Applied linear operators and spectral methods/Lecture 4

< Applied linear operators and spectral methods

More on spectral decompositions

In the course of the previous lecture we essentially proved the following theorem:

Theorem:

1) If a n\times n matrix \mathbf{A} has n linearly independent real or complex eigenvectors, the \mathbf{A} can be diagonalized. 2) If \mathbf{T} is a matrix whose columns are eigenvectors then \mathbf{T}\mathbf{A}\mathbf{T}^{-1}= \boldsymbol{\Lambda} is the diagonal matrix of eigenvalues.

The factorization \mathbf{A} = \mathbf{T}^{-1}\boldsymbol{\Lambda}\mathbf{T} is called the spectral representation of \mathbf{A}.

Application

We can use the spectral representation to solve a system of linear homogeneous ordinary differential equations.

For example, we could wish to solve the system

 
\cfrac{d\mathbf{u}}{dt} = \mathbf{A}\mathbf{u} = \begin{bmatrix}-2 & 1 \\ 1 & -2 \end{bmatrix} 
\begin{bmatrix} u_1 \\ u_2 \end{bmatrix}

(More generally \mathbf{A} could be a n \times n matrix.)

Comment:

Higher order ordinary differential equations can be reduced to this form. For example,


  \cfrac{d^2u_1}{dt^2} + a~\cfrac{du_1}{dt} = b~u_1

Introduce


  u_2 = \cfrac{du_1}{dt}

Then the system of equations is


  \begin{align}
    \cfrac{du_1}{dt} & = u_2 \\
    \cfrac{du_2}{dt} & = b~u_1 - a~u_2 
  \end{align}

or,


  \cfrac{d\mathbf{u}}{dt} = \begin{bmatrix} 0 & 1 \\ b & -a \end{bmatrix} 
      \begin{bmatrix} u_1 \\ u_2 \end{bmatrix} = \mathbf{A} \mathbf{u}

Returning to the original problem, let us find the eigenvalues and eigenvectors of \mathbf{A}. The characteristic equation is


  det(\mathbf{A} - \lambda~\mathbf{I}) = 0

o we can calculate the eigenvalues as


  (2+\lambda)(2+\lambda) - 1 = 0 \quad \implies \quad
  \lambda^2 + 4\lambda + 3 = 0 \qquad \implies \qquad \lambda_1 = -1, \qquad \lambda_2= -3

The eigenvectors are given by


  (\mathbf{A} - \lambda_1~\mathbf{I})\mathbf{n}_1 = \mathbf{0} ~;~~ (\mathbf{A} - \lambda_2~\mathbf{I})\mathbf{n}_2 = \mathbf{0}

or,


  -n_1^1 + n_2^1 = 0 ~;~~ n_1^1 - n_2^1 = 0 ~;~~
  n_1^2 + n_2^2 = 0 ~;~~ n_1^2 + n_2^2 = 0

Possible choices of \mathbf{n}_1 and \mathbf{n}_2 are


  \mathbf{n}_1 = \begin{bmatrix} 1 \\ 1 \end{bmatrix} ~;~~
  \mathbf{n}_2 = \begin{bmatrix} 1 \\ -1 \end{bmatrix}

The matrix \mathbf{T} is one whose columd are the eigenvectors of \mathbf{A}, i.e.,


  \mathbf{T} = \begin{bmatrix} 1 & 1 \\ 1 & -1 \end{bmatrix}

and


  \boldsymbol{\Lambda} = \mathbf{T}^{-1} \mathbf{A} \mathbf{T} = \begin{bmatrix} -1 & 0 \\ 0 & -3 \end{bmatrix}

If \mathbf{u} = \mathbf{T} \mathbf{u}^{'} the system of equations becomes


  \cfrac{d\mathbf{u}^{'}}{dt} = \mathbf{T}^{-1} \mathbf{A} \mathbf{T} \mathbf{u}^{'} = \boldsymbol{\Lambda}~\mathbf{u}^{'}

Expanded out


  \cfrac{du_1^{'}}{dt} = - u_1^{'} ~;~~ \cfrac{du_2^{'}}{dt} = - 3~u_2^{'}

The solutions of these equations are


  u_1^{'} = C_1~e^{-t} ~;~~ u_2^{'} = C_2~e^{-3t}

Therefore,


  \mathbf{u} = \mathbf{T}~\mathbf{u}^{'} = \begin{bmatrix} C_1~e^{-t} + C_2~e^{-3t} \\ C_1~e^{-t} - C_2~e^{-3t}
     \end{bmatrix}

This is the solution of the system of ODEs that we seek.

