We saw in the previous lecture that \(A \in M_{n \times n}\) is diagonalizable if $A$ has $n$ linearly independent eigenvectors \(\beta = \{v_1, v_2,...,v_n\}\). In that case,

$$ \begin{align*} [L_A]_{\beta}^{\beta} = \begin{pmatrix} \lambda_1 & \cdots & 0 \\ \vdots & \ddots & \vdots \\ 0 & \cdots & \lambda_n \end{pmatrix}, Av_j = \lambda_j v_j \end{align*} $$

\(v_j\) is an eigenvector and \(\lambda_j\) is called an eigenvalue. We developed a plan to find these and eigenvalues and eigenvectors.

  1. Find these eigenvalues \(\lambda_1, ..., \lambda_2\)
  2. Find a basis for \(\beta_i\) for each eigenspace \(E_{\lambda_i}\)
  3. If \(\sum_{j=1}^{k}\dim(E_{\lambda_i}) = n\), then collect all the separate bases and that's our basis \(\beta = \beta_1 \cup ... \cup \beta_k\). This step works because we proved in the last lecture the corollary \(\lambda_1 \neq \lambda_2 \rightarrow \beta_1 \cup \beta_2\) is linearly independent.


Examples of Non-Diagonalizable Matrices

Of course the plan above might not work and there are different ways, a matrix could fail to be diagonalized.

$$ \begin{align*} A = \begin{pmatrix} 0 & -1 \\ 1 & 0 \end{pmatrix} \end{align*} $$

This matrix has no eigenvalues so it can’t be diagonalized.

$$ \begin{align*} B = \begin{pmatrix} 1 & 1 \\ 0 & 1 \end{pmatrix} \end{align*} $$

This matrix has only one eigen value.

$$ \begin{align*} det(A - tI_2) = \begin{pmatrix} 1-t & 1 \\ 0 & 1-t \end{pmatrix} = (1-t)^2 = 0 \rightarrow t = 1 \end{align*} $$

This implies

$$ \begin{align*} E_{\lambda_1} = N(B - 1I_n) = N \left( \begin{pmatrix} 0 & 1 \\ 0 & 0 \end{pmatrix} \right) = \{ (t, 0) \ | \ t \in \mathbf{R} \} \end{align*} $$

So we don’t have enough linearly independent vectors in \(\beta\).



Is There a Better Diagonlization Test?

Today, we will refine our answer to the question “Is \(A\) diagonalizable?”

Definition
A polynomial \(f(t)\) splits over \(\mathbf{R}\) if there are scalars \(c,a_1,...,a_k \in \mathbf{R}\) such that $$ \begin{align*} f(t) = c(t-a_1)(t - a_2)...(t-a_k) \end{align*} $$


In other words, can completely factorize the polynomial?

Example 1: \(t^2 + 1\) doesn’t split over \(\mathbf{R}\). It does however split over \(\mathbf{C} as (t - i)(t + i)\).

Example 2: \(t^2 - 2t + 1 = (t - 1)(t - 1)\) splits over \(\mathbf{R}\).

So now what does splitting have to do with diagonalizability? The following theorem explains this

Theorem 1
If \(A\) is diagonalizable, then its characteristic polynomial splits over \(\mathbf{R}\)


Note that the the converse is false. If the characteristic polynomial splits over \(\mathbf{R}\), it doesn’t necessarily means that \(A\) is diagonalizable. (See example 2 above)



Eigenvalues of Similar Matrices

Before going into the proof of theorem 1 above, the following proposition will be useful

Proposition
If \(B = Q^{-1}AQ\), then $$ \begin{align*} \det(B - tI_n) = \det(A - tI_n) \end{align*} $$


This means that if \(A\) and \(B\) are similar, then they have the same eigenvalues.

