Jordan–Chevalley decomposition

From Wikipedia, the free encyclopedia

In mathematics, the Jordan–Chevalley decomposition, named after Camille Jordan and Claude Chevalley, expresses a linear operator as the sum of its commuting semisimple part and its nilpotent part. The multiplicative decomposition expresses an invertible operator as the product of its commuting semisimple and unipotent parts. The decomposition is easy to describe when the Jordan normal form of the operator is given, but it exists under weaker hypotheses than the existence of a Jordan normal form. Analogues of the Jordan-Chevalley decomposition exist for elements of linear algebraic groups, Lie algebras, and Lie groups, and the decomposition is an important tool in the study of these objects.

Decomposition of a linear operator[]

Consider linear operators on a finite-dimensional vector space over a field. An operator T is semisimple if every T-invariant subspace has a complementary T-invariant subspace (if the underlying field is algebraically closed, this is the same as the requirement that the operator be diagonalizable). An operator x is nilpotent if some power xm of it is the zero operator. An operator x is unipotent if x − 1 is nilpotent.

Now, let x be any operator. A Jordan–Chevalley decomposition of x is an expression of it as a sum

x = xs + xn,

where xs is semisimple, xn is nilpotent, and xs and xn commute. Over a perfect field,[1] such a decomposition exists (cf. #Proof of uniqueness and existence), the decomposition is unique, and the xs and xn are polynomials in x with no constant terms.[2][3] In particular, for any such decomposition over a perfect field, an operator that commutes with x also commutes with xs and xn.

If x is an invertible operator, then a multiplicative Jordan–Chevalley decomposition expresses x as a product

x = xs · xu,

where xs is semisimple, xu is unipotent, and xs and xu commute. Again, over a perfect field, such a decomposition exists, the decomposition is unique, and xs and xu are polynomials in x. The multiplicative version of the decomposition follows from the additive one since, as is easily seen to be invertible,

and is unipotent. (Conversely, by the same type of argument, one can deduce the additive version from the multiplicative one.)

If x is written in Jordan normal form (with respect to some basis) then xs is the endomorphism whose matrix contains just the diagonal terms of x, and xn is the endomorphism whose matrix contains just the off-diagonal terms; xu is the endomorphism whose matrix is obtained from the Jordan normal form by dividing all entries of each Jordan block by its diagonal element.

Proof of uniqueness and existence[]

The uniqueness follows from the fact are polynomial in x: if is another decomposition such that and commute, then , and both commute with x, hence with . Now, the sum of commuting semisimple (resp. nilpotent) endomorphisms is again semisimple (resp. nilpotent). Since the only operator which is both semisimple and nilpotent is the zero operator it follows that and .

We show the existence. Let V be a finite-dimensional vector space over a perfect field k and an endomorphism.

First assume the base field k is algebraically closed. Then the vector space V has the direct sum decomposition where each is the kernel of , the generalized eigenspace and x stabilizes , meaning . Now, define so that, on each , it is the scalar multiplication by . Note that, in terms of a basis respecting the direct sum decomposition, is a diagonal matrix; hence, it is a semisimple endomorphism. Since is then whose -th power is zero, we also have that is nilpotent, establishing the existence of the decomposition.

(Choosing a basis carefully on each , one can then put x in the Jordan normal form and are the diagonal and the off-diagonal parts of the normal form. But this is not needed here.)

The fact that are polynomials in x follows from the Chinese remainder theorem. Indeed, let be the characteristic polynomial of x. Then it is the product of the characteristic polynomials of ; i.e., Also, (because, in general, a nilpotent matrix is killed when raised to the size of the matrix). Now, the Chinese remainder theorem applied to the polynomial ring gives a polynomial satisfying the conditions

(for all i).

(There is a redundancy in the conditions if some is zero but that is not an issue; just remove it from the conditions.)

The condition , when spelled out, means that for some polynomial . Since is the zero map on , and agree on each ; i.e., . Also then with . The condition ensures that and have no constant terms. This completes the proof of the algebraically closed field case.

If k is an arbitrary perfect field, let be the absolute Galois group of k. By the first part, we can choose polynomials over such that is the decomposition into the semisimple and nilpotent part. For each in ,

Now, is a polynomial in ; so is . Thus, and commute. Also, the application of evidently preserves semisimplicity and nilpotency. Thus, by the uniqueness of decomposition (over ), and . Hence, are -invariant; i.e., they are endomorphisms (represented by matrices) over k. Finally, since contains a -basis that spans the space containing , by the same argument, we also see that have coefficients in k. This completes the proof.

Short proof using abstract algebra[]

(Jacobson 1979) proves the existence of a decomposition as a consequence of the Wedderburn principal theorem. (This approach is not only short but also makes the role of the assumption that the base field be perfect clearer.)

Let V be a finite-dimensional vector space over a perfect field k, an endomorphism and the subalgebra generated by x. Note that A is a commutative Artinian ring. The Wedderburn principal theorem states: for a finite-dimensional algebra A with the Jacobson radical J, if is separable, then the natural surjection splits; i.e., contains a semisimple subalgebra such that is an isomorphism.[4] In the setup here, is separable since the base field is perfect (so the theorem is applicable) and J is also the nilradical of A. There is then the vector-space decomposition . In particular, the endomorphism x can be written as where is in and in . Now, the image of x generates ; thus is semisimple and is a polynomial of x. Also, is nilpotent since is nilpotent and is a polynomial of x since is.

