Jordan–Chevalley decomposition: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
Citation bot (talk | contribs)
Alter: author1-link. Add: publisher. | Use this bot. Report bugs. | Suggested by Whoop whoop pull up | Category:Linear algebra | #UCB_Category 43/287
EPhrimal (talk | contribs)
Rewrite of almost entire article to (1) include potentially diagonalisable operators in discussion (2) add motivation and context (3) follow Wikipedia style more.
Line 1: Line 1:
In mathematics, the '''Jordan–Chevalley decomposition''', named after [[Camille Jordan]] and [[Claude Chevalley]], expresses a [[linear operator]] as the sum of its commuting [[Semi-simple operator|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 group]]s, [[Lie algebra]]s, and [[Lie group]]s, and the decomposition is an important tool in the study of these objects.
In [[mathematics]], specifically [[linear algebra]], the '''Jordan–Chevalley decomposition''', named after [[Camille Jordan]] and [[Claude Chevalley]], expresses a [[linear operator]] in a unique way as the sum of two other linear operators which are simpler to understand. Specifically, one part is [[Diagonalizable matrix|potentially diagonalisable]] and the other is [[Nilpotent matrix|nilpotent]]. The two parts are [[polynomial]]s in the operator, which makes them behave nicely in algebraic manipulations.

The decomposition has a short description when the [[Jordan normal form]] of the operator is given, but it exists under weaker hypotheses than are needed for the existence of a Jordan normal form. Hence the Jordan-Chevalley decomposition can be seen as a generalisation of the Jordan normal form, which is also reflected in many proofs of it.

It is closely related to the [[Wedderburn principal theorem]] about [[associative algebra]]s, which also leads to several analogues in [[Lie algebra]]s. Analogues of the Jordan-Chevalley decomposition also exist for elements of [[Linear algebraic group]]s and [[Lie group]]s via a multiplicative reformulation. The decomposition is an important tool in the study of all of these objects, and was developed for this purpose.

In many texts, the potentially diagonalisable part is also characterised as the [[Semi-simple operator|semisimple]] part.

== Introduction ==

A basic question in linear algebra is whether an operator on a finite-dimensional [[vector space]] can be [[Diagonalizable matrix|diagonalised]]. For example, this is closely related to the [[eigenvalue]]s of the operator. In several contexts, one may be dealing with many operators which are not diagonalisable. Even over an algebraically closed field, a diagonalisation may not exist. In this context, the [[Jordan normal form]] achieves the best possible result to obtain something similar to a diagonalisation. For linear operators over a [[field]] which is not [[algebraically closed field|algebraically closed]], there may be no eigenvector at all. The latter point is not the main concern in discussions on the Jordan-Chevalley decomposition. To avoid this problem, one can instead consider potentially diagonalisable operators, which are those that admit a diagonalisation over some field (or equivalently over the [[algebraic closure]] of the field under consideration).

The operators which are „the furthest away“ from being diagonalisable are [[Nilpotent matrix|nilpotent operators]]. An operator (or more generally an element of a [[ring]]) <math> x </math> is said to be nilpotent when there is some positive integer <math> m \geq 1 </math> such that <math> x^m = 0 </math>. In several contexts in [[abstract algebra]], it is the case that the presence of nilpotent elements of a ring make them much more complicated to work with.{{Citation needed|date=January 2024}} To a large extent, this is also the case for linear operators. The remarkable fact is that one can „separate out“ the nilpotent part of an operator which causes it to be not potentially diagonalisable. This is what is described by the Jordan-Chevalley decomposition.

Historically, the Jordan-Chevalley decomposition was motivated by the applications to the theory of [[Lie algebra|Lie algebras]] and [[linear algebraic group]]s<ref>{{Cite journal |last=Danielle Couty, Jean Esterle, Rachid Zarouf. |date=2010. |title=Décomposition effective de Jordan-Chevalley et ses retombées en enseignement. |url=https://hal.science/file/index/docid/525465/filename/Jordan-Chevalley.pdf |journal= |via=HAL}}</ref><ref name=":0">{{Cite journal |last=Couty |first=Danielle |last2=Esterle |first2=Jean |last3=Zarouf |first3=Rachid |date=2011, 129, pp.29–49. |title=Décomposition effective de Jordan-Chevalley et ses retombées en enseignement |url=https://hal.science/hal-00525465/file/Jordan-Chevalley.pdf |journal=Gazette des Mathématiciens |pages=15-19 |via=HAL}}</ref>, as described in sections below.


== Decomposition of a linear operator ==
== Decomposition of a linear operator ==
Consider linear operators on a finite-dimensional [[vector space]] over a field. An operator <math> T </math> is [[Semisimple operator|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 matrix|diagonalizable]]). An operator ''x'' is ''[[nilpotent]]'' if some power ''x''<sup>''m''</sup> of it is the zero operator. An operator ''x'' is ''[[unipotent]]'' if ''x''&nbsp;−&nbsp;1 is nilpotent.


Let <math> K </math> be a [[field]], <math> V </math> a finite-dimensional [[vector space]] over <math> K </math>, and <math> T </math> a linear operator over <math> V </math> (equivalently, a [[matrix]] with entries from <math> K </math>). If the [[minimal polynomial]] of <math> T </math> splits over <math> K </math> (for example if <math> K </math> is algebraically closed), then <math> T </math> has a [[Jordan normal form]] <math> T = SJS^{-1} </math>. If <math> D </math> is the diagonal of <math> J </math>, let <math> R = J - D </math> be the remaining part. Then <math> T = SDS^{-1} + SRS^{-1} </math> is a decomposition where <math> SDS^{-1} </math> is diagonalisable and <math> SRS^{-1} </math> is nilpotent. This restatement of the normal form as an additive decomposition not only makes the numerical computation more stable{{Citation needed|date=January 2024}}, but can be generalised to cases where the minimal polynomial of <math> T </math> does not split.
Now, let ''x'' be any operator. A Jordan–Chevalley decomposition of ''x'' is an expression of it as a sum
:''x'' = ''x''<sub>''s''</sub> + ''x''<sub>''n''</sub>,


If the minimal polynomial of <math> T </math> splits into \textit{distinct} linear factors, then <math> T </math> is diagonalisable. Therefore, if the minimal polynomial of <math> T </math> is at least [[separable polynomial|separable]], then <math> T </math> is potentially diagonalisable. The Jordan-Chevalley decomposition is concerned with the more general case where the minimal polynomial of <math> T </math> is a product of separable polynomials.
where ''x''<sub>''s''</sub> is semisimple, ''x''<sub>''n''</sub> is nilpotent, and ''x''<sub>''s''</sub> and ''x''<sub>''n''</sub> commute. Over a [[perfect field]],<ref>In fact, the proof goes through if the quotient <math>k[x]/\text{nil}</math> is a separable algebra; see [[#Short proof using abstract algebra]].</ref> such a decomposition exists (cf. [[#Proof of uniqueness and existence]]), the decomposition is unique, and the ''x''<sub>''s''</sub> and ''x''<sub>''n''</sub> are polynomials in ''x'' with no constant terms.<ref name=Humphreys>{{harvnb|Humphreys|1972|loc=Prop. 4.2, p. 17}} for the algebraically closed field case.</ref><ref>{{harvnb|Waterhouse|1979|loc=Ch. 9, Exercise 1.}}</ref> In particular, for any such decomposition over a perfect field, an operator that commutes with ''x'' also commutes with ''x''<sub>''s''</sub> and ''x''<sub>''n''</sub>.


If ''x'' is an invertible operator, then a multiplicative Jordan–Chevalley decomposition expresses ''x'' as a product
Let <math> x: V \to V </math> be any linear operator on the finite-dimensional vector space <math> V </math> over the field <math> K </math>. A Jordan–Chevalley decomposition of <math> x </math> is an expression of it as a sum
:<math> x = x_s + x_n </math> ,
:''x'' = ''x''<sub>''s''</sub> · ''x''<sub>u</sub>,
where <math> x_s </math> is potentially diagonalisable, <math> x_n </math> is nilpotent, and <math> x_s x_n = x_n x_s </math>.


{{Math theorem|name=Jordan-Chevalley decomposition|math_statement=
where ''x''<sub>''s''</sub> is semisimple, ''x''<sub>u</sub> is unipotent, and ''x''<sub>''s''</sub> and ''x''<sub>''u''</sub> commute. Again, over a perfect field, such a decomposition exists, the decomposition is unique, and ''x''<sub>''s''</sub> and ''x''<sub>u</sub> are polynomials in ''x''. The multiplicative version of the decomposition follows from the additive one since, as <math>x_s</math> is easily seen to be invertible,
Let <math> x: V \to V </math> be any operator on the finite-dimensional vector space <math> V </math> over the field <math> K </math>. Then <math> x </math> admits a Jordan-Chevalley decomposition [[if and only if]] the minimal polynomial of <math> x </math> is a product of separable polynomials. Moreover, in this case, there is a unique Jordan-Chevalley decomposition, and <math> x_s </math> (and hence also <math> x_n </math>) can be written as a polynomial (with coefficients from <math> K </math>) in <math> x </math> with zero constant coefficient.
:<math>x = x_s + x_n = x_s\left(1 + x_s^{-1}x_n\right)</math>
}}


Several proofs are discussed in <ref name=":0" />. A short direct version of the argument is given in <ref>{{Citation |last=Geck |first=Meinolf |title=On the Jordan--Chevalley decomposition of a matrix |date=2022-05-18 |url=http://arxiv.org/abs/2205.05432 |access-date=2024-01-09 |doi=10.48550/arXiv.2205.05432}}</ref>.
and <math>1 + x_s^{-1}x_n</math> is unipotent. (Conversely, by the same type of argument, one can deduce the additive version from the multiplicative one.)