Proof:

Observe that

$$ \begin{align*} 1 = \det(I_n) = \det(QQ^{-1}) = \det(Q)\det(Q^{-1}). \end{align*} $$

Therefore,

$$ \begin{align*} \det(B - tI_n) &= \det(Q^{-1}AQ - tI_n) \\ &= \det(Q^{-1}AQ - QtI_nQ^{-1}) \\ &= \det(Q (A - tI_n)Q^{-1}) \text{ (factor out Q on the left ..) }\\ &= \det(Q) \det((A - tI_n)) \det(Q^{-1}) \\ &= \det(Q) \det(Q^{-1}) \det((A - tI_n)) \text{ (they are just real numbers)}\\ &= \det((A - tI_n)). \ \blacksquare \\ \end{align*} $$

Note here that in step 2, \(QtI_nQ^{-1} = tQQ^{-1} = tQtQ^{-1} = tI_n\)



Proof of Theorem 1

Suppose that \(A\) is diagonalizable. This means that there exists a basis \(\beta\) such that

$$ \begin{align*} [L_A]_{\beta}^{\beta} = \begin{pmatrix} \lambda_1 & \cdots & 0 \\ \vdots & \ddots & \vdots \\ 0 & \cdots & \lambda_n \end{pmatrix}, Av_j = \lambda_j v_j \end{align*} $$

If we let \(\alpha\) be the standard basis, then know that there exists a change of coordinate matrix \(Q = [I]^{\alpha}_{\beta}\) such that

$$ \begin{align*} [L_A]_{\beta}^{\beta} &= [I]_{\alpha}^{\beta}[L_A]_{\alpha}^{\alpha}[I]^{\alpha}_{\beta} \\ &= Q^{-1}AQ. \end{align*} $$

This means that \([L_A]_{\beta}^{\beta}\) and \(A\) are similar matrices. But by the previous proposition, this means that we have

$$ \begin{align*} \det ([L_A]_{\beta}^{\beta} -tI_n) &= \det(A - tI_n). \end{align*} $$

On the other hand,

$$ \begin{align*} \det ([L_A]_{\beta}^{\beta} -tI_n) &= \begin{pmatrix} \lambda_1 -t & \cdots & 0 \\ \vdots & \ddots & \vdots \\ 0 & \cdots & \lambda_n -t \end{pmatrix} \\ &= (\lambda_1 - t)...(\lambda_n - t). \ \blacksquare \end{align*} $$




Algebraic and Geometric Multiplicities of Eigenvalues

There is another condition for diagonalizability but we need a few more definitions.

Definition
The algebraic multiplicity of \(\lambda\) is the number of times \(\lambda - t\) divides \(\det(A - tI_n)\).


An an example if \(A = \begin{pmatrix} 1 & 1 \\ 0 & 1 \end{pmatrix}\) then, \(\lambda = 1\) has algebraic multiplicity 2 because \(\det(A - I_n) = (1-t)^2\).

Definition
The geometric multiplicity of \(\lambda\) is \(\dim(E_{\lambda})\).


For the same example above. The geometric multiplicity of \(\lambda = 1\) is 1 because the nullspace is spanned by one vector.

Given these definitions we can now introduce the following theorem.

Theorem 2
Geometric multiplicity of \(\lambda \leq\) the algebraic multiplicity of \(\lambda\).


Proof
Let \(\lambda\) be an eigenvalue of \(A \in M_{n \times n}\) and let the geometric multiplicity of \(\lambda\) be \(k\). Now the goal is to relate the characteristic polynomial of \(\lambda\) to the dimension of \(E_{\lambda}\).

Let \(\{v_1,...,v_k\}\) be a basis of \(E_{\lambda}\). We know that \(k \leq n\) so extend this basis to a basis for \(\mathbf{R}^n\).



Theorem
\(A\) is diagonalizable if and only if
  • \(\det(A - tI_n)\) splits over \(\mathbf{R}\)
  • For each eigenvalue \(\lambda\), geometric multiplicity = algebraic multiplictiy





References

  • Video Lectures from Math416 by Ely Kerman.