(Y_x)\ra_{F(x)}~. $$ Then $g$ is a Pseudo-Riemannian metric on $N$. The metric $g$ is called the pull-back metric of the metric $\la.,.\ra$ by $F$. For example $$ F(r,\theta,\vp)\colon=(r\cos\vp\sin\theta,r\sin\vp\sin\theta,r\cos\theta) $$ maps $N\colon=\R\sm\{0\}\times(0,\pi)\times(0,2\pi)$ into the three dimensional Euclidean space $M=(\R^3,\can)$ and the pull-back of $\can$ by $F$ is $$ g(E^r,E^r)=1, g(E^\theta,E^\theta)=r^2, g(E^\vp,E^\vp)=r^2\sin^2\theta $$ and $g(E^r,E^\theta)=g(E^r,E^\vp)=g(E^\theta,E^\vp)=0$.
Put $y\colon=(r,\theta,\vp)\in N$ and $x\colon=F(y)$, then the matrix of $T_yF$ with respect to the bases $E^r,E^\theta,E^\vp$ and $E^1,E^2,E^3$ of $T_yN$ and $T_x\R^3$, respectively, is the Jacobi matrix $DF(y)$. For all vector fields $X$ and $Y$ on $N$ we have $$ g(X,Y)_y =\can(TF_y(X),T_yF(Y)) =\can(T_yF^*T_yF(X),Y)~. $$ As the matrix of $T_yF^*T_yF$ with respect to the basis $E^r,E^\theta,E^\vp$ is $DF(y)^tDF(y)$, the matrix of metric coefficients $$ \left(\begin{array}{ccc} g(E^r,E^r)&g(E^r,E^\theta)&g(E^r,E^\vp)\\ g(E^\theta,E^r)&g(E^\theta,E^\theta)&g(E^\theta,E^\vp)\\ g(E^\vp,E^r)&g(E^\vp,E^\theta)&g(E^\vp,E^\vp) \end{array}\right) $$ is just $DF(y)^tDF(y)$. The metric $g$ is called the Euclidean metric in polar coordinates $(r,\theta,\vp)$: $r$ is the radial part, $\theta$ the latitude - here the north pole has latitude $0$ - and $\vp$ the longitude.
Put $N\colon=\R\sm\{0\}\times\R\times(0,2\pi)$, $M$ three dimensional Minkowski space $\R_1^3$ and $F:N\rar M$: $$ F(r,\theta,\vp)\colon=(r\cosh\theta,r\sinh\theta\cos\vp,r\sinh\theta\sin\vp)~. $$ Compute the pull-back metric $h$ of $F:N\rar M$. Suggested solution.
Suppose $N$ carries the pull-back metric $h$ of $(M,g)$ under $F:N\rar M$. Then for any curve $c:I\rar N$ the curve $c:I\rar(N,h)$ and the curve $F\circ c:I\rar(M,g)$ have the same length.
In particuar for any triple of smooth functions $(r,\theta,\vp):I\rar\R^+\times(0,\pi)\times(0,2\pi)$ the length of the curve $$ t\mapsto(r(t)\cos\vp(t)\cos\theta(t),r(t)\sin\vp(t)\cos\theta(t),r(t)\sin\theta(t)) $$ in Euclidean space $\R^3$ is given by $$ \int\sqrt{r^\prime(t)^2+r(t)^2(\theta^\prime(t)^2+\vp^\prime(t)^2\sin^2\theta(t))}\,dt~. $$
Two Pseudo-Riemannian manifolds $(N,h)$ and $(M,g)$ are said to be locally isometric if there is a smooth mapping $F:N\rar M$ such that the pull-back metric of $g$ by $F$ equals $h$, i.e. for all vector fields $X$ and $Y$ on $N$ and all $y\in N$: $$ h(X_y,Y_y)=g(T_yF(X_y),T_y(Y_y))~. $$ $(N,h)$ and $(M,g)$ are said to be isometric if in addition $F:N\rar M$ is a bijection.

What's the difference between the vector space $\R_1^{n+1}$ and the manifold $\R_1^{n+1}$

These spaces are of course different but we are somehow used to identify a bunch of things: for each pair $x,y$ in Minkowski space $M=\R_1^{n+1}$ (i.e. the manifold $\R_1^{n+1}$) there is a map called parallel translation: $P_{x\to y}:T_xM\rar T_yM$, $\sum v_jE_x^j\mapsto\sum v_jE_y^j$ and there is also an obvious map $J$ mapping $T_0M$ onto the vector space $\R_1^{n+1}$.

All these mappings are isometries of Lorentz spaces.

If you like to identify all tangent vectors $\sum v_jE_x^j\in T_xM$, $x\in M$, by means of these mappings with the vector $v\colon=(v_0,\ldots,v_n)$ in the vector space $\R_1^{n+1}$, you get the picture you are probably familiar with! For instance, the 'length' of a curve $c(t)=(c_0(t),\ldots,c_n(t))$ in the vector space $\R_1^{n+1}$ is simply defined by $\int\norm{c^\prime(t)}\,dt$, where $c^\prime(t)=(c_0^\prime(t),\ldots,c_n^\prime(t))$ is again a vector in $\R_1^{n+1}$. We'd get the same formula by just identifying in e.g.
exam the tangent vector $c^\prime(t)=\sum c_j^\prime(t)E_{c(t)}^j$ with the vector $(c_0^\prime(t),\ldots,c_n^\prime(t))$ in the vector space $\R_1^{n+1}$. However, these identifications only work for Euclidean spaces and Minkowski spaces and usually do not work for any other Riemannian or Lorentz manifold!
← Introduction → Lorentz Spaces and Relativity

<Home>
Last modified: Wed Nov 13 15:52:24 CET 2024