If <math> K </math> is a [[perfect field]], then every polynomial is a product of separable polynomials (since every polynomial is a product of its irreducible factors, and these are separable over a perfect field). So in this case, the Jordan-Chevalley decomposition always exists. Moreover, over a perfect field, a polynomial is separable if and only if it is square-free. Therefore an operator is potentially diagonalisable if and only if its minimal polynomial is square-free. In general (over any field), the minimal polynomial of a linear operator is square-free if and only if the operator is [[Semisimple operator|semisimple]]<ref>{{Cite web |last=Conrad |first=Keith |date=accessed Jan 2024 |title=Semisimplicity |url=https://kconrad.math.uconn.edu/blurbs/linmultialg/semisimple.pdf |url-status=live |website=Expository papers}}</ref>. (In particular, the sum of two commuting semisimple operators is always semisimple over a perfect field. The same statement is not true over general fields.) The property of being semisimple is more relevant than being potentially diagonalisable in most contexts where the Jordan-Chevalley decomposition is applied, such as for Lie algebras.{{Citation needed|date=January 2024}} For these reasons, many texts restrict to the case of perfect fields.
If ''x'' is written in [[Jordan normal form]] (with respect to some basis) then ''x''<sub>''s''</sub> is the endomorphism whose matrix contains just the diagonal terms of ''x'', and ''x''<sub>''n''</sub> is the endomorphism whose matrix contains just the off-diagonal terms; ''x''<sub>''u''</sub> 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 ===
=== Proof of uniqueness and necessity ===
The uniqueness follows from the fact <math>x_s, x_n</math> are polynomial in ''x'': if <math>x = x_s' + x_n'</math> is another decomposition such that <math>x'_s</math> and <math>x'_n</math> commute, then <math>x_s - x_s' = x_n' - x_n</math>, and both <math>x_s', x_n'</math> commute with ''x'', hence with <math>x_s, x_n</math> since they are polynomials in <math>x</math>. The sum of commuting nilpotent endomorphisms is nilpotent, and over a perfect field the sum of commuting semisimple endomorphisms is again semisimple. Since the only operator which is both semisimple and nilpotent is the zero operator it follows that <math>x_s = x_s'</math> and <math>x_n = x_n'</math>.


That <math> x_s </math> and <math> x_n </math> are polynomials in <math> x </math> implies in particular that they commute with any operator that commutes with <math> x </math>. This observation underlies the uniqueness proof.
We show the existence. Let ''V'' be a finite-dimensional vector space over a perfect field ''k'' and <math>x : V \to V</math> an endomorphism.


First assume the base field ''k'' is algebraically closed. Then the vector space ''V'' has the direct sum decomposition <math display="inline">V = \bigoplus_{i=1}^r V_i</math> where each <math>V_i</math> is the kernel of <math>(x - \lambda_i I)^{m_i}</math>, the [[generalized eigenspace]] and ''x'' stabilizes <math>V_i</math>, meaning <math>x \cdot V_i \subset V_i</math>. Now, define <math>x_s : V \to V</math> so that, on each <math>V_i</math>, it is the scalar multiplication by <math>\lambda_i</math>. Note that, in terms of a basis respecting the direct sum decomposition, <math>x_s</math> is a diagonal matrix; hence, it is a semisimple endomorphism. Since <math>x - x_s : V_i \to V_i</math> is then <math>x - \lambda_i I : V_i \to V_i</math> whose <math>m_i</math>-th power is zero, we also have that <math>x_n := x - x_s</math> is nilpotent, establishing the existence of the decomposition.
Let <math>x = x_s + x_n</math> be a Jordan-Chevalley decomposition in which <math> x_s </math> and (hence also) <math> x_n </math> are polynomials in <math> x </math>. Let <math>x = x_s' + x_n'</math> be any Jordan-Chevalley decomposition. Then <math>x_s - x_s' = x_n' - x_n</math>, and <math>x_s', x_n'</math> both commute with <math> x </math>, hence with <math>x_s, x_n</math> since these are polynomials in <math>x</math>. The sum of commuting nilpotent operators is again nilpotent, and the sum of commuting potentially diagonalisable operators again potentially diagonalisable (because they are simultaneously diagonalisable over the algebraic closure of <math> K </math>). Since the only operator which is both potentially diagonalisable and nilpotent is the zero operator it follows that <math>x_s - x_s' = 0 = x_n - x_n'</math>.


To show that the condition that <math> x </math> have a minimal polynomial which is a product of separable polynomials is necessary, suppose that <math> x = x_s + x_n </math> is some Jordan-Chevalley decomposition. Letting <math> p </math> be the minimal polynomial of <math> x_s </math>, one can check using the [[binomial theorem]] that <math> p(x_s + x_n) </math> can be written as <math> x_n y </math> where <math> y </math> is some polynomial in <math> x_s, x_n </math>. Moreover, for some <math> \ell \geq 1 </math>, <math> x_n^\ell = 0 </math>. Thus <math> p(x)^\ell = x_n^\ell y^\ell = 0 </math> and so the minimal polynomial of <math> x </math> must divide <math> p^\ell </math>. As <math> p </math> is a product of separable polynomials (namely of <math> p </math>), so is the minimal polynomial.
(Choosing a basis carefully on each <math>V_i</math>, one can then put ''x'' in the Jordan normal form and <math>x_s, x_n</math> are the diagonal and the off-diagonal parts of the normal form. But this is not needed here.)


=== Concrete example for non-existence ===
The fact that <math>x_s, x_n</math> are polynomials in ''x'' follows from the [[Chinese remainder theorem]]. Indeed,

let <math>f(t) = \operatorname{det}(t I - x)</math> be the [[characteristic polynomial]] of ''x''. Then it is the product of the characteristic polynomials of <math>x : V_i \to V_i</math>; i.e., <math display="inline">f(t) = \prod_{i=1}^r (t - \lambda_i)^{d_i}</math>, where <math>d_i = \dim V_i.</math> Also, <math>d_i \ge m_i</math> (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 <math>k[t]</math> gives a polynomial <math>p(t)</math> satisfying the conditions
If the ground field is not [[perfect field|perfect]], then a Jordan–Chevalley decomposition may not exist, as it is possible that the minimal polynomial is not a product of separable polynomials. The simplest such example is the following. Let <math> p </math> be a prime number, let <math>k</math> be an imperfect field of characteristic <math>p,</math> (e. g. <math> k = \mathbb{F}_p(t) </math>) and choose <math>a \in k</math> that is not a <math>p</math>th power. Let <math>V = k[X]/\left(X^p - a\right)^2,</math> let <math>x = \overline X</math> be the image in the quotient and let <math>T</math> be the <math>k</math>-linear operator given by multiplication by <math>x</math> in <math>V</math>. Note that the minimal polynomial is precisely <math> \left(X^p - a\right)^2 </math>, which is inseparable and a square. By the necessity of the condition for the Jordan-Chevalley decomposition (as shown in the last section), this operator does not have a Jordan-Chevalley decomposition. It can be instructive to see concretely why there is at least no decomposition into a square-free and a nilpotent part.

{{Collapse top|title=Concrete argument for non-existence of a Jordan-Chavelley decomposition}}
Note that <math> T </math> has as its invariant <math>k</math>-linear subspaces precisely the ideals of <math>V</math> viewed as a ring, which correspond to the ideals of <math>k[X]</math> containing <math>\left(X^p - a\right)^2</math>. Since <math>X^p - a</math> is irreducible in <math>k[X],</math> ideals of <math> V </math> are <math>0,</math> <math>V</math> and <math>J = \left(x^p - a\right)V.</math> Suppose <math>T = S + N</math> for commuting <math>k</math>-linear operators <math>S</math> and <math>N</math> that are respectively semisimple (just over <math>k</math>, which is weaker than semisimplicity over an algebraic closure of <math>k</math> and also weaker than being potentially diagonalisable) and nilpotent. Since <math>S</math> and <math>N</math> commute, they each commute with <math>T = S + N</math> and hence each acts <math>k[x]</math>-linearly on <math>V</math>. Therefore <math>S</math> and <math>N</math> are each given by multiplication by respective members of <math>V</math> <math>s = S(1)</math> and <math>n = N(1),</math> with <math>s + n = T(1) = x</math>. Since <math>N</math> is nilpotent, <math>n</math> is nilpotent in <math>V,</math> therefore <math>\overline n = 0</math> in <math>V/J,</math> for <math>V/J</math> is a field. Hence, <math>n\in J,</math> therefore <math>n = \left(x^p - a\right)h(x)</math> for some polynomial <math>h(X) \in k[X]</math>. Also, we see that <math>n^2 = 0</math>. Since <math>k</math> is of characteristic <math>p,</math> we have <math>x^p = s^p + n^p = s^p</math>. On the other hand, since <math>\overline x = \overline s</math> in <math>A/J,</math> we have <math>h\left(\overline s\right) = h\left(\overline x\right),</math> therefore <math>h(s) - h(x)\in J</math> in <math>V.</math> Since <math>\left(x^p - a\right)J = 0,</math> we have <math>\left(x^p - a\right)h(x) = \left(x^p - a\right)h(s).</math> Combining these results we get <math>x = s + n = s + \left(s^p - a\right)h(s).</math> This shows that <math>s</math> generates <math>V</math> as a <math>k</math>-algebra and thus the <math>S</math>-stable <math>k</math>-linear subspaces of <math>V</math> are ideals of <math>V,</math> i.e. they are <math>0,</math> <math>J</math> and <math>V.</math> We see that <math>J</math> is an <math>S</math>-invariant subspace of <math>V</math> which has no complement <math>S</math>-invariant subspace, contrary to the assumption that <math>S</math> is semisimple. Thus, there is no decomposition of <math>T</math> as a sum of commuting <math>k</math>-linear operators that are respectively semisimple and nilpotent.
{{Collapse bottom}}

If instead of with the polynomial <math> \left(X^p - a\right)^2 </math>, the same construction is performed with <math> {X^p} - a </math>, the resulting operator <math> T </math> still does not admit a Jordan-Chevalley decomposition by the main theorem. However, <math> T </math> is semi-simple. The trivial decomposition <math> T = T + 0 </math> hence expresses <math> T </math> as a sum of a semisimple and a nilpotent operator, both of which are polynomials in <math> T </math>.