Most "generic" matrices have linearly independent eigenvectors. Generally a matrix will have n distinct eigenvalues unless there are symmetries that lead to repeated values.

Theorem

If \mathbf{A} has k distinct eigenvalues then it has k linearly independent eigenvectors.

Proof:

We prove this by induction.

Let \mathbf{n}_j be the eigenvector corresponding to the eigenvalue \lambda_j. Suppose \mathbf{n}_1, \mathbf{n}_2, \dots, \mathbf{n}_{k-1} are linearly independent (note that this is true for k = 2). The question then becomes: Do there exist \alpha_1, \alpha_2, \dots, \alpha_k not all zero such that the linear combination


  \alpha_1~\mathbf{n}_1 + \alpha_2~\mathbf{n}_2 + \dots + \alpha_k~\mathbf{n}_k = 0

Let us multiply the above by (\mathbf{A} - \lambda_k~\mathbf{I}). Then, since \mathbf{A}~\mathbf{n}_i = \lambda_i~\mathbf{n}_i, we have


  \alpha_1~(\lambda_1 - \lambda_k)~\mathbf{n}_1 + \alpha_2~(\lambda_2 - \lambda_k)~\mathbf{n}_2 + \dots + 
  \alpha_{k-1}~(\lambda_{k-1} - \lambda_k)~\mathbf{n}_{k-1} + 
  \alpha_k~(\lambda_k - \lambda_k)~\mathbf{n}_k = \mathbf{0}

Since \lambda_k is arbitrary, the above is true only when


  \alpha_1 = \alpha_2 = \dots = \alpha_{k-1} = 0

In thast case we must have


  \alpha_k~\mathbf{n}_k = \mathbf{0} \quad \implies \quad \alpha_k = 0

This leads to a contradiction.

Therefore \mathbf{n}_1, \mathbf{n}_2, \dots, \mathbf{n}_k are linearly independent. \qquad \square

Another important class of matrices which are diagonalizable are those which are self-adjoint.

Theorem

If \boldsymbol{A} is self-adjoint the following statements are true


  1. \langle \boldsymbol{A}\mathbf{x}, \mathbf{x} \rangle is real for all \mathbf{x}.
  2. All eigenvalues are real.
  3. Eigenvectors of distinct eigenvalues are orthogonal.
  4. There is an orthonormal basis formed by the eigenvectors.
  5. The matrix \boldsymbol{A} can be diagonalized (this is a consequence of the previous statement.)


Proof

1) Because the matrix is self-adjoint we have


  \langle \boldsymbol{A}\mathbf{x}, \mathbf{x} \rangle = \langle \mathbf{x}, \boldsymbol{A}\mathbf{x} \rangle

From the property of the inner product we have


 \langle \mathbf{x}, \boldsymbol{A}\mathbf{x} \rangle = \overline{\langle \boldsymbol{A}\mathbf{x}, \mathbf{x} \rangle}

Therefore,


  \langle \boldsymbol{A}\mathbf{x}, \mathbf{x} \rangle =  \overline{\langle \boldsymbol{A}\mathbf{x}, \mathbf{x} \rangle}

which implies that \langle \boldsymbol{A}\mathbf{x}, \mathbf{x} \rangle is real.

2) Since \langle \boldsymbol{A}\mathbf{x}, \mathbf{x} \rangle is real, \langle \boldsymbol{I}\mathbf{x}, \mathbf{x} \rangle = \langle \mathbf{x},\mathbf{x} \rangle is real. Also, from the eiegnevalue problem, we have


  \langle \boldsymbol{A}\mathbf{x}, \mathbf{x} \rangle = \lambda~\langle \mathbf{x}, \mathbf{x} \rangle

Therefore, \lambda is real.

3) If (\lambda,\mathbf{x}) and (\mu, \mathbf{y}) are two eigenpairs then


  \lambda~\langle \mathbf{x}, \mathbf{y} \rangle = \langle \boldsymbol{A}\mathbf{x}, \mathbf{y} \rangle

Since the matrix is self-adjoint, we have


  \lambda~\langle \mathbf{x}, \mathbf{y} \rangle = \langle \mathbf{x}, \boldsymbol{A}\mathbf{y} \rangle = \mu~\langle \mathbf{x}, \mathbf{y} \rangle

Therefore, if \lambda \ne \mu \ne 0, we must have


  \langle \mathbf{x}, \mathbf{y} \rangle = 0

Hence the eigenvectors are orthogonal.

