Nilpotent matrix

From HandWiki
Short description: Mathematical concept in algebra

In linear algebra, a nilpotent matrix is a square matrix N such that

[math]\displaystyle{ N^k = 0\, }[/math]

for some positive integer [math]\displaystyle{ k }[/math]. The smallest such [math]\displaystyle{ k }[/math] is called the index of [math]\displaystyle{ N }[/math],[1] sometimes the degree of [math]\displaystyle{ N }[/math].

More generally, a nilpotent transformation is a linear transformation [math]\displaystyle{ L }[/math] of a vector space such that [math]\displaystyle{ L^k = 0 }[/math] for some positive integer [math]\displaystyle{ k }[/math] (and thus, [math]\displaystyle{ L^j = 0 }[/math] for all [math]\displaystyle{ j \geq k }[/math]).[2][3][4] Both of these concepts are special cases of a more general concept of nilpotence that applies to elements of rings.

Examples

Example 1

The matrix

[math]\displaystyle{ A = \begin{bmatrix} 0 & 1 \\ 0 & 0 \end{bmatrix} }[/math]

is nilpotent with index 2, since [math]\displaystyle{ A^2 = 0 }[/math].

Example 2

More generally, any [math]\displaystyle{ n }[/math]-dimensional triangular matrix with zeros along the main diagonal is nilpotent, with index [math]\displaystyle{ \le n }[/math][citation needed]. For example, the matrix

[math]\displaystyle{ B=\begin{bmatrix} 0 & 2 & 1 & 6\\ 0 & 0 & 1 & 2\\ 0 & 0 & 0 & 3\\ 0 & 0 & 0 & 0 \end{bmatrix} }[/math]

is nilpotent, with

[math]\displaystyle{ B^2=\begin{bmatrix} 0 & 0 & 2 & 7\\ 0 & 0 & 0 & 3\\ 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0 \end{bmatrix} ;\ B^3=\begin{bmatrix} 0 & 0 & 0 & 6\\ 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0 \end{bmatrix} ;\ B^4=\begin{bmatrix} 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0 \end{bmatrix}. }[/math]

The index of [math]\displaystyle{ B }[/math] is therefore 4.

Example 3

Although the examples above have a large number of zero entries, a typical nilpotent matrix does not. For example,

[math]\displaystyle{ C=\begin{bmatrix} 5 & -3 & 2 \\ 15 & -9 & 6 \\ 10 & -6 & 4 \end{bmatrix} \qquad C^2=\begin{bmatrix} 0 & 0 & 0 \\ 0 & 0 & 0 \\ 0 & 0 & 0 \end{bmatrix} }[/math]

although the matrix has no zero entries.

Example 4

Additionally, any matrices of the form

[math]\displaystyle{ \begin{bmatrix} a_1 & a_1 & \cdots & a_1 \\ a_2 & a_2 & \cdots & a_2 \\ \vdots & \vdots & \ddots & \vdots \\ -a_1-a_2-\ldots-a_{n-1} & -a_1-a_2-\ldots-a_{n-1} & \ldots & -a_1-a_2-\ldots-a_{n-1} \end{bmatrix} }[/math]

such as

[math]\displaystyle{ \begin{bmatrix} 5 & 5 & 5 \\ 6 & 6 & 6 \\ -11 & -11 & -11 \end{bmatrix} }[/math]

or

[math]\displaystyle{ \begin{bmatrix} 1 & 1 & 1 & 1 \\ 2 & 2 & 2 & 2 \\ 4 & 4 & 4 & 4 \\ -7 & -7 & -7 & -7 \end{bmatrix} }[/math]

square to zero.