=== Proof of existence via Galois theory ===

This is one of the most typical proofs of the Jordan Chevalley decomposition. It has the advantage that it describes quite precisely how close one can get to a Jordan Chevalley decomposition.

Above it was observed that if <math> x </math> has a Jordan normal form (i. e. if the minimal polynomial of <math> x </math> splits), then it has a Jordan Chevalley decomposition. In this case, one can also see directly that <math> x_n </math> (and hence also <math> x_s </math>) is a polynomial in <math> x </math>. Indeed, it suffices to check this for the decomposition of the Jordan matrix <math> J = D + R </math>. This is a technical argument, but does not require any tricks beyond the [[Chinese remainder theorem]].

{{Collapse top|title=Proof (Jordan-Chevalley decomposition from Jordan normal form)}}
In the Jordan normal form, we have written <math> V = \bigoplus_{i = 1}^r V_i </math> where <math> r </math> is the number of Jordan blocks and <math> x |_{V_i} </math> is one Jordan block. Now let <math>f(t) = \operatorname{det}(t I - x)</math> be the [[characteristic polynomial]] of <math> x </math>. Because <math> f </math> splits, it can be written as <math> f(t) = \prod_{i=1}^r (t - \lambda_i)^{d_i} </math>, where <math> r </math> is the number of Jordan blocks, <math> \lambda_i </math> are the distinct eigenvalues, and <math> d_i </math> are the sizes of the Jordan blocks, so <math> d_i = \dim V_i </math>. Now, the Chinese remainder theorem applied to the polynomial ring <math>k[t]</math> gives a polynomial <math>p(t)</math> satisfying the conditions
:<math>p(t) \equiv 0 \bmod t,\, p(t) \equiv \lambda_i \bmod (t - \lambda_i)^{d_i}</math> (for all i).
:<math>p(t) \equiv 0 \bmod t,\, p(t) \equiv \lambda_i \bmod (t - \lambda_i)^{d_i}</math> (for all i).
(There is a redundancy in the conditions if some <math>\lambda_i</math> is zero but that is not an issue; just remove it from the conditions.) The condition <math>p(t) \equiv \lambda_i \bmod (t - \lambda_i)^{d_i}</math>, when spelled out, means that <math>p(t) - \lambda_i = g_i(t) (t - \lambda_i)^{d_i}</math> for some polynomial <math>g_i(t)</math>. Since <math>(x - \lambda_i I)^{d_i}</math> is the zero map on <math>V_i</math>, <math>p(x)</math> and <math>x_s</math> agree on each <math>V_i</math>; i.e., <math>p(x) = x_s</math>. Also then <math>q(x) = x_n</math> with <math>q(t) = t - p(t)</math>. The condition <math>p(t) \equiv 0 \bmod t</math> ensures that <math>p(t)</math> and <math>q(t)</math> have no constant terms. This completes the proof of the theorem in case the minimal polynomial of <math> x </math> splits.
{{Collapse bottom}}


This fact can be used to deduce the Jordan-Chevalley decomposition in the general case. Let <math> x </math> have minimal polynomial <math> p </math> and assume this is a product of separable polynomials. This condition is equivalent to demanding that there is some <math> q </math> such that <math> q | p </math> and <math> p | q^m </math> for some <math> m \geq 1 </math>. Let <math> L </math> be the [[splitting field]] of <math> q </math>, over which also <math> p </math> splits. Therefore, by the argument just given, <math> x </math> has a Jordan-Chevalley decomposition <math> x = {c(x)} + {(x - {c(x)})} </math> where <math> c </math> is a polynomial with coefficients from <math> L </math>, <c(x)> is diagonalisable (over <math> L </math>) and <math> x - c(x) </math> is nilpotent.
(There is a redundancy in the conditions if some <math>\lambda_i</math> is zero but that is not an issue; just remove it from the conditions.)


Let <math> \sigma </math> be a field automorphism of <math> L </math> which fixes <math> K </math>. Then
The condition <math>p(t) \equiv \lambda_i \bmod (t - \lambda_i)^{d_i}</math>, when spelled out, means that <math>p(t) - \lambda_i = g_i(t) (t - \lambda_i)^{d_i}</math> for some polynomial <math>g_i(t)</math>. Since <math>(x - \lambda_i I)^{d_i}</math> is the zero map on <math>V_i</math>, <math>p(x)</math> and <math>x_s</math> agree on each <math>V_i</math>; i.e., <math>p(x) = x_s</math>. Also then <math>q(x) = x_n</math> with <math>q(t) = t - p(t)</math>. The condition <math>p(t) \equiv 0 \bmod t</math> ensures that <math>p(t)</math> and <math>q(t)</math> have no constant terms. This completes the proof of the algebraically closed field case.
<math display="block">c(x) + (x-{c(x)}) = x = {\sigma(x)} = {\sigma({c(x)})} + {\sigma(x- {c(x)} )}</math>
Here <math>\sigma(c(x)) = \sigma(c)(x)</math> is a polynomial in <math>x</math>, so is <math> x - c(x) </math>. Thus, <math>\sigma(c(x))</math> and <math>\sigma(x - c(x))</math> commute. Also, <math> \sigma (c(x)) </math> is potentially diagonalisable and <math> \sigma({x - c(x)}) </math> is nilpotent. Thus, by the uniqueness of the Jordan-Chevalley decomposition (over <math>L</math>), <math>\sigma(c(x)) = c(x)</math> and <math>\sigma(c(x)) = c(x)</math>. Because <math> q </math> is separable, the extension <math> L/K </math> is [[Galois]], meaning an element of <math> L </math> fixed by all automorphisms of <math> L </math> which fix <math> K </math> must lie in <math> K </math>. That is, <math> x_s, x_n </math> are endomorphisms (represented by matrices) over <math> K </math>. Finally, since <math>\left\{1, x, x^2, \dots\right\}</math> contains a <math>\overline{k}</math>-basis that spans the space containing <math>x_s, x_n</math>, by the same argument, we also see that <math> c </math> has coefficients in <math> K </math>. This completes the proof. [[Q.E.D.]]


If ''k'' is an arbitrary perfect field, let <math>\Gamma = \operatorname{Gal}\left(\overline{k}/k\right)</math> be the absolute Galois group of ''k''. By the first part, we can choose polynomials <math>p, q</math> over <math>\overline{k}</math> such that <math>x = p(x) + q(x)</math> is the decomposition into the semisimple and nilpotent part. For each <math>\sigma</math> in <math>\Gamma</math>,
This argument describes precisely why the condition on <math> x </math> is required in the theorem and how close to a Jordan-Chevalley decomposition one can get in general: If <math> p </math> is the minimal polynomial of <math> x </math>, and <math> L </math> is the splitting field of <math> p </math>, then <math> x </math> admits a Jordan-Chevalley decomposition over the subfield of <math> L </math> of elements fixed by the entire [[Galois group]] <math> \operatorname{Gal}\left(L/K\right)</math> (but not over smaller subfields).
<math display="block">x = \sigma(x) = \sigma(p(x)) + \sigma(q(x)) = p(x) + q(x).</math>


== Relations to the theory of algebras ==
Now, <math>\sigma(p(x)) = \sigma(p)(x)</math> is a polynomial in <math>x</math>; so is <math>\sigma(q(x))</math>. Thus, <math>\sigma(p(x))</math> and <math>\sigma(q(x))</math> commute. Also, the application of <math>\sigma</math> evidently preserves semisimplicity and nilpotency. Thus, by the uniqueness of decomposition (over <math>\overline{k}</math>), <math>\sigma(p(x)) = p(x)</math> and <math>\sigma(q(x)) = q(x)</math>. Hence, <math>x_s = p(x), x_n = q(x)</math> are <math>\Gamma</math>-invariant; i.e., they are endomorphisms (represented by matrices) over ''k''. Finally, since <math>\left\{1, x, x^2, \dots\right\}</math> contains a <math>\overline{k}</math>-basis that spans the space containing <math>x_s, x_n</math>, by the same argument, we also see that <math>p, q</math> have coefficients in ''k''. This completes the proof. [[Q.E.D.]]


=== Separable algebras ===
=== Short proof using abstract algebra ===
{{harv|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.)


