Generalised Eigenstuff

Let V be a vector space over a field \mathbb K, and T : V \to V be a linear transformation.

Recall that a nonzero vector \mathbf v \in V is an eigenvector of T : V \to V with eigenvalue \lambda \in \mathbb K if T(\mathbf v) = \lambda \mathbf v. Equivalently, \mathbf v \in \ker(\lambda I_V - T). If V is finite dimensional with \dim(V) = n, then V has a basis K of n linearly independent vectors and T is similar to the linear transformation [T]_K : \mathbb K^n \to \mathbb K^n. Then \mathbf \lambda is an eigenvalue of T if and only if c_T(\lambda) \equiv c_{[T]_K}(\lambda) = 0. We call c_T the characteristic polynomial of T. By the Cayley-Hamilton theorem, c_T(T) = O_V.

Let p \in \mathbb K[x] be a monic polynomial (i.e. the highest power of x has coefficient 1). If p = c_T, then p(T) = O_V. What is the “smallest” polynomial p such that p(T) = O_V? We call this the minimal polynomial, and formulate it rigorously as follows.

Let V be a finite-dimensional.

Theorem 1. For any linear transformation T : V \to V, there exists a unique monic polynomial, denoted m_T, such that m_T(T) = O_V, and for any p \in \mathbb K[x] such that p(T) = O_V, m_T \mid p.

Proof. By the Cayley-Hamilton theorem, there exists a monic polynomial p = c_T \in \mathbb K[x] such that p(T) = O_V. By the well-ordering principle, the set

\{ n \in \mathbb N : (\exists p \in \mathbb K_n[x]\backslash \mathbb K_{n-1}[x]\quad\! p(T) = 0) \} \subseteq \mathbb N^+

has a minimum element m. Consider the monic polynomials p_1 \in \mathbb K_{m}[x]\backslash \mathbb K_{m-1}[x], p_2 \in \mathbb K[x] such that p_1(T) = p_2(T) = O_V. Use polynomial long division to compute q \in \mathbb K[x], r \in \mathbb K_{m-1}[x] such that

p_2 = q \cdot p_1 + r \quad \Rightarrow \quad r = p_2 - q \cdot p_1.

Therefore, r(T) = 0. By definition of p_1,p_2, r = 0. For uniqueness, if p_2 \in \mathbb K_{m-1}[x] is monic, then p_1 \mid p_2 and p_2 \mid p_1. Find polynomials q_1, q_2 \in \mathbb K[x] such that

p_2 = p_1 \cdot q_1\quad \text{and} \quad p_1 = p_2 \cdot q_2.

Combining both decompositions, q_1 \cdot q_2 = 1. Since p_1,p_2 are monic, so are q_1, q_2, so that q_1 = q_2 = 1, and hence, p_1 = p_2.

Suppose T has distinct eigenvalues \lambda_1,\dots,\lambda_k.

Theorem 2. If T has characteristic polynomial

c_T(x) = (x - \lambda_1)^{\alpha_1} \cdot \cdots \cdot (x-\lambda_k)^{\alpha_k},

then there exist unique integers 0 \leq \beta_i \leq \alpha_i such that T has minimal polynomial

m_T(x) = (x - \lambda_1)^{\beta_1} \cdot \cdots \cdot (x-\lambda_k)^{\beta_k},\quad 1 \leq \beta_i \leq \alpha_i.

For each eigenvalue \lambda_i, define the subspace K_{\lambda_i} := \ker((T - \lambda_i)^{\beta_i}), where we abbreviate T - \lambda I_V \equiv T - \lambda. Note that if \beta_i = 1, then E_{\lambda_i} = K_{\lambda_i}. Then

V = K_{\lambda_1} \oplus \cdots \oplus K_{\lambda_k}.

Proof. For the first claim, by definition, m_T \mid c_T, and

m_T(x) = (x - \lambda_1)^{\beta_1} \cdot \cdots \cdot (x-\lambda_k)^{\beta_k},\quad 0 \leq \beta_i \leq \alpha_i.

It suffices to prove that \beta_1 \geq 1 without loss of generality. Let \mathbf w \in E_{\lambda_1} and define W := \mathrm{span}\{\mathbf w\}. Then (T|_W - \lambda_1) = O_W. Hence,

(x-\lambda_1) \mid m_{T|_W}(x) \quad \text{and}\quad m_{T|_W}(x) \mid (x - \lambda_1),

so that m_{T|_W}(x) = x-\lambda_1. Furthermore, it is not hard to see by definition that m_{T|_W} \mid m_T. For the second claim, fix \mathbf v \in V. Then (m_T(T))(\mathbf v) = \mathbf 0. If \mathbf v \notin K_{\lambda_i} for any i, then (m_T(T))(\mathbf v) \neq 0, a contradiction.

Corollary 1. Under the conditions in Theorem 2, the following propositions hold:

  • E_{\lambda_i} \subseteq K_{\lambda_i},
  • K_{\lambda_i} is T-invariant,
  • m_{ T|_{ K_{\lambda_i} } } = (x - \lambda_i)^{\beta_i},
  • c_{ T|_{ K_{\lambda_i} } } = (x - \lambda_i)^{\alpha_i},
  • \dim(K_{\lambda_i}) = \alpha_i.

Proof. Use an intermediate result that for T-invariant subspaces W_1, W_2 \subseteq V,

m_{T|_{W_1 + W_2}} = \mathrm{lcm}\{ m_{T|_{W_1}}, m_{T|_{W_2}} \}.

Corollary 2. Under the conditions in Theorem 2, the following propositions are equivalent:

  • T is diagonalisable,
  • m_T(x) = (x-\lambda_1)\cdots (x - \lambda_k),
  • for each i, \dim(K_{\lambda_i}) = \alpha_i,
  • V = E_{\lambda_1} \oplus \cdots \oplus E_{\lambda_k}.

Recall that E_{\lambda_i} = K_{\lambda_i} when \beta_i = 1. In this regard, any vector in K_{\lambda_i} = \ker( (T - \lambda_i)^{\beta_i} ) is a generalised eigenvector of T with generalised eigenvalue \lambda_i. That is, (T - \lambda_i)^{\beta_i}(\mathbf v) = 0, which reduces to an “actual” eigenvector precisely when \beta_i = 1. Now, if

\mathbb V = E_{\lambda_1} \oplus \cdots \oplus E_{\lambda_k},

then T is diagonalisable and vice-versa. That means T is similar to a diagonal matrix [T]_K for some basis K of V, which is a relatively simple transformation. On the other hand, in general, if c_T can be factored into linear factors, then

V = K_{\lambda_1} \oplus \cdots \oplus K_{\lambda_k}.

Based on this decomposition, would there a suitably defined (and sufficiently simple) matrix (i.e. linear transformation) [T]_K for some basis K of V such that T is similar to [T]_K? The search for this answer will lead us to the climatic conclusion of this eigen-discussion—the Jordan normal form.

—Joel Kindiak, 11 Mar 25, 1615H

,

Published by


Leave a comment