4) This part is a bit more involved. We need to define a manifold first.

Linear manifold

A linear manifold (or vector subspace) \mathcal{M} \in \mathcal{S} is a subset of \mathcal{S} which is closed under scalar multiplication and vector addition.

Examples are a line through the origin of n-dimensional space, a plane through the origin, the whole space, the zero vector, etc.

Invariant manifold

An invariant manifold \mathcal{M} for the matrix \boldsymbol{A} is the linear manifold for which \mathbf{x} \in \mathcal{M} implies \boldsymbol{A}\mathbf{x} \in \mathcal{M}.

Examples are the null space and range of a matrix \boldsymbol{A}. For the case of a rotation about an axis through the origin in n-space, invaraiant manifolds are the origin, the plane perpendicular to the axis, the whole space, and the axis itself.

Therefore, if \mathbf{x}_1, \mathbf{x}_2, \dots,\mathbf{x}_m are a basis for \mathcal{M} and \mathbf{x}_{m+1}, \dots, \mathbf{x}_n are a basis for \mathcal{M}_\perp (the perpendicular component of \mathcal{M}) then in this basis \boldsymbol{A} has the representation


  \boldsymbol{A} = \begin{bmatrix} x & x & | & x & x \\
                        x & x & | & x & x \\
                        - & - & - & - & - \\
                        0 & 0 & | & x & x \\
                        0 & 0 & | & x & x  \end{bmatrix}

We need a matrix of this form for it to be in an invariant manifold for \boldsymbol{A}.

Note that if \mathcal{M} is an invariant manifold of \boldsymbol{A} it does not follow that \mathcal{M}_\perp is also an invariant manifold.

Now, if \boldsymbol{A} is self adjoint then the entries in the off-diagonal spots must be zero too. In that case, \boldsymbol{A} is block diagonal in this basis.

Getting back to part (4), we know that there exists at least one eigenpair (\lambda_1, \mathbf{x}_1) (this is true for any matrix). We now use induction. Suppose that we have found (n-1) mutually orthogonal eigenvectors \mathbf{x}_i with \boldsymbol{A}\mathbf{x}_i = \lambda_i\mathbf{x}_i and \lambda_i are real, i = 1, \dots, k-1. Note that the \mathbf{x}_is are invariant manifolds of \boldsymbol{A} as is the space spanned by the \mathbf{x}_is and so is the manifold perpendicular to these vectors).

We form the linear manifold


  \mathcal{M}_k = \{\mathbf{x} | \langle \mathbf{x}, \mathbf{x}_j \rangle = 0 ~~ j = 1, 2, \dots, k-1\}

This is the orthogonal component of the k-1 eigenvectors \mathbf{x}_1, \mathbf{x}_2, \dots, \mathbf{x}_{k-1} If \mathbf{x} \in \mathcal{M}_k then


  \langle \mathbf{x}, \mathbf{x}_j \rangle = 0 \quad \text{and} \quad
  \langle \boldsymbol{A}\mathbf{x}, \mathbf{x}_j \rangle = \langle \mathbf{x}, \boldsymbol{A}\mathbf{x}_j \rangle = \lambda_j\langle \mathbf{x}, \mathbf{x}_j \rangle = 0

Therefore \boldsymbol{A}\mathbf{x} \in \mathcal{M}_k which means that \mathcal{M}_k is invariant.

Hence \mathcal{M}_k contains at least one eigenvector \mathbf{x}_k with real eigenvalue \lambda_k. We can repeat the procedure to get a diagonal matrix in the lower block of the block diagonal representation of \boldsymbol{A}. We then get n distinct eigenvectors and so \boldsymbol{A} can be diagonalized. This implies that the eigenvectors form an orthonormal basis.

5) This follows from the previous result because each eigenvector can be normalized so that \langle \mathbf{x}_i, \mathbf{x}_j \rangle = \delta_{ij}.

We will explore some more of these ideas in the next lecture.

Resource type: this resource contains a lecture or lecture notes.
Action required: please create Category:Applied linear operators and spectral methods/Lectures and add it to Category:Lectures.
This article is issued from Wikiversity - version of the Monday, September 13, 2010. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.