The Jordan-Chevalley decomposition is very closely related to the [[Wedderburn principal theorem]] in the following formulation<ref>{{Cite book |url=https://books.google.com/books?id=ZKGq4IQHhHUC&q=wedderburn+principal+theorem&pg=PA143 |title=Ring Theory |date=18 April 1972 |publisher=Academic Press |isbn=9780080873572}}</ref>:
Let ''V'' be a finite-dimensional vector space over a perfect field ''k'', <math>x : V \to V</math> an endomorphism and <math>A = k[x] \subset \operatorname{End}(V)</math> 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 <math>A/J</math> is separable, then the natural surjection <math>p: A \to A/J</math> splits; i.e., <math>A</math> contains a [[semisimple algebra|semisimple subalgebra]] <math>B</math> such that <math>p|_B : B \overset{\sim} \to A/J</math> is an isomorphism.<ref>{{Cite book|url=https://books.google.com/books?id=ZKGq4IQHhHUC&q=wedderburn+principal+theorem&pg=PA143|title = Ring Theory|date = 18 April 1972| publisher=Academic Press |isbn = 9780080873572}}</ref> In the setup here, <math>A/J</math> 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 <math>A = B \oplus J</math>. In particular, the endomorphism ''x'' can be written as <math>x = x_s + x_n</math> where <math>x_s</math> is in <math>B</math> and <math>x_n</math> in <math>J</math>. Now, the image of ''x'' generates <math>A/J \simeq B</math>; thus <math>x_s</math> is semisimple and is a polynomial of ''x''. Also, <math>x_n</math> is nilpotent since <math>J</math> is nilpotent and is a polynomial of ''x'' since <math>x_s</math> is. <math>\square</math>

{{Math theorem
| name = Wedderburn principal theorem
| math_statement = Let <math> A </math> be a finite-dimensional associative algebra over the field <math> K </math> with Jacobson radical <math> J </math>. Then <math> A/J </math> is separable [[if and only if]] <math> A </math> has a separable subalgebra <math> B </math> such that <math> A = B \oplus J </math>.
}}

Usually, the term „separable“ in this theorem refers to the general concept of a [[separable algebra]] and the theorem might then be established as a corollary of a more general high-powered result<ref>{{Cite book |last=Cohn |first=Paul M. |title=Further Algebra and Applications |date=2012 |publisher=Springer London |year=2002 |isbn=978-1-85233-667-7}} </ref>. However, if it is instead interpreted in the more basic sense that every element have a separable minimal polynomial, then this statement is essentially equivalent to the Jordan-Chevalley decomposition as described above. This gives a different way to view the decomposition, and for instance {{harv|Jacobson|1979}} takes this route for establishing it.

{{Collapse top|title=Proof of equivalence between Wedderburn principal theorem and Jordan-Chevalley decomposition}}
To see how the Jordan-Chevalley decomposition follows from the Wedderburn principal theorem, let <math> V </math> be a finite-dimensional vector space over the field <math> K </math>, <math>x : V \to V</math> an endomorphism with a minimal polynomial which is a product of separable polynomials and <math>A = K[x] \subset \operatorname{End}(V)</math> the subalgebra generated by <math> x </math>. Note that <math> A </math> is a commutative [[Artinian ring]], so <math> J </math> is also the nilradical of <math> A </math>. Moreover, <math>A/J</math> is separable, because if <math> a \in A </math>, then for minimal polynomial <math> p </math>, there is a separable polynomial <math> q </math> such that <math> q | p </math> and <math> p | q^m </math> for some <math> m \geq </math>. Therefore <math> q(a) \in J </math>, so the minimal polynomial of the image <math> a + J \in A/J </math> divides <math> q </math>, meaning that it must be separable as well (since a divisor of a separable polynomial is separable). There is then the vector-space decomposition <math> A = B \oplus J </math> with <math> B </math> separable. In particular, the endomorphism <math> x </math> can be written as <math>x = x_s + x_n</math> where <math>x_s \in B</math> and <math>x_n \in J</math>. Moreover, both elements are, like any element of </math> A </math>, polynomials in <math> x </math>.

Conversely, the Wedderburn principal theorem in the formulation above is a consequence of the Jordan-Chevalley decomposition. If <math> A </math> has a separable subalgebra <math> B </math> such that <math> A = B \oplus J </math>, then of course <math> A/J \cong B </math> is separable. Conversely, if <math> A/J </math> is separable, then any element of <math> A </math> is a sum of a separable and a nilpotent element. As shown above in #{proof for uniqueness}, this implies that the minimal polynomial will be a product of separable polynomials. Let <math> x \in A </math> be arbitrary, define the operator <math> T_x: A \to A, a \mapsto ax </math>, and note that this has the same minimal polynomial as <math> x </math>. So it admits a Jordan-Chevalley decomposition, where both operators are polynomials in <math> T_x </math>, hence of the form <math> T_s, T_n </math> for some <math> s, n \in A </math> which have separable and nilpotent minimal polynomials, respectively. Moreover, this decomposition is unique. Thus if <math> B </math> is the subalgebra of all separable elements (that this is a subalgebra can be seen by recalling that <math> s </math> is separable if and only if <math> T_s </math> is potentially diagonalisable), <math> A = B \oplus J </math> (because <math> J </math> is the ideal of nilpotent elements).
{{Collapse bottom}}

Over perfect fields, this result simplifies. Indeed, <math> A/J </math> is then always separable in the sense of minimal polynomials: If <math> a \in A </math>, then the minimal polynomial <math> p </math> is a product of separable polynomials, so there is a separable polynomial <math> q </math> such that <math> q | p </math> and <math> p | q^m </math> for some <math> m \geq 1 </math>. Thus <math> q(a) \in J </math>. So in <math> A/J </math>, the minimal polynomial of <math> a + J </math> divides <math> q </math> and is hence separable. The crucial point in the theorem is then not that <math> A/J </math> is separable (because that condition is vacuous), but that it is semisimple, meaning its [[radical]] is trivial.
The same statement is true for Lie algebras, but only in characteristic zero. This is the content of [[Levi decomposition|Levi’s theorem]]. (Note that the notions of semisimple in both results do indeed correspond, because in both cases this is equivalent to being the sum of simple subalgebras, at least in the finite-dimensional case.)

=== Preservation under representations ===

The crucial point in the proof for the Wedderburn principal theorem above is that an element <math> x \in A </math> corresponds to a linear operator <math> T_x: A \to A </math> with the same properties. In the theory of Lie algebras, this corresponds to the adjoint representation of a Lie algebra <math> \mathfrak{g} </math>. This decomposed operator has a Jordan-Chevalley decomposition <math>\operatorname{ad}(x) = \operatorname{ad}(x)_s + \operatorname{ad}(x)_n</math>. Just as in the associative case, this corresponds to a decomposition of <math> x </math>, but polynomials are not available as a tool. One context in which this does makes sense is the restricted case where <math> \mathfrak{g} </math> is contained in the Lie algebra <math> \mathfrak{gl}(V) </math> of the endomorphisms of a finite-dimensional vector space <math> V </math> over the perfect field <math> K </math>. Indeed, any Lie algebra can be realised in this way.{{Citation needed|date=January 2024}}

If <math>x = x_s + x_n</math> is the Jordan decomposition, then <math>\operatorname{ad}(x) = \operatorname{ad}(x_s) + \operatorname{ad}(x_n)</math> is the Jordan decomposition of the adjoint endomorphism <math>\operatorname{ad}(x)</math> on the vector space <math>\mathfrak{g}</math>. Indeed, first, <math>\operatorname{ad}(x_s)</math> and <math>\operatorname{ad}(x_n)</math> commute since <math>[\operatorname{ad}(x_s), \operatorname{ad}(x_n)] = \operatorname{ad}([x_s, x_n]) = 0</math>. Second, in general, for each endomorphism <math>y \in \mathfrak{g}</math>, we have:

# If <math>y^m = 0</math>, then <math>\operatorname{ad}(y)^{2m-1} = 0</math>, since <math>\operatorname{ad}(y)</math> is the difference of the left and right multiplications by ''y''.
# If <math>y</math> is semisimple, then <math>\operatorname{ad}(y)</math> is semisimple, since semisimple is equivalent to potentially diagonalisable over a perfect field (if <math> y </math> is diagonal over the basis <math> \{b_1, \dots, b_n \} </math>, then <math> \operatorname{ad}(y) </math> is diagonal over the basis consisting of the maps <math> M_{ij} </math> with <math> b_i \mapsto b_j </math> and <math> b_k \mapsto 0 </math> for <math> k \neq 0 </math>).<ref>This is not easy to see in general but is shown in the proof of {{harv|Jacobson|1979|loc=Ch. III, § 7, Theorem 11.}}. Editorial note: we need to add a discussion of this matter to "[[semisimple operator]]".</ref>

Hence, by uniqueness, <math>\operatorname{ad}(x)_s = \operatorname{ad}(x_s)</math> and <math>\operatorname{ad}(x)_n = \operatorname{ad}(x_n)</math>.