Example 5

Perhaps some of the most striking examples of nilpotent matrices are [math]\displaystyle{ n\times n }[/math] square matrices of the form:

[math]\displaystyle{ \begin{bmatrix} 2 & 2 & 2 & \cdots & 1-n \\ n+2 & 1 & 1 & \cdots & -n \\ 1 & n+2 & 1 & \cdots & -n \\ 1 & 1 & n+2 & \cdots & -n \\ \vdots & \vdots & \vdots & \ddots & \vdots \end{bmatrix} }[/math]

The first few of which are:

[math]\displaystyle{ \begin{bmatrix} 2 & -1 \\ 4 & -2 \end{bmatrix} \qquad \begin{bmatrix} 2 & 2 & -2 \\ 5 & 1 & -3 \\ 1 & 5 & -3 \end{bmatrix} \qquad \begin{bmatrix} 2 & 2 & 2 & -3 \\ 6 & 1 & 1 & -4 \\ 1 & 6 & 1 & -4 \\ 1 & 1 & 6 & -4 \end{bmatrix} \qquad \begin{bmatrix} 2 & 2 & 2 & 2 & -4 \\ 7 & 1 & 1 & 1 & -5 \\ 1 & 7 & 1 & 1 & -5 \\ 1 & 1 & 7 & 1 & -5 \\ 1 & 1 & 1 & 7 & -5 \end{bmatrix} \qquad \ldots }[/math]

These matrices are nilpotent but there are no zero entries in any powers of them less than the index.[5]

Example 6

Consider the linear space of polynomials of a bounded degree. The derivative operator is a linear map. We know that applying the derivative to a polynomial decreases its degree by one, so when applying it iteratively, we will eventually obtain zero. Therefore, on such a space, the derivative is representable by a nilpotent matrix.

Characterization

For an [math]\displaystyle{ n \times n }[/math] square matrix [math]\displaystyle{ N }[/math] with real (or complex) entries, the following are equivalent:

  • [math]\displaystyle{ N }[/math] is nilpotent.
  • The characteristic polynomial for [math]\displaystyle{ N }[/math] is [math]\displaystyle{ \det \left(xI - N\right) = x^n }[/math].
  • The minimal polynomial for [math]\displaystyle{ N }[/math] is [math]\displaystyle{ x^k }[/math] for some positive integer [math]\displaystyle{ k \leq n }[/math].
  • The only complex eigenvalue for [math]\displaystyle{ N }[/math] is 0.

The last theorem holds true for matrices over any field of characteristic 0 or sufficiently large characteristic. (cf. Newton's identities)

This theorem has several consequences, including:

  • The index of an [math]\displaystyle{ n \times n }[/math] nilpotent matrix is always less than or equal to [math]\displaystyle{ n }[/math]. For example, every [math]\displaystyle{ 2 \times 2 }[/math] nilpotent matrix squares to zero.
  • The determinant and trace of a nilpotent matrix are always zero. Consequently, a nilpotent matrix cannot be invertible.
  • The only nilpotent diagonalizable matrix is the zero matrix.

See also: Jordan–Chevalley decomposition.

Classification

Consider the [math]\displaystyle{ n \times n }[/math] (upper) shift matrix:

[math]\displaystyle{ S = \begin{bmatrix} 0 & 1 & 0 & \ldots & 0 \\ 0 & 0 & 1 & \ldots & 0 \\ \vdots & \vdots & \vdots & \ddots & \vdots \\ 0 & 0 & 0 & \ldots & 1 \\ 0 & 0 & 0 & \ldots & 0 \end{bmatrix}. }[/math]

This matrix has 1s along the superdiagonal and 0s everywhere else. As a linear transformation, the shift matrix "shifts" the components of a vector one position to the left, with a zero appearing in the last position:

[math]\displaystyle{ S(x_1,x_2,\ldots,x_n) = (x_2,\ldots,x_n,0). }[/math][6]

This matrix is nilpotent with degree [math]\displaystyle{ n }[/math], and is the canonical nilpotent matrix.

Specifically, if [math]\displaystyle{ N }[/math] is any nilpotent matrix, then [math]\displaystyle{ N }[/math] is similar to a block diagonal matrix of the form

[math]\displaystyle{ \begin{bmatrix} S_1 & 0 & \ldots & 0 \\ 0 & S_2 & \ldots & 0 \\ \vdots & \vdots & \ddots & \vdots \\ 0 & 0 & \ldots & S_r \end{bmatrix} }[/math]

where each of the blocks [math]\displaystyle{ S_1,S_2,\ldots,S_r }[/math] is a shift matrix (possibly of different sizes). This form is a special case of the Jordan canonical form for matrices.[7]

For example, any nonzero 2 × 2 nilpotent matrix is similar to the matrix

[math]\displaystyle{ \begin{bmatrix} 0 & 1 \\ 0 & 0 \end{bmatrix}. }[/math]

That is, if [math]\displaystyle{ N }[/math] is any nonzero 2 × 2 nilpotent matrix, then there exists a basis b1b2 such that Nb1 = 0 and Nb2 = b1.