Nilpotency criterion[]

The Jordan decomposition can be used to characterize nilpotency of an endomorphism. Let k be an algebraically closed field of characteristic zero, the endomorphism ring of k over rational numbers and V a finite-dimensional vector space over k. Given an endomorphism , let be the Jordan decomposition. Then is diagonalizable; i.e., where each is the eigenspace for eigenvalue with multiplicity . Then for any let be the endomorphism such that is the multiplication by . Chevalley calls the replica of given by . (For example, if , then the complex conjugate of an endomorphism is an example of a replica.) Now,

Nilpotency criterion — [5] is nilpotent (i.e., ) if and only if for every . Also, if , then it suffices the condition holds for complex conjugation.

Proof: First, since is nilpotent,

.

If is the complex conjugation, this implies for every i. Otherwise, take to be a -linear functional followed by . Applying that to the above equation, one gets:

and, since are all real numbers, for every i. Varying the linear functionals then implies for every i.

A typical application of the above criterion is the proof of Cartan's criterion for solvability of a Lie algebra. It says: if is a Lie subalgebra over a field k of characteristic zero such that for each , then is solvable.

Proof:[6] Without loss of generality, assume k is algebraically closed. By Lie's theorem and Engel's theorem, it suffices to show for each , is a nilpotent endomorphism of V. Write . Then we need to show:

is zero. Let . Note we have: and, since is the semisimple part of the Jordan decomposition of , it follows that is a polynomial without constant term in ; hence, and the same is true with in place of . That is, , which implies the claim given the assumption.

Counterexample to existence over an imperfect field[]

If the ground field is not perfect, then a Jordan–Chevalley decomposition may not exist. Example: Let p be a prime number, let be imperfect of characteristic , and choose in that is not a th power. Let , let and let be the -linear operator given by multiplication by in . This has as its invariant -linear subspaces precisely the ideals of viewed as a ring, which correspond to the ideals of containing . Since is irreducible in , ideals of V are , and . Suppose for commuting -linear operators and that are respectively semisimple (just over , which is weaker than semisimplicity over an algebraic closure of ) and nilpotent. Since and commute, they each commute with and hence each acts -linearly on . Therefore and are each given by multiplication by respective members of and , with . Since is nilpotent, is nilpotent in , therefore in , for is a field. Hence, , therefore for some polynomial . Also, we see that . Since is of characteristic , we have . Also, since in , we have , therefore in . Since , we have . Combining these results we get . This shows that generates as a -algebra and thus the -stable -linear subspaces of are ideals of , i.e. they are , and . We see that is an -invariant subspace of which has no complement -invariant subspace, contrary to the assumption that is semisimple. Thus, there is no decomposition of as a sum of commuting -linear operators that are respectively semisimple and nilpotent. Note that minimal polynomial of is inseparable over and is a square in . It can be shown that if minimal polynomial of linear operator is separable then has Jordan-Chevalley decomposition and that if this polynomial is product of distinct irreducible polynomials in , then is semisimple over .

Analogous decompositions[]

The multiplicative version of the Jordan-Chevalley decomposition generalizes to a decomposition in a linear algebraic group, and the additive version of the decomposition generalizes to a decomposition in a Lie algebra.

Lie algebras[]

Let denote the Lie algebra of the endomorphisms of a finite-dimensional vector space V over a perfect field. If is the Jordan decomposition, then is the Jordan decomposition of on the vector space . Indeed, first, and commute since . Second, in general, for each endomorphism , we have:

  1. If , then , since is the difference of the left and right multiplications by y.
  2. If is semisimple, then is semisimple.[7]

Hence, by uniqueness, and .

If is a finite-dimensional representation of a semisimple finite-dimensional complex Lie algebra, then preserves the Jordan decomposition in the sense: if , then and .[8]

Real semisimple Lie algebras[]

In the formulation of Chevalley and Mostow, the additive decomposition states that an element X in a real semisimple Lie algebra g with Iwasawa decomposition g = kan can be written as the sum of three commuting elements of the Lie algebra X = S + D + N, with S, D and N conjugate to elements in k, a and n respectively. In general the terms in the Iwasawa decomposition do not commute.

Linear algebraic groups[]

Let be a linear algebraic group over a perfect field. Then, essentially by definition, there is a closed embedding . Now, to each element , by the multiplicative Jordan decomposition, there are a pair of a semisimple element and a unipotent element a priori in such that . But, as it turns out,[9] the elements can be shown to be in (i.e., they satisfy the defining equations of G) and that they are independent of the embedding into ; i.e., the decomposition is intrinsic.

When G is abelian, is then the direct product of the closed subgroup of the semisimple elements in G and that of unipotent elements.[10]

Real semisimple Lie groups[]

The multiplicative decomposition states that if g is an element of the corresponding connected semisimple Lie group G with corresponding Iwasawa decomposition G = KAN, then g can be written as the product of three commuting elements g = sdu with s, d and u conjugate to elements of K, A and N respectively. In general the terms in the Iwasawa decomposition g = kan do not commute.

References[]

  1. ^ In fact, the proof goes through if the quotient is a separable algebra; see #Short proof using abstract algebra.
  2. ^ Humphreys 1972, Prop. 4.2, p. 17 for the algebraically closed field case.
  3. ^ Waterhouse, Ch. 9, Exercise 1.
  4. ^ Ring Theory. 18 April 1972. ISBN 9780080873572.
  5. ^ Serre, LA 5.17. Lemma 6.7. The endomorphism
  6. ^ Serre, LA 5.19. Theorem 7.1.
  7. ^ This is not easy to see but is shown in the proof of (Jacobson, Ch. III, § 7, Theorem 11.). Editorial note: we need to add a discussion of this matter to "semisimple operator".
  8. ^ Fulton & Harris, Theorem 9.20.
  9. ^ Waterhouse, Theorem 9.2.
  10. ^ Waterhouse, Theorem 9.3.
Retrieved from ""