The adjoint representation is a very natural and general representation of any Lie algebra. The argument above illustrates (and indeed proves) a general principle which generalises this: If <math>\pi: \mathfrak{g} \to \mathfrak{gl}(V)</math> is ''any'' finite-dimensional representation of a [[semisimple Lie algebra|semisimple]] finite-dimensional Lie algebra over a perfect field, then <math>\pi</math> preserves the Jordan decomposition in the following sense: if <math>x = x_s + x_n</math>, then <math>\pi(x_s) = \pi(x)_s</math> and <math>\pi(x_n) = \pi(x)_n</math>.<ref>{{Cite web |last=Weber |first=Brian |date=02 Oct 2012 |title=Lecture 8 - Preservation of the Jordan Decomposition
and Levi’s Theorem |url=https://www2.math.upenn.edu/~brweber/Courses/2012/Math650/Notes/L8_JordanLevi.pdf |access-date=09 Jan 2024 |website=Course Notes}}</ref><ref>{{harvnb|Fulton|Harris|1991|loc=Theorem 9.20.}}</ref>


=== Nilpotency criterion ===
=== Nilpotency criterion ===

The Jordan decomposition can be used to characterize nilpotency of an endomorphism. Let ''k'' be an algebraically closed field of characteristic zero, <math>E = \operatorname{End}_\mathbb{Q}(k)</math> the endomorphism ring of ''k'' over rational numbers and ''V'' a finite-dimensional vector space over ''k''. Given an endomorphism <math>x : V \to V</math>, let <math>x = s + n</math> be the Jordan decomposition. Then <math>s</math> is diagonalizable; i.e., <math display="inline">V = \bigoplus V_i</math> where each <math>V_i</math> is the eigenspace for eigenvalue <math>\lambda_i</math> with multiplicity <math>m_i</math>. Then for any <math>\varphi\in E</math> let <math>\varphi(s) : V \to V</math> be the endomorphism such that <math>\varphi(s) : V_i \to V_i</math> is the multiplication by <math>\varphi(\lambda_i)</math>. Chevalley calls <math>\varphi(s)</math> the '''replica''' of <math>s</math> given by <math>\varphi</math>. (For example, if <math>k = \mathbb{C}</math>, then the complex conjugate of an endomorphism is an example of a replica.) Now,
The Jordan decomposition can be used to characterize nilpotency of an endomorphism. Let ''k'' be an algebraically closed field of characteristic zero, <math>E = \operatorname{End}_\mathbb{Q}(k)</math> the endomorphism ring of ''k'' over rational numbers and ''V'' a finite-dimensional vector space over ''k''. Given an endomorphism <math>x : V \to V</math>, let <math>x = s + n</math> be the Jordan decomposition. Then <math>s</math> is diagonalizable; i.e., <math display="inline">V = \bigoplus V_i</math> where each <math>V_i</math> is the eigenspace for eigenvalue <math>\lambda_i</math> with multiplicity <math>m_i</math>. Then for any <math>\varphi\in E</math> let <math>\varphi(s) : V \to V</math> be the endomorphism such that <math>\varphi(s) : V_i \to V_i</math> is the multiplication by <math>\varphi(\lambda_i)</math>. Chevalley calls <math>\varphi(s)</math> the '''replica''' of <math>s</math> given by <math>\varphi</math>. (For example, if <math>k = \mathbb{C}</math>, then the complex conjugate of an endomorphism is an example of a replica.) Now,


{{math_theorem
{{math_theorem|name=Nilpotency criterion|math_statement|<ref>{{harvnb|Serre|1992|loc=LA 5.17. Lemma 6.7.}} The endomorphism </ref><math>x</math> is nilpotent (i.e., <math>s = 0</math>) if and only if <math>\operatorname{tr}(x\varphi(s)) = 0</math> for every <math>\varphi \in E</math>. Also, if <math>k = \mathbb{C}</math>, then it suffices the condition holds for <math>\varphi = </math> complex conjugation.}}
| name = Nilpotency criterion|math_statement|<ref>{{harvnb|Serre|1992|loc=LA 5.17. Lemma 6.7.}} The endomorphism </ref><math>x</math> is nilpotent (i.e., <math>s = 0</math>) if and only if <math>\operatorname{tr}(x\varphi(s)) = 0</math> for every <math>\varphi \in E</math>. Also, if <math>k = \mathbb{C}</math>, then it suffices the condition holds for <math>\varphi = </math> complex conjugation.
}}


''Proof:'' First, since <math>n \varphi(s)</math> is nilpotent,
''Proof:'' First, since <math>n \varphi(s)</math> is nilpotent,
Line 66: Line 130:
is zero. Let <math>\mathfrak{g}' = \mathfrak{gl}(V)</math>. Note we have: <math>\operatorname{ad}_{\mathfrak{g}'}(x) : \mathfrak{g} \to D \mathfrak{g}</math> and, since <math>\operatorname{ad}_{\mathfrak{g}'}(s)</math> is the semisimple part of the Jordan decomposition of <math>\operatorname{ad}_{\mathfrak{g}'}(x)</math>, it follows that <math>\operatorname{ad}_{\mathfrak{g}'}(s)</math> is a polynomial without constant term in <math>\operatorname{ad}_{\mathfrak{g}'}(x)</math>; hence, <math>\operatorname{ad}_{\mathfrak{g}'}(s) : \mathfrak{g} \to D \mathfrak{g}</math> and the same is true with <math>\varphi(s)</math> in place of <math>s</math>. That is, <math>[\varphi(s), \mathfrak{g}] \subset D \mathfrak{g}</math>, which implies the claim given the assumption. <math>\square</math>
is zero. Let <math>\mathfrak{g}' = \mathfrak{gl}(V)</math>. Note we have: <math>\operatorname{ad}_{\mathfrak{g}'}(x) : \mathfrak{g} \to D \mathfrak{g}</math> and, since <math>\operatorname{ad}_{\mathfrak{g}'}(s)</math> is the semisimple part of the Jordan decomposition of <math>\operatorname{ad}_{\mathfrak{g}'}(x)</math>, it follows that <math>\operatorname{ad}_{\mathfrak{g}'}(s)</math> is a polynomial without constant term in <math>\operatorname{ad}_{\mathfrak{g}'}(x)</math>; hence, <math>\operatorname{ad}_{\mathfrak{g}'}(s) : \mathfrak{g} \to D \mathfrak{g}</math> and the same is true with <math>\varphi(s)</math> in place of <math>s</math>. That is, <math>[\varphi(s), \mathfrak{g}] \subset D \mathfrak{g}</math>, which implies the claim given the assumption. <math>\square</math>


==== Real semisimple Lie algebras ====
===Counterexample to existence over an imperfect field ===
If the ground field is not [[perfect field|perfect]], then a Jordan–Chevalley decomposition may not exist. Example: Let ''p'' be a prime number, let <math>k</math> be imperfect of characteristic <math>p,</math> and choose <math>a</math> in <math>k</math> that is not a <math>p</math>th power. Let <math>V = k[X]/\left(X^p - a\right)^2,</math> let <math>x = \overline X</math>and let <math>T</math> be the <math>k</math>-linear operator given by multiplication by <math>x</math> in <math>V.</math> This has as its invariant <math>k</math>-linear subspaces precisely the ideals of <math>V</math> viewed as a ring, which correspond to the ideals of <math>k[X]</math> containing <math>\left(X^p - a\right)^2.</math> Since <math>X^p - a</math> is irreducible in <math>k[X],</math> ideals of ''V'' are <math>0,</math> <math>V</math> and <math>J = \left(x^p - a\right)V.</math> Suppose <math>T = S + N</math> for commuting <math>k</math>-linear operators <math>S</math> and <math>N</math> that are respectively semisimple (just over <math>k,</math> which is weaker than semisimplicity over an algebraic closure of <math>k</math>) and nilpotent. Since <math>S</math> and <math>N</math> commute, they each commute with <math>T = S + N</math> and hence each acts <math>k[x]</math>-linearly on <math>V.</math> Therefore <math>S</math> and <math>N</math> are each given by multiplication by respective members of <math>V</math> <math>s = S(1)</math> and <math>n = N(1),</math> with <math>s + n = T(1) = x.</math> Since <math>N</math> is nilpotent, <math>n</math> is nilpotent in <math>V,</math> therefore <math>\overline n = 0</math> in <math>V/J,</math> for <math>V/J</math> is a field. Hence, <math>n\in J,</math> therefore <math>n = \left(x^p - a\right)h(x)</math> for some polynomial <math>h(X) \in k[X].</math> Also, we see that <math>n^2 = 0.</math> Since <math>k</math> is of characteristic <math>p,</math> we have <math>x^p = s^p + n^p = s^p.</math> Also, since <math>\overline x = \overline s</math> in <math>A/J,</math> we have <math>h\left(\overline s\right) = h\left(\overline x\right),</math> therefore <math>h(s) - h(x)\in J</math> in <math>V.</math> Since <math>\left(x^p - a\right)J = 0,</math> we have <math>\left(x^p - a\right)h(x) = \left(x^p - a\right)h(s).</math> Combining these results we get <math>x = s + n = s + \left(s^p - a\right)h(s).</math> This shows that <math>s</math> generates <math>V</math> as a <math>k</math>-algebra and thus the <math>S</math>-stable <math>k</math>-linear subspaces of <math>V</math> are ideals of <math>V,</math> i.e. they are <math>0,</math> <math>J</math> and <math>V.</math> We see that <math>J</math> is an <math>S</math>-invariant subspace of <math>V</math> which has no complement <math>S</math>-invariant subspace, contrary to the assumption that <math>S</math> is semisimple. Thus, there is no decomposition of <math>T</math> as a sum of commuting <math>k</math>-linear operators that are respectively semisimple and nilpotent. Note that minimal polynomial of <math>T</math> is inseparable over <math>k</math> and is a square in <math>k[X].</math> It can be shown that if minimal polynomial of <math>k</math> linear operator <math>L</math> is separable then <math>L</math> has Jordan-Chevalley decomposition and that if this polynomial is product of distinct irreducible polynomials in <math>k[X],</math> then <math>L</math> is semisimple over <math>k.</math>


In the formulation of Chevalley and [[George Mostow|Mostow]], the additive decomposition states that an element ''X'' in a real [[semisimple Lie algebra]] '''g''' with [[Iwasawa decomposition]] '''g''' = '''k''' ⊕ '''a''' ⊕ '''n''' 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.
== 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.


== Multiplicative decomposition ==
=== Lie algebras ===
Let <math>\mathfrak{gl}(V)</math> denote the Lie algebra of the endomorphisms of a finite-dimensional vector space ''V'' over a perfect field. If <math>x = x_s + x_n</math> is the Jordan decomposition, then <math>\operatorname{ad}(x) = \operatorname{ad}(x_s) + \operatorname{ad}(x_n)</math> is the Jordan decomposition of <math>\operatorname{ad}(x)</math> on the vector space <math>\mathfrak{gl}(V)</math>. Indeed, first, <math>\operatorname{ad}(x_s)</math> and <math>\operatorname{ad}(x_n)</math> commute since <math>[\operatorname{ad}(x_s), \operatorname{ad}(x_n)] = \operatorname{ad}([x_s, x_n]) = 0</math>. Second, in general, for each endomorphism <math>y \in \mathfrak{gl}(V)</math>, we have:
# If <math>y^m = 0</math>, then <math>\operatorname{ad}(y)^{2m-1} = 0</math>, since <math>\operatorname{ad}(y)</math> is the difference of the left and right multiplications by ''y''.
# If <math>y</math> is semisimple, then <math>\operatorname{ad}(y)</math> is semisimple.<ref>This is not easy to see but is shown in the proof of {{harv|Jacobson|1979|loc=Ch. III, § 7, Theorem 11.}}. Editorial note: we need to add a discussion of this matter to "[[semisimple operator]]".</ref>


If <math> x </math> is an invertible linear operator, it may be more convenient to use a multiplicative Jordan-Chevalley decomposition. This expresses <math> x </math> as a product
Hence, by uniqueness, <math>\operatorname{ad}(x)_s = \operatorname{ad}(x_s)</math> and <math>\operatorname{ad}(x)_n = \operatorname{ad}(x_n)</math>.
:<math> x = x_s \cdot x_u </math>,
where <math> x_s </math> is potentially diagonalisable, and <math> x_u - 1 </math> is nilpotent (one also says that <math> x_u </math> is unipotent).


The multiplicative version of the decomposition follows from the additive one since, as <math>x_s</math> is invertible (because the sum of an invertible operator and a nilpotent operator is invertible)
If <math>\pi: \mathfrak{g} \to \mathfrak{gl}(V)</math> is a finite-dimensional representation of a [[semisimple Lie algebra|semisimple]] finite-dimensional complex<!-- need a ref for a non-complex base --> Lie algebra, then <math>\pi</math> preserves the Jordan decomposition in the sense: if <math>x = x_s + x_n</math>, then <math>\pi(x_s) = \pi(x)_s</math> and <math>\pi(x_n) = \pi(x)_n</math>.<ref>{{harvnb|Fulton|Harris|1991|loc=Theorem 9.20.}}</ref>
:<math>x = x_s + x_n = x_s\left(1 + x_s^{-1}x_n\right)</math>
and <math>1 + x_s^{-1}x_n</math> is unipotent. (Conversely, by the same type of argument, one can deduce the additive version from the multiplicative one.)


The multiplicative version is closely related to decompositions encountered in a linear algebraic group. For this it is again useful to assume that the underlying field <math> K </math> is perfect because then the Jordan-Chevalley decomposition exists for all matrices.
==== Real semisimple Lie algebras ====
In the formulation of Chevalley and [[George Mostow|Mostow]], the additive decomposition states that an element ''X'' in a real [[semisimple Lie algebra]] '''g''' with [[Iwasawa decomposition]] '''g''' = '''k''' ⊕ '''a''' ⊕ '''n''' 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 ===
=== Linear algebraic groups ===

Let <math>G</math> be a [[linear algebraic group]] over a perfect field. Then, essentially by definition, there is a closed embedding <math>G \hookrightarrow \mathbf{GL}_n</math>. Now, to each element <math>g \in G</math>, by the multiplicative Jordan decomposition, there are a pair of a semisimple element <math>g_s</math> and a unipotent element <math>g_u</math> ''a priori'' in <math>\mathbf{GL}_n</math> such that <math>g = g_s g_u = g_u g_s</math>. But, as it turns out,<ref>{{harvnb|Waterhouse|1979|loc=Theorem 9.2.}}</ref> the elements <math>g_s, g_u</math> can be shown to be in <math>G</math> (i.e., they satisfy the defining equations of ''G'') and that they are independent of the embedding into <math>\mathbf{GL}_n</math>; i.e., the decomposition is intrinsic.
Let <math>G</math> be a [[linear algebraic group]] over a perfect field. Then, essentially by definition, there is a closed embedding <math>G \hookrightarrow \mathbf{GL}_n</math>. Now, to each element <math>g \in G</math>, by the multiplicative Jordan decomposition, there are a pair of a semisimple element <math>g_s</math> and a unipotent element <math>g_u</math> ''a priori'' in <math>\mathbf{GL}_n</math> such that <math>g = g_s g_u = g_u g_s</math>. But, as it turns out,<ref>{{harvnb|Waterhouse|1979|loc=Theorem 9.2.}}</ref> the elements <math>g_s, g_u</math> can be shown to be in <math>G</math> (i.e., they satisfy the defining equations of ''G'') and that they are independent of the embedding into <math>\mathbf{GL}_n</math>; i.e., the decomposition is intrinsic.


When ''G'' is abelian, <math>G</math> is then the direct product of the closed subgroup of the semisimple elements in ''G'' and that of unipotent elements.<ref>{{harvnb|Waterhouse|1979|loc=Theorem 9.3.}}</ref>
When ''G'' is abelian, <math>G</math> is then the direct product of the closed subgroup of the semisimple elements in ''G'' and that of unipotent elements.<ref>{{harvnb|Waterhouse|1979|loc=Theorem 9.3.}}</ref>

==== Real semisimple Lie groups ====
==== 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.
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 ==
== References ==