This classification theorem holds for matrices over any field. (It is not necessary for the field to be algebraically closed.)

Flag of subspaces

A nilpotent transformation [math]\displaystyle{ L }[/math] on [math]\displaystyle{ \mathbb{R}^n }[/math] naturally determines a flag of subspaces

[math]\displaystyle{ \{0\} \subset \ker L \subset \ker L^2 \subset \ldots \subset \ker L^{q-1} \subset \ker L^q = \mathbb{R}^n }[/math]

and a signature

[math]\displaystyle{ 0 = n_0 \lt n_1 \lt n_2 \lt \ldots \lt n_{q-1} \lt n_q = n,\qquad n_i = \dim \ker L^i. }[/math]

The signature characterizes [math]\displaystyle{ L }[/math] up to an invertible linear transformation. Furthermore, it satisfies the inequalities

[math]\displaystyle{ n_{j+1} - n_j \leq n_j - n_{j-1}, \qquad \mbox{for all } j = 1,\ldots,q-1. }[/math]

Conversely, any sequence of natural numbers satisfying these inequalities is the signature of a nilpotent transformation.

Additional properties

  • If [math]\displaystyle{ N }[/math] is nilpotent of index [math]\displaystyle{ k }[/math] , then [math]\displaystyle{ I+N }[/math] and [math]\displaystyle{ I-N }[/math] are invertible, where [math]\displaystyle{ I }[/math] is the [math]\displaystyle{ n \times n }[/math] identity matrix. The inverses are given by
    [math]\displaystyle{ \begin{align} (I + N)^{-1} &= \displaystyle\sum^k_{m=0}\left(-N\right)^m = I - N + N^2 - N^3 + N^4 - N^5 + N^6 - N^7 + \cdots +(-N)^k \\ (I - N)^{-1} &= \displaystyle\sum^k_{m=0}N^m = I + N + N^2 + N^3 + N^4 + N^5 + N^6 + N^7 + \cdots + N^k \\ \end{align} }[/math]
  • If [math]\displaystyle{ N }[/math] is nilpotent, then
    [math]\displaystyle{ \det (I + N) = 1. }[/math]

    Conversely, if [math]\displaystyle{ A }[/math] is a matrix and

    [math]\displaystyle{ \det (I + tA) = 1\!\, }[/math]
    for all values of [math]\displaystyle{ t }[/math], then [math]\displaystyle{ A }[/math] is nilpotent. In fact, since [math]\displaystyle{ p(t) = \det (I + tA) - 1 }[/math] is a polynomial of degree [math]\displaystyle{ n }[/math], it suffices to have this hold for [math]\displaystyle{ n+1 }[/math] distinct values of [math]\displaystyle{ t }[/math].
  • Every singular matrix can be written as a product of nilpotent matrices.[8]
  • A nilpotent matrix is a special case of a convergent matrix.

Generalizations

A linear operator [math]\displaystyle{ T }[/math] is locally nilpotent if for every vector [math]\displaystyle{ v }[/math], there exists a [math]\displaystyle{ k\in\mathbb{N} }[/math] such that

[math]\displaystyle{ T^k(v) = 0.\!\, }[/math]

For operators on a finite-dimensional vector space, local nilpotence is equivalent to nilpotence.

Notes

  1. (Herstein 1975)
  2. (Beauregard Fraleigh)
  3. (Herstein 1975)
  4. (Nering 1970)
  5. Mercer, Idris D. (31 October 2005). "Finding "nonobvious" nilpotent matrices". self-published; personal credentials: PhD Mathematics, Simon Fraser University. http://www.idmercer.com/nilpotent.pdf. 
  6. (Beauregard Fraleigh)
  7. (Beauregard Fraleigh)
  8. R. Sullivan, Products of nilpotent matrices, Linear and Multilinear Algebra, Vol. 56, No. 3

References

External links