{{reflist}}
{{reflist}}
*{{citation|last=Chevalley|first= Claude|title= Théorie des groupes de Lie. Tome II. Groupes algébriques|publisher= Hermann|year= 1951 |oclc=277477632}}
*{{citation |last=Chevalley |first=Claude |title=Théorie des groupes de Lie. Tome II. Groupes algébriques |publisher=Hermann |year=1951 |oclc=277477632}}
*{{Fulton-Harris}}
*{{Fulton-Harris}}
*{{citation|first=Sigurdur|last= Helgason|title=Differential geometry, Lie groups, and symmetric spaces|year=1978|publisher=Academic Press|isbn= 0-8218-2848-7}}
*{{citation |first=Sigurdur |last=Helgason |title=Differential geometry, Lie groups, and symmetric spaces |year=1978 |publisher=Academic Press |isbn=0-8218-2848-7}}
*{{citation|first=James E.|last= Humphreys|title=Linear Algebraic Groups|publisher=Springer|year=1981|isbn=0-387-90108-6|series=Graduate texts in mathematics|volume=21}}
*{{citation |first=James E. |last=Humphreys |title=Linear Algebraic Groups |publisher=Springer |year=1981 |isbn=0-387-90108-6 |series=Graduate texts in mathematics |volume=21}}
* {{citation | first1=James E. | last1=Humphreys | year=1972 | title=Introduction to Lie Algebras and Representation Theory | publisher=Springer | isbn=978-0-387-90053-7 | url-access=registration | url=https://archive.org/details/introductiontoli00jame }}
* {{citation |first1=James E. |last1=Humphreys |year=1972 |title=Introduction to Lie Algebras and Representation Theory |publisher=Springer |isbn=978-0-387-90053-7 |url-access=registration |url=https://archive.org/details/introductiontoli00jame}}
*{{citation |author1-link=Nathan Jacobson |last1=Jacobson |first1=Nathan |title=Lie algebras |orig-year=1962|publisher=Dover |year=1979 |isbn=0-486-63832-4}}
*{{citation |author1-link=Nathan Jacobson |last1=Jacobson |first1=Nathan |title=Lie algebras |orig-year=1962 |publisher=Dover |year=1979 |isbn=0-486-63832-4}}
*{{citation|last=Lazard|first=M.|title=Théorie des répliques. Critère de Cartan (Exposé No. 6)|journal=Séminaire "Sophus Lie"|year=1954|volume=1|url=http://www.numdam.org/numdam-bin/recherche?nci=SSL_1954-1955__1__A9_0&format=short|archive-url=https://archive.today/20130704093910/http://www.numdam.org/numdam-bin/recherche?nci=SSL_1954-1955__1__A9_0&format=short|url-status=dead|archive-date=2013-07-04}}
*{{citation |last=Lazard |first=M. |title=Théorie des répliques. Critère de Cartan (Exposé No. 6) |journal=Séminaire "Sophus Lie" |year=1954 |volume=1 |url=http://www.numdam.org/numdam-bin/recherche?nci=SSL_1954-1955__1__A9_0&format=short |archive-url=https://archive.today/20130704093910/http://www.numdam.org/numdam-bin/recherche?nci=SSL_1954-1955__1__A9_0&format=short |url-status=dead |archive-date=2013-07-04}}
* {{citation|first=G. D.|last=Mostow|title=Factor spaces of solvable groups|journal= Ann. of Math.|volume= 60|year=1954|issue=1|pages= 1–27|doi=10.2307/1969700|jstor=1969700}}
* {{citation |first=G. D. |last=Mostow |title=Factor spaces of solvable groups |journal=Ann. of Math. |volume=60 |year=1954 |issue=1 |pages=1–27 |doi=10.2307/1969700 |jstor=1969700}}
* {{citation|last=Mostow|first= G. D.|title= Strong rigidity of locally symmetric spaces|series= Annals of Mathematics Studies|volume= 78|publisher= Princeton University Press|year= 1973 |isbn=0-691-08136-0 |url=https://books.google.com/books?id=xT0SFmrFrWoC }}
* {{citation |last=Mostow |first=G. D. |title=Strong rigidity of locally symmetric spaces |series=Annals of Mathematics Studies |volume=78 |publisher=Princeton University Press |year=1973 |isbn=0-691-08136-0 |url=https://books.google.com/books?id=xT0SFmrFrWoC}}
* {{Lang Algebra|edition=3r}}
* {{Lang Algebra|edition=3r}}
*{{citation|title=Lie algebras and Lie groups: 1964 lectures given at Harvard University|
*{{citation |title=Lie algebras and Lie groups: 1964 lectures given at Harvard University |volume=1500 |series=Lecture Notes in Mathematics |first=Jean-Pierre |last=Serre |edition=2nd |publisher=Springer-Verlag |year=1992 |isbn=978-3-540-55008-2}}
volume=1500|series= Lecture Notes in Mathematics|first=Jean-Pierre|last= Serre|edition=2nd|publisher=Springer-Verlag|year= 1992|isbn=978-3-540-55008-2}}
* {{citation |last=Varadarajan |first=V. S. |title=Lie groups, Lie algebras, and their representations |series=Graduate Texts in Mathematics |volume=102 |publisher=Springer-Verlag |year=1984 |isbn=0-387-90969-9}}
* {{citation|last=Varadarajan|first= V. S.|title= Lie groups, Lie algebras, and their representations| series=Graduate Texts in Mathematics|volume= 102|publisher= Springer-Verlag|year= 1984|isbn=0-387-90969-9}}
* {{Citation |last1=Waterhouse |first1=William |author1-link=William C. Waterhouse |title=Introduction to affine group schemes |publisher=[[Springer-Verlag]] |location=Berlin, New York |series=Graduate Texts in Mathematics |isbn=978-0-387-90421-4 |year=1979 |volume=66 |doi=10.1007/978-1-4612-6217-6 |mr=0547117}}
* {{Citation | last1=Waterhouse | first1=William | author1-link=William C. Waterhouse | title=Introduction to affine group schemes | publisher=[[Springer-Verlag]] | location=Berlin, New York | series=Graduate Texts in Mathematics | isbn=978-0-387-90421-4 | year=1979 | volume=66 | doi=10.1007/978-1-4612-6217-6 | mr=0547117}}


{{DEFAULTSORT:Jordan-Chevalley Decomposition}}
{{DEFAULTSORT:Jordan-Chevalley Decomposition}}

Revision as of 14:44, 9 January 2024

In mathematics, specifically linear algebra, the Jordan–Chevalley decomposition, named after Camille Jordan and Claude Chevalley, expresses a linear operator in a unique way as the sum of two other linear operators which are simpler to understand. Specifically, one part is potentially diagonalisable and the other is nilpotent. The two parts are polynomials in the operator, which makes them behave nicely in algebraic manipulations.

The decomposition has a short description when the Jordan normal form of the operator is given, but it exists under weaker hypotheses than are needed for the existence of a Jordan normal form. Hence the Jordan-Chevalley decomposition can be seen as a generalisation of the Jordan normal form, which is also reflected in many proofs of it.

It is closely related to the Wedderburn principal theorem about associative algebras, which also leads to several analogues in Lie algebras. Analogues of the Jordan-Chevalley decomposition also exist for elements of Linear algebraic groups and Lie groups via a multiplicative reformulation. The decomposition is an important tool in the study of all of these objects, and was developed for this purpose.

In many texts, the potentially diagonalisable part is also characterised as the semisimple part.

Introduction

A basic question in linear algebra is whether an operator on a finite-dimensional vector space can be diagonalised. For example, this is closely related to the eigenvalues of the operator. In several contexts, one may be dealing with many operators which are not diagonalisable. Even over an algebraically closed field, a diagonalisation may not exist. In this context, the Jordan normal form achieves the best possible result to obtain something similar to a diagonalisation. For linear operators over a field which is not algebraically closed, there may be no eigenvector at all. The latter point is not the main concern in discussions on the Jordan-Chevalley decomposition. To avoid this problem, one can instead consider potentially diagonalisable operators, which are those that admit a diagonalisation over some field (or equivalently over the algebraic closure of the field under consideration).

The operators which are „the furthest away“ from being diagonalisable are nilpotent operators. An operator (or more generally an element of a ring) is said to be nilpotent when there is some positive integer such that . In several contexts in abstract algebra, it is the case that the presence of nilpotent elements of a ring make them much more complicated to work with.[citation needed] To a large extent, this is also the case for linear operators. The remarkable fact is that one can „separate out“ the nilpotent part of an operator which causes it to be not potentially diagonalisable. This is what is described by the Jordan-Chevalley decomposition.

Historically, the Jordan-Chevalley decomposition was motivated by the applications to the theory of Lie algebras and linear algebraic groups[1][2], as described in sections below.

Decomposition of a linear operator

Let be a field, a finite-dimensional vector space over , and a linear operator over (equivalently, a matrix with entries from ). If the minimal polynomial of splits over (for example if is algebraically closed), then has a Jordan normal form . If is the diagonal of , let be the remaining part. Then is a decomposition where is diagonalisable and is nilpotent. This restatement of the normal form as an additive decomposition not only makes the numerical computation more stable[citation needed], but can be generalised to cases where the minimal polynomial of does not split.

If the minimal polynomial of splits into \textit{distinct} linear factors, then is diagonalisable. Therefore, if the minimal polynomial of is at least separable, then is potentially diagonalisable. The Jordan-Chevalley decomposition is concerned with the more general case where the minimal polynomial of is a product of separable polynomials.

Let be any linear operator on the finite-dimensional vector space over the field . A Jordan–Chevalley decomposition of is an expression of it as a sum

,

where is potentially diagonalisable, is nilpotent, and .

Jordan-Chevalley decomposition — Let be any operator on the finite-dimensional vector space over the field . Then admits a Jordan-Chevalley decomposition if and only if the minimal polynomial of is a product of separable polynomials. Moreover, in this case, there is a unique Jordan-Chevalley decomposition, and (and hence also ) can be written as a polynomial (with coefficients from ) in with zero constant coefficient.

Several proofs are discussed in [2]. A short direct version of the argument is given in [3].

If is a perfect field, then every polynomial is a product of separable polynomials (since every polynomial is a product of its irreducible factors, and these are separable over a perfect field). So in this case, the Jordan-Chevalley decomposition always exists. Moreover, over a perfect field, a polynomial is separable if and only if it is square-free. Therefore an operator is potentially diagonalisable if and only if its minimal polynomial is square-free. In general (over any field), the minimal polynomial of a linear operator is square-free if and only if the operator is semisimple[4]. (In particular, the sum of two commuting semisimple operators is always semisimple over a perfect field. The same statement is not true over general fields.) The property of being semisimple is more relevant than being potentially diagonalisable in most contexts where the Jordan-Chevalley decomposition is applied, such as for Lie algebras.[citation needed] For these reasons, many texts restrict to the case of perfect fields.

Proof of uniqueness and necessity

That and are polynomials in implies in particular that they commute with any operator that commutes with . This observation underlies the uniqueness proof.

Let be a Jordan-Chevalley decomposition in which and (hence also) are polynomials in . Let be any Jordan-Chevalley decomposition. Then , and both commute with , hence with since these are polynomials in . The sum of commuting nilpotent operators is again nilpotent, and the sum of commuting potentially diagonalisable operators again potentially diagonalisable (because they are simultaneously diagonalisable over the algebraic closure of ). Since the only operator which is both potentially diagonalisable and nilpotent is the zero operator it follows that .

To show that the condition that have a minimal polynomial which is a product of separable polynomials is necessary, suppose that is some Jordan-Chevalley decomposition. Letting be the minimal polynomial of , one can check using the binomial theorem that can be written as where is some polynomial in . Moreover, for some , . Thus and so the minimal polynomial of must divide . As is a product of separable polynomials (namely of ), so is the minimal polynomial.

Concrete example for non-existence

If the ground field is not perfect, then a Jordan–Chevalley decomposition may not exist, as it is possible that the minimal polynomial is not a product of separable polynomials. The simplest such example is the following. Let be a prime number, let be an imperfect field of characteristic (e. g. ) and choose that is not a th power. Let let be the image in the quotient and let be the -linear operator given by multiplication by in . Note that the minimal polynomial is precisely , which is inseparable and a square. By the necessity of the condition for the Jordan-Chevalley decomposition (as shown in the last section), this operator does not have a Jordan-Chevalley decomposition. It can be instructive to see concretely why there is at least no decomposition into a square-free and a nilpotent part.

Concrete argument for non-existence of a Jordan-Chavelley decomposition

Note that 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 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 also weaker than being potentially diagonalisable) 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 . On the other hand, 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.

If instead of with the polynomial , the same construction is performed with , the resulting operator still does not admit a Jordan-Chevalley decomposition by the main theorem. However, is semi-simple. The trivial decomposition hence expresses as a sum of a semisimple and a nilpotent operator, both of which are polynomials in .

Proof of existence via Galois theory

This is one of the most typical proofs of the Jordan Chevalley decomposition. It has the advantage that it describes quite precisely how close one can get to a Jordan Chevalley decomposition.

Above it was observed that if has a Jordan normal form (i. e. if the minimal polynomial of splits), then it has a Jordan Chevalley decomposition. In this case, one can also see directly that (and hence also ) is a polynomial in . Indeed, it suffices to check this for the decomposition of the Jordan matrix . This is a technical argument, but does not require any tricks beyond the Chinese remainder theorem.

Proof (Jordan-Chevalley decomposition from Jordan normal form)

In the Jordan normal form, we have written where is the number of Jordan blocks and is one Jordan block. Now let be the characteristic polynomial of . Because splits, it can be written as , where is the number of Jordan blocks, are the distinct eigenvalues, and are the sizes of the Jordan blocks, so . 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 theorem in case the minimal polynomial of splits.

This fact can be used to deduce the Jordan-Chevalley decomposition in the general case. Let have minimal polynomial and assume this is a product of separable polynomials. This condition is equivalent to demanding that there is some such that and for some . Let be the splitting field of , over which also splits. Therefore, by the argument just given, has a Jordan-Chevalley decomposition where is a polynomial with coefficients from , <c(x)> is diagonalisable (over ) and is nilpotent.

Let be a field automorphism of which fixes . Then

Here is a polynomial in , so is . Thus, and commute. Also, is potentially diagonalisable and is nilpotent. Thus, by the uniqueness of the Jordan-Chevalley decomposition (over ), and . Because is separable, the extension is Galois, meaning an element of fixed by all automorphisms of which fix must lie in . That is, are endomorphisms (represented by matrices) over . Finally, since contains a -basis that spans the space containing , by the same argument, we also see that has coefficients in . This completes the proof. Q.E.D.

This argument describes precisely why the condition on is required in the theorem and how close to a Jordan-Chevalley decomposition one can get in general: If is the minimal polynomial of , and is the splitting field of , then admits a Jordan-Chevalley decomposition over the subfield of of elements fixed by the entire Galois group (but not over smaller subfields).

Relations to the theory of algebras

Separable algebras

The Jordan-Chevalley decomposition is very closely related to the Wedderburn principal theorem in the following formulation[5]:

Wedderburn principal theorem — Let be a finite-dimensional associative algebra over the field with Jacobson radical . Then is separable if and only if has a separable subalgebra such that .

Usually, the term „separable“ in this theorem refers to the general concept of a separable algebra and the theorem might then be established as a corollary of a more general high-powered result[6]. However, if it is instead interpreted in the more basic sense that every element have a separable minimal polynomial, then this statement is essentially equivalent to the Jordan-Chevalley decomposition as described above. This gives a different way to view the decomposition, and for instance (Jacobson 1979) takes this route for establishing it.

Proof of equivalence between Wedderburn principal theorem and Jordan-Chevalley decomposition

To see how the Jordan-Chevalley decomposition follows from the Wedderburn principal theorem, let be a finite-dimensional vector space over the field , an endomorphism with a minimal polynomial which is a product of separable polynomials and the subalgebra generated by . Note that is a commutative Artinian ring, so is also the nilradical of . Moreover, is separable, because if , then for minimal polynomial , there is a separable polynomial such that and for some . Therefore , so the minimal polynomial of the image divides , meaning that it must be separable as well (since a divisor of a separable polynomial is separable). There is then the vector-space decomposition with separable. In particular, the endomorphism can be written as where and . Moreover, both elements are, like any element of </math> A </math>, polynomials in .

Conversely, the Wedderburn principal theorem in the formulation above is a consequence of the Jordan-Chevalley decomposition. If has a separable subalgebra such that , then of course is separable. Conversely, if is separable, then any element of is a sum of a separable and a nilpotent element. As shown above in #{proof for uniqueness}, this implies that the minimal polynomial will be a product of separable polynomials. Let be arbitrary, define the operator , and note that this has the same minimal polynomial as . So it admits a Jordan-Chevalley decomposition, where both operators are polynomials in , hence of the form for some which have separable and nilpotent minimal polynomials, respectively. Moreover, this decomposition is unique. Thus if is the subalgebra of all separable elements (that this is a subalgebra can be seen by recalling that is separable if and only if is potentially diagonalisable), (because is the ideal of nilpotent elements).

Over perfect fields, this result simplifies. Indeed, is then always separable in the sense of minimal polynomials: If , then the minimal polynomial is a product of separable polynomials, so there is a separable polynomial such that and for some . Thus . So in , the minimal polynomial of divides and is hence separable. The crucial point in the theorem is then not that is separable (because that condition is vacuous), but that it is semisimple, meaning its radical is trivial. The same statement is true for Lie algebras, but only in characteristic zero. This is the content of Levi’s theorem. (Note that the notions of semisimple in both results do indeed correspond, because in both cases this is equivalent to being the sum of simple subalgebras, at least in the finite-dimensional case.)

Preservation under representations

The crucial point in the proof for the Wedderburn principal theorem above is that an element corresponds to a linear operator with the same properties. In the theory of Lie algebras, this corresponds to the adjoint representation of a Lie algebra . This decomposed operator has a Jordan-Chevalley decomposition . Just as in the associative case, this corresponds to a decomposition of , but polynomials are not available as a tool. One context in which this does makes sense is the restricted case where is contained in the Lie algebra of the endomorphisms of a finite-dimensional vector space over the perfect field . Indeed, any Lie algebra can be realised in this way.[citation needed]

If is the Jordan decomposition, then is the Jordan decomposition of the adjoint endomorphism 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, since semisimple is equivalent to potentially diagonalisable over a perfect field (if is diagonal over the basis , then is diagonal over the basis consisting of the maps with and for ).[7]

Hence, by uniqueness, and .

The adjoint representation is a very natural and general representation of any Lie algebra. The argument above illustrates (and indeed proves) a general principle which generalises this: If is any finite-dimensional representation of a semisimple finite-dimensional Lie algebra over a perfect field, then preserves the Jordan decomposition in the following sense: if , then and .[8][9]

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 — [10] 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:[11] 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.

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.

Multiplicative decomposition

If is an invertible linear operator, it may be more convenient to use a multiplicative Jordan-Chevalley decomposition. This expresses as a product

,

where is potentially diagonalisable, and is nilpotent (one also says that is unipotent).

The multiplicative version of the decomposition follows from the additive one since, as is invertible (because the sum of an invertible operator and a nilpotent operator is invertible)

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

The multiplicative version is closely related to decompositions encountered in a linear algebraic group. For this it is again useful to assume that the underlying field is perfect because then the Jordan-Chevalley decomposition exists for all matrices.

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,[12] 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.[13]

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. ^ Danielle Couty, Jean Esterle, Rachid Zarouf. (2010.). "Décomposition effective de Jordan-Chevalley et ses retombées en enseignement" (PDF) – via HAL. {{cite journal}}: Check date values in: |date= (help); Cite journal requires |journal= (help)CS1 maint: multiple names: authors list (link)
  2. ^ a b Couty, Danielle; Esterle, Jean; Zarouf, Rachid (2011, 129, pp.29–49.). "Décomposition effective de Jordan-Chevalley et ses retombées en enseignement" (PDF). Gazette des Mathématiciens: 15–19 – via HAL. {{cite journal}}: Check date values in: |date= (help)
  3. ^ Geck, Meinolf (2022-05-18), On the Jordan--Chevalley decomposition of a matrix, doi:10.48550/arXiv.2205.05432, retrieved 2024-01-09
  4. ^ Conrad, Keith (accessed Jan 2024). "Semisimplicity" (PDF). Expository papers. {{cite web}}: Check date values in: |date= (help)CS1 maint: url-status (link)
  5. ^ Ring Theory. Academic Press. 18 April 1972. ISBN 9780080873572.
  6. ^ Cohn, Paul M. (2012). Further Algebra and Applications. Springer London. ISBN 978-1-85233-667-7. {{cite book}}: Check date values in: |year= / |date= mismatch (help)
  7. ^ This is not easy to see in general but is shown in the proof of (Jacobson 1979, Ch. III, § 7, Theorem 11.). Editorial note: we need to add a discussion of this matter to "semisimple operator".
  8. ^ Weber, Brian (02 Oct 2012). "Lecture 8 - Preservation of the Jordan Decomposition and Levi's Theorem" (PDF). Course Notes. Retrieved 09 Jan 2024. {{cite web}}: Check date values in: |access-date= and |date= (help); line feed character in |title= at position 53 (help)
  9. ^ Fulton & Harris 1991, Theorem 9.20.
  10. ^ Serre 1992, LA 5.17. Lemma 6.7. The endomorphism
  11. ^ Serre 1992, LA 5.19. Theorem 7.1.
  12. ^ Waterhouse 1979, Theorem 9.2.
  13. ^ Waterhouse 1979, Theorem